content
stringlengths
1
15.9M
\section{Introduction} Recent observations reveal the widespread existence of magnetic fields in the universe and are producing much firmer estimates of their strengths in interstellar and intergalactic space. They also appear to be a common property of the intracluster medium of galaxy clusters, extending well beyond the core regions (see \cite{K} and references therein). Strengths of ordered magnetic fields in the intracluster medium of cooling flow clusters exceed those typically associated with the interstellar medium of the Milky Way, suggesting that galaxy formation and even cluster dynamics are, at least in some cases, influenced by magnetic forces. Furthermore, reports of Faraday rotation associated with high redshift Lyman-$\alpha$ absorption systems seem to imply that dynamically significant magnetic fields may be present in condensations at high redshift \cite{KPZ}. The more we look for extragalactic magnetic fields, the more ubiquitous we find them to be. Large-scale magnetic fields introduce new ingredients into the standard, but nevertheless uncertain, picture of the early universe. They seem unlikely to survive an epoch of inflation, but it is conceivable that large-scale fields and magnetic inhomogeneities could be generated at the end of that era or in subsequent phase transitions (see, e.g., \cite{R}). Studies of magnetogenesis are partly motivated by the need to explain the origin of large-scale galactic fields. Typical spiral galaxies have magnetic fields of the order of a few $\mu$G coherent over the plane of their disc. Such fields could arise from a relatively large primordial seed field, amplified by the collapse of the protogalaxy, or by a much weaker one that has been strengthened by the galactic dynamo. Provided that this mechanism is efficient, the seed can be as low as $\sim10^{-30}$G at present \cite{dlt}. However, in the absence of nonlinear amplification, seeds of the order of $10^{-12}$G or even $10^{-8}$G are required \cite{Ku}. Determining whether the origin of galactic and cluster magnetic fields is primordial or post-recombination is a difficult task, since strong amplification in these virialized systems overwhelms any traces of their earlier history. In contrast, magnetic effects on the cosmic microwave background (CMB) anisotropies, or any magnetic presence away from clusters and galaxies, can provide better insight into these early phases. If large-scale magnetic fields are present throughout the universe today, their structure and spectrum should bear clearer signatures of their past. Thus, improved direct observations, such as high resolution Faraday rotation maps and the study of extragalactic cosmic rays, may help in this respect \cite{O}. For example, we would like to know whether or not the intergalactic voids are permeated by a widespread magnetic field, and whether there is magnetic field evolution in galaxies. If large-scale magnetic fields were present in the early universe, were they dynamically significant, and if so, how have they affected the formation and evolution of the observed structure? It is known that element abundances constrain the strength of a primordial field at the nucleosynthesis epoch \cite{gr}. Stronger limits on a primordial magnetic field are imposed via CMB anisotropies, since the field distorts the acoustic peaks and induces Faraday rotation in the polarization \cite{adgr,cmb,dky}. In this article we assume the existence of a large-scale ordered magnetic field of primordial origin {\em a priori}, and we investigate the magnetic effects on density inhomogeneities. Specifically, we analyze magnetized density perturbations, magnetized cosmic vortices (i.e., rotational instabilities), and magnetized shape distortion. Magnetized density perturbations were studied by Ruzmaikina and Ruzmaikin \cite{RR} in Newtonian theory, while Wasserman \cite{W} looked at the rotational behavior of a magnetized fluid. Kim et al. \cite{kor} derive a magnetized Jeans length, assuming that there are no density perturbations in the absence of the field. In a relativistic treatment, Battaner et al. \cite{bfj} investigated magnetized structure formation in the radiation era. Jedamzik et al. and Subramanian \& Barrow \cite{sb} have considered magnetic dissipative effects at recombination. We generalize aspects of these previous treatments by giving a fully relativistic analysis of the scalar and vector contributions of the magnetic field to the evolution of density inhomogeneity. We consider not only density perturbations and rotational instabilities, but also the shape-distortion effects of the field. Density perturbations are found explicitly in the radiation and dust eras, including a new solution that shows how the relativistic magneto-curvature coupling acts to enhance growth on superhorizon scales. The existence of the magneto-curvature coupling was first identified by Tsagas and Barrow \cite{TB}. The nonadiabatic magnetic effect on the modes is clearly identified, including the magnetized isocurvature modes. New solutions are also found for rotational instabilities, which are significantly affected by the field, and we show that magnetic effects actively generate shape distortion in the density distribution. We follow the relativistic analysis of cosmic electromagnetic fields given by Ellis \cite{E2}, and we use the covariant and gauge-invariant approach to perturbations \cite{EB,EB2,MT}. A covariant and gauge-invariant analysis of magnetized density perturbations was first developed by Tsagas and Barrow \cite{TB,T}, whose results we extend. We adopt the usual approximation that in the background, which is a spatially flat Friedmann-Robertson-Walker (FRW) model,\footnote{ Spatial flatness is necessary for the gauge-invariance of all the perturbative variables \cite{TB}.} the ordered large-scale magnetic field is too weak to destroy spatial isotropy. The weak field approximation is an acceptable physical approximation when the field energy density is a small fraction of the isotropically distributed dominant energy density (see \cite{weak} for further discussion). If one demands strict mathematical homogeneity in the background, i.e., if one refuses to accept a coherent test field in the background, then one must adopt a Bianchi I model for the background. However, this is a highly complicated approach, which in the end will give results that are practically indistinguishable from those with an FRW background. (We are currently completing calculations that confirm this statement \cite{tm}.) The standard assumption of very high conductivity is also made, so that we can ignore large-scale electric fields, while maintaining the desired coupling between the fluid and the magnetic field. We use a single perfect fluid model, which is reasonable in the radiation era, but does not apply during recombination, while after last scattering, it means that our solutions only apply to a baryon-dominated universe. (See \cite{ck} for a discussion of the effects of cold dark matter (CDM) potential wells on the field.) The assumption of a weak background magnetic field ensures that the field terms in our linearized equations are first-order. Current limits on the strength of a primordial magnetic field show that its influence is secondary relative to that of the dominant matter component. In practice, this allows for the possibility that second-order fluid terms can have a strength comparable to that of the linear magnetic terms. Nevertheless, such second-order terms do not contain any further information regarding the lowest order influence of the field on gravitational instability. We ignore the second-order fluid terms, even though they may be of comparable magnitude to the first-order magnetic effects. This approach is consistent at linear order, and allows us to isolate the lowest order magnetic effects on gravitational instability. Our aim is to identify the sources of the magnetic effects, calculate their impact to lowest order, and discuss their implications for the evolution of density inhomogeneities. In Sec. II, we outline the formalism and the main equations that govern the coupled evolution of density inhomogeneity, the magnetic field and the curvature. Sec. III considers magnetized density perturbations, identifying the nonadiabatic effects of the field. We find a new solution on superhorizon scales in the radiation era, showing how the magneto-curvature coupling slightly enhances growth. In the radiation era, a small damping effect is wrongly predicted when the magneto-curvature coupling is ignored. On subhorizon scales, magneto-sonic waves in the radiation era have a slightly increased frequency, leading to a decrease in the spacing of CMB acoustic peaks. In the dust era, the growing mode on small scales is slightly damped by magnetic effects. We also find the pure-magnetic density perturbations, i.e., the fluctuations created in a smooth plasma at magnetogenesis. These include growing modes. Magnetized isocurvature modes are characterized, and found explicitly on superhorizon scales. These modes all decay, in the radiation and dust eras. Magnetized cosmic vortices are considered in Sec. IV. We show that magnetized vorticity is scale-dependent , and that the field generates precession in the rotational vector. We solve exactly for the rotational instabilities, showing that they propagate as Alfv\'{e}n (vector) waves on small scales during the radiation era. After recombination, such vortices persist for longer than non-magnetized vortices. Sec. V investigates magnetized shape distortion, showing that the field is as an active source of distortion. Purely magnetic distortion on superhorizon scales in the dust era is shown to have a growing mode. Conclusions are given in Sec. VI. We use units with $c=1=8\pi G$ and our signature is $(-+++)$; spacetime indices are $a,b,\cdots$ and (square) round brackets enclosing indices denote (anti-)symmetrization. \section{Cosmic magnetohydrodynamics} As noted above, the cosmic magnetic field $B^a$ must have weak energy density $\rho_{\rm mag}={1\over2}B_aB^a$ to be consistent with observational limits, so that $c_{\rm a}\ll1$, where $c_{\rm a}$ is the Alfv\'{e}n speed. The Alfv\'{e}n speed, which effectively leads to a nonadiabatic increase in the sound speed, is given by \begin{equation} c_{\rm a}^2= {B^2\over\rho} \,, \label{1} \end{equation} where $\rho$ is the energy density of the cosmic fluid. In the limit of vanishing density inhomogeneity, i.e., in the background, the field is uniform, but its weak magnitude means that it does not disturb the background isotropy, so that the magnetic anisotropic stress is negligible in the background. In the actual inhomogeneous universe the field's influence propagates via: \begin{enumerate} \item[] the background energy density and pressure ($p_{\rm mag}={1\over3}\rho_{\rm mag}$) occur in terms of the form $c_{\rm a}^2{\cal P}$, where ${\cal P}$ is a perturbed quantity, and despite their weakness, they can have observable consequences (e.g., a change in the spacing of CMB acoustic peaks in the radiation era); \item[] spatial gradients of $\rho_{\rm mag}$ couple with gradients of $\rho$ and thus alter the fluctuations of $\rho$ (in particular, introducing nonadiabatic modes); \item[] the background direction of the field introduces anisotropy by picking out preferred directions in perturbed vector and tensor fields, and preferred directional derivatives of perturbed scalar/ vector/ tensor fields, leading to effects such as Faraday rotation; \item[] the background direction of the field is also the source of the magneto-curvature coupling, via terms of the form $K_{abcd}B^d$, where $K_{abcd}$ is the part of the curvature tensor which vanishes in the background; \item[] fluctuations in the direction of the field generate new anisotropies that can source magnetized vortices (leading in particular to Alfv\'{e}n waves) and shape distortion. \end{enumerate} We include all of these aspects in our analysis, so that we incorporate the full range of scalar (magnetic energy density and isotropic pressure) and vector (anisotropic stress) effects of the field, allowing for fluctuations in both the magnitude and direction. In order to provide a transparent relativistic generalization of Newtonian analysis, and to use variables that as far as possible have a direct physical interpretation, we adopt a covariant Lagrangian approach \cite{E2,EB,EB2,MT,M1,mb}. This continues and develops the work of \cite{TB}. In particular, we discuss in detail the physical meaning and implications of the density perturbation solutions, and we extend the investigation to cover magnetized cosmic vortices and shape distortion. The cosmic perfect fluid defines a unique four-velocity $u_a$ (with $u_au^a=-1$), and then $h_{ab}=g_{ab}+u_au_b$, where $g_{ab}$ is the spacetime metric, projects into the local rest spaces of comoving observers. The projection of a vector is $V_{\langle a\rangle}=h_a{}^bV_b$, and a projected second rank tensor $S_{ab}$ splits irreducibly as \[ S_{ab}={\textstyle{1\over3}}Sh_{ab} +\varepsilon_{abc}S^c+ S_{\langle ab\rangle }\,, \] where $S\equiv h_{ab}S^{ab}$ is the spatial trace, $S_a\equiv{1\over2}\varepsilon_{abc}S^{bc}$ is the spatial vector dual to the skew part of $S_{ab}$, and $S_{\langle ab\rangle}\equiv[h_{(a}{}^ch_{b)}{}^d-{1\over3}h^{cd} h_{ab}]S_{cd}$ is the projected symmetric tracefree (PSTF) part. Here $\varepsilon_{abc}=\eta_{abcd}u^d$ is the projection of $\eta_{abcd}$, the spacetime alternating tensor. The covariant derivative splits into a comoving time derivative $\dot{J}_{a\cdots b}=u^c\nabla_cJ_{a\cdots b}$, and a covariant spatial derivative $\mbox{D}_cJ_{a\cdots b}=h_c{}^dh_a{}^e\cdots h_b{}^f\nabla_dJ_{e\cdots f}$. Then we define a covariant spatial divergence and curl that generalize the Newtonian operators to curved spacetime \cite{M1}: \begin{eqnarray*} \mbox{div}\,V= \mbox{D}^aV_a\,,&~~& (\mbox{div}\,S)_a= \mbox{D}^bS_{ab}\,,\\ \mbox{curl}\,V_a=\varepsilon_{abc}\mbox{D}^bV^c\,,&~~& \mbox{curl}\,S_{ab}=\varepsilon_{cd(a}\mbox{D}^cS_{b)}{}^d\,. \end{eqnarray*} The fluid kinematics are described by the expansion $\Theta=\mbox{div}\,u$, four-acceleration $A_a=\dot{u}_a$, vorticity $\omega_a=-{1\over2}\mbox{curl}\,u_a$ and shear $\sigma_{ab}=\mbox{D}_{\langle a}u_{b\rangle}$. Local curvature is described by the Ricci tensor $R_{ab}$, while nonlocal tidal forces and gravitational radiation are described by the electric and the magnetic parts of the Weyl tensor, $E_{ab}=E_{\langle ab\rangle }= C_{acbd}u^cu^d$ and $H_{ab}=H_{\langle ab\rangle }={1\over2} \varepsilon_{acd}C^{cd}{}{}_{be}u^e$. The magnetized perfectly conducting fluid has energy density $\rho$ and isotropic pressure $p$. The magnetic field is $B_a=B_{\langle a\rangle}$, with energy density, isotropic pressure and anisotropic stress given respectively by \begin{equation} \rho_{\rm mag}={\textstyle{1\over2}}B_aB^a\,,~p_{\rm mag}= {\textstyle{1\over3}}\rho_{\rm mag}\,,~ \pi_{ab}=-B_{\langle a}B_{b\rangle}\,. \label{emt'}\end{equation} Then the total energy-momentum tensor is \begin{equation} T_{ab}=\left(\rho+\rho_{\rm mag}\right)u_au_b+ \left(p+p_{\rm mag}\right)h_{ab}+\pi_{ab} \,. \label{emt} \end{equation} Notice that the absence of an electric field means that there is no energy flux (Poynting vector). The magnetic field appears from Eq. (\ref{emt}) to behave like a radiation fluid with anisotropic stress. However, this fluid picture does not fully encompass the vector properties of the field, and in particular, its coupling to the curvature. In the background, $B_a$ is weak enough not to affect the isotropy, i.e. the anisotropic stress is negligible in the background, and $\rho_{\rm mag}\ll \rho$. The background expansion is $\Theta=3H$, where $H=\dot{a}/a$ is the Hubble rate. The background is covariantly characterized by \begin{eqnarray} &&\mbox{D}_a\Theta=\mbox{D}_a\rho=\mbox{D}_ap=\mbox{D}_bB_a=0 \,, \label{nograd}\\ &&A_a=\omega_a=0 \,, \label{novec}\\ &&\sigma_{ab}=E_{ab}=H_{ab}={\cal R}_{ab}=\pi_{ab}=0 \,, \label{noten} \end{eqnarray} where \[ {\cal R}_{ab}= h_a{}^{c}h_b{}^{d}R_{cd}+R_{acbd}u^cu^d+\mbox{D}_cu_a\mbox{D}_bu^c-\Theta\mbox{D}_bu_a\,, \] with $R_{ab}$ the Ricci tensor and $R_{abcd}$ the Riemann tensor. Note that ${\cal R}_{ab}$ is the intrinsic 3-Ricci tensor of spatial hypersurfaces only if $\omega_a=0$; otherwise there are no such hypersurfaces orthogonal to $u^a$ \cite{E2}. Quantities that vanish in the background are gauge-invariant, and they covariantly describe linear deviations from homogeneity and anisotropy. We collect below the linearized evolution and constraint equations given in \cite{TB}, rewritten in the streamlined formalism of \cite{M1}, which considerably simplifies the equations and facilitates analysis of their properties. The following covariant identities \cite{MT} are used in deriving the equations (assuming a flat background with vanishing cosmological constant): \begin{eqnarray} {\rm curl}\,\mbox{D}_af &=&-2\dot{f}\omega_a \,, \label{id2} \\ \left(a\mbox{D}_af\right)^{\displaystyle{\cdot}} &=& a\mbox{D}_a\dot{f}+a\dot{f}A_a \,, \label{a14}\\ \mbox{D}^2\left(\mbox{D}_af\right) &=&\mbox{D}_a\left(\mbox{D}^2f\right) +2\dot{f}{\rm curl}\,\omega_a \,, \label{a19}\\ \left(a\mbox{D}_aJ_{b\cdots}\right)^{\displaystyle{\cdot}} &=& a\mbox{D}_a\dot{J}_{b\cdots}\,, \label{id1}\\ \mbox{D}_{[a}\mbox{D}_{b]}V_c &=&0= \mbox{D}_{[a}\mbox{D}_{b]}S^{cd} \,, \label{a17}\\ {\rm div}\,{\rm curl}\, V &=& 0 \,,\label{a20}\\ ({\rm div}\,{\rm curl}\, S)_{a} &=& {\textstyle{1\over2}}{\rm curl}\, ({\rm div}\, S)_{a}\,,\label{a18}\\ {\rm curl}\,{\rm curl}\, V_a &=& \mbox{D}_a ({\rm div}\, V) -\mbox{D}^2V_a\,, \label{id3}\\ {\rm curl}\,{\rm curl}\, S_{ab} &=& {\textstyle{3\over2}}\mbox{D}_{\langle a}({\rm div}\, S)_{b\rangle}-\mbox{D}^2S_{ab}\,,\label{a23} \end{eqnarray} where the vectors and tensors vanish in the background, $S_{ab}=S_{\langle ab\rangle}$. The magnetic field itself does not vanish in the background, so that its projected derivatives do not commute at linear order: the vector identity in Eq. (\ref{a17}) is replaced by \[ \mbox{D}_{[a}\mbox{D}_{b]}B_c={\textstyle{1\over2}}{\cal R}_{dcba}B^d-\varepsilon_{abd}\omega^d\dot{B}_c\,, \] where ${\cal R}_{abcd}$ is formed from $R_{abcd}$ and the kinematic quantities \cite{TB}. This non-commutativity is the root of the magneto-curvature coupling found in \cite{TB}. \subsection{Maxwell's equations} In the infinite conductivity limit, Maxwell's equations \cite{E2,mb} provide three constraints, \begin{eqnarray} \mbox{div}\,B&=&0 \,,\label{Max1}\\ \mbox{curl}\,B_a&=&\varepsilon_{abc}B^bA^c+j_a \,,\label{m'}\\ \omega^aB_a &=& {\textstyle{1\over2}}q\,, \label{m''} \end{eqnarray} where $j_a$ is the current and $q$ the charge density generated by fluctuations, and one propagation equation \begin{equation} \dot{B}_{\langle a\rangle}=-{\textstyle{2\over3}}\Theta B_a+ \sigma_{ab}B^b+ \varepsilon_{abc}B^b\omega^c \,, \label{Max2} \end{equation} which is the covariant form of the induction equation. Note that $\dot{B}_{\langle a\rangle}=\dot{B}_a-A_bB^bu_a$, and $B^aA_a=0$ to first order only in the case of a pressure-free perfect fluid \cite{TB}. Contracting Eq. (\ref{Max2}) with $B^a$, and neglecting the second order term $\sigma_{ab}\pi^{ab}$, we deduce the radiation-like evolution law of the magnetic energy density, \begin{equation} (B^2)^{\displaystyle{\cdot}}+{\textstyle{4\over3}}\Theta(B^2)=0 \,. \label{B^2ev} \end{equation} We can also derive the evolution of the anisotropic stress from Eq. (\ref{Max2}): \begin{equation} \dot{\pi}_{ab}=-4H\pi_{ab}-{\textstyle{2\over3}}c_{\rm a}^2\rho\sigma_{ab}\,,\label{pi} \end{equation} \subsection{Conservation laws} Energy density conservation is expressed via the equation of continuity \ \dot{\rho}+\Theta(1+w)\rho=0 \,, \ where $w=p/\rho$. Notice the absence of magnetic terms in this equation, since field energy conservation holds separately as a consequence of Maxwell's equations, as shown in Eq. (\ref{B^2ev}). The two energy conservation equations imply \[ \left(c_{\rm a}^2\right)^{\displaystyle{\cdot}}=(w-{\textstyle{1\over3}})\Theta c_{\rm a}^2\,. \] On the other hand, the field does enter conservation of momentum density \begin{equation} (1+w)\rho A_a+c_{\rm s}^2\mbox{D}_a\rho+ \varepsilon_{abc}B^b\mbox{curl}\,B^c=0 \,, \label{mdc} \end{equation} where $c_{\rm s}^2=\dot{p}/\dot{\rho}$ is the adiabatic sound-speed squared (with $\mbox{D}_ap=c_{\rm s}^2\mbox{D}_a\rho$). This equation reflects the momentum density exchange between the fluid and the field. The magnetic field is a source of acceleration (provided that curl $B_a$ is not parallel to $B_a$); it can destroy the geodesic motion of the matter even in the absence of pressure. \subsection{Kinematic equations} Evolution of the expansion is governed by the Raychaudhuri equation \begin{equation} \dot{\Theta}+{\textstyle{1\over3}}\Theta^2+ {\textstyle{1\over2}}(1+3w)\rho-\frac{c_{\rm a}^2}{3(1+w)}{\cal R}+ \frac{1}{2(1+w)a^2}\left(2c_{\rm s}^2\Delta+c_{\rm a}^2{\cal B}\right)- \Lambda=0 \,, \label{Ray} \end{equation} where $\Lambda$ is the cosmological constant, ${\cal R}=h^{ab}{\cal R}_{ab}$ is the projected curvature scalar, and \[ \Delta=a\mbox{D}^a\Delta_a\,,~~\Delta_a={a\mbox{D}_a\rho\over\rho}\,,~~ {\cal B}={a^2\mbox{D}^a{\cal B}_a\over B^2}\,,~~ {\cal B}_a=\mbox{D}_aB^2\,, \] describe perturbations in the fluid and field energy densities. Note that the overall magnetic effect includes a coupling to the projected curvature, $c_{\rm a}^2{\cal R}$. Magnetic influence on cosmic rotation is encoded in the vorticity propagation equation \begin{equation} \dot{\omega}_a+2H\omega_a=-{\textstyle{1\over2}} \mbox{curl}\,A_a \,, \label{v'} \end{equation} which may be rewritten, after eliminating the acceleration term via Eq. (\ref{mdc}), as \begin{equation} \dot{\omega}_a+\left(2-3c_{\rm s}^2\right)H\omega_a= -\frac{1}{2(1+w)\rho}B^b\mbox{D}_b\mbox{curl}\,B_a \,. \label{dotom} \end{equation} Thus there is a magnetically induced vorticity component parallel to $\mbox{curl}\,B_a$. The effect disappears if the directional derivative $B^b\mbox{D}_b\mbox{curl}\,B_a$ vanishes, i.e., when $\mbox{curl}\,B_a$ does not change along the magnetic force lines. Kinematic anisotropies evolve via the shear propagation equation \begin{eqnarray} \dot{\sigma}_{ab}+2H\sigma_{ab}&=& -\frac{c_{\rm s}^2}{a(1+w)}\mbox{D}_{\langle a}\Delta_{b\rangle}- \frac{1}{2(1+w)\rho}\mbox{D}_{\langle a}{\cal B}_{b\rangle}+ \frac{c_{\rm a}^2}{3(1+w)}{\cal R}_{\langle ab\rangle} \nonumber\\&\mbox{}&+ {\textstyle\frac{1}{2}}\pi_{ab}+ \frac{1}{(1+w)\rho}B^c\mbox{D}_c\mbox{D}_{\langle a}B_{b\rangle}- E_{ab} \,. \label{dotsh} \end{eqnarray} The direct magnetic effects propagate through the field's anisotropic stress ($\pi_{ab}$), as well as via anisotropies in the distribution of magnetic energy density ($\mbox{D}_{\langle a}{\cal B}_{b\rangle}$) and of the field vector itself ($\mbox{D}_{\langle a}B_{b\rangle}$). The latter effect vanishes when $\mbox{D}_{\langle a}B_{b\rangle}$ is invariant along the magnetic force lines. Also, the coupling between the field and the projected curvature has led to an extra magneto-geometrical contribution, $c_{\rm a}^2{\cal R}_{\langle ab\rangle}$. The kinematic quantities also obey constraint equations: \begin{eqnarray} \left(\mbox{div}\,\sigma\right)_a&=& {\textstyle\frac{2}{3}}\mbox{D}_a\Theta+\mbox{curl}\,\omega_a \,, \label{shcon}\\ \mbox{div}\,\omega&=&0 \,, \label{vorcon}\\ H_{ab}&=&\mbox{curl}\,\sigma_{ab}+ \mbox{D}_{\langle a}\omega_{b\rangle} \,. \label{Hcon} \end{eqnarray} \subsection{Curvature} The electric and magnetic Weyl tensors obey Maxwell-like equations \cite{E2,mb}: \begin{eqnarray} \dot{E}_{ab}+3HE_{ab}-\mbox{curl}\,H_{ab} &=& -{\textstyle{1\over2}}\rho(1+w)\sigma_{ab}+3H\pi_{ab} \,,\label{gem1}\\ \dot{H}_{ab}+3HH_{ab}+\mbox{curl}\,E_{ab} &=&{\textstyle{1\over2}} \mbox{curl}\,\pi_{ab}\,,\label{gem2}\\ \left({\rm div}\,E\right)_a&=&{\textstyle{1\over3}}\mbox{D}_a\rho +{\textstyle{1\over6}}{\cal B}_a -{\textstyle{1\over2}}\left({\rm div}\,\pi\right)_a \,,\label{gem3}\\ \left({\rm div}\,H\right)_a&=&(1+w)\rho\omega_a\,,\label{gem4} \end{eqnarray} where we have used Eq. (\ref{pi}). For a magnetized fluid the projected curvature tensor ${\cal R}_{ab}$ is not in general symmetric, but has the form ${\cal R}_{ab}={\cal R}_{\langle ab\rangle}+{\textstyle{1\over3}}{\cal R}h_{ab}+\varepsilon_{abc}{\cal R}^c$, where \cite{TB} \begin{eqnarray} {\cal R}_{\langle ab\rangle}&=& \mbox{D}_{\langle a}A_{b\rangle}- \frac{1}{a^3}\left(a^3\sigma_{ab}\right)^{\displaystyle{\cdot}}+\pi_{ab}\,,\label{3R_ab}\\ {\cal R}_a &=&\mbox{curl}\,A_a -\frac{1}{a^3}\left(a^3\omega_{a}\right)^{\displaystyle{\cdot}}\,,\\ {\cal R}&=&2\left(\rho-{\textstyle{1\over3}}\Theta^2 +\Lambda\right)\label{3R''} \end{eqnarray} Note that the background relation $3H^2=\rho+\Lambda$ ensures that ${\cal R}$ vanishes in the background, so that it is gauge-invariant. By Eq. (\ref{v'}), the vector part of ${\cal R}_{ab}$ is simply ${\cal R}_a=-H\omega_a$, which vanishes when the vorticity vanishes. The dimensionless curvature perturbation ${\cal K}=a^2{\cal R}$ has comoving gradient \cite{TB} \begin{equation} a\mbox{D}_a{\cal K}=2\rho a^2\Delta_a+a^3{\cal B}_a-4Ha^3\mbox{D}_a\Theta\,, \label{r}\end{equation} which evolves as \cite{TB} \begin{equation} \left(a\mbox{D}_a{\cal K}\right)^{\displaystyle{\cdot}}= {2aH\over\rho(1+w)}\mbox{D}^2\left( 2\rho c_{\rm s}^2\Delta_a+a{\cal B}_a\right) +24a^3H^2c_{\rm s}^2{\rm curl}\,\omega_a\,. \label{r'}\end{equation} \subsection{Evolution of inhomogeneities} The key gauge-invariant quantities describing inhomogeneity are the comoving spatial gradients of the fluid density, the field density, the expansion and the field vector: \begin{equation} \Delta_a={a\mbox{D}_a\rho\over\rho}\,,~~ {\cal B}_a=\mbox{D}_aB^2\,,~~ \Theta_a=a\mbox{D}_a\Theta\,, ~~B_{ab}=a\mbox{D}_bB_a\,. \label{grad} \end{equation} Their propagation equations are \cite{TB} \begin{eqnarray} \dot{\Delta}_a&=&3wH\Delta_a+(1+w)\Theta_a+ \frac{3aH}{\rho}\varepsilon_{abc}B^b\mbox{curl}\,B^c \,, \label{dotDel_a}\\ \dot{\Theta}_a&=&-2H\Theta_a- {\textstyle{1\over2}}{\rho}\Delta_a- \frac{c_{\rm s}^2}{1+w}\mbox{D}^2\Delta_a- {\textstyle{1\over2}}{a}{\cal B}_a- \frac{a}{2(1+w)\rho}\mbox{D}^2{\cal B}_a \nonumber\\&\mbox{}&+ {\textstyle\frac{3}{2}}a\varepsilon_{abc}B^b\mbox{curl}\,B^c -\left[6c_{\rm s}^2+\frac{4c_{\rm a}^2}{1+w}\right]aH\mbox{curl}\,\omega_a \,,\label{dotThe_a}\\ \dot{B}_{ab}&=&-2HB_{ab}+ \frac{c_{\rm s}^2H}{1+w}\left[3B_{\langle a}\Delta_{b\rangle}+ B_{[a}\Delta_{b]}\right]-B_{\langle a}\Theta_{b\rangle}- B_{[a}\Theta_{b]} \nonumber\\&&{} +aB^c\mbox{D}_c\sigma_{ab}+a\varepsilon_{ab}{}{}^dB^c\mbox{D}_{\langle c}\omega_{d\rangle}-aB_{[a}{\rm curl}\,\omega_{b]}-a\varepsilon_{acd}B^cH^d{}_b \,. \label{dotB_ab} \end{eqnarray} Note how the magnetic field couples to the magnetic Weyl curvature via the last term in Eq. (\ref{dotB_ab}). By eliminating the expansion gradients from the time derivative of Eq. (\ref{dotDel_a}), we arrive at \cite{TB} \begin{eqnarray} \ddot{\Delta}_a&=&-\left(2+3c_{\rm s}^2-6w\right)H\dot{\Delta}_a+ {\textstyle{1\over2}}\left[\left(1-6c_{\rm s}^2+8w-3w^2\right)\rho -2\left(3c_{\rm s}^2-5w\right)\Lambda\right]\Delta_a \nonumber\\&&{}+ c_{\rm s}^2\mbox{D}^2\Delta_a +\frac{a}{2\rho}\mbox{D}^2{\cal B}_a+ {3a\over\rho}\left[\left(c_{\rm s}^2-w\right)\rho+ \left(1+c_{\rm s}^2\right)\Lambda\right]\varepsilon_{abc}B^b\mbox{curl}\,B^c \nonumber\\&&{}+ {\textstyle{1\over2}}(1+w)a{\cal B}_a+ \left(\frac{3aH}{\rho}\right)\varepsilon_{abc}B^b\mbox{curl}\, \dot{B}^{\langle c\rangle}+ \left[6(1+w)c_{\rm s}^2+4c_{\rm a}^2\right]aH\mbox{curl}\,\omega_a \,.\label{ddotDel_a} \end{eqnarray} In the Newtonian limit, Eq. (\ref{ddotDel_a}) recovers the results of \cite{RR}. The relativistic correction terms are the last three terms on the right hand side, i.e., the terms with ${\cal B}_a$, $\varepsilon_{abc}B^b{\rm curl}\,\dot{B}^{\langle c\rangle}$, and ${\rm curl}\,\omega_a$. \section{Magnetized density perturbations} The equations (\ref{dotDel_a})--(\ref{dotB_ab}) provide the basis for a complete description of coupled density-magnetic inhomogeneities. We begin by isolating the evolution equations for the density perturbation scalars $\Delta$ (a covariant alternative to the density contrast $\delta\rho/\rho$) and ${\cal B}$ (describing fluctuations in the magnetic energy density), and the curvature perturbation ${\cal K}$: \[ \Delta=a\mbox{D}^a\Delta_a\,,~~{\cal B}={a^2\mbox{D}^a{\cal B}_a\over B^2}\,,~~ {\cal K}=a^2{\cal R}\,. \] The required evolution equations are \cite{TB}: \begin{eqnarray} \ddot{\Delta}&=&-\left(2+3c_{\rm s}^2-6w\right)H\dot{\Delta}+{\textstyle{1\over2}} \left[\left(1-6c_{\rm s}^2+8w-3w^2\right)\rho-2 \left(3c_{\rm s}^2-5w\right)\Lambda\right]\Delta \nonumber\\&&{} +c_{\rm s}^2\mbox{D}^2\Delta -{\textstyle{1\over2}} \left[\left(1-3c_{\rm s}^2+2w\right)\rho-\left(1+3c_{\rm s}^2 \right)\Lambda\right]c_{\rm a}^2{\cal B}+{\textstyle{1\over2}} c_{\rm a}^2\mbox{D}^2{\cal B} \nonumber\\&&{}+{\textstyle{1\over3}} \left[\left(2-3c_{\rm s}^2+3w\right)\rho-\left(1+3c_{\rm s}^2\right) \Lambda\right]c_{\rm a}^2{\cal K} \,,\label{ddotDel}\\ \dot{{\cal B}}&=& \frac{4}{3(1+w)}\dot{\Delta}+ \frac{4(c_{\rm s}^2-w)H}{1+w}\Delta \,, \label{dotcB}\\ \dot{{\cal K}}&=& \frac{4c_{\rm s}^2H}{1+w}\Delta +\frac{2c_{\rm a}^2H}{1+w}{\cal B} \,. \label{dotcK} \end{eqnarray} This system of equations governs the coupling between density fluctuations $\Delta$, magnetic fluctuations ${\cal B}$ and curvature fluctuations ${\cal K}$. Eq. (\ref{dotcK}) shows that ${\cal K}$ grows if $\Delta$ and ${\cal B}$ are growing, while Eq. (\ref{dotcB}) shows that if $c_{\rm s}^2\geq w$ (which holds in the radiation and dust eras), then ${\cal B}$ grows in concert with growing $\Delta$. The magnetic field introduces a direct effect, via the term $c_{\rm a}^2{\cal K}$ in Eq. (\ref{ddotDel}), of the curvature on the density perturbations. In the non-magnetized case, there are two modes of $\Delta$, which is governed by the single second-order equation (\ref{ddotDel}) with $c_{\rm a}=0$. In this case, the evolution of $\Delta$ is independent of ${\cal K}$, and ${\cal K}$ is determined once $\Delta$ is known, via Eq. (\ref{dotcK}). Magnetism introduces two additional modes, since the system has four degrees of freedom. These modes are nonadiabatic, and can source density perturbations, i.e., even when $\Delta(t_0)=0=\dot{\Delta}(t_0)$, magnetic effects will lead to $\Delta\neq0$ for $t>t_0$. If one omits the magneto-curvature effect, then the evolution equation for ${\cal K}$, Eq. (\ref{dotcK}), is uncoupled from the system, which can then be decoupled via a third-order equation in $\Delta$. Neglecting the magneto-curvature effect thus removes one of the additional nonadiabatic modes. For zero cosmological constant, we can solve the system analytically in the radiation and dust eras, treating super- and sub-horizon scales separately. Some solutions were given in \cite{TB}. There, however, magneto-curvature effects were neglected in three out of the four cases. Here, we generalize some of the solutions to incorporate the magneto-curvature coupling, and we show that the magneto-curvature coupling cannot in general be neglected, since it leads to important qualitative differences in the behavior of $\Delta$. \subsection{Radiation era} During the radiation era, $w=c_{\rm s}^2={1\over3}$, $\rho=\rho_0(a_0/a)^4$, and the Alfv\'{e}n speed does not change along the fluid flow, i.e., $\dot{c}_{\rm a}=0$, reflecting the radiation-like evolution of the magnetic energy density, as given by Eq. (\ref{B^2ev}). For the Fourier modes with comoving wave-number $k$, we get \begin{eqnarray} \left({a\over a_0}\right)^2\Delta'' &=& \left[2-{\textstyle{1\over3}}\left(\frac{ k}{k_{{\rm h}0}}\right)^2 \left({a\over a_0}\right)^2\right] \Delta- \left[1+{\textstyle{1\over2}} \left(\frac{ k}{k_{{\rm h}0}}\right)^2\left({a\over a_0}\right)^2 \right]c_{\rm a}^2{\cal B}+ 2c_{\rm a}^2{\cal K} \,, \label{rnddotDel}\\ {\cal B}'&=& \Delta'\,,\label{rn'}\\ \left({a\over a_0}\right){\cal K}'&=&\Delta+ {\textstyle{3\over2}}c_{\rm a}^2{\cal B} \,, \label{rndotcB-dotcK} \end{eqnarray} where a prime denotes $d/d(a/a_0)$ and $k_{{\rm h}0}=a_0H_0$ is the comoving wavenumber of the horizon at $a_0$. \subsubsection{Superhorizon scales and the curvature coupling} In the long wavelength limit $ k \ll k_{{\rm h}0}$, the system has the power-law solution \begin{eqnarray} \Delta&=&C_{(0)}+ \sum_{\alpha}C_{(\alpha)}\left({a\over a_0}\right)^\alpha \,,\label{lrnDel}\\ {\cal B} &=& -\left({2\over3 c_{\rm a}^2}\right)C_{(0)}+ \sum_{\alpha}C_{(\alpha)}\left({a\over a_0}\right)^\alpha \,,\label{lrn'}\\ {\cal K} &=&-\left({4\over3 c_{\rm a}^2}\right)C_{(0)}+\left(1+{\textstyle{3\over2}}c_{\rm a}^2\right) \sum_{\alpha}{C_{(\alpha)}\over\alpha}\left({a\over a_0}\right)^\alpha \,, \label{lrn''} \end{eqnarray} where $C_{(0)}$ and $C_{(\alpha)}$ are constants, and the parameter $\alpha$ satisfies the cubic equation \begin{equation} \alpha^3-\alpha^2-\left(2-c_{\rm a}^2\right)\alpha -(2+3c_{\rm a}^2)c_{\rm a}^2=0 \,. \label{lrDelz} \end{equation} The cubic has one positive and two negative roots. One of the negative roots corresponds to a decaying nonadiabatic mode. The other nonadiabatic mode is the $C_{(0)}$-mode, which is constant. The remaining cubic roots correspond to the magnetized versions of the standard adiabatic modes, one growing and one decaying. Since $c_{\rm a}^2$ is small, we can find the roots perturbatively. The zero-order roots are $0,-1,2$ (the $\alpha=0$ solution is spurious in the non-magnetized case). To lowest order, we find that: \begin{equation} \alpha=\left\{\begin{array}{r} 0-c_{\rm a}^2+O\left(c_{\rm a}^4\right)\,,\\ \\ -1+c_{\rm a}^2+O\left(c_{\rm a}^4\right)\,,\\ \\ 2+{\textstyle{1\over2}}c_{\rm a}^4 +O\left(c_{\rm a}^6\right)\,. \end{array}\right. \label{nc}\end{equation} Thus the adiabatic growing mode of the non-magnetized case is slightly {\em enhanced} by magnetic effects (the enhancement is not felt to lowest order in $c_{\rm a}^2$); the adiabatic decaying mode decays less rapidly by virtue of magnetic effects; the decaying nonadiabatic mode decays very slowly; and the final, nonadiabatic, mode is constant. To lowest order \begin{equation} \Delta= C_{(+)}\left({a\over a_0}\right)^2+C_{(1-)}\left({a\over a_0}\right)^{-1+c_{\rm a}^2}+C_{(0)}+ C_{(2-)}\left({a\over a_0}\right)^{-c_{\rm a}^2} \,.\label{new} \end{equation} The magnetic and curvature fluctuations are given by equations (\ref{lrn'}) and (\ref{lrn''}), with $\alpha$ given by Eq. (\ref{nc}). This new solution in Eq. (\ref{new}) can be compared with the solution that arises when the magneto-curvature coupling term $c_{\rm a}^2{\cal K}$ is ignored in Eq. (\ref{lrnDel}) \cite{TB}. Then the last term in Eq. (\ref{lrDelz}) falls away, leading to the quadratic $\alpha^2-\alpha-\left(2-c_{\rm a}^2\right)=0$. To lowest order \[ \alpha=\left\{\begin{array}{r} -1+{\textstyle{1\over3}}c_{\rm a}^2 \,,\\ \\ 2-{\textstyle{1\over3}}c_{\rm a}^2 \,, \end{array}\right. \] so that the density perturbation is given by \ \Delta=C_{(+)}\left({a\over a_0}\right)^{2-{1\over3}c_{\rm a}^2}+ C_{(-)}\left({a\over a_0}\right)^{-1+{1\over3}c_{\rm a}^2} +C_{(0)}\,, \ and the magnetic fluctuations are \[ {\cal B} = \Delta -\left[1-\left({2\over c_{\rm a}^2}\right)\right]C_{(0)}\,. \] Clearly, omitting the magneto-curvature coupling has a significant qualitative impact. Not only is one of the nonadiabatic modes ($C_{(2-)}$) removed, as expected, but we also find that the growing mode is slightly {\em damped}, at odds with the correct solution in Eq. (\ref{new}). Thus the magneto-curvature coupling, which was identified in general in \cite{TB}, turns out to have a crucial role in increasing (even though it is only by a small amount) the standard adiabatic modes of density perturbations on large scales in the radiation era. It is not reasonable to omit the magneto-curvature coupling in this case. \subsubsection{Subhorizon scales and magneto-sonic waves} At the opposite end of the wavelength spectrum, when $k \gg k_{{\rm h}0}$, we differentiate Eq. (\ref{rnddotDel}) and use Eq. (\ref{rndotcB-dotcK}) to decouple the system. Integrating once we get \ 6\left({a\over a_0}\right)^2\Delta''+ 2\left(\frac{ k}{k_{{\rm h}0}}\right)^2\left({a\over a_0}\right)^2 \left(1+{\textstyle{3\over2}} c_{\rm a}^2\right)\Delta= 6C_{{\cal K}}- 3C_{{\cal B}}c_{\rm a}^2\left({k\over k_{{\rm h}0}}\right)^2 \left({a\over a_0}\right)^2 \,, \ where $C_{{\cal K}}$ is an additional constant associated with curvature effects. (We have ignored higher order terms in $c_{\rm a}^2$, given the weakness of the magnetic field.) This has solution (to lowest order in $c_{\rm a}^2$) \begin{eqnarray} \Delta&=&\left[C_{(1)}-C_{{\cal K}} {\rm Si}\left(\beta {k\over k_{{\rm h}0}}{a\over a_0}\right) \right]\sin\left(\beta {k\over k_{{\rm h}0}}{a\over a_0}\right)\nonumber\\ &&{}+\left[C_{(2)}-C_{{\cal K}} {\rm Ci}\left(\beta {k\over k_{{\rm h}0}}{a\over a_0}\right) \right]\cos\left(\beta {k\over k_{{\rm h}0}}{a\over a_0}\right) -C_{{\cal B}}c_{\rm a}^2 \,, \label{srnDel} \end{eqnarray} where $C_{(i)}$ are constants, Si and Ci are the sine and cosine integral functions,\footnote{ ${\rm Si}(x)=\int_0^xt^{-1}\sin t\,dt$ and ${\rm Ci}(x)=\gamma+\ln x+\int_0^xt^{-1}(\cos t-1)dt$, where $\gamma= 0.578\cdots$ is Euler's constant \cite{AS}. } and \begin{equation} \beta=c_{\rm s}\left(1+{\textstyle{3\over4}}c_{\rm a}^2\right) \,, \label{srDelz} \end{equation} where $c_{\rm s}=1/\sqrt{3}$ is the adiabatic sound speed. Thus $\beta$ is the magnetized (nonadiabatic) sound speed of magneto-sonic waves. These waves differ slightly in amplitude and frequency from the adiabatic acoustic waves. The magneto-curvature coupling, reflected in the nonadiabatic $C_{{\cal K}}$ mode, has the effect of slightly modulating the amplitude of acoustic oscillations, with the effect decreasing as $a/a_0$ increases. The main magnetic effect is on the frequency. Comparing our result in Eq. (\ref{srDelz}) to the standard solutions of magnetic-free models (see, e.g., \cite{P}), we see that the field has increased the frequency of acoustic oscillations. Since $a\propto\sqrt{t}$, the magnetized acoustic frequency is \begin{equation} \nu_{\rm ac,mag}=\nu_{\rm ac}\left(1+{\textstyle{3\over2}}c_{\rm a}^2 \right)~\mbox{ where }~\nu_{\rm ac}={H_0\over3\pi }\left({k\over k_{{\rm h}0}}\right)^2\,. \label{3}\end{equation} This magnetic correction results from the the ``tensioning" effect of magnetic force lines in the plasma, which produces a nonadiabatic increase of the sound speed via a contribution from the Alfv\'{e}n speed. As a result, the magnetic influence brings the acoustic peaks of short-wavelength radiation density oscillations closer, producing in principle an observable signature on CMB anisotropies \cite{adgr}. An additional effect comes from the nonadiabatic constant mode in Eq. (\ref{srnDel}). Its presence suggests that the average value of the density contrast is {\em nonzero}, unlike the magnetic-free case. \subsection{Dust era} After recombination, in a baryon-dominated cold matter background, $w=0=c_{\rm s}^2$, $a=a_0 (t/t_0)^{2/3}$, $H=2/3t$ and $\rho=4/3t^2$. The Alfv\'{e}n speed is no longer constant, but by Eq. (\ref{B^2ev}) varies as \ c_{\rm a}^2=\left(c_{\rm a}^2\right)_0 \left({t_0\over t}\right)^{2/3}\,, \ reflecting the fact that the magnetic energy density drops faster than that of nonrelativistic matter. Thus magnetic effects grow weaker as the expansion of the universe proceeds beyond recombination. The equations for the Fourier modes become \begin{eqnarray} \Delta''&=&-{\textstyle{4\over3}}\left({t_0\over t}\right)\Delta'+ {\textstyle{2\over3}}\left({t_0\over t}\right)^2\Delta\nonumber\\ &&{}- {\textstyle{2\over3}}\left(c_{\rm a}^2\right)_0\left({t_0\over t}\right)^{8/3}\left[ 1+{\textstyle{1\over3}} \left(\frac{k}{k_{{\rm h}0}}\right)^2\left({t\over t_0}\right)^{2/3} \right]{\cal B}+ {\textstyle{8\over9}} \left(c_{\rm a}^2\right)_0\left({t_0\over t}\right)^{8/3}{\cal K} \,, \label{dnddotDel} \\ {\cal B}' &=& {\textstyle{4\over3}}\Delta' \,, \label{dn'} \\ {\cal K}' &=& {\textstyle{4\over3}} \left(c_{\rm a}^2\right)_0\left({t_0\over t}\right)^{5/3} {\cal B}\,, \label{dndotcB-dotcK} \end{eqnarray} where a prime denotes $d/d(t/t_0)$. Thus \begin{eqnarray} {\cal B}&=&{\textstyle{4\over3}}\left(\Delta+C_{{\cal B}}\right)\,, \label{dnc'} \\ {\cal K}'&=&{\textstyle{16\over9}} \left(c_{\rm a}^2\right)_0\left({t_0\over t}\right)^{5/3} \left[\Delta+C_{{\cal B}}\right] \,, \label{dncB-dotcK} \end{eqnarray} where $\dot{C}_{{\cal B}}=0$. We can now decouple the system. \subsubsection{Superhorizon scales} For long wavelength fluctuations, we get \begin{equation} 9\left({t\over t_0}\right)^3\Delta'''+36\left({t\over t_0}\right)^2 \Delta''+14\left({t\over t_0}\right)\Delta'- 4\Delta =0 \,, \label{ldndddotDel} \end{equation} to lowest order in $c_{\rm a}^2$. Note that curvature effects are quadratic in $c_{\rm a}^2$ and do not contribute at this level. In fact equations (\ref{dnddotDel}), (\ref{dndotcB-dotcK}) guarantee that, to lowest order in $c_{\rm a}^2$, curvature has no effect on magnetised disturbances in the dust distribution. We can solve Eq. (\ref{ldndddotDel}), which is of Euler-type: \begin{equation} \Delta=C_{(+)}\left({t\over t_0}\right)^{2/3}+C_{(1-)} \left({t\over t_0}\right)^{-1}+ C_{(2-)}\left({t\over t_0}\right)^{-2/3} \,. \label{lldnDel} \end{equation} Thus, the field has simply added the nonadiabatic decaying mode $C_{(2-)}$ to the evolution of superhorizon density perturbations, while the non-magnetized adiabatic modes are unchanged. The growth of large-scale matter aggregations proceeds virtually unaffected by the presence of the field or by curvature complexities. Magnetic effects on superhorizon scales in the dust era do not change the adiabatic growing mode to lowest order in $c_{\rm a}^2$. The adiabatic decaying mode is also unchanged (unlike the radiation case). However, a new nonadiabatic decaying mode arises, which decays less rapidly than the adiabatic mode. \subsubsection{Subhorizon scales} On subhorizon scales \begin{eqnarray*} 9\left({t\over t_0}\right)^3\Delta'''+36\left({t\over t_0}\right)^2 \Delta''+ 14\left({t\over t_0}\right)\left[1+{\textstyle{4\over21}} \left(c_{\rm a}^2\right)_0\left({k\over k_{{\rm h}0}}\right)^2 \right]\Delta'&& \nonumber\\ {}-4\left[1-{\textstyle{4\over9}}\left(c_{\rm a}^2\right)_0 \left({k\over k_{{\rm h}0}}\right)^2\right] \Delta =-{\textstyle{16\over9}}\left(c_{\rm a}^2\right)_0 \left({k\over k_{{\rm h}0}}\right)^2C_{{\cal B}} \,,~&& \end{eqnarray*} where again we have ignored terms of higher order in $(c_{\rm a}^2)_0$. The solution is \begin{eqnarray} \Delta&=&C_{(+)}\left({t\over t_0}\right)^{\alpha_+}+ C_{(-)}\left({t\over t_0}\right)^{\alpha_-}+ C_{({\cal B}-)}\left({t\over t_0}\right)^{-2/3}\nonumber\\ &&{}+ C_{\cal B}\left[\frac{4\left(c_{\rm a}^2\right)_0 k ^2} {4\left(c_{\rm a}^2\right)_0 k ^2-9k_{{\rm h}0}^2} \right] \,, \label{sdnDel} \end{eqnarray} with \begin{equation} \alpha_\pm={\textstyle{1\over6}}\left[-1\pm5\sqrt{1-{\textstyle{32\over75}} \left(c_{\rm a}^2\right)_0\left({k \over k_{{\rm h}0}}\right)^2 }\,\right] \,. \label{sdnz12} \end{equation} The magnetic influence is expressed in two ways: additional decaying ($C_{({\cal B}-)}$) and constant ($C_{{\cal B}}$) nonadiabatic modes; and modification of the non-magnetized adiabatic modes ($C_{(\pm)}$). The net effect is to inhibit the growth of matter aggregations, as noted also in the Newtonian case \cite{RR}. Note that the magnetic effects, direct or indirect, become less important after matter-radiation equality, due to the decrease of the Alfv\'{e}n speed. The damping of the growing mode is greater on smaller scales. Indeed there is a minimum scale, below which the solution in Eq. (\ref{sdnDel}) oscillates, since the magnetic pressure balances gravitational infall. This magnetic Jeans scale follows from Eq. (\ref{sdnz12}): \begin{equation} \lambda_{\rm mJ}(t_0)={\textstyle{4\over5}}\pi\sqrt{6}\lambda_{\rm a}(t_0)\,,\label{mj} \end{equation} where \ \lambda_{\rm a}=c_{\rm a}t={\textstyle{2\over3}}c_{\rm a}\lambda_{\rm h} \ is the Alfv\'{e}n horizon, with $\lambda_{\rm h}=H^{-1}$ the Hubble scale. In fact, given the weakness of the magnetic field, it is likely that kinetic pressure cannot be ignored near the magnetic Jeans scale. In this case, a more sophisticated analysis is necessary, to incorporate nonrelativistic pressure effects in baryonic matter. On scales well above the Alfv\'{e}n horizon (equivalently, magnetic Jeans scale) but well within the Hubble horizon, i.e., for \[ k_{{\rm h}0}\ll k\ll k_{{\rm a}0}~\mbox{ where }~k_{\rm a}={3\over2c_{\rm a}}k_{\rm h}\,, \] we find that the magnetized corrections $\alpha_\pm$ of the adiabatic exponents are \[ \alpha_\pm=\left\{\begin{array}{r} {\textstyle{2\over3}}-{\textstyle{2\over5}}(k/ k_{{\rm a}0})^2\,,\\ \\ -1+{\textstyle{2\over5}}(k/ k_{{\rm a}0})^2\,.\\ \end{array}\right. \] The way in which magnetic effects act to increase the adiabatic Jeans length may be qualitatively understood as follows. Consider a tube of magnetic force-lines with instantaneous cross-sectional area $\delta S$. In a perfectly conducting medium the field remains frozen into the fluid, i.e., the magnetic force-lines always connect the same particles \cite{E2}. More precisely, the induction equation (\ref{Max2}) shows that $a^3B^a$ is a connecting vector. Thus the volume of the tube is given by \[ \delta V=\delta\ell\,\delta S\propto a^3B\,\delta S\,. \] However, we also have that in general, $\delta V\propto a^3$. It follows that \[ \left(B\,\delta S \right)^{\displaystyle{\cdot}}=0\,. \] The conservation law in Eq. (\ref{B^2ev}) shows that $B\propto a^{-1}$; thus \[ \delta S\propto a^2\,, \] so that the cross section of the flux tube increases as the expansion redshifts the energy density of the field. Thus the field acts against gravitational infall. \subsection{Pure-magnetic and magnetized isocurvature perturbations} We have seen that the magnetic field introduces nonadiabatic modes in the density perturbations. This means that the field itself can generate fluctuations in the density, even when there are no primordial density fluctuations. Thus, if \begin{equation} \Delta(t_0)=0=\dot{\Delta}(t_0)\,, \label{pm}\end{equation} where $t_0$ is the epoch of magnetogenesis in the early radiation era, then nonzero $\Delta$ will arise purely from the magnetic field; in the absence of magnetogenesis, Eq. (\ref{pm}) would imply $\Delta=0$ for $t>t_0$. These nonadiabatic pure magnetic density perturbations can be found explicitly from the solutions given above. On superhorizon scales (assuming that the field is created on these scales at $t_0$), the general solution in Eq. (\ref{new}) implies with the initial conditions in Eq. (\ref{pm}) that the pure-magnetic nonadiabatic mode is (to lowest order in $c_{\rm a}^2$) \begin{eqnarray} \Delta_{\rm pm}& =& -{\textstyle{1\over3}}\left[C_{(0)}+C_{(2-)}\right] \left({a\over a_0}\right)^{2}\left[1+2 \left({a\over a_0}\right)^{-3+c_{\rm a}^2}\right] \nonumber\\ &&{}+C_{(0)}+C_{(2-)}\left({a\over a_0}\right)^{-c_{\rm a}^2}\,. \label{new'} \end{eqnarray} The pure-magnetic density perturbations have a dominant growing mode of the same strength as in the non-magnetized adiabatic case. The decaying modes are in fact the isocurvature part of the pure-magnetic density perturbations, as we now show. Equation (\ref{pm}) is often taken to characterize isocurvature perturbations, but it does so only in specific cases \cite{ent}. For magnetized perturbations, this is not the isocurvature condition. Isocurvature density perturbations are those for which the curvature perturbation of the initial hypersurface orthogonal to the fluid flow is spatially constant, i.e., $(a\mbox{D}_a{\cal K})(t_0)=0$. (There is also the implicit condition that $\omega_a=0$, which is necessary for the existence of the spatial hypersurface.) Taking the the comoving divergence of Eqs. (\ref{r}) and (\ref{dotDel_a}), we find the condition for magnetized isocurvature perturbations: \begin{equation} \dot{\Delta}+{\textstyle{3\over2}}(1-w)H\Delta={\textstyle{3\over4}}(1-w) c_{\rm a}^2H{\cal B}-c_{\rm a}^2H{\cal K}~\mbox{ at }~t=t_0\,.\label{iso1} \end{equation} In the non-magnetized case $c_{\rm a}=0$, it is clear that this condition is satisfied by Eq. (\ref{pm}), but when $c_{\rm a}>0$, then Eq. (\ref{pm}) does not characterize isocurvature perturbations. As an example, consider the implication of the magnetized isocurvature condition in the dust era, on scales well above the Alfv\'{e}n horizon but well within the Hubble horizon, i.e., $k_{{\rm h}0}\ll k\ll k_{{\rm a}0}$. Then Eqs. (\ref{iso1}) and (\ref{sdnDel}) give, to lowest order, \[ C_{(+)}={\textstyle{2\over5}}\left(c_{\rm a}^2\right)_0\left[C_{(-)} +C_{{\cal B}}\right]-{\textstyle{1\over5}}C_{{\cal K}}\,. \] On superhorizon scales, Eq. (\ref{iso1}) holds for all $t$ by virtue of Eq. (\ref{r'}), which implies $(a\mbox{D}_a{\cal K})^{\displaystyle{\cdot}}=0$. The magnetized isocurvature condition then selects a sub-class of the general superhorizon solutions found above. In the radiation era, we find (to lowest order in $c_{\rm a}^2$) \begin{equation} \Delta_{\rm iso}=C_{(1-)}\left(\frac{t}{t_0}\right)^{-\frac{1}{2}c_a^2}+ C_{(2-)}\left(\frac{t}{t_0}\right)^{-\frac{1}{2}+\frac{1}{2}c_a^2} \,. \label{iso4}\end{equation} Equation (\ref{iso4}) arises from the general superhorizon solution Eq. (\ref{new}) by eliminating the non-decaying modes. In the dust era, \begin{equation} \Delta_{\rm iso}=C_{(1-)}\left(\frac{t}{t_0}\right)^{-1}+ C_{(2-)}\left(\frac{t}{t_0}\right)^{-\frac{2}{3}} \,, \label{iso5}\end{equation} which is a special case of the general superhorizon dust solution Eq. (\ref{lldnDel}), once again without the growing mode. It is clear from Eqs. (\ref{iso4}) and (\ref{iso5}) that magnetized isocurvature perturbations on superhorizon scales are purely decaying. They are very different from the pure-magnetic solution Eq. (\ref{new'}) that is based on the initial conditions in Eq. (\ref{pm}). The latter has a constant and a growing mode. Magnetized nonadiabatic perturbations on superhorizon scales can contribute to the growing mode and generate a constant mode, whereas the magnetized isocurvature perturbations are purely decaying. \section{Magnetized cosmic vortices and Alfv\'{e}n waves} In the previous section, we generalized the results given in \cite{TB}, which itself provided a relativistic extension of previous work on magnetized density perturbations. A general inhomogeneous perturbation is characterized not only by its magnitude, i.e. the density perturbation $\Delta$, but also by its rotation and deformation properties, as described in general terms in \cite{EB2,MT}. Recently, these properties were investigated in CDM, and it was shown how the small stresses (isotropic and anisotropic) from residual velocity dispersion can have an important effect on rotation and deformation, even though the effect on density perturbations is effectively negligible \cite{mtm}. The evolution equations for rotational and deformation variables in an imperfect fluid were derived in \cite{MT}. The evolution equations for inhomogeneities were coupled to causal transport equations for viscosity and heat conduction. By Eq. (\ref{emt'}), a magnetized perfect fluid can be considered as an imperfect fluid with anisotropic stress, and the equations of \cite{MT} may be specialized to this case. However, the system needs to be completed by evolution equations for the magnetic stress, which are determined by Maxwell's equations. Here we investigate the coupled equations governing rotational and deformational inhomogeneity in the fluid and magnetic field. The comoving gradient of the density inhomogeneity $\Delta_a$ splits irreducibly as \ a\mbox{D}_b\Delta_a= {\textstyle{1\over3}}\Delta h_{ab}+ \varepsilon_{abc}W^c+ \xi_{ab} \,. \ The density perturbation is the comoving divergence, the rotational part is given by the comoving curl and the deformation part is the comoving PSTF derivative: \ \Delta=a\mbox{D}^a\Delta_a\,,~~ W_a=-{\textstyle{1\over2}}a\,{\rm curl}\,\Delta_{a}\,,~~ \xi_{ab}=a\mbox{D}_{\langle a}\Delta_{b\rangle}\,. \ The vector $W_a$ governs rotational instabilities in the density distribution of the matter, and ${\rm div}\, W=0$. On the other hand, $\xi_{ab}$ determines the volume-true anisotropic distortion, with $h^{ab}\xi_{ab}=0$. Both quantities describe differential, i.e., infinitesimal, properties. Here we focus on rotation, and in the next section we look at anisotropic deformations. A fundamental property of rotational perturbations is that they are proportional to the vorticity vector. This arises from the identity Eq. (\ref{id2}) which ensures that the curl of any gradient field, such as $\Delta_a$, derives from the vorticity. It follows that \begin{equation} W_a=-3a^2H(1+w)\omega_a\,. \label{w}\end{equation} Not only $W_a$, but also rotational perturbations in magnetic density and expansion inhomogeneities, are parallel to the vorticity: \[ {\rm curl}\,{\cal B}_a\equiv{\rm curl}\,\mbox{D}_aB^2={4\over3(1+w)}W_a\,,~ {\rm curl}\,\Theta_a\equiv{\rm curl}\,a\mbox{D}_a\Theta=-{\dot{H}\over(1+w)H}W_a\,. \] Thus all rotational instability in the magnetized medium arises from vorticity. As we have seen from the vorticity propagation equation (\ref{dotom}), magnetic inhomogeneities can source vorticity. Rewriting Eq. (\ref{dotom}) using Eq. (\ref{w}), we have \begin{equation} \dot{W}_a+{\textstyle{3\over2}}(1-w)HW_a=\left( \frac{3a^2H}{2\rho}\right)B^b\mbox{D}_b\mbox{curl}\,B_a \,. \label{dotW_a} \end{equation} Thus the field is a source of vorticity provided that its curl varies along its force lines. Furthermore, the effect of the field is to induce precession of the rotational vector $W_a$. In the absence of the field, $\dot{W}_a$ remains parallel to $W_a$, so that the initial direction is preserved along the fluid flow. By contrast, in the magnetized case, $\dot{W}_a$ is no longer parallel to $W_a$, and the initial direction changes along the fluid flow. The rate of precession is \begin{equation} \nu_{\rm prec}={|B^b\mbox{D}_b\mbox{curl}\,B_a|\over2(1+w)\rho|\omega_a|}\,. \label{prec} \end{equation} Equation (\ref{dotW_a}) shows how the magnetic field can become a source of density vortices. However, the field effect upon pre-existing rotational perturbations is not clear yet. To quantify the magnetic influence on $W_a$ we need to go one step further and obtain a decoupled equation for the evolution of $W_a$. We take the curl of Eq. (\ref{ddotDel_a}), using the above results and the identities in Eqs. (\ref{id2}), (\ref{id1}) and (\ref{id3}), and we arrive at the required evolution equation (with $\Lambda=0$): \begin{equation} \ddot{W}_a+\left(4-3w\right)H\dot{W}_a+ {\textstyle{1\over2}}\left[1-7w+3c_{\rm s}^2(1+w)\right]\rho W_a= \left[\frac{c_{\rm a}^2}{3(1+w)}\right]\mbox{D}^2W_a \,. \label{ddotW_a} \end{equation} This is a wave equation for $W_a$, with signal speed $v_{\rm a}$ given by\footnote{ A similar equation was derived in \cite{MT} for a fluid with shear viscosity; in that case $v^2=\eta/[\tau\rho(1+w)]$, where $\eta$ is the viscosity and $\tau$ is the causal relaxation time. } \[ v_{\rm a}^2={c_{\rm a}^2\over 3(1+w)}\,. \] Propagating solutions of this equation are Alfv\'{e}n waves (compare \cite{cmb}), i.e., incompressible, vector waves, as opposed to the compressible, scalar magneto-sonic waves. In the non-magnetized case, the signal speed vanishes, and no wave solutions exist. We note also that the only scale-dependence arising in the wave equation is via the magnetic $c_{\rm a}^2$ term. Thus magnetized vortices are scale-dependent, unlike the non-magnetized case. Decomposing the solenoidal vector $W_a$ into Fourier modes ${\cal W}$, Eq. (\ref{ddotW_a}) gives \begin{equation} \ddot{\cal W}+\left(4-3w\right)H\dot{\cal W}+ \left\{{\textstyle{1\over2}} \left[1-7w+3c_{\rm s}^2(1+w)\right]\rho+ \frac{ c_{\rm a}^2k^2}{3(1+w)a^2}\right\}{\cal W}=0\,. \label{nddotW_a} \end{equation} Clearly, $W_a$ can only grow if the term in square brackets becomes negative. If $\dot{w}=0$, as in the radiation and dust eras, then the quantity $1-7w+3(1+w)c_{\rm s}^2=(3w-1)(w-1)$ becomes negative when ${1\over3}<w<1$. Thus it is only when matter stiffer than radiation dominates the universe (and is coupled to the magnetic field), that vortices in the density distribution can grow (in the linear regime). Furthermore, the presence of the magnetic field ensures that such growth occurs only on scales larger than the critical ``rotational Jeans" wavelength \ \lambda_{\rm rJ}=2\pi c_{\rm a}\left[{\frac{2}{3(3w-1)(1-w^2)\rho}} \right]^{1/2} \,, \ which is small due to the weakness of the field. In the radiation era, Eq. (\ref{nddotW_a}) becomes \ {\cal W}''+{2}\left({a_0\over a}\right){\cal W}'+ \left({k\over k_{{\rm a}0}}\right)^2{\cal W}=0 \,, \ where a prime denotes $d/d(a/a_0)$, and $k_{\rm a}=2k_{\rm h}/c_{\rm a}$ is the wavenumber of the Alfv\'{e}n horizon $\lambda_{\rm a}={1\over2}c_{\rm a}\lambda_{\rm h}$. This has the general solution \begin{equation} {\cal W}=\frac{a_0}{a} \left[C_{(1)}\cos\left( {k\over k_{{\rm a}0}}\frac{a}{a_0}\right) +C_{(2)}\sin\left( {k\over k_{{\rm a}0}}{a\over a_0}\right)\right] \,, \label{rW_a} \end{equation} which describes Alfv\'{e}n waves. The Alfv\'{e}n frequency \begin{equation} \nu_{\rm a}={H_0\over\pi}\left({k\over k_{{\rm a}0}}\right)^2={\textstyle{1\over2}}\delta\nu_{\rm ac}\,, \label{osc}\end{equation} where $\delta\nu_{\rm ac}=\nu_{\rm ac,mag}-\nu_{\rm ac}$ is the excess magnetic acoustic frequency given in Eq. (\ref{3}). Thus local differential vortices in the density distribution are ``flip-flopping" in concert with the acoustic oscillations in the density perturbations. Alfv\'{e}n waves are a purely magnetic effect, arising from the fluctuations in the magnetic field direction. These waves have decaying amplitude, in common with non-magnetized (and non-propagating) vector perturbations. On scales well beyond the Alfv\'{e}n horizon, $k\ll k_{{\rm a}0}$, the oscillatory behavior is not felt, and Eq. (\ref{rW_a}) gives, to lowest order, \[ {\cal W}=C_{(1)}\left[{a_0\over a}-{\textstyle{1\over2}}\left({k\over k_{{\rm a}0}} \right)^2\left({a\over a_0}\right)\right]+C_{(2)} \left({ k\over k_{{\rm a}0}}\right)\,. \] On superhorizon scales, the oscillations disappear: ${\cal W}\rightarrow C_{(1)}(a_0/a)$, regaining the standard non-magnetized result. Thus, before matter-radiation equality, and on scales much larger than the Alfv\'{e}n horizon, density vortices evolve unaffected by the presence of a cosmological magnetic field. After equality, in the matter-dominated dust era, Eq. (\ref{nddotW_a}) becomes \begin{equation} 3\left(\frac{t}{t_0}\right)^2{\cal W}''+8\left({t\over t_0}\right) {\cal W}'+\left[2+ \left({k\over k_{{\rm a}0}}\right)^2\right]{\cal W}=0 \,, \label{dnddW_a} \end{equation} where a prime denotes $d/d(t/t_0)$. The solution is \begin{equation} {\cal W}=C_{(+)}\left({t\over t_0}\right)^{\alpha_+}+ C_{(-)}\left({t\over t_0}\right)^{\alpha_-}\,, \label{dW_a} \end{equation} where \begin{equation} \alpha_\pm={\textstyle{1\over6}}\left[-5 \pm\sqrt{1-12 \left( {k\over k_{{\rm a}0}}\right)^2}\,\right] \,. \label{dz12} \end{equation} Therefore, any rotational instabilities present in the density distribution of the dust die away with time, as they do in non-magnetized cosmologies. The field effect on a given mode $k$ is to reduce the depletion rate of $W_a$ by an amount proportional to the initial Alfv\'{e}n speed squared, $(c_{\rm a}^2)_0$. Thus, magnetized dust universes will contain more residual vortices than magnetic-free ones. However, the effect is confined within a narrow wavelength band beyond the Alfv\'{e}n horizon $\lambda_{\rm a}$. On much larger scales, the field influence becomes negligible, and $W_a\propto t_0/t$ as in non-magnetized models. On scales with $k>k_{{\rm a}0}/\sqrt{12}$, i.e. within a few times the Alfv\'{e}n scale, Eq. (\ref{dz12}) shows that the density vortices oscillate as Alfv\'{e}n waves. \section{Magnetized shape-distortion} We monitor anisotropic deformation (shape distortion) in the density distribution of the medium through the PSTF tensor $\xi_{ab}=a\mbox{D}_{\langle a}\Delta_{b\rangle}$. This is associated with density variations that do not represent matter aggregations, since the associated divergence of $\Delta_a$ is zero, but rather describe changes in the local anisotropy pattern of the density gradients. Distortion in the density is coupled to distortion in the expansion and the magnetic energy density, defined via the PSTF tensors \begin{equation} \vartheta_{ab}=a^2\mbox{D}_{\langle a}\mbox{D}_{b\rangle}\Theta\,,~~ \beta_{ab}= \frac{a^2}{B^2}\mbox{D}_{\langle a}\mbox{D}_{b\rangle}B^2 \,, \label{vthe_ab-bet_ab} \end{equation} which vanish in the background and are thus gauge-invariant. The propagation equations for $\xi_{ab}$, $\vartheta_{ab}$ and $\beta_{ab}$ follow from the comoving PSTF derivatives of Eqs. (\ref{dotDel_a})--(\ref{dotB_ab}): \begin{eqnarray} \dot{\xi}_{ab}&=& 3wH\xi_{ab}- (1+w)\vartheta_{ab}+ {\textstyle{3\over2}}c_{\rm a}^2H\beta_{ab}- c_{\rm a}^2H\kappa_{ab}- 3H\mu_{ab} \,, \label{dotxi}\\ \dot{\vartheta}_{ab}&=&-2H\vartheta_{ab}- {\textstyle{1\over2}}{\rho}\xi_{ab}- \frac{c_{\rm s}^2}{1+w}\mbox{D}^2\xi_{ab}+ {\textstyle{1\over4}}{c_{\rm a}^2\rho}\beta_{ab}- \frac{c_{\rm a}^2}{2(1+w)}\mbox{D}^2\beta_{ab}\nonumber\\ &&{}-{\textstyle{1\over2}}{c_{\rm a}^2\rho}\kappa_{ab}- {\textstyle{3\over2}}\rho\mu_{ab}- \left(6c_{\rm s}^2+\frac{4c_{\rm a}^2}{1+w}\right)H\varpi_{ab} \,, \label{dotvth}\\ \dot{\beta}_{ab}&=&\frac{4}{3(1+w)}\dot{\xi}_{ab}+ \frac{4(c_{\rm s}^2-w)H}{1+w}\xi_{ab} \,. \label{dotbet} \end{eqnarray} The additional gauge-invariant PSTF tensors \ \kappa_{ab}=a^2{\cal R}_{\langle ab\rangle}\,,~~ \varpi_{ab}=a^2\mbox{D}_{\langle a}\mbox{curl}\,\omega_{b\rangle}\,,~~ \mu_{ab}={a\over\rho}B^c\mbox{D}_cB_{\langle ab\rangle} \,, \ respectively describe distortions caused by projected curvature, rotation and by anisotropies in the distribution of the magnetic field gradients. The first is due to the natural coupling of the field to the curvature and is given by \ \kappa_{ab}=a^2\left({\textstyle{1\over2}}\pi_{ab}- H\sigma_{ab}+ E_{ab}\right)\,, \ obtained from Eq. (\ref{3R_ab}) by means of the shear propagation equation (\ref{dotsh}). The second arises from the fluid flow, which generally is not hypersurface orthogonal. It has no impact on deformation if the rather special condition $\mbox{D}_{\langle a}\mbox{curl}\,\omega_{b\rangle}=0$ holds. Finally, the effect of $\mu_{ab}$ vanishes when any anisotropies present in the distribution of $B_{ab}\equiv a\mbox{D}_bB_a$ remain invariant along the magnetic force-lines, that is when $B^c\mbox{D}_cB_{\langle ab\rangle}=0$. Note that both the scalar and vector aspects of the field contribute to shape distortion. Equations (\ref{dotxi}) and (\ref{dotvth}) combine to provide a second order differential equation, also obtained by taking the comoving PSTF derivative of Eq. (\ref{ddotDel_a}). With $\Lambda=0$, we have \begin{eqnarray} \ddot{\xi}_{ab}&=&-\left(2+3c_{\rm s}^2-6w\right)H\dot{\xi}_{ab}+ {\textstyle{1\over2}} \left(1-6c_{\rm s}^2+8w-3w^2\right)\rho\xi_{ab}+ c_{\rm s}^2\mbox{D}^2\xi_{ab} \nonumber\\&\mbox{}&-{\textstyle{1\over2}} \left(1-3c_{\rm s}^2+2w\right)c_{\rm a}^2\rho\beta_{ab}+ {\textstyle{1\over2}}c_{\rm a}^2\mbox{D}^2\beta_{ab}+{\textstyle{1\over3}} \left(2-3c_{\rm s}^2+3w\right)\left(c_{\rm a}^2\rho\kappa_{ab}+3\rho\mu_{ab}\right) \nonumber\\&\mbox{}&+ \left[6(1+w)c_{\rm s}^2+2c_{\rm a}^2\right]H\varpi_{ab} \,. \label{ddotxi_ab} \end{eqnarray} This is coupled to Eq. (\ref{dotbet}) for the growth of infinitesimal distortions in the magnetic energy density. The other source terms, namely $\kappa_{ab}$, $\mu_{ab}$ and $\varpi_{ab}$, evolve according to \begin{eqnarray} \dot{\kappa}_{ab}&=&-\frac{1}{1+w}\dot{\xi}_{ab}+ \frac{(c_{\rm s}^2+3w)H}{1+w}\xi_{ab}+ \frac{2c_{\rm a}^2H}{1+w}\beta_{ab}- \frac{4H}{1+w}\mu_{ab}-2\varpi_{ab}- \mbox{D}^2\sigma_{ab} \,, \label{dotkap}\\ \dot{\mu}_{ab}&=&-\left[4-3(1+w)\right]H\mu_{ab}+ \frac{c_{\rm a}^2}{3(1+w)}\dot{\xi}_{ab}+ \frac{(c_{\rm s}^2-w)c_{\rm a}^2H}{1+w}\xi_{ab} \nonumber\\ &&{}- \frac{c_{\rm a}^4H}{2(1+w)}\beta_{ab}+ \frac{c_{\rm a}^4H}{3(1+w)}\kappa_{ab}+ {\textstyle{1\over3}}c_{\rm a}^2a^2\mbox{D}^2\sigma_{ab} \,, \label{dotmu}\\ \dot{\varpi}_{ab}&=&-(2-3c_{\rm s}^2)H\varpi_{ab}+ \frac{a}{2(1+w)\rho}\mbox{D}_c\mbox{D}_{\langle a}B^d\mbox{D}_{|d|}\left[B_{b\rangle}{}^c-B^c{}_{b\rangle}\right]\,. \label{dotvpi} \end{eqnarray} We used the propagation equations (\ref{gem1}) and (\ref{pi}) for $E_{ab}$ and $\pi_{ab}$, which imply \[ \dot{E}_{ab}=-3HE_{ab}+ {\textstyle{3\over2}}H\pi_{ab}- {\textstyle{1\over2}}(1+w)\rho\sigma_{ab}- \mbox{D}^2\sigma_{ab}+ \frac{1}{a^2}\left(\vartheta_{ab}+2\varpi_{ab}\right) \,, \] on using the constraint equations (\ref{shcon}) and (\ref{Hcon}). Equation (\ref{dotmu}) requires Eqs. (\ref{Max2}) and (\ref{dotxi}). Finally, to obtain the evolution formula of $\varpi_{ab}$ we have successively taken the comoving curl and the comoving PSTF derivative of Eq. (\ref{dotom}). The system of equations (\ref{dotbet})--(\ref{dotvpi}) provides in principle a complete description of linear infinitesimal shape distortion generated by magnetic effects, provided we have a prescription for the $\mbox{D}^2\sigma_{ab}$ terms in Eqs. (\ref{dotkap}) and (\ref{dotmu}), and for the last term on the right of Eq. (\ref{dotvpi}). Even without these terms, the system is too complicated to analyze in general. However, it is clear in general terms how magnetic effects will actively generate distortion. We can illustrate this by comparing with the non-magnetized case in a simple example. For simplicity, consider superhorizon scales in the dust era (neglecting vorticity). Suppose that at a given event $(t_0,\vec{x}_0)$, we have no initial distortion or rate of distortion: \begin{equation} (\xi_{ab})_0=0=(\dot{\xi}_{ab})_0\,, \label{d1}\end{equation} In the non-magnetized case, the distortion system collapses to the single equation \[ \ddot{\xi}_{ab}=-2H\dot{\xi}_{ab}+ {\textstyle{3\over2}}H^2\xi_{ab}\,. \] It follows that along the fluid flow line through $\vec{x}_0$, no distortion is generated: \[ \xi_{ab}(t,\vec{x}_0)=0\,. \] The evolution is purely {\em passive}, or inertial, i.e., distortion can only develop if it is there {\em a priori}. In the magnetized case, by contrast, distortion is {\em actively} and nonadiabatically generated by magnetic effects. Equation (\ref{ddotxi_ab}) shows that \[ H_0^{-2}(\ddot{\xi}_{ab})_0=\left(c_{\rm a}^2\right)_0\left[-{\textstyle{3\over2}}\left(\kappa_{ab}\right)_0+ 2\left(\kappa_{ab}\right)_0\right]+6\left(\mu_{ab}\right)_0\,, \] so that $(\ddot{\xi}_{ab})_0\neq0$. Distortion is immediately generated along the flow line. In fact, the distortion has a growing mode, as we now show. The superhorizon scalar modes of the distortion tensors satisfy a system that follows from Eqs. (\ref{dotbet}) and (\ref{ddotxi_ab})--(\ref{dotmu}): \begin{eqnarray} \xi''&=&-{\textstyle{4\over3}}\left({t_0\over t}\right)\xi'+ {\textstyle{2\over3}}\left({t_0\over t}\right)^2\xi- {\textstyle{2\over9}}\left(c_{\rm a}^2\right)_0\left({t_0\over t}\right)^{8/3} [3\beta-4\kappa]+ {\textstyle{8\over3}}\left({t_0\over t}\right)^2\mu \,, \label{dnddotxi}\\ \beta'&=&{\textstyle{4\over3}}\xi' \,, \label{dndotbet}\\ \kappa'&=&-\xi'+{\textstyle{4\over3}}\left(c_{\rm a}^2\right)_0 \left({t_0\over t}\right)^{5/3}\beta- {\textstyle{8\over3}}\left({t_0\over t}\right)\mu \,, \label{dndotkap}\\ \mu'&=&-{\textstyle{2\over3}}\left({t_0\over t}\right)\mu+{\textstyle{1\over3}}\left(c_{\rm a}^2\right)_0\left({t_0\over t}\right)^{2/3}\xi'\,. \label{dndotmu} \end{eqnarray} Equations (\ref{dndotbet}) and (\ref{dndotmu}) integrate to \begin{equation} \beta={\textstyle{4\over3}}\left(\xi+\Gamma_\beta\right)\,,~~ \mu={\textstyle{4\over3}}\left(c_{\rm a}^2\right)_0 \left({t_0\over t}\right)^{2/3} \left[\xi+\Gamma_\mu\right] \,, \label{dnbetmu} \end{equation} where $\dot{\Gamma}_\beta=0= \dot{\Gamma}_\mu$. Then Eq. (\ref{dnbetmu}) transforms Eq. (\ref{dndotkap}) into \begin{equation} \kappa'=-\xi'+{\textstyle{8\over9}}\left(c_{\rm a}^2\right)_0 \left({t_0\over t}\right)^{2/3}\left[\xi+2\Gamma_\beta-\Gamma_\mu\right]\,. \label{dndotkap1} \end{equation} According to Eq. (\ref{dnbetmu}), the effect of any anisotropies present in the distribution of $B_{ab}\equiv a\mbox{D}_aB_b$ on $\xi$ decreases after matter-radiation equilibrium. Since $\mu$ is a key source of shape-distortion, we expect the evolution of $\xi$ to approach that of $\Delta$ as the universe expands. Equation (\ref{dnddotxi}) gives \[ 9\left({t\over t_0}\right)^3\xi'''+36\left({t\over t_0}\right)^2\xi''+ 14\left({t\over t_0}\right)\xi'- 4\xi=0 \,, \] on using Eqs. (\ref{dnbetmu}) and (\ref{dndotkap1}). This has the same form as the corresponding density perturbation equation (\ref{ldndddotDel}), and the solution is thus of the form Eq. (\ref{lldnDel}). Imposing the initial conditions in Eq. (\ref{d1}), we find that \begin{equation} \xi(t,\vec{x}_0)=\Gamma\left[\left({t\over t_0}\right)^{2/3}+4\left({t\over t_0} \right)^{-1}-5\left({t\over t_0}\right)^{-2/3}\right] \,, \label{lldnxi} \end{equation} where $\Gamma$ is a constant. Thus the shape distortion has a growing mode along the fluid flow line, due purely to magnetic effects; in the absence of the magnetic field, $\Gamma=0$. (A similar situation arises in the simpler case of distortion generated by velocity dispersion \cite{mtm}.) \section{Conclusion} We have given a fully general relativistic treatment of the scalar and vector effects of a weak large-scale magnetic field on cosmological density inhomogeneity. This refines the results of \cite{TB} on magnetized density perturbations, and extends that work to analyze magnetized vortices and shape distortion in the density distribution. Our covariant Lagrangian approach allows us to derive gauge-invariant evolution equations for all these aspects of density inhomogeneity in the general linear case, i.e., incorporating all fluctuations of, and couplings between, the field, the fluid and the curvature. In summary, magnetized density perturbations are governed by Eqs. (\ref{ddotDel})--(\ref{dotcK}), magnetized density vortices are governed by Eq. (\ref{ddotW_a}), and magnetized shape distortion is governed by Eqs. (\ref{dotbet})--(\ref{dotvpi}). We give the solutions in closed form for magnetized density perturbations and vortices, in the radiation and dust eras. Some of the scalar solutions and all of the vector solutions are new. For magnetized shape distortion, we found a special solution with a growing mode. Given the overall weakness of the field, the magnetic effects described here are secondary relative to those of the matter. In some cases, second-order fluid effects may be comparable in strength to first-order magnetic effects. However, their presence does not affect the field impact, which remains the same. Thus we can neglect the second-order effects in a consistent linear analysis that probes the lowest order magnetic effect on density inhomogeneity. To lowest order, the magnetic influence on gravitational instability may be summarised as follows: \begin{enumerate} \item The magneto-curvature coupling, which is a direct consequence of the field's vectorial nature, first identified in \cite{TB}, has an important influence. In particular, on superhorizon scales in the radiation era, this coupling slightly enhances the growing mode of density perturbations, as shown in Eqs. (\ref{lrnDel}) and (\ref{nc}): \[ \Delta= C_{(+)}\left({a\over a_0}\right)^{2+{1\over2}c_{\rm a}^4} +C_{(1-)}\left({a\over a_0}\right)^{-1+c_{\rm a}^2}+C_{(0)}+ C_{(2-)}\left({a\over a_0}\right)^{-c_{\rm a}^2}\,. \] When the coupling is neglected, the growing mode is incorrectly found to be damped relative to the non-magnetized case. \item Magneto-sonic waves in the radiation era are given in exact form in Eq. (\ref{srnDel}), \begin{eqnarray*} \Delta&=&\left[C_{(1)}-C_{{\cal K}} {\rm Si}\left(\beta {k\over k_{{\rm h}0}}{a\over a_0}\right) \right]\sin\left(\beta {k\over k_{{\rm h}0}}{a\over a_0}\right)\nonumber\\ &&{}+\left[C_{(2)}-C_{{\cal K}} {\rm Ci}\left(\beta {k\over k_{{\rm h}0}}{a\over a_0}\right) \right]\cos\left(\beta {k\over k_{{\rm h}0}}{a\over a_0}\right) -C_{{\cal B}}c_{\rm a}^2\,. \end{eqnarray*} This shows the nonadiabatic modulation of the amplitude and increase in the frequency of acoustic oscillations. These effects, together with the nonzero average value implied by the $C_{{\cal B}}$ term, have potentially important implications for the CMB acoustic peaks, some of which have been investigated in \cite{adgr}. \item In the dust era, subhorizon magnetized density perturbations are given exactly in Eq. (\ref{sdnDel}), which leads to the magnetized Jeans scale in Eq. (\ref{mj}): \[ \lambda_{\rm mJ}(t_0)={\textstyle{8\over15}}\pi\sqrt{6} \left(c_{\rm a}\right)_0\lambda_{{\rm h}0}\,. \] On scales such that $\lambda_{\rm mJ}\ll\lambda\ll\lambda_{\rm h}$, the density perturbations are \[ \Delta=C_{(+)}\left({t\over t_0}\right)^{{2\over3}-\epsilon}+ C_{(-)}\left({t\over t_0}\right)^{-1+\epsilon}+ C_{({\cal B}-)}\left({t\over t_0}\right)^{-2/3} -{\textstyle{5\over2}}\epsilon C_{\cal B}\,, \] where $\epsilon={2\over5}(k/k_{{\rm a}0})^2$. This shows the small damping effect on the adiabatic growing mode, as well as the new nonadiabatic modes. These results imply small modifications to structure formation in the linear regime. However, they are limited by the fact that we have neglected any non-baryonic matter or cosmological term. \item Pure-magnetic density fluctuations, which are induced in an initially smooth fluid by magnetogenesis, are given on superhorizon scales by Eq. (\ref{new'}). This solution would be important in any attempt to model large-scale structure formation as seeded by magnetogenesis. \item Magnetized isocurvature perturbations are characterized by Eq. (\ref{iso1}). On superhorizon scales, these modes are purely decaying. \item The field is a source of incompressible rotational instabilities, and the condition for this to happen is given via Eq. (\ref{dotW_a}). Magnetized density vortices are shown be scale-dependent and to precess, at a rate given by Eq. (\ref{prec}). The general propagation equation for these vortices (i.e., incorporating all relevant effects) is given by Eq. (\ref{ddotW_a}): \[ \ddot{W}_a+\left(4-3w\right)H\dot{W}_a+ {\textstyle{1\over2}}\left[1-7w+3c_{\rm s}^2(1+w)\right]\rho W_a= \left[\frac{c_{\rm a}^2}{3(1+w)}\right]\mbox{D}^2W_a \,. \] In the radiation era, the Alfv\'{e}n wave solutions are given exactly in Eq. (\ref{rW_a}). The Alfv\'{e}n frequency and wave-speed are \[ \nu_{\rm a}={H_0\over\pi}\left({k\over k_{{\rm a}0}}\right)^2\,,~~ v_{\rm a}={\textstyle{1\over2}}c_{\rm a}\,. \] These results generalize some of the theoretical results of Durrer et al. \cite{dky}, who then go further and apply the results to determine the effect of Alfv\'{e}n wave modes on CMB anisotropies. \item After recombination, magnetized density vortices are given exactly in Eq. (\ref{dW_a}). They decay like their adiabatic counterparts, but at a slower rate, so that rotational instability persists for longer in a magnetic universe. This will have a small effect on structure formation in the linear regime. \item Finally, we have investigated for the first time magnetic effects on infinitesimal shape distortion in the density distribution. The magnetic influence is manifold. Anisotropies in the field energy density, together with those in the distribution of the magnetic vector itself are direct sources of density deformation. The field's coupling to curvature and rotation also acts as an indirect source of magnetically induced shape distortions. Following the evolution of shape-distortion along the worldline of a fluid element, we showed that the field is an {\em active} source of distortion. On superhorizon scales, we showed via a special solution of the shape-distortion system that there is a growing mode of magnetized shape-distortion [see Eq. (\ref{lldnxi})]. Unlike the magnetic effects on density and rotational perturbations, which are small corrections of the non-magnetized results, magnetic effects on shape-distortion constitute a significant change from the non-magnetized (and {\em passive}) case. (A similar statement applies in the case of velocity dispersion effects in CDM \cite{mtm}.) The results on magnetized shape-distortion have potentially important implications for (linear) structure formation. Not only is distortion actively generated once scales re-enter the Hubble horizon and begin to collapse, but it is also actively generated while the scales are beyond the horizon. Of course, the shape distortion in the linear regime will be overwhelmed by effects that arise during the nonlinear stages of collapse. \end{enumerate} \[ \] {\bf Acknowledgments:} CGT is supported by PPARC. We thank Marco Bruni and David Matravers for useful discussions.
\section{Introduction} The concept of deformation quantization of a symmetric manifold $M$ has been defined by Bayen, Flato, Fronsdal, Lichnerovich, and Sternheimer in \cite{BFFLS}. Deformation quantization means a formal $*$-product $$f_1*f_2=f_1 \cdot f_2+\sum_{k=1}^\infty C_k(f_1,f_2)t^k,\qquad f_1,f_2 \in C^\infty(M)$$ with some additional properties, where $C_k:C^\infty(M)\times C^\infty(M)\to C^\infty(M)$ are bidifferential operators. In the special case of the unit disc in ${\mathbb C}$ with $SU_{1,1}$-invariant symplectic structure, a formal $*$-product and explicit fopmulae for $C_k$, $k \in{\mathbb N}$, are derivable by a method of Berezin \cite{CGR, B}. Our intention is to replace the ordinary disc with its q-analogue. We are going to produce $U_q \mathfrak{su}_{11}$-invariant formal deformation of our quantum disc and to obtain an explicit formula for $C_k$, $k \in{\mathbb N}$, using a q-analogue of the Berezin method \cite{SSV5}. Our work is closely related to the paper of Klimek and Lesniewski \cite{KL} on two-parameter deformation of the unit disc. The explicit formulae for $C_k$ we provide below work as a natural complement to the results of this paper. \bigskip \section{Covariant symbols of linear operators} Everywhere in the sequel the field of complex numbers ${\mathbb C}$ is assumed as a ground field. Let also $q \in(0,1)$. Consider the well known algebra ${\rm Pol}({\mathbb C})_q$ with two generators $z,z^*$ and a single commutation relation $z^*z=q^2zz^*+1-q^2$. Our intention is to produce a formal $*$-product \begin{equation}f_1*f_2=f_1 \cdot f_2+\sum_{k=1}^\infty C_k(f_1,f_2)t^k,\qquad f_1,f_2 \in{\rm Pol}({\mathbb C})_q, \end{equation} (with some remarkable properties) to be given by explicit formulae for bilinear operators $C_k:{\rm Pol}({\mathbb C})_q \times{\rm Pol}({\mathbb C})_q \to{\rm Pol}({\mathbb C})_q$. We describe in this section the method of producing this $*$-product whose idea is due to F. Berezin. It was explained in \cite{SSV1} that the vector space $D({\mathbb U})_q'$ of formal series $\displaystyle \sum_{j,k=0}^\infty a_{jk}z^jz^{*k}$ with complex coefficients is a q-analogue of the space of distributions in the unit disc ${\mathbb U}=\{z \in{\mathbb C}|\;|z|<1 \}$. Equip this space of formal series with the topology of coefficientwise convergence. Since $\{z^jz^{*k}\}_{j,k \in{\mathbb Z}_+}$ constitute a basis in the vector space ${\rm Pol}({\mathbb C})_q$, ${\rm Pol}({\mathbb C})_q$ admits an embedding into $D({\mathbb U})_q'$ as a dense linear subvariety. Consider the unital subalgebra ${\mathbb C}[z]_q \subset{\rm Pol}({\mathbb C})_q$ generated by $z \in{\rm Pol}({\mathbb C})_q$. Let $\alpha>0$. We follow \cite{KL} in equipping the vector space ${\mathbb C}[z]_q$ with the scalar product $(z^j,z^k)_\alpha=\delta_{jk}\dfrac{(q^2;q^2)_k}{(q^{4 \alpha+2};q^2)_k}$, $j,k \in{\mathbb Z}_+$, where $(a;q^2)_k=(1-a)(1-q^2a)\ldots(1-q^{2(k-1)}a)$. Let $L_a^2(d \nu_ \alpha)_q$ be the a completion of ${\mathbb C}[z]_q$ with respect to the norm $\| \psi \|_\alpha=(\psi,\psi)_\alpha^{1/2}$. It was demonstrated in \cite{KL} that the Hilbert space $L_a^2(d \nu_ \alpha)_q$ is a q-analogue of the weighted Bergman space. Let $\widehat{z}$ be a linear operator of multiplication by $z$: $$\widehat{z}:L_a^2(d \nu_ \alpha)_q \to L_a^2(d \nu_ \alpha)_q;\qquad \widehat{z}:\psi(z)\mapsto z \cdot \psi(z),$$ and denote by $\widehat{z}^*$ the adjoint operator in $L_a^2(d \nu_ \alpha)_q$ to $\widehat{z}$. The definition of the scalar product in $L_a^2(d \nu_ \alpha)_q$ implies that the operators $\widehat{z}$, $\widehat{z}^*$ are bounded. Equip the space ${\cal L}_\alpha$ of bounded linear operators in $L_a^2(d \nu_ \alpha)_q$ with the weakest topology in which all the linear functionals $$l_{\psi_1,\psi_2}:{\cal L}_\alpha \to{\mathbb C},\qquad l_{\psi_1,\psi_2}:A \mapsto(A \psi_1,\psi_2)_\alpha,\qquad \psi_1,\psi_2 \in{\mathbb C}[z]_q$$ are continuous. The following proposition is a straightforward consequence of the definitions (see the proof in \cite{SSV1}). \medskip \begin{proposition} Given any bounded linear operator $\widehat{f}$ in the Hilbert space $L_a^2(d \nu_ \alpha)_q$, there exists a unique formal series $f=\sum \limits_{j,k=0}^\infty a_{jk}z^jz^{*k}\in D({\mathbb U})_q'$ such that $\widehat{f}=\sum \limits_{j,k=0}^\infty a_{jk}\widehat{z}^j \widehat{z}^{*k}$. \end{proposition} \medskip Thus we get an injective linear map ${\cal L}_\alpha \to D({\mathbb U})_q'$, $\widehat{f}\mapsto f$. The distribution $f$ is called a {\sl covariant symbol} of the linear operator $\widehat{f}$. \medskip {\sc Remark 2.2.} For an arbitrary $f \in{\rm Pol}({\mathbb C})_q$, there exists a unique operator $\widehat{f}\in{\cal L}_\alpha$ with the covariant symbol $f$. Specifically, for $f=\sum \limits_{j,k=0}^{N(f)}a_{jk}z^jz^{*k}$, one has $\widehat{f}=\sum \limits_{j,k=0}^{N(f)}a_{jk}\widehat{z}^j \widehat{z}^{*k}$. \medskip We follow F. Berezin in producing the $*$-product of covariant symbols using the ordinary product of the associated linear operators. Let $f_1,f_2 \in{\rm Pol}({\mathbb C})_q$ and $\widehat{f}_1,\widehat{f}_2 \in{\cal L}_\alpha$ be the operators whose covariant symbols are $f_1$, $f_2$. Under the notation $t=q^{4 \alpha}$, let $m_t(f_1,f_2)$ stand for the covariant symbol of the product $\widehat{f}_1 \cdot \widehat{f}_2$ of the linear maps $\widehat{f}_1$, $\widehat{f}_2$. Evidently, we have constructed a bilinear map $m_t:{\rm Pol}({\mathbb C})_q \times{\rm Pol}({\mathbb C})_q \to D({\mathbb U})_q'$. The $*$-product $f_1*f_2$ of $f_1,f_2 \in{\rm Pol}({\mathbb C})_q$ is to be introduced by replacement of the one-parameter family of distributions $m_t(f_1,f_2)$, $t \in(0,1)$, with its asymptotic expansion as $t \to 0$. \bigskip \section{\boldmath$*$-Product} The term {\sl 'order one differential calculus over the algebra ${\rm Pol}({\mathbb C})_q$'} stand for a ${\rm Pol}({\mathbb C})_q$-bimodule $\Omega^1({\mathbb C})_q$ equipped with a linear map $d:{\rm Pol}({\mathbb C})_q \to \Omega^1({\mathbb C})_q$ such that \\ i) $d$ satisfies the Leibniz rule $d(f_1f_2)=df_1 \cdot f_2+f_1 \cdot df_2$ for any $f_1,f_2 \in{\rm Pol}({\mathbb C})_q$,\\ ii) $\Omega^1({\mathbb C})_q$ is a linear span of $f_1 \cdot df_2 \cdot f_3$, $f_1,f_2,f_3 \in{\rm Pol}({\mathbb C})_q$ (see \cite{KS}). One can find in \cite{SV} a construction of that kind of order one differential calculus for a wide class of prehomogeneous vector spaces $V$. In the case $V={\mathbb C}$ we deal with this calculus is well known; it can be described in terms of the following commutation relations: $$z \cdot dz=q^{-2}dz \cdot z,\qquad z^*dz^*=q^2dz^*z^*,\qquad z^*dz=q^2dz \cdot z^*,\qquad z \cdot dz^*=q^{-2}dz^*z.$$ The partial derivatives $\dfrac{\partial^{(r)}}{\partial z}$, $\dfrac{\partial^{(r)}}{\partial z^*}$, $\dfrac{\partial^{(l)}}{\partial z}$, $\dfrac{\partial^{(l)}}{\partial z^*}$ are linear operators in ${\rm Pol}({\mathbb C})_q$ given by $$df=\frac{\partial^{(r)}f}{\partial z}dz+\frac{\partial^{(r)}f} {\partial z^*}dz^*=dz \frac{\partial^{(l)}f}{\partial z}+dz^*\frac{\partial^{(l)}f}{\partial z^*},$$ with $f \in{\rm Pol}({\mathbb C})_q$. Let $\widetilde{\square}:{\rm Pol}({\mathbb C})_q^{\otimes 2}\to{\rm Pol}({\mathbb C})_q^{\otimes 2}$, $m_0:{\rm Pol}({\mathbb C})_q^{\otimes 2}\to{\rm Pol}({\mathbb C})_q$ be linear operators given by $$\widetilde{\square}(f_1 \otimes f_2)=\left(\frac{\partial^{(r)}f_1}{\partial z^*}\otimes 1 \right)\cdot q^{-2}(1-(1+q^{-2})z^*\otimes z+q^{-2}z^{*2}\otimes z^2)\cdot \left(1 \otimes \frac{\partial^{(l)}f_2}{\partial z}\right),$$ $m_0(f_1 \otimes f_2)=f_1f_2$, with $f_1,f_2 \in{\rm Pol}({\mathbb C})_q$. \medskip \begin{theorem}\label{ae}For all $f_1,f_2 \in{\rm Pol}({\mathbb C})_q$, the following asymptotic expansion in $D({\mathbb U})_q'$ is valid: $$m_t(f_1,f_2)\sim_{_{\!\!\!\!\!\!\!\!\!t \to 0}}f_1*f_2,\qquad{\rm with}$$ \begin{equation}\label{stpr}f_1*f_2=f_1 \cdot f_2+\sum_{k=1}^\infty C_k(f_1,f_2)t^k \in{\rm Pol}({\mathbb C})_q[[t]], \end{equation} \begin{equation}\label{Ck}C_k(f_1,f_2)=m_0 \left(\left(p_k \left(\widetilde{\square}\right)- p_{k-1}\left(\widetilde{\square}\right)\right)(f_1 \otimes f_2)\right), \end{equation} and $p_k(x)$, $k \in{\mathbb Z}_+$, are polynomials given by \begin{equation}\label{pk}p_k(x)=\sum_{j=0}^k \frac{(q^{-2k};q^2)_j}{(q^2;q^2)^2_j}q^{2j} \prod_{i=0}^{j-1}(1-q^{2i}((1-q^2)^2 x+1+q^2)+q^{4i+2}). \end{equation} \end{theorem} \medskip This statement is to be proved in the next section, using the results of \cite{SSV5} on a q-analogue of the Berezin transform \cite{UU}. We are grateful to H. T. Koelink who attracted our attention to the fact that the polynomials $p_k(x)$ differ from the polynomials of Al-Salam -- Chihara \cite{K} only by normalizing multiples and a linear change of the variable $x$. \bigskip \section{A q-analogue of the Berezin transform} Remind the notation $t=q^{4 \alpha}$, with $q \in(0,1)$, $\alpha>0$. Consider the linear map ${\rm Pol}({\mathbb C})_q \to{\cal L}_\alpha$ which sends a polynomial $\stackrel{\circ}{f}=\sum \limits_{jk}b_{jk}z^{*j}z^k$ to the linear operator $\widehat{f}=\sum \limits_{jk}b_{jk}\widehat{z}^{*j}\widehat{z}^k$. The polynomial $\stackrel{\circ}{f}$ will be called a {\sl contravariant symbol} of the linear operator $\widehat{f}$. Note that our definitions of covariant and contravariant symbols agree with the conventional ones, as one can observe from \cite{SSV5} (specifically, see proposition 6.6 and lemma 7.2 of that work). The term {\sl 'q-transform of Berezin'} will be stand for the linear operator $B_{q,t}:{\rm Pol}({\mathbb C})_q \to D({\mathbb U})_q'$, $B_{q,t}:\stackrel{\circ}{f}\mapsto f$, which sends the contravariant symbols of linear operators $\widehat{f}=\sum \limits_{jk}b_{jk}\widehat{z}^{*j}\widehat{z}^k$ to their covariant symbols. \medskip {\sc Remark 4.1.} It is easy to extend the operators $B_{q,t}$ onto the entire {\sl 'space of bounded functions in the quantum disc'} via a non-standard approach to their construction (see \cite{SSV5}). \medskip \cite[proposition 5.5]{SSV5} imply \medskip \begin{proposition}\label{Bqtae}Given arbitrary $\stackrel{\circ}{f}\in{\rm Pol}({\mathbb C})_q$, the following {\sl asymptotic expansion} in the topological vector space $D({\mathbb U})_q'$ is valid: $$B_{q,t}\stackrel{\circ}{f}\;\sim_{_{\!\!\!\!\!\!\!\!\!t \to 0}}\;\stackrel{\circ}{f}+\sum_{k=1}^\infty((p_k(\square)\stackrel{\circ}{f}- p_{k-1}(\square)\stackrel{\circ}{f})t^k,$$ with $\square$ being a q-analogue of the Laplace-Beltrami operator $$\square f \stackrel{\rm def}{=}(1-zz^*)^2 \frac{\partial^{(l)}}{\partial z^*}\frac{\partial^{(l)}f}{\partial z}=q^2 \frac{\partial^{(r)}}{\partial z^*}\frac{\partial^{(r)}f}{\partial z}(1-zz^*)^2,$$ with $f \in{\rm Pol}({\mathbb C})_q$ and $p_k$, $k \in{\mathbb Z}_+$, being polynomials given by (\ref{pk}). \end{proposition} \medskip It follows from the definition of the bilinear maps $m_t$, $t \in(0,1)$, that for all $i,j,k,l \in{\mathbb Z}_+$, $f_1,f_2 \in{\rm Pol}({\mathbb C})_q$, $$m_t(z^if_1,f_2)=z^im_t(f_1,f_2),$$ $$m_t(f_1,f_2z^{*l})=m_t(f_1,f_2)z^{*l},$$ $$m_t(z^{*j},z^k)=B_{q,t}(z^{*j}z^k).$$ Hence for all $i,j,k,l \in{\mathbb Z}_+$ one has \begin{equation}\label{mtBqt}m_t((z^iz^{*j}),(z^kz^{*l}))= z^iB_{q,t}(z^{*j}z^k)z^{*l}. \end{equation} We are about to deduce theorem \ref{ae} from (\ref{mtBqt}) and proposition \ref{Bqtae}. In fact, one can easily demonstrate as in \cite[proposition 8.3]{SSV5} that $$\square(f_2(z^*)\cdot f_1(z))=q^2 \frac{\partial^{(r)}f_2(z^*)}{\partial z^*}(1-zz^*)^2 \frac{\partial^{(l)}f_1(z)}{\partial z}=$$ $$=\frac{\partial^{(r)}f_2(z^*)}{\partial z^*}q^{-2}(1-(1+q^{-2})z^*z+ q^{-2}z^{*2}z^2)\frac{\partial^{(l)}f_1(z)}{\partial z}$$ for arbitrary polynomials $f_1(z)$, $f_2(z^*)$. What remains is to compare this expression for $\square$ with the definition of $\widetilde{\square}$ and apply the fact that $\{z^iz^{*j}\}_{i,j \in{\mathbb Z}_+}$ constitute a basis in the vector space ${\rm Pol}({\mathbb C})_q$. \bigskip \section{A formal associativity} \begin{proposition}\label{m}The multiplication in ${\rm Pol}({\mathbb C})_q[[t]]$ given by the bilinear map $$m:{\rm Pol}({\mathbb C})_q[[t]]\times{\rm Pol}({\mathbb C})_q[[t]]\to{\rm Pol}({\mathbb C})_q[[t]],$$ \begin{equation}\label{mf}m:\sum_{j=0}^\infty a_jt^j \times \sum_{k=0}^\infty b_kt^k \mapsto \sum_{i=0}^\infty \left(\sum_{j+k=i}a_j*b_k \right)t^i\;\footnotemark, \end{equation} \footnotetext{The outward sum clearly converges in the topological vector space ${\rm Pol}({\mathbb C})_q[[t]]$.} with $\{a_j \}_{j \in{\mathbb Z}_+},\{b_k \}_{k \in{\mathbb Z}_+}\in{\rm Pol}({\mathbb C})_q$, is associative. \end{proposition} \medskip {\bf Proof.} Introduce the algebra ${\rm End}_{\mathbb C}({\mathbb C}[z]_q)$ of all linear operators in the vector space ${\mathbb C}[z]_q$, and the algebra ${\rm End}_{\mathbb C}({\mathbb C}[z]_q)[[t]]$ of formal series with coefficients in ${\rm End}_{\mathbb C}({\mathbb C}[z]_q)$. To prove our statement, it suffices to establish an isomorphism of the algebra ${\rm Pol}({\mathbb C})_q[[t]]$ equipped with the multiplication $m$ and a subalgebra of ${\rm End}_{\mathbb C}({\mathbb C}[z]_q)[[t]]$ given the standard multiplication. Let ${\cal I}:{\rm Pol}({\mathbb C})_q \to{\rm End}_{\mathbb C}({\mathbb C}[z]_q)[[t]]$ be such a linear operator that for all $j,k,m \in{\mathbb Z}_+$ $${\cal I}(z^jz^{*k}):z^m \mapsto \left \{\begin{array}{ccl}\dfrac{(q^{2m};q^{-2})_k}{(tq^{2m};q^{-2})_k}z^{m-k+j}&,& k \le m \\ 0 &,& k>m \end{array}\right..$$ (More precisely, one should replace the rational function $1/(tq^{2m};q^{-2})_k$ of an indeterminate $t$ with its Teylor expansion.) The following lemma follows from the construction of \cite[section 7]{SSV5}. \medskip \begin{lemma}\label{Q}The linear map $$Q:{\rm Pol}({\mathbb C})_q[[t]]\to{\rm End}_{\mathbb C}({\mathbb C}[z]_q)[[t]],$$ $$Q:\sum_{j=0}^\infty f_jt^j \mapsto \sum_{j=0}^\infty{\cal I}(f_j)t^j\;\footnote{The convergence of the series $\sum \limits_{j=0}^\infty{\cal I}(f_j)t^j$ in the space ${\rm End}_{\mathbb C}({\mathbb C}[z]_q)[[t]]$ is obvious.},\qquad \{f_j \}_{j \in{\mathbb Z}_+}\subset{\rm Pol}({\mathbb C})_q,$$ is injective, and for all $\psi_1,\psi_2 \in{\rm Pol}({\mathbb C})_q[[t]]$ one has $Qm(\psi_1,\psi_2)=(Q \psi_1)\cdot(Q \psi_2)$. \end{lemma} \medskip Lemma \ref{Q} implies the associativity of the multiplication $m$ in ${\rm Pol}({\mathbb C})_q[[t]]$. Thus, proposition \ref{m} is proved. \hfill $\blacksquare$ \medskip Define a linear operator $*$ in ${\rm Pol}({\mathbb C})_q[[t]]$ by $$\left(\sum_{j=0}^\infty f_jt^j \right)^*=\sum_{j=0}^\infty f_j^*t^j,\qquad \{f_j \}_{j \in{\mathbb Z}_+}\subset{\rm Pol}({\mathbb C})_q.$$ \medskip \begin{proposition} $*$ is an involution in ${\rm Pol}({\mathbb C})_q[[t]]$ equipped by $m$ as a multiplication: $$m(\psi_1,\psi_2)^*=m(\psi_2^*,\psi_1^*),\qquad \psi_1,\psi_2 \in{\rm Pol}({\mathbb C})_q[[t]].$$ \end{proposition} \smallskip {\bf Proof.} For all $f_1,f_2 \in{\rm Pol}({\mathbb C})_q$ one has $$(m_0(f_1 \otimes f_2))^*=m_0(f_2^*\otimes f_1^*),$$ $$\widetilde{\square}^{21}(f_1 \otimes f_2)^{*\otimes*}=\widetilde{\square}(f_1^*\otimes f_2^*)$$ with $\widetilde{\square}^{21}=c_0 \square c_0$, and $c_0$ being the flip of tensor multiples. What remains is to observe that the coefficients of $p_n(x)$, $n \in{\mathbb Z}_+$, are real, and to apply (\ref{mf}), (\ref{stpr}), (\ref{Ck}). \hfill $\blacksquare$ \bigskip \section{\boldmath $U_q \mathfrak{su}_{1,1}$-invariance} Remind some well known results on the quantum group $SU_{1,1}$ and the quantum disc (see, for example, \cite{CP, SSV2}). The quantum universal enveloping algebra $U_q \mathfrak{sl}_2$ is a Hopf algebra over ${\mathbb C}$ determined by the generators $K$, $K^{-1}$, $E$, $F$, and the relations $$KK^{-1}=K^{-1}K=1,\qquad K^{\pm 1}E=q^{\pm 2}EK^{\pm 1},\qquad K^{\pm 1}F=q^{\mp 2}FK^{\pm 1},$$ $$EF-FE=(K-K^{-1})/(q-q^{-1}).$$ Comultiplication $\Delta:U_q \mathfrak{sl}_2 \to U_q \mathfrak{sl}_2 \otimes U_q \mathfrak{sl}_2$, counit $\varepsilon:U_q \mathfrak{sl}_2 \to{\mathbb C}$ and antipode $S:U_q \mathfrak{sl}_2 \to U_q \mathfrak{sl}_2$ are given by $$\Delta(K^{\pm 1})=K^{\pm 1}\otimes K^{\pm 1},\qquad \Delta(E)=E \otimes 1+K \otimes E,\qquad \Delta(F)=F \otimes K^{-1}+1 \otimes F,$$ $$\varepsilon(E)=\varepsilon(F)=\varepsilon(K^{\pm 1}-1)=0,$$ $$S(K^{\pm 1})=K^{\mp 1},\qquad S(E)=-K^{-1}E,\qquad S(F)=-FK.$$ The structure of Hopf algebra allows one to define a tensor product of $U_q \mathfrak{sl}_2$-modules and a tensor product of their morphisms. Thus, we obtain a tensor category of $U_q \mathfrak{sl}_2$-modules. Consider an algebra $F$ equipped also with a structure of $U_q \mathfrak{sl}_2$-module. $F$ is called a $U_q \mathfrak{sl}_2$-module algebra if the multiplication $$m_F:F \otimes F \to F,\qquad m_F:f_1 \otimes f_2 \mapsto f_1f_2,\qquad f_1,f_2 \in F,$$ is a morphism of $U_q \mathfrak{sl}_2$-modules. (In the case $F$ has a unit, the above definition should also include its invariance: $\xi \cdot 1=\varepsilon(\xi)1$, $\xi \in U_q \mathfrak{sl}_2$). The following relations determine a structure of $U_q \mathfrak{sl}_2$-module algebra on ${\mathbb C}[z]_q$: \begin{equation}\label{Czq}K^{\pm 1}z=q^{\pm 2}z,\qquad Fz=q^{1/2},\qquad Ez=-q^{1/2}z^2. \end{equation} Equip $U_q \mathfrak{sl}_2$ with an involution: $$E^*=-KF,\qquad F^*=-EK^{-1},\qquad (K^{\pm 1})^*=K^{\pm 1},$$ and let $U_q \mathfrak{su}_{1,1}$ stand for the Hopf $*$-algebra produced this way. An involutive algebra $F$ is said to be $U_q \mathfrak{su}_{1,1}$-module algebra if it is $U_q \mathfrak{sl}_2$-module algebra, and the involutions in $F$ and $U_q \mathfrak{su}_{1,1}$ agree as follows: \begin{equation}\label{inv}(\xi f)^*=(S(\xi))^*f^*,\qquad \xi \in U_q \mathfrak{su}_{1,1},\;f \in F. \end{equation} (\ref{Czq}) determines a structure of $U_q \mathfrak{su}_{1,1}$-module algebra in ${\rm Pol}({\mathbb C})_q$. Thus, each of the vector spaces ${\rm Pol}({\mathbb C})_q$, ${\rm Pol}({\mathbb C})_q[[t]]$ is equipped with a structure of $U_q \mathfrak{su}_{1,1}$-module. \medskip \begin{proposition} ${\rm Pol}({\mathbb C})_q[[t]]$ with the multiplication defined above and the involution $*$ is a $U_q \mathfrak{su}_{1,1}$-module algebra. \end{proposition} \smallskip {\bf Proof.} Since ${\rm Pol}({\mathbb C})_q$ is a $U_q \mathfrak{su}_{1,1}$-module algebra, (\ref{inv}) is valid for $F={\rm Pol}({\mathbb C})_q$. Hence it is also true for $F={\rm Pol}({\mathbb C})_q[[t]]$. What remains is to prove that ${\rm Pol}({\mathbb C})_q[[t]]$ is a $U_q \mathfrak{sl}_2$-module algebra. For that, by a virtue of (\ref{stpr}), (\ref{Ck}), it suffices to demonstrate that the linear maps $m_0$ and $\widetilde{\square}$ are morphisms of $U_q \mathfrak{sl}_2$-modules. As for $m_0$, this property has already been mentioned. So we need only to consider $\widetilde{\square}$. Given any polynomials $f_1(z^*)$, $f_2(z)$, it follows from $\square(f_1(z^*)f_2(z))=\sum \limits_{jk}b_{jk}z^{*j}z^k$, $b_{jk}\in{\mathbb C}$, that $\widetilde{\square}(f_1(z^*)\otimes f_2(z))=\sum \limits_{jk}b_{jk}z^{*j}\otimes z^k$, and $${\widetilde{\square}}(g_1(z)f_1(z^*)\otimes f_2(z)g_2(z^*))=(g_1(z)\otimes 1)\widetilde{\square}(f_1(z^*)\otimes f_2(z))(1 \otimes g_2(z^*)).$$ Thus, it suffices to prove that $\square$ is a morphism $U_q \mathfrak{sl}_2$-modules. This latter result is obtained in \cite{SSV2} (It is a consequence of $U_q \mathfrak{su}_{1,1}$-invariance of the differential calculus in the quantum disc considered there). \hfill $\blacksquare$ \medskip {\sc Remark 6.2.} The works \cite{SSV1, SSV2} deal with the $U_q \mathfrak{su}_{1,1}$-module algebra $D({\mathbb U})_q$ of 'finite functions in the quantum disc'. (The space $D({\mathbb U})_q'$ mentioned in this work is dual to $D({\mathbb U})_q$). The relations (\ref{stpr}) -- (\ref{pk}) determine a formal deformation of $D({\mathbb U})_q$ in the class of $U_q \mathfrak{su}_{1,1}$-module algebras, that is, it allows one to equip $D({\mathbb U})_q[[t]]$ with a structure of $U_q \mathfrak{su}_{1,1}$-module algebra over the ring ${\mathbb C}[[t]]$. \bigskip \section{Concluding notes} We have demonstrated that the method of Berezin allows one to produce a formal deformation for a q-analogue of the unit disc. In \cite{SV}, q-analogues for arbitrary bounded symmetric domains were constructed. We hope in that essentially more general setting, the method of Berezin will help remarkable results to be obtained. \bigskip
\section{Introduction} One-dimensional quantum spin systems exhibit remarkable physical properties. One of the most interesting case is the s=1 antiferromagnetic Heisenberg spin chain. Contrary to the s=1/2 case, this system has a gap as predicted by Haldane\cite{FDM} and a finite spin correlation length. It is an example of a system which is disordered at zero temperature due to quantum fluctuations. The original conjecture has been checked experimentally\cite{exp}, numerically\cite{num} as well as analytically\cite{schulz,AKLT,STN}. Although the ground state is disordered in the sense that spin correlations decay exponentially, there is a hidden topological order\cite{DNR} that is revealed in the bulk of the chain only by nonlocal observables or by ground state degeneracy in an open geometry. This hidden order is most clearly seen in the VBS wavefunction which is an approximate ground state of the spin-1 chain\cite{AKLT}. To construct this wavefunction one has first to write each spin s=1 as a triplet of two fictitious spins s=1/2. Then one couples nearest-neighbor spins s=1/2 into singlets. This leads to a function which is obviously singlet and translation invariant. It is an excellent approximation of the true ground state\cite{nous}. The perfect crystalline pattern of singlets is the hidden order. K. Hida has given an appealing picture of the Haldane gap and the VBS ground state by considering an alternating ferro-antiferromagnetic s=1/2 chain\cite{Hida1}. The Hamiltonian is given by (see figure 1)~: \begin{equation} {\mathcal H}= J_{AF}\sum_{n} {\bf S}_{2n}\cdot {\bf S}_{2n+1} +J_{F} \sum_{n} {\bf S}_{2n+1}\cdot {\bf S}_{2n+2}. \label{faf} \end{equation} Here ${\bf S}_{i}$ are s=1/2 spin operators, $J_{AF}$ is {\it positive} and $J_{F}$ is {\it negative}. In what follows, we set $J_{AF}=1$ and $J_{F}=-\gamma$. The family of systems defined by Eq.(\ref{faf}) has simple limiting cases. For $\gamma =0$ we have a set of decoupled pairs of spins that have a trivial ground state~: all pairs are locked in singlets. When $\gamma\rightarrow\infty$, the s=1/2 are coupled by pairs into s=1 states and we get a chain of spins s=1. Hida has studied numerically the gap of the system as a function of $\gamma$. He has used Lanczos diagonalization techniques to evaluate the gap and he showed that there is no phase transition as a function of $\gamma$. As a consequence, the Haldane gap of the limit $\gamma\rightarrow\infty$ is continuously connected to the trivial gap of the decoupled limit $\gamma =0$. The excitation spectrum also evolves smoothly. So the alternating chain offers a simple physical picture of both the Haldane gap and the hidden topological order. There is another approach to quantum spin chains which is the continuum field theory known as the nonlinear $\sigma$ model (NL$\sigma$M). Introduced originally by Haldane\cite{FDM}, this field theory includes a topological term\cite{Aff} $\theta =2\pi s$ for a spin-s chain. While this term can be discarded when s is integer, it is responsible for masslessness when $\theta =\pi $ (mod $2\pi$), i.e. for half-integer spin chain. This approach has been recently applied \cite{Sierra,Ita,sene} to spin ladders where there is also a parity effect which is given by the number of legs of the ladder. Indeed, spin ladders with even number of legs are generically gapped while odd-numbered ones are gapless. This effect can be explained in the NL$\sigma$M framework with a topological term which is $\theta =2\pi s\times n_{l}$ where $n_{l}$ is the number of legs. It is also of great interest to consider generalized spin ladders with various types of bond alternation and/or additional exchange couplings. For example, a spin ladder with a diagonal coupling interpolates smoothly\cite{White} between the s=1 chain and the two-leg spin ladder. Previous investigations\cite{Takano,Koga,Fukui1,Fukui2} of these more general situations have used the NL$\sigma$M but with the impossibility to treat cases including ferro bonds as in the ferro-antiferro chain introduced by Hida. In this work, we present a generalized NL$\sigma$M that includes additional massive modes and we show that this model is able to reproduce the s=1 limit of alternating chain of Hida. We discuss the construction of this new effective theory in the path-integral framework although it could also be done in the hamiltonian formalism with identical results. In Sect. \ref{1} we treat Hida's chain case. We compute the spin-wave spectrum, motivate the NL$\sigma$M approach and compute an approximate spin gap formula for the whole range of $\gamma$. In sect. \ref{2}, we apply the same formalism to White's mapping from the spin-1 chain to the antiferromagnetic two-leg ladder. Sect. \ref{3} contains our conclusions. \section{Ferro-antiferromagnetic spin chain} \label{1} \subsection{Spin-wave spectrum} A NL$\sigma$M for the spin chain is a non perturbative field theory which is built on the lowest energy modes of the classical theory of the chain. We first study this alternating spin chain at the classical level. The equations of motion for the ferro-antiferromagnetic spin chain follow from $\frac{d}{dt}{\bf n}_i=i[{\cal H},{\bf n}_i]$ and are given by~: \begin{equation} \left\{ \begin{array}{l} \frac{d}{dt}{\bf n}_{2i}=-{\bf n}_{2i}\times(-\gamma {\bf n}_{2i+1} +{\bf n}_{2i-1}), \\ \frac{d}{dt}{\bf n}_{2i+1}={\bf n}_{2i+1}\times(-\gamma {\bf n}_{2i} +{\bf n}_{2i+2}), \end{array} \right. \end{equation} where ${\bf n}_i$ is the classical spin vector at site $i$. Then we linearize those equations around the Neel order configuration ${\bf n}_{2i}=(-1)^is {\bf u}+{\bf \alpha}_i$ and ${\bf n}_{2i+1}=(-1)^is {\bf u}+{\bf \beta}_i$ where $\bf u$ is an arbitrary unitary vector and ${\bf \alpha}_i$ and ${\bf \beta}_i$ are orthogonal planar vector fields. Note that $\bf u$ breaks the natural O(3) symmetry of the hamiltonian. We represent both ${\bf \alpha}_i$ and ${\bf \beta}_i$ by complex fields $\xi_i$ and $\chi_i$. The linearized equations of motion then reads~: \begin{equation} \left\{ \begin{array}{l} \frac{d}{dt}{\xi}_{n}=is(-1)^n(-\gamma \chi_n+\chi_{n-1}+(1+\gamma)\xi_n), \\ \frac{d}{dt}{\chi}_{n}=is(-1)^n(-\gamma \xi_n+\xi_{n+1}+(1+\gamma)\chi_n). \end{array} \right. \end{equation} A plane wave Ansatz for those fields is~: \begin{equation} \left\{ \begin{array}{l} \xi_n=e^{i(\omega+kn)}(a(k)+(-1)^n b(k)), \\ \chi_n=e^{i(\omega+kn)}(c(k)+(-1)^n d(k)). \end{array} \right. \end{equation} The eigenstates $(a,b,c,d)(k)$ of energy $\omega$ and momentum $k$ are the eigenvectors of the following matrix~: \begin{equation} s.\left[ \begin{array}{cccc} 0 & 1+\gamma & 0 & -\gamma -e^{-ik} \\ 1+\gamma & 0 & -\gamma+e^{-ik} & 0 \\ 0 & -\gamma-e^{ik} & 0 & 1+\gamma \\ -\gamma+e^{ik} & 0 & 1+\gamma & 0 \\ \end{array} \right] , \end{equation} with eigenvalues $\omega(k)$. The energy $\omega$ is a solution of its characteristic polynomial \begin{equation} \left(\frac{\omega}{s}\right)^4-4\gamma(1+\gamma) \left(\frac{\omega}{s}\right)^2+4\gamma^2 \sin^2 k=0. \end{equation} The dispersion relation is given by~: \begin{equation} \omega(k)=s\sqrt{2\gamma(1+\gamma)\pm 2\gamma\sqrt{(1+\gamma)^2-\sin^2 k}}. \end{equation} The two lowest positive excitations correspond to momenta $k_c=0$ and $k_c=\pi$. Linearizing the dispersion relation around $k_c$, we obtain for those modes the linear relation $\omega=s\sqrt{\frac{\gamma}{1+\gamma}}|k-k_c|$. Hence the velocity of those spin waves is $v=s\sqrt{\frac{\gamma}{1+\gamma}}$. They stand for the two Goldstone modes of the (classically) broken $O(3)$ symmetry. We will ground our NL$\sigma$M on them, assuming they are slowing varying modes, in an theory where the rotational symmetry is unbroken. We also reasonably assume that the massive modes do not interfere significantly with the Goldstone modes, their energy being of the order of $2s\sqrt{\gamma(1+\gamma)}$. \subsection{The NL$\sigma$M mapping} Here we briefly recall the derivation of the NL$\sigma$M for an antiferromagnetic spin-s chain with $s$ even. So consider a spin-s antiferromagnetic spin chain with s even, characterized by the hamiltonian ${\cal H}=\sum_i J_{AF}{\bf S}_i.{\bf S}_{i+1}$. This spin chain can be mapped onto a NL$\sigma$M . The mapping can be performed within the lagrangian formalism if one uses coherent states representation for the spin operators. The discrete action is made of the exchange interaction terms and of the Berry phases $W[{\bf n}_i]$ of the spin vector field ${\bf n}_i$. \begin{equation} S=\int dt \sum_i Js^2 {\bf n}_i.{\bf n}_{i+1}+ \sum_i s.W[{\bf n}_i]. \end{equation} The coherent state vectors ${\bf n}_i$ are expanded according to \begin{equation} \left\{ \begin{array}{l} {\bf n}_{2i}={\bf l}_{2i}-s\Phi_{2i} \\ {\bf n}_{2i+1}={\bf l}_{2i+1}+s\Phi_{2i+1}, \\ \end{array} \right. \end{equation} where the vector field $\Phi_i$ (local staggered magnetization) is unitary and the field ${\bf l}_i$ satisfy ${\bf n}_i.{\bf l}_i=0$. Those fields can be hinted at from the study of modes in the semi-classical spin chain. Now suppose ${\bf l}_i$ is of order of magnitude $a$, the lattice spacing. This representation of the spin chain allows for a continuous form for the action $S$, which depends on the fields ${\bf l}(x,t)$ and $\Phi(x,t)$. Because we supposed ${\bf l}$ is of magnitude $a$, the action is quadratic in ${\bf l}$ so that the field can be integrated out. Now recall that the spin magnitude s is even. Then the Berry phases of the spin coherent states partially gather to form a topological term $\theta=2\pi s$, the effect of which is null in the action. The final effective action in the field $\Phi$ is~: \begin{equation} S=\int dtdx \frac{1}{2g}\left(\frac{1}{c}(\partial_t \Phi)^2 -c(\partial_x\Phi)^2\right) , \end{equation} where $g$ is the coupling constant of the NL$\sigma$M : $g=2/s$ and $c$ is the velocity $c=2J_{AF}s$. \subsection{Effective NL$\sigma$M } Let us consider now the alternating spin chain. The main difficulty we encounter in trying to represent the Hida chain is the fact that it is inhomogeneous, so that writing an adequate continuous action out of the discrete hamiltonian is not a trivial task. Indeed, the fundamental microscopic structure is a block of two spins (call them $1$ and $2$, say). It will turn out that it is better to choose pairs of spins naturally coupled by the ferromagnetic link, preferably to the antiferromagnetic one. This choice corresponds anyway to Hida's idea of pairing those spins s=1/2 to make them an effective spin-1 in the limit where $\gamma$ goes to infinity (large ferromagnetic coupling). The most rigorous way we could imagine to handle the continuous limit would be to introduce two coherent states vector fields, one for each of the two sites. Unfortunately, this would yield an intricate action, due to the appearance of several equally contributing massive modes. In particular, the Berry phases contribution of those coherent states would not be any more easily recognizable as a topological invariant. Now consider a two-block structure, made of the four spins ${\bf S}_{2i}^1,{\bf S}_{2i}^2,{\bf S}_{2i+1}^1$, and ${\bf S}_{2i+1}^2$. The pairs ${\bf S}_{2i}^1,{\bf S}_{2i+1}^1$ and ${\bf S}_{2i}^2,{\bf S}_{2i+1}^2$ are appropriate candidates to generate two NL$\sigma$M models. A first (incorrect) idea, sustained by our wish to go to the continuous limit, is to assume that they form the same NL$\sigma$M . This assumption would be correct in the limiting case $\gamma$ goes to infinity, but far too crude in any other case. To cure this, at least partially, we add one extra field $\Delta_i$ to the two semi-classical NL$\sigma$M slow modes ${\bf l}_i$ and $\Phi_i$. It represents small quantum fluctuations, remnants of massive modes, that may bring about effective corrections to the NL$\sigma$M action. Accordingly, the coherent states fields are decomposed as~: \begin{equation} \left\{ \begin{array}{l} {\bf n}_{2i}^1={\bf l}_{2i}-s\Phi_{2i}+a\Delta_{2i} \\ {\bf n}_{2i}^2={\bf l}_{2i}-s\Phi_{2i}-a\Delta_{2i} \\ {\bf n}_{2i+1}^1={\bf l}_{2i+1}+s\Phi_{2i+1}-a\Delta_{2i+1} \\ {\bf n}_{2i+1}^2={\bf l}_{2i+1}+s\Phi_{2i+1}+a\Delta_{2i+1} . \end{array} \right. \end{equation} Here $s$ is the spin magnitude and $a$ is the lattice spacing. Note that this lattice spacing is of the length of a two-spin block. The amplitude of the quantum fluctuation is of the order of this lattice spacing. This assumption make the problem a tractable one, since we do not need to consider derivatives of the field $\Delta_i$ when expanding the action. We enforce it by setting $a$ as a prefactor of the fluctuating field. Note that the standard momentum field ${\bf l}$ is implicitly assumed to be of order $a$. We intentionally chose only one fluctuation field, contrary to Senechal's scheme \cite{sene} where the coherent state fields are decomposed on as many possible independent fluctuation fields. We now justify the choice of this particular field by means of a path integral reasoning. We suspect that some highly fluctuating paths are contributing in the action. Indeed, exchange couplings vary on the microscopic scale by a macroscopic amount, and do not behave smoothly with respect to the position. As a consequence some irregular paths might be energetically favorable. Yet a straightforward continuous limit of the action would not retain them since they are not spatially regular. That is why we should enforce some possible (non derivable) contributing fluctuation in the paths. Now let us see why we chose this particular field $\Delta_i$. The fields ${\bf l}_i$ and $\phi_i$ parametrize the variation of the path between the two-block pattern $(2i,2i+1)$, spatially indexed by $i$. So we need one field to represent the variation inside the two-spin blocks. In the block $2i$, the coherent state vectors ${\bf n}_{2i}^1$ and ${\bf n}_{2i}^2$ differ by an amount of $2a\Delta_i$. Since the microscopic pattern is a two-spin block, we have then no choice than to make the coherent state vectors ${\bf n}_{2i+1}^1$ and ${\bf n}_{2i+1}^2$ differ by an amount of $-2a\Delta_i$, because the variation on the scale of the two-spin block are already taken into account within the NL$\sigma$M fields. And this exhausts contributing infinitesimal fluctuations of the path. The action of the spin chain we get through the coherent states representation is of the form \begin{equation} S=\int dt \sum_i s^2 {\bf n}_{2i}.{\bf n}_{2i+1}-\gamma \sum_i s^2{\bf n}_{2i+1}.{\bf n}_{2i+2}+ \sum_i s.W[{\bf n}_i]. \end{equation} One can expand the coupling terms in the action, then goes to the continuous limit, which yields \begin{equation} S_c=\int dt\frac{dx}{a} \left[-4a^2(1+\gamma)\Delta^2-4l^2 +4a^2s\partial_x\Phi.\Delta-a^2 s^2(\partial_x \Phi)^2 \right]. \end{equation} We emphasize the fact that the ferromagnetic exchange terms are of course expanded from its aligned configuration contrary to the antiferromagnetic exchange terms which are expanded from the Neel order. That is the reason why NL$\sigma$ models derived for antiferromagnetic ladder or alternating spin chain \cite{Takano,Koga,Fukui1,Fukui2} cannot be straightforwardly applied to the present cases~: they are not built upon the same semiclassical configurations. Note that since the measure element $dx$ is a two-spin block, it is equal to $a$. The Berry phases of the spins are of the form $\int dtdx \delta{\bf n}.{\bf n}\wedge \partial_t {\bf n}$, where $\delta{\bf n}$ is the spatial variation of the field ${\bf n}$, and give \begin{equation} S_b=\int dt\frac{dx}{a} s(4{\bf l}+2as\partial_x \Phi). \Phi\wedge \partial_t \Phi. \end{equation} Since we sum up the contributions for a double two-spin block, we must divide the whole sum by a factor $2$. We end up with \begin{equation} S=\int dx dt \left[ -2(1+\gamma) \Delta^2-2{\bf l}^2-2s\Delta.\partial_x \Phi -\frac{1}{2}s^2(\partial_x \phi)^2 \right] +\int dx dt \left[ s^2(\partial_x \Phi).\Phi\wedge\partial_t \Phi +2s{\bf l}.\Phi\wedge\partial_t \Phi \right], \end{equation} where we have set $a=1$ for commodity. We recover a spin-2s topological $\theta$-term with $\theta=4\pi s$. This has the consequence that the topological term does not contribute, so that the spin chain is likely to be gapped. We next integrate over the fluctuation fields, that is to say $\bf l$ and $\Delta$. We then obtain the NL$\sigma$M action \begin{equation} S=\int dtdx \left[ \frac{1}{2}(\partial_t \Phi)^2 -\frac{1}{2}s^2 \frac{\gamma}{1+\gamma}(\partial_x \Phi)^2 \right]. \end{equation} The standard parameters of this NL$\sigma$M are the coupling constant $g$ and the velocity $c$ given by \begin{equation} g=\frac{1}{s}\sqrt{\frac{1+\gamma}{\gamma}} \qquad \mbox{and} \qquad c=s \sqrt{\frac{\gamma}{1+\gamma}}. \end{equation} Note that $c$ perfectly matches the velocity $v$ we found for the classical spin waves. When we make $\gamma$ goes to infinity the coupling constant $g$ goes to $1/s$ and the velocity goes to $s$. These are the expected parameters for the 2s-NL$\sigma$M . The Hida's chain is build up of spin one half so that for $s=1/2$ $g$ goes to $2$ and $c$ to $1/2$ in units of the antiferromagnetic exchange coupling. Those are the parameters of a spin-1 chain of exchange $J_{AF}/4$. This is the limit found by Hida for the alternating chain. Let us recall why the magnitude of the exchange coupling should be so. Indeed in this limit the ferromagnetic pairs are in a triplet state. Because of rotation invariance, the limiting hamiltonian can be written as an effective spin-1 chain. The coupling can only be quadratic ${\bf S}_i.{\bf S}_j$ or quartic $({\bf S}_i.{\bf S}_j)^2$. The quartic term is excluded in this limit because it corresponds to second order perturbation theory in $\gamma^{-1}$. The prefactor in front of the spin-1 coupling can be determined with the help of the calculation of a single matrix element, which is easily done on the all-spin-up configuration and yields $J_{\tiny \mbox{eff}}=J_{AF}/4$. So that in this limit, our NL$\sigma$M is consistent with Hida's argument. \subsection{Energy gap evaluation in the large-$N$ limit} To evaluate the spin gap, we compute the mass generated by the NL$\sigma$M in the limiting case of a large number of components for the field $\Phi$($N\rightarrow \infty$). In that limit a closed expression can be obtained for it, thanks to the large N saddle-point approximation. So we made $\Phi$ a N-component field and rescaled it by a factor $\sqrt{N}$. We then enforce the unitary constraint on $\Phi$ by means of a conjugate field $\lambda$, so that the unconstrained partition function is \begin{equation} Z=\int D\Phi D \lambda e^{-S+i\int dxdt \lambda(\Phi^2-N)}. \end{equation} The saddle point equation for the complete action can now be safely derived, and we may look for a constant solution for $\lambda$ that we will call $im^2$. In the limit of large $N$, $m^2$ is the mass generated by the NL$\sigma$M . \begin{equation} \int \frac{dk}{2\pi}\frac{d\omega}{2\pi} \frac{1}{\frac{c}{g}k^2+\frac{1}{gc}\omega^2+m^2}=1. \end{equation} In the limit $\gamma$ goes to $0$, the chain is totally dimerized and the velocity $c$ goes to $0$. Because of this unusual feature, and also because we wish to derive results valid for a large range of $\gamma$, we will not resort to a radial cut-off in the euclidean space-time of the NL$\sigma$M . Rather we will first integrate on frequencies, then integrate on the momenta, with a large-momenta cut-off $\Lambda$. So that instead of the usual radial cut-off, we integrate over a strip in the plane $(k,\omega)$ along the $\omega$-axis. The reason for this is that when $\gamma$ goes to 0, the prefactor $cg^{-1}$ of $k^2$ goes also to zero whereas the prefactor $(cg)^{-1}$ of $\omega^2$ remains constant so that large frequencies are more and more relevant and must not be cut off. In the process (which actually corresponds to the decoupling of dimers), we lost the space dimension of space-time. The first integration is over $\omega$ and gives \begin{equation} g\int_{-\Lambda}^\Lambda \frac{dk}{4\pi} \frac{1}{\sqrt{k^2+\frac{g}{c}m^2}}=1. \end{equation} Then integrating over $k$ we can extract the mass of the NL$\sigma$M \begin{equation} m=\sqrt{\frac{c}{g}} \frac{\Lambda}{\sinh(2\pi s \sqrt{\frac{\gamma}{1+\gamma}})}. \end{equation} In the standard derivation of the mass generated by the NL$\sigma$M (see for example \cite{Auerbach}) one would exchange the function hyperbolic sine for the function exponential. Indeed, in order for this computation to make sense we must have $\sqrt{gc^{-1}}m << \Lambda$. Since in the NL$\sigma$M related to the spin-2s chain, $\sqrt{gc^{-1}}$ is finite, the (then meaningless) hyperbolic sine function can be replaced with the exponential function. Yet for our alternating chain, it can't be done since in the limit $\gamma$ goes to $0$ the argument of the function vanishes. Hence to encompass the full range of $\gamma$ we must retain the hyperbolic sine function. Now we can evaluate the energy gap~: \begin{equation} \Delta_s=\sqrt{gc}.m=\Lambda \frac{s \sqrt{\frac{\gamma}{1+\gamma}}} {\sinh(2\pi s \sqrt{\frac{\gamma}{1+\gamma})}}. \end{equation} Whatever the spin magnitude $s$, the gap of an antiferromagnetic pair of spins is equal to the gap between the triplet state and the singlet, i.e. $J_{AF}$. Hence when $\gamma \rightarrow 0$, we can determine that $\Delta$ goes to $\Lambda/(2\pi)$. This statement allows us to determine the cut-off which appears to be $2\pi$ (actually $2\pi/a$ but we set $a=1$). We can then write~: \begin{equation} \Delta_s=\frac{2\pi s \sqrt{\frac{\gamma}{1+\gamma}}} {\sinh(2\pi s \sqrt{\frac{\gamma}{1+\gamma})}}. \end{equation} We have obtained a {\em self-contained}, estimate of the chain gap for the spin magnitude $s$ . For the spin-1 chain we obtain $\Delta_1=\frac{4\pi}{\sinh \pi}$, which is $1.08$ far from the numerically \cite{Golinelli} known $0.41$. We expect a better result for the spin-2 chain, closer to the "large spin limit". We obtain $\Delta_2=\frac{8\pi}{\sinh 2\pi}$, which is $0.094$ fairly close to the numerically \cite{Jolicoeur} known value of $0.085$. As for the correlation lengths, we obtain $\xi_1\sim 2$ to compare with the numerically known $\xi_1\sim 6$ whereas we obtained $\xi_2\sim 43$ to compare with the numerical value $\xi_2\sim 49$ . The figures labeled 3 and 4 are drawings of the curves of the spin gap w.r.t. $-\gamma$ in the range $\gamma \in [0,1]$ and $-1/\gamma$ in the range $\gamma \in [1,\infty[$. We did so in order to compare our results to Hida's presentation of his numerical computation \cite{Hida1}. Our result agrees qualitatively on the whole range of $\gamma$ including the limiting case $\gamma$ goes to zero. To check that our NL$\sigma$M approach is still valid in the limit $\gamma$ goes to zero, we can also compute the correlation length. It can be read on the saddle point equation~: $\xi=\sqrt{\frac{c}{g}}\frac{1}{m}$. Then we get~: \begin{equation} \xi=\frac{1}{2\pi}\sinh\left( 2\pi s \sqrt{\frac{\gamma}{1+\gamma}} \right). \end{equation} We check that the correlation length goes to zero like $s\sqrt \gamma$ when $\gamma \rightarrow 0$, which is expected since at this point the chain is made of decoupled dimers. \section{An alternating spin ladder} \label{2} In Ref.~(\cite{White}), S.~R. White introduced several mappings that interpolate smoothly between spin ladders and a spin-1 chain within the Haldane'spin gap phase. Here we treat one of these mappings. It consists in an antiferromagnetic spin ladder with an additional diagonal bond in every plaquette formed by legs and rungs (see figure 2). The exchange coupling associated to this bond is ferromagnetic (we will denote it as $D$) and does not introduce any frustration in the ladder. The hamiltonian is given by \begin{equation} {\cal H}=J\sum_{n;a=1,2} {\bf S}^a_{n} {\bf S}^a_{n+1} +K \sum_{n} {\bf S}^1_{n} {\bf S}^2_{n} +D \sum_{n} {\bf S}^1_{n} {\bf S}^2_{n+1} \end{equation} where all exchange couplings are chosen positive. When $D$ goes to $0$ we recover the usual antiferromagnetic spin ladder. Whereas in the limit $-D$ goes to infinity, the pairs of ferromagnetically bounded spin are in a triplet state. The effective hamiltonian can then be expanded by rotational invariance in terms of spin-1 couplings. The effective antiferromagnetic coupling constant is then evaluated on any matrix element of the hamiltonian and gives $J_{\tiny \mbox{eff}}=(2J+K)/4$. So this limit corresponds to an antiferromagnetic spin-1 chain. Since in the process, we remain in the Haldane gapped phase, we may apply our scheme to obtain an estimate of the gap as a function of the diagonal coupling $D$. \subsection{Effective NL$\sigma$M } The microscopic pattern of the ladder is composed of the four spins ${\bf S}_{2i}^1,{\bf S}_{2i}^2, {\bf S}_{2i+1}^1$, and ${\bf S}_{2i+1}^2$ forming a square and, apart from the diagonal bond, three bonds: one on each leg, and the last one on one of the two rungs closing the square (see figure 2). As for the Hida's chain case calculations are done on a doubled cell. Possible candidates to spin pairs forming a NL$\sigma$M are the nearest neighbours on the same leg. Pairs of site linked by a rung contribute to the same NL$\sigma$M . Then, we will need only one extra fluctuation field to describe of fluctuating paths inside the elementary cell, {\em before} going to the continuous limit. Accordingly the coherent states fields are decomposed as~: \begin{equation} \left\{ \begin{array}{l} {\bf n}_{2i}^1={\bf l}_{2i}-s\Phi_{2i}+a\Delta_{2i} \\ {\bf n}_{2i}^2={\bf l}_{2i}-s\Phi_{2i}+a\Delta_{2i} \\ {\bf n}_{2i+1}^1={\bf l}_{2i+1}+s\Phi_{2i+1}-a\Delta_{2i+1} \\ {\bf n}_{2i+1}^2={\bf l}_{2i+1}+s\Phi_{2i+1}-a\Delta_{2i+1} \end{array} \right. \end{equation} In the following we set $J=1$, $K=\rho$ and $D=-\delta$ so that the coupling constants are expressed in units of the longitudinal antiferromagnetic coupling $J$. One can then expand the coupling terms in the action, which yields~: \begin{equation} S_c=\int dt\frac{dx}{a} \left[-4a^2(\rho+\delta)\Delta^2-4(2+\rho){\bf l}^2 +4a^2\delta s\partial_x\Phi.\Delta-a^2 (\rho+\delta) s^2(\partial_x \Phi)^2 \right] \end{equation} with the same care for the ferromagnetic bonds as was done previously. Like for the Hida's chain the measure element $dx$ is a two-spin block and is equal to $a$. The Berry phases of the spins are of the form $\int dtdx \delta{\bf n}.{\bf n}\wedge \partial_t {\bf n}$ and give \begin{equation} S_b=\int dt\frac{dx}{a} 2 s {\bf l}.\Phi\wedge \partial_t \Phi. \end{equation} Since we sum up the contributions for a double two-spin block, we must divide the whole sum by $2$. We end up with~: \begin{equation} S=\int dx dt \left[ -2(2+\rho){\bf l}^2-2(\rho+\delta)\Delta^2-2s\Delta. \partial_x \Phi-\frac{1}{2}s^2(2+\delta)(\partial_x \Phi)^2 \right] +\int dxdt \left[ 2{\bf l}s\Phi\wedge\partial_t \Phi \right] \end{equation} where we have set $a=1$. We next integrate over the fluctuation fields, that is to ${\bf l}$ and $\Delta$. So that we finally obtain the NL$\sigma$M action \begin{equation} S=\int dtdx \left[ \frac{1}{2+\rho}(\partial_t \Phi)^2 -\frac{1}{2}s^2(2+\frac{\delta\rho}{\delta+\rho})(\partial_x \Phi)^2 \right]. \end{equation} The standard parameter of this NL$\sigma$M are \begin{equation} g=\frac{1}{s}\sqrt{\frac{2+\rho}{2+\frac{\delta\rho}{\delta+\rho}}} \qquad \mbox{and} \qquad c=s \sqrt{(2+\rho)\left(2+\frac{\delta \rho}{\delta+\rho}\right)}. \end{equation} Now in order to stick to White's notations, we set $\rho=1$ to get~: \begin{equation} g=\frac{1}{s}\sqrt{\frac{3(1+\delta)}{2+3\delta}} \qquad c=\frac{s}{2}\sqrt{\frac{3(2+3\delta)}{1+\delta}} . \end{equation} These are the coupling and the velocity of the effective NL$\sigma$M . \subsection{Energy gap evaluation in the large $N$-limit} A line of reasoning similar to Hida's chain treatment can be applied to this NL$\sigma$M . In the large N-component limit, we can derive a saddle point equation and obtain the generated mass. The energy gap is then given by \begin{equation} \Delta_L^s=\Lambda.c.\exp(-\frac{2\pi}{g}). \end{equation} Note that contrary to the Hida's chain case, it is not meaningful to stick to the hyperbolic sine function, since it is not more precise than the function exponential. With the previously computed coupling constant $g$ and velocity $c$, we obtain \begin{equation} \Delta_L^s=\Lambda.\frac{s}{2}\sqrt{\frac{3(2+3\delta)}{1+\delta}} \exp\left(-2\pi s \sqrt{\frac{2+3\delta}{3(1+\delta)}} \right). \end{equation} Since when $\delta \rightarrow \infty$, we should recover 3/4 of the gap $\Delta_C$ of the spin-1 chain, we can rewrite it as~: \begin{equation} \frac{\Delta_L^s}{\frac{3}{4}.\Delta_C^s}= \sqrt{\frac{2+3\delta}{3(1+\delta)}} \exp \left[ 2\pi s\left(1- \sqrt{\frac{2+3\delta}{3(1+\delta)}}\right) \right] . \end{equation} Hence we can relate the spin gap of the antiferromagnetic spin-1 chain to the spin gap of the antiferromagnetic two-leg spin $1/2$ ladder by the formula~: \begin{equation} \frac{\Delta_L}{\Delta_C}= \sqrt{\frac{3}{8}} \exp \left[ \pi\left( 1-\sqrt{\frac{2}{3}}\right) \right]. \end{equation} With the value $\Delta_C \simeq 0.41$, we obtain the estimate $\Delta_L \simeq 0.45$, close to the known\cite{Barnes} value $0.50$. In figure 5, we have drawn the curve of the normalized spin gap $\Delta_L/(3/4.\Delta_C)$ w.r.t. $1-2/\pi.\arctan \delta$. We can compare the curve with data from White\cite{White}. Not only does our result agree qualitatively, but it is also quantitatively quite good. \section{Conclusion} \label{3} We have constructed non-linear sigma models appropriate to the description of the properties of some generalized spin chains including ferromagnetic exchanges. In the case of the alternating ferro-antiferromagnetic spin chain, our treatment reproduces correctly the smooth crossover from decoupled dimers to the Haldane phase. This approach may be of relevance to the study of the compound CuNb$_{2}$O$_{6}$ which is such an alternating chain and has a spin gap\cite{Kodama}. We have also treated a ladder including ferro bonds so that the Haldane phase can be reached. Here again there is good agreement with numerical data. \acknowledgments We thank A. V. Chubukov for useful discussions and collaboration at an early stage of this work.
\section{Introduction} In recent times the decoherence of a coherent superposition state has acquired a new dimension \cite{{one},{two},{three},{four},{five}} because of the requirement of the stability of such a superposition. The stability has been investigated for certain systems. The decoherence rates have been calculated and even measured \cite{six} in the context of Cat like states \cite{seven} for the radiation field in a cavity. The decoherence issues are also very significant in the context of quantum computation \cite{{eight},{nine},{ten}}. Clearly the stability of coherent superpositions requires methods for slowing down the decoherence. Several proposals exist in the literature \cite{{eleven},{twelve},{thirteen}}. These involve for example use of a sequence of pulses\cite{twelve} or engineering of the density of states associated with the reservoir or even changing the reservoir interaction from a single photon to multiphoton (or more generally multiboson) interaction \cite{{eleven},{fourteen}}. Other proposals involve feedback methods \cite{thirteen}. It may be added that spontaneous emission in many systems is also a cause of decoherence. We now understand reasonably well, how to inhibit spontaneous emission either by manipulating the density of states \cite{fifteen} or by using external fields \cite{{sixteen},{seventeen}}. The methods based on external fields could be especially useful for slowing down decoherence. In this letter, we discuss a method based on the frequency modulation \cite{{nineteen},{twenty},{twenty one}} of the system-heat bath (environment) coupling. We specifically assume large frequency modulation and take the modulation index $m$ to have a value given by $J_0 (m) = 0$. Under these conditions, we demonstrate considerable slowing down of the decay and decoherence rates. We present a physical basis for this slowing down. We present several examples including the heating of a trapped ion. Our method is useful only if the correlation time $\kappa^{-1}$ of the heat bath is larger than the rate of frequency modulation i.e. $\nu^{-1}$. It must be noted that a recent experimental proposal to control decoherence [6(b)] also depends on a coherent coupling with a bath (second single mode cavity) and works under similar conditions \cite{note}. In order to appreciate the basic idea of using frequency modulation we consider a two state system, where the states $|a\rangle$ and $|b\rangle$ are coupled by some field. We also assume that the state $|b\rangle$ decays at the rate $2\kappa$. This simple model [Fig.1] can describe many physical situations. For example, it can represent an excited atom in a cavity \cite{twenty two} in which the photon leaks out at the rate $2\kappa.$ In this case the states $|a\rangle$ and $|b\rangle$ will correspond to $|e,o\rangle$, $|g,1\rangle,$ where $|o\rangle$ and $|1\rangle$ represent vacuum and one photon state respectively and where $|e\rangle$ and $|g\rangle$ represent the excited and ground states of the atom. It can also describe a situation where the state $|b\rangle$ could be an excited state coupled to ionization continuum. The probability amplitudes $C_a$ and $C_b$ for the states $|a\rangle$ and $|b\rangle$ obey the equations \begin{eqnarray} \dot{C}_a&=& - igC_b,\nonumber\\ \dot{C}_b&=& - \kappa C_b - ig^*C_a. \end{eqnarray} We have removed any fast time dependence by working in an appropriate frame. If $\kappa$ is large, then as is well known \begin{eqnarray} |C_a|^2 &\cong& \exp\{-2\Gamma t\}, \nonumber \\ \Gamma &=& |g|^2/\kappa ;~ \kappa\gg g. \end{eqnarray} The decay of the state $|a\rangle$ arises from the decay of the state $|b\rangle$. In the opposite limit $(\kappa\rightarrow 0)$ one gets oscillatory behavior, which in the cavity context is known as the vacuum field Rabi oscillation. We now consider the effect of a phase modulation $m\sin\nu t$ on the decay of the state $|a\rangle:$ we assume a modulation of the coupling constant \begin{equation} g\rightarrow g\exp\{- im\sin\nu t\}. \end{equation} Here $m$ and $\nu$ give respectively, the amplitude and the frequency of the modulation. Equation (1) is no longer amenable to analytical solutions. In Fig. 2, we display the excited state population for different values of $\nu$ and $m$ chosen to be a zero of the Bessel function of order zero \begin{equation} J_o (m) = 0. \end{equation} This choice of $m$ will become clear in the analysis to follow. In Fig. 2 we also show the behavior in the absence of modulation. We observe that under the condition(4) the decay of the excited state population is considerably slowed as the modulation frequency increases. This clearly demonstrates {\it how a frequency modulation can slow down the effects of decay.} We thus have a {\it method of controlling relaxation / decay} by frequency modulation. We now explain the observed numerical behavior for large $\nu$. Using (1), we can easily derive the following integro-differential equation for the amplitude of the excited state \begin{equation} \dot{C_a}\equiv -|g|^2 e^{-i\Phi(t)}\int_0^t e^{-\kappa(t-\tau)+i\Phi(\tau)}C_a (\tau)d\tau. \end{equation} We use \begin{equation} e^{-i\Phi(t)} = \sum_{l=-\infty}^{+\infty} J_l(m)e^{-il\nu t}, \end{equation} and we assume that (i)~ $\nu$ is large (ii)~$ C_a (\tau)$ varies slowly with~ $\tau$~and carry out a long time average denoted by over bar to get \begin{eqnarray} \frac{\partial}{\partial t}{\rm(lnC_a)}\equiv -|g|^2 \int_0^t e^{-\kappa\tau} \overline{e^{-i\Phi(t)}e^{i\Phi(t-\tau)}}~ d\tau,\nonumber \\ \equiv -|g|^2\sum_{p}J_p^2 (m)(\kappa+ip\nu)^{-1}. \end{eqnarray} In order to slow down the decoherence, we need to remove the $\nu=0$ term in (7). This can be achieved by imposing the condition (4) whence (7) reduces to \begin{equation} \frac {\partial}{\partial t}{\rm(ln C_{a})} \approx -2|g|^2 J_1^2(m)(\kappa^2 +\nu^2)^{-1}\kappa. \end{equation} Therefore the decay of the excited state occurs at a modified rate $2\tilde\Gamma$ with \begin{equation} \frac{\tilde\Gamma}{\Gamma}\cong J_1^2 (m)\left(\frac{2\kappa^2}{\kappa^2+\nu^2}\right). \end{equation} The decay factor (9) agrees very well with the behavior shown in the Fig. 2 for $20\pi$ as then $\nu \gg \kappa$. The very fast oscillations do not show up on the scale of the Fig. 2. The result (9) can be understood by noting that - (i) the factor in the parenthesis in (9) is just the factor that one would have obtained with a detuned interaction between the states $|a\rangle$ and $|b\rangle$; (ii) the Bessel function represents the strength of the first side band. We next demonstrate that the above idea applies rather generally. We consider the usual microscopic treatment of the heat bath \cite{twenty three} with the {\it modulation of the system heat bath coupling}. For the purpose of illustration, we consider a spin system [raising and lowering operators $S^+$ and $S^-$] interacting say with a dc and ac magnetic field in Z direction so that the unperturbed Hamiltonian is $(\omega_0 - m\nu \cos \nu t)S^z$. The energy separation gets modulated - such modulations are routinely used (see e.g. Noel {\it et al} \cite{nineteen}). In the interaction picture the interaction with the heat bath can be written as \begin{equation} H_I(t)=(S^+e^{+i\omega_0 t-i\Phi(t)} R^-(t)+{\rm H.c.}), \end{equation} where $R^-(t)$ is the appropriate operator for the heat bath. As usual \cite{twenty three} we will assume that the coupling of the bath to the system is weak. The heat bath is characterized in terms of the correlation functions: \begin{eqnarray} \langle R^-(t)\rangle &=& 0,\nonumber\\ \langle R^+(t+\tau) R^-(t)\rangle &=& C^{+-} (\tau),\nonumber\\ \langle R^-(t+\tau) R^+(t)\rangle &=& C^{-+} (\tau),\nonumber\\ \langle R^-(t+\tau) R^-(t)\rangle &=& 0. \end{eqnarray} The Fourier transforms of~$C^{+-}$~ and~$ C^{-+}$~ are related via the fluctuation dissipation theorem. We can now do the standard calculation \cite{twenty three} to derive a master equation for the reduced density matrix $\rho$ of the system alone. We quote the result of this calculation. \begin{eqnarray} \frac{\partial \rho}{\partial t} = -(S^+S^-~\rho -S^-~\rho S^+) \int_0^t d\tau ~C^{-+}(\tau)e^{+i\omega_0\tau}e^{-i\Phi(t)}e^{i\Phi (t-\tau)},\nonumber\\ -(\rho~S^-S^+ - S^+~\rho~ S^-)\int_0^t d\tau~C^{+-} (-\tau)e^{+i\omega_0\tau} e^{i\Phi(t-\tau)-i\Phi(t)},\nonumber\\ {\rm +~terms~with~subscripts}~\pm~\rightarrow~\mp, ~\omega_0 ~\rightarrow -~\omega_0,~\Phi~\rightarrow -\Phi. \end{eqnarray} First of all we note, that if the bath correlations were like delta correlations ~$C^{-+}(\tau) = 2\delta (\tau)C^{-+}$,~ then the master equation (12) does {\it not depend on the modulation~$\Phi$}. Clearly, the {\it bath correlation time $\tau_c$ has to be at least of the order of the time} associated with the modulation. Under the fundamental condition (4), the time average in (12) can be {\it approximated by} \begin{equation} \overline{e^{-i\Phi(t)+i\Phi(t-\tau)}} \cong 2J_1^2 (m)~\cos~\nu~\tau \end{equation} and then (12) reduces to \begin{eqnarray} \frac{\partial\rho}{\partial t}\equiv -2(S^+S^- \rho - S^-\rho~ S^+) \int_0^\infty d\tau C^{-+}(\tau)e^{+ i\omega_0\tau}~\cos~\nu~\tau~ J_1^2 (m),\nonumber\\ -2~(~\rho~ S^-S^+ - S^+ \rho~S^-)\int_0^\infty d\tau C^{+-} (-\tau)e^{+i\omega_0\tau}~\cos~\nu~\tau~ J_1^2(m),\nonumber\\ {\rm +~terms~ with}~\pm~\rightarrow~\mp,~\omega_0~\rightarrow~-\omega_0. \end{eqnarray} The standard master equation corresponds to the limits $\nu\rightarrow 0, J_1^2 \rightarrow 1$. It is clear that if $\nu$ is large enough compared to frequency scale of $C^{-+}(\tau)e^{i\omega_0\tau}$ then the real part of the integral in (14) will be approximately zero and {\it decoherence effectively does not exist}. In particular, if~ $C^{-+}(\tau) ~= ~C_0^{-+}~ e^{-\kappa\tau-i\omega\tau}, ~C^{+-}(-\tau)~=~ C_0^{-+}~e^{-\kappa\tau - i\omega\tau},$~ then \begin{eqnarray} \frac{\partial\rho}{\partial t} &=& -\frac{2(\kappa-i\Delta)J_1^2(m)}{(\kappa - i\Delta)^2+\nu^2} \{C_0^{-+}(S^+S^-~\rho - S^-~\rho~ S^+) \nonumber \\ &+& C_0^{+-}(\rho~S^-S^+ - S^+\rho~S^-)\}{\rm + c.c.}.,\Delta = (\omega_0 - \omega). \end{eqnarray} Clearly, the relaxation coefficients in the master equation are modified by factors like (9). Hence the relaxation is much slower. In particular, the relaxation of the coherence~$\langle S^\pm\rangle$~ will be on a much longer time scale. For large $\nu$ compared to $\kappa$ and $\Delta$, the relaxation time is very large. We next consider the application of the above ideas to the decoherence of an ion in a trap. This is important in many applications of ion traps such as in connection with the production of Cat states and in quantum computation. In particular, we consider the possibility of {\it reducing} the {\it heating} of the {\it ground state of trapped ions}. In a recent letter, James \cite{ten} has considered a model for heating produced by a stochastic field {\rm E(t)}. In terms of the annihilation and creation operators $a, a^{\dagger}$ associated with the ionic motion the heating is described by the Hamiltonian \begin{equation} H_1 = i\hbar[u(t)a^\dagger - u^*(t)a], \end{equation} where $u(t) = iZE(t)e^{i\omega_0 t}/\sqrt{2M\hbar\omega_0}$. The field $E(t)$ is a Gaussian stochastic process. The time scale of the stochastic field is taken to be comparable to the time scale of the ionic motion. Hence this model is {\it outside} the usual {\it markovian limit}. We now consider the effect of an external modulation so that effectively, $u(t)\rightarrow u(t)e^{-i\Phi(t)}$. Following James' work the fidelity ${\rm F(t)}$ of the ground state is given by \begin{equation} F(t)=[1+2\langle|v(t)|^2\rangle + \langle|v(t)|^2\rangle^2 - |\langle v^2(t)\rangle|^2]^{-1/2}, \end{equation} \begin{equation} v(t) = \frac{iZ}{\sqrt{2M\hbar\omega_0}}\int_0^t E(t^\prime)e^{-i\Phi(t^\prime)+i\omega_0 t^\prime}dt^\prime. \end{equation} The mean values in (17) can be obtained from (18) by assuming exponential correlation for $E(t):$ \begin{eqnarray} \langle u(t)u^*(t^\prime)\rangle &=&\frac{\Omega^2}{2} e^{-\kappa|t-t^{\prime}|}\nonumber\\ \langle |v|^2 \rangle &=&\frac{\Omega^2}{2}\sum_{- \infty}^{+\infty}\sum J_n (m) J_p (m) I (\omega_0-n\nu ~; ~-\omega_0+p\nu) \nonumber \\ \langle v^2\rangle &=& - \frac{\Omega^2}{2}\sum_{-\infty}^{+\infty}\sum J_n (m) J_p (m)I(\omega_0 - n\nu ~;~ \omega_0 - p\nu) \end{eqnarray} where the integral $I(\omega_\alpha, \omega_\beta)$ is found to be \begin{eqnarray} I(\omega_\alpha,\omega_\beta) &=& (i\omega_\alpha + i\omega_\beta)^{-1} \left[(\kappa + i\omega_\beta)^{-1}e^{i(\omega_\alpha+\omega_\beta)t} -(\kappa - i\omega_\alpha)^{-1} \right]\nonumber \\ &+& (i\omega_\alpha - \kappa)^{-1}(-i\omega_\beta -\kappa)^{-1} e ^{i\omega_\alpha t- \kappa t}\nonumber \\ &+& ~{\rm terms~ with}~ \alpha\Leftrightarrow\beta. \end{eqnarray} Note that $\omega_\alpha + \omega_\beta$ can vanish in which case, a limiting procedure leads to \begin{equation} I (\omega_\alpha, -\omega_\alpha) = (\kappa - i\omega_\alpha)^{-1}~ t + (i\omega_\alpha - \kappa)^{-2} \left(e^{i\omega_\alpha t-\kappa t}-1\right) + {\rm c.c.} \end{equation} We show the fidelity factor $F$ in Fig. 3, both in the absence and presence of the modulation. We choose a parameter domain in which {\it fidelity} was being {\it degraded} rather fast. Clearly, if we assume large frequency modulation and condition (4), then as the figure shows, there is considerable {\it improvement in the fidelity} under frequency modulation of the stochastic field {\rm E(t)} responsible for heating the trapped ion. In conclusion, we have shown how the appropriate modulation of the system-heat bath interaction can slow down the decay as well as the decoherence to a very large extent. This happens as generally, the decoherence is determined by the spectral components of the bath correlation functions. If system-bath interaction is modulated, then the decoherence is determined by the spectral components, which are shifted by the multiples of the modulation frequency. If the modulation frequency is large compared to the width of the bath correlations, then we would get much smaller decoherence rate. Finally, note that we have a method to control the effects of decoherence since the modulation depth and frequency can be varied. The author thanks Sunish Menon for the plots.
\section{introduction} It is widely accepted that (co-)quasitriangular Hopf algebra is a good algebraic notion which expresses ``quantum groups.'' For, example, each lattice model $w$ of vertex type (and of face type) without spectral parameter naturally generates a coquasitriangular (CQT) Hopf (face) algebra, thanks to the FRT construction and the Hopf closure (or Hopf envelope) construction. The former construction assigns $w$ to the CQT bialgebra (or face algebra) ${\frak A}(w)$ (cf. \cite{RTF}, \cite{LarsonTowber}, \cite{Schauenburg}, \cite{gd}), while the latter construction assigns some CQT bialgebra (or face algebra) $\frak H$ to the CQT Hopf (face) algebra $\mathrm{Hc} (\frak H)$ (\cite{Phung}, \cite{gsg}). However, to give applications of CQT Hopf (face) algebras $\frak H$ to low-dimensional topology, we need one additional structure on these, which is called the {\it ribbon functional} on $\frak H$, a dual notion of the ribbon element. It is known that there exists a Drinfeld's double of a finite-dimensional Hopf algebra, which has no ribbon element (cf. \cite{KauffmanRadford} Proposition 7). Also, it is known that the ribbon functional of a CQT Hopf algebra is not necessarily unique even if it exists. Hence it is natural to investigate sufficient conditions for the existence of the ribbon functional on CQT Hopf (face) algebras, and to develop the classification theory of the ribbon functionals. One of the purpose of this paper is to prove the existence of the ribbon functionals on CQT Hopf face algebras of the form $\mathrm{Hc} ({\frak A}(w))$ (cf. Theorem \ref{existrib}). This result implies that each $w$ produces a ribbon category, and therefore, it implies that $w$ generates a family of link invariants. We note that when $w = \Check{R}$ is a vertex model, this family contains the link invariant constructed by Reshetikhin \cite{Reshetikhin} (see Remark at the end of Section 6). As byproducts, we also obtain several useful results on the ribbon functionals on CQT Hopf face algebras. The other purpose of this paper is to give the classification of the braidings and the ribbon functionals on the function algebras $\mathrm{Fun} (G_q)$ of the quantized classical groups $G_q = GL_q(N)$, $SL_q(N)$, $SO_q(N)$, $O_q(N)$ and $Sp_q(N)$, and also, on some Hopf face algebras ${\frak S}(A_{N-1};t)_{\epsilon}$ which are closely related to the $SU(N)_L$-topological quantum field theories. The braiding of these Hopf (face) algebras is not unique. However, the non-uniqueness is explained using certain gradings of these algebras via cyclic groups $\Gamma$. The ribbon functionals of these algebras always exist and the number of those is at most two. We note that the proof of the former is very similar to that of the classification of the braidings of $\mathrm{Fun} (\mathrm{Mat}_q (N))$ due to Takeuchi \cite{Takeuchicocycle}, while the proof of the latter essentially depends on our general theory for $\mathrm{Hc} ({\frak A}(w))$. The paper is organized as follows. In Section 2, we recall basic concepts of the face algebra. The notion of the face algebra generalizes that of the bialgebra, and is necessary to study lattice models of face type and the corresponding link invariants in the framework of the quantum group theory. In Section 3, we recall the relation between lattice models and face algebras. In Section 4, we recall the Hopf closure construction which is the the main tool of this paper. In Section 5, we give a study of group-like elements of the dual algebras ${\frak A}(w)^{\circ}$ and $\mathrm{Hc} (\frak H)^{\circ}$. It plays a crucial role for our study of the ribbon functionals. In Section 6, we give several results on the ribbon functionals of $\mathrm{Hc} (\frak H)$ and its quotients. In Section 7 and Section 8, we give the results for quantized classical groups and the algebras ${\frak S}(A_{N-1};t)_{\epsilon}$ stated above. The author is grateful to Dr. S. Suzuki for his useful informations. \begin{equation*} \end{equation*} \par\noindent {\it Notation}. Throughout this paper, we use Sweedler's sigma notation for coalgebras $C$ and their right comodules $U$, such as $(\Delta \otimes \mathrm{id}) (\Delta (a)) = \sum_{(a)} a_{(1)} \otimes a_{(2)} \otimes a_{(3)}$ (cf. \cite{Montgomery}). Also, we denote by $\rho_U$ the coaction $U \to U \otimes C;$ $u \mapsto \sum_{(u)} u_{(0)} \otimes u_{(1)}$, and by $\pi_U$ the left action of $C^*$ on $U$ given by $\pi_U (X) u = \sum_{(u)} u_{(0)} X(u_{(1)})$ $(u \in U, X \in C^*)$. For a linear operator $A$ on a vector space $W$ with basis $\{ \bold p \}$, we define its matrix $[A^{\bold p}_{\bold q}]_{\bold p \bold q}$ by $A \bold q = \sum_{\bold p} \bold p A^{\bold p}_{\bold q}$. \section{face algebras} Let $\frak H$ be an algebra over a field $\Bbb K$ equipped with a coalgebra structure $(\frak H,\Delta,\varepsilon)$. Let ${\EuScript V}$ be a finite nonempty set and let $e_{\frak H,i} = e_i$ and ${{\stackrel{\scriptscriptstyle\circ}{e}}_{\frak H,i} = {\stackrel{\scriptscriptstyle\circ}{e}_i}}$ $(i \in {\EuScript V})$ be elements of ${\frak H}$. We say that $({\frak H}, \{ e_i,{\stackrel{\scriptscriptstyle\circ}{e}_i} \})$ is a {\em ${\EuScript V}$-face algebra} if the following relations are satisfied: \begin{equation} \Delta (ab) = \Delta(a) \Delta(b), \label{D(ab)} \end{equation} \begin{equation} e_ie_j = \delta_{ij}e_i, \quad {\stackrel{\scriptscriptstyle\circ}{e}_i} {\stackrel{\scriptscriptstyle\circ}{e}_j} = \delta_{ij} {\stackrel{\scriptscriptstyle\circ}{e}_i}, \quad e_i {\stackrel{\scriptscriptstyle\circ}{e}_j} = {\stackrel{\scriptscriptstyle\circ}{e}_j} e_i, \label{ee} \end{equation} \begin{equation} \sum_{k \in \EuScript V} \, e_k = 1 = \sum_{k \in {\EuScript V}} {\stackrel{\scriptscriptstyle\circ}{e}_k}, \label{sume} \end{equation} \begin{equation} \Delta({\stackrel{\scriptscriptstyle\circ}{e}_i} e_j) = \sum_{k \in {\EuScript V}} {\stackrel{\scriptscriptstyle\circ}{e}_i} e_k \otimes {\stackrel{\scriptscriptstyle\circ}{e}_k} e_j, \quad \varepsilon ({\stackrel{\scriptscriptstyle\circ}{e}_i} e_j) = \delta _{ij} , \label{D(ee)} \end{equation} \begin{equation} \varepsilon (ab) = \sum_{k \in {\EuScript V}} \varepsilon (ae_k) \varepsilon ({\stackrel{\scriptscriptstyle\circ}{e}_k} b) \label{e(ab)} \end{equation} for each $a,b \in \frak H$ and $i,j \in \EuScript V$. We call elements $e_i$ and ${\stackrel{\scriptscriptstyle\circ}{e}_i}$ {\it face idempotents} of $\frak H$. We denote by $\frak{E} = \frak{E}_{\frak H}$ the subalgebra of $\frak H$ generated by face idempotents. It is known that bialgebra is an equivalent notion of $\EuScript V$-face algebra with $\#(\EuScript V) = 1$. For a $\EuScript V$-face algebra, we have the following formulas: \begin{equation} \label{e(eae)} \varepsilon ({\stackrel{\scriptscriptstyle\circ}{e}_i} a) = \varepsilon (e_i a), \quad \varepsilon (a {\stackrel{\scriptscriptstyle\circ}{e}_i}) = \varepsilon (a e_i), \end{equation} \begin{equation} \label{ae(eae)} \sum_{(a)} a_{(1)} \varepsilon (e_i a_{(2)} e_j) = e_i a e_j, \end{equation} \begin{equation} \label{e(eae)a} \sum_{(a)} \varepsilon (e_i a_{(1)} e_j)a_{(2)} = {\stackrel{\scriptscriptstyle\circ}{e}_i} a {\stackrel{\scriptscriptstyle\circ}{e}_j}, \end{equation} \begin{equation} \label{D(a)} \Delta (a) = \sum_{k,l \in \V} \sum_{(a)} e_k a_{(1)} e_l \otimes {\stackrel{\scriptscriptstyle\circ}{e}_k} a_{(2)} {\stackrel{\scriptscriptstyle\circ}{e}_l}, \end{equation} \begin{equation} \label{eae*a} \sum_{(a)} e_i a_{(1)} e_j \otimes a_{(2)} = \sum_{(a)} a_{(1)} \otimes {\stackrel{\scriptscriptstyle\circ}{e}_i} a_{(2)} {\stackrel{\scriptscriptstyle\circ}{e}_j}, \end{equation} \begin{equation} \label{D(eeaee)} \Delta ({\stackrel{\scriptscriptstyle\circ}{e}_i} e_j a {\stackrel{\scriptscriptstyle\circ}{e}}_{i'} e_{j'}) = \sum_{(a)} {\stackrel{\scriptscriptstyle\circ}{e}_i} a_{(1)} {\stackrel{\scriptscriptstyle\circ}{e}}_{i'} \otimes e_j a_{(2)} e_{j'} \end{equation} for each $a \in \frak H$ and $i,j,i',j' \in \EuScript V$. For a ${\EuScript V}$-face algebra $\frak H$, its {\it dual face algebra} ${\frak H}^{\circ}$ \cite{fdd} is defined to be the dual coalgebra of $\frak H$ equipped with product and face idempotents given by $\langle XY,\, a \rangle$ $=$ $\sum_{(a)} \langle X,\, a_{(1)} \rangle \langle Y,\, a_{(2)} \rangle$ $(X, Y \in \frak H^{\circ}, a \in \frak H)$ and \begin{equation} \label{eHoi} \langle e_{{\frak H}^{\circ},i},\, a \rangle = \varepsilon (a e_{\frak H,i}), \quad \langle {\stackrel{\scriptscriptstyle\circ}{e}}_{{\frak H}^{\circ},i},\, a \rangle = \varepsilon (e_{\frak H,i} a) \quad (a \in \frak H, i \in \EuScript V). \end{equation} Let $x^+$, $x^-$, $e^+$ and $e^-$ be elements of an arbitrary algebra $A$. We say that $x^-$ is an {\em $(e^+,e^-)$-generalized inverse} of $x^+$ if the following four relations are satisfied: \begin{equation} x^{\mp} x^{\pm} = e^{\pm}, \quad x^{\pm} x^{\mp} x^{\pm} = x^{\pm}. \label{gen.inv.def} \end{equation} We note that the $(e^+,e^-)$-generalized inverse of $x^+$ is unique if it exists. We say that a linear map $S\!: {\frak H} \to {\frak H}$ is an {\em antipode} of ${\frak H}$, or ${\frak H}$ is a {\em Hopf} ${\EuScript V}$-{\em face algebra} if $S$ is the $(E^+, E^-)$-generalized inverse of $\mathrm{id}_{\frak H}$ with respect to the convolution product of $\mathrm{End}_{\Bbb K} (\frak H)$, where \begin{equation} \label{E+adef} E^+ (a) = \sum_{k \in \V} \varepsilon (ae_k) e_k, \quad E^- (a) = \sum_{k \in \V} \varepsilon (e_k a) {\stackrel{\scriptscriptstyle\circ}{e}_k} \quad (a \in \frak H). \end{equation} An antipode of a $\EuScript V$-face algebra is an antialgebra-anticoalgebra map, which satisfies \begin{equation} S({\stackrel{\scriptscriptstyle\circ}{e}_i} e_j) = {\stackrel{\scriptscriptstyle\circ}{e}_j} e_i \quad (i,j \in {\EuScript V}). \label{S(ee)} \end{equation} Let ${\frak H}$ be a ${\EuScript V}$-face algebra and let ${\cal R}^+ = {\cal R}^+_{\frak H}$ be an element of $({\frak H} \otimes{\frak H})^*$ with $(m^* (1), (m^{\text{op}})^* (1))$-generalized inverse ${\cal R}^- = {\cal R}^-_{\frak H}$, where $m\!: {\frak H} \otimes {\frak H} \to {\frak H}$ denotes the product of ${\frak H}$. We say that ${\cal R}^+$ is a {\em braiding} of $\frak H$ or $(\frak H,{\cal R}^{\pm})$ is a {\em coquasitriangular} (or {\em CQT}) ${\EuScript V}$-face algebra if the following relations are satisfied: \begin{equation} \label{RmXR} \quad {\cal R}^+ m^*(X) {\cal R}^- = (m^{\text{op}})^*(X) \quad (X \in \frak H^*), \end{equation} \begin{equation} \label{mid(R)} (m {\otimes}\text{id})^* ({\cal R}^+) = {\cal R}^+_{13} {\cal R}^+_{23}, \quad (\text{id}{\otimes}m)^* ({\cal R}^+) = {\cal R}^+_{13} {\cal R}^+_{12}. \end{equation} Here for $Z \in (\frak H \otimes \frak H)^*$ and $\{ i,j,k \} = \{1,2,3 \}$, we define $Z_{ij} \in (\frak H^{\otimes 3})^*$ by $Z_{ij} (a_1, a_2, a_3) = Z(a_i, a_j) \varepsilon (a_k)$ $(a_1, a_2, a_3 \in \frak H)$. The braiding ${\cal R}^+$ satisfies the following relations: \begin{equation} {\cal R}^{\pm}_{12} {\cal R}^{\pm}_{13} {\cal R}^{\pm}_{23} = {\cal R}^{\pm}_{23} {\cal R}^{\pm}_{13} {\cal R}^{\pm}_{12}, \label{RRR} \end{equation} \begin{equation} \cal{R}^{\mp}_{23} {\cal R}^{\pm}_{12} {\cal R}^{\pm}_{13} = {\cal R}^{\pm}_{13} {\cal R}^{\pm}_{12} \cal{R}^{\mp}_{23}, \quad \cal{R}^{\mp}_{13} \cal{R}^{\mp}_{23} {\cal R}^{\pm}_{12} = {\cal R}^{\pm}_{12} \cal{R}^{\mp}_{23} \cal{R}^{\mp}_{13}, \label{RmpRR} \end{equation} \begin{equation} {\cal R}^+ ({\stackrel{\scriptscriptstyle\circ}{e}_i} e_j a {\stackrel{\scriptscriptstyle\circ}{e}_k} e_l,\, b) = {\cal R}^+ (a,\, {\stackrel{\scriptscriptstyle\circ}{e}_j} e_l b {\stackrel{\scriptscriptstyle\circ}{e}_i} e_k), \label{Rpeeee} \end{equation} \begin{equation} {\cal R}^- ({\stackrel{\scriptscriptstyle\circ}{e}_i} e_j a {\stackrel{\scriptscriptstyle\circ}{e}_k} e_l,\, b) = {\cal R}^- (a,\, {\stackrel{\scriptscriptstyle\circ}{e}_k} e_i b {\stackrel{\scriptscriptstyle\circ}{e}_l} e_j), \label{Rmeeee} \end{equation} \begin{equation} {\cal R}^+ ({\stackrel{\scriptscriptstyle\circ}{e}_i} e_j,\, a) = \varepsilon (e_j a e_i), \quad {\cal R}^+ (a,\, {\stackrel{\scriptscriptstyle\circ}{e}_i} e_j) = \varepsilon (e_i a e_j), \label{Rpeea} \end{equation} \begin{equation} {\cal R}^- ({\stackrel{\scriptscriptstyle\circ}{e}_i} e_j,\, a) = \varepsilon(e_i a e_j), \quad {\cal R}^- (a,\, {\stackrel{\scriptscriptstyle\circ}{e}_i} e_j) = \varepsilon(e_j a e_i) \label{Rmeea} \end{equation} for each $a,b \in \frak H$ and $i,j,k,l \in \EuScript V$. If $\frak H$ is a Hopf $\EuScript V$-face algebra, then we have: \begin{equation} (S{\otimes}\text{id})^* ({\cal R}^+) = {\cal R}^-, \quad (\text{id}{\otimes}S)^* ({\cal R}^-) = {\cal R}^+. \label{Sid(R)} \end{equation} \begin{prop} \label{altbraiding} Let $(\frak H, {\cal R}^{\pm})$ be a CQT $\EuScript V$-face algebra.\\ \rom{(1)} Then, ${\cal R}^-_{21}\!: a \otimes b \mapsto {\cal R}^-(b, a)$ gives another braiding of $\frak H$.\\ % \rom{(2)} Let $\Gamma$ be a semigroup and $\frak H = \bigoplus_{\gamma \in \Gamma} \frak H_{\gamma}$ a decomposition of $\frak H$ such that $\frak H_{\gamma} \frak H_{\delta} \subset \frak H_{\gamma \delta}$ and that $\Delta ( \frak H_{\gamma} ) \subset \frak H_{\gamma} \otimes \frak H_{\gamma}$. Let $\chi\!: \Gamma \times \Gamma \to \Bbb K^{\times}$ be a map such that $\chi (\gamma_1 \gamma_2, \delta)$ $=$ $\chi (\gamma_1, \delta) \chi (\gamma_2, \delta)$, $\chi (\gamma, \delta_1 \delta_2)$ $=$ $\chi (\gamma, \delta_1) \chi (\gamma, \delta_2)$. Then, there exists a new braiding ${\cal R}^+_{\chi}$ of $\frak H$ given by \begin{equation} \label{Rchidef} {\cal R}^{\pm}_{\chi} (a, b) = \chi (\gamma, \delta)^{\pm 1} {\cal R}^{\pm} (a, b) \quad (a \in \frak H_{\gamma}, b \in \frak H_{\delta}). \end{equation} % If $(\frak H, {\cal R}^{\pm})$ is closable, then so is $(\frak H, {\cal R}^{\pm}_{\chi})$. \end{prop} \begin{pf} This is straightforward. \end{pf} Let $({\frak H},{\cal R}^{\pm})$ be a CQT Hopf $\EuScript V$-face algebra and $\cal{V}$ an invertible central element of ${\frak H}^*$. We say that $\cal{V}$ is a {\em ribbon functional} of $\frak H$, or $(\frak H,\cal{V})$ is a {\em coribbon Hopf} $\EuScript V$-{\em face algebra} if \begin{equation} \label{m(V)} m^*({\cal V}) = {\cal R}^- {\cal R}^-_{21} ({\cal V} \otimes {\cal V}), \end{equation} \begin{equation} \label{S(V)} S^* (\cal{V}) = \cal{V}. \end{equation} A map $f\!: \frak H \to \frak K$ between $\EuScript V$-face algebras is called a {\em map of $\EuScript V$-face algebras} if it is both an algebra and a coalgebra map such that $f(e_i) = e_i$, $f({\stackrel{\scriptscriptstyle\circ}{e}_i}) = {\stackrel{\scriptscriptstyle\circ}{e}_i}$ for each $i \in \EuScript V$. If both $\frak H$ and $\frak K$ have antipode, then we have \begin{equation} f(S(a)) = S(f(a)) \quad (a \in \frak H). \label{f(S(a))} \end{equation} A map $f\!: \frak H \to \frak K$ of $\EuScript V$-face algebras between CQT $\EuScript V$-face algebras is called a {\em map of CQT $\EuScript V$-face algebras} if \begin{equation} \label{mapCQTdef} (f \otimes f)^* ({\cal R}^+_{\frak K}) = {\cal R}^+_{\frak H}. \end{equation} An ideal $\frak{I}$ of a $\EuScript V$-face algebra $\frak H$ is called a {\it biideal} if it is a coideal of the underlying coalgebra of $\frak H$. If in addition, $\frak H$ is a CQT $\EuScript V$-face algebra and $\frak{I}$ satisfies ${\cal R}^{\pm}(\frak{I},\frak H) = {\cal R}^{\pm}(\frak H, \frak{I}) = 0$, then $\frak{I}$ is called a {\it CQT biideal} of $\frak H$. For each $\EuScript V$-face algebra (resp. CQT $\EuScript V$-face algebra) $\frak H$ and its biideal (resp. CQT biideal) $\frak{I}$, the quotient $\frak H / \frak{I}$ becomes a $\EuScript V$-face algebra (resp. CQT $\EuScript V$-face algebra) in an obvious manner. \section{lattice models and comodules} Let $\EuScript G$ be a finite oriented graph with set of vertices ${\EuScript V}$ = ${{\EuScript G}^0}$. For an edge ${\bold p}$, we denote by ${\frak {s} (\bold p)}$ and ${\frak {r} (\bold p)}$ its {\em source} ({\em start}) and its {\em range} ({\em end}) respectively. For each $m \geq 1$, we denote by ${{\EuScript G}^m = {\coprod}_{i,j \in \EuScript V}{\EuScript G}_{ij}^m}$ the set of {\em paths} of $\EuScript G$ of {\em length} $m$, that is, $\bold p \in {\EuScript G}_{ij}^m$ if $\bold p$ is a sequence $(\bold p_1, \ldots, \bold p_m)$ of edges of $\EuScript G$ such that $\frak {s} (\bold p):= \frak {s} (\bold p_1) = i$, $\frak {r} (\bold p_n) = \frak {s}(\bold p_{n+1})\,$ $(1 \leq n < m)$ and $\frak {r} (\bold p):= \frak {r}(\bold p_m) = j$. Also, we set $\frak {s} (i) = i = \frak {r} (i)$, $\EuScript G_{ii}^0 = \{ i\}$ and $\EuScript G_{ij}^0 = \emptyset$ for each $i \in \EuScript V$ and $j \ne i$. Let ${\frak H (\EuScript G)}$ be the linear span of the symbols ${e{\bold p \choose \bold q}}$ \( ( \bold p,\bold q \in {\EuScript G}^m, m \geq 0) \). Then ${\frak H (\EuScript G)}$ becomes a $\EuScript V$-face algebra by setting \begin{equation} {\stackrel{\scriptscriptstyle\circ}{e}_i} = \sum_{j \in \EuScript V}e{i \choose j}, \quad e_j = \sum_{i \in \EuScript V}e{i \choose j}, \label{eHGi} \end{equation} \begin{equation} e{\bold p \choose \bold q} e{\bold r \choose \bold s} = {\delta}_{\frak {r}(\bold p) \frak {s}(\bold r)} \, {\delta}_{\frak {r}(\bold q) \frak {s}(\bold s)} \: e{\bold p\cdot\bold r \choose \bold q\cdot\bold s}, \label{epqers} \end{equation} \begin{equation} \Delta \left( e{\bold p \choose \bold q} \right) = \sum_{{\bold t} \in {\EuScript G}^m} e{\bold p \choose {\bold t}} \otimes e{{\bold t} \choose \bold q}, \quad \varepsilon \left( e{\bold p \choose \bold q} \right) = {\delta}_{\bold p \bold q} \label{D(epq)} \end{equation} for each $\bold p,\bold q \in {\EuScript G}^m$ and $ \bold r,\bold s \in {\EuScript G}^n$ $( m,n \geq 0)$. Here for paths $\bold p = (\bold p_1,\ldots,\bold p_m)$ and $\bold r = (\bold r_1,\ldots,\bold r_n)$, we set $\bold p \cdot \bold r = (\bold p_1,\ldots,\bold p_m,\bold r_1,\ldots,\bold r_n)$ if $\frak {r}(\bold p) = \frak {s}(\bold r)$ and $m, n \geq 1$, and also, we set $\frak {s} (\bold p) \cdot \bold p = \bold p = \bold p \cdot \frak {r} (\bold p)$ for each $\bold p \in \EuScript G^m$ $(m \geq 0)$. We say that a quadruple $\left( \bold r \frac[0pt]{\bold p}{\bold q} \bold s \right)$ is a {\it face} if $\bold p,\bold q, \bold r, \bold s \in {\EuScript G}^1$ and \begin{equation} \frak {s} (\bold p) = \frak {s} (\bold r), \quad \frak {r} (\bold p) = \frak {s} (\bold s), \quad \frak {r} (\bold r) = \frak {s} (\bold q), \quad \frak {r} (\bold q) = \frak {r} (\bold s). \label{facecond} \end{equation} We say that $(\EuScript G, w)$ is a {\it face model} (or $\EuScript V$-{\it face model} ) over ${\Bbb K}$ if $w$ is a map which assigns a scalar $w \!\! \left[ \bold r \frac[0pt]{\bold p}{\bold q} \bold s \right] \in {\Bbb K}$ to each face $\left( \bold r \frac[0pt]{\bold p}{\bold q} \bold s \right)$ of $\EuScript G$. A face model $(\EuScript G, w)$ is called a {\em vertex model} if $\#(\EuScript V) = 1$. For convenience, we set $w \!\! \left[ \bold r \frac[0pt]{\bold p}{\bold q} \bold s \right] = 0$ unless $\bold p, \bold q, \bold r, \bold s \in \EuScript G^1$ satisfy \eqref{facecond}. For a face model $(\EuScript G,w)$, we identify $w$ with the linear operator on $\Bbb K \EuScript G^2 := \bigoplus_{\bold p \in \EuScript G^2} \Bbb K \bold p$ given by \begin{equation} \label{w(pq)} w(\bold p, \bold q) = \sum_{(\bold r, \bold s) \in {\EuScript G}^2} w \!\! \left[ \bold r \frac[0pt]{\bold p}{\bold s} \bold q \right] (\bold r, \bold s) \quad ( (\bold p, \bold q) \in \EuScript G^2). \end{equation} A face model is called {\it invertible} if $w$ is invertible as an operator on $\Bbb K \EuScript G^2$. For an invertible face model $(\EuScript G, w)$, we define another face model $(\EuScript G, w^{-1})$, using the identification \eqref{w(pq)}. An invertible face model is called {\it star-triangular} (or {\it Yang-Baxter}) if $w$ satisfies the braid relation $w_1 w_2 w_1 = w_2 w_1 w_2$, where $w_1$ and $w_2$ denote linear operators on $\Bbb K \EuScript G^3$ defined by $w_1(\bold p, \bold q, \bold r) = w(\bold p, \bold q) \otimes \bold r$ and $w_2(\bold p, \bold q, \bold r) = \bold p \otimes w(\bold q, \bold r)$. Here we identify $(\bold p, \bold q, \bold r) \in \EuScript G^3$ with $\bold p \otimes \bold q \otimes \bold r \in (\Bbb K \EuScript G^1)^{\otimes 3}$. For a face model $(\EuScript G, w)$, we define the algebra $\frak A (\EuScript G, w)= {\frak A}(w)$ to be the quotient of $\frak H (\EuScript G)$ modulo the following relations: \begin{equation} \sum_{(\bold c, \bold d) \in {\EuScript G}^2} w \!\! \left[ \bold a \frac[0pt]{\bold c}{\bold b} \bold d \right] e{\bold c \cdot \bold d \choose \bold p \cdot \bold q} = \sum_{(\bold r, \bold s) \in {\EuScript G}^2} w \!\! \left[ \bold r \frac[0pt]{\bold p}{\bold s} \bold q \right] e{\bold a \cdot \bold b \choose \bold r \cdot \bold s} \quad ( (\bold p, \bold q),\, (\bold a, \bold b) \in {\EuScript G}^2). \label{relAw} \end{equation} Then ${\frak A}(w)$ has a unique structure of $\EuScript V$-face algebra such that the projection $\frak H (\EuScript G) \to {\frak A}(w)$ is a map of $\EuScript V$-face algebras. For each $n \geq 0$, $\Bbb K \EuScript G^n$ becomes a comodule of $\frak{A}_n (w)$ via $\rho (\bold q) = \sum_{\bold p \in \EuScript G^n} \bold p \otimes e\!\binom{\bold p}{\bold q}$, where the subcoalgebra $\frak{A}_n (w)$ of ${\frak A}(w)$ is defined as the linear span of the elements of the form $e\!\binom{\bold p}{\bold q}$ $(\bold p, \bold q \in \EuScript G^n)$. If $(\EuScript G,w)$ is star-triangular, then there exist unique bilinear pairings ${\cal R}^{\pm}$ on ${\frak A}(w)$ such that $({\frak A}(w), {\cal R}^{\pm})$ is a CQT $\EuScript V$-face algebra and that \begin{equation} {\cal R}^+ \left( e{\bold p \choose \bold q},\; e{\bold r \choose \bold s} \right) = w \!\! \left[ \bold r \frac[0pt]{\bold q}{\bold p} \bold s \right] \label{UnivR=W} \end{equation} for each $\bold p,\bold q, \bold r, \bold s \in {\EuScript G}^1$ (cf. \cite{LarsonTowber}, \cite{gd}, \cite{Schauenburg}, \cite{fb}). We call ${\cal R}^+$ the {\it canonical braiding} of ${\frak A}(w)$. For a vertex model $w = \Check{R}$, ${\frak A}(w)$ coincides with FRT bialgebra $A_R$, where $R = P \Check{R}$ and $P(\bold p, \bold q) = (\bold q, \bold p)$. Let $\tilde{\EuScript G}$ be the orientation-reversed graph of $\EuScript G$ and let $\tilde{}: \EuScript G^m \to \tilde{\EuScript G}^m \, ;\; \bold p \mapsto \tilde{\bold p}$ $(m \geq 0)$ be the canonical bijection which satisfies $\widetilde{\bold p \cdot \bold q} = \tilde{\bold q} \cdot \tilde{\bold p}$, $\frak {s}(\tilde{\bold p}) = \frak {r}({\bold p})$ and $\frak {r}(\tilde{\bold p}) = \frak {s}({\bold p})$. We also define a new graph $\G_{\mathrm{LD}}$ by setting $\G_{\mathrm{LD}}^0 = \EuScript V$ and $\G_{\mathrm{LD}}^1 = \EuScript G^1 \coprod \tilde{\EuScript G}^1$. Let $\EuScript G \bar{\times} \tilde{\EuScript G}$ and $\tilde{\EuScript G} \bar{\times} \EuScript G$ denote subsets of $\G_{\mathrm{LD}}^2$ consisting of elements of the form $\bold p \cdot \tilde{\bold q}$ and $\tilde{\bold p} \cdot \bold q$ $(\bold p,\bold q \in \EuScript G^1)$ respectively. We define linear operators $w_{\mathrm{LD}}, w_{\mathrm{LD}}^-\!: \Bbb K (\tilde{\EuScript G} \bar{\times} \EuScript G) \to \Bbb K (\EuScript G \bar{\times} \tilde{\EuScript G})$ by \begin{equation} w_{\mathrm{LD}} (\tilde{\bold p} \cdot \bold q) = \sum_{\bold r,\bold s} w_{\mathrm{LD}} \!\! \left[ \bold r \frac[0pt]{\tilde{\bold p}}{\tilde{\bold s}} \bold q \right] \bold r \cdot \tilde{\bold s}\, ; \quad w_{\mathrm{LD}} \!\! \left[ \bold r \frac[0pt]{\tilde{\bold p}}{\tilde{\bold s}} \bold q \right] =: w^{-1} \!\! \left[ \bold p \frac[0pt]{\bold q}{\bold r} \bold s \right], \end{equation} \begin{equation} w_{\mathrm{LD}}^- (\tilde{\bold p} \cdot \bold q) = \sum_{\bold r,\bold s} w_{\mathrm{LD}}^- \!\! \left[ \bold r \frac[0pt]{\tilde{\bold p}}{\tilde{\bold s}} \bold q \right] \bold r \cdot \tilde{\bold s} \, ; \quad w_{\mathrm{LD}}^- \!\! \left[ \bold r \frac[0pt]{\tilde{\bold p}}{\tilde{\bold s}} \bold q \right] =: w \!\! \left[ \bold p \frac[0pt]{\bold q}{\bold r} \bold s \right]. \end{equation} We say that a star-triangular $\EuScript V$-face model $(\EuScript G,w)$ is {\it closable} if both $w_{\mathrm{LD}}$ and $w_{\mathrm{LD}}^-$ are invertible. In this case, we define a new $\EuScript V$-face model $(\G_{\mathrm{LD}},w_{\mathrm{LD}})$ by extending $w_{\mathrm{LD}}$ on $\Bbb K \G_{\mathrm{LD}}^2$ via $w_{\mathrm{LD}} |_{\Bbb K \EuScript G^2} = w$, $w_{\mathrm{LD}} |_{\Bbb K (\EuScript G \bar{\times} \tilde{\EuScript G})} = (w_{\mathrm{LD}}^-)^{-1}$ and \begin{equation} w_{\mathrm{LD}} \!\! \left[ \tilde{\bold r} \frac[0pt]{\tilde{\bold p}}{\tilde{\bold q}} \tilde{\bold s} \right] = w \!\! \left[ \bold s \frac[0pt]{\bold q}{\bold p} \bold r \right] \quad (\bold p, \bold q, \bold r, \bold s \in \EuScript G^1). \end{equation} We call $w_{\mathrm{LD}}$ the {\it Lyubashenko double} of $w$. As in case $(\EuScript G,w)$ is a vertex model, $(\G_{\mathrm{LD}},w_{\mathrm{LD}})$ is a star-triangular face model. Let $\frak H$ be a $\EuScript V$-face algebra and $U$ its (right) comodule. We define its {\it face space decomposition} $U = \bigoplus_{i,j \in {\EuScript V}}U(i,j)$ by $U(i,j) = \pi_U ({\stackrel{\scriptscriptstyle\circ}{e}}_i e_j) (U)$. Let $V$ be another $\frak H$-comodule. We define the {\it truncated tensor product} $U \overline{\otimes} V$ to be the vector space \begin{equation} U \overline{\otimes} V = {\bigoplus}_{i,j,k \in {\EuScript V}}U(i,k) \otimes V(k,j) \end{equation} equipped with the $\frak H$-comodule structure given by \begin{equation} \rho_{U \bar{\otimes} V} (\overline{u \otimes v} ) = \sum_{(u),(v)} \left( u_{(0)} \otimes v_{(0)} \right) \otimes u_{(1)}v_{(1)}, \end{equation} where $\bar{}\!: U \otimes V \to U \bar{\otimes} V$ denotes the projection $\sum_k \pi_U (e_k) \otimes \pi_V ({\stackrel{\scriptscriptstyle\circ}{e}_k}) $. For $\frak H$-comodules $U, U^{\prime},$ $V, V^{\prime}$ and maps $f \in \mathrm{End}_{\pi (\frak{E})} (U, U^{\prime})$, $g \in \mathrm{End}_{\pi (\frak{E})} (V, V^{\prime})$, we set \begin{equation} f \bar{\otimes} g = (f \otimes g) |_{U \overline{\otimes} V}, \end{equation} where $\frak{E} = \frak{E}_{\frak H^{\circ}}$. If both $f$ and $g$ are comodule maps, then so is $f \bar{\otimes} g$. The category $\bold{Com}_{\frak H}$ of all $\frak H$-comodules becomes a monoidal category via $\bar{\otimes}$ and the category $\bold{Com}^f_{\frak H}$ of all finite-dimensional $\frak H$-comodules becomes its sub monoidal category. The category $\bold{Com}^f_{\frak H}$ is rigid if and only if $\frak H$ has a bijective antipode. Next, suppose $\frak H$ has a braiding ${\cal R}^{\pm}$. Then $\bold{Com}_{\frak H}$ and $\bold{Com}^f_{\frak H}$ become braided categories via the functorial isomorphism $c_{UV}\!: U \overline{\otimes} V$ $\cong V \overline{\otimes} U$ given by \begin{equation} c_{UV} (\overline{u \otimes v}) = \sum_{(u),(v)} v_{(0)} \otimes u_{(0)} {\cal R}^+(u_{(1)},v_{(1)}). \end{equation} If, in addition, $\frak H$ has a ribbon functional $\cal{V}$, then $\bold{Com}^f_{\frak H}$ becomes a ribbon category (see e.g. \cite{Kassel}) via twist $\theta_{U}\!: U \cong U$ given by $\theta_{U} = \pi_U (\cal{V}^{-1})$. Conversely, we have the following. \begin{prop}[\cite{LarsonTowber}, \cite{Kassel}] \label{correspalgcom} Let $\frak H$ be a $\EuScript V$-face algebra such that either $\bold{Com}_{\frak H}$ or $\bold{Com}^f_{\frak H}$ is a braided monoidal category with braiding $\{ c_{UV} \}$. Then, $\frak H$ becomes a CQT $\EuScript V$-face algebra via % \begin{gather} \label{Rp=eec} {\cal R}^+ (a, b) = \sum_{k,l \in \V} ( \varepsilon \otimes \varepsilon ) \circ c_{LM} ({\stackrel{\scriptscriptstyle\circ}{e}_k} a e_l \otimes e_l b {\stackrel{\scriptscriptstyle\circ}{e}_k} ), \\ \label{Rm=eec} {\cal R}^- (b, a) = \sum_{k,l \in \V} ( \varepsilon \otimes \varepsilon ) \circ (c_{ML})^{-1} ({\stackrel{\scriptscriptstyle\circ}{e}_k} a e_l \otimes e_l b {\stackrel{\scriptscriptstyle\circ}{e}_k} ), \end{gather} % where $L$ and $M$ denote arbitrary finite-dimensional sub $\frak H$-comodules of $\frak H$ such that ${\stackrel{\scriptscriptstyle\circ}{e}_i} a e_j \in \L$, $e_j b {\stackrel{\scriptscriptstyle\circ}{e}_i} \in M$ $(i, j \in \EuScript V)$. % If, in addition, $\frak H$ has an antipode and $\bold{Com}^f_{\frak H}$ is a ribbon category with twist $\{ \theta_{U} \}$, then $\frak H$ becomes a coribbon Hopf $\EuScript V$-face algebra via % \begin{equation} \label{V=etheta-} \cal{V}^{\pm 1} (a) = \varepsilon ( \theta_L^{\mp 1} (a) ). \end{equation} % \end{prop} \begin{pf} To begin with, we note that the existence of such $L$ and $M$ follows from the fundamental theorem of coalgebras, and that \eqref{Rp=eec}-\eqref{V=etheta-} do not depend on the choice of $L$ and $M$ because of the naturality of $c$ and $\theta$. Here, we will give a proof of the last assertion. Let $\cal{V}^{\pm} \in \frak H^*$ be as in \eqref{V=etheta-} and let $U$ be a finite-dimensional $\frak H$-comodule. For each $u^* \in U^*$, we define the $\frak H$-comodule map $F_{u^*}\!:U \to \frak H$ by $F_{u^*}(u)$ $=$ $\sum_{(u)} \langle u^*, u_{(0)} \rangle u_{(1)}$ $(u \in U)$. Then, we have % \begin{align} \label{u*thetam} \langle u^*,\, \theta_U^{\mp 1} (u) \rangle = & \varepsilon \circ F_{u^*} \circ \theta_U^{\mp 1} (u) = \varepsilon \circ \theta_{\mathrm{Im} (F_{u^*})}^{\mp 1} \circ F_{u^*} (u) \\ = & \langle u^*,\, \pi_U (\cal{V}^{\pm 1}) u \rangle, \nonumber \end{align} % or equivalently, \begin{equation} \label{theta=piV-} \theta_U^{\mp 1} (u) = \pi_U (\cal{V}^{\pm 1}) u. \end{equation} % Rewriting $\langle u^*,\, \theta_U (u) \rangle$ $=$ $\langle \theta_{U^{\lor}} (u^*), u \rangle$ via this equality, we obtain $S (\cal{V}) = \cal{V}$. Let $a$ and $b$ elements of $\frak H$ and let $L$ and $M$ be as above. Since $\overline{a \otimes b} = \sum_k a e_k \otimes e_k b$ by \eqref{e(ab)}-\eqref{ae(eae)}, we have % \begin{equation} \pi_{L \bar{\otimes} M } (\cal{V}) (\overline{a \otimes b}) = \sum_{(a), (b)} a_{(1)} \otimes b_{(1)} \langle \cal{V}, a_{(2)} b_{(2)} \rangle \end{equation} % by \eqref{D(eeaee)}. Using \eqref{theta=piV-} and the equality $\theta_{L \overline{\otimes} M}^{-1}$ $=$ $c_{L \overline{\otimes} M}^{-1} \circ c_{M \overline{\otimes} L}^{-1} \circ (\theta_{L}^{-1} \bar{\otimes} \theta_{M}^{-1})$, we see that the left-hand side of the above equality is % \begin{align} \label{cc(ab)VaVb} &\sum_{(a), (b)} c_{L \overline{\otimes} M}^{-1} \circ c_{M \overline{\otimes} L}^{-1} (\overline{a_{(1)} \otimes b_{(1)}}) \cal{V} (a_{(2)}) \cal{V} (b_{(2)}) \\ = & \sum_{(a), (b)} a_{(1)} \otimes b_{(1)} {\cal R}^- (a_{(2)}, b_{(2)}) {\cal R}^- (b_{(3)}, a_{(3)}) \cal{V} (a_{(4)}) \cal{V} (b_{(4)}), \end{align} where \eqref{cc(ab)VaVb} follows from the fact that $\theta_L$ and $\theta_M$ commute with the action of the face idempotents of $\frak H^{\circ}$. Taking the image via $\varepsilon \otimes \varepsilon$, we get \eqref{m(V)}. \end{pf} Let $U$ be a finite-dimensional comodule of a CQT $\EuScript V$-face algebra $\frak H$. For each $i,j \in \EuScript V$, choose a basis $\EuScript G^1_{ij}$ of $U (i,j)$. Let $\EuScript G$ be the oriented graph with set of vertexes $\EuScript V$ and the set of edges $\EuScript G^1: = \coprod_{ij} \EuScript G^1_{ij}$. Then we obtain a star-triangular $\EuScript V$-face model $(\EuScript G, w_U)$ be setting \begin{equation} \label{wUdef} c_{UU} \left( \bold p \otimes \bold q \right) = \sum_{(\bold r,\bold s) \in \EuScript G^2} w_U \!\! \left[ \bold r \frac[0pt]{\bold p}{\bold s} \bold q \right] \bold r \otimes \bold s. \quad ((\bold p, \bold q ) \in \EuScript G^2). \end{equation} \section{Drinfeld functionals and the Hopf closure} Let $(\frak H,{\cal R}^{\pm})$ be a CQT ${\EuScript V}$-face algebra. We say that ${\frak H}$ is {\em closable} (or ${\frak H}$ is a {\em CCQT} ${\EuScript V}$-face algebra) if there exist both $(\cal{F}^+, \cal{F}^-)$-generalized inverse ${\cal Q}^-$ of ${\cal R}^+$ and $(\cal{F}^-, \cal{F}^+)$-generalized inverse ${\cal Q}^+$ of ${\cal R}^-$ in the algebra $({\frak H}\otimes{\frak H}^{\mathrm{cop}})^*$, where $\cal{F}^{\pm}$ denote bilinear forms on $\frak H$ defined by \begin{equation} \cal{F}^+ (a,\, b) = \sum_{k \in \V} \varepsilon (e_k a) \varepsilon(e_k b), \quad \cal{F}^- (a,\, b) = \sum_{k \in \V} \varepsilon (a e_k) \varepsilon(b e_k) \quad (a,b \in \frak H). \end{equation} We call ${\cal Q}^{\pm}$ {\em Lyubashenko forms} of $\frak H$. The Lyubashenko forms of a CQT $\EuScript V$-face algebra are unique if they exist. If $\frak H$ has an antipode, then $\frak H$ is closable with Lyubashenko forms given by \begin{equation} \label{Qp=Rm} {\cal Q}^+(a,b) = {\cal R}^- (S(a), b), \quad {\cal Q}^-(a,b) = {\cal R}^+ (a, S(b)) \quad (a,b \in \frak H). \end{equation} For a star-triangular face model $(\EuScript G,w)$, ${\frak A}(w)$ is closable if and only if $(\EuScript G,w)$ is closable. In this case, Lyubashenko forms ${\cal Q}^{\pm}$ of ${\frak A}(w)$ satisfy \begin{equation} {\cal Q}^+ \left( e{\bold p \choose \bold q},\; e{\bold r \choose \bold s} \right) = w_{\mathrm{LD}}^{-1} \!\! \left[ \tilde{\bold q} \, \frac[0pt]{\bold s}{\bold r} \, \tilde{\bold p} \right], \quad {\cal Q}^- \left( e{\bold p \choose \bold q},\; e{\bold r \choose \bold s} \right) = w_{\mathrm{LD}} \!\! \left[ \tilde{\bold s} \, \frac[0pt]{\bold q}{\bold p} \, \tilde{\bold r} \right] \label{Qp=W} \end{equation} for each $\bold p,\bold q, \bold r,\bold s \in {\EuScript G}^1$. For a CCQT ${\EuScript V}$-face algebra ${\frak H}$, we define linear functionals ${\cal U}_{\nu}$ $(\nu = 1,2)$ on ${\frak H}$ via \begin{equation} {\cal U}_1 (a) = \sum_{(a)} {\cal Q}^- (a_{(2)},a_{(1)}), \quad {\cal U}_2 (a) = \sum_{(a)} {\cal Q}^+ (a_{(1)},a_{(2)}) \quad (a \in \frak H) \label{Udef} \end{equation} and call them {\em Drinfeld functionals} of ${\frak H}$. The Drinfeld functionals of a CCQT $\EuScript V$-face algebra $\frak H$ are invertible in ${\frak H}^*$ and satisfy the following relations\rom{:} \begin{equation} \label{U-} {\cal U}_1^{-1}(a) = \sum_{(a)} {\cal Q}^+ (a_{(2)},a_{(1)}), \quad {\cal U}_2^{-1}(a) = \sum_{(a)} {\cal Q}^- (a_{(1)},a_{(2)}), \end{equation} \begin{equation} \label{UURp} ({\cal U}_{\nu} \otimes {\cal U}_{\nu}) {\cal R}^{\pm} = {\cal R}^{\pm} ({\cal U}_{\nu} \otimes {\cal U}_{\nu}), \end{equation} \begin{equation} \label{U(ee)} {\cal U}_{\nu}^{\pm 1}({\stackrel{\scriptscriptstyle\circ}{e}_i} e_j) = {\delta}_{ij}, \quad {\cal U}_{\nu}^{\pm}({\stackrel{\scriptscriptstyle\circ}{e}_i} a {\stackrel{\scriptscriptstyle\circ}{e}_j}) = {\cal U}_{\nu}^{\pm}(e_i a e_j), \end{equation} \begin{equation} \label{U1U2} {\cal U}_1 {\cal U}_2 = {\cal U}_2 {\cal U}_1, \end{equation} \begin{equation} \label{m(U)} m^*({\cal U}) = {\cal R}^- {\cal R}^-_{21} ({\cal U} \otimes {\cal U}) = ({\cal U} \otimes {\cal U}) {\cal R}^- {\cal R}^-_{21}, \end{equation} \begin{equation} m^*({\cal U}^{-1}) = {\cal R}^+_{21} {\cal R}^+ ({\cal U} \otimes {\cal U} )^{-1} = ({\cal U} \otimes {\cal U} )^{-1} {\cal R}^+_{21} {\cal R}^+ \label{m(Um)} \end{equation} for each $\nu = 1,2$, $a \in \frak H$ and $i,j \in \EuScript V$, where $\cal{U}$ stands for $\cal{U}_1$ or $\cal{U}_2^{-1}$. % % % % % Let $f\!: \frak H \to \frak K$ be a map of CQT $\EuScript V$-face algebras. If $\frak K$ is closable with Lyubashenko forms ${\cal Q}^{\pm}_{\frak K}$ and Drinfeld functionals ${\cal{U}_{\nu}}_{\frak K}$, then $\frak H$ is also closable with Lyubashenko forms and Drinfeld functionals given by \begin{equation} \label{f(U)=U} {\cal Q}^{\pm}_{\frak H} = (f \otimes f)^* ({\cal Q}^{\pm}_{\frak K}), \quad {\cal{U}_{\nu}}_{\frak H} = f^* ({\cal{U}_{\nu}}_{\frak K}). \end{equation} Next, we recall the {\it Hopf closure} (or {\it Hopf envelope}) construction of CQT Hopf $\EuScript V$-face algebras. It is introduced by Phung Ho Hai \cite{Phung} for bialgebras, and independently, by \cite{gsg} for face algebras. Let $\frak H$ be a CCQT $\EuScript V$-face algebra. We denote by ${\frak H}^{\mathrm{bop}}$ its biopposite $\EuScript V$-face algebra, that is, ${\frak H}^{\mathrm{bop}}$ is a $\EuScript V$-face algebra equipped with the opposite product and the opposite coproduct of $\frak H$ together with the face idempotents \begin{equation} {\stackrel{\scriptscriptstyle\circ}{e}}_{\frak H^{\mathrm{bop}},i} = e_{\frak H,i}, \quad e_{{\frak H}^{\mathrm{bop}},i} = {\stackrel{\scriptscriptstyle\circ}{e}}_{{\frak H},i}. \end{equation} Let $\sigma\! : \frak H \to {\frak H}^{\mathrm{bop}}$ be the canonical anti-isomorphism, which satisfies \begin{equation} \sigma(e_{\frak H,i}) = {\stackrel{\scriptscriptstyle\circ}{e}}_{{\frak H}^{\mathrm{bop}},i}, \quad \sigma(\stackrel{\scriptscriptstyle\circ}{e}_{\frak H,i}) = e_{{\frak H}^{\mathrm{bop}},i} \quad (i \in \EuScript V). \end{equation} Then \begin{equation} \hat{\frak H} := \frak H \otimes_{\frak{E}} \frak H^{\mathrm{bop}} = \bigoplus_{k,l \in \EuScript V} \frak H {\stackrel{\scriptscriptstyle\circ}{e}}_k e_l \otimes \sigma (\frak H {\stackrel{\scriptscriptstyle\circ}{e}_l} e_k) \end{equation} becomes a $\EuScript V$-face algebra by setting \begin{equation} (a {\otimes}_{\frak E} \sigma(b))(c {\otimes}_{\frak E} \sigma(d)) = \sum_{(b),(c)} {\cal R}^- (b_{(1)},c_{(3)}) {\cal Q}^+ (b_{(3)},c_{(1)}) a c_{(2)} {\otimes}_{\frak E} \sigma(d b_{(2)}), \label{asbcsd} \end{equation} \begin{equation} \Delta (a {\otimes}_{\frak E} \sigma(b)) = \sum_{(a),(b)} (a_{(1)} {\otimes}_{\frak E} \sigma(b_{(2)})) \otimes (a_{(2)} {\otimes}_{\frak E} \sigma(b_{(1)})), \end{equation} \begin{equation} \varepsilon (a {\otimes}_{\frak E} \sigma(b)) = \sum_{k \in \V} {\varepsilon (a e_k) \varepsilon (b e_k)}, \end{equation} \begin{equation} e_{\hat{\frak H},i} = e_{\frak H,i} {\otimes}_{\frak E} \sigma(1_{\frak H}), \quad \stackrel{\scriptscriptstyle\circ}{e}_{\hat{\frak H},i} = \stackrel{\scriptscriptstyle\circ}{e}_{\frak H,i} {\otimes}_{\frak E} \sigma(1_{\frak H}) \end{equation} for each $a,b,c,d \in \frak H$ and $i \in \EuScript V$, Let ${\frak J}$ be the ideal of $\hat{\frak H}$ generated by all elements of the form: \begin{equation} \sum_{(a)} (1 {\otimes}_{\frak E} \sigma(a_{(1)}))(a_{(2)} {\otimes}_{\frak E} 1) - \sum_{k \in \V} \varepsilon (a e_k) e_k, \label{Idef1} \end{equation} \begin{equation} \sum_{(a)} a_{(1)} {\otimes}_{\frak E} \sigma(a_{(2)}) - \sum_{k \in \V} \varepsilon (e_k a) {\stackrel{\scriptscriptstyle\circ}{e}_k} \quad (a \in \frak H). \label{Idef2} \end{equation} It is easy to verify that $\frak{J}$ becomes a biideal. We denote the quotient $\EuScript V$-face algebra $\hat{\frak H} / {\frak J}$ by $\mathrm{Hc}({\frak H})$ and call it the {\em Hopf closure} of $\frak H$. For simplicity, we denote an element $a {\otimes}_{\frak E} \sigma(b) + \frak{J}$ of $\mathrm{Hc}({\frak H})$ by $a \sigma(b)$ for each $a,b \in \frak H$. The Hopf closure $\mathrm{Hc}({\frak H})$ has a unique structure of CQT Hopf $\EuScript V$-face algebra such that the canonical map $\iota\!: \frak H \to \mathrm{Hc}({\frak H});$ $a \mapsto a {\otimes}_{\frak E} 1 + \frak{J}$ $(a \in \frak H)$ is a map of CQT $\EuScript V$-face algebras. Explicitly, the antipode of $\mathrm{Hc}({\frak H})$ is given by \begin{align} \label{SHcH} S(a \sigma(b)) = \sum_{(b)} \cal{U}_{\nu} (b_{(1)}) b_{(2)} \sigma (a) \cal{U}_{\nu}^{-1} (b_{(3)}) \quad (\nu = 1,2). \end{align} When $\frak H$ is a bialgebra, the underlying Hopf algebra of $\mathrm{Hc}({\frak H})$ agrees with the Hopf envelope of $\frak H$ in the sense of Manin \cite{Manin}. The Hopf closure has the following universal mapping property. \begin{thm} \label{UMP} Let $\frak H$ be a CCQT $\EuScript V$-face algebra and $\frak K$ a CQT Hopf $\EuScript V$-face algebra. Let $f\!: \frak H \to \frak K$ be a map of CQT $\EuScript V$-face algebras. Then there exists a unique map $\bar{f}\!:\mathrm{Hc}({\frak H}) \to \frak K$ of CQT $\EuScript V$-face algebras such that $f = \bar{f} \circ \iota$, where $\iota\!: \frak H \to \mathrm{Hc}({\frak H})$ is given by $\iota (a) = a {\otimes}_{\frak E} 1 + \frak{J}$ $(a \in \frak H)$. Explicitly, we have % \begin{equation} \label{barf(asb)} \bar{f} (a \sigma (b)) = f (a) S(f (b)). \end{equation} % \end{thm} \begin{prop} \label{univA} Let $\frak H$ be a CQT $\EuScript V$-face algebra \rom{(}resp. CQT Hopf $\EuScript V$-face algebra\rom{)} and $U$ its finite-dimensional comodule. Let $(\EuScript G, w_U)$ be a face model given by \eqref{wUdef}. Then there exists a unique map $f\!: \frak{A}(w_U) \to \frak H$ \rom{(}resp. $f\!: \mathrm{Hc}(\frak{A}(w_U)) \to \frak H$\rom{)} of CQT $\EuScript V$-face algebras such that $(\mathrm{id}_{\Bbb K \EuScript G^1} \otimes f) \circ \rho_{\frak H}$ $=$ $\rho_{\frak{A}(w_U)}$ \rom{(}resp. $(\mathrm{id}_{\Bbb K \EuScript G^1} \otimes f) \circ \rho_{\frak H}$ $=$ $\rho_{\mathrm{Hc}(\frak{A}(w_U))}$\rom{)}. \end{prop} \begin{pf} See \cite{fb} for a proof of the assertion for $\frak{A}(w_U)$. The assertion for $\mathrm{Hc}( \frak{A}(w_U) )$ follows from that of $\frak{A}(w_U)$ and the universal mapping property of $\mathrm{Hc}$. \end{pf} % \begin{prop} For each CQT Hopf $\EuScript V$-face algebra ${\frak H}$, we have\rom{:} % \begin{equation} \label{S(U)} S^*({\cal U}_1^{\pm 1}) = {\cal U}_2^{\mp 1}, \quad S^*({\cal U}_2^{\pm 1}) = {\cal U}_1^{\mp 1}, \end{equation} % \begin{equation} \label{UXU-} {\cal U}_{\nu} X {\cal U}_{\nu}^{-1} =(S^2)^*(X) \quad (X \in {\frak H}^*,\, \nu = 1, 2). \end{equation} % In particular, $S$ is bijective and ${\cal U}_1{\cal U}_2^{-1}$ is a central element of ${\frak H}^*$. % \end{prop} \begin{pf} (cf. Drinfeld \cite{Drinfeld}). The relation \eqref{S(U)} follows from \eqref{Qp=Rm}, \eqref{Sid(R)} and \eqref{U-}. Substituting $\sum_{(c)} c_{(2)}$ $\otimes$ $S(c_{(1)})$ into ${\cal R}^+ m^* (X) = (m^{\mathrm{op}})^* (X) {\cal R}^+$, we obtain % \begin{align*} \sum_{(c)} \cal{U}_1 (c_{(2)}) c_{(3)} S(c_{(1)}) = & \sum_{(c)} S(c_{(2)}) c_{(3)} {\cal Q}^- (c_{(4)},c_{(1)}) \\ = & \sum_{(c)} \sum_{k \in \V} e_k {\cal Q}^- (c_{(2)} {\stackrel{\scriptscriptstyle\circ}{e}_k},c_{(1)}) \\ = & \sum_{k \in \V} \cal{U}_1 ({\stackrel{\scriptscriptstyle\circ}{e}_k}c) e_k, \end{align*} % where the second equality follows from \eqref{e(eae)a} and the third equality follows from \eqref{Rmeeee} and \eqref{D(eeaee)}. Using this relation, we compute % \begin{align*} \sum_{(c)} S^2 (c_{(1)}) \cal{U}_1 (c_{(2)}) = & \sum_{(c)} \sum_{k \in \V} \cal{U}_1 ({\stackrel{\scriptscriptstyle\circ}{e}_k}c_{(2)}) e_k S^2(c_{(1)}) \\ = & \sum_{(c)} \cal{U}_1 (c_{(3)}) c_{(4)} S(S(c_{(1)}) c_{(2)}) \\ = & \sum_{(c)} \sum_{k \in \V} \cal{U}_1 (c_{(1)}{\stackrel{\scriptscriptstyle\circ}{e}_k}) c_{(2)}{\stackrel{\scriptscriptstyle\circ}{e}_k} \\ = & \sum_{(c)} \cal{U}_1 (c_{(1)}) c_{(2)}, \end{align*} % where the first equality follows from \eqref{D(a)} and \eqref{S(ee)} and the last equality follows from \eqref{U(ee)} and \eqref{D(a)}. Substituting this into $X \in {\frak H}^*$, we get $((S^2)^*(X) \cal{U}_1)(c) = (\cal{U}_1 X)(c)$, which proves \eqref{UXU-} for $\nu = 1$. \end{pf} \section{Group-like functionals} Let $g$ be an element of a $\EuScript V$-face algebra $\frak H$. We say that $g$ is {\it group-like} if \begin{equation} \label{D(g)} \Delta (g) = \sum_{k \in \V} g e_k \otimes g {\stackrel{\scriptscriptstyle\circ}{e}_k}, \end{equation} \begin{equation} \label{gee} g {\stackrel{\scriptscriptstyle\circ}{e}_i} e_j = {\stackrel{\scriptscriptstyle\circ}{e}_i} e_j g, \quad \varepsilon (g {\stackrel{\scriptscriptstyle\circ}{e}_i} e_j) = \delta_{ij} \end{equation} for each $i,j \in \EuScript V$. We say that a linear functional ${\cal G}$ on $\frak H$ is {\it group-like} if it is group-like as an element of the dual face algebra $\frak H^{\circ}$. Explicitly, ${\cal G}$ is group-like if and only if it satisfies \begin{equation} \label{Gab} {\cal G}(ab) = \sum_{k \in \V} {\cal G} (ae_k) {\cal G}({\stackrel{\scriptscriptstyle\circ}{e}_k} b), \end{equation} \begin{equation} \label{Geae} {\cal G}({\stackrel{\scriptscriptstyle\circ}{e}_i}a{\stackrel{\scriptscriptstyle\circ}{e}_j}) = {\cal G}(e_i a e_j), \end{equation} \begin{equation} \label{Gee} {\cal G}({\stackrel{\scriptscriptstyle\circ}{e}_i}e_j) = \delta _{ij} \end{equation} for each $a,b \in \frak H$ and $i,j \in {\EuScript V}$. We say that $\cal{G}$ is {\it invertible} if it is invertible as an element of the dual algebra $\frak H^*$. We denote by $\mathrm{GLF} (\frak H)$ the set of all group-like functionals of $\frak H$, and by $\mathrm{GLF} (\frak H)^{\times}$ the set of all invertible group-like functionals. Note that \begin{equation} \mathrm{GLF} (\frak H) = \mathrm{Hom}_{\, \Bbb{K}\! - \! Alg} (\frak H, \Bbb{K}) \end{equation} if $\frak H$ is a bialgebra. \begin{lem} \label{GLFisfunc} \rom{(1)} The correspondence $\frak H \mapsto \mathrm{GLF} (\frak H)$ defines a contravariant functor from the category of $\EuScript V$-face algebras to the category of semigroups. \\ \rom{(2)} Let $\frak H$ be a $\EuScript V$-face algebra and $\frak{I}$ its biideal. Then the projection $p\!: \frak H \to \frak K = \frak H / \frak{I}$ gives % \begin{equation} p^*\!: \mathrm{GLF}( \frak K ) \cong \{ \cal{G} \in \mathrm{GLF}( \frak H )\, |\, \cal{G} (\frak{I}) = 0 \}. \end{equation} \rom{(3)} If $\frak H$ has an antipode, then $\mathrm{GLF} (\frak H) = \mathrm{GLF} (\frak H)^{\times}$ and % \begin{equation} \label{S(G)} S^*({\cal G}) = {\cal G}^{-1}. \end{equation} % for each ${\cal G} \in \mathrm{GLF} (\frak H)$. \end{lem} \begin{pf} The proof of Part (1) is straightforward. Taking the dual of $0 \to \frak{I} \to \frak H \to \frak K \to 0$, we obtain % \begin{equation} p^*\!: \frak K^* \cong \{ X \in \frak H^*\, |\, X (\frak{I}) = 0 \}. \end{equation} It is straightforward to verify that $\cal{M} \in \mathrm{GLF} (\frak K)$ if and only if $p^* (\cal{M}) \in \mathrm{GLF} (\frak H)$ for each $\cal{M} \in \frak K^*$. This proves Part (2). See \cite{cpt} Proposition 7.1 for a proof of Part (3). \end{pf} \begin{lem} \label{cqtglf} Let $\frak H$ be a CQT $\EuScript V$-face algebra. \\ \rom{(1)} For each group-like functional $\cal{G}$ on $\frak H$, we have % \begin{equation} \label{GGRpm} (\cal{G} \otimes \cal{G}) {\cal R}^{\pm} = {\cal R}^{\pm} (\cal{G} \otimes \cal{G}). \end{equation} % Hence, for each $\frak H$-comodules $U$ and $V$, we have % \begin{equation} \label{GGcMN} (\pi_{V} (\cal{G}) \bar{\otimes} \pi_{U} (\cal{G})) c_{UV} = c_{UV} (\pi_{U} (\cal{G}) \bar{\otimes} \pi_{V} (\cal{G})). \end{equation} % \rom{(2)} If $\frak H$ is closable, then % \begin{equation} \label{U1U2isglf} \cal{U}_1 \cal{U}_2 \in \mathrm{GLF} (\frak H). \end{equation} % \end{lem} \begin{pf} Since ${\cal R}^+ = (m^{\mathrm{op}})^* (1) {\cal R}^+$, we have $(\cal{G} \otimes \cal{G}) {\cal R}^+$ $=$ $(m^{\mathrm{op}})^* (\cal{G}) {\cal R}^+$. Hence the first assertion of Part (1) follows from \eqref{RmXR}. The second assertion follows from the first assertion. Part (2) follows from \eqref{U(ee)} and \eqref{m(U)}-\eqref{m(Um)}. % \end{pf} Let $\frak H$ be a $\EuScript V$-face algebra and $\cal{G}$ its group-like functional. We define $\mathrm{coad}(\cal{G})\!: \frak H \to \frak H$ by \begin{equation} \label{coaddef} \mathrm{coad}(\cal{G}) (a) = \sum_{(a)} \cal{G}^{-1} (a_{(1)}) a_{(2)} \cal{G} (a_{(3)}) \quad (a \in \frak H). \end{equation} Using \eqref{Gab}-\eqref{Gee} and \eqref{eae*a}, we see that $\mathrm{coad}(\cal{G})$ is an automorphism of $\frak H$. \begin{prop} \label{GLFAwiso} For $\frak H =$ $\frak{H} (\EuScript G)$ or ${\frak A}(w)$, the map $\frak H^* \to \mathrm{End} (\Bbb K \EuScript G^1);$ $X \mapsto \pi_{\Bbb K \EuScript G^1} (X)$ gives the following semigroup isomorphisms\rom{:} \begin{equation} \label{GLFHG} \mathrm{GLF} (\frak{H} (\EuScript G)) \cong \mathrm{End}_{\pi (\frak{E})} (\Bbb K \EuScript G^1), \end{equation} \begin{equation} \label{GLFAw} \mathrm{GLF} ({\frak A}(w)) \cong \bigl\{ G \in \mathrm{End}_{\pi (\frak{E})} (\Bbb K \EuScript G^1) \bigm| (G \bar{\otimes} G) w = w (G \bar{\otimes} G) \bigr\}, \end{equation} where $\frak{E} = \frak{E}_{\frak H^{\circ}}$ is as in Sect. 2. \end{prop} \begin{pf} For each element $G$ of the right-hand side of \eqref{GLFHG}, we define a linear functional $\cal{G}$ $=$ $\cal{G}^{\frak H}_G$ on $\frak{H} (\EuScript G)$ by setting \begin{equation} \label{Gepq} \cal{G} \left( e \binom{i}{j} \right) = \delta_{ij}, \quad \cal{G} \left( e \binom{\bold p}{\bold q} \right) = G^{\bold p_1}_{\bold q_1} \cdots G^{\bold p_m}_{\bold q_m} \end{equation} for each paths $\bold p = (\bold p_1, \ldots, \bold p_m)$ and $\bold q = (\bold q_1, \ldots, \bold q_m)$ of length $m > 0$ and $i, j \in \EuScript V$. It is straightforward to verify that $\cal{G}$ is a group-like functional of $\frak{H} (\EuScript G)$. Hence $\pi_{\Bbb K \EuScript G^1}$ gives a surjection $\mathrm{GLF} (\frak{H} (\EuScript G)) \to \mathrm{End}_{\pi (\frak{E})} (\Bbb K \EuScript G^1)$. Conversely, for $\cal{G} \in \mathrm{GLF} (\frak{H} (\EuScript G))$, set $G = \pi_{\Bbb K \EuScript G^1} (\cal{G})$. Then by \eqref{Gab}-\eqref{Gee}, we have \eqref{Gepq}. Thus we get the isomorphism \eqref{GLFHG}. Next we show \eqref{GLFAw}. By \eqref{GGcMN}, $\pi_{\Bbb K \EuScript G^1}$ defines a well-defined map from $\mathrm{GLF} ({\frak A}(w))$ to the right-hand side of \eqref{GLFAw}. Hence it suffices to construct the inverse of this map. Let $G$ be an element of the right-hand side of \eqref{GLFAw} and let $\cal{G}^{\frak H}_G \in \mathrm{GLF} (\frak{H} (\EuScript G))$ be as above. By \eqref{GGcMN}, we have \begin{equation} \label{GrelAw} \cal{G}^{\frak H}_G \left( \sum_{(\bold c, \bold d) \in {\EuScript G}^2} w \!\! \left[ \bold a \frac[0pt]{\bold c}{\bold b} \bold d \right] e{\bold c \cdot \bold d \choose \bold p \cdot \bold q} - \sum_{(\bold r, \bold s) \in {\EuScript G}^2} w \!\! \left[ \bold r \frac[0pt]{\bold p}{\bold s} \bold q \right] e{\bold a \cdot \bold b \choose \bold r \cdot \bold s} \right) = 0 \end{equation} for each $(\bold p, \bold q),\, (\bold a, \bold b) \in {\EuScript G}^2$. By \eqref{Gab}, this shows that $\cal{G}^{\frak H}_G$ vanishes on the biideal $\mathrm{Ker} (\frak{H} (\EuScript G) \to {\frak A}(w))$ and that it induces an element of $\mathrm{GLF} ({\frak A}(w))$. This completes the proof of \eqref{GLFAw}. \end{pf} \begin{prop} \label{GLFHcHHiso} For each CCQT $\EuScript V$-face algebra, the canonical map $\iota\!: \frak H \to \mathrm{Hc}({\frak H})$ induces the isomorphism % \begin{equation} \iota^*\!: \mathrm{GLF} (\mathrm{Hc} (\frak H)) \cong \mathrm{GLF}^{\times} (\frak H), \end{equation} whose inverse $\cal{G} \mapsto \cal{G}_{\mathrm{Hc}}$ is given by % \begin{equation} \label{Gasb} \cal{G}_{\mathrm{Hc}} (a \sigma (b)) = \sum_{k \in \V} \cal{G}(a e_k) \cal{G}^{-1} (b e_k). \end{equation} \end{prop} \begin{pf}% By Lemma \ref{GLFisfunc} (1), it suffices to show that \eqref{Gasb} gives the inverse of the correspondence $\iota^*$. It is easy to verify that there exists a linear functional $\hat{\cal{G}} \in \hat{\frak H}^*$ which sends $a {\otimes}_{\frak E} \sigma (b)$ to the right-hand side of \eqref{Gasb} and that $\hat{\cal{G}}$ satisfies \eqref{Geae} and \eqref{Gee}. Using \eqref{Gab} for $\cal{G}^{\pm 1}$, we obtain % \begin{equation} \hat{\cal{G}} ((a {\otimes}_{\frak E} 1) x (1 {\otimes}_{\frak E} \sigma (d))) = \sum_{i,j \in \EuScript V} \cal{G} (a e_i) \hat{\cal{G}} ({\stackrel{\scriptscriptstyle\circ}{e}_i} x e_j) \cal{G}^{-1} (d e_j) \quad (a, d \in \frak H, x \in \hat{\frak H}). \end{equation} % By replacing $x$ with $(1 {\otimes}_{\frak E} \sigma (b))(c {\otimes}_{\frak E} \sigma(1))$, we obtain % \begin{multline} \label{Gasbcsd} \hat{\cal{G}} ( (a {\otimes}_{\frak E} \sigma (b))(c {\otimes}_{\frak E} \sigma(d)) ) = \sum_{i,j \in \EuScript V} \cal{G} (a e_i) \hat{\cal{G}} ( (1 {\otimes}_{\frak E} \sigma (b e_i))(c e_j {\otimes}_{\frak E} \sigma(1)) ) \cal{G}^{-1} (d e_j). \end{multline} On the other hand, using \eqref{Geae}, \eqref{eae*a} and \eqref{Rmeeee}, we obtain % \begin{multline} \hat{\cal{G}} ( (1 {\otimes}_{\frak E} \sigma (b))(c {\otimes}_{\frak E} \sigma(1))) = \sum_{k \in \V} \sum_{(b),(c)} {\cal R}^- (b_{(1)} e_k , c_{(3)} {\stackrel{\scriptscriptstyle\circ}{e}_k}) {\cal Q}^+ (b_{(3)}, c_{(1)}) \cal{G}^{-1} (b_{(2)}) \cal{G} (c_{(2)}) \\ = \sum_{(b),(c)} \langle (1 \otimes \cal{G}) {\cal R}^- (\cal{G}^{-1} \otimes 1),\, b_{(1)} \otimes c_{(2)} \rangle {\cal Q}^+ (b_{(2)}, c_{(1)}). \end{multline} By \eqref{GGRpm}, the right-hand side of the above equality is % \begin{multline} \sum_{(b),(c)} \cal{G}^{-1} (b_{(1)}) {\cal R}^- ( b_{(2)}, c_{(2)}) {\cal Q}^+ (b_{(3)}, c_{(1)}) \cal{G} (c_{(3)}) = \sum_{k \in \V} \cal{G}^{-1} (e_k b) \cal{G} (e_k c). \end{multline} Hence the right-hand side of \eqref{Gasbcsd} is % \begin{equation} \sum_{i,j,k \in \EuScript V} \cal{G} (a e_i) \cal{G}^{-1} (e_k a e_i) \cal{G} (e_k c e_j) \cal{G}^{-1} (d e_j) = \sum_{k \in \V} \hat{\cal{G}} ((a {\otimes}_{\frak E} \sigma (b)) e_k) \hat{\cal{G}} (e_k (c {\otimes}_{\frak E} \sigma (d))). \end{equation} % Thus $\hat{\cal{G}}$ is a group-like functional of $\hat{\frak H}$. Using \eqref{Gab} for $\hat{\cal{G}}$, we compute \begin{multline} \hat{\cal{G}} \left( {\stackrel{\scriptscriptstyle\circ}{e}_i} e_j \left( \sum_{(a)} (1 {\otimes}_{\frak E} \sigma (a_{(1)})) (a_{(2)} {\otimes}_{\frak E} 1) \right) {\stackrel{\scriptscriptstyle\circ}{e}_k} e_l \right) \\ = \sum_{m \in \EuScript V} \sum_{(a)} \cal{G}^{-1} ({\stackrel{\scriptscriptstyle\circ}{e}}_m a_{(1)} {\stackrel{\scriptscriptstyle\circ}{e}_j} e_i ) \cal{G} ({\stackrel{\scriptscriptstyle\circ}{e}}_m a_{(2)} {\stackrel{\scriptscriptstyle\circ}{e}_k} e_l) = \sum_{(a)} \delta_{ij} \delta_{kl} \cal{G}^{-1} (a_{(1)} {\stackrel{\scriptscriptstyle\circ}{e}_j}) \cal{G} (a_{(2)} e_l) \\ = \delta_{ij} \delta_{kl} \delta_{jl} \varepsilon (a e_l) = \hat{\cal{G}} \left( {\stackrel{\scriptscriptstyle\circ}{e}_i} e_j \left( \sum_{m \in \EuScript V} \varepsilon (a e_m) e_m \right) {\stackrel{\scriptscriptstyle\circ}{e}_k} e_l \right) \qquad\qquad\qquad \end{multline} % for each $i,j,k,l \in \EuScript V$ and $a \in \frak H$, where the second equality follows from \eqref{Geae} and \eqref{D(a)} and the third equality follows from \eqref{D(eeaee)}. By repeating similar calculation, we see that $\hat{\cal{G}}$ induces a group-like functional $\cal{G}_{\mathrm{Hc}}$ on $\mathrm{Hc}({\frak H})$. Now it is straightforward to verify that $\cal{G} \mapsto \cal{G}_{\mathrm{Hc}}$ gives the inverse of $\iota^*$. \end{pf} Combining Proposition \ref{GLFAwiso} and Proposition \ref{GLFHcHHiso}, we obtain the group isomorphism \begin{equation} \label{Phidef} \Phi\!: \bigl\{ G \in \mathrm{Aut}_{\pi (\frak{E})} (\Bbb K \EuScript G^1) \bigm| (G \bar{\otimes} G) w = w (G \bar{\otimes} G)\, \bigr\} \cong \mathrm{GLF} (\mathrm{Hc} ({\frak A}(w))) \end{equation} for each star-triangular face model $(\EuScript G, w)$. \section{A classification theory of ribbon functionals} % % % % % \begin{lem} For a coribbon Hopf $\EuScript V$-face algebra ${\frak H}$, we have % \begin{equation} \label{V(eae)} \cal{V}^{\pm 1} ({\stackrel{\scriptscriptstyle\circ}{e}_i} a {\stackrel{\scriptscriptstyle\circ}{e}_j}) = \cal{V}^{\pm 1} (e_i a e_j), \quad \cal{V}^{\pm 1} ({\stackrel{\scriptscriptstyle\circ}{e}_i} e_j) = \delta_{ij}, \end{equation} % \begin{equation} \label{V2} \cal{V}^2 = \cal{U}_1 \cal{U}_2^{-1}, \end{equation} % \begin{equation} \label{m(V-)} m^* (\cal{V}^{-1}) = (\cal{V} \otimes \cal{V})^{-1} {\cal R}^+_{21} {\cal R}^+. \end{equation} % \end{lem} \begin{pf} The first equality of \eqref{V(eae)} follows from the fact that $\cal{V}$ commutes with the face idempotents of $\frak H^{\circ}$. Using \eqref{m(V)} and \eqref{Rmeea}, we obtain % \begin{align} \cal{V} (a) & = \sum_{j,k \in \EuScript V} \sum_{(a)} {\cal R}^- (e_j, a_{(1)}) {\cal R}^- (a_{(2)}, {\stackrel{\scriptscriptstyle\circ}{e}_j} e_k) \cal{V} ({\stackrel{\scriptscriptstyle\circ}{e}_k}) \cal{V} (a_{(3)})\\ & = \sum_{k \in \V} \cal{V} ({\stackrel{\scriptscriptstyle\circ}{e}_k}) \cal{V} ({\stackrel{\scriptscriptstyle\circ}{e}_k} a) \nonumber\\ & = \langle \sum_{k \in \V} \cal{V} ({\stackrel{\scriptscriptstyle\circ}{e}_k}) {\stackrel{\scriptscriptstyle\circ}{e}_k} \cal{V}, a \rangle, \nonumber \end{align} % which implies $\sum_k \cal{V} ({\stackrel{\scriptscriptstyle\circ}{e}_k}) {\stackrel{\scriptscriptstyle\circ}{e}_k} = 1$. Since $\{ {\stackrel{\scriptscriptstyle\circ}{e}_k} \}$ is linearly independent by the second equality of \eqref{D(ee)}, this proves $\cal{V} ({\stackrel{\scriptscriptstyle\circ}{e}_k}) = 1$, or the second equality of \eqref{V(eae)}. By a similar discussion to \cite{Kassel} page 351, we obtain $\pi_M (\cal{V}^2) =$ $\pi_M (\cal{U}_1 \cal{U}_2^{-1})$ for every $\frak H$-comodule $M$. Hence \eqref{V2} follows from the fundamental theorem of coalgebras (cf. \cite{Sweedler} page 46). Using the fact that ${\cal R}^-$ is the $(m^* (1), (m^{\mathrm{op}})^* (1))$-generalized inverse of ${\cal R}^+$, we obtain % \begin{equation} \label{RRRR} {\cal R}^- {\cal R}^-_{21} {\cal R}^+_{21} {\cal R}^+ = m^* (1) = {\cal R}^+_{21} {\cal R}^+ {\cal R}^- {\cal R}^-_{21}. \end{equation} % Hence the right-hand side of \eqref{m(V-)} is the inverse of $m^* (\cal{V})$ in the algebra $m^* (1) (\frak H^{\otimes 2})^*$ $m^* (1)$. This proves \eqref{m(V-)}. \end{pf} \begin{prop} Let ${\frak H}$ be a CQT Hopf $\EuScript V$-face algebra and ${\cal V}$ an invertible element of $\frak H^*$. Then $(\frak H,\cal{V})$ is a coribbon Hopf $\EuScript V$-face algebra if and only if $\cal{M} = \cal{U}_1 \cal{V}^{-1}$ is group-like and satisfies the following relations\rom: % \begin{equation} \label{MXM-} \qquad {\cal M} X {\cal M}^{-1} = (S^2)^* (X) \quad (X \in \frak H^*), \end{equation} % \begin{equation} \label{M2=U1U2} \cal{M}^2 = \cal{U}_1 \cal{U}_2. \qquad \end{equation} % \end{prop} \begin{pf} To begin with, we note that the equivalence of $\cal{V} \in Z(\frak H^*)$ and \eqref{MXM-} follows from \eqref{UXU-}, and that that of \eqref{m(V)} and \eqref{D(g)} for $g = \cal{M}$ follows from \eqref{m(U)}, \eqref{m(V-)} and \eqref{RRRR}. Suppose $\cal{V}$ is a ribbon functional. Then the relation \eqref{M2=U1U2} follows from \eqref{V2} and \eqref{U1U2}, while the first (resp. second) relation of \eqref{gee} for $g = \cal{M}$ follows from \eqref{UXU-} and \eqref{S(ee)} (resp. \eqref{U(ee)} and the second relation of \eqref{V(eae)}). Conversely, if $\cal{M}$ satisfies the above conditions, \eqref{S(V)} follows from \eqref{V2}, \eqref{U1U2} and \eqref{S(G)}. \end{pf} For a coribbon Hopf $\EuScript V$-face algebra $(\frak H, \cal{V})$, we call $\cal{M} = \cal{U}_1 \cal{V}^{-1}$ the {\it modified ribbon functional} on $\frak H$ corresponding to $\cal{V}$. For each CQT Hopf $\EuScript V$-face algebra $\frak H$, we denote by $\mathrm{Rib}(\frak H)$ the set of all ribbon functionals on $\frak H$ and by $\mathrm{MRib}(\frak H)$ the set of all modified ribbon functionals on $\frak H$. \begin{prop} \label{MRibaltbraiding} Let $(\frak H, {\cal R}^{\pm})$ be a CQT Hopf $\EuScript V$-face algebra. \rom{(1)} We have % \begin{equation} \label{MRibR-21} \mathrm{MRib}((\frak H, {\cal R}^{\mp}_{21})) = \mathrm{MRib}((\frak H, {\cal R}^{\pm})). \end{equation} % \rom{(2)} Let $\frak H_{\gamma}$ $(\gamma \in \Gamma)$ and $\chi$ be as in Proposition \ref{altbraiding}. Then we have % \begin{equation} \label{MRibRchi} \mathrm{MRib}((\frak H, {\cal R}^{\pm}_{\chi})) = \mathrm{MRib}((\frak H, {\cal R}^{\pm})). \end{equation} \end{prop} \begin{pf} Let $\cal{U}_i$, $\cal{U}_i^{\prime}$ and $\cal{U}_{i, \chi}$ $(i = 1,2)$ be the Drinfeld functionals of $(\frak H, {\cal R}^{\pm})$, $(\frak H, {\cal R}^{\mp}_{21})$ and $(\frak H, {\cal R}^{\pm}_{\chi})$ respectively. Then we have $\cal{U}_1^{\prime}$ $=$ $\cal{U}_2$, $\cal{U}_2^{\prime}$ $=$ $\cal{U}_1$ and % \begin{equation} \cal{U}_{1, \chi} (a) = \chi (\gamma, \gamma)^{-1} \cal{U}_1 (a), \quad \cal{U}_{2, \chi} (a) = \chi (\gamma, \gamma) \cal{U}_2 (a) \quad (a \in \frak H_{\gamma}). \end{equation} % Hence the assertions follows from the definition of the modified ribbon functional and \eqref{U1U2}. \end{pf} \begin{thm} \label{clasRibHcAw} For each closable star-triangular face model $(\EuScript G, w)$, the map $\pi_{\Bbb K \EuScript G^1}$ gives the following bijection\rom{:} % \begin{equation} \label{Ribiso} \mathrm{Rib}(\mathrm{Hc} ({\frak A}(w))) \cong \{ V \in \mathrm{Aut}_{\mathrm{Hc} ({\frak A}(w))} (\Bbb K \EuScript G^1)\, |\, V^2 = \pi_{\Bbb K \EuScript G^1} (\cal{U}_1 \cal{U}_2^{-1}) \}. \end{equation} % Equivalently, $\pi_{\Bbb K \EuScript G^1}$ gives % \begin{equation} \label{MRibiso} \mathrm{MRib}(\mathrm{Hc} ({\frak A}(w))) \cong \{ M\, |\, M \pi_{\Bbb K \EuScript G^1} (\cal{U}_1)^{-1} \in \mathrm{Aut}_{\mathrm{Hc} ({\frak A}(w))} (\Bbb K \EuScript G^1),\, M^2 = \pi_{\Bbb K \EuScript G^1} (\cal{U}_1 \cal{U}_2) \}. \end{equation} % \end{thm} \begin{pf} Let $M$ be an element of the right-hand side of \eqref{MRibiso}. By \eqref{UURp}, $\pi_{\Bbb K \EuScript G^1} (\cal{U}_{\nu}) \bar{\otimes}$ $\pi_{\Bbb K \EuScript G^1} (\cal{U}_{\nu})$ commutes with $w$ for each $\nu = 1, 2$. Hence $M$ belongs to the right-hand side of \eqref{GLFAw}. Set $\cal{M} = \Phi (M)$, where $\Phi$ is as in \eqref{Phidef}. Since $\pi_{\Bbb K \EuScript G^1} (\cal{M}^2)$ $=$ $\pi_{\Bbb K \EuScript G^1} (\cal{U}_1 \cal{U}_2)$, we have $\cal{M}^2$ $=$ $\cal{U}_1 \cal{U}_2$ by Lemma \ref{cqtglf} (3). By \eqref{UXU-}, we have $\mathrm{coad}(\cal{M}) (e \binom{\bold p}{\bold q})$ $=$ $S^{-2} (e \binom{\bold p}{\bold q})$, for each $\bold p, \bold q \in \EuScript G^1$. Since $\mathrm{coad}(\cal{M})$ is an automorphism and $e \binom{\bold p}{\bold q}$, $S(e \binom{\bold p}{\bold q})$ $(\bold p, \bold q \in \EuScript G^1)$ and $e \binom{i}{j}$ $(i,j \in \EuScript V)$ generate $\mathrm{Hc} ({\frak A}(w))$, this shows that $\mathrm{coad}(\cal{M}) = S^{-2}$. Thus $\cal{M}$ is a modified ribbon functional of $\mathrm{Hc} ({\frak A}(w))$. Conversely, it is clear that $\pi_{\Bbb K \EuScript G^1}$ maps the left-hand side of \eqref{MRibiso} into the right-hand side of \eqref{MRibiso}. Thus we get the theorem. \end{pf} \begin{thm}[\cite{Reshetikhin}] \label{existrib} For each closable star-triangular face model $(\EuScript G, w)$ over an algebraically closed field $\Bbb K$ of $\mathrm{ch} \Bbb K \ne 2$, $\mathrm{Hc} ({\frak A}(w))$ has a ribbon functional. % \end{thm} \begin{pf} It suffices to construct a linear operator $V$ which belongs to the right-hand side of \eqref{Ribiso}. Let $A$ be the operator $\pi_{\Bbb K \EuScript G^1} (\cal{U}_1 \cal{U}_2^{-1})$ and $A = S + N$ its Jordan decomposition, that is, $S$ is a diagonalizable operator and $N$ is a nilpotent operator such that $SN = NS$. Let $\lambda_i$ $(1 \leq i \leq k)$ be (mutually distinct) eigenvalues of $S$ and $P_i$ the projection corresponding to $\lambda_i$. It is known that $P_i = f_i (A)$ and $N = g(A)$ for some polynomials $f_i, g \in \Bbb K [X]$. Let $ \sqrt{\lambda_i}$ be a square root of $\lambda_i$ and define a operator $V$ by $V = \sum_i \sqrt{\lambda_i} P_i h( S^{-1} N)$, where $h \in \Bbb K [X]$ is defined by % \begin{equation} h(X) = 1 + \sum_{n = 0}^{\sharp \EuScript G^1} (-1)^n 2^{-2n - 1} \frac{1}{n + 1} \binom{2n}{n} X^{n + 1}. \end{equation} Then, we have $V^2 = A$. Since $\cal{U}_1 \cal{U}_2^{-1}$ is a central element of $\mathrm{Hc} ({\frak A}(w))^*$ and $V$ is a polynomial of $A$, we have $V \in \mathrm{Aut}_{\mathrm{Hc} ({\frak A}(w))} (\Bbb K \EuScript G^1)$. By the theorem above, this proves the existence of a ribbon functional on $\mathrm{Hc} ({\frak A}(w))$. \end{pf} Let $(\EuScript G, w)$ be a closable star-triangular face model. We say that $(\EuScript G, w)$ is ({\it absolutely}) {\it irreducible} if $\Bbb K \EuScript G^1$ is (absolutely) irreducible as an $\mathrm{Hc}({\frak H})$-comodule. As an immediate consequence of the Theorem \ref{clasRibHcAw} and Schur's Lemma, we have the following. \begin{thm} \label{cardRib} Let $(\EuScript G, w)$ be an irreducible closable star-triangular face model over an algebraically closed field. Then we have $\sharp \mathrm{Rib}(\mathrm{Hc} ({\frak A}(w))) = 2$ if $\mathrm{ch} \Bbb K \ne 2$ and $\sharp \mathrm{Rib}(\mathrm{Hc} ({\frak A}(w))) = 1$ if $\mathrm{ch} \Bbb K = 2$. \end{thm} \begin{thm} \label{Mcrit} Let $(\EuScript G,w)$ be an absolutely irreducible closable star-triangular face model. Suppose $M \in GL (\Bbb K \EuScript G^1)$ satisfies $\sum_{\bold r \bold s} M^{\bold p}_{\,\bold r} e \binom{\bold r}{\bold s} (M^{-1})^{\bold s}_{\bold q}$ $=$ $S^2 (e \binom{\bold p}{\bold q})$ and $\mathrm{Tr} (M) = \mathrm{Tr} (M^{-1}) \ne 0$. Then we have % \begin{equation} \mathrm{MRib} (\mathrm{Hc} ({\frak A}(w))) = \{ \Phi (\pm M) \}. \end{equation} % \end{thm} \begin{pf} By Schur's lemma, we have $\pi (\cal{U}_{\nu})$ $=$ $c_{\nu} M$ for some nonzero constant $c_{\nu}$ $(\nu = 1,2)$. Since $\mathrm{Tr} \pi (\cal{U}_1)$ $=$ $\mathrm{Tr} \pi (\cal{U}_2^{-1})$ by \eqref{Udef} and \eqref{U-}, we obtain $c_1 \mathrm{Tr} (M)$ $=$ $c_2^{-1} \mathrm{Tr} (M^{-1})$. Therefore $M$ belongs to the right-hand side of \eqref{MRibiso}. \end{pf} \begin{prop} \label{clasquotient} Let $(\EuScript G, w)$ be a closable star-triangular face model and let $\frak K = \mathrm{Hc} ({\frak A}(w)) / \frak{I}$ be a quotient CQT Hopf $\EuScript V$-face algebra of $\frak H:= \mathrm{Hc} ({\frak A}(w))$ such that $\Bbb K \EuScript G^1$ is absolutely irreducible as a $\frak K$-comodule. Then the projection $p\!: \frak H \to \frak K$ gives the isomorphism % \begin{equation} \label{RibH} p^*\!: \mathrm{MRib} (\frak K) \cong \{ \cal{M} \in \mathrm{MRib} (\mathrm{Hc} ({\frak A}(w)))\, |\, \cal{M}(\frak{I}) = 0 \}. \end{equation} % \end{prop} \begin{pf} % We prove the assertion by using Lemma \ref{GLFisfunc} (2). Let $\cal{M}$ be a group-like functional on $\frak K$. It suffices to verify that $\mathrm{coad}(\cal{M})$ $=$ $S^{-2}$ if and only if $\mathrm{coad}(p^* (\cal{M}))$ $=$ $S^{-2}$. Since $\mathrm{coad}(\cal{M})(p (a))$ $=$ $p (\mathrm{coad}(p^* (\cal{M}))(a))$ for each $a \in \frak H$, the ``if''-part is obvious. Suppose $\mathrm{coad}(\cal{M})$ $=$ $S^{-2}$ and set $M: = \pi_{\Bbb K \EuScript G^1} (\cal{M})$. Since $\pi_{\Bbb K \EuScript G^1}^{\frak H} (p^* (\cal{M})) = M$, we have $\Phi (M) = p^* (\cal{M})$. On the other hand, using \eqref{f(U)=U} and Schur's Lemma, we see that $M \pi_{\Bbb K \EuScript G^1}^{\frak H} (\cal{U}_1)^{-1}$ is a scalar multiple of the identity operator. Hence $M$ belongs to the right-hand side of \eqref{MRibiso}. By Theorem \ref{clasRibHcAw}, this proves the proposition. \end{pf} Let $\frak H$ be a CQT Hopf $\EuScript V$-face algebra. We say that $\frak H$ is {\em monogenerated} if there exists an absolutely irreducible $\frak H$-comodule $U$ such that $\frak H$ is generated by ${\stackrel{\scriptscriptstyle\circ}{e}_i} e_j$ $(i,j \in \EuScript V)$, the image $C$ of the corepresentation $\mathrm{End}(U)^* \to \frak H$ and $S(C)$, as an algebra. \begin{lem} Let $\frak H$ be a CQT Hopf $\EuScript V$-face algebra over $\Bbb K$ and $\Bbb{F}$ a field extension of $\Bbb K$. Then $\frak H \otimes \Bbb{F}$ naturally becomes a CQT Hopf $\EuScript V$-face algebra over $\Bbb{F}$ and there exists an injection $\mathrm{Rib} (\frak H) \to \mathrm{Rib} (\frak H \otimes \Bbb{F});$ $\cal{V} \mapsto \cal{V}_{\Bbb{F}}$ given by $\cal{V}_{\Bbb{F}} (a \otimes 1_{\Bbb{F}})$ $=$ $\cal{V} (a)$. \end{lem} \begin{pf} This is straightforward. \end{pf} \begin{prop} \label{ribestimate} For each monogenerated CQT Hopf $\EuScript V$-face algebra $\frak H$, we have $\sharp \mathrm{Rib}(\frak H) \leq 2$ if $\mathrm{ch} \Bbb K \ne 2$ and $\sharp \mathrm{Rib}(\frak H) \leq 1$ if $\mathrm{ch} \Bbb K = 2$ \end{prop} \begin{pf} Let $(\EuScript G, w_U)$ and $f\!: \mathrm{Hc} (\frak{A}(w_U)) \to \frak H$ be as in Proposition \ref{univA}. Since $\frak H$ is monogenerated, $f$ is surjective for a suitable absolutely irreducible comodule $U$. Now the assertion is an immediate consequence of \eqref{RibH}, Theorem \ref{cardRib} and the lemma above. \end{pf} \noindent {\it Remark.} (1) To construct a link invariant via a lattice model $(\EuScript G, w)$, it is usual to assume that $(\EuScript G, w)$ is ``enhanced'' in the sense of \cite{TuraevYB} (cf. \cite{ADW}, \cite{Jones}, \cite{TuraevYB}). Theorem \ref{existrib} says that the assumption is superfluous provided that $(\EuScript G, w)$ is closable. For vertex models, this was first proved by Reshetikhin \cite{Reshetikhin}. \\ (2) Combining Theorem \ref{existrib} with the categorical framework of the link invariant \cite{Turaev3mfd}, we obtain an invariant of framed links colored by comodules of $\mathrm{Hc} ({\frak A}(w))$, for each closable star-triangular face model $(\EuScript G, w)$. Choosing the $\mathrm{Hc} ({\frak A}(w))$-comodule $\Bbb K \EuScript G^1$ as a color, we obtain an invariant $I_w (L)$ of framed links $L$ which agrees with the known one. However, if $(\EuScript G, w)$ is constructed from a (four-weight) spin model $(W_i)$ (\cite{Jones}, \cite{Bannai^2}), $I_w (L)$ does not agree with the known invariant $Z_{(W_i)} (L)$. In fact we have $I_w (L) = Z_{(W_i)} (L) Z_{(W_i)}^* (L)$, where $Z_{(W_i)}^* (L)$ is the ``dual invariant'' of $Z_{(W_i)} (L)$. \section{Quantized classical groups} Let $X_l$ be one of the Dynkin diagram of type $A_l$, $B_l$, $C_l$ or $D_l$, where $l \geq 1$ if $X = A$ and $l \geq 2$ if $X = B, C$ or $D$. We define integers $N$ and $\nu$ by \begin{equation} N = \begin{cases} l + 1 & (X = A) \\ 2l + 1 & (X = B) \\ 2l & (X = C, D), \\ \end{cases} \quad \nu = \begin{cases} 0 & (X = A) \\ - 1 & (X = B, D) \\ 1 & (X = C). \end{cases} \end{equation} For $X = B, C, D$ and $1 \leq i \leq N$, we set $i^{\prime} = N + 1 - i$ and $\bar{i} = i - \sigma_i \nu / 2$, where \begin{equation} \sigma_i = \begin{cases} 1 & (1 \leq i < (N+1)/2) \\ 0 & (i = (N+1)/2) \\ - 1 &((N+1)/2 < i \leq N), \end{cases} \quad \epsilon_i = \begin{cases} 1 & (1 \leq i \leq (N+1)/2) \\ - \nu & ((N+1)/2 \leq i \leq N). \end{cases} \end{equation} Also we set $\sigma_i \equiv 1$ for $X = A$. Let $\Check{R} = \Check{R}_q (X_l)$ be Jimbo's solution of the Yang-Baxter equation of type $X_{l}$: \begin{equation} \label{RAdef} \Check{R}_q (A_l) = q^{-1} \sum_{r = 1}^{N} E_{rr} \otimes E_{rr} + \sum_{r \ne s} E_{rs} \otimes E_{sr} - (q - q^{-1}) \sum_{r > s} E_{rr} \otimes E_{ss}, \end{equation} \begin{multline} \label{RBCDdef} \Check{R}_q (X_l) = \sum_{r;\, r \not= r'} (q^{-1} E_{rr} \otimes E_{rr} + q E_{rr'} \otimes E_{r'r}) + \sum_{r;\, r = r'} E_{rr} \otimes E_{rr} + \\ \sum_{r,s;\, r \ne s,s'} E_{rs} \otimes E_{sr} + (q - q^{-1}) \sum_{r > s} (- E_{rr} \otimes E_{ss} + \epsilon_r \epsilon_s q^{\overline{r} - \overline{s}} E_{rs'} \otimes E_{r's}) \quad (X = B, C, D), \end{multline} where for $X =$ $A, C, D$ (resp. $X= B)$, $q$ (resp. $q^{1/2}$) denotes a non-zero number such that $q^2 \not= 1$, and $E_{rs} \in \mathrm{Mat} (N, \Bbb{K})$ denote the matrix units. For $X =$ $B, C, D$, we also set $\lambda = - \nu q^{-N - \nu}$. For $1 \leq i, j \leq N$, we denote by $t_{ij}$ the element $e {i \choose j}$ of $\frak{A} (\Check{R})$, or its image by an arbitrary bialgebra map. For each $\eta \in \Bbb K^{\times}$, we denote by ${\cal R}^+_{\eta}$ the canonical braiding of the FRT bialgebra $\frak{A} (\eta \Check{R})$ or its Hopf closure $\mathrm{Hc} (\frak{A} (\eta \Check{R}))$. Since $\frak{A} (\Check{R})$ (resp. $\mathrm{Hc} (\frak{A} (\Check{R}))$) is isomorphic to $\frak{A} (\eta \Check{R})$ (resp. $\mathrm{Hc} (\frak{A} (\eta \Check{R}))$) as a bialgebra, we regard $\{ {\cal R}^+_{\eta} \}$ as a one-parameter family of braidings of $\frak{A} (\Check{R})$ (resp. $\mathrm{Hc} (\frak{A} (\Check{R}))$). \begin{thm}[Takeuchi \cite{Takeuchicocycle}] \label{clasbrFRT} Any braidings of $\frak{A} (\Check{R}_q (X_{l}))$ are either of the form ${\cal R}^+_{\eta}$ or of the form $({\cal R}^-_{\eta})_{21}$, where $\eta \in \Bbb K^{\times}$. \end{thm} \begin{pf} For $X = A$, this theorem has been proved by M. Takeuchi \cite{Takeuchicocycle}. Here we give a proof for $X = B, C, D$ by imitating his arguments. It is well known that the operators $g\!:= \Check{R}$ and $e\!:= (g - g^{-1}) / \mu + 1$ give a representation of the Birman-Murakami-Wenzl algebra on $(\Bbb K^N)^{\otimes 3}$ (cf. \cite{BirmanWenzl}, \cite{Murakami}), where $\mu = q - q ^{-1}$. That is, we have the following formulas: % \begin{equation} \label{minpolyR} (g_i - \lambda^{-1})(g_i + q )(g_i - q^{-1}) = 0 \quad (i = 1, 2), \end{equation} % \begin{equation} g_1 g_2 g_1 = g_2 g_1 g_2, \quad e_1 g_2 e_1 = \lambda e_1, \quad e_2 g_1 e_2 = \lambda e_2, \end{equation} % where, as usual, we set $f_1$ $=$ $f \otimes \mathrm{id}_{\Bbb K^N}$ and $f_2 = \mathrm{id}_{\Bbb K^N} \otimes f$ for $f \in \mathrm{End}_{\Bbb K} ((\Bbb K^N)^{\otimes 2})$. As consequences of these relations, we also obtain the following formulas: % \begin{equation} \label{gi2} g_i^2 = - \mu g_i + \lambda^{-1} \mu e_i + 1, \quad e_i^2 = \zeta e_i, \end{equation} % \begin{equation} \label{eigi} e_i g_i = g_i e_i = \lambda^{-1} e_i, \end{equation} % \begin{equation} e_i e_j e_i = e_i, \quad e_i g_j e_i = \lambda e_i, \end{equation} % \begin{equation} e_i e_j g_i = e_i g_j - \mu e_i e_j + \mu e_i, \quad g_i e_j e_i = g_j e_i - \mu e_j e_i + \mu e_i, \end{equation} % \begin{equation} e_i g_j g_i = e_i e_j, \quad g_i g_j e_i = e_j e_i, \end{equation} % \begin{equation} g_i e_j g_i - g_j e_i g_j = \mu (e_i g_j + g_j e_i - e_j g_i - g_i e_j) + \mu^2 (e_i - e_j) \end{equation} % for $(i,j)$ $=$ $(1,2)$, $(2,1)$, where $\zeta = $ $- (\lambda - \lambda^{-1}) \mu^{-1} + 1$. By \eqref{gi2} and \eqref{eigi}, we see that $\{ g, e, 1 \}$ is a linear basis of the algebra $\langle g \rangle$. Let $\cal{B}$ be a braiding of $\frak{A} (\Check{R}))$ and $\Check{B} \in \mathrm{End} ((\Bbb K^N)^{\otimes 2})$ the corresponding solution of the Yang-Baxter equation. Since $\frak{A}_2 (\Check{R})^*$ is the commutant of the algebra $\langle g \rangle$ in $\mathrm{End}_{\Bbb K} ((\Bbb K^N)^{\otimes 2})$, $\Check{B}$ belongs to the double commutant of $\langle g \rangle$. By \cite{Jacobson} page 202, this implies $\Check{B}$ $\in \langle g \rangle$. Hence $\Check{B}$ is of the form $a g + b e + c$ for some $a, b, c \in \Bbb K$. Rewriting the Yang-Baxter equation for $\Check{B}$ via the formulas above, we obtain % \begin{multline} ( \mu a^2 b + a b^2 ) X + ( - \mu a^2 c + a c^2 ) Y \\ + \{ b^3 + \mu^2 a^2 b + (\lambda + 2 \mu) a b^2 + \lambda^{-1} \mu a^2 c + \zeta b^2 c + b c^2 + 2 \lambda^{-1} a b c \} Z = 0, \end{multline} % where % \begin{equation} X = e_1 g_2 + g_2 e_1 - e_2 g_1 - g_1 e_2, \quad Y = g_1 - g_2, \quad Z = e_1 - e_2. \end{equation} % Since $X, Y, Z$ are linearly independent, we obtain three algebraic equations for $a$, $b$ and $c$. Solving these, we see that $\Check{B}$ is proportional to either $g$, $g^{-1} = g - \mu e + \mu$, $1 + \alpha e$ or $1$, where $\alpha$ denotes a solution of $x^2 + \zeta x + 1 = 0$. Suppose $\Check{B} = \eta$ or $\eta (1 + \alpha e)$ for some $\eta \in \Bbb K^{\times}$. Then using \eqref{mid(R)}, we obtain % \begin{equation} \cal{B} (t_{12} t_{21},\, t_{22}) = \eta^2, \quad \cal{B} (t_{21} t_{12},\, t_{22}) = 0. \end{equation} % On the other hand, substituting $t_{21} \otimes t_{12}$ into \eqref{RmXR}, we obtain $t_{21} t_{12} = t_{12} t_{21}$, a contradiction. Therefore $\Check{B}$ is proportional to either $g$ or $g^{-1}$. This completes the proof of the theorem. \end{pf} The following lemma allows us to apply our general results developed in Sect. 6 to the Hopf closures. \begin{lem} For each $q$ and $\eta$, $\Bbb K^N = \Bbb K \EuScript G^1$ is absolutely irreducible as a comodule of $\mathrm{Hc} ( \frak{A} ( \eta \Check{R}_q (X_l) ))$. In particular, $\mathrm{Hc} ( \frak{A} ( \eta \Check{R}_q (X_l) ))$ is monogenerated.\\ \end{lem} Since the proof of this lemma is quite similar to that of Lemma \ref{irrforqcg} below, we omit it. Next, we determine the ribbon functionals of the Hopf closure of $\frak{A} (\eta \Check{R}_q (X_l))$. We note that the following result immediately follows from Theorem \ref{Mcrit} and the formula \eqref{S2tij} given below, except for the case $\sum_i q^{2i - N - 1 - \sigma_i \nu} = 0$. \begin{prop} \label{clasribHc} For each $\eta \in \Bbb K^{\times}$, $\mathrm{Hc} (\frak{A} ( \eta \Check{R}_q (X_l) ))$ has exactly two \rom{(}resp. one\rom{)} modified ribbon functionals $\cal{M}_{\pm}$ given by % \begin{equation} \label{M+-def} \cal{M}_{\pm} (t_{ij}) = \pm \delta_{ij} q^{2i - N - 1 - \sigma_i \nu} \end{equation} if $\mathrm{ch} \Bbb K \ne 2$ \rom{(}resp. $\mathrm{ch} \Bbb K = 2$\rom{)}. \end{prop} \begin{pf} We will prove this result using Proposition \ref{ribestimate} and Theorem \ref{clasRibHcAw}. Using \eqref{S2tij} and \eqref{UXU-}, we obtain % \begin{gather} (\pi (\cal{U}_1) \otimes \mathrm{id}) \circ \rho \circ \pi (\cal{U}_1)^{-1} (u_j) = \sum_i u_i \otimes q^{2(i -j) - (\sigma_i - \sigma_j) \nu}\, t_{ij} \\ = (M \otimes \mathrm{id}) \circ \rho \circ M^{-1} (u_j), \end{gather} % where $M:= \mathrm{diag} (q^{2i - N - 1 - \sigma_i \nu})_i$. This shows that $M \pi (\cal{U}_1)^{-1}$ commutes with the coaction of $\mathrm{Hc} (\frak{A} ( \eta \Check{R}_q (X_l) ))$ on $\Bbb K \EuScript G^1$. Hence, it suffices to verify that \begin{equation} \label{piU1U2=M2} \pi (\cal{U}_1 \cal{U}_2) = M^2. \end{equation} % By Schur's lemma, we have $\pi (\cal{U}_{\nu}) = c_{\nu} M$ for some constant $c_{\nu} \in\Bbb K^{\times}$. Suppose $X = B, C$ or $D$. Using \eqref{SBCD}, we compute % \begin{equation} \cal{U}_1 (t_{11}) = \sum_{k=1}^{N} \epsilon_1 \epsilon_k q^{\overline{1} - \overline{k}} \eta \Check{R} \left( k^{\prime} \frac[0pt]{1}{k} 1^{\prime} \right) = \eta \Check{R} \left( N \frac[0pt]{1}{1} N \right) = \eta q. \end{equation} % Using \eqref{S(U)} and \eqref{SBCD}, we also obtain % \begin{equation} \cal{U}_2^{-1} (t_{NN}) = \cal{U}_1 (t_{11}) = \eta q. \end{equation} This proves $c_1 = \eta q^{N + \nu} = c_2^{-1}$, or \eqref{piU1U2=M2} for $X = B, C$, $D$. When $X = A$, \eqref{piU1U2=M2} is proved by computing the Lyubashenko double of $\Check{R}_q (A_l)$ explicitly. \end{pf} Hereafter, we assume that $q^2 \ne -1$ and that $\lambda^{-1} \ne q^{-1}, -q$ if $X = B, C$ or $D$. By \eqref{minpolyR}, this implies \begin{equation} \Bbb K^N \otimes \Bbb K^N = \begin{cases} \mathrm{Ker} (\Check{R} - q^{-1}) \oplus \mathrm{Ker} (\Check{R} + q) & (X = A)\\ \mathrm{Ker} (\Check{R} - q^{-1}) \oplus \mathrm{Ker} (\Check{R} + q) \oplus \mathrm{Ker} (\Check{R} - \lambda^{-1}) & (X = B, C, D)\\ \end{cases} \end{equation} as $\frak{A} (\Check{R})$-comodules. To give the definition of the quantized classical groups, we recall the definition of the (quantum) determinant of $\frak{A} ( \Check{R})$. Let $\Omega = \Omega (X_{l})$ be the following q-analogue of the exterior algebra: \begin{equation} \Omega (X_{l}) = \begin{cases} T( \Bbb K^N ) / (\mathrm{Ker} (\Check{R} - q^{-1})) & (X = A, C) \\ T( \Bbb K^N ) / (\mathrm{Im} (\Check{R} + q)) & (X = B, D). \end{cases} \end{equation} More explicitly, we have \begin{equation} \Omega (A_{l}) = \bigl\langle u_i \; (1 \leq i \leq N) \, \bigm\vert \, u_i^2 = 0,\, q u_i u_j + u_j u_i = 0 \; (i < j) \bigr\rangle, \end{equation} \begin{multline} \Omega (X_{l}) = \bigl\langle u_i \; (1 \leq i \leq N) \, \bigm\vert \, u_i^2 = 0 \; (i \not= (N+1)/2), \\ q u_i u_j + u_j u_i = 0 \; (i < j, i \not= j'), \; \eta_i = 0 \; (1 \leq i \leq (N+1)/2) \bigr\rangle \\ (X = B, C, D). \end{multline} Here for $X =$ $B$, $C$, $D$ and $1 \leq i \leq (N+1)/2$, we set \begin{equation} \eta_i = \begin{cases} u_{i'} u_i + u_i u_{i'} - (q - q^{-1}) \sum_{j = 1}^{i-1} q^{j - i + 1} u_j u_{j'} & (X = B, D,\, i \leq l) \\ u_{l+1} u_{l+1} - (q^{1/2} - q^{- 1/2}) \sum_{j = 1}^{l} q^{j - l} u_j u_{j'} & (X = B, \, i = l+1) \\ u_{i'} u_i + q^2 u_i u_{i'} + (q - q^{-1}) \sum_{j = i+1}^{l} q^{j - i + 1} u_j u_{j'} & (X = C,\, i \leq l). \end{cases} \end{equation} Then $\Omega (X_{l})$ becomes an $\frak{A} (\Check{R}_q (X_l))$-comodule algebra via $u_j \mapsto \sum_i u_i \otimes t_{ij}$. For $0 \leq r \leq N$, $\Omega_r := \sum_{i_1, \ldots, i_r} \Bbb K u_{i_1} \cdots u_{i_r}$ is a $\binom{N}{r}$-dimensional subcomodule of $\Omega$. In particular, $\Omega_N$ $=$ $\Bbb K u_1 u_2 \ldots u_N$ is one-dimensional and determines the group-like element $\det \in \frak{A} (\Check{R})$ via the coaction $u_1 \ldots u_N \mapsto u_1 \ldots u_N \otimes \det$. For $X = B, C, D$, $\frak{A} (\Check{R})$ has another group-like element $\mathrm{quad}$ which is determined by its coaction on the one-dimensional comodule \begin{equation} \mathrm{Ker} (\Check{R} - \lambda^{-1}) = \Bbb K \sum_i \epsilon_i q^{\overline{i} + 1 / 2} u_i \otimes u_{i'}. \end{equation} By \cite{qcg} Proposition 5.4-5.5 and the universal mapping property of the Hopf closure and the localization construction, we have \begin{equation} \label{HcA} \mathrm{Hc} ( \frak{A} ( \eta \Check{R}_q (A_l) )) \cong \frak{A} ( \eta \Check{R}_q (A_l) ) [\mathrm{det}^{-1}] =: \mathrm{Fun} \left( GL_{q} (N) \right)_{\eta}, \end{equation} \begin{align} \label{HcBCD} \mathrm{Hc} ( \frak{A} ( \eta \Check{R}_q (X_l) )) & \cong \frak{A} ( \eta \Check{R}_q (X_l) ) [\mathrm{quad}^{-1}] \\ & \cong \frak{A} ( \eta \Check{R}_q (X_l) ) [\mathrm{det}^{-1}] \quad (X = B, C, D). \end{align} The biideal $(\det - 1)$ becomes a CQT biideal of $\frak{A} (\eta \Check{R})$ if and only if \begin{equation} \eta^N = % \begin{cases} q & (X = A) \\ 1 & (X = B, C, D), \end{cases} \end{equation} while $(\mathrm{quad} - 1)$ becomes a CQT biideal if and only if $\eta = \pm 1$ (cf. \cite{gd}). Now we define the {\it function algebra of the quantized classical groups} (cf. \cite{TakeuchiMat}, \cite{qcg}, \cite{gd}) to be the CQT bialgebras given by \begin{equation} \label{SLdef} \mathrm{Fun} \left( SL_{q} (N) \right)_{\eta} = \frak{A} (\eta \Check{R}_q (A_l)) / (\mathrm{det} -1 ) \quad (\eta^N = q), \end{equation} \begin{multline} \label{SOdef} \mathrm{Fun} \left( SO_{q} (N) \right)_{\eta} = \frak{A} (\eta \Check{R}_q (X_{l})) / (\mathrm{det} - 1, \mathrm{quad} - 1) \\ (\eta = 1\, \mathrm{if}\, X = B\, \mathrm{and}\, \eta = \pm 1\, \mathrm{if}\, X = D), \end{multline} \begin{equation} \label{Odef} \mathrm{Fun} \left( O_{q} (N) \right)_{\eta} = \frak{A} (\eta \Check{R}_q (X_{l})) / (\mathrm{quad} - 1) \quad (\eta = \pm 1, X = B, D) \end{equation} \begin{equation} \label{Spdef} \mathrm{Fun} \left( Sp_{q} (N) \right)_{\eta} = \frak{A} (\eta \Check{R}_q (C_{l})) / (\mathrm{quad} - 1) \quad (\eta = \pm 1). \end{equation} See \cite{qcg} for a justification of these definitions in case $\Bbb K = \Bbb C$ and $q$ is transcendental over $\Bbb{Q}$. For $G_q = SL_q (N), SO_q (N)$, etc., we denote by $\mathrm{Fun} (G_q)$ the underlying bialgebra of $\mathrm{Fun} (G_q)_{\eta}$, and by ${\cal R}^+_{\eta, G_q}$ the braiding of $\mathrm{Fun} (G_q)_{\eta}$. Each of these algebras has an antipode. For example, the antipode of the algebras given in \eqref{SOdef}-\eqref{Spdef} is given by \begin{equation} \label{SBCD} S(t_{ij}) = \epsilon_i \epsilon_j\, q^{\overline{i} - \overline{j}}\, t_{j'i'}. \end{equation} The square of the antipode of the algebras given in \eqref{HcA}, \eqref{HcBCD}, \eqref{SLdef}-\eqref{Spdef} is given by \begin{equation} \label{S2tij} S^2 (t_{ij}) = q^{2(i -j) - (\sigma_i - \sigma_j) \nu}\, t_{ij}. \end{equation} \begin{lem} \label{irrforqcg} Let $F$ be either $\mathrm{Hc} (\frak{A} (\Check{R}))$ or one of the algebras given in \eqref{SLdef}-\eqref{Spdef}. Then each of the $F$-comodules $\Bbb K^N$, $\mathrm{Ker} (\Check{R} - q^{-1})$ and $\mathrm{Ker} (\Check{R} + q)$ are absolutely irreducible. In particular, $F$ is monogenerated.\\ \end{lem} \begin{pf} Since $\Bbb K$ is arbitrary, it suffices to show the irreducibility of these comodules. Here we give a proof for $W := \mathrm{Ker} (\Check{R}_q (C_l) + q)$. To simplify the computation, it is convenient to identify $W$ with its image via the projection $(\Bbb K^N)^{\otimes 2} \to \Omega_2$. Following \cite{RTF}, we define $K_i, E_i, F_i \in F^*$ $(1 \leq i \leq l)$ by % \begin{equation} \label{Kidef} K_i = {\cal R}^-_{\eta} (t_{ii},\, -), \end{equation} % \begin{equation} E_i = \begin{cases} - \eta^{-1} (q - q^{-1})^{-1} {\cal R}^+_{\eta} (-,\, t_{i+1\,\, i}) & (1 \leq i < l)\\ - \eta^{-1} q^{-1} (q^2 - q^{-2})^{-1} {\cal R}^+_{\eta} (-,\, t_{l+1\,\, l}) & (i = l), \\ \end{cases} \end{equation} % \begin{equation} F_i = \begin{cases} \eta (q - q^{-1})^{-1} {\cal R}^-_{\eta} (t_{i\,\, i+1},\,-) & (1 \leq i < l)\\ \eta q (q^2 - q^{-2})^{-1} {\cal R}^-_{\eta} (t_{l\,\, l+1},\,-) & (i = l). \\ \end{cases} \end{equation} % Then these elements belong to the dual Hopf algebra $F^{\circ}$ (cf. \cite{LarsonTowber}) and satisfy % \begin{equation} \pi_{\Bbb K^N} (K_i) = \eta^{-1} \sum_{k=1}^{N} q^{\delta_{ki} - \delta_{ki^{\prime}}} E_{kk}, \end{equation} % \begin{equation} \pi_{\Bbb K^N} (E_i) = E_{i\, i+1} - q E_{(i+1)^{\prime}\, i^{\prime}} \quad (i < j), \quad \pi_{\Bbb K^N} (E_l) = E_{l\,\, l+1}, \end{equation} % % \begin{equation} \pi_{\Bbb K^N} (F_i) = E_{i+1\, i} - q^{-1} E_{i^{\prime}\, (i+1)^{\prime}} \quad (i < j), \quad \pi_{\Bbb K^N} (F_l) = E_{l+1\,\, l}, \end{equation} % \begin{equation} \Delta (K_i) = K_i \otimes K_i, \end{equation} % \begin{equation} \Delta (E_i) = E_i \otimes K_i^{-1} + K_{i + 1}^{-1} \otimes E_i, \quad \Delta (F_i) = F_i \otimes K_{i + 1} + K_i \otimes F_i, \end{equation} % where $K_{l+1}$ is given by \eqref{Kidef}. As a $\langle K_i \rangle$-module, $W$ is the direct sum of the mutually non-isomorphic, non-trivial comodules $\Bbb K u_i u_j$ $(j \ne i, i^{\prime})$ and the trivial comodule $T = \bigoplus_{i=1}^{l-1} \Bbb K (q u_{i} u_{i^{\prime}} - u_{i+1} u_{(i+1)^{\prime}})$. Hence any non-zero subcomodule $M$ of $W$ contains a vector $v \ne 0$ which belongs to one of these $\langle K_i \rangle$-modules. By verifying $T \cap (\bigcap_i \mathrm{Ker}\pi (E_i))$ $=$ $0$, we see that $u_1 u_2 \in \Bbb K E_{i_1} \cdots E_{i_k} v$ for some $i_1, \ldots, i_k$. Also, by verifying $T = \sum_{i=1}^{l-1} \Bbb K F_i (u_i u_{(i+1)^{\prime}})$, we see that $u_1 u_2$ generates $W$ as an $\langle F_i \rangle$-module. Thus, $W$ is irreducible as a $\langle K_i, E_i, F_i \rangle$-module, and also, it is irreducible as an $F$-comodule. % \end{pf} \begin{thm} \label{clasbrqcg} \rom{(1)} Any braidings of $\mathrm{Hc} (\frak{A} (\Check{R}_q (X_{l})))$ are either of the form ${\cal R}^+_{\eta}$ or of the form $({\cal R}^-_{\eta})_{21}$, where $\eta \in \Bbb K^{\times}$.\\ \rom{(2)} Let $G_q$ be either $SL_q (N)$, $SO_q (N)$, $O_q (N)$ or $Sp_q (N)$. Then, any braiding of $\mathrm{Fun} (G_q)$ is either of the form ${\cal R}^+_{\eta, G_q}$ or of the form $({\cal R}^-_{\eta, G_q})_{21}$, where $\eta$ is as in \eqref{SLdef}-\eqref{Spdef}. \end{thm} \begin{pf} Let $\Check{B}$ be a solution of the Yang-Baxter equation, which corresponds to a braiding of one of the above algebras $F$. By Lemma \ref{irrforqcg}, $\mathrm{End}_F ((\Bbb K^N)^{\otimes 2})$ is spanned by two or three projections onto eigenspaces of $\Check{R}$, according to $X = A$ or $X = B, C, D$. By linear algebra, these projections are polynomials of $\Check{R}$. Therefore, we have $\mathrm{End}_F ((\Bbb K^N)^{\otimes 2})$ $=$ $\langle \Check{R} \rangle$. Hence, by the discussions in the proof of Theorem \ref{clasbrFRT}, we see that $\Check{B}$ is proportional to either $\Check{R}$ or $\Check{R}^{-1}$. Thus this theorem follows from the result of \cite{gd} stated above. \end{pf} Let $F$ be one of the Hopf algebras given in \eqref{SLdef}-\eqref{Spdef}. We define the cyclic group $\Gamma$ $=$ $\Gamma_F$ as follows: \begin{equation} \Gamma = \begin{cases} \Bbb Z / N \Bbb Z & ( G_q = SL_q(N))\\ \Bbb Z / 2 \Bbb Z & ( G_q = O_q(N),\, SO_q (2 l),\, Sp_q (N))\\ \{ 1 \} & ( G_q = SO_q (2 l + 1)).\\ \end{cases} \end{equation} For $F$ $=$ $\frak{A} (\Check{R})$ and $\mathrm{Hc} (\frak{A} (\Check{R}))$, we also set $\Gamma_F = \Bbb Z$. Since $\det \in \frak{A}_N (\Check{R})$ and $\mathrm{quad} \in \frak{A}_2 (\Check{R})$, the grading $\frak{A} (\Check{R})$ $=$ $\bigoplus_n \frak{A}_n (\Check{R})$ naturally induces a $\Gamma$-grading of $F$ satisfying the properties stated in Proposition \ref{altbraiding} (2). Now we can restate our classification theorems for braidings as follows. \begin{cor} \label{clasbrviaGamma} Let $F$ be one of the bialgebras treated in Theorem \ref{clasbrFRT} and Theorem \ref{clasbrqcg} and let $\Gamma$ be the cyclic group defined as above. Then, any braiding of $F$ is either of the form ${\cal R}^+_{\chi}$ or of the form $({\cal R}^-_{\chi})_{21}$, where $\chi$ is as in Proposition \ref{altbraiding}. In particular, $\mathrm{MRib} (F)$ does not depend on the choice of the braiding of $F$ \rom{(}cf. Proposition \ref{MRibaltbraiding}\rom{)}. \end{cor} Next, we give the classification theorem of the ribbon functionals for the algebras given in \eqref{SLdef}-\eqref{Spdef}. \begin{lem} Let $\cal{M}_{\pm}$ be as in \eqref{M+-def}. Then we have % \begin{equation} \cal{M}_{\pm} (\mathrm{det}) = (\pm 1)^N, \quad \cal{M}_{\pm} (\mathrm{quad}) = 1. \end{equation} % \end{lem} \begin{pf} We calculate % \begin{multline} \cal{M}_{\pm} (\mathrm{det}) u_1 \cdots u_N = \pi_{\Omega_N} (\cal{M}_{\pm}) (u_1 \cdots u_N) = (\cal{M}_{\pm} u_1) \cdots (\cal{M}_{\pm} u_N) \\ = \prod_i \left( \pm q^{2i - N -1 - \sigma_i \nu} \right) u_1 \cdots u_N = (\pm 1)^N u_1 \cdots u_N. \qquad \end{multline} % The proof of the second formula is similar. \end{pf} In view of the universal mapping property of the Hopf closure, we see that we may replace $\frak{A} (\eta \Check{R})$ in \eqref{SLdef}-\eqref{Spdef} with $\mathrm{Hc}( \frak{A} (\eta \Check{R}))$. Hence, as an immediate consequence of Proposition \ref{clasquotient} and the lemma above, we obtain the following. \begin{thm} \rom{(1)} % Let $G_q$ be either $SL_{q} (N)$, $SO_{q} (N)$ or $Sp_{q} (N)$, and let $p\!: \mathrm{Hc} (\frak{A} ( \eta \Check{R}_q (X_l) )) \to$ $\mathrm{Fun} (G_q)_{\eta}$ denote the projection. Then we have % \begin{equation} % \mathrm{MRib} \left(\mathrm{Fun} (G_q)_{\eta} \right) = % \begin{cases} \{ \cal{M}_+ \circ p,\, \cal{M}_- \circ p \} & (N \in 2 \Bbb Z) \\ \{ \cal{M}_+ \circ p \} & (N \in 1 + 2 \Bbb Z), \end{cases} \end{equation} where $\eta$ is as in \eqref{SLdef}-\eqref{Spdef}. \\ \rom{(2)} We have % \begin{equation} \mathrm{MRib} \left( \mathrm{Fun} \left( O_{q} (N) \right)_{\pm 1} \right) = \{ \cal{M}_+ \circ r,\, \cal{M}_- \circ r \}, \end{equation} where $r\!: \mathrm{Hc}( \frak{A} ( \pm \Check{R}_q (X_l) )) \to$ $\mathrm{Fun} \left( O_{q} (N) \right)_{\pm 1}$ denotes the projection. \end{thm} \section{SOS algebras} Let $N \geq 2$ and $L \geq 2$ be integers. Let $\cal{C}$ be an $\Bbb C$-abelian semisimple rigid monoidal category whose simple objects $L_{\lambda}$ are indexed by the following set of partitions: \begin{equation} \label{Vdef} \EuScript V = {\EuScript V}_{NL}:= \bigl\{ \lambda = (\lambda_1, \ldots, \lambda_N) \in \Bbb Z^N \bigm| L \geq {\lambda}_1 \geq \dots \geq {\lambda}_{N} = 0 \bigr\}. \end{equation} We say that $\cal{C}$ is an $SU(N)_L$-{\it category} if the structure constants of its Grothendiek ring agree with the fusion rules $N^{\nu}_{\lambda \mu}$ of $SU(N)_L$-WZW models. The $SU(N)_L$-categories play crucial roles to construct $SU(N)_L$-topological quantum field theories, or corresponding invariants of 3-manifolds (cf. \cite{Turaev3mfd}). It is known that two $SU(N)_L$-categories are equivalent to each other up to a ``twist'' of the associativity constraint (cf. Kazhdan-Wenzl \cite{KazhdanWenzl}). In \cite{fut}, we have constructed a coribbon Hopf $\EuScript V_{NL}$-face algebra $\frak{S} = {\frak S}(A_{N-1};t)_{\epsilon}$ such that $\bold{Com}_{\frak{S}}^f$ is an $SU(N)_L$-category. In this section, we determine the braiding and the ribbon structure of $\bold{Com}_{\frak{S}}^f$ or equivalently, those of $\frak{S}$ (cf. Proposition \ref{correspalgcom}). To begin with, we recall the definition of $SU(N)_L$-SOS model. For each $1 \leq i \leq N$, we set $\hat{i} = (\delta_{1i}, \ldots, \delta_{Ni}) \in \Bbb Z^N$. For $m \geq 0$, we define the subset ${\EuScript G}^m$ of ${\EuScript V}^{m+1}$ by \begin{equation} {\EuScript G}^m = {\EuScript V}^{m+1} \cap \bigl\{ \bold p = (\lambda\, |\, i_1, \ldots, i_m) \bigm| \lambda\in \EuScript V,\, 1 \leq i_1, \ldots, i_m \leq N \bigr\}, \end{equation} where for $\lambda \in \Bbb Z^N$ and $1 \leq i_1, \ldots, i_m \leq N$, we set \begin{equation} (\lambda\, |\, i_1, \ldots, i_m) = (\lambda, \lambda + \hat{i}_1, \ldots, \lambda + \hat{i}_1 + \cdots + \hat{i}_m), \end{equation} and we identify $(\lambda_1 +1, \cdots, \lambda_N + 1) \in \Bbb Z^N$ with $\lambda \in \EuScript V$. Then $(\EuScript V,{\EuScript G}^1)$ defines an oriented graph $\EuScript G = {\EuScript G}_{N,L}$ and ${\EuScript G}^m$ is identified with the set of paths of $\EuScript G$ of length $m$. For $\bold p = (\lambda\, |\, i,j)$, we set ${\bold p}^{\dag} = (\lambda\, |\, j,i)$ and $d(\bold p) = {\lambda}_i - {\lambda}_j + j - i$. We define subsets ${\EuScript G}^2[\to]$, ${\EuScript G}^2[\;\downarrow\;]$ and ${\EuScript G}^2[\searrow]$ of ${\EuScript G}^2$ by \begin{gather} {\EuScript G}^2[\to] = \bigl\{ \bold p \in {\EuScript G}^2 \bigm| \bold p^{\dag} = \bold p \bigr\}, \quad {\EuScript G}^2[\;\downarrow\;] = \bigl\{ \bold p \in {\EuScript G}^2 \bigm| \bold p^{\dag} \not\in \EuScript G^2 \bigr\}, \\ {\EuScript G}^2[\searrow] = \bigl\{ \bold p \in {\EuScript G}^2 \bigm| \bold p \not= \bold p^{\dag} \in \EuScript G^2 \bigr\}. \end{gather} Let $t \in\Bbb C$ be a primitive $2(N+L)$-th root of $1$. Let $\epsilon$ be either $1$ or $-1$ and $\zeta$ a nonzero parameter. We define a star-triangular face model $(\EuScript G,w_{N,t,\epsilon}) = ({\EuScript G}_{N,L},w_{N,t, \epsilon, \zeta})$ by setting \begin{equation} w_{N,t,\epsilon} \! \begin{bmatrix} \lambda & \lambda + \hat{i} \\ \lambda + \hat{i} & \lambda + \hat{i} + \hat{j} \end{bmatrix} = - \zeta^{-1} t^{-d(\bold p)} \frac{1}{[d(\bold p)]}, \end{equation} \begin{equation} w_{N,t,\epsilon} \! \begin{bmatrix} \lambda & \lambda + \hat{i} \\ \lambda + \hat{j} & \lambda + \hat{i} + \hat{j} \end{bmatrix} = \zeta^{-1} \epsilon \, \frac{[d(\bold p)-1]}{[d(\bold p)]}, \end{equation} \begin{equation} w_{N,t,\epsilon} \! \begin{bmatrix} \lambda & \lambda + \hat{k} \\ \lambda + \hat{k} & \lambda + 2 \hat{k} \end{bmatrix} = \zeta^{-1} t \end{equation} for each $\bold p = (\lambda\, |\, i,j) \in \EuScript G^2[\searrow] \amalg \EuScript G^2 [\,\downarrow\,]$ and $(\lambda\, |\, k,k) \in \EuScript G^2[\to]$, where $[n] = (t^n-t^{-n})/(t-t^{-1})$ for each $n \in \Bbb Z$. We call $(\EuScript G,w_{N,t,\epsilon})$ {\it $SU(N)_L$-SOS model} (without spectral parameter) \cite{JMO}. Now the $SU(N)_L$-SOS algebra ${\frak S}(A_{N-1};t)_{\epsilon}$ is defined as the following quotient of the FRT construction ${\frak A}(w_{N,t,\epsilon})$: \begin{equation} {\frak S}(A_{N-1};t)_{\epsilon}:= {\frak A}(w_{N,t,\epsilon}) / (\det - 1), \end{equation} where the group-like element ${\det} = \sum_{\lambda, \mu \in \EuScript V} \det \binom{\lambda}{\mu}$ of ${\frak A}(w_{N,t,\epsilon})$ is defined by \begin{equation} \label{detformula} \det \binom{\lambda}{\mu} = \frac{D(\mu)}{D(\lambda)} \sum_{\bold p \in \EuScript G^N_{\lambda \lambda}} (- \epsilon)^{\EuScript{L} (\bold p) + \EuScript{L} (\bold q)} e \binom{\bold p}{\bold q}. \end{equation} Here $\bold q$ denotes an arbitrary element of $\EuScript G^N_{\mu \mu}$, and $\EuScript{L} \!: \EuScript G^m \to \Bbb Z_{\geq 0}$ and $D( \lambda ) \in \Bbb C$ are given by \begin{equation} \EuScript{L} (\lambda\, |\, i_1, \ldots, i_m) = \mathrm{Card} \{ (k,l) | 1 \leq k < l \leq N, i_k < i_l \}, \end{equation} \begin{equation} D(\lambda) = \prod_{1 \leq i < j \leq N} \frac{[d(\lambda\, |\, i,j)]}{[d(0\, |\, i,j)]} \quad (\lambda \in \EuScript V). \end{equation} The canonical braiding of $\frak{A} (w_{N,t,\epsilon, \zeta})$ induces the braiding ${\cal R}^+_{\zeta}$ of ${\frak S}(A_{N-1};t)_{\epsilon}$ if and only if $\zeta$ satisfies $\zeta^N = \epsilon^{N-1} t$. The face algebra ${\frak S}(A_{N-1};t)_{\epsilon}$ has an antipode whose square is given by \begin{equation} S^2 \left( e \binom{\bold p}{\bold q} \right) = \frac{D(\frak {r} (\bold p)) D(\frak {s} (\bold q))}{D(\frak {s} (\bold p)) D(\frak {r} (\bold q))} e \binom{\bold p}{\bold q} \quad (\bold p, \bold q \in \EuScript G^m, m \geq 0). \end{equation} The $\frak{S}$-comodule $\Bbb K \EuScript G^1$ is irreducible, while the $\frak{S}$-comodule $\Bbb K \EuScript G^2$ has the irreducible decomposition: \begin{equation} \Bbb K \EuScript G^2 = \mathrm{Ker} (w_{N,t,\epsilon, \zeta} - \zeta^{-1} t) \oplus \mathrm{Ker} (w_{N,t,\epsilon, \zeta} + \zeta^{-1} t^{-1}). \end{equation} The proof of the following result is quite similar to that of Theorem \ref{clasbrqcg} and Takeuchi \cite{Takeuchicocycle} Lemma 2.4, hence we omit it. \begin{thm} Any braiding of ${\frak S}(A_{N-1};t)_{\epsilon}$ is either of the form ${\cal R}^+_{\zeta}$ or of the form $({\cal R}^-_{\zeta})_{21}$, where $\zeta$ denotes a solution of $\zeta^N = \epsilon^{N-1} t$. \end{thm} Similarly to Corollary \ref{clasbrviaGamma}, we can rewrite the result above in terms of the $\Bbb Z / N \Bbb Z$-grading of ${\frak S}(A_{N-1};t)_{\epsilon}$ induced by $\frak{A} (w_{N,t,\epsilon})$ $=$ $\bigoplus_n \frak{A}_n (w_{N,t,\epsilon})$. \begin{thm} When $N$ is odd, ${\frak S}(A_{N-1};t)_{\epsilon}$ has exactly one ribbon functional. The corresponding modified ribbon functional $\cal{M}_+$ is given by % \begin{equation} \label{M+SOS} \cal{M}_+ = \sum_{k,l \in \V} \frac{D(l)}{D(k)}\, {\stackrel{\scriptscriptstyle\circ}{e}_k} e_l. \end{equation} % When $N$ is even, ${\frak S}(A_{N-1};t)_{\epsilon}$ has exactly two ribbon functionals. The corresponding modified ribbon functionals $\cal{M}_{\pm}$ are given by \eqref{M+SOS} and % \begin{equation} \left\langle \cal{M}_-, e\! \binom{\bold p}{\bold q} \right\rangle = \delta_{\bold p \bold q} (-1)^m \frac{D(\frak {r} (\bold p))}{D(\frak {s} (\bold p))} \quad (\bold p, \bold q \in \EuScript G^m, m \geq 0). \end{equation} % % % \end{thm} \begin{pf} By Theorem \ref{Mcrit}, we have $\mathrm{MRib} (\mathrm{Hc} ( {\frak A}(w_{N,t,\epsilon})))$ $=$ $\{ \Phi (\pm M) \}$, where $M \in GL (\Bbb K \EuScript G^1)$ is given by $M \bold p = D(\frak {r} (\bold p)) D(\frak {s} (\bold p))^{-1} \bold p$ $(\bold p \in \EuScript G^1)$. Since % \begin{align*} \left\langle \Phi (\pm M),\, e \binom{\bold p}{\bold q} \right\rangle = & \left( \pm \delta_{\bold p_1 \bold q_1} \frac{D(\frak {r} (\bold p_1))}{D(\frak {s} (\bold p_1))} \right) \cdots \left( \pm \delta_{\bold p_m \bold q_m} \frac{D(\frak {r} (\bold p_m))}{D(\frak {s} (\bold p_m))} \right) \\ = & (\pm 1)^m \delta_{\bold p \bold q} \frac{D(\frak {r} (\bold p))}{D(\frak {s} (\bold p))} \end{align*} % for each $\bold p = (\bold p_1 \ldots \bold p_m)$, $\bold q = (\bold q_1, \ldots \bold q_m) \in \EuScript G^m$, we have % \begin{equation} \langle \Phi (\pm M),\, \det - 1\rangle = \mathrm{Card} (\EuScript V) ( (\pm 1)^N - 1). \end{equation} % By Proposition \ref{clasquotient}, this proves the assertion. \end{pf}
\section{Introduction} \label{usec1} In 1959, Buchdahl$^{\cite{Buch}}$ first pointed out a relativistic effect that a perfect fluid (uncharged) ball cannot have a non-singular static configuration while its radius $R$ is not larger than 9/8 of its horizon size $R_{g}$ which is twice as large as its mass $M$ in natural units, if the energy density is assumed not to increase outwards. For an incompressible ball, pressure singularity will emerge at its center if $R$=(9/8)$R_{g}$. In 1993, de Felice and Yu$^{\cite{Felice2}}$ found that if a configuration with inner (central) singular boundary is acceptable, the shell shaped static configuration can exist for any radius which is larger than $R_g$. Although some theoretically favorable arguments have been discussed$^{\cite {Felice3}\cite{Felice4}}$, the singular configuration seems exotic. For charged balls, the pressure gradient will be partially balanced by the electric force inside. Thus the central pressure of a dense configuration will be weakened and the pressure singularity might be avoided. In 1995, de Felice, Yu and Fang$^{\cite{Felice}}$ did find a series of non-singular static configurations with $R$ arbitrarily approaching the corresponding horizon size $R_{+}=M+\sqrt{M^2-Q_0^2}$ in the case that $Q_0\to M$, where $Q_0$ is the charge of the ball. The motivation of this work was to find the critical value of ratio $Q_{0}/M$ for the existence of series of regular static configurations with $R\to R_{+}$. Finally we find out that in case $Q_{0}/M$ is less than 1, no such configuration exists. That is to say, $Q_{0}=M$ is the only case in which the series of regular static configurations can contract to approach its horizon size. In section \ref{usec2}, the general analysis of the problem is given. In section \ref{usec3}, we pose the theorem and prove it. In section \ref{usec4}, the physical meaning of the result is discussed. \section{The Argument} \label{usec2} In the spherically symmetric case, the metric takes the form \begin{equation} ds^2=-e^\nu dt^2+e^\lambda dr^2+r^2d\theta ^2+r^2\sin^2\theta d\varphi ^2. \label{u1} \end{equation} Here and here after the natural units system with $G=c=1$ is used. Assuming the source is made of charged perfect fluid, Einstein's equations for a static configuration will have the form$^{\cite{Beken}}$: \begin{eqnarray} e^{-\lambda }\left(\frac{\lambda ^{\prime }}r-\frac 1{r^2}\right)+\frac 1{r^2} &=&\frac{% Q^2}{r^4}+8\pi \rho \label{uEin1}, \\ -e^{-\lambda }\left(\frac{\nu ^{\prime }}r+\frac 1{r^2}\right)+\frac 1{r^2} &=&\frac{Q^2% }{r^4}-8\pi p \label{uEin2}, \\ -\frac{e^{-\lambda }}2\left(\nu ^{\prime \prime }+\frac{\nu ^{\prime }{}^2}2+ \frac{\nu ^{\prime }-\lambda ^{\prime }}r-\frac{\nu ^{\prime }\lambda ^{\prime }}2\right) &=&-\frac{Q^2}{r^4}-8\pi p, \label{uEin3} \end{eqnarray} where $\rho(r)$ and $p(r)$ are its energy density and isotropic pressure respectively, and $Q(r)$ is the electric charge within the radius $r$. The primes here stand for the derivatives with respect to $r$. In general, an equation of state should be taken as input. Instead, we will consider the energy density $\rho(r)$ as an arbitrary input. The charge distribution should depend on the electromagnetic property of the medium. Since that property can be arbitrarily assigned, we will consider $Q(r)$ as another arbitrary input. Thus the set of equations (\ref{uEin1}) to (\ref{uEin3}) is a complete set for solving $\nu(r)$, $\lambda(r)$ and $p(r)$. For convenience, we define a new variable $m(r)$ as \begin{eqnarray} m(r) &=&\int_0^r4\pi x^2\rho (x)dx+\frac 12\int_0^r\frac{Q^2(x)}{x^2 }dx+\frac{Q^2(r)}{2r} \label{u5}. \end{eqnarray} Let $r=R$ be the surface of the ball, the global parameters are \begin{eqnarray} M &=&m(R), \label{u6} \\ Q_0 &=&Q(R), \label{u7} \end{eqnarray} where $Q_0$ is the total charge of the ball. By using the variable $m(r)$, equation (\ref{uEin1}) can be solved out as \begin{equation} e^{-\lambda(r)}=1-\frac{2m}r+\frac{Q^2}{r^2}.\label{u8} \end{equation} At the surface, it suits the Reissner-Nordstr\"om metric in standard form. Thus $M$ is called the total mass of the ball. By eliminating pressure $p$ from equation (\ref{uEin2}) and (\ref{uEin3}), we have \begin{equation} \left[e^{\nu -\lambda\over 2}\frac{\nu ^{\prime }}r\right]^{\prime }=\frac{1}{r^2}e^{ \lambda +\nu\over 2}\left(\frac{8Q^2}{r^3}-\frac{6m}{r^2}+8\pi \rho r\right). \label{u9} \end{equation} As mentioned, $\rho(r)$ and $Q(r)$ are considered as inputs, then (\ref{u9}) is a second order differential equation for $\nu(r)$ only. We want our interior solution suits the exterior Reissner-Nordstr\"om metric at the surface. So the boundary conditions for $\nu(r)$ are as what follow: \begin{eqnarray} e^{\nu(R)}&=&1-\frac{2M}{R}+\frac{Q_0^2}{R^2}, \label{u10} \\ \left[e^{\nu(R)}\right]^{\prime}&\equiv&{\left[e^{\nu(r)}\right]^{\prime}}_{r=R}={2\over R}\left({M\over R}-{Q_0^2\over R^2}\right). \label{u11} \end{eqnarray} We will study the cases with $Q_0<M$ only. In these cases, the exterior horizon size $R_+$ can be expressed by $Q_0$ and $M$ as \begin{equation} R_{+}=M+\sqrt{M^2-Q_0^2}. \label{u12} \end{equation} For assigned $R$, $Q_0$ and $M$, the solutions of equation (\ref{u9}) which satisfy the boundary conditions (\ref{u10}) and (\ref{u11}) will be called a Mathematical Solution Set (MSS). Each element in the set is defined by specified inputs of $\rho(r)$ and $Q(r)$. Surely, the solution is not always physically acceptable. We assign the physically acceptable conditions (PAC) as what follow: \begin{eqnarray} R&>&R_+, \label{u13} \\ \rho(r)&\ge& 0, \label{u14} \\ e^{\lambda(r)}&>&0, \label{u15} \\ e^{\nu(r)}&>&0. \label{u16} \end{eqnarray} These are all necessary conditions. However, we will prove that while the radius $R$ approaches its corresponding horizon size $R_+$, there will be no element in MSS which satisfies PAC. By other words, for $Q_0<M$, there is a $R_0>R_+$ such that NO STATIC PHYSICAL CONFIGURATION WITH $R<R_0$ EXISTS. Since only the cases with $Q_0<M$ will be considered and only $R>R_+$ needs to be studied, we further define \begin{eqnarray} Q_0&\equiv&\sqrt{1-\Delta^2}M, \label{u17} \\ R&\equiv&(1+\epsilon)R_+\ =\ (1+\epsilon)(1+\Delta)M, \label{u18} \end{eqnarray} where $0<\Delta\le 1$ and $\epsilon>0$. Then the boundary values of $e^\nu$ and $e^\lambda$ are \begin{eqnarray} e^{\nu(R)} &=&\frac{\epsilon +(2+\epsilon )\Delta }{(1+\epsilon )^2(1+\Delta )}\epsilon, \label{u19} \\ e^{\lambda(R)}&=&e^{-\nu(R)}, \label{u20} \\ \left[e^{\nu(R)}\right]'&=&\frac{2(\epsilon +\Delta)}{(1+\epsilon)^3(1+\Delta)^2M} \label{u21}. \end{eqnarray} While $\Delta$ is fixed as a finite quantity and $\epsilon$ approaches zero, we see that $e^{\nu(R)}$ is infinitesimal, $[e^{\nu(R)}]'$ is finite and $e^{\lambda(R)}$ is infinite. \section{The Proof} \label{usec3} The main theorem that we want to prove is as what follows: THEOREM: For $0<\Delta\le 1$, an $\epsilon_0$ (corresponding to $R_0$) can be found so that while $\epsilon<\epsilon_0$, any Element in MSS which satisfies PAC (\ref{u13}), (\ref{u14}) and (\ref{u15}) will always violates PAC (\ref{u16}). Some preparations are needed. For any element in MSS, integrating both sides of (\ref{u9}) from $r$ to $R$, we get \begin{equation} \left[e^{\frac \nu 2}\right]^{\prime }=\frac r2e^{\frac \lambda 2}\left[\frac{2}{R^2}\left({M\over R}- \frac{Q_0^2}{R^2}\right)+F_{(r)}\right], \label{u22} \end{equation} where the boundary conditions have been used and $F(r)$ is defined as \begin{equation} F(r)=-\int_r^R\frac{1}{x^2}e^{\frac{\lambda +\nu }2}\left[\frac{8Q^2}{x^3}- \frac{6m}{x^2}+8\pi \rho x\right]dx. \label{u23} \end{equation} We study the behavior of $e^{\nu(r)\over 2}$ and $F(r)$ from the boundary towards its center. At the boundary, we have $e^{\nu(R)}>0$ and $F(R)=0$. Therefore, there is an interval near $R$ in which \begin{eqnarray} e^{\nu(r)\over 2}&>&0, \label{u24} \\ F(r)&>&-{1\over R^2}\left({M\over R}-{Q_0^2\over R^2}\right). \label{u25} \end{eqnarray} LEMMA: For any element of MSS which satisfies PAC (\ref{u13}), (\ref{u14}) and (\ref{u15}), an $\epsilon_1>0$ can be found such that while $\epsilon<\epsilon_1$, if (\ref{u24}) remains valid in $[\beta R, R]$ and (\ref{u25}) is valid in $(\beta R, R]$, it is impossible to have \begin{equation} F(\beta R)=-{1\over R^2}\left({M\over R}-{Q_0^2\over R^2}\right), \label{u26} \end{equation} where $\beta<1$ is a positive number. We prove it in the following way: Suppose that $F(r)$ reaches $-{1\over R^2}\left({M\over R}-{Q_0^2\over R^2}\right)$ at $r=\beta R$, then an $\epsilon_1>0$ can be found such that while $\epsilon<\epsilon_1$, $e^{\nu(\beta R)\over 2}$ will be negative which contradicts (\ref{u24}). PROOF OF THE LEMMA: In view of (\ref{u25}), (\ref{u22}) shows that $[e^{\nu(r)\over 2}]'>0$ for $r\in(\beta R,R]$. Then we have \begin{equation} 0<e^{\nu(r)\over 2}\le e^{\nu(R)\over 2}\quad{\rm for\ }r\in[\beta R,R]. \label{u301} \end{equation} Rewrite (\ref{u5}) as \begin{equation} (2rm-Q^2)'=4\pi\rho r^2-{Q^2\over 2r^2} \label{u27}, \end{equation} then equation (\ref{u8}) leads to \begin{equation} \left[e^{-\frac{\lambda }2}\right]^{\prime }=e^{\frac \lambda 2}\left[\frac m{r^2}-4\pi \rho r-\frac{Q^2}{r^3}\right]. \label{u28} \end{equation} On the other hand, the equations (\ref{u8}) and (\ref{u14}) show \begin{eqnarray} e^\lambda\ge 1, \label{u29} \\ 1>{2m\over r}-{Q^2\over r^2}\ge 0. \label{u30} \end{eqnarray} By using formula (\ref{u28}), $F(r)$ turns to be \begin{eqnarray} F(r) &=&\int_r^R\frac{10}{x^2}e^{\frac \nu 2}d(e^{-\frac \lambda 2})+\int_r^R\frac{1}{x^2}e^{\frac{\nu +\lambda }2}\left[32\pi \rho x-\frac{4}{x^2}\left(m- \frac{Q^2}{2x}\right)\right]dx. \label{u31} \end{eqnarray} Integrating the first term of RHS by parts and using the boundary condition (\ref{u20}), we get \begin{eqnarray} F(r)&=&\frac{10}{R^2}e^{\nu(R)} -\frac{10}{r^2}e^{\frac{\nu(r)-\lambda(r)}2} +20\int_r^R\frac{1}{x^3}e^{\frac{\nu -\lambda }2}dx \nonumber \\ & &-10\int_r^R\frac{1}{x^2}e^{\frac{-\lambda}2}d(e^{\frac \nu 2})+\int_r^R \frac{1}{x^2}e^{\frac{\nu +\lambda }2}\left[32\pi \rho x-\frac{4 }{x^2}\left(m-\frac{Q^2}{2x}\right)\right]dx. \label{u32} \end{eqnarray} By taking away the positive terms in RHS, the equality turns to be an inequality: \begin{equation} F(r)>-\frac{10}{r^2}e^{\frac{\nu(r)-\lambda(r)}2} -10\int_r^R\frac{1}{x^2}e^{\frac{-\lambda }2}d(e^{\frac \nu 2}) -\int_r^R\frac{4}{x^4}e^{\frac{\nu +\lambda }2}\left(m-\frac{Q^2}{2x}\right)dx \quad {\rm for}\ r\in[\beta R,R].\label{u33} \end{equation} Due to inequalities (\ref{u301}), (\ref{u29}), (\ref{u30}) and $r\ge\beta R$, we further have \begin{equation} F(r)>-{20\over (\beta R)^2}e^{\nu(R)\over 2}-{2\over (\beta R)^3}e^{\nu(R)\over 2}\int_r^R e^{\lambda\over 2}dx \quad {\rm for}\ r\in[\beta R,R]. \label{u34} \end{equation} Suppose $F(r)$ reaches $-{1\over R^2}\left({M\over R}-{Q_0^2\over R^2}\right)$ at $r=\beta R$. Using (\ref{u34}), we get \begin{eqnarray} {1\over R^2}({M\over R}-{Q_0^2\over R^2})=-F(\beta R)<{20\over (\beta R)^2}e^{\nu(R)\over 2}+{2\over (\beta R)^3}e^{\nu(R)\over 2}\int_{\beta R}^R e^{\lambda\over 2}dr. \label{u35} \end{eqnarray} Now we consider the upper limit of $e^{\nu(\beta R)\over 2}$. By integrating (\ref{u22}) from $\beta R$ to $R$ and using the condition (\ref{u25}), we have \begin{eqnarray} e^{\nu(\beta R)\over 2}&=&e^{\nu(R)\over 2} -\int_{\beta R}^R\frac r2e^{\frac\lambda 2}\left\{\frac{1}{R^2}\left({M\over R}-\frac{Q_0^2}{R^2}\right) +\left[\frac{1}{R^2}\left({M\over R}-\frac{Q_0^2}{R^2}\right)+F(r)\right]\right\}dr \nonumber \\ &<&e^{\nu(R)\over 2}-{\beta\over 2R}\left({M\over R}-\frac{Q_0^2}{R^2}\right)\int_{\beta R}^R e^{\lambda\over 2}dr, \label{u36} \end{eqnarray} where $r\ge\beta R$ is also used. Solving out the integral $\int_{\beta R}^R e^{\lambda\over 2}dr$ from (\ref{u35}) and substituting the result into (\ref{u36}), we find inequality (\ref{u36}) becomes \begin{eqnarray} e^{\nu(\beta R)\over 2} &<&e^{\nu(R)\over 2}-{\beta\over 2}\left({M\over R}-\frac{Q_0^2}{R^2}\right)\left[{{\beta}^3\over 2}e^{-\nu(R)\over 2}\left({M\over R}-\frac{Q_0^2}{R^2}\right)-10\beta\right]. \label{u37} \end{eqnarray} So far, the radius $R$ is arbitrary. What we want to prove is that a $\epsilon_1$ (corresponding to $R_1$) does exist such that while $\epsilon<\epsilon_1$, the RHS of (\ref{u37}) will be always negative. This result implies the invalidity of (\ref{u26}). The negativity of the RHS of (\ref{u37}) means \begin{equation} 4e^{\nu(R)}\left({M\over R}-{Q_0^2\over R^2}\right)^{-1}+20\beta^2e^{\nu(R)\over 2} <\beta^4\left({M\over R}-{Q_0^2\over R^2}\right). \label{u38} \end{equation} Since it is a dimensionless formula, we express it by parameters $\Delta$ and $\epsilon$, that is \begin{equation} 20\beta^2\sqrt{\epsilon(\epsilon+2\Delta+\epsilon\Delta)(1+\Delta)}(1+\epsilon)(\Delta+\epsilon)+4\epsilon(\epsilon+2\Delta+\epsilon\Delta)(1+\Delta)(1+\epsilon)^2 <\beta^4(\epsilon+\Delta)^2. \label{u39} \end{equation} For $\epsilon=0$, the LHS is zero and RHS is a positive finite quantity, so the inequality (\ref{u39}) is valid. Therefore, a positive $\epsilon_1$ exists such that (\ref{u39}) is valid for any $\epsilon<\epsilon_1$. That ends our proof. COROLLARY: For $\epsilon<\epsilon_1$, if (\ref{u24}) is valid in the interval $[\beta R, R]$, then (\ref{u25}) must be valid in the same interval. There is no need to prove the corollary. We turn to prove the main theorem. PROOF OF THE THEOREM: For any element in MSS which satisfies PAC (\ref{u13}),(\ref{u14}) and (\ref{u15}), we want to prove that if PAC (\ref{u16}) is also assumed to be valid in some interval $[\beta R,R]$, for small enough $\epsilon$, a contrary result will emerge. By the corollary, the validity of PAC (\ref{u16}) leads to the validity of (\ref{u25}). Then we still have the inequality (\ref{u36}). Here we use $e^\lambda\ge 1$ to evaluate the integral in (\ref{u36}) as \begin{equation} \int_{\beta R}^R e^{\lambda\over 2}dr\ge (1-\beta)R. \label{u41} \end{equation} Substituting it into (\ref{u36}), we see that if \begin{equation} 2e^{\nu(R)\over 2}<\beta(1-\beta)\left({M\over R}-{{Q_0}^2\over R^2}\right) \label{u42} \end{equation} is valid, $e^{\nu(\beta R)\over 2}$ will certainly be negative. It is contrary to the validity of PAC (\ref{u16}). We rewrite inequality (\ref{u42}) by using parameters $\Delta$ and $\epsilon$ \begin{equation} 2(1+\epsilon)\sqrt{\epsilon(\epsilon+2\Delta+\Delta\epsilon)} <{\beta(1-\beta)(\epsilon+\Delta)\over \sqrt{1+\Delta}}. \label{u43} \end{equation} Evidently, it is valid for $\epsilon=0$. Then $\epsilon_2$ can be chosen as \begin{equation} \epsilon_2={\beta^2(1-\beta)^2\Delta^2\over 16(1+\Delta)(1+3\Delta)}. \label{u44} \end{equation} It is not difficult to see, (\ref{u43}) will always be valid for $\epsilon<\epsilon_2$. Finally, we have to choose $\epsilon_0$ as \begin{equation} \epsilon_0=\min\{\epsilon_1,\epsilon_2 \} \label{u45} \end{equation} to assure the validity of the lemma and its corollary. However, such a $\epsilon_0$ does exist. It ends the proof of our theorem. \section{Discussions} \label{usec4} A. Pressure singularity. As we know, for an incompressible perfect fluid (uncharged) ball, the pressure diverges and $e^{\nu}$ reaches zero at the center while $R={9\over 8}R_g$. If $R<{9\over 8}R_g$, the pressure singularity emerges at some $r_0>0$. While $R$ is smaller, the corresponding $r_0$ will be larger. In fact, the same thing happens in the charged case. Equation (\ref{uEin2}) can be rewritten as \begin{equation} \nu ^{\prime }=2\frac{4\pi rp+\frac m{r^2}-\frac{Q^2}{r^3}}{1-\frac{2m}r+ \frac{Q^2}{r^2}}. \label{u46} \end{equation} For each configuration which satisfies PAC (\ref{u13})---(\ref{u15}) and has null $e^{\nu \over 2}$ inside, suppose \begin{equation} \lim_{r\to {r_0}_+}e^{\nu(r)\over 2}=0 \label{u461} \end{equation} and $e^{\nu(r)\over 2}$ is positive for $r\in (r_0,R]$. If there is an upper limit for pressure $p$ in $[r_0,R]$, from equation (\ref{u46}), we see that $\nu(R)-\nu(r)$ will be finite while $r\to {r_0}_+$. So $e^{\nu(r)}$ approaches a positive quantity while $r\to {r_0}_+$ which contradicts (\ref{u461}). It implies that pressure $p$ must diverge at some point in $[r_0,R]$, and equation (\ref{u46}) shows that $\nu^{\prime }$ diverges at the same point. However, equation (\ref{u22}) shows that $[e^{\frac \nu 2}]^{\prime }={1\over 2}e^{\nu\over 2}\nu'$ is finite in $[r_0,R]$, then we have that $e^{\nu(r)\over 2}$ approaches zero while $p$ approaches infinity. Because $e^{\nu\over 2}$ is a positive function in $(r_0,R]$, we deduce that pressure approaches infinity at the same point where $e^{\nu \over 2}$ approaches zero for the first time from the boundary inwards. Qualitatively, pressure increases inwards from $p(R)=0$ and the electric force tends to weaken the increasing of pressure. As we expected, this effect may avoid the emergence of the pressure singularity. Our theorem shows that while $Q_0<M$, the electric force is not strong enough to avoid the pressure singularity. \medskip \noindent B. The case of $Q_0=M$. In this case, our proof fails to be valid since the $\epsilon_0>0$ cannot be found. In fact, the opposite can be proved. If the given $\rho(r)$ and $Q(r)$ satisfy \begin{eqnarray} e^{-\lambda }(r) &\ge &1-\frac{2r^2}{R^2}\left({M\over R}-{Q_0^2\over 2R^2}\right), \label{u47} \\ \frac{6m}{r^2} &\le &\frac{8Q^2}{r^3}+8\pi r\rho, \label{u48} \end{eqnarray} elements in MSS satisfying PAC (\ref{u13})---(\ref{u16}) do exist for $R\to R_+$. The solution found by de Felice et al. offers an example. Other series of models can also be produced. However, it is worthy of noting that PAC (\ref{u13})---(\ref{u16}) are necessary conditions only. So it does not imply that static physical configurations do exist while $R\to R_+$. \medskip \noindent C. The involved assumptions. We conclude that no static physical configuration exists for $R\to R_+$ if $Q_0<M$. Here only some naive assumptions are used. They are 1: the charged ball is made of perfect fluid, 2: The energy density $\rho(r)$ is non-negative everywhere. \noindent No particular assumption for property of the medium is involved.
\section{Introduction} A plasma instability theory for the prompt emission of gamma-ray bursts is presented.\cite{brainerd4} In this theory, a relativistic shell with $\Gamma \gg 1$ passes through the interstellar medium. Two plasma instabilities, the filamentation instability and the two-stream instability, generate a magnetic field and heat the electrons to relativistic energies. The heated electrons emit synchrotron radiation in the radio to optical bands and synchrotron self-Compton radiation from the optical to gamma-ray bands. This theory produces the observed prompt gamma-ray emission seen in all bursts, and the prompt optical emission seen in GRB~990123. The magnetic field generated by the filamentation instability is calculated from first principals. Lower limits on $\Gamma$ and $n_{ism}$ arise from the requirement that the model efficiently radiate gamma-rays. The limit on density requires each gamma-ray burst to be surrounded by a medium that is optically thick to Compton attenuation. The limit on $\Gamma$ suggests that there exists a class of transient that produces optical and ultraviolet emission but no gamma-ray emission. One of the more interesting aspects of the theory is that the plasma instabilities cannot satisfy the Rankine-Hugoniot conditions, so a shock is not produced by these instabilities. As a consequence, the interstellar medium remains in place after passage of the relativistic shell. This permits the interstellar medium to interact with multiple relativistic shells to produce the complex time structure seen in gamma-ray burst light profiles. \section{Plasma Instabilities} Two instabilities arise when a plasma streams through a second plasma at a highly relativistic velocity. The first is the filamentation instability, while the second is the two-stream instability.\cite{davidson1} Of these, the former has the higher growth rate. The growth rate of the filamentation instability as measured in the shell rest frame is \begin{equation} \gamma^{\prime}_{f} \approx {1 \over 2} \sqrt{ { 4 \pi e^2 \over m } } \, n_{ism}^{{1 \over 2}} \, , \end{equation} where $n_{ism}$ is the number density of the interstellar medium in the ISM rest frame, and $m$ is the mass of the filamenting plasma component of the ISM. The filamentation occurs for wave numbers perpendicular to the velocity vector that obey the inequality \begin{equation} k_{\perp} > \omega_{p,e,shell}^{\prime}/c \, . \end{equation} In other words, the length scale of the filamentation is set by the electron plasma frequency of the shell as measured in the rest frame of the shell. \begin{figure} \centering \epsfig{figure=06_JBrainerd1_1.ps, width=4.0in} \caption{Fraction of energy that goes into magnetic field and ion thermalization through the filamentation instability. This is measured in the rest frame of the shell. The energy available is the energy of the interstellar medium streaming through the shell. The curves are plotted as the ratio $n_{ism}/n_{shell}^{\prime}$. The solid curve gives the fraction of energy in magnetic field. This fraction is independent of $\Gamma$. The dotted curve gives the fraction of energy that goes into ion thermal energy. The upper curve is for $\Gamma = 10^4$, while the lower is for $\Gamma = 10^3$. The squares mark the value of $\eta$ that one expects from the Rankine-Hugoniot relations for $\Gamma = 10^3$, while the diamonds mark this relation for $\Gamma = 10^4$. } \end{figure} The filament is a magnetic pinch, with a toroidal magnetic field that is formed when the particles in the filament collapse to the center of the filament. The filament grows until its growth rate equals the bounce frequency of a particle across the magnetic pinch. This saturation defines the maximum magnetic field that can be generated.\cite{davidson2,lee} For both ions and electrons, the maximum field strength is the same \begin{equation} {B^{\prime \, 2} \over 8 \pi } = { m_e c^2 n_{ism}^2 \Gamma^2 \over n_{shell}^{\prime} } \end{equation} One finds that for $n_{shell}^{\prime}/n_{ism} = \Gamma$, which is a natural value for the efficient emission of radiation, the magnetic field is $B^{\prime} = 0.14 \, \hbox{G}$ when $\Gamma = 10^3$ and $n_{ism} = 1 \, \hbox{cm}^{-3}$, and $B^{\prime} = 45.4 \, \hbox{G}$ when $\Gamma = 10^3$ and $n_{ism} = 10^5 \, \hbox{cm}^{-3}$. The energy extracted from the interstellar medium and converted into magnetic and thermal energy is small. This is shown in Figure~1. As a consequence, even when the Rankine-Hugoniot condition on $n_{shell}^{\prime}$ of $n_{shell}^{\prime}/n_{ism} = \Gamma$ is satisfied, the fraction of kinetic energy that is converted to thermal and magnetic energy is tiny, so that the interstellar medium continues to stream through the shell. \begin{figure} \centering \epsfig{figure=06_JBrainerd1_2.ps, width=4.0in} \caption{Characteristic observed photon energies for $n_{ISM} = 1 \, \hbox{cm}^{-3}$. The solid curves give characteristic energies for single scattering synchrotron self-Compton emission. The dotted curves are for synchrotron emission. The upper pair are for $\Gamma = 10^4$, while the lower pair are for $\Gamma = 10^3$. Note that each of these curves peak at $\eta \approx \Gamma^{-1}$. Only the synchrotron self-Compton emission is capable of producing gamma-ray emission. } \end{figure} \begin{figure} \centering \epsfig{figure=06_JBrainerd1_3.ps, width=4.0in} \caption{Characteristic observed photon energies for $n_{ISM} = 10^{5} \, \hbox{cm}^{-3}$. These curves are as in Figure 3a. As in the previous case, only the synchrotron self-Compton emission is capable of producing gamma-ray emission, although the synchrotron emission can produce hard x-ray emission for case of $\Gamma = 10^4$. } \end{figure} The two-stream instability acts on the electrons to bring them into an equilibrium with the two ion streams. This instability grows at the rate of \begin{equation} \gamma^{\prime}_{2s} \approx {1 \over 2} \sqrt{{ 4 \pi e^2 \over m_e}} n_{ism}^{{1 \over 2}} \Gamma^{-1} \, . \end{equation} The growth rate of the electron two-stream instability is smaller than the ion filamentation instability by the factor of $m_{p}^{1/2}/m_e^{1/2} \Gamma \approx 0.04$. The two stream instability will grow until the electron Lorentz factor is of order $\Gamma$. At this point, the system becomes stable. This instability therefore only converts $m_e/m_p$ of the total energy of the interstellar medium into thermal energy. \section{Radiative Mechanisms} The two radiative mechanisms at work in this theory are synchrotron emission and synchrotron self-Compton emission. The characteristic energy of each process is given in Figures 2 and 3 for $n_{ism} = 1 \, \hbox{cm}^{-3}$ and $n_{ism} = 10^5 \, \hbox{cm}^{-3}$. This characteristic energy defines the maximum energy of the continuum as defined by the maximum energy of the electron distribution. The minimum energy of the synchrotron emission is given by the cyclotron energy, which is \begin{equation} \nu = 2.68 \times 10^{8} \, \hbox{Hz} \, n_{ism}^{{1 \over 2}} \Gamma_3^{{3 \over 2}} \, , \end{equation} where $\Gamma_3 = \Gamma/10^3$. This value is smaller than the characteristic synchrotron energy by the factor of $\Gamma^2$. The synchrotron self-Compton energy range is determined by a single scattering, since more than one scattering takes the photons to the characteristic electron energy. The minimum energy is set by $\Gamma^2$ times the cyclotron resonance energy, so that the low end of the synchrotron self-Compton continuum overlaps the high end of the synchrotron continuum. The characteristic Compton scattering energy is a factor of $\Gamma^2$ larger than this. From the figures, one sees that for $\Gamma > 10^3$, the energy range of the Compton scattered radiation extends above $1 \hbox{MeV}$. For a lower value of $\Gamma$, the cooling occurs predominately at optical and ultraviolet wavelengths. Because gamma-ray bursts are identified by their gamma-ray emission, bursts with $\Gamma < 10^3$ will not be observed. This suggests that there exists a class of burst phenomena with optical and ultraviolet emission, but no gamma-ray emission. \section{Selection Effects} The emission of gamma-rays defines one selection effect. In order to produce gamma-rays, $\Gamma > 10^3$. This provides an explanation for the absence of photon-photon pair creation in gamma-ray bursts, as demonstrated by the absence of gamma-ray bursts with thermal spectra. \begin{figure} \centering \epsfig{figure=06_JBrainerd1_4.ps, width=4.0in} \caption{Lower limits on the interstellar medium number density. This is as a function of the shell Lorentz factor $\Gamma$. The solid curve is for radiative processes radiating at a rate equal to the rate at which energy is transferred to electrons from the streaming ions. The dotted curve is for radiative processes radiating at 10\% of the rate at which energy is transferred to electrons. To efficiently radiate energy, the interstellar medium number density must be above these curves. The provides a selection effect: only burst sources surrounded by high density material will produce observable gamma-ray bursts. } \end{figure} A second selection effect that comes into play in this theory concerns the efficient production of gamma-rays. If the density is too low, the rate of radiative cooling is far below the rate at which electrons are thermalized. Such bursts would be dim relative to their high-density counterparts. The most observable bursts are therefore those with a sufficiently high $n_{ism}$ to efficiently convert electron thermal energy into gamma-rays. This lower limit on $n_{ism}$ can be written as \begin{equation} n_{ism} > 8.22 \times 10^{4} \hbox{cm}^{-3} {\cal M}_{27}^{-{1 \over 2}} \Gamma_3^{-{13 \over 4}} f_{emis}^{-{3 \over 2}} \left( { R \over R_0 } \right)^{{3 \over 4}} \end{equation} where $f_{emis}$ is the fraction of electrons at the characteristic energy, where $R/R_0 \approx \left( m_p/m_e \right)^{1/3}$ is the distance traveled when $\Gamma$ drops by a factor of 2 relative to the distance traveled when the ISM is swept up, all under the assumption that the heating of the electrons represent the maximum energy loss, and where ${\cal M}_{27}$ is the mass of the shell per unit ster radian in units of $10^{27} \, \hbox{gm}$. The limit given by this last equation is plotted in Figure 4 as a solid line. A similar curve that gives emission that is 10\% effective at radiating the energy converted into thermal energy is plotted as a dotted line. The importance of this limit is that the gamma-ray burst mechanism requires a high density interstellar medium to operate efficiently. When the burst radiates efficiently, the interstellar medium is of sufficient density to attenuate the gamma-ray spectrum through Compton scattering.\cite{brainerd1} This provides the plasma instability theory with a mechanism that gives the spectrum a characteristic energy of several hundred keV, despite the broad spectral range of the synchrotron self-Compton continuum, and its strong dependence on $\Gamma$. Such a mechanism is required to correctly reproduce the observed spectra.\cite{brainerd2,brainerd3} \section{Discussion} The plasma instability theory discussed above is a new mechanism to produce prompt gamma-ray burst emission. It has several unique features. \begin{itemize} \item The plasma instability theory produces the prompt gamma-ray emission without creating a shock. This implies that additional gamma-ray bursts can occur in the interstellar medium when new relativistic shells are ejected by the source, because the region is not cleared of material. \item The requirement that the mechanism efficiently produce gamma-rays introduces selection effects, so that bursts have $\Gamma > 10^3$ and $n_{ism} > 10^5 \hbox{cm}^{-3}$. \item There exists a class of optical transient that has no gamma-ray emission. These bursts differ from gamma-ray bursts in having $\Gamma < 10^3$. \item The lower limit on $n_{ism}$ ensures that gamma-ray bursts are always in region in which Compton attenuation by the surrounding interstellar medium occurs. \end{itemize} The next step in developing this theory is to undertake a numerical study of the plasma instabilities. The goal of the study will be to confirm the analytic results discussed above, and to provide a precise calculation of the electron distribution produced by the instabilities. A second aspect of the theory that will be examined is the afterglow radiation produced in this theory; the cooling of the region behind the relativistic shells gives a light curve that differs from the light curve created by the decelerating shell, adding a complexity to the afterglow from this theory that is not present in the shock theories of afterglow radiation. Finally, calculations of model spectra and their comparison to observed spectra is will be undertaken. In particular, a comparison of the optical and gamma-ray spectrum of this model to that of GRB~990123\cite{akerlof} is planned. This study will test whether the prompt optical and gamma-ray emission are part of a single synchrotron self-Compton continuum. \section*{References}
\section{Introduction} Since Penzias \& Wilson published their groundbreaking report of the existence of the CBR (1965), experimentalists have been checking to see if the CBR is polarized. CBR polarization experiments are challenging. Predicted polarization signals are an order of magnitude below the levels at which CBR anisotropy experiments have only just begun to detect signals. Typical polarization signals are a few $\mu$K. The most significant challenges include achieving adequate statistical sensitivity, limiting systematic errors sufficiently to detect the small polarization signal, and discriminating the CBR polarization from foreground sources such as galactic synchrotron radiation. To date, no polarization has been detected, but several experiments which have just begun operation or plan to within the year are designed to reach the sensitivities required to detect polarized signals of about the size predicted by CDM models. Figure \ref{fig-lvsnu} shows the angular resolution and frequency coverage of the new and recent experiments. Some details about the experiments are listed in Table \ref{tbl-1}. \begin{figure}[p] \epsfxsize=2.75cm \epsfbox[70 520 170 620]{figure1.eps} \vspace{5.0in} \caption{ A figure schematically indicating the multipole and frequency coverage of current and upcoming experiments. The frequency coverage shown indicates the {\em range} of frequencies probed by the experiments; in no case is the frequency coverage continuous. See Table \ref{tbl-1} for more information. The lower figure shows the E-polarization spectrum and the absolute value of the TE correlation expected for a typical CDM model. (The actual parameters are $\Omega_b = 0.05$, $\Lambda = 0$, $h = 0.65$, and $\Omega_{tot}=1$.) For an explanation of ``E" polarization, see, for example, Zaldarriaga \& Seljak, 1997. The spectra were generated with CMBFAST (Seljak \& Zaldarriaga 1996).} \label{fig-lvsnu} \end{figure} After a brief inspection of Figure \ref{fig-lvsnu}, one sees that the experimentalist is faced with a difficult choice. Should s/he design an experiment to search for a CBR polarization signal at large angles, where the spectrum codes information about reionization and the existence of tensor fluctuations? Such an experiment has the potential of enormous scientific payback if the CBR exhibits the coherent oscillation anisotropy spectrum predicted by models with adiabatic initial fluctuations amplified by a period of inflation. Also, a horn antenna can be used to attain beamwidths of a few degrees, without the need for reflectors which might introduce certain systematic errors. However, the aforementioned models predict that the CBR polarization anisotropy will be very small, less than $0.1$~$\mu$K, at angular scales larger than a couple of degrees. Extant limits at those angular scales are more than three orders of magnitude greater than the expected signal. The other choice for the experimentalist is to design an experiment to probe the polarization anisotropy at smaller angular scales, where the predicted signal from standard CDM is larger. These experiments must contend with polarized offsets due to the use of reflectors (or lenses). At present, groups are pursuing both lines of attack, as indicated in Figure \ref{fig-lvsnu}. The space-based MAP and Planck satellites plan to collect data from large fractions of the sky with small resolutions, which allows probing of small and large angular scales at once. Ground-based experiments complement the satellite missions because of their ability to probe specific regions of sky more deeply. MAP, for example, is expected to only make a statistical detection of the polarization anisotropy, while ground-based experiments should be able to detect polarization directly. Also, since no detections of polarization have yet been claimed, experimenters and analysts may not yet have beaten down all the relevant systematic errors which will plague the measurements; the ground-based experiments will address systematic effects as they become evident. Experience on the ground may prove helpful to the satellite experiments. A further distinction is that the current crop of ground-based experiments compare the orthogonal linear polarization components within a single beam at a time rather than comparing them between beams separated on the sky. \begin{figure}[p] \epsfxsize=3.0cm \epsfbox[120 520 220 620]{figure2.eps} \vspace{6.0in} \caption{The state of the field: approximate experimental limits on the CBR polarization. The same CDM model as in Figure \ref{fig-lvsnu} is sketched in for reference.}\label{fig-polimit} \end{figure} \section{Overview of Techniques} Figure \ref{fig-polimit} presents current experimental limits on the CBR polarization. The first experiments dedicated toward detecting or limiting the CBR polarization (Caderni et al.\ 1978; Nanos 1979; Lubin \& Smoot 1979, 1981; Lubin, Melese \& Smoot 1983) modulated the polarization of the incoming signal (either with a Faraday switch or a mechanical rotating analyzer) and locked in at the modulation frequency to measure $Q$ or $U\!$, depending on the orientation of the polarimeter. The increased sensitivity of bolometer-based and HEMT\footnote{HEMT stands for high electron mobility transistor; amplifiers using HEMTs have very low noise (Pospieszalski 1992, 1995).}-amplifier-based radiometers being built today is such that the systematic effects associated with the Faraday switch may be too large to tolerate. (Faraday switches are very sensitive to temperature variations and external magnetic field variations.) The modulation frequency of a mechanically rotating analyzer is limited in practice. Several new methods of Dicke-switching are described below. One other experimental limit appears in Figure \ref{fig-polimit}, the limit from the Saskatoon anisotropy experiment (Wollack et al.\ 1993), at $\ell\approx 80$. This experiment collected temperature differences between patches on the sky, $\Delta T_i$. Measurements of $\Delta T_i$ were made in two orthogonal linear polarizations, so that the polarization anisotropy limit comes from comparing the two sets of $\Delta T_i$. Thus, the variance of the polarization anisotropy measured includes a component due to the spatial variation of the (unpolarized) atmosphere. In other words, the best limit on the plot comes from an experiment optimized to measure the temperature anisotropy, not the polarization!\footnote{In fact, later data from Saskatoon at a slightly larger $\ell$ give an even lower limit to the polarization, but the full analysis has not yet been completed. (Page 1999).} The current crop of polarization experiments (with the exception of some satellite missions which do not have to contend with atmospheric spatial variations) measure the difference between two orthogonal linear polarizations within a single beam at a time. Receivers used in the current experiments dedicated to polarimetery of the CBR may be divided into two categories: 1) bolometer pairs read out in an AC bridge, such as the bolometer-bridge polarimeter invented for use in the Polatron experiment described below and 2) HEMT-amplifier correlation receivers. Correlation receivers (Fujimoto, 1964; Rohlfs \& Wilson, 1996) comprise two types also: a) receivers in which the two signals proportional to the electric fields from the two orthogonal polarizations are multiplied together directly in a nonlinear device and b) receivers in which the two signals are split and recombined with appropriate phase shifts into four signals which are subsequently detected and differenced to give $Q$ and $U$. The former type of correlation receiver is typically less bulky and capable of larger bandwidths than the latter, but requires care to ensure sufficient dynamic range. All these receivers may be Dicke-switched rapidly. In particular, the correlation receivers may be phase-switched at several kilohertz, fast enough to overcome the $1/f$ knee of even the highest frequency (90~GHz) HEMT amplifiers. The switching is achieved by inserting a relative phase shift of $0\deg$ or $180\deg$ between the two incoming signals; the output of a correlation receiver is proportional to $\cos\theta$, where $\theta$ is the total phase shift between the two incoming signals. Direct-multiplication correlators are used in the POLAR and PIQU experiments, while the Milano polarimeter and the SPOrt experiment use detect-and-difference correlation schemes. (Peter Timbie pioneered the use of direct-multiplication correlation receivers for CBR anisotropy in the 1980's: reference Timbie.) \section {Polarization Experiments at the Turn of the Millenium} Due to the authors' familiarity with three of the current ground-based polarization experiments, we devote unequal attention to those in what follows. For completeness, we mention five other experiments of which we are aware. The frequency coverage and $\ell$ coverage of these experiments is depicted in Figure \ref{fig-lvsnu}. Much of the other pertinent information is condensed in Table \ref{tbl-1}. First we discuss ground-based experiments, beginning with the experiment sensitive to the smallest angular scales. Then we discuss space-based experiments briefly. \subsection{The VLA 8.4~GHz CBR Project.} Partridge et al.\ (1997) recently completed imaging and analysis of a 40~arcmin$^2$ field at 8.44~GHz with 6\arcsec resolution, using the VLA. The signals were collected in circular polarization, and all four Stokes parameters recovered. The data came from 159 hours of observations; such a project is unlikely to be repeated. This work follows up on earlier work at lower resolution (Fomalont et al.\ 1993) and at lower frequency (Partridge et al.\ 1988). These three papers contain the only published CBR polarization limits from interferometry to date; the results are summarized in Figure \ref{fig-polimit}. \subsection{The Polatron.} The Polatron experiment is being built by a team comprising members from Caltech, Stanford University and Queen Mary and Westfield College. The Polatron will be used to search for CBR polarization on arcminute angular scales where the amplitude of the polarization power spectrum is expected to peak. Observations will be made at a frequency of 100\,GHz where confusion from polarized foregrounds is expected to be a minimum. The corrugated entrance feed of the Polatron will be located at the Cassegrain focus of the 5.5\,m dish at the Owens Valley Radio Observatory (OVRO), generating a $2.5'$ beam on the sky. A broad-band OMT (orthomode transducer, see Chattopadhyay et al.\ 1999) splits the incoming signal into two orthogonal linear polarizations. The backend of the polarimeter feeds two bolometers, each of which detects one of the linear polarization states that comes from the OMT. Metal-mesh resonant filters located between the OMT and the bolometers define a 20\% passband centered at 96\,GHz. An AC-bridge readout circuit (see, for example, Holzapfel et al.\ 1997) will be used to difference the outputs from the two bolometers. At any instance, this differencing scheme permits the measurement of a single Stokes parameter (Q or U). By rotating the plane of polarization by 45$^\circ$ with a quartz half-wave plate located in front of the entrance feed, a second Stokes parameter (U or Q) is then measured. The silicon nitride metal-mesh bolometers (Bock et al.\ 1996) that will be used in the Polatron are identical to those that will be flown as part of the Planck Surveyor High Frequency Instrument (HFI) in 2006. The bolometers will be operated at 250\,mK where they will have NEPs significantly lower than the background photon noise limit. In 1 second of integration on a single $2.5'$ pixel, the Polatron is expected to measure each Stokes parameter to a precision of 700\,$\mu$K. The Polatron will be commissioned in the summer of 1999 and will make its first observations in the winter of 1999/2000. In 6 months it is expected that the Polatron will observe 850 $2.5'$ pixels to a sensitivity of 8\,$\mu$K per pixel in each of Q and U. In a standard CDM model, this will be sufficient to detect rms polarization at 5$\sigma$. Over its projected four-year lifetime, the Polatron plans to map out the polarization anisotropy between $\ell=200$ and $\ell=2600$. \subsection{The PIQU Experiment.} A group at Princeton University has designed an experiment nicknamed PIQU, for Princeton IQU, a reference to the Stokes parameters the experiment eventually plans to measure. PIQU will measure the polarization at small angular scales at two frequencies, 40~GHz and 90~GHz, with multiple horns in the focal plane. The experiment uses a 1.4~m off-axis parabola fed with a corrugated horn antenna to provide a beamwidth of $0\fdg 23$ for the 90~GHz radiometer. The RF signals from the two arms of an OMT, corresponding to two linear orthogonal polarizations from the sky, are mixed down to a 2--18~GHz IF, split into three sub-bands, and then directly multiplied together in a broad bandwidth mixer. The front end of the instrument uses HEMT amplifiers cooled to about 12~K. Phase-switching at 4~kHz is done in one LO line. Phase tuners in the IF lines and the LO line allow balancing of the phase to $\pm 20\deg$, such that the bandwidth degradation is less than 10\%. The phase-one instrument, PIQ-90, measures only one of $Q$ and $U$, chosen to be the polarization with the most symmetry with respect to the experimental apparatus, in an effort to minimize systematic effects. This instrument operates at 90~GHz and will be deployed to the roof of Princeton in the spring of 1999. A separate 40~GHz cryostat is being developed and will be operated in a follow-up data run subsequently. The 40~GHz receiver shares the IF components of the 90~GHz receiver. In subsequent years, multiple horns will be placed in the focal plane, using a larger cryostat, so that multiple multifrequency measurements may be made simultaneously. \subsection {The POLAR Experiment.}\label{sec-polar} The Polarizaton Observations of Large Angular Regions (POLAR) experiment has been designed and built by the Observational Cosmology Group at the University of Wisconsin--Madison. POLAR is designed to measure the CBR's $Q$ and $U$ Stokes parameters in several broad bands between 26 and 100 GHz. POLAR observes the CBR polarization at large ($7^{\circ}$ FWHM) angular scales--comparable to the COBE satellite. At these angular scales the rms CBR polarization signal is expected to be quite small ($<1 \mu$K) unless the Universe was reionized at an early epoch corresponding to a $z\sim100$. The raw system noise of POLAR is a factor of 100 smaller than the instrument that was used to set the current upper limits on large angular scales (Lubin, Melese \& Smoot, 1983). The primary goal of POLAR is to reach a sensitivity level of $\Delta$T$_{Pol}/$T$_{CBR}\le 10^{- 6}$. At this level, POLAR will either detect polarization in the CBR or place a tight constraint on the epoch of reionization. POLAR employs a correlation radiometer that uses a corrugated feed horn to couple the CBR into an orthomode transducer which decomposes the incoming radiation into two linear polarizations. These two linear polarized components are amplified using two HEMT amplifiers cooled to $\sim$15 K in a cryocooler. After an additional stage of ambient temperature amplification, the two parallel signal chains are mixed down to an intermediate frequency band (2-12 GHz) where they are multiplied. The IF is subdivided into three equal bands for additional foreground discrimination. The output of the correlator is proportional to $Q\sin 2\alpha +U\cos 2\alpha$ where $\alpha$ is the orientation of the polarimeter. POLAR began observations in the Ka-band (26-36 GHz) in September 1998 at the University of Wisconsin's Pine Bluff Observatory. POLAR observes $\sim 36$ spots directly overhead in a strip at a declination of $43\deg$. Later in 1999, POLAR will add either a W-band (90-100 GHz) or Q-band (35-45~GHz) polarimeter (in the same cryostat) which will observe simultaneously with the Ka-band system. This additional polarimeter will help significantly in discriminating against foreground contamination. For additional information regarding POLAR, see Keating et al.\ 1998 and {\it http://cmb.physics.wisc.edu/polar/.} Future plans for the POLAR team include collaborating with UC-Santa Barbara to develop a small angular scale ($\approx 10'$) polarimeter. \subsection{The Milano Polarimetry Project.} Sironi et al.\ (1998) have built and operated a 33 GHz polarimeter coupled to a $14\deg$ corrugated feed horn. The correlation receiver uses a phase discriminator comprising several $90\deg$ hybrid tees (which output an incoming signal on two ports which have a $90\deg$ phase difference between them) and one $180\deg$ hybrid tee. Both $Q$ and $U$ are observed simultaneously. An extension to the feed horn allows data to be taken at $7\deg$; data have also been collected in this mode (Sironi 1999). Future plans include mounting the polarimeter at the Cassegrain focus of the 2.6 m dish at Testa Grigia (Sironi et al. 1998), which would allow measurements with a $1\deg$ beam. \subsection{Space-based Experiments.} The MAP and Planck Surveyor satellites are designed to measure the CBR temperature anisotropy across the entire sky. Both will also measure polarization. MAP, which is set to launch in fall of 2000, should make a statistical detection of polarization anisotropy. Planck, with proposed launch date of 2007, will be sensitive enough to detect polarization directly for most CDM models. SPOrt is an experiment dedicated to measuring CBR polarization. SPOrt is meant to be deployed on the International Space Station in 2001 or 2002. The MAP satellite (Jarosik et al. 1998) will use polarization-sensitive radiometers to observe the whole sky, with a frequency-dependent resolution ranging from $0\fdg21$ to $0\fdg93$. Data will be collected in five bands, from 22~GHz to 90~GHz. The receivers are phase-switched at 2.5~kHz, and use HEMT amplifiers. Corrugated feed horns couple to OMTs to separate two orthogonal linear polarizations. The primary reflectors are back to back $1.4\mbox{ m }\times1.6\mbox{ m}$ dishes. The difference data consist of $T_A - T_{B^\prime}$ and $T_B - T_{A^\prime}$ where $A$ and $A^\prime$ designate the two orthogonal linear polarizations from horn A, and similarly for $B$ and $B^\prime$. From such pairs of numbers one can derive a full-sky map of the polarization of the sky (Wright, Hinshaw, \& Bennett 1995). Most of the radiometers are completely constructed and tested. More information is available at {\it http://map.gsfc.nasa.gov/html/technical\_info.html}. The Planck satellite (Tauber 1998) has two sets of receivers, the LFI set and the HFI set, for ``Low-Frequency Instrument" and ``High-Frequency Instrument." The HEMT-based LFI radiometers are all linearly polarized, and include four bands between 30 and 100~GHz. The angular resolution for the LFI ranges from $0\fdg55$ to $0\fdg20$. The HFI receivers use bolometers. Three of the six HFI frequency bands are linearly polarized (143~GHz, 217~GHz and 545~GHz). The resolution for polarized bands ranges from $0\fdg13$ to $0\fdg08$. More information is available at {\it http://astro.estec.esa.nl/SA-general/Projects/Planck/}. The SPOrt project (Cortiglioni, et al. 1998) aims to measure the polarization of over 80\% of the sky in $7\deg$ patches in several frequency bands between 20 and 90~GHz. The receivers are correlation receivers of the detect-and-difference variety, using phase-switching, and measuring $Q$ and $U$ simultaneously. \section{Summary and Discussion} In addition to the CBR brightness spectrum and the temperature anisotropy, the CBR polarization represents the third treasure trove of information encoded in the CBR. The small polarization signal raises large barriers that undermine the ability to extract this information. These barriers include obtaining the requisite statistical sensitivity, minimizing systematic errors, and discriminating against foreground contamination. (See Tegmark, these proceedings.) The experiments described above overcome these barriers in a variety of different ways. Long integration times (roughly 10 hours/pixel/detector) with ever more sensitive detectors will yield the statistical sensitivity needed to detect CBR polarization. These experiments will have to minimize systematic effects at levels that have yet to be charted. The minimization of these systematic effects impacts the design and implementation of experiments in many different ways as demonstrated by the variety of experiments described above. The foregrounds that could contaminate CBR polarization are not well characterized. Polarized galactic synchrotron radiation is the only known foreground, and its intensity (much less its polarization) has yet to be measured at high galactic latitudes at the relevant frequencies. The best way to marginalize this ignorance is to perform multifrequency measurements, as temperature anisotropy experiments have shown. Since the {\em best} overall approach has yet to be determined, the community will invariably benefit from the variety of experimental approaches described above. \begin{table} \caption{Parameters of ongoing CBR polarization experiments are compiled here. References are given in the text, except in those cases where the information is communicated by the authors.} \label{tbl-1} \begin{center}\scriptsize \begin{tabular}{lllll} Experiment & beamsize\tablenotemark{a} & frequency & Receiver\tablenotemark{b} & Site \\ & & GHz & &\\ \tableline VLA & $0\fdg 02$ & 8.44 & int & NM desert \\ POLATRON & $0\fdg 04$ & 96 & bolo br & OVRO \\ PLANCK HFI & $0\fdg 08$ & 143, 217, 545 & bolo & space L2 \\ PIQU & $0\fdg 22$ & 40, 92 & HCdm & Princeton, NJ \\ PIQU2\tablenotemark{c} & $\la 0\fdg 2$ & 40, 92 & HCdm & high alt site \\ MAP & $0\fdg 23$ & 22, 33, 40, 61, 98 & HTP & space L2 \\ PLANCK LFI & $0\fdg 20$ & 30,44,70,100 & HTP & space L2 \\ POLAR & $7\deg$ & 30 \& 40 or 90 & HCdm & Madison, WI\\ SBUW\tablenotemark{c} & $\la 0\fdg 2$ & 30 \& 40 or 90 & HCdm & high alt US \\ SPORT & $7\deg$ & 22, 32, 60, 90 & HCdd & space station \\ Milano & $7\deg, 14\deg$ & 33 & HCdd & Antarctica\\ Milano2\tablenotemark{c} & $1\deg$ & 33 & HCdd & Alps\\ \end{tabular} \end{center} \tablenotetext{a}{The smallest beamsize is given; most of the experiments have frequency-dependent beamsizes.} \tablenotetext{b}{Receiver types include {\bf int} (inteferometer), {\bf bolo br} (bolometers in AC bridge), {\bf bolo} (bolometers), {\bf HCdm} (HEMT correlation receivers with direct multiplication), {\bf HCdd} (HEMT correlation receivers using recombination of signals with appropriate phases, followed by detection and differencing), and {\bf HTP} (HEMT total power -- see text and the MAP website for more information.)} \tablenotetext{c} {These three experiments are planned upgrades to existing experiments; the exact parameters are not yet determined. SBUW is a collaboration between UC--Santa Barbara and the existing POLAR team.} \end{table}
\section{Introduction} Galaxy clusters play an important role in the models for structure formation based on the gravitational instability hypothesis. They are the most extended gravitationally bound systems in the Universe. For this reason the study of their properties is a useful tool to constrain the parameters entering in the definition of the cosmological scenarios. In particular, their abundance and spatial distribution (also as a function of redshift) have been used to obtain estimates of the mass fluctuation amplitude and of the density parameter $\Omega_0$ (e.g. Eke, Cole \& Frenk 1996; Viana \& Liddle 1996; Mo, Jing \& White 1996; Oukbir, Bartlett \& Blanchard 1997; Eke et al. 1998; Sadat, Blanchard \& Oubkir 1998; Viana \& Liddle 1999; Borgani, Plionis \& Kolokotronis 1999; Borgani et al. 1999). In the past years many different groups have compiled deep cluster surveys in the optical band, which have been used to compute the clustering properties of galaxy clusters. The first results showed that clusters are strongly correlated, with a correlation length $r_0\approx 20-25\ h^{-1}$ Mpc, a factor 4-5 larger than that obtained for local galaxies (e.g. Bahcall \& Soneira 1983; Postman, Huchra \& Geller 1992). Here $h$ represents the Hubble constant $H_0$ in units of 100 km s$^{-1}$ Mpc$^{-1}$. Sutherland (1988) suggested the existence of a possible strong effect due to the spurious presence of galaxies acting as interlopers (see also Dekel et al. 1989; van Haarlem, Frenk \& White 1997). New analyses of optical catalogues, taking into account this projection effect (Dalton et al. 1992; Nichol et al. 1992; Dalton et al. 1994; Croft et al. 1997), led to a smaller value for the cluster correlation length ($r_0\approx 13-18\ h^{-1}$ Mpc). A way to overcome the projection problem is the use of data obtained in the X-ray region of the spectrum. In fact, in this band, galaxy clusters have a strong emission produced by thermal bremsstrahlung, which allows to detect them also at high redshifts. Starting from the eighties, different space missions produced extended cluster catalogues which have been mainly used to compute their X-ray luminosity function. In particular, the $ROSAT$ satellite provided a good opportunity to build a reliable all-sky survey, which was performed in the soft (0.1 -- 2.4 keV) X-ray band. First studies of the clustering properties in small samples of X-ray selected galaxy clusters have been performed by Lahav et al. (1989), Nichol, Briel \& Henry (1994) and Romer et al. (1994). More recently the $ROSAT$ data have been correlated with the Abell-ACO cluster catalogue (Abell, Corwin \& Olowin 1989) to produce the X-ray Brightest Abell Cluster sample (XBACs; Ebeling et al. 1996), for which estimates of the two-point correlation function have been recently obtained (Abadi, Lambas \& Muriel 1998; Borgani, Plionis \& Kolokotronis 1999). The corresponding values for $r_0$ are in the range $r_0\approx 20-26\ h^{-1}$ Mpc. A smaller amplitude of the correlation function is obtained from the preliminary analyses of the REFLEX sample (Collins et al. 1999), which is also obtained by the $ROSAT$ All-Sky Survey (RASS) data. In this paper we estimate the clustering properties for the RASS1 Bright Sample (De Grandi et al. 1999), which is another X-ray cluster catalogue obtained from the RASS. In this case the clusters are spectroscopically searched in a preliminary list of candidates produced by correlating the X-ray data with regions of galaxy overdensity in the southern sky. In this way, the resulting catalogue is not affected by the selection biases present in the Abell-ACO cluster catalogue. The RASS1 Bright Sample is used to test a theoretical model for the correlation function of X-ray clusters in flux-limited samples (see also Moscardini et al. 1999). This model makes use of the technique introduced by Matarrese et al. (1997) and Moscardini et al. (1998), which allows a detailed modelling of the redshift evolution of clustering, accounting both for the non-linear dynamics of the dark matter distribution and for the redshift evolution of the bias factor. A characteristic feature of this technique is that it takes into full account light-cone effects, which are relevant in analysing the clustering of even moderate redshift objects (see also Matsubara, Suto \& Szapudi 1997; de Laix \& Starkman 1998; Yamamoto \& Suto 1999). The plan of the paper is as follows. In Section 2 we summarize the characteristics of the RASS1 Bright Sample used in the following clustering analysis. In Section 3 we discuss the method used to compute the observational two-point correlation function for the RASS1 Bright Sample and we present the results. In Section 4 we introduce our theoretical model to estimate the correlations of the X-ray clusters in the framework of different cosmological models and we compare our predictions to the observational results. Conclusions are drawn in Section 5. \section {The Sample} The RASS1 Bright Sample (De Grandi et al. 1999), contains 130 clusters of galaxies selected from the first processing of the $ROSAT$ All-Sky Survey (RASS) data (Voges 1992). This sample was constructed as part of an ESO Key Programme (Guzzo et al. 1995) aimed at surveying all southern RASS candidates, which is now known as the REFLEX cluster survey (B\"ohringer et al. 1998; Guzzo et al. 1999). The identification of RASS cluster candidates was performed by means of different optically and X-ray based methods. First, candidates were found as overdensities in the galaxy density distribution at the position of the X-ray sources using the COSMOS optical object catalogue (e.g. Heydon-Dumbleton, Collins \& MacGillivray 1989). Then, correlating all RASS sources with the ACO cluster catalogue, and, finally, selecting all RASS X-ray extended sources. X-ray fluxes were remeasured using the steepness ratio technique (De Grandi et al. 1997), specifically developed for estimating fluxes from both extended and pointlike objects. A number of selections aimed at improving the completeness of the final sample lead to the RASS1 Bright Sample. Considering the intrinsic biases and incompletenesses introduced by the X-ray flux selection and source identification processes, the overall completeness of the sample is estimated to be $\magcir 90$ per cent. The RASS1 Bright Sample is count-rate-limited in the $ROSAT$ hard band (0.5 -- 2.0 keV), so that due to the distribution of Galactic absorption its effective flux limit varies between 3.05 and $4\times 10^{-12}$ erg cm$^{-2}$ s$^{-1}$ over the selected area. This covers a region of approximately 2.5 sr within the Southern Galactic Cap, i.e. $\delta<2.5^o$ and $b<-20^o$, with the exclusion of patches with RASS exposure times lower than 150 s and of the Magellanic Clouds area. The exact sky map covered by the sample is shown in Figure 2 of De Grandi et al. (1999). The redshift distribution for our whole sample is presented in the left panel of Figure~\ref{fi:nz} while the X-ray luminosity $L_X$ as a function of the redshift for each cluster is shown in the right one. It is possible to notice that 66 per cent of the clusters have $z<0.1$ but the redshift distribution has a tail up to $z\simeq 0.3$. \begin{figure*} \centering \psfig{figure=fig1.ps,height=8.cm,width=16cm,angle=0} \caption{Left panel: the redshift distribution of the RASS1 Bright Sample. Right panel: the X-ray luminosity $L_X$ (in the $ROSAT$ hard band 0.5 -- 2.0 keV) vs. the redshift $z$ for each cluster of the sample. The solid line shows the X-ray luminosity corresponding to the limiting flux $S_{\rm lim}= 3\times 10^{-12}$ erg s$^{-1}$ cm$^{-2}$ in an Einstein-de Sitter model.} \label{fi:nz} \end{figure*} \section{The 2-point correlation function} Before computing the clustering properties of our sample, we have to derive for each cluster the comoving radial distance $r$ from the observer, given the redshift of each source. To this goal we use the standard relation (neglecting the effect of peculiar motions) \begin{equation} r(z) = {{c} \over H_0 \sqrt{|\Omega_{0\cal R}|}} {\cal S} \left(\sqrt{|\Omega_{0\cal R}|} \int_0^z dz' \left[\left( 1+z' \right)^2 \left(1+\Omega_{\rm 0m} z'\right) - z'\left(2+z'\right) \Omega_{0\Lambda}\right]^{-1/2} \right) \;, \label{eq:x_z} \end{equation} where $\Omega_{0\cal R} \equiv 1 - \Omega_{\rm 0m} - \Omega_{0\Lambda}$, with $\Omega_{\rm 0m}$ and $\Omega_{0\Lambda}$ the density parameters for the non-relativistic matter and cosmological constant components, respectively. In this formula, for an open universe model, $\Omega_{0\cal R}>0$ and ${\cal S}(x)\equiv \sinh (x)$, for a closed universe, $\Omega_{0\cal R}<0$ and ${\cal S}(x)\equiv \sin (x)$, while in the Einstein-de Sitter (EdS) case, $\Omega_{0\cal R} = 0$ and ${\cal S}(x) \equiv x$. To compute the spatial two-point correlation function $\xi(r)$ we adopt both the Landy \& Szalay (1993) estimator, \begin{equation} \xi(r)= {{N_r (N_r-1)}\over{N_c (N_c-1)}} {{DD(r)}\over{RR(r)}} - {{N_r-1}\over {N_c}} {{DR(r)}\over{RR(r)}}+1\ , \label{eq:xi_ls} \end{equation} and the Davis \& Peebles (1983) estimator, \begin{equation} \xi(r)= 2 {{N_r}\over {N_c-1}} {{DD(r)}\over{RR(r)}}-1 \ . \label{eq:xi_dp} \end{equation} In the previous formulas $N_r$ is the number of random points and $N_c$ that of clusters, DD is the number of distinct cluster-cluster pairs, DR is the number of cluster-random pairs and RR refers to random-random pairs with separation between $r$ and $r + \Delta r$. The random catalogue contains a number of sources 1,000 times larger than the real catalogue (i.e. $N_r=1000 N_c$). To generate this sample we have extracted randomly coordinates from the surveyed area (see Figure 2 in De Grandi et al. 1999), assigning to each position a random flux drawn from the observed number counts (Figure 8 in De Grandi et al. 1999). We decided to retain the source in the catalogue if its flux is larger than the flux limit at the choosen position. We adopt two different methods to assign the random redshifts: in the first we scramble the observed redshifts of the clusters in the sample; in the second we generate them randomly from the observed redshift distribution binned in intervals of 0.01 in $z$. The results obtained with these two different methods are practically indistinguishable. The same happens also if we use the two previous estimators for the two-point correlation function (eqs.\ref{eq:xi_ls}-\ref{eq:xi_dp}). The errorbars have been estimated by using the bootstrap method with 50 resamplings. We find that the errors obtained in this way are in many cases larger than $\sqrt{3}$ times the Poissonian estimates, which are often used as an analytical approximation of the bootstrap errors (Mo, Jing \& B\"orner 1992). This is particularly true at small separations. In the left panel of Figure~\ref{fi:xitot} we show the correlation function computed for the whole catalogue. We present results obtained by using both an EdS model and two models with $\Omega_{\rm 0m}=0.3$ (with and without cosmological constant). The results are quite similar and the differences are always small. \begin{figure*} \centering \psfig{figure=fig2.ps,height=8.cm,width=16cm,angle=0} \caption{Left panel: the two-point correlation function of the whole RASS1 Bright Sample. Filled circles, open circles and open squares refer to results obtained by assuming $(\Omega_{\rm 0m},\Omega_{0\Lambda})=(1,0)$, $(\Omega_{\rm 0m},\Omega_{0\Lambda})=(0.3,0)$ and $(\Omega_{\rm 0m},\Omega_{0\Lambda})=(0.3,0.7)$, respectively. The corresponding best-fit relations are shown by heavy solid line, light solid line and dotted line, respectively. The 1-$\sigma$ errorbars obtained by the bootstrap method are shown only for the EdS case for clarity; their sizes in the other cases are similar. Right panel: confidence contours (68.3, 95.4 and 99.73 per cent confidence levels) of the fitting parameters $r_0$ and $\gamma$ for the two-point correlation function of the whole RASS1 Bright Sample. Heavy solid lines and filled circle refer to confidence levels and to the best fitting value obtained for the EdS case; light solid lines and open circle show the corresponding results for $(\Omega_{\rm 0m},\Omega_{0\Lambda})=(0.3,0)$ while dotted lines and open square refer to $(\Omega_{\rm 0m},\Omega_{0\Lambda})=(0.3,0.7)$ model.} \label{fi:xitot} \end{figure*} The correlation function has been fitted by adopting the power-law relation \begin{equation} \xi(r)=(r/r_0)^{-\gamma}\ . \label{eq:powlaw} \end{equation} The best-fit parameters have been obtained by using a maximum likelihood estimator based on Poisson statistics and unbinned data (Croft et al. 1997; see also Borgani, Plionis \& Kolokotronis 1999). Unlike the usual $\chi^2$-minimization, this method allows to avoid the uncertainties due to the binsize, to the position of the bin centres and to the bin scale (linear or logarithmic). To build the estimator, it is necessary to estimate the predicted probability distribution of cluster pairs, given a choice for the correlation length $r_0$ and the slope $\gamma$. By using all the distances between the cluster-random pairs, we can compute the number of pairs $g(r)dr$ in arbitrarily small bins $dr$ and use it to predict the mean number of cluster-cluster pairs $h(r)dr$ in that interval as \begin{equation} h(r)dr= {{N_c-1}\over{2N_r}} [1+\xi(r)] g(r)dr\ , \end{equation} where the correlation function $\xi$ is modelled with a power-law as in eq.(\ref{eq:powlaw}). Actually the previous equation holds only for the Davis \& Peebles (1983) estimator [eq.(\ref{eq:xi_dp})] but, since we obtain very similar results using different estimators, we can safely apply it here. Now it is possible to use all the distances between the $N_p$ cluster-cluster pairs data to build a likelihood. In particular, the likelihood function ${\cal L}$ is defined as the product of the probabilities of having exactly one pair at each of the intervals $dr$ occupied by the cluster-cluster pairs data and the probability of having no pairs in all the other intervals. Assuming a Poisson distribution, one finds \begin{equation} {\cal L}= \prod_i^{N_p} \exp[-h(r)dr] h(r)dr \prod_{j\ne i} \exp[-h(r)dr]\ , \end{equation} where $j$ runs over all the intervals $dr$ where there are no pairs. It is convenient to define the usual quantity $S=-2 \ln {\cal L}$ which can be written, once we retain only the terms depending on the model parameters $r_0$ and $\gamma$, as \begin{equation} S=2\int^{r_{\rm max}}_{r_{\rm min}} h(r)dr -2\sum_i^{N_p} \ln[h(r_i)]\ . \end{equation} The integral in the previous equation is computed over the range of scales where the fit is made. We will adopt 5 and 80 $h^{-1}$ Mpc for $r_{\rm min}$ and $r_{\rm max}$, respectively. By minimizing $S$ one can obtain the best-fitting parameters $r_0$ and $\gamma$; the confidence levels are defined by computing the increase $\Delta_S$ with respect the minimum value of $S$ and assuming a $\chi^2$ distribution for $\Delta_S$. By applying this maximum likelihood method to the RASS1 Bright Sample with the assumption of an EdS model, we find $r_0= 21.5^{+3.4}_{-4.4}\ h^{-1}$ Mpc and $\gamma=2.11^{+0.53}_{-0.56}$ (95.4 per cent confidence level with one fitting parameter). Since the redshift distribution is shallow, the values obtained in other cosmologies are quite similar: for $(\Omega_{\rm 0m},\Omega_{0\Lambda})=(0.3,0)$ we find $r_0= 21.4^{+3.4}_{-4.6} \ h^{-1}$ Mpc and $\gamma=2.17^{+0.55}_{-0.56}$, while for $(\Omega_{\rm 0m},\Omega_{0\Lambda})=(0.3,0.7)$ we find $r_0= 22.1^{+3.6}_{-4.7} \ h^{-1}$ Mpc and $\gamma=2.06^{+0.54}_{-0.56}$. The best-fit relations are also shown in Figure~\ref{fi:xitot}. Notice that a $\chi^2$-minimization procedure gives similar results, but with larger errorbars. In the right panel of the same figure we show the contour levels corresponding to $\Delta_S$ equal to 2.30, 6.31 and 11.8. Assuming that $\Delta_S$ is distributed as a $\chi^2$ distribution with two degrees of freedom, they correspond to 68.3, 95.4 and 99.73 per cent confidence levels, respectively. Notice that by assuming a Poisson distribution the method considers all pairs as independent, neglecting their clustering. Consequently the resulting errobars can be underestimated (see the discussion in Croft et al. 1997). Our results are somewhat larger than those derived by Romer et al. (1994) who found $r_0 = 13-15\ h^{-1}$ Mpc by analysing a sample of galaxy clusters also selected from the RASS in a similar region of sky ($22^h <$ RA $< 3^h$, $-50^o < \delta < 2^o$, $|b|>-40^o$). A partial explanation of this difference is related to the deeper limiting flux ($S_{\rm lim}\simeq 10^{-12}$ erg s$^{-1}$ cm$^{-2}$ in the 0.1 -- 2.4 keV band) of their catalogue: as we will discuss in the next section, the correlation length is expected to depend on the characteristics defining the surveys, such as their limits in flux and/or luminosity. However we have to remind that this early sample was derived drawing on X-ray information from the $ROSAT$ standard analysis software (SASS), which was not optimazed for the analysis of extended sources (for a more detailed discussion see e.g. De Grandi et al. 1997), and this source of incompleteness was not included in the analysis of Romer et al. (1994). Moreover in their analysis the sample sky coverage (i.e., the surveyed area as a function of the flux limit) was not discussed. Previous analyses of the XBACs sample, which is a flux-limited catalogue of X-ray Abell clusters with a limiting flux $S_{\rm lim}= 5\times 10^{-12}$ erg s$^{-1}$ cm$^{-2}$ in the 0.1 -- 2.4 keV band, gave compatible amplitudes for the correlation function: $r_0=21.1^{+1.6}_{-2.3}\ h^{-1}$ Mpc (Abadi, Lambas \& Muriel 1998) and $r_0=26.0^{+4.1}_{-4.7}\ h^{-1}$ Mpc (Borgani, Plionis \& Kolokotronis 1999; errorbars in this case are 2-$\sigma$ uncertainties). Preliminary analyses of the clustering properties of the REFLEX sample (Collins et al. 1999; Guzzo et al. 1999), which has a limiting flux $S_{\rm lim}= 3\times 10^{-12}$ erg s$^{-1}$ cm$^{-2}$ in the 0.1 -- 2.4 keV band, lead to a smaller correlation length ($r_0\simeq 18 \ h^{-1}$ Mpc). Also in this case, the discrepancy is probably a consequence of the deeper limiting flux. \section{Comparison with theoretical models} \subsection{Structure formation models} In the following analysis we consider five models, all normalized to reproduce the local cluster abundance. In particular we will adopt the normalizations obtained by Eke, Cole \& Frenk (1996) by analysing the temperature distribution of X-ray clusters (Henry \& Arnaud 1991). All our models belong to the general class of Cold Dark Matter (CDM) scemarios; their linear power-spectrum can be represented as $P_{\rm lin}(k,0) \propto k^n T^2(k)$, where, for the CDM transfer function $T(k)$, we use the Bardeen et al. (1986) fit. In particular, we consider three different EdS models, for which the power-spectrum amplitude corresponds to $\sigma_8=0.52$ (here $\sigma_8$ is the r.m.s. fluctuation amplitude in a sphere of $8 h^{-1}$ Mpc). They are: a version of the standard CDM (SCDM) model with shape parameter (see its definition in Sugiyama 1995) $\Gamma=0.45$ and spectral index $n=1$; the so-called $\tau$CDM model, with $\Gamma=0.21$ and $n=1$; a tilted model (TCDM), with $n=0.8$ and $\Gamma=0.41$, corresponding to a high (10 per cent) baryonic content. We also consider an open CDM model (OCDM), with matter density parameter $\Omega_{\rm 0m}=0.3$ and $\sigma_8=0.87$ and a low-density flat CDM model ($\Lambda$CDM), with $\Omega_{\rm 0m}=0.3$, with $\sigma_8=0.93$. Except for SCDM, which is shown as a reference model, all these models are also consistent with COBE data; for TCDM consistency is achieved by taking into account the possible contribution of gravitational waves to large-angle CMB anisotropies. A summary of the parameters of the cosmological models used in this paper is given in Table \ref{t:models}. \begin{table} \centering \caption[]{The parameters of the cosmological models. Column 2: the present matter density parameter $\Omega_{\rm 0m}$; Column 3: the present cosmological constant contribution to the density $\Omega_{0\Lambda}$; Column 4: the primordial spectral index $n$; Column 5: the Hubble parameter $h$; Column 6: the shape parameter $\Gamma$; Column 7: the spectrum normalization $\sigma_8$; Column 8: the value of the parameter $\eta$ in the luminosity-temperature relation required to reproduced the observed $\log N$--$\log S$ (see text for details).} \tabcolsep 4pt \begin{tabular}{lccccccc} \\ \\ \hline \hline Model & $\Omega_{\rm 0m}$ & $\Omega_{0\Lambda}$ & $n$ & $h$ & $\Gamma$ & $\sigma_8$ & $\eta$ \\ \hline SCDM & 1.0 & 0.0 & 1.0 & 0.50 & 0.45 & 0.52 & -0.8 \\ $\tau$CDM & 1.0 & 0.0 & 1.0 & 0.50 & 0.21 & 0.52 & 0.0 \\ TCDM & 1.0 & 0.0 & 0.8 & 0.50 & 0.41 & 0.52 & -0.3 \\ OCDM & 0.3 & 0.0 & 1.0 & 0.65 & 0.21 & 0.87 & -0.3 \\ $\Lambda$CDM & 0.3 & 0.7 & 1.0 & 0.65 & 0.21 & 0.93 & -0.2 \\ \hline \end{tabular} \label{t:models} \end{table} \subsection{The method} Theoretical predictions for the spatial two-point correlation function in the RASS1 Bright Sample have been here obtained in the framework of the above cosmological models by using a method presented in more detail in (Moscardini et al. 1999). Here we only give a short description. Matarrese et al. (1997; see also Moscardini et al. 1998) developed an algorithm to describe clustering in our past light-cone, where the non-linear dynamics of the dark matter distribution and the redshift evolution of the bias factor are taken into account. In the present paper we adopt a more refined formula which better accounts for the light-cone effects (see Moscardini et al. 1999). The observed spatial correlation function $\xi_{\rm obs}$ in a given redshift interval ${\cal Z}$ is given by the exact expression \begin{equation} \xi_{\rm obs}(r) = { \int_{\cal Z} d z_1 d z_2 {\cal N}(z_1) r(z_1)^{-1} {\cal N}(z_2) r(z_2)^{-1} ~\xi_{\rm obj}(r;z_1,z_2) \over \bigl[ \int_{\cal Z} d z_1 {\cal N}(z_1) r(z_1)^{-1} \bigr]^2 } \;, \label{eq:xifund} \end{equation} where $\xi_{\rm obj}(r,z_1,z_2)$ is the correlation function of pairs of objects at redshifts $z_1$ and $z_2$ with comoving separation $r$ and ${\cal N}(z)$ is the actual redshift distribution of the catalogue. A related approach to the study of correlations on the light-cone hypersurface has been recently presented by Yamamoto \& Suto (1999) and Nishioka \& Yamamoto (1999) within linear theory and by Matsubara, Suto \& Szapudi (1997) in the non-linear regime. An accurate approximation for $\xi_{\rm obj}$ over the scales considered here is \begin{equation} \xi_{\rm obj}(r,z_1,z_2) \approx b_{\rm eff}(z_1) b_{\rm eff}(z_2) \xi_{\rm m}(r,z_{\rm ave}) \;, \end{equation} where $\xi_m$ is the dark matter covariance function and $z_{\rm ave}$ is an intermediate redshift between $z_1$ and $z_2$, for which an excellent approximation is obtained through $D_+(z_{\rm ave}) = D_+(z_1)^{1/2}D_+(z_2)^{1/2}$ (Porciani 1997), with $D_+(z)$ the linear growth factor of density fluctuations. In our treatment we disregard the effect of redshift-space distortions. Some analytical expressions have been obtained in the mildly non-linear regime, by using either the Zel'dovich approximation (Fisher \& Nusser 1996) or higher order perturbation theory (Heavens, Matarrese \& Verde 1998). The complicating role of the cosmological redshift-space distortions on the evolution of the bias factor has been considered by Suto et al. (1999). A rough estimate of the effect of redshift-space distortions can be obtained within linear theory and the distant-observer approximation (Kaiser 1987; see Zaroubi \& Hoffman 1996 for an extension of this formalism to all-sky surveys). In this case the enhancement of the redshift-space averaged power spectrum is given by the factor $1+2\beta/3+\beta^2/5$, where $\beta\simeq \Omega_{\rm 0m}^{0.6}/b_{\rm eff}$ and $b_{\rm eff}$ is the effective bias (see below). Plionis \& Kolokotronis (1998), by analysing the XBACs catalogue and using linear perturbation theory to relate the X-ray cluster dipole to the Local Group peculiar velocity, found $\beta\simeq 0.24\pm 0.05$. Adopting this approach, Borgani, Plionis \& Kolokotronis (1999) conclude that the overall effect of redshift-space distortions is a small change of the correlation function, which expressed in terms of $r_0$ corresponds to an $\simeq 8$ per cent increase. The effective bias $b_{\rm eff}$ appearing in the previous equation can be expressed as a weighted average of the `monochromatic' bias factor $b(M,z)$ of objects of some given intrinsic property $M$ (like mass, luminosity, ...), as follows \begin{equation} b_{\rm eff}(z) \equiv {\cal N}(z)^{-1} \int_{\cal M} d\ln M' ~b(M',z) ~{\cal N}(z,M')\, , \label{eq:b_eff} \end{equation} where ${\cal N}(z,M)$ is the number of objects actually present in the catalogue with redshift in the range $z,~z+dz$ and $M$ in the range $M,~M+dM$, whose integral over $\ln M$ is ${\cal N}(z)$. In our analysis of cluster correlations we will use for ${\cal N}(z)$ in eq.(\ref{eq:xifund}) the observed one, while in the theoretical calculation of the effective bias we will take the ${\cal N}(z,M)$ predicted by the model described below. This phenomenological approach is self-consistent, in that our theoretical model for ${\cal N}(z,M)$ will be required to reproduce the observed cluster abundance and their $\log N$--$\log S$ relation. For the cluster population it is extremely reasonable to assume that structures on a given mass scale are formed by the hierarchical merging of smaller mass units; for this reason we can consider clusters as being fully characterized at each redshift by the mass $M$ of their hosting dark matter haloes. In this way their comoving mass function $\bar n(z,M)$ can be computed using an approach derived from the Press-Schechter technique. Moreover, it is possible to adopt for the monochromatic bias $b(M,z)$ the expression which holds for virialized dark matter haloes (e.g. Mo \& White 1996; Catelan et al. 1998). Recently, a number of authors have shown that the Press-Schechter (1974) relation does not provide an accurate description of the halo abundance both in the large and small-mass tails (e.g. Sheth \& Tormen 1999). Also, the simple Mo \& White (1996) bias formula has been shown not to correctly reproduce the correlation of low mass haloes in numerical simulations. Several alternative fits have been recently proposed (Jing 1998; Porciani, Catelan \& Lacey 1999; Sheth \& Tormen 1999; Jing 1999). In this paper we adopt the relations recently introduced by Sheth \& Tormen (1999), which have been shown to produce an accurate fit of the distribution of the halo populations in the GIF simulations (Kauffmann et al. 1999). They read \begin{equation} \bar n(z,M) = \sqrt{2 a A^2 \over \pi} \ {{3 H_0^2 \Omega_{\rm 0m}} \over {8\pi G}} \ { \delta_c \over M D_+(z) \sigma_M } \ \biggl[ 1 + \biggl( {{D_+(z) \sigma_M} \over {\sqrt{a} \delta_c}} \biggr)^{2p} \biggr] \ \bigg| {d \ln \sigma_M \over d \ln M} \bigg| \exp \biggl[ -{a \delta_c^2 \over 2 D_+^2(z) \sigma^2_M} \biggr] \; \label{eq:ps2} \end{equation} and \begin{equation} b(M,z) = 1 + {1 \over \delta_c } \biggl( {a \delta_c^2 \over \sigma_M^2 D_+^2(z)} - 1\biggr) + {2 p \over \delta_c } \biggl( {1 \over { 1+[\sqrt{a} \delta_c / (\sigma_M D_+(z))]^{2p}}} \biggr) \; . \label{eq:b_mono2} \end{equation} Here $\sigma^2_M$ is the mass-variance on scale $M$, linearly extrapolated to the present time ($z=0$), and $\delta_c$ the critical linear overdensity for spherical collapse. Following Sheth \& Tormen (1999), we adopt their best-fit parameters $a=0.707$, $p=0.3$ and $A\approx 0.3222$, while the standard (Press \& Schechter and Mo \& White) relations are recovered for $a=1$, $p=0$ and $A=1/2$. Notice that \begin{equation} {\cal N}(z,M) = 4\pi r^2(z) {dr \over dz} \left[1 + {H_0^2 \over c^2} ~\Omega_{0\cal R} ~r^2(z)\right]^{-1/2} {\bar n}(z,M) ~\phi(z,M) \;, \end{equation} where $\phi(z,M)$ is the isotropic catalogue selection function, which also accounts for the catalogue sky coverage, as detailed below. The last ingredient entering in our computation of the correlation function is the redshift evolution of the dark matter covariance function $\xi_{\rm m}$. As in Matarrese et al. (1997) and Moscardini et al. (1998) we use an accurate method, based on the Hamilton et al. (1991) ansatz, to evolve $\xi_{\rm m}$ into the fully non-linear regime. In particular, we use the fitting formula given by Peacock \& Dodds (1996). In order to predict the abundance and clustering of X-ray clusters in the RASS1 Bright Sample we need to relate X-ray cluster fluxes into a corresponding halo mass at each redshift. The given band flux $S$ corresponds to an X-ray luminosity $L_X=4\pi d_L^2 S$ in the same band, where $d_L=(1+z) r(z)$ is the luminosity distance. To convert $L_X$ into the total luminosity $L_{\rm bol}$ we perform band and bolometric corrections by means of a Raymond-Smith code, where an overall ICM metallicity of $0.3$ times solar is assumed. We translate the cluster bolometric luminosity into a temperature, adopting the empirical relation $T = {\cal A} \ L_{\rm bol}^{\cal B} \ (1+z)^{-\eta}$, where the temperature is expressed in keV and $L_{\rm bol}$ is in units of $10^{44} h^{-2}$ erg s$^{-1}$. In the following analysis we assume ${\cal A}=4.2$ and ${\cal B}=1/3$; these values allow a good representation of the local data for temperatures larger than $\approx 1$ keV (e.g. David et al. 1993; White, Jones \& Forman 1997; Markevitch 1998). Analysing a catalogue of local compact groups, Ponman et al. (1996) showed that at lower temperatures the $L_{\rm bol}-T$ relation has a steeper slope (${\cal B}\approx 0.1$). For these reasons we prefer to fix a minimum value for the temperature at $T=1$ keV. Moreover, even if observational data are consistent with no evolution in the $L_{\rm bol}-T$ relation out to $z \approx 0.4$ (Mushotzky \& Scharf 1997), a redshift evolution described by the parameter $\eta$ has been introduced to reproduce the observed $\log N$--$\log S$ relation (Rosati et al. 1998; De Grandi et al. 1999) in the range $2 \times 10^{-14} \le S \le 2 \times 10^{-11}$ (see also Kitayama \& Suto 1997; Borgani et al. 1999). The values of $\eta$ required for SCDM, $\tau$CDM, TCDM, OCDM and $\Lambda$CDM models are reported in Table \ref{t:models}. A general discussion of the effects of different choices of the parameters entering in this method (e.g. the scatter in the $L_{\rm bol}-T$ relation) is presented elsewhere (Moscardini et al. 1999). Finally, with the standard assumption of virial isothermal gas distribution and spherical collapse, it is possible to convert the cluster temperature into the mass of the hosting dark matter halo, namely (e.g. Eke, Cole \& Frenk 1996) \begin{equation} T = {7.75 \over \beta_{\rm TM}} {\left(M\over {10^{15} h^{-1} M_\odot}\right)}^{2/3} (1+z) {\left( \Omega_{\rm 0m} \over \Omega_{\rm m}(z)\right) }^{1/3} \left({\Delta_{\rm vir}(z) \over {178}}\right)^{1/3} \ . \end{equation} The quantity $\Delta_{\rm vir}$ represents the mean density of the virialized halo in units of the critical density at that redshift (e.g. Bryan \& Norman 1998 for fitting formulas). We assume $\beta_{\rm TM}=1.17$, which is in agreement with the results of different hydrodynamical simulations (Bryan \& Norman 1998; Gheller, Pantano \& Moscardini 1998; Frenk et al. 1999). Once the relation between observed flux and halo mass at each redshift is established we account for the RASS1 Bright Sample sky coverage $\Omega_{\rm sky}(S)$ (see Figure 7 in De Grandi et al. 1999) by simply setting $ 4 \pi \phi(z,M) = \Omega_{\rm sky}[S(z,M)]$. \subsection{Results} In Figure~\ref{fi:theor} we compare our predictions for the RASS1 spatial correlation function in different cosmological models to the observational data. All the EdS models here considered predict too small an amplitude. Their correlation lengths are smaller than the observational results: we find $r_0\simeq 11.5, 12.8, 14.8 \ h^{-1}$ Mpc for SCDM, TCDM and $\tau CDM$, respectively. On the contrary, both the OCDM and $\Lambda$CDM models are in much better agreement with the data and their predictions are always inside the 1-$\sigma$ errorbars ($r_0\simeq 18.4, 18.6 \ h^{-1}$ Mpc, respectively). In order to quantify the differences between the model predictions and the observational data, we use again the maximum likelihood approach. The minimum value for $S$ is obtained for the $\Lambda$CDM model. A similar value, corresponding to $\Delta_S=0.7$, is obtained for the OCDM model, while for the EdS models the resulting $S$ are much larger: $\Delta_S=10.6, 19.3, 26.8$ for $\tau CDM$, TCDM and SCDM models, respectively. To evaluate what is the effect of neglecting the description of clustering in the past light-cone, as usually done in previous analyses, we estimate the cluster correlation function as $\xi(r)=b^2_{\rm eff}(z_{\rm med})\xi_{\rm m}(r,z_{\rm med})$, where $z_{\rm med}$ is the median redshift of the catalogue. For the RASS1 Bright Sample we have $z_{\rm med}\simeq 0.08$. The resulting values of the correlation lengths obtained in this way are $r_0\simeq 13.5, 15.5, 18.2, 21.9, 22.4 \ h^{-1}$ Mpc, for SCDM, TCDM, $\tau CDM$, OCDM and $\Lambda CDM$, respectively. They are typically 20 per cent higher than the estimates obtained by our method. This difference is due to the fact that in the past light-cone formalism the matter correlation functions (and bias factors) are weighted by a factor ${\cal N}(z)/r(z)$, for which the average value on the whole sample corresponds neither to the median nor to the mode of the redshift distribution. Actually, the presence of the comoving radial distance $r(z)$ at the denominator tends to shift the ``effective redshift'' to smaller values of $z$. As a consequence, the true cluster correlations, which are indeed measured in our past light-cone, have typically smaller amplitudes than those estimated at the median redshift of the catalogue. Of course, the importance of this effect becomes larger when deeper surveys are considered. \begin{figure} \centering \psfig{figure=fig3.ps,height=10.cm,width=16cm,angle=0} \caption{Comparison of the observed spatial correlation for clusters in the RASS1 Bright Sample (shown as in Figure~\ref{fi:xitot}) with the predictions of the various theoretical models: SCDM (solid line), $\tau$CDM (dotted line), TCDM (short-dashed line), OCDM (long-dashed line) and $\Lambda$CDM (dotted-dashed line). } \label{fi:theor} \end{figure} In order to study the possible dependence of the clustering properties of the X-ray clusters on the observational characteristics defining the survey, we use our model to predict the values of the correlation length $r_0$ in catalogues where we vary the limiting X-ray flux $S_{\rm lim}$ or luminosity $L_{\rm lim}$ (both defined in the energy band 0.5 -- 2 keV). Notice that this analysis can be related to the study of the richness dependence of the cluster correlation function. In fact, a change in the observational limits implies a change in the expected mean intercluster separation $d_c$. Bahcall \& West (1992) found that the Abell clusters data are consistent with a linear relation $r_0=0.4 d_c$, while a milder dependence is resulting from the analysis of the APM clusters (Croft et al. 1997). Our results are shown in Figure~\ref{fi:slim}. All the cosmological models display a similar trend: in the flux and luminosity intervals here considered, the correlation length $r_0$ has a slow growth with $S_{\rm lim}$ (left panel), and a more marked one with $L_{\rm lim}$ (right panel). For example, for OCDM and $\Lambda CDM$ the correlation length changes from $r_0 \simeq 18$ to $r_0 \simeq 21 h^{-1}$ Mpc, when the limiting flux varies from $S_{\rm lim}= 10^{-12}$ to $S_{\rm lim}=10^{-11}$ erg s$^{-1}$ cm$^{-2}$, and from $r_0\simeq 18$ to $r_0\simeq 30 h^{-1}$ Mpc, when the limiting luminosity varies from $L_{\rm lim}= 10^{42}$ to $L_{\rm lim}= 10^{44} h^{-2}$ erg s$^{-1}$. The values of $r_0$ for the EdS models have similar variations but are always smaller. We can compare these predictions to the results obtained by computing the two-point correlation function in the RASS1 Bright Sample with the same cuts in X-ray flux or luminosity. The estimates of $r_0$ obtained in this way are presented in Table \ref{t:r0_lim} (where we also report the number of clusters inside each subsample) and shown in the figure (open squares). In both panels the whole catalogue is represented by the square on the right. Because of the small number of clusters in these subsamples, we prefer to fit the correlation function by fixing the value of the slope $\gamma=1.8$; we find that our values of $r_0$ are only slightly dependent on this assumption. The errorbars shown in the figure correspond to an increase $\Delta_S = 4$ with respect to the minimum value of $S$. With the assumption that $\Delta_S$ is distributed as a $\chi^2$ distribution with a single degree of freedom, this corresponds to 95.4 per cent confidence level. By analysing the trend with changing limiting flux, we find that the observed values of $r_0$ are almost constant even if $S_{\rm lim}$ changes by a factor larger than 2. This result is in agreement with what Borgani, Plionis \& Kolokotronis (1999) obtained for XBACs. We notice that OCDM and $\Lambda$CDM are able to reproduce the amount of clustering shown by the RASS1 Bright Sample, while all the EdS models strongly underpredict the amplitude of the correlation function. The situation is slightly different when we analyse luminosity-limited catalogues. The RASS1 Bright Sample suggests a small increase of $r_0$ with $L_{\rm lim}$, even if the hypothesis of a constant correlation length cannot be rejected. This is consistent with a similar analysis made by Abadi, Lambas \& Muriel (1998) on the XBACs catalogue. As shown in the left panel of Figure~\ref{fi:slim}, our models always tend to predict smaller correlations, even if the non-EdS models are still marginally consistent with the observational data. \begin{figure} \centering \psfig{figure=fig4.ps,height=8.cm,width=16cm,angle=0} \caption{The behaviour of the correlation length $r_0$ as a function of the limiting X-ray flux $S_{\rm lim}$ (left panel) and luminosity $L_{\rm lim}$ (right panel). The open squares and errorbars (95.4 per cent confidence level) refer to different subsamples of RASS1 Bright Sample here analysed. The prediction of the various theoretical models are shown as in Figure~\ref{fi:theor}: SCDM (solid line), $\tau$CDM (dotted line), TCDM (short-dashed line), OCDM (long-dashed line) and $\Lambda$CDM (dotted-dashed line). } \label{fi:slim} \end{figure} \begin{table} \centering \caption[]{The correlation length $r_0$ in flux-limited (left) and luminosity-limited (right) subsamples of the RASS1 Bright Sample. Column 1: the limiting X-ray flux $S_{\rm lim}$ (in units of $10^{-11}$ erg s$^{-1}$ cm$^{-2}$) or luminosity $L_{\rm lim}$ (in units of $10^{44} h^{-2}$ erg s$^{-1}$); Column 2: the number of clusters in the subsample; Column 3: the correlation length $r_0$ (in units of $h^{-1}$ Mpc) as obtained from the maximum likelihood analysis with fixed slope $\gamma=1.8$ (errorbars correspond to 95.4 per cent confident level) } \tabcolsep 4pt \begin{tabular}{ccccccc} \\ \\ \hline \hline $S_{\rm lim}$ & no. of clusters & $r_0$ &\hspace{2.truecm} & $L_{\rm lim}$ & no. of clusters & $r_0$ \\ \hline 0.3 & 126 & $21.2^{+4.0}_{-4.1}$ && 0.01& 126 & $21.2^{+4.0}_{-4.1}$\\ 0.5 & 78 & $23.7^{+5.2}_{-5.4}$ && 0.03& 122 & $22.6^{+4.2}_{-4.2}$\\ 0.7 & 41 & $18.1^{+9.3}_{-11.1}$ &&0.05& 118 & $23.1^{+4.3}_{-4.3}$\\ & & & &0.10 & 115 & $24.3^{+4.4}_{-4.5}$ \\ & & & &0.15 & 106 & $24.9^{+5.1}_{-5.4}$ \\ & & & &0.20 & 98 & $27.2^{+6.0}_{-6.1}$ \\ & & & &0.25 & 89 & $30.2^{+7.1}_{-7.2}$ \\ & & & &0.30 & 84 & $32.5^{+7.6}_{-7.7}$ \\ & & & &0.40 & 74 & $32.8^{+9.5}_{-9.9}$ \\\hline \end{tabular} \label{t:r0_lim} \end{table} \section{Conclusions} In this paper we have studied the two-point correlation function $\xi(r)$ of a flux-limited sample of X-ray galaxy clusters, the RASS1 Bright Sample. These observational results have been used to test a theoretical model predicting the clustering properties of X-ray clusters in flux-limited surveys in the framework of different cosmological scenarios. Our main results are: \begin{itemize} \item Assuming an Einstein-de Sitter model, $\xi(r)$ can be well fitted using the standard power-law relation $\xi=(r/r_0)^{-\gamma}$, with $r_0= 21.5^{+3.4}_{-4.4} h^{-1}$ Mpc and $\gamma=2.11^{+0.53}_{-0.56}$ (95.4 per cent confidence levels with one fitting parameter). The values obtained in models with matter density parameter $\Omega_{\rm 0m}=0.3$ are quite similar. \item The amplitude of the correlation function is almost constant when the RASS1 Bright Sample is analysed with different limiting fluxes in the range $S_{\rm lim}=3-7 \times 10^{-12}$ erg s$^{-1}$ cm$^{-2}$ while it displays a slightly increasing trend when computed in catalogues with an increasing X-ray luminosity limit in the range $L_{\rm lim}=0.01-0.4 \times 10^{44}h^{-2}$ erg s$^{-1}$. \item The comparison with the predictions of our theoretical models shows that the Einstein-de Sitter models are unable to reproduce the observational results obtained for the whole RASS1 Bright Sample. On the contrary, good agreement is found for the models with matter density parameter $\Omega_{\rm 0m}=0.3$, both with and without a cosmological constant. \item Our models are also able to reproduce the behaviour of $r_0$ with $S_{\rm lim}$ and $L_{\rm lim}$, but the observed amount of clustering is reproduced only by the open and $\Lambda$ models. \end{itemize} In conclusion, we believe that the method presented here leads to robust predictions on the clustering of X-ray clusters; its future application to new and deeper catalogues will allow to provide useful constraints on the cosmological parameters. \section*{Acknowledgments.} This work has been partially supported by Italian MURST, CNR and ASI. We are grateful to Stefano Borgani, Enzo Branchini, Houjun Mo, Ornella Pantano, Piero Rosati, Bepi Tormen and Elena Zucca for useful discussions. We also thank the referee, C. Collins, for comments which allowed us to improve the presentation of this paper.
\section{Introduction} The theory of joins of (geometric) simplicial complexes as given by Brown, \cite{Top:Brown}, or Spanier, \cite{:Span}, reveals the join operation to be a basic geometric construction. It is used in the development of several areas of geometric topology (cf. Hudson, \cite{Hudson}) whilst also being applied to the basic properties of polyhedra relating to homology. The theories of geometric and abstract simplicial complexes run in a largely parallel way and when describing the theory, expositions often choose which aspect -- abstract combinatorial or geometric -- to emphasise at each step. Historically in algebraic topology geometric simplicial complexes, as tools, were largely replaced by CW complexes whilst the combinatorial abstract complex became part of simplicial set theory. In the process, joins were negelected and there does not seem to be a well known definition of the join of two simplicial sets. Within the setting of simplicial set theory, the ordinal sum plays a strange r\^ole. This operation takes two ordinals and concatenates them, so $[m]or[n] = [m+n+1]$, where \linebreak$[m] = \{ 0 < 1 < \cdots < m \}$, so it is fundamental for the combinatorics of ordinals. In the literature on simplicial set theory it seems rarely to be mentioned, yet it is sometimes there but hidden, for instance, in the $\overline{W}$-construction for simplicial groups (see May, \cite{Algtop:May}) or simplicial groupoids (see Dwyer and Kan, \cite{DandK}). In this context it occurs through the use of the Artin-Mazur codiagonal, \cite{AandM}, which assigns to a bisimplicial set or group, a much smaller model of the homotopy type than does the diagonal. (The diagonal is intuitively easier to use and tends to be ``wheeled out'' whenever passage from bisimplicial objects to simplicial objects is needed; however, it may not always be the most efficient tool to use.) This codiagonal is linked with the total DEC functor (Illusie \cite{Cot:Ill}, Duskin \cite{:Dus}, Porter, \cite{:Porter}, Bullejos et al \cite{Bull}), which can be given explicitly in terms of the ordinal sum. In this brief note, it is shown that the ordinal sum leads naturally to a ``join'' operation on {\em augmented} simplicial sets, and the relation of this join to the geometric join is studied. \section{Definitions}\label{defn} It will be assumed that the reader is conversant in general with basic simplicial set theory, in particular, the definition of the singular complex of a topological space, and the geometric realisation of a singular complex. On the subject of notation, note that the simplicial set which is called the $n$-simplex, $\triangle [n]$\glossary{$\triangle [n]$}, is the representable functor, $\Delta (-,[n])$. The simplicial set $\triangle [n]$ will be referred to as the {\it standard} $n$-{\it simplex}. The category of finite ordinals will be denoted $\Delta $: the ordinal $\{0 < 1 < \cdots < n\}$ will be denoted $[n]$ with the empty set being denoted by $[-1]$.\\ \noindent {\bf Definition \ref{defn} (i)}\\ Let $f_i:[p_i] \rightarrow [q_i]$ for $i = 0, 1$. Define the ``ordinal sum'' functor, \\ $or:\Delta ^{2} \longrightarrow \Delta$, as follows:-\[ or([p_0],[p_1]) = [p_0 + p_1 + 1] \]\[ or(f_0,f_1) = \left\{\begin{array}{lcl} f_0(k) & \mbox{ if } & 0 \leq k \leq p_0 \\ f_1(k-p_0-1)+q_0+1 & \mbox{ if } & p_0+1 \leq k \\\end{array} \right. \] \\ Note that $[-1]$ is a two sided identity for the operation on objects.\\ \noindent {\bf Definition \ref{defn} (ii)} \\ An augmented simplicial set is a simplicial set, $X$, together with an augmentation, that is, a set $X_{-1}$ and a morphism $q_X:X_0\longrightarrow X_{-1}$, where $q_Xd_0 = q_Xd_1$. There is an obvious forgetful functor from the category, $ASS$, of augmented simplicial sets to that $SS$, of simplicial sets, cf. Duskin \cite{:Dus}.\\ \noindent {\bf Definition \ref{defn} (iii)} \\ The {\it canonical augmentation} of a simplicial set has $X_{-1} = \pi_0(X)$ and $q_X$ the coequaliser of $$X_1 \begin{array}{l} \stackrel{d_0}{\longrightarrow} \\ \stackrel{d_1}{\longrightarrow} \end{array} X_0 .$$ This augmentation is left adjoint to the forgetful functor.\\ \noindent {\bf Definition \ref{defn} (iv)} \\ The {\it trivial augmentation} of a simplicial set has $X_{-1} = \ast $, the one point set, and $q_X$ the unique (trivial) morphism $X_0 \longrightarrow \ast$. This augmentation is right adjoint to the forgetful functor.\\ \noindent {\bf Definition \ref{defn} (v)} \\ The {\it geometric realisation} defined on augmented simplicial sets is the composition of the forgetful functor to simplicial sets and the usual geometric realisation functor to topological spaces. (This is the only reasonable definition of a geometric realisation on augmented simplicial sets, as the codomain of the augmentation is, in some sense, the image of the empty set.) \noindent {\bf Definition \ref{defn} (vi)} \\ The {\it singular complex functor} from topological spaces to augmented simplicial sets is the composition of the normal singular complex functor, which is right adjoint to the geometric realisation functor, with the trivial augmentation functor, right adjoint to the forgetful functor. It is automatic that the two functors so defined are adjoint. \section{Combinatorial Join} \label{monoid} The following is our proposed definition for a join of augmented simplicial sets.\\ \noindent {\bf Definition \ref{monoid} (i)}\\ Let the {\it join} of two augmented simplicial sets $X$ and $Y$ be denoted $X \odot Y $. The set of $n$-simplices, $(X \odot Y)_n$, is:- \[ \bigsqcup_{i=-1}^n X_{n-1-i} \times Y_i \] the face maps are given by:- \[d_i^n(x,y) = \left\{ \begin{array}{cl} (d_i^px,y) & \mbox{ \ if \ } 0 \leq i \leq p \\(x,d_{i-p-1}^{n-p-1}y) & \mbox{ \ if \ } p < i \leq n \end{array} \right. \] where $(x,y) \in X_p \times Y_{n-p-1}$, and $d_0^0$ is the augmentation (of $X$ or $Y$);\\ lastly, the degeneracies are:- \[s_i^{n-1}(x,y) = \left\{ \begin{array}{cl} (s_i^px,y) & \mbox{ \ if \ } 0 \leq i \leq p \\(x,s_{i-p-1}^{n-p-2}y) & \mbox{ \ if \ } p < i \leq n-1 \end{array} \right. \] where $(x,y) \in X_p \times Y_{n-p-2}$. There is also a coend definition for $\odot$:\[ X \odot Y \cong\int^{p,q} (X_p \times Y_q) \cdot \triangle([p]or[q]) \] \noindent {\bf Remark}\\ It is trivial to prove that $\triangle [n] \odot \triangle [m] \cong \triangle [n+m+1]$. It is also true that $\odot $ is an associative operation, but the simplest proof requires a number of constructions and results associated with the join which are not of immediate interest here. \section{Topological Join} \label{top} The following definition is a generalisation of the concept of join for two suitable subspaces of a vector space. The {\it topological join} thus defined is discussed in some detail in chapter 5, section 7 of \cite{Top:Brown}. Results proved there will be used here without proof: the notation for this section is largely taken from there. We work within the category of compactly generated spaces.\\ \noindent {\bf Definition \ref{top} (i)} \\ Consider two topological spaces ${\cal U}$ and ${\cal V}$, and construct a set of 4-tuples $(r,u,s,v)$, where $u \in {\cal U},\; v \in {\cal V}, \; r,s \in [0,1]$ and $r + s = 1$: in the case that $r = 0$, the $u$ will be ignored, and in the case that $s = 0$, the $v$ will be ignored. This set will be suggestively called ${\cal U} \ast {\cal V}$. There are obvious projections from this set of 4-tuples: \\ $p_{\cal U}:{\cal U} \ast {\cal V} \rightarrow {\cal U}$, \rule{5pt}{0pt} $p_{\cal V}:{\cal U} \ast {\cal V} \rightarrow {\cal V}$, \rule{5pt}{0pt} $p_r:{\cal U} \ast {\cal V} \rightarrow (0,1] $ and $p_s:{\cal U} \ast {\cal V} \rightarrow (0,1]$ \\ which are termed the {\it coordinate functions} of ${\cal U} \ast {\cal V}$. The first two are obviously defined, the last two take a point $(r,u,s,v) \in {\cal U} \ast {\cal V} $ to $r$ and $s$ respectively. The {\it topological join} of ${\cal U}$ and ${\cal V}$ is defined as the set ${\cal U} \ast {\cal V}$ together with the initial topology with respect to the {\it coordinate functions}. Thus a function with codomain ${\cal U} \ast {\cal V}$ is a continuous function if and only if its composite with each of the coordinate functions is continuous. \medskip To compare the combinatorial and topological join operations, we will need more precision on the construction of the geometric realisation. There are a number of different constructive definitions of geometric realisation. The process is essentially the following: (i) take one copy of $\triangle ^n$ for each non-degenerate $n$-simplex of $X$;\\ and then (ii) glue them all together using the face and degeneracy maps of the simplicial set $X$ (see \cite{Cat:Mac}). \\ Explicitly we have: Let $X$ be a simplicial set. Define $RX$ by:\\ \[ RX = \sqcup_{n \in {\mathbb N}}\sqcup_{x \in X_n} \triangle ^n_x \] Define an equivalence relation on $RX$ as generated by the following relation:\\ writing $({\bf p},x)$ for $(p_0,\cdots ,p_m) \in \triangle ^m_x$ and $({\bf q},y)$ for $ (q_0,\cdots ,q_n) \in \triangle ^n_y$ then $({\bf p},x) \sim ({\bf q},y)$ if either \\ \rule{20mm}{0mm} $d_ix = y$ and $\delta _i(q_0,\cdots ,q_n) = (p_0,\cdots , p_m)$ or \\ \rule{20mm}{0mm} $s_ix = y$ and $\sigma _i(q_0,\cdots ,q_n) = (p_0,\cdots, p_m)$,\\ where the $\delta_i$ and $\sigma_i$ are the continuous maps given by face inclusion and folding in the usual way. Then $ |X| \cong {RX}/{\sim} $ where $ {RX}/{\sim } $ has the identification topology. \begin{proposition} \label{join} \[\triangle ^p \ast \triangle ^q \cong \triangle ^{p+q+1} \] \end{proposition} {\bf Proof} \\ Consider the vector space ${\mathbb R}^{p+q+1}$ and the two compact convex subsets: \[ X = \{ \, (x_0,x_1,\cdots, x_p,0,\cdots ,0) \, | \, \sum_{i=0}^p x_i = 1 \, \} \] \[ Y = \{ \, (0,\cdots,0,y_0,y_1,\cdots ,y_q) \, | \, \sum_{j=0}^q y_j = 1 \, \} \] First note that $X \cong \triangle ^p$ and $Y \cong \triangle ^q$. Furthermore, it is clear that no two lines in the set $U = \{ \, r{\bf x} + (1-r){\bf y} \, | \, 0 \leq r \leq 1, \, {\bf x} \in X, \, {\bf y} \in Y \, \} $ intersect except at endpoints. Thus $X \ast Y = U$. However, $U$ is the subset of ${\mathbb R} ^{p+q+1} $ given by \[ \{\, (rx_0,\cdots ,rx_p, (1-r)y_0,\cdots ,(1-r)y_q) \, | \, \sum_{i=0}^p rx_i + \sum_{j=0}^q (1-r)y_j = 1 \, \}. \] That is, $U$ is the affine $(p+q+1)$-simplex. Therefore $\triangle ^p \ast \triangle ^q \cong \triangle ^{p+q+1}$. \hfill $\square $ \\ When we form $\triangle [p] \odot \triangle [q]$, we obtain, on varying $p$ and $q$, a bicosimplicial object in $SS$. (In general if $\cal C$ is a category, a cosimplicial object in $\cal C$ is a functor from $\Delta$ to $\cal C$, whilst a bicosimplicial object is a functor from $\Delta \times \Delta$ to $\cal C$.) Similarly $\triangle^p \ast \triangle^q$ is a bicosimplicial space. \begin{lemma} There is a natural isomorphism \[ |\triangle [p]| \ast |\triangle [q]| \cong |\triangle [p] \odot \triangle [q]| \] of bisimplicial spaces. \label{join2} \end{lemma} {\bf Proof}\\ Recall $|\triangle [m]| := \triangle ^m $. Since $\triangle [p] \odot \triangle [q] \cong \triangle ([p]or[q]) = \triangle [p+q+1]$, the isomorphism exists for each pair $(p,q)$. Now $\{\triangle ^n\}_{n \in {\mathbb N}}$ has an obvious cosimplicial structure, and the isomorphism is easily seen to be an isomorphism of bicosimplicial spaces. \hfill $\square $ \\ \begin{theorem} \rule{0pt}{12pt} \\ Let $X$ and $Y$ be trivially augmented simplicial sets. Then \[ |X \odot Y| \cong |X| \ast |Y| \] \label{join3} \end{theorem} {\bf Proof}\\ (The following is a direct geometric proof: we will comment later on the categorical aspects.) Recall that \[ |X| \ast |Y| \, := \, \left\{ \begin{array}{ll} r[(p_0,\cdots,p_m)_x] \rule{24pt}{0pt} s.t. & \sum_{i=0}^m p_i = 1 \mbox{ , } \sum_{i=0}^n q_i = 1 \mbox{ , } \\ \rule{12pt}{0pt} + s[(q_0, \cdots q_n)_y] \, & x \in X_m \mbox{ , } y \in Y_n \mbox{ , } r + s = 1 \\ & p_i,q_i,r,s \geq 0 \\ & \mbox{and } [-] \mbox{ denotes equivalence class} \end{array} \right\} \] It should also be noted that if $r = 0$, the point from $|X|$ is ignored and similarly, if $s = 0,$ the point from $|Y|$ is ignored.\\ Define a map $ f:|X| \ast |Y| \longrightarrow |X \odot Y| $ as follows: \[ f(r[(p_0,\cdots ,p_m)_x] + s[(q_0,\cdots ,q_n)_y]) \mapsto [(rp_0,\cdots ,rp_m,sq_0,\cdots ,sq_n)_{x,y}] \] The function $f$ is well defined since if $r = 0$, the point $x$ is ignored, similarly if $s = 0$. This means that for any $y$, it must be true that $(0,\cdots 0,q_0,\cdots ,q_n)_{(x,y)} \sim (0,\cdots 0,q_0,\cdots q_n)_{(x',y)}$ for all $x,x' \in X$. This will be true exactly when the augmentation of both $X$ and $Y$ are trivial as was required. A moment's thought then will show that the function $f$ respects the relation and so is well defined. Continuity is also trivial to check. The obvious inverse function is also continuous under the definition of the topology on $|X| \ast |Y|$. Thus the two spaces are homeomorphic. \hfill $\square $ \\ \noindent{\bf Remarks}\\ (i) It may seem slightly contrived that the condition ``trivially augmented'' should be needed, however consider the following example: Let $X := \triangle [0] \sqcup \triangle [0]$ together with the {\it canonical} augmentation, and consider $X \odot X$. The result is the disjoint union of four unit intervals -- that is, $\triangle [1] \sqcup \triangle [1] \sqcup \triangle [1] \sqcup \triangle [1] $: Ideally, the result should be homotopically equivalent to a $1$-sphere. (ii) The Theorem above is in fact a simple consequence of a categorical argument which shows a different aspect of the necessity for having a trivial augmentation. The simplicial complex functor to augmented simplicial sets needs to specify an augmentation, and for the functor to be right adjoint to the geometric realisation functor, the augmentation must be the trivial one (since the trivial augmentation is right adjoint to the forgetful functor from augmented simplicial sets to simplicial sets). Thus the condition `trivially augmented' merely requires that the augmented simplicial sets are related to the geometric realisation functor upon which the theorem depends. The result is now seen to depend just on left adjoints interacting nicely with the coends in the geometric realisation and join functors. \section{Simplicial Spheres.} Recall (from \cite{Top:Brown}) that \[ {\bf S}^p \ast {\bf S}^q \cong {\bf S}^{p+q+1} . \] This essentially says that the $n$-sphere in the category of topological spaces is the join of $n+1$ copies of the $0$-sphere. There are several simplicial models for the n-sphere. For instance, Gabriel and Zisman, \cite{GandZ}, p.26, define the simplicial circle, $\Omega$, to be the coequaliser of the pair of morphisms $$\diagram \triangle [0]\rto<1ex>^{\delta_0}\rto<-1ex>_{\delta_1} & \triangle[1], \enddiagram$$ and the suspension of a pointed simplicial set $X$ to be $\Omega \wedge X$. This gives an n-sphere as being $\bigwedge^n\Omega$, obtained from the n-cube $\triangle[1]^n$ by collapsing the `boundary' of the cube to a point. Other authors form a simplicial sphere by collapsing the boundary $\partial \triangle[n]$ of the n-simplex to a point. The join operation suggests another form. Consider the simplicial set formed as the disjoint union of two copies of $\triangle [0]$ and augmented trivially. This will be denoted by ${\bf S}^0$ and will be referred to as the simplicial $0$-sphere. Then ${\bf S}^0 \odot {\bf S}^0$ has four non-degenerate $1$-simplices connected to each other in a ``diamond'' as below:-\\ \begin{picture}(300,115) \put(150,55){\vector(1,1){47}} \put(150,55){\vector(1,-1){47}} \put(250,55){\vector(-1,1){47}} \put(250,55){\vector(-1,-1){47}} \put(150,55){\circle*{5}} \put(200,105){\circle*{5}} \put(200,5){\circle*{5}} \put(250,55){\circle*{5}} \end{picture} Define the simplicial $n$-sphere, ${ S}^n \in ob{\mathit ASS}$, as follows:- \[S^n : = \underbrace{ {S}^0 \odot \cdots \odot { S}^0}_{n+1} \] It is clear from the definition of combinatorial join and of the simplicial $0$-sphere that the simplicial $n$-sphere is a triangulation of the topological $n$-sphere. In fact, theorem~\ref{join3} gives explicitly that \[ |{ S}^n| \cong {\bf S}^n. \] Moreover this model clearly satisfies \[ S^p \odot S^q \cong S^{p+q+1}\] unlike the other models. Thus if we write $\Sigma^n = \triangle[n] / \partial\triangle[n]$ then $\Sigma^p \odot \Sigma^q$ has one non-degenerate simplex in each of the dimensions 1, p + 1, q + 1, and p + q + 1, and two non-degenerate simplices in dimension 0 and so `looks' totally unlike $\Sigma^{p+q+1}$. The combinatorial join forms part of a closed monoidal structure on the category of augmented simplicial sets, $ASS$. (The `internal hom' is given by \[ [X,Y]_n = ASS(X,Dec^{n+1}Y),\] where $Dec$ is the d\'ecalage functor (see Duskin, \cite{:Dus} ).) It is therefore possible to define augmented analogues of the loopspace construction that are compatible with the join.
\section{Prologue} The satellite missions {\sl MAP} and {\sl Planck} dominate any view of future measurements of the anisotropy of the Cosmic Microwave Background. We will attempt to look beyond and around those two experiments and see what sorts of physical questions other future projects might address. The reader has several advantages over the authors which we will not try to counter. First, many of the experiments which are in the near term future for us will be in the present or past for the reader, so we do not focus on evaluating detailed {\it anticipated\/} technical capabilities for a short list of such experiments. Readers who wish to pursue that approach might start at {\tt http://www.astro.ubc.ca/people/scott/cmb.html}, or other similar web-pages for up to date information and links. Second, readers of the rest this volume will be in a better position than we are to make informed judgements about the ideal strategies for measuring, avoiding or understanding foreground sources. Therefore, even though we think that this aspect of anisotropy will be an increasingly important and sophisticated part of the field, we have not put much emphasis on it here. As a crude aid to understanding how well future experiments are equipped to cope with foreground sources we have included a column giving the number of independent frequency channels for each experiment listed in Table~1. A view of the present situation, indicated in Figure~1 (see Smoot \& Scott 1998 for more details), sets the context for our view of the future. Even at a casual and sceptical glance these experiments seem to be converging on a power spectrum which has a peak in it. This is confirmed by careful quantitative analysis of combined data sets (Bond, Jaffe \& Knox~1998). Collectively these CMB measurements already tell us a number of fundamental things about the sort of Universe that we live in (see Lawrence, Scott \& White~1999). The prospects for future measurements look very bright indeed. Announcements of the value of $\Omega_0$, for example, are likely to (continue to) come from experiments carried out from the best terrestrial sites or suspended from stratospheric balloons, during the next few years. However, the full belief of the community in any detailed cosmological conclusions will and should await the satellite results. \begin{figure} \epsfxsize=13cm \epsfysize=12cm \epsfbox{halpern_scott_fig1_col.ps} \caption{A summary of the current anisotropy data, here presented as the extrapolated value of the quadrupole moment for a flat power spectrum, plotted against the multipole moment, which is roughly an inverse angle, $\ell\simeq135^\circ/\theta$. The dotted line is the flat power spectrum fit to {\sl COBE} slone. The solid line is the power spectrum for standard Cold Dark Matter, as an example model. We give a separate list of references at the end, and there are several other recent reviews which discuss current experiments in more detail (e.g.~Lawrence 1998).} \end{figure} Despite the steadily improving quality of experiments, we believe that none of the more recent experiments in Figure~1 would have stood as a convincing {\it discovery\/} of primordial anisotropy had it not been for {\sl COBE} (this remark certainly applies to our own experiment, {\sl BAM}, as much as to any other experiment). What was critical in the discovery was the understanding of the roles of galactic contributions and systematic errors, provided by {\sl COBE}'s all-sky coverage and comparatively stable operating environment. It was also crucial for the discovery process that the DMR on {\sl COBE} and the {\sl FIRS} balloon program showed a consistent fluctuation amplitude {\it and}, later analysis showed, correlated structures observed at very different wavelengths. Many experimenters had reassured themselves by making plots showing the similarity of the {\sl FIRS} and DMR power spectra, before the end of the day on which the DMR results were announced. There is a lesson arising from the history of the measurement of the intensity spectrum of the CMB which may be useful here. There were plenty of experiments prior to 1990 which appeared to have sufficient sensitivity to perform useful measurements, many of these with no obvious source of systematic error.\footnote{We will decline to provide examples here, reminded as we are of Winston Churchill's failed attempt to maintain parliamentary courtesy when he said that half of the members present were not asses.} The successful 1990 experiments (Gush et al.~1990, Mather et al.~1990) were performed {\it out of the atmosphere\/}, they were {\it differential\/} and they were carried out with {\it a fanatical attention to avoiding systematic errors\/} as the primary design guideline. The results were clear and reliable enough to render moot any lingering debates about inconsistencies between previous experiments. One should not be surprised to see a very similar scenario play itself out in the near term anisotropy measurements. \section{Near Term Future Experiments} Table~1 lists the properties of some future experiments. The list is meant to be illustrative of current planning; experiments are included which are past the proposal stage and for which no results are yet available. Some of the listed experiments already have data. Of course many experiments which have already produced some results and are therefore not on this list will also produce future results. All of the listed experiments involve dedicated, custom-built instrumentation. The control of systematic errors which this allows puts these experiments well ahead of attempts to use existing general purpose facilities. \begin{table} \caption{Several Future Anisotropy Experiments} \begin{center \begin{tabular}{llll} \tableline Snappy & Frequency &$\ell$-range & Number of \\ Acronym & Coverage (GHz)& & freq. channels \\ \tableline \multicolumn{4}{r}{Single Dish Telescopes}\\ \tableline {\sl MAT} & 26--46, 140--150 & 30--850 & 3 \\ {\sl MAXIMA}& 150--420 & 50--700 & 4 \\ {\sl BOOMERanG} & 100--800 & 10--700 & 4 \\ {\sl BEAST } & 25--90 & 10--800 & \\ {\sl TopHat } & 150--720 & 10--700 & 5 \\ {\sl ACBAR } &150--450 & 60--2500 & 4 \\ \tableline \multicolumn{4}{r}{Interferometers}\\ \tableline {\sl VSA }& 26--36 & 130--1800& 6 \\ {\sl CBI }& 26--36 & 300--3000& 10 \\ {\sl DASI} & 26--36 & 125--700 & 5 \\ {\sl MINT} & 140--250 & 1000--3000 & 10 \\ \tableline \multicolumn{4}{r}{Satellites}\\ \tableline {\sl MAP} & 20--106 & 1--800& 5 \\ {\sl Planck} & 30--850 & 1--1500 & 4 + 6 \\ \end{tabular} \end{center} \end{table} Sufficient sensitivity is achievable, sometimes through great technical effort. The various detection technologies available each impose constraints on experimental design and come with their own set of sources of potential systematic errors. Any serious discussion of specific systematic errors is beyond the scope of this article but we include some naive examples to illustrate the problem. Either a $100\,$mK change in the temperature of an {\it ideal\/} aluminum mirror or a $200\,$mK change in the atmosphere above a stratospheric balloon causes a radiative signal {\it 25 times larger\/} per pixel than the {\sl MAP} systematic error budget! \subsection{Systematic Errors} The careful CMB experimenter is not paranoid, but knows that the Universe is {\it in fact\/} trying to ruin the experiment. The standard answer to the question of what level of systematic error is tolerable is that there is no systematic way to handle systematic errors and, therefore, that {\it any\/} level of systematic error is a concern. We will ignore this good advice for a moment and try to estimate an answer. If the goal of an experiment is to get a rough estimate of the power spectrum of the sky, a systematic error amounting to $10\%$ of the signal amplitude contributes about $1\%$ to the power spectrum. Even if the signals are correlated in some surprising way and this estimate is wrong by a factor of a few, the effect is not likely to mask the presence of the main acoustic peak, for example. This fact is what has allowed us to get so far without a better understanding of diffuse foreground sources. On the other hand, there are important questions whose answer requires correlating many pixels in a map together in order to pick out a fairly weak efect. Measuring amplitudes of non-Gaussian statistics of a map or searching for intensity-polarization correlations are examples. In these cases the requirement for what level of signal systematic errors can contribute to a map becomes very stringent. The amplitude of systematic errors should be below the experiment's single pixel variance divided by the square root of the number of pixels to be averaged. As a numerical example, in an experiment with $0.13^\circ$ pixels and $30\,\mu$K variance averaging 1/10 of the sky, one needs to know that systematic errors are less than $0.06\,\mu$K rms for the average value not to be tainted. This is 50 times better than any experiment we have heard of. The lesson is: to produce maps of the CMB which merit careful scrutiny, avoid systematic errors like the plague.\footnote{Winston Churchill also said `One ought never to turn one's back on a threatened danger and try to run away from it. If you do that, you will double the danger. But if you meet it promptly and without flinching, you will reduce the danger by half'.} \subsection {Detection Techniques} Detectors fall into two broad categories: coherent detectors, in which the radiative electric field, including its phase, is amplified before detection; and incoherent detectors, which measure total radiative power within some frequency band. There are two types of very low noise coherent amplifier: InP high electron-mobility transistors (HEMTs); and superconductor-insulator-superconductor (SIS) mixers. HEMTs can be operated at temperatures as warm as room temperature. The noise performance gets better as they are cooled, down to $\simeq4\,$K, although amplifiers exhibit gain fluctuations at low temperature. Recently HEMT amplifiers have been made to work at frequencies well above $100\,$GHz -- noise performance is better at lower frequencies. SIS mixers are typically quieter than HEMTs and can operate at frequencies as high as $1\,$THz. However they must be cooled to $4\,$K to operate. Either of these coherent amplifiers can be used in a single telescope where the signal is amplified and detected, or as part of an interferometer in which case amplified signals from several telescopes are each multiplied with a local oscillator signal yielding lower frequency outputs which are then correlated to produce interference fringes. The advantages of coherent detectors are that they are fast, stable, not sensitive to microphonic pick-up and involve simple cryogenics. HEMTs also have the important practical advantage that many aspects of detector performance can be verified at room temperature, which greatly speeds up new instrument development. The disadvantage is that they are not as sensitive to broad band signals as incoherent detectors are. \begin{table} \caption{Detection strategies} \begin{center \begin{tabular}{llll} \tableline Snappy & Detectors & Striking & Location \\ Acronym & & Feature & \\ \tableline \multicolumn{4}{r}{Single Dish Telescopes}\\ \tableline {\sl MAT } &HEMTs and SIS & Has data & Chile 17{,}000$^\prime$\\ {\sl MAXIMA}& $100\,$mK Bolos. &Has data & Balloon\\ {\sl BOOMERanG} &$300\,$mK Bolos. &First CMB LDB flt.& Balloon, LDB\tablenotemark{a}\\ {\sl BEAST } & & & Balloon, LDB\\ {\sl TopHat } & Bolometers &Tel. {\it above\/} balloon & Balloon, LDB\\ {\sl ACBAR } &$300\,$mK Bolos. & Imaging array & S. Pole 10{,}000$^\prime$\\ \tableline \multicolumn{4}{r}{Interferometers}\\ \tableline {\sl VSA } &HFETs &14 antennae & Tenerife\\ {\sl CBI }& HEMTs at $6\,$K &13 antennae & Chile, 17{,}000$^\prime$\\ {\sl DASI} &Cooled HEMTs & 13 elements & S. Pole, 10{,}000$^\prime$\\ {\sl MINT} &SIS &6 antennae &Chile\\ \tableline \multicolumn{4}{r}{Satellites}\\ \tableline {\sl MAP} & HEMTs $<95\,$K & Differential tels. & Space, L2\tablenotemark{b}\\ {\sl Planck} &HEMTs at $20\,$K & & Space, L2\\ &$0.1\,$K Bolos. & & \\ \end{tabular} \end{center} \tablenotetext{a}{LDB = Long Duration Balloon} \tablenotetext{b}{L2 = Earth-Sun L$_2$ Lagrange point} \end{table} Incoherent detectors, in this case bolometers, can be an order of magnitude more sensitive than HEMT and SIS systems. They can be made to operate with background limited performance (BLIP), where fundamental thermodynamic fluctuations in the incident radiation field dominate over detector noise. In addition they can be made to be sensitive to a broad range of wavelengths. However, physical device size scales with wavelength and so it is easier to make small bolometers sensitive. Typically bolometers are designed for frequencies above $50\,$GHz. Bolometers are often susceptible to microphonic and radio-frequency pick-up. They are non-linear and therefore they must be characterized in their experimental operating condition, which can be very difficult for balloon and satellite experiments. They need cumbersome cryogenics to reach their operating temperatures of $0.3\,$K or colder. However, their extraordinary sensitivity and broad frequency coverage often outweigh these disadvantages. Table~2 lists some detector properties for the experiments in Table~1. \subsubsection{Interferometers} The idea of building a dedicated interferometer to study anisotropy of the CMB is not new, but improvements in detectors, and especially in broad bandwidth correlators has made this a very promising option, which is being actively pursued by several groups. Interferometers do a good job of rejecting the effects of atmospheric variations compared to beam-chopped single telescope instruments. Measurements take place essentially instantaneously, on time scales associated with the interference bandwidth, and on these time scales the atmosphere does not vary. Also, interferometers measure at slightly higher angular resolution than a single telescope of the same overall size, and in any case can easily be built for higher resolution than the currently planned space missions. This advantage will be important in exploring the expected Sunyaev-Zel'dovich forest, especially if they can also be made to work above $200\,$GHz. \begin{figure} \epsfxsize=8cm \epsfysize=8cm \epsfbox{halpern_scott_fig2.ps} \caption{The proposed pointed platform of {\sl MINT}, with four $30\,$cm aperture telescopes mounted on a single $1.5\,$m platform, illustrating how CMB interferometers are very different from the {\sl VLA}. Drawing courtesy of W.B. Dorwart, Princeton University.} \end{figure} Unlike the case for conventional radio interferometers, the individual telescopes here are crowded quite close together to keep angular resolution modest. Often, all the telescopes are mounted on a single pointed platform, which eliminates the need for signal delays before the correlators. See White et al.~(1997) for an analysis of the performance of these interferometers for measuring anisotropies. \subsubsection{Satellites} Assuming that neither suffers any serious mishap, {\sl MAP} and {\sl Planck} will produce {\it definitive\/} measurements of the primary anisotropy of the CMB, at a reliability level which the other experiments can not attain. The reliability arises form the long observation period, complete sky coverage and, primarily, the extraordinarily good observing environment at L2. Even during the 90-odd day period during which it makes its way past the moon and out to L2 to start the nominal observation program, {\sl MAP} will be in a much more thermally and radiatively stable observing environment than any previous CMB experiment. \section{After {\sl MAP} and {\sl Planck}} What will the important experimental questions be after {\sl MAP} and {\sl Planck} succeed? Clearly, measurements of the polarization of the CMB, which are explored elsewhere in this volume, will be very exciting. We also expect that studying diffuse foreground emission will become very exciting and active, as our ability to measure and identify these sources of emission develops. However, that topic is covered in the whole rest of this book so we need not consider it further here! For the remainder of this article we will discuss various ideas for what might be conceivable in the more long term future.\footnote{Ignoring the sound advice of Winston Churchill, who said `It is a mistake to try to look too far ahead. The chain of destiny can only be grasped one link at a time.'} \subsection{Anisotropy} \subsubsection{Statistics} Can phases contain information which is not {\it more\/} easily seen in the power spectrum? In principle, of course the answer is yes. But in practice, it seems clear that the smart money has to be on the negative answer. So although it would always be foolish to neglect to search for other signals, we expect that the vast majority of the primary anisotropy information will be contained in the power spectrum. Partly this is because the signals seem likely to be close to Gaussian, but also because the power spectrum (or the variance as a function of scale) is such a robust quantity -- specific patterns on the sky may require lots of phase-correlation to produce them, but much of that simply specifies the specific realization, rather than containing information about the underlying model. The supremacy of the power spectrum will certainly cease to be true for foreground signals, or indeed for a range of astrophysical processing effects that come in at smaller angular scales. One could imagine mounting a specific search for, e.g.~point or line sources on the sky, as specific examples of non-Gaussian signals. One question to ask, then, is what sort of strategy one would design to carry this out most efficiently ({\it and\/} convincingly). We find it hard to see how to avoid the conclusion that you would end up making a map, perhaps deeper and with higher resolution than otherwise, but a map nevertheless. Hence we suspect that the search for non-Gaussian signals is unlikely to be a strong driver for the design of future experiments, even if it plays a stronger role in the data analysis. A great deal of effort has been going into the study of non-Gaussian signals, e.g.~using Minkowski functionals, wavelets, etc. Given how many such tests have already been applied to {\sl COBE}, we imagine that {\it every\/} reasonable statistic will be measured for all future large data-sets. In particular we foresee an increased interest in the investigation of non-Gaussian statistics for various sorts of {\it foreground\/} signal. \subsubsection{Angular scales} Ignoring foregrounds, how far out in $\ell$ is far enough? {\sl Planck} seems sufficient for the primary signal. But that may change, depending on what we learn about foregrounds and the secondary signals, caused by various astrophysical effects, which conceptually lie in the `grey-area' between background and foregrounds. There seems to be a growing interest in these astrophysical signals at small scales, and we see no reason for that to change. It may be that the smallest angular scales are ultimately best probed with interferometers. We expect there to be secondary signal information down to the angular scales of distant galaxies, i.e. $\ell\simeq{\rm few}\times10^4$. \subsubsection{A CMB Deep Field} What might we learn from a CMB deep field? Of course, irrespective of the answer to that question, it will be done anyway! Non-Gaussian signals from higher-order effects at small-scales would certainly show up in such a map. On scales where there has been significant growth of structure, and certainly on non-linear scales, we would expect there to be significant non-Gaussianity. There seems little doubt that at some point, when the instrumentation has matured, it will be worthwhile to carry out such a CMB Deep Field. Exactly how non-Gaussian (or in other ways surprising) the small-scale signals turn out to be will determine how far beyond `cosmic variance' it is worth integrating. \subsubsection{Sunyaev-Zel'dovich effects} We are sure that, motivated by the impressive results of today's experimenters, investigation of S-Z effects will continue to grow as a sub-field. Particularly exciting is the idea of `blank sky' searches for the `S-Z Forest', or ionized gas tracing out the Cosmic Web. The thermal Sunyaev-Zel'dovich effect, or inverse Compton scattering of the CMB photons through hot gas, gives a temperature fluctuation of roughly $(kT_{\rm e}/m_{\rm e}c^2)\tau$, where $T_{\rm}$ and $m_{\rm e}$ are the temperature and mass of the electron, respectively, and $\tau$ is the optical depth through the ionized gas. There is also a spectral shape, distinct from the CMB blackbody, of a well-known form (see e.g.~Sunyaev \& Zel'dovich~1980) Detailed studies of the thermal S-Z effect for particular clusters will provide constraints on the morphology, clumping, thermal state of the gas and projected mass, amongst other things. The power spectrum of these fluctuations peaks at $\ell\simeq2000$ typically, with $\left\langle Q_{\rm flat}\right\rangle$ amplitude of a few $\mu$K, although with great variation between models. Detailed investigation of this power spectrum might further distinguish between cosmological models, and between ideas for cluster formation. The power spectrum for the kinematic effect, and for related effects (due to variations in potential, for example) are generally much smaller. Several of these `higher-order' Sunyaev-Zel'dovich type effects are potentially measurable for {\it individual\/} clusters, and will doubtless be attempted in the future (see the review by Birkinshaw~1998). Certainly the kinematic effect (which depends on the line-of-sight velocity and is $\sim(v_{\parallel}/c)\tau$) can be measured for some clusters. However, this effect has the same spectrum as a CMB fluctuation, and so the small-angle CMB anisotropies act as a source of `noise', making is difficult to measure the velocities to better than a few hundred ${\rm km}\,{\rm s}^{-1}$. One further effect uses the polarization in the CMB scattered by the kinematic S-Z effect, which depends on the cluster's transverse velocity (actually $\sim (v^2/c^2)\tau$). In principle, together with the kinematic S-Z effect itself, this gives a means of estimating the full 3 dimensional velocity of clusters. Although difficult to measure, this polarization signal has a frequency dependence which may help to disentangle it from other effects (Audit \& Simmons~1999). There are other spectral signals expected from non-thermal electron populations, for example in the lobes of radio sources. However, the utilisation of such measurements to study the lobe properties will require extremely high angular resolution. \subsubsection{Other secondary effects} There are several other known secondary effects (see, e.g.~Hu et al.~1995, and other contributions to this volume), and surely many other {\it unknown\/} ones! One effect which has been studied in some detail is a second-order coupling between density and velocity, usually referred to as the Vishniac effect. In a sense this can be thought of as specific case of the kinematic Sunyaev-Zel'dovich effect. For Cold Dark Matter type models the signal is typically $\simeq1\,\mu$K and peaking at $\ell\simeq{\rm few}\times10^3$ (e.g.~Hu \& White~1996, Jaffe \& Kamionkowski~1998). Although certainly difficult to measure, it is nevertheless feasible, and worth pursuing, since measurement can give direct information about reionization. Additional structure ar small scales (e.g.~from an isocurvature mode) could increase the signal. In addition there will be polarization effects, although these are likely to be {\it really\/} small. Patchy reionization (discussed elsewhere in this volume) is just another S-Z effect, and tends to be dominated by the kinematic source from moving bubbles of gas as the Universe undergoes reionization. The amplitude of this signal seems likely to be smaller than for the Vishniac effect, although it is as yet unclear what the predictions will be for realistic models which include inhomogeneous reionization (with radiative transfer through voids etc.), distributions of sources, and other complications. Again there may be polarization effects, and correlations with other signals, which, in principle, could be used to pull the signal out. In addition there may also be a measurable S-Z signal from the Ly$\,\alpha$ forest, on scales well below an arcminute, and with amplitude perhaps as high as a few $\mu$K (Loeb~1996). Rees-Sciama, or varying potential fluctuations tend to be rather small in amplitude ($<10^{-6}$ in fractional temperature change), but not negligibly so. Here again there are a number of effects, in particular those caused by time-varying potentials in the light-crossing time, and those caused by potentials moving across the line of sight (e.g.~Tuluie, Laguna \& Anninos~1996). These will have CMB-like spectra, and the signal will be dominated by non-linear structures (meaning that the statistics will be highly non-Gaussian). The effects may peak at relatively small angles $\ell\simeq{\rm few}\times100$, but there they will be well below the primary signal, and hard to disentangle. So the best prospects for detection may be at smaller scales, where the primary power spectrum is falling off. Detection may also be easier using correlations with other signals. And certainly such signals are unlikely to be Gaussian, and so may be teased out of the data by looking at their statistics (e.g.~Spergel \& Goldberg~1999). Gravitational lensing affects the CMB power spectrum by smearing the anisotropies, thereby smoothing out features in the power spectrum. The temperature field is affected by this smearing, so that combinations of derivatives can be used to extract the lensing signal directly, at least in principle (Seljak \& Zaldarriaga~1999). The projected matter field can be reconstructed through a combination of this technique and correlations with other signals (Zaldarriaga \& Seljak~1999). For example, the large angle signal caused by the variation in gravitational potential (the `ISW effect') may be correlated with the lensing signal in open or $\Lambda$-dominated models. However, the level of such a correlated signal is not likely to be large. One can easily imagine searching for all sorts of other correlations, for example the lensing signal with S-Z signals, with surveys at other wavelengths, e.g.~large-scale structure, X-ray maps, etc. \subsubsection{Spatial-spectral signals} At the moment the only significant signal which mixes both spatial and spectral deviations is the S-Z effect. Although we have no specific ideas, we imagine that other such effects, involving perhaps different scattering processes, are likely to be developed in the near future. Although we expect the effect to be quite small, we mention as an example that Rayleigh scattering, which would spectrally filter anisotropy signals, has been omitted from calculations. In addition there could in principle be resonant line scattering from molecules in clouds at high redshift. Searches for such mixed spatial-spectral signals seem likely to become more important as multi-frequency data-sets improve in quality and quantity. \subsection{Non-anisotropy} Non-anisotropy measurements are heroically hard to do; certainly such things are worth pursuing, but the immediate returns are not as obvious as for the current anisotropy prospects. On the other hand, we expect that effort will fairly soon return to this direction when the `easy' results have been mined from the primary power spectrum. Here we simply list a number of possibilities. Figure~3 shows the form of some of the standard distortions to the CMB spectrum. \begin{figure} \epsfxsize=11cm \epsfbox{halpern_scott_fig3.ps} \caption{Shapes for some theoretical possibilities for spectral distortions. The amplitudes here are arbitrary. However, the FIRAS 95\% confidence limits for the amplitudes of these distortions are that the chemical potential is less than $15\,$mK at the peak and that free-free is less than $10\,$mK at $4\,$GHz. At those frequencies the Galaxy is perhaps 100 times brighter. See Smoot (1997) for further discussion.} \end{figure} A FIRAS/COBRA style total emission measurement of the spectrum of the sky will almost entirely be dominated by foregrounds outside of the 20 to $400\,$GHz frequency window in which the spectrum is already well measured. The present FIRAS limits allow parametrized spectral distortions as large as 10s of mK, easily larger than the measurement uncertainty in a careful experiment, but 100 to 1000 times dimmer than diffuse galactic emission at those wavelengths! Perhaps a multi-frequency measurement with appropriate angular resolution and sky coverage will allow a reliable extrapolation to zero galactic emission, but it will not be easy. Details for commonly considered distortions are listed below. \subsubsection{Compton distortions} $y$-distortions have essentially already been measured, in the sense that the sum of all the S-Z detected structure will give the uniform Compton-distortion over the sky. Certainly this gives a lower limit, which seems likely to be the bulk of the detectable signal (barring unforeseen exotic processes). The size of this signal is estimated to be $y\sim10^{-6}$ (e.g.~Colafrancesco et al.~1997), depending on the cosmology. After the {\sl Planck} mission (and S-Z investigations from ground-based interferometers) we will have a very precise estimate for the uniform $y$-distortion (and indeed an estimate of its power spectrum as well). Between this underlying signal, and the FIRAS upper limit on a full-sky distortion, there will remain only a rather narrow window to search for possible isotropic $y$-distortions from other sources (such as late energy injection, unrelated to cluster formation). Since there are no immediate candidates for such processes, and the window is rather narrow, we don't see this as a particularly strong motivation for mounting a next generation FIRAS mission. \subsubsection{Free-free emission} For late energy releases, free-free emission leads to a distortion in the CMB spectrum, which increases towards lower frequency. This seems to be the type of distortion which is most feasible to measure in the near future for realistic models of the Universe. The best upper limits at the moment imply free-free optical depths of order $Y_{\rm ff}\la10^{-5}$ (e.g.~Nordberg \& Smoot~1998). Since this distortion increases at lower frequencies, then it is best investigated at the lowest frequency at which foreground signals can be dealt with, which means somewhere around $5\,$GeV. The expected signal at these frequencies may be as high as $100\,\mu$K, corresponding to $Y_{\rm ff}$ only about an order of magnitude below the current limits. The planned experiments ARCADE and DIMES (Kogut~1996) may be able to reach into the parameter space for realistic models, and help us understand more about the early ionized stages of the intergalactic medium. One nice thing about free-free is that lowering the temperature of the ionized medium {\it increases\/} the distortion (approximately $\propto n_{\rm e}^2/\sqrt{T_{\rm e}}$), even although it decreases the Compton distortion. Hence good limits on $Y_{\rm ff}$ imply either low reionization redshifts or high electron temperatures, and limits on $y$ would restrict $T_{\rm e}$, so that direct limits on $z_{\rm reion}$ could be obtained. \subsubsection{Chemical potential} Current limits on $\mu$-type distortions are at the $10^{-5}$ level. Note that this allows about $15\,$mK at $1\,$GHz within the error budget of the measurements, which is about 0.1\% of the galactic signal. So pushing that limit further down is going to be tricky! The way to do this would presumably be to make a spectral map of the sky and extrapolate to zero galaxy (essentially what FIRAS did). So how big could a signal be? Some amount of $\mu$-distortion is unavoidable, since it is generated by the damping of small-scale perturbations. For realistic models the value is likely to be around ${\rm few}\times10^{-8}$ (Hu, Scott \& Silk~1994), which seems unlikely ever to be measurable. Of course various exotic processes, including energy injection at redshifts $z\sim10^5$ could give much higher values of $\mu$. Limits could be set by experiments which also constrain free-free signals. However, we see no compelling reason currently to invest heavily in future experiments seeking to measure $\mu$ itself. Of course, if any hint of signal were to turn up then that would be extremely exciting (since unexpected) -- in that case further investigation of the turn-off in the distortion at low frequencies would probe an otherwise unexplored early epoch. \subsubsection{Recombination lines} When the Universe recombined, every atom emitted at least one Lyman photon, or else got from the first excited state to the ground state via the two-photon process (see Seager, Scott \& Sasselov~1999 for more details). This is a lot of photons, waiting there to be discovered! Mere measurement of the background of these photons would be an unprecedented confirmation of the Big Bang paradigm, that the Universe became neutral at $z\simeq1000$. Further investigation of these recombination lines would be a direct probe of the recombination process, and might provide further cosmological constraints. For example, the strengths of the residual Ly$\,\alpha$ feature and the two-photon feature will depend on the baryon density and on the expansion rate, hence allowing measurements of $\Omega_{\rm B}$ and $H$ at $z\simeq1000$. The problem is that the main recombination lines lie at wavelengths $\lambda\simeq150\,\mu$m, where the signal from the galaxy is orders of magnitude stronger. The way to try to find the signal is then presumably to have enough spatial information to be able to extrapolate to zero Galaxy, and at the same time to have adequate spectral information to distinguish the relatively wide spectral feature. If all else failed it might be possible to rely on the dipole to extract the cosmological signal, but that would be even more difficult. So we might envisage an experiment with reasonable sky coverage, low angular resolution, but at least 3 spectral channels (say in the range 100--200$\,\mu$m) to extract the wide line. The spectral resolution would have to be good enough to distinguish this from a roughly isotropic component of warm interstellar dust -- but that should be possible given that the spectral shape of the recombination lines is calculable (Dell'Antonio \& Rybicki~1993, Boschan \& Biltzinger~1998). \subsubsection{21\,cm line studies} If the Universe became reionized at redshifts between 5 and 20 there should be a spectral feature due to red-shifted $21\,$cm emission from neutral hydrogen which appears today at $70$ to $240\,$MHz (see Shaver et al.~1999). This emission can be seen against the CMB provided that there are spatial or spectral signatures (e.g.~Tozzi, et al.~1999) and a mechanism which decoupled the electron spin temperature from the CMB. In principle, such studies, using the proposed Square Kilometer Array for example, could provide information about the processes that marked the end of the so-called Dark Ages, i.e.~the reionization process and the formation of the first structures. This endeavor is sometimes called `cosmic tomography'. \subsubsection{Other diagnostics of the `Dark Ages'} There are of course other ways of probing the end of the Dark Ages, and even into that epoch, many of which might come from entirely different wavelengths, for example the near infra-red with {\sl NGST}. However, we imagine that the microwave band will continue to be important in furnishing new ideas for exploring the domain between $z=5$ and $z=1000$. One recent speculative idea involves searching for masers which may come from structures at either the recombination or reionization epochs (Spaans \& Norman~1997). There will surely be other such ideas in the coming years. \subsubsection{Measurements of $T_{\rm CMB}(z)$} The currently best value for the CMB temperature is $T_0=2.725\pm0.001\,$K (Mather et al.~1999). It seems unclear why anyone should care about a more precise measurement than that! Before the existence of the CMB was even suspected, there was evidence for excess excitation in line ratios of certain molecules, notably cyanogen , in the interstellar medium (McKellar,~1941). This method has more recently been used to constrain the CMB temperature at high redshifts ($z\sim{\rm few}$) using line with excitation temperatures of the relevant energy (e.g.~Songaila et al.~1994, Roth \& Bauer~1999). Measurements of other line ratios etc.~can be used to set limits on the variation of fundamental physical constants (e.g.~Webb et al.~1999). In a similar way, detailed measurement of the blackbody shape indicates that certain combinations of fundamental constants have not varied much since $z\sim1000$. \begin{figure} \epsfxsize=14cm \epsfbox{halpern_scott_fig4_col.ps} \caption{A compilation of recent constraints on extragalactic diffuse background radiation. In terms of total energy the CMB and Far-Infrared Backgrounds dominate. The data are collected primarily from Ressel \& Turner~1990, Smoot~1997, Lagache et al.~1999, Hauser et al.~1998, Dwek \& Arendt~1998, Pozzetti et al.~1998, Leinert et al.~1998, Miyaji et al.~1998, Sreekumar et al.~1998, and Kappadath et al.~1999. In this colour version lower limits are shown in red and upper limits in blue.} \end{figure} \section{Epilogue} Assuming that {\sl MAP} and {\sl Planck} are fully successful, and that the current suite of ground- and balloon-based experiments also return exquisite data, what then? Will this be then end of the study of the CMB?\footnote{Churchill warned that `success is never final'. He also pointed out that `it is a good thing for an uneducated man to read books of quotations'.} Eventually we can imagine a time when the primordial anisotropies have been measured so accurately that there are diminishing returns from further generations of satellite missions, and when small scale measurements, involving non-Gaussian signals, mixed spatial-spectral signals, and other complications, have moved firmly into the regime of `messy astrophysics'. However, there will be further primordial information to unlock from ever more ambitious polarization experiments. Certainly the CMB should not be looked at in isolation -- although it is the dominant diffuse extragalactic background, there are several others to study (see Figure~4). And if that doesn't leave the future still filled with exciting and challenging possibilities, there's always the cosmic neutrino background! \acknowledgments We thank the editors for their patience.
\section*{Introduction} The purpose of the Leiden--Berkeley Deep Survey (hereafter ``the LBDS'') was to gain a better understanding of the nature of faint radio galaxies and quasars, and to determine their cosmological evolution. Several high latitude fields in the selected areas SA28, SA57, SA68 and an area in Hercules had been selected for the purpose of faint galaxy and quasar photometry, and a collection of good multi-color prime focus photographic plates had been acquired. Nine of these fields were then surveyed with the Westerbork Synthesis Radio Telescope at 21 cm (1.412 GHz), reaching a 5-$\sigma$ limiting flux density of 1~mJy \cite{Windhorst84a}. Following this selection of the radio sample, 171 of the radio sources (53\%) were identified on the photographic plates, whilst for the Hercules fields there were 47 out of 73 sources identified \cite{Windhorst84b,Kron85}. Presented here are the results of an extensive optical/infrared investigation of the two Hercules fields, with the aim of completing the identification and redshift content of this sub-sample\cite{Waddington98}. A cosmology of ${\rm H_0}=50$~km s$^{-1}$ Mpc$^{-1}$, $\Omega_0=1$ and $\Lambda=0$ is assumed throughout. \section*{The data} The Hercules field was observed on the 200~inch Hale telescope at Palomar Observatory between 1984 and 1988. Multiple observations were made through Gunn $g$, $r$ and $i$ filters over the six runs. After processing and stacking of the multiple-epoch images, optical counterparts for 22 of the sources were found, leaving only four sources unidentified to $r\simeq 26$. Near-infrared observations have been made of the entire subsample at $K$, yielding 60/73 detections down to $K\simeq 19$--21. Half of the sources have been observed in $H$ and approximately one-third in $J$. Observations of the brighter sources were made by Thuan et~al.~(1984)\cite{Thuan84} and by Neugebauer et~al.\ and Katgert et~al.\ (priv.\ comm.). $K$-band observations of the sample were completed by the present authors at UKIRT. Figure~1 presents the optical and infrared magnitude distributions. For those sources without CCD observations, photographic magnitudes from Kron et~al.~(1985)\cite{Kron85} have been transformed to the Gunn system\cite{Windhorst91}. It is seen that the distribution turns over at $r \sim 22$, a consequence of evolution in the redshift and/or luminosity distributions of the radio sources. The the $r$-band magnitude distribution is essentially unchanged from this milli-Jansky survey down to micro-Jansky surveys, a thousand times fainter in radio flux\cite{Windhorst98}. \begin{figure} \centerline{% \epsfig{file=waddington_fig1a.ps,width=7cm,height=8cm} \ \ \ \ \ \epsfig{file=waddington_fig1b.ps,width=7cm,height=8cm}} \vspace{10pt} \caption{Magnitude distributions for the LBDS Hercules sample. Shaded histograms show the sources with $S_{1.4}\ge 2$~mJy. Arrows denote 3-$\sigma$ upper limits at $H$ and $K$.} \label{figureone} \end{figure} Prior to the start of the current work, only 16 of the 73 sources in the LBDS Hercules fields had redshifts published in the literature. Another 16 sources had unpublished redshifts. The author and collaborators have successfully observed a further 17 sources during the past few years, using both the 4.2~m William Herschel Telescope\cite{Waddington98} and the 10~m W. M. Keck Telescope\cite{Dunlop96,Spinrad97,Dey97}. This brings the total number of redshifts to 49 out of 73 sources (67\%). Photometric redshifts were calculated for the remaining one-third of the sample. Using the spectral population synthesis models of Jimenez et~al.~(1998)\cite{Jimenez98}, synthetic $griJHK$ magnitudes were computed and fitted to the observed magnitudes, giving the most-probable redshift and a measure of its uncertainty. Comparison of the estimated and the true redshifts for those sources with spectroscopic observations, showed that the average difference was $\sim 0.1$ in $z$. \section*{The 1.4~GHz radio luminosity function and the redshift cut-off} Dunlop and Peacock~(1990)\cite{Dunlop90} used a sample of radio sources brighter than 0.1~Jy at 2.7~GHz to investigate the radio luminosity function. They concluded that the comoving density of both flat- and steep-spectrum sources suffers a cut-off at redshifts $z\simeq 2$--4. This conclusion was drawn from the behavior of both free-form and simple parametric models (PLE/LDE), and the model-independent, banded $V/V_{\rm max}$ test. However, the results were crucially dependent upon the accuracy of their redshift estimates in the Parkes Selected Regions (PSR). \begin{figure} \centerline{% \epsfig{file=waddington_fig2a.ps,width=8cm,% height=8cm} \ \ \ \ \ \epsfig{file=waddington_fig2b.ps,width=6cm,% height=8cm}} \vspace{10pt} \caption{{\bf [Left]} The cumulative redshift distribution of all sources in the 2-mJy Hercules sample. The bold histogram is computed from the best-fit photometric redshift distribution, the lighter histograms correspond to the lower and upper limits to the photometric redshifts. Lines are the model RLFs of [1]. {\bf [Right]} The observed radio luminosity function for the 2-mJy Hercules sample, for each of the three photometric redshift distributions.} \label{figuretwo} \end{figure} With a flux limit $\sim 100\times$ fainter than the PSR, the LBDS is well-suited to test the reliability of those RLF models and the redshift cut-off, via its potential to detect powerful radio galaxies at very high redshifts. In figure~2 [left] the cumulative redshift distribution of the LBDS Hercules sample (only sources with $S_{1.4}\ge 2$~mJy) is compared with the predictions of \cite{Dunlop90}. It is seen that two of the free-form models (FF-4 and FF-5) provide a reasonable fit to the data over all redshifts. The ``bump'' in the best-fit histogram at $0.4\mathrel{\hbox{\rlap{\hbox{\lower5pt\hbox{${^\sim}$}}}\hbox{\lower0.5pt\hbox{${^<}$}}}} z \mathrel{\hbox{\rlap{\hbox{\lower5pt\hbox{${^\sim}$}}}\hbox{\lower0.5pt\hbox{${^<}$}}}} 1$ is due to two spikes in the redshift distribution, that may be the result of possible large-scale structures (sheets) along the line of sight. The observed 1.4~GHz luminosity function presented in figure~2 [right] was also compared with the models. It was found that the two models which fit the cumulative counts (FF-4 and FF-5) do not predict the observed {\it luminosity\/} dependence of the data nearly as well as the overall redshift dependence. The observed RLF shows some indication that it turns over at $z\simeq 0.5$--1.5, and that the redshift of this cut-off is a function of the radio luminosity. However, the small number of sources makes it difficult to separate the redshift and luminosity dependence of the RLF sufficiently to be certain of this trend. The full results of this project are presented in \cite{Waddington98}, and in forthcoming papers by the author and collaborators. \noindent {\bf Acknowledgments:} Many people have contributed data and knowledge to this project. In particular, I thank James Dunlop, Rogier Windhorst and John Peacock for their assistance. The financial support of the PPARC is acknowledged.
\section{Introduction} \label{sec:intro} The idea of using a solid elastic sphere as a gravitational wave (GW) antenna is almost as old as that of using cylindrical bars: as far back as 1971 Forward published a paper \cite{fo71} in which he assessed some of the potentialities offered by a spherical solid for that purpose. It was however Weber's ongoing philosophy and practice of using bars which eventually prevailed and developped up to the present date, with the highly sophisticated and sensitive ultracryogenic systems currently in operation ---see \cite{amaldi} and \cite{gr14} for reviews and bibliography. With few exceptions~\cite{ad75,wp77}, spherical detectors fell into oblivion for years, but interest in them strongly re-emerged in the early 1990's, and an important number of research articles have been published since which address a wide variety of problems in GW spherical detector science. At the same time, international collaboration has intensified, and prospects for the actual construction of large spherical GW observatories (in the range of $\sim$100 tons) are being currently considered in several countries \footnote{ There are collaborations in Brazil, Holland, Italy and Spain.}, even in a variant {\it hollow\/} shape \cite{vega}. A spherical antenna is obviously omnidirectional but, most important, it is also a natural {\it multimode\/} device, i.e., when suitably monitored, it can generate information on all the GW amplitudes and incidence direction \cite{nadja}, a capability which possesses no other {\it individual\/} GW detector, whether resonant or interferometric \cite{dt}. Furthermore, a spherical antenna could also reveal the eventual existence of {\it monopole\/} gravitational radiation, or set thresholds on it \cite{maura}. The theoretical explanation of these facts is to be found in the unique matching between the GW amplitude structure and that of the sphere oscillation eigenmodes: a general {\it metric\/} GW generates a {\it tidal\/} field of forces in an elastic body which is given in terms of the ``electric'' components $R_{0i0j}(t)$ of the Riemann tensor at its centre of mass by the following formula \cite{lobo}: \begin{equation} {\bf f}_{\rm GW}({\bf x},t)\ \ \ = \sum_{\stackrel{\scriptstyle l=0\ {\rm and}\ 2}{m=-l,...,l}}\, {\bf f}^{(lm)}({\bf x})\,g^{(lm)}(t) \label{1.1} \end{equation} where ${\bf f}^{(lm)}({\bf x})$ are ``tidal form factors'', while $g^{(lm)}(t)$ are specific linear combinations of the Riemann tensor components $R_{0i0j}(t)$ which carry all the {\it dynamical\/} information on the GW's monopole ($l\/$\,=\,0) and quadrupole ($l\/$\,=\,2) amplitudes. It is precisely these amplitudes which a GW detector is aimed to measure. On the other hand, a free elastic sphere has two families of oscillation eigenmodes, so called {\it toroidal\/} and {\it spheroidal\/} modes, and modes within either family group into ascending series of $l\/$-pole harmonics, each of whose frequencies is (2$l\/$+1)-fold degenerate ---see \cite{lobo} for full details. It so happens that {\it only\/} monopole and/or quadrupole spheroidal modes can possibly be excited by an incoming {\it metric\/} GW \cite{bian}, and their GW driven amplitudes are directly proportional to the wave amplitudes $g^{(lm)}(t)$ of equation (\ref{1.1}). It is this very fact which makes of the spherical detector such a natural one for GW observations \cite{lobo}. In addition, a spherical antenna has a significantly higher absorption {\it cross section\/} than a cylinder of like fundamental frequency, and also presents good sensitivity at the {\it second\/} quadrupole harmonic \cite{clo}. In order to monitor the GW induced deformations of the sphere {\it motion sensors\/} are required. In cylindrical bars, current state of the art technology is based upon {\it resonant transducers\/} \cite{pia,hamil}. A resonant transducer consists in a small (compared to the bar) mechanical device possessing a resonance frequency accurately tuned to that of the cylinder. This {\it frequency matching\/} causes back-and-forth {\it resonant energy transfer\/} between the two bodies (bar and resonator), which results in turn in {\it mechanically amplified\/} oscillations of the smaller resonator. The philosophy of using resonators for motion sensing is directly transplantable to a spherical detector ---only a {\it multiple\/} set rather than a single resonator is required if its potential capabilities as a multimode system are to be exploited to satisfaction. A most immediate question in a multiple motion sensor system is: {\it where\/} should the sensors be? The answer to this basic question naturally depends on design and purpose criteria. Merkowitz and Johnson (M\&J) made a very appealing proposal consisting in a set of 6 identical resonators coupling to the {\it radial\/} motions of the sphere's surface, and occupying the positions of the centres of the 6 non-parallel pentagonal faces of a truncated icosahedron \cite{jm93,jm95}. One of the most remarkable properties of such layout is that there exist 5 linear combinations of the resonators' readouts which are directly proportional to the 5 quadrupole GW amplitudes $g^{(2m)}(t)$ of equation (\ref{1.1}). M\&J call these combinations {\it mode channels\/}, and they therefore play a fundamental role in GW signal deconvolution in a real, {\it noisy\/} system \cite{m98,lms}. In addition, a reduced scale prototype antenna ---called {\sl TIGA\/}, for {\sl T\/}runcated {\sl I\/}cosahedron {\sl G\/}ravitational {\sl A}ntenna--- was constructed at Lousiana State University, and its working experimentally put to test~\cite{phd}. The remarkable success of this experiment in almost every detail \cite{jm96,jm97,jm98} stands as a vivid proof of the practical feasibility of a spherical GW detector \cite{sfera}. Despite its success, the theoretical model proposed by M\&J to describe the system dynamics is based upon a simplifying assumption that the resonators {\it only\/} couple to to the quadrupole vibration modes of the sphere \cite{jm93,jm95}. While this is seen {\it a posteriori\/} of experimental measurements to be a very good approximation \cite{phd,jm97}, a deeper {\it physical\/} reason which explains {\it why\/} this happens is missing so far. The original motivation for the research we present in this article was to develop a new and more general approach for the analysis of the resonator problem, very much in the spirit of the methodology and results of reference \cite{lobo}; this, we thought, would not only provide the necessary tools for a rigorous analysis of the system dynamics, but also contribute to improve our understanding of the physics of the spherical GW detector. Pursuing this programme, we succeeded in setting up and solving the equations of motion for the coupled system of sphere plus resonators. The most important characteristic of our solution is that it is expressible as a {\it perturbative series expansion in ascending powers of the small parameter $\eta^{1/2}$\/}, where $\eta\/$ is the ratio between the average resonator's mass and the sphere's mass. The dominant (lowest) order terms in this expansion appear to exactly reproduce Merkowitz and Johnson's equations \cite{jm95}, whence a quantitative assessment of their degree of accuracy, as well as of the range of validity of their underlying hypotheses obtains; if further precision is required then a well defined procedure of going to next (higher) order terms is unambiguously prescribed by the system equations. Beyond this, though, the simple and elegant algebra which emerges out of the general scheme has enabled us to explore different resonator layouts, alternative to the unique {\sl TIGA\/} of M\&J. In particular we have found one \cite{ls,lsc} requiring 5 rather than 6 resonators per quadrupole mode sensed and possessing the remarkable property that {\it mode channels\/} can be constructed from the system readouts, i.e., five linear combinations of the latter which are directly proportional to the five quadrupole GW amplitudes. We have called this distribution {\sl PHC\/} ---see below for full details. The intrinsically perturbative nature of our approach makes it also particularly well adapted to assess the consequences of small defects in the system structure, such as for example symmetry breaking due to suspension attachments, small resonator mistunings and mislocations, etc.\ We have applied them with outstanding success to account for the reported frequency measurements of the {\sl LSU TIGA\/} prototype \cite{phd}, which was diametrically drilled for suspension purposes; in particular, discrepancies between measured and calculated values (generally affecting only the {\it fourth\/} decimal place) are precisely of the theoretically predicted order of magnitude. We have also applied our methods to analyse the stability of the spherical detector to several mistuned parameters, with the result that it is not very sensitive to small construction errors. This conforms again to experimental reports \cite{jm98}, but has the advantage that the argument depends on {\it analytic\/} mathematical work rather than on computer simulated procedures, the only ones available to this date to our knowledge ---see e.g.\ \cite{jm98} or \cite{ts}. The paper is structured as follows: in section~\ref{sec:GE} we present the main physical hypotheses of the model, and the general equations of motion. In section~\ref{sec:gff} we set up a Green function approach to solve those equations, and in section~\ref{sec:srgw} we apply it to assess the system response to both monopole and quadrupole GW signals. In section~\ref{sec:PHC} we describe in detail the {\sl PHC\/} layout, including its frequency spectrum and {\it mode channels\/}. Section~\ref{sec:hs} contains a few brief considerations on the system response to a hammer stroke calibration signal, and finally in section~\ref{sec:symdef} we study how the different parameter mistunings affect the detector's behaviour. The paper closes with a summary of conclusions, and three appendices where the heavier mathematical details are made precise for the interested reader. \section{General equations} \label{sec:GE} With minor improvements, we shall use the notation of references \cite{lobo} and \cite{ls}, some of which is now briefly recalled. We consider a solid sphere of mass $\cal M\/$, radius $R\/$, (uniform) density $\varrho\/$, and elastic Lam\'e coefficients $\lambda$ and $\mu\/$, endowed with a set of $J\/$ resonators of masses $M_a\/$ and resonance frequencies $\Omega_a\/$ ($a\/$\,=\,1,\ldots,$J\/$), respectively. We shall model the latter as {\it point masses\/} attached to one end of a linear spring, whose other end is rigidly linked to the sphere at locations~${\bf x}_a\/$ ---see Figure \ref{fig1}. The system degrees of freedom are given by the {\it field\/} of elastic displacements ${\bf u}({\bf x},t)$ of the sphere plus the {\it discrete\/} set of resonator spring deformations $z_a(t)$; equations of motion need to be written down for them, of course, and this is our next concern in this section. \begin{figure} \label{fig1} \psfig{file=fig1.ps,height=17cm,width=12cm,rheight=6.8cm,bbllx=-3cm,bblly=-7.4cm,bburx=14.2cm,bbury=19cm} \caption{Schematic diagramme of the coupling model between a solid sphere and a resonator. The notation is that in the text, but subindices have been dropped for clarity. The dashed-dotted arc line on the left indicates the position of the {\it undeformed\/} sphere's surface, and the solid arc its {\it actual\/} position.} \end{figure} We shall assume that the resonators only move radially, and also that Classical Elasticity theory \cite{ll70} is sufficiently accurate for our purposes\footnote{ We clearly do not expect relativistic motions in extremely small displacements at typical frequencies in the range of 1 kHz.}. In these circumstances we have \cite{ls} \begin{eqnarray} \varrho\,\frac{\partial^2 {\bf u}}{\partial t^2} & = & \mu\nabla^2 {\bf u} + (\lambda+\mu)\,\nabla(\nabla{\bf\cdot}{\bf u}) + {\bf f}({\bf x},t) \label{2.1.a} \\*[0.7 em] \ddot{z}_a(t) & = & -\Omega_a^2\, \left[z_a(t)-u_a(t)\right]+\xi_a^{\rm external}(t)\ , \qquad a=1,\ldots,J \label{2.1.b} \end{eqnarray} where ${\bf n}_a\/$\,$\equiv$\,${\bf x}_a/R\/$ is the outward pointing normal at the the $a\/$-th resonator's attachment point, and \begin{equation} u_a(t)\equiv{\bf n}_a\!\cdot\!{\bf u}({\bf x}_a,t)\ ,\qquad a=1,\ldots,J \label{3.8} \end{equation} is the {\it radial\/} deformation of the sphere's surface at ${\bf x}_a\/$. A dot (\,$\dot{}$\,) is an abbreviation for time derivative. The term in square brackets in (\ref{2.1.b}) is thus the spring deformation ---$q(t)$ in Figure \ref{fig1}. ${\bf f}({\bf x},t)$ in the rhs of (\ref{2.1.a}) contains the {\it density\/} of all {\it non-internal\/} forces acting on the sphere, which we expediently split into a component due the resonators' {\it back action\/} and an external action {\it proper\/}, which can be a GW signal, a calibration signal, etc. Thus \begin{equation} {\bf f}({\bf x},t) = {\bf f}_{\rm resonators}({\bf x},t) + {\bf f}_{\rm external}({\bf x},t) \label{2.2} \end{equation} Finally, $\xi_a^{\rm external}(t)$ in the rhs of (\ref{2.1.b}) is the force per unit mass (acceleration) acting on the $a\/$-th resonator due to {\it external\/} agents. Since we are making the hypothesis that the resonators are point masses the following holds: \begin{equation} {\bf f}_{\rm resonators}({\bf x},t) = \sum_{a=1}^J M_a\Omega_a^2\,\left[\,z_a(t)-u_a(t)\right]\, \delta^{(3)}({\bf x}-{\bf x}_a)\,{\bf n}_a \label{2.3} \end{equation} where $\delta^{(3)}\/$ is the three dimensional Dirac density function. The {\it external\/} forces we shall be considering in this paper will be {\it gravitational wave\/} signals, and also a simple calibration signal, a perpendicular {\it hammer stroke\/}. GW driving terms, we recall from~(\ref{1.1}), can be written \begin{equation} {\bf f}_{\rm GW}({\bf x},t) = {\bf f}^{(00)}({\bf x})\,g^{(00)}(t)\ +\ \sum_{m=-2}^2\,{\bf f}^{(2m)}({\bf x})\,g^{(2m)}(t) \label{2.4} \end{equation} for a general {\it metric\/} wave ---see \cite{lobo} for explicit formulas and technical details. While the spatial coefficients ${\bf f}^{(lm)}({\bf x})$ are pure {\it form factors\/} associated to the {\it tidal\/} character of a GW excitation, it is the time dependent factors $g^{(lm)}(t)$ which carry the specific information on the incoming GW. The purpose of a GW detector is to determine the latter coefficients on the basis of suitable measurements. If a GW sweeps the observatory then the resonators themselves will also be affected, of course. They will be driven, relative to the sphere's centre, by a tidal acceleration which, since they only move radially, is given by \begin{equation} \xi_a^{\rm GW}(t) = c^2\,R_{0i0j}(t)\,x_{a,i}n_{a,j}\ , \qquad a=1,\ldots,J \label{c.1} \end{equation} where $R_{0i0j}(t)$ are the ``electric'' components of the GW Riemann tensor at the centre of the sphere. These can be easily manipulated to give\footnote{ $Y_{lm}({\bf n})$ are spherical harmonics \protect\cite{Ed60} ---see also the multipole expansion of $R_{0i0j}(t)$ in reference \cite{lobo}.} \begin{equation} \xi_a^{\rm GW}(t) = R\, \sum_{\stackrel{\scriptstyle l=0\ {\rm and}\ 2}{m=-l,...,l}}\, Y_{lm}({\bf n}_a)\,g^{(lm)}(t)\ ,\qquad a=1,\ldots,J \label{c.4} \end{equation} where $R\/$ is the sphere's radius. We shall also be later considering in this paper the response of the system to a particular {\it calibration\/} signal, consisting in a hammer stroke with intensity ${\bf f}_0$, delivered perpendicularly to the sphere's surface at point ${\bf x}_0$: \begin{equation} {\bf f}_{\rm stroke}({\bf x},t) = {\bf f}_0\, \delta^{(3)}({\bf x}-{\bf x}_0)\,\delta(t) \label{2.5} \end{equation} which we have modeled as an impulsive force in both space and time variables. Unlike GW tides, a hammer stroke will be applied on the sphere's surface, so it has no {\it direct\/} effect on the resonators. In other words, \begin{equation} \xi_a^{\rm stroke}(t) = 0\ ,\qquad a=1,\ldots,J \label{c.5} \end{equation} Our fundamental equations thus finally read: \begin{eqnarray} \varrho \frac{\partial^2 {\bf u}}{\partial t^2} & = & \mu\nabla^2 {\bf u} + (\lambda+\mu)\,\nabla(\nabla{\bf\cdot}{\bf u}) + \nonumber \\ & & \sum_{b=1}^J M_b\Omega_b^2\,\left[z_b(t)-u_b(t)\right]\, \delta^{(3)}({\bf x}-{\bf x}_b)\,{\bf n}_b + {\bf f}_{\rm external}({\bf x},t) \label{2.6.a} \\*[0.7 em] \ddot{z}_a(t) & = & -\Omega_a^2\, \left[z_a(t)-u_a(t)\right] + \xi_a^{\rm external}(t)\ , \qquad a=1,\ldots,J \label{2.6.b} \end{eqnarray} where ${\bf f}_{\rm external}({\bf x},t)$ will be given by either (\ref{2.4}) or (\ref{2.5}), as the case may be. Likewise, $\xi_a^{\rm external}(t)$ will be given by (\ref{c.4}) or (\ref{c.5}), respectively. The remainder of this paper will be concerned with finding solutions to the system of coupled differential equations (\ref{2.6.a}) and (\ref{2.6.b}), and with their meaning and consequences. \section{Green function formalism} \label{sec:gff} In order to solve equations~(\ref{2.6.a})-(\ref{2.6.b}) we shall resort to Green function formalism. The essentials of this procedure in the context of the present problem can be found in detail in reference~\cite{lobo}; more specific technicalities are given in appendix~\ref{app:a}. By means of such formalism equations~(\ref{2.6.a})-(\ref{2.6.b}) become the following integro-differential system: \begin{eqnarray} u_a(t) & = & u_a^{\rm external}(t) + \sum_{b=1}^J\,\eta_b\,\int_0^t K_{ab}(t-t')\,\left[\,z_b(t')-u_b(t')\right]\,dt' \label{3.7.a}\\[1 ex] \ddot{z}_a(t) & = & \xi_a^{\rm external}(t) -\Omega_a^2\,\left[\,z_a(t)-u_a(t)\right]\ , \qquad a=1,\ldots,J \label{3.7.b} \end{eqnarray} where $u_a^{\rm external}(t)$\,$\equiv$\, ${\bf n}_a\!\cdot\!{\bf u}^{\rm external}({\bf x}_a,t)$, and ${\bf u}^{\rm external}({\bf x},t)$ is the {\it bare\/} (i.e., without attached resonators) sphere's response to the external forces ${\bf f}_{\rm external}({\bf x},t)$ in the rhs of~(\ref{2.6.a}). $K_{ab}(t)$ is a {\it kernel matrix\/} defined by the following weighted sum of diadic products of wavefunctions\footnote{ We shall often use the capitalised index $N\/$ to imply the multiple index $\{nlm\}$ which characterises the sphere's wavefunctions.}: \begin{equation} K_{ab}(t) = \Omega_b^2\,\sum_N\,\omega_N^{-1}\, \left[{\bf n}_b\!\cdot\!{\bf u}_N^*({\bf x}_b)\right] \left[{\bf n}_a\!\cdot\!{\bf u}_N({\bf x}_a)\right]\,\sin\omega_Nt \label{3.10} \end{equation} Finally, we have defined the mass ratios of the resonators to the entire sphere \begin{equation} \eta_b\equiv \frac{M_b}{\cal M}\ ,\qquad b=1,\ldots,J \label{3.11} \end{equation} which will be {\it small parameters\/} in a real device. Before proceeding further, let us briefly pause for a qualitative inspection of equations~(\ref{3.7.a}) and~(\ref{3.7.b}). Equation (\ref{3.7.a}) shows that the sphere's surface deformations $u_a(t)$ are made up of two contributions: one due to the action of {\it external\/} agents (GWs or other), contained in $u_a^{\rm external}(t)$, and another one due to coupling to the resonators. The latter is commanded by the small parameters $\eta_b\/$, and correlates to {\it all\/} of the sphere's spheroidal eigenmodes through the kernel matrix $K_{ab}(t)$. This has consequences for GW detectors, for even though GWs may only couple to quadrupole and monopole\footnote{ Monopole modes only exist in scalar-tensor theories of gravity, such as e.g. Brans--Dicke \protect\cite{bd61}; General Relativity does not belong in this category.} spheroidal modes of the {\it free\/} sphere \cite{lobo,bian}, attachment of resonators causes, as we see, coupling between these and the other modes of the antenna, and conversely, these modes back-act on the former. As we shall shortly prove, such effects can be minimised by suitable {\it tuning\/} of the resonators' frequencies. \subsection{Laplace transform domain equations} We now take up the problem of solving these equations. Equation (\ref{3.7.a}) is an integral equation belonging in the general category of Volterra equations \cite{tricomi}, but the usual iterative solution to it by repeated substitution of $u_b(t)$ into the kernel integral is not viable here due to the {\it dynamical\/} contribution of $z_b(t)$, which is in turn governed by the {\it differential\/} equation~(\ref{3.7.b}). A better suited method to solve this {\it integro-differential\/} system is to Laplace-transform it. We denote the Laplace transform of a generic function of time $f(t)$ with a {\it caret\/} (\,$\hat{}$\,) on its symbol, e.g., \begin{equation} \hat{f}(s) \equiv \int_0^\infty f(t)\,e^{-st}\,dt \label{3.12} \end{equation} and make the assumption that the system is at rest before an instant of time, $t\/$\,=\,0, say, or \begin{equation} {\bf u}({\bf x},0)={\bf\dot u}({\bf x},0)=z_a(0)=\dot z_a(0) = 0 \label{3.14} \end{equation} Equations (\ref{3.7.a}) and (\ref{3.7.b}) then adopt the equivalent form \begin{eqnarray} \hat u_a(s) & = & \hat u_a^{\rm external}(s) - \,\sum_{b=1}^J \eta_b\,\hat K_{ab}(s)\, \left[\hat z_b(s)-\hat u_b(s)\right] \label{3.13.a} \\*[0.7 em] s^2\,\hat{z}_a(s) & = & \hat\xi_a^{\rm external}(s) - \Omega_a^2\,\left[\hat z_a(s)-\hat u_a(s)\right]\ ,\qquad a=1,\ldots,J \label{3.13.b} \end{eqnarray} for which use has been made of the {\it convolution theorem\/} for Laplace transforms\footnote{ This theorem states, it is recalled, that the Laplace transform of the convolution product of two functions is the arithmetic product of their respective Laplace transforms.}. A further simplification is accomplished if we consider that we shall in practice be only concerned with the {\it measurable\/} quantities \begin{equation} q_a(t)\equiv z_a(t)-u_a(t) \ ,\qquad a=1,\ldots,J \label{3.15} \end{equation} representing the resonators' actual elastic deformations ---cf.\ Figure \ref{fig1}. It is readily seen that these verify the following: \begin{equation} \sum_{b=1}^J \left[\delta_{ab} + \eta_b\,\frac{s^2}{s^2+\Omega_a^2}\, \hat K_{ab}(s)\right]\,\hat q_b(s) = -\frac{s^2}{s^2+\Omega_a^2}\, \hat u_a^{\rm external}(s) + \frac{\hat\xi_a^{\rm external}(s)} {s^2+\Omega_a^2}\ ,\qquad a=1,\ldots,J \label{3.16} \end{equation} Equations (\ref{3.16}) constitute a significant simplification of the original problem, as they are a set of just $J\/$ {\it algebraic\/} rather than integral or differential equations. We must solve them for the unknowns $\hat q_a(s)$, then perform {\it inverse Laplace transforms\/} to revert to $q_a(t)$. We do this next. \section{System response to a Gravitational Wave} \label{sec:srgw} Our concern now is the actual system response when it is acted upon by an incoming GW. We shall calculate it by making a number of simplifying assumptions, more precisely: \begin{enumerate} \item[\sf i)] The detector is perfectly spherical. \item[\sf ii)] The resonators have identical masses and resonance frequencies. \item[\sf iii)] The resonators' frequency is accurately matched to one of the sphere's oscillation eigenfrequencies. \end{enumerate} As we shall see below (section \ref{sec:symdef}), a real system can be appropriately treated as one which deviates by definite amounts from this idealised construct. Therefore detailed knowledge of the ideal system behaviour is essential for all purposes. This is the justification for the above simplifications. The wavefunctions ${\bf u}_{nlm}({\bf x})$ of an elastic sphere can be found in reference~\cite{lobo} in full detail, and we shall keep the notation of that paper for them. The Laplace transform of the kernel matrix~(\ref{3.10}) can thus be expressed as ---see equation (\ref{A3.20}) in appendix~\ref{app:a}: \begin{equation} \hat K_{ab}(s) = \sum_{nl}\,\frac{\Omega_b^2}{s^2+\omega_{nl}^2}\, \left|A_{nl}(R)\right|^2\,\frac{2l+1}{4\pi}\, P_l({\bf n}_a\!\cdot\!{\bf n}_b) \equiv \sum_{nl}\,\frac{\Omega_b^2}{s^2+\omega_{nl}^2}\,\chi_{ab}^{(nl)} \label{4.2} \end{equation} where the last term simply {\it defines\/} the quantities $\chi_{ab}^{(nl)}$. Note that the sums here stretch across the {\it entire\/} spectrum of the solid sphere. Our next assumption that all the resonators are {\it identical\/} simply means that \begin{equation} \eta_1=\,\ldots\,=\eta_J\equiv\eta\ ,\qquad \Omega_1=\,\ldots\,=\Omega_J\equiv\Omega \label{4.5} \end{equation} The third hypothesis makes reference to the fundamental idea behind using resonators, which is to have them tuned to one of the frequencies of the sphere's spectrum. We express this by \begin{equation} \Omega = \omega_{n_0l_0} \label{4.6} \end{equation} where $\omega_{n_0l_0}$ is a specific and {\it fixed\/} frequency of the spheroidal spectrum. In a GW detector it will only make sense to choose $l_0$\,=\,0 or $l_0$\,=\,2, as only monopole and quadrupole sphere modes couple to the incoming signal; in practice, $n_0$ will refer to the first or perhaps second harmonic \cite{clo}. We shall however keep the generic expression (\ref{4.6}) for the time being in order to encompass all the possibilities with a unified notation. Based on the above hypotheses, equation~(\ref{3.16}) can be rewritten in the form \begin{equation} \sum_{b=1}^J\,\left[\delta_{ab} + \eta\,\sum_{nl}\, \frac{\Omega^2s^2}{(s^2+\Omega^2)(s^2+\omega_{nl}^2)}\,\chi_{ab}^{(nl)} \right]\,\hat q_b(s) = -\frac{s^2}{s^2+\Omega^2}\, \hat u_a^{\rm GW}(s) + \frac{\hat\xi_a^{\rm GW}(s)} {s^2+\Omega^2}\ ,\qquad (\Omega = \omega_{n_0l_0}) \label{4.8} \end{equation} where $\hat\xi_a^{\rm GW}(s)$ is the Laplace transform of~(\ref{c.4}), i.e., \begin{equation} \hat\xi_a^{\rm GW}(s) = R\, \sum_{\stackrel{\scriptstyle l=0\ {\rm and}\ 2}{m=-l,...,l}}\, Y_{lm}({\bf n}_a)\,\hat g^{(lm)}(s)\ ,\qquad a=1,\ldots,J \label{4.85} \end{equation} As mentioned at the end of the previous section, we must now invert the matrix in the lhs of~(\ref{4.8}), which will give us an expression for $\hat q_a(s)$, then find the {\it inverse Laplace transform\/} of these functions to revert back to the time domain. A simple glance at the equation suffices however to grasp the unsurmountable difficulties of accomplishing this {\it analytically\/}. Thankfully, though, a {\it perturbative\/} approach is applicable when the masses of the resonators are small compared to the mass of the whole sphere, i.e., when the inequality \begin{equation} \eta\ll 1 \label{4.10} \end{equation} holds. We shall henceforth assume that this is the case, as also is with cylindrical bar resonant transducers. It is shown in appendix~\ref{app:b} that the perturbative series happens in ascending powers of $\eta^{1/2}$, rather than $\eta\/$ itself, and that the lowest order contribution has the form \begin{equation} \hat q_a(s) = \eta^{-1/2}\,\sum_{l,m}\,\hat\Lambda_a^{(lm)}(s;\Omega)\, \hat g^{(lm)}(s) + O(0) \ ,\qquad a=1,\ldots,J \label{6.8} \end{equation} where $O(0)$ stands for terms of order $\eta^0$ or smaller. Here, $\hat\Lambda_a^{(lm)}(s;\Omega)$ is a {\it transfer function matrix\/} which relates {\it linearly\/} the system response $\hat q_a(s)$ to the GW amplitudes $\hat g^{(lm)}(s)$, in the usual sense that $q_a(t)$ is given by the {\it convolution product\/} of the signal $g^{(lm)}(t)$ with the time domain expression, $\Lambda_a^{(lm)}(t;\Omega)$, of $\hat\Lambda_a^{(lm)}(s;\Omega)$. The detector is thus seen to act as a {\it linear filter\/} on the GW signal, whose frequency response is characterised by the properties of $\hat\Lambda_a^{(lm)}(s;\Omega)$. More specifically, the filter has a number of characteristic frequencies which correspond to the {\it imaginary parts of the poles\/} of $\hat\Lambda_a^{(lm)}(s;\Omega)$. As also shown in appendix~\ref{app:b}, these frequencies are the symmetric pairs \begin{equation} \omega_{a\pm}^2 = \Omega^2\,\left(1\pm\sqrt{\frac{2l+1}{4\pi}}\, \left|A_{n_0l_0}(R)\right|\,\zeta_a\,\eta^{1/2}\right) + O(\eta)\ , \qquad a=1,\ldots,J \label{5.2} \end{equation} where $\zeta_a^2\/$ is the $a\/$-th eigenvalue of the Legendre matrix \begin{equation} P_{l_0}({\bf n}_a\!\cdot\!{\bf n}_b)\ ,\qquad a,b=1,\,\ldots,J \label{5.25} \end{equation} associated to the multipole ($l_0$) selected for tuning ---see (\ref{4.6}). These frequency pairs correspond to {\it beats\/}, typical of resonantly coupled oscillating systems ---we shall find them again in section~\ref{sec:hs} in a particularly illuminating example. Another very important fact is also neatly displayed by equation~(\ref{6.8}): the resonators' motions are {\it mechanically amplified\/} by a factor $\eta^{-1/2}$ relative to the driving amplitudes $\hat g^{(lm)}(s)$. This is the counterpart, in our multimode system, of a similar behaviour known to happen in monomode cylindrical antennas \cite{pia}. The specific form of the transfer function matrix $\hat\Lambda_a^{(lm)}(s;\Omega)$ depends on both the selected mode to tune the resonator frequency $\Omega$, and on the resonator distribution geometry. We now come to a discussion of these. \subsection{Monopole gravitational radiation sensing} General Relativity, as is well known, forbids monopole GW radiation. More general {\it metric\/} theories, e.g. Brans-Dicke \cite{bd61}, do however predict this kind of radiation. It appears that a spherical antenna is potentially sensitive to monopole waves, so it can serve the purpose of thresholding, or eventually detecting them. This clearly requires that the resonator set be tuned to a monopole harmonic of the sphere, i.e., \begin{equation} \Omega = \omega_{n0}\ ,\qquad (l_0=0) \label{6.9} \end{equation} where $n\/$ tags the chosen harmonic ---most likely the first ($n\/$\,=\,1) in a thinkable device. Since $P_0(z)$\,$\equiv$\,1 (for all $z\/$) the eigenvalues of $P_0({\bf n}_a\!\cdot\!{\bf n}_b)$ are, clearly, \begin{equation} \zeta_1^2=J\ ,\qquad \zeta_2^2=\,\ldots\,=\zeta_J^2=0 \label{6.10} \end{equation} for {\it any resonator distribution\/}. The tuned mode frequency thus splits into a {\it single\/} strongly coupled pair: \begin{equation} \omega_\pm^2 = \Omega^2\,\left(1\pm\sqrt{\frac{J}{4\pi}}\, \left|A_{n0}(R)\right|\,\eta^{1/2}\right) + O(\eta)\ , \qquad \Omega=\omega_{n0} \label{6.11} \end{equation} The $\Lambda$-matrix of equation (\ref{6.8}) is seen to be in this case \begin{equation} \hat\Lambda_a^{(lm)}(s;\omega_{n0}) = (-1)^J\,\frac{a_{n0}}{\sqrt{J}}\, \frac{1}{2}\,\left[\left(s^2+\omega_+^2\right)^{-1} - \left(s^2+\omega_-^2\right)^{-1}\right]\,\delta_{l0}\,\delta_{m0} \label{6.12} \end{equation} whence the system response is \begin{equation} \hat q_a(s) = \eta^{-1/2}\,\frac{(-1)^J}{\sqrt{J}}\,a_{n0}\, \frac{1}{2}\,\left[\left(s^2+\omega_+^2\right)^{-1} - \left(s^2+\omega_-^2\right)^{-1}\right]\,\hat g^{(00)}(s) + O(0)\ , \qquad a=1,\ldots,J \label{6.13} \end{equation} {\it regardless of resonator positions\/}. The overlap coefficient $a_{n0}$ is calculated by means of formulas given in \cite{lobo}, and has dimensions of length. By way of example, $a_{10}/R\/$\,=\,0.214, and $a_{20}/R\/$\,=\,$-$0.038 for the first two harmonics. A few interesting facts are displayed by equation (\ref{6.13}). First, as we have already stressed, it is seen that if the resonators are tuned to a monopole {\it detector\/} frequency then only monopole {\it wave amplitudes\/} couple strongly to the system, even if quadrupole radiation amplitudes are significantly high at the observation frequencies $\omega_\pm\/$. Also, the amplitudes $\hat q_a(s)$ are equal for all $a\/$, as corresponds to the spherical symmetry of monopole sphere's oscillations, and are proportional to $J^{-1/2}$, a factor we should indeed expect as an indication that GW {\it energy\/} is evenly distributed amongst all the resonators. A {\it single\/} transducer suffices to experimentally determine the only monopole GW amplitude $\hat g^{(00)}(s)$, of course, but (\ref{6.13}) provides the system response if more than one sensor is mounted on the antenna for whatever reasons. \subsection{Quadrupole gravitational radiation sensing} We now consider the more interesting case of quadrupole motion sensing. We thus take \begin{equation} \Omega = \omega_{n2}\ ,\qquad (l_0=2) \label{6.14} \end{equation} where $n\/$ labels the chosen harmonic ---most likely the first ($n\/$\,=\,1) or the second ($n\/$\,=\,2) in a practical system. The evaluation of the $\Lambda$-matrix is now considerably more involved \cite{serrano}, yet a remarkably elegant form is found for it: \begin{equation} \hat\Lambda_a^{(lm)}(s;\omega_{n2}) = (-1)^N\,\sqrt{\frac{4\pi}{5}}\, a_{n2}\,\sum_{b=1}^J\,\left\{\sum_{\zeta_c\neq 0}\,\frac{1}{2}\left[ \left(s^2+\omega_{c+}^2\right)^{-1} - \left(s^2+\omega_{c-}^2\right)^{-1} \right]\,\frac{v_a^{(c)}v_b^{(c)*}}{\zeta_c}\right\}\, Y_{2m}({\bf n}_b)\,\delta_{l2} \label{6.15} \end{equation} where $v_a^{(c)}$ is the $c\/$-th normalised eigenvector of $P_2({\bf n}_a\!\cdot\!{\bf n}_b)$, associated to the {\it non-null\/} eigenvalue $\zeta_c^2$. Let us stress that equation (\ref{6.15}) explicitly shows that at most 5 pairs of modes, of frequencies $\omega_{c\pm}$, couple strongly to quadrupole GW amplitudes, {\it no matter how many resonators in excess of 5 are mounted on the sphere\/}. The tidal overlap coefficients $a_{2n}\/$ can also be calculated, and give for the first two harmonics \cite{ls} \begin{equation} \frac{a_{12}}{R} = 0.328\ ,\qquad\frac{a_{22}}{R} = 0.106 \label{6.16} \end{equation} The system response is thus \begin{eqnarray} \hat q_a(s) & = & \eta^{-1/2}\,(-1)^J\,\sqrt{\frac{4\pi}{5}}\,a_{n2}\, \sum_{b=1}^J\,\left\{\sum_{\zeta_c\neq 0}\,\frac{1}{2}\left[ \left(s^2+\omega_{c+}^2\right)^{-1} - \left(s^2+\omega_{c-}^2\right)^{-1} \right]\,\frac{v_a^{(c)}v_b^{(c)*}}{\zeta_c}\right\}\times \nonumber \\[0.5 em] & & \hspace*{4 cm} \times\sum_{m=-2}^2\,Y_{2m}({\bf n}_b)\,\hat g^{(2m)}(s) + O(0)\ , \qquad a=1\,\ldots,J \label{6.17} \end{eqnarray} Equation (\ref{6.17}) is {\it completely general\/}, i.e., it is valid for any resonator configuration over the sphere's surface, and for any number of resonators. It describes precisely how all 5 GW amplitudes $\hat g^{(2m)}(s)$ interact with all 5 strongly coupled system modes; like before, {\it only quadrupole wave amplitudes\/} are seen in the detector (to leading order) when $\Omega$\,=\,$\omega_{n2}$, even if the incoming wave carries significant monopole energy at the frequencies $\omega_{c\pm}$. The degree of generality and algebraic simplicity of (\ref{6.17}) is new in the literature. As we shall now see, it makes possible a systematic search for different resonator distributions and their properties. \section{The {\sl PHC\/} configuration} \label{sec:PHC} Merkowitz and Johnson's {\sl TIGA\/} \cite{jm93} is highly symmetric, and is the minimal set with maximum degeneracy, i.e., all the non-null eigenvalues $\zeta_a\/$ are equal. To accomplish this, however, 6 rather than 5 resonators are required on the sphere's surface. Since there are just 5 quadrupole GW amplitudes one may wonder whether there are alternative layouts with {\it only\/} 5 resonators. Equation (\ref{6.17}) is completely general, so it can be searched for an answer to this question. In reference \cite{ls} we made a specific proposal, which we now describe in more detail. In pursuing a search for 5 resonator sets we found that distributions having a sphere diameter as an axis of {\it pentagonal symmetry\/}\footnote{ By this we mean resonators are placed along a {\it parallel\/} of the sphere every 72$^\circ$.} exhibit a rather appealing structure. More specifically, let the resonators be located at the spherical positions \begin{equation} \theta_a = \alpha \qquad ({\rm all}\,\ a)\ ,\qquad \varphi_a = (a-1)\,\frac{2\pi}{5}\ ,\qquad a=1,\ldots,5 \end{equation} The eigenvalues and eigenvectors of $P_2({\bf n}_a\!\cdot\!{\bf n}_b)$ are easily calculated: \begin{eqnarray} & \zeta_0^2 = \frac{5}{4}\,\left(3\,\cos^2\alpha-1\right)^2\ ,\qquad \zeta_1^2 = \zeta_{-1}^2 = \frac{15}{2}\,\sin^2\alpha\,\cos^2\alpha \ ,\qquad\zeta_2^2 = \zeta_{-2}^2 = \frac{15}{8}\,\sin^4\alpha & \label{6.24.a} \\[1 em] & v_a^{(m)} = \sqrt{\frac{4\pi}{5}}\,\zeta_m^{-1}\,Y_{2m}({\bf n}_a)\ , \qquad m=-2,\ldots,2\ ,\ \ a=1,\ldots,5 & \label{6.24.b} \end{eqnarray} so the $\Lambda$-matrix is also considerably simple in structure in this case: \begin{equation} \hat\Lambda_a^{(lm)}(s;\omega_{n2}) = -\sqrt{\frac{4\pi}{5}}\,a_{n2}\, \zeta_m^{-1}\,\frac{1}{2}\left[\left(s^2+\omega_{m+}^2\right)^{-1} - \left(s^2+\omega_{m-}^2\right)^{-1}\right]\,Y_{2m}({\bf n}_a)\,\delta_{l2} \qquad ({\sl PHC}) \label{6.25} \end{equation} where we have used the notation \begin{equation} \omega_{m\pm}^2 = \Omega^2\,\left(1\pm\sqrt{\frac{5}{4\pi}}\, \left|A_{n2}(R)\right|\,\zeta_m\,\eta^{1/2}\right) + O(\eta)\ , \qquad m=-2,\ldots,2 \label{6.26} \end{equation} As we see from these formulas, the {\it five\/} expected pairs of frequencies actually reduce to {\it three\/}, so pentagonal distributions keep a certain degree of degeneracy, too. The most important distinguishing characteristic of the general {\it pentagonal\/} layout is best displayed by the explicit system response: \begin{eqnarray} & \hat q_a(s) = -\eta^{-1/2}\,\sqrt\frac{4\pi}{5}\,a_{n2} & \left\{\, \frac{1}{2\zeta_0}\left[ \left(s^2+\omega_{0+}^2\right)^{-1} - \left(s^2+\omega_{0-}^2\right)^{-1} \right]\,Y_{20}({\bf n}_a)\,\hat g^{(20)}(s)\right. \nonumber \\ & & + \;\frac{1}{2\zeta_1}\left[ \left(s^2+\omega_{1+}^2\right)^{-1} - \left(s^2+\omega_{1-}^2\right)^{-1} \right]\,\left[ Y_{21}({\bf n}_a)\,\hat g^{(11)}(s) + Y_{2-1}({\bf n}_a)\,\hat g^{(1\,-1)}(s)\right] \label{6.27} \\ & & + \left.\frac{1}{2\zeta_2}\left[ \left(s^2+\omega_{2+}^2\right)^{-1} - \left(s^2+\omega_{2-}^2\right)^{-1} \right]\,\left[ Y_{22}({\bf n}_a)\,\hat g^{(22)}(s) + Y_{2-2}({\bf n}_a)\,\hat g^{(2\,-2)}(s)\right]\right\} \nonumber \end{eqnarray} This equation indicates that {\it different wave amplitudes selectively couple to different detector frequencies\/}. This should be considered a very remarkable fact, for it thence follows that simple inspection of the system readout {\it spectrum\/}\footnote{ In a noiseless system, of course} immediately reveals whether a given wave amplitude $\hat g^{2m}(s)$ is present in the incoming signal or not. Pentagonal configurations also admit {\it mode channels\/}, which are easily constructed from (\ref{6.27}) thanks to the orthonormality property of the eigenvectors (\ref{6.24.b}): \begin{equation} \hat y^{(m)}(s)\equiv\sum_{a=1}^5\,v_a^{(m)*}\hat q_a(s) = \eta^{-1/2}\,a_{n2}\, \frac{1}{2}\left[\left(s^2+\omega_{m+}^2\right)^{-1} - \left(s^2+\omega_{m-}^2\right)^{-1}\right]\,\hat g^{(2m)}(s) + O(0) \ ,\qquad m=-2,\ldots,2 \label{6.28} \end{equation} These are almost identical to the {\sl TIGA\/} mode channels \cite{jm95}, the only difference being that each mode channel comes now at a {\it single specific\/} frequency pair $\omega_{m\pm}$. {\it Mode channels\/} are fundamental in signal deconvolution algorithms in noisy systems \cite{m98,lms}. Pentagonal resonator configurations should thus be considered non-trivial candidates for a real GW detector. \begin{figure} \psfig{file=fig4.ps,height=15cm,width=12cm,rheight=9.2cm,bbllx=-1.8cm,bblly=-3.8cm,bburx=18.4cm,bbury=24.7cm} \caption{The three distinct eigenvalues $\zeta_m\/$ ($m\/$\,=\,0,1,2) as functions of the distance of the resonator parallel's co-latitude $\alpha\/$ relative to the axis of symmetry of the distribution, cf. equation (\protect\ref{6.24.a}). \label{fig3}} \end{figure} \begin{figure} \psfig{file=fig5a.ps,height=11cm,width=7cm} \vspace*{-11 cm} \hspace*{9.7 cm} \psfig{file=fig5b.ps,height=11cm,width=7cm} \vspace*{-1.2 cm} \caption{To the left, the {\it pentagonal hexacontahedron\/} shape. Certain faces are marked to indicate resonator positions in a specific proposal ---see text--- as follows: a {\it square\/} for resonators tuned to the first quadrupole frequency, a {\it triangle\/} for the second, and a {\it star\/} for the monopole. On the right we see the (pentagonal) face of the polyhedron. A few details about it: the confluence point of the dotted lines at the centre is the tangency point of the {\it inscribed\/} sphere to the {\sl PHC\/}; the labeled angles have values $\alpha\/$\,=\,61.863$^\circ$, $\beta\/$\,=\,87.205$^\circ$; the angles at the $T\/$-vertices are all equal, and their value is 118.1366$^\circ$, while the angle at $P\/$ is 67.4536$^\circ$; the ratio of a long edge (e.g. $PT_1$) to a short one (e.g. $T_1T_2$) is 1.74985, and the radius of the inscribed sphere is {\it twice\/} the long edge of the pentagon, $R\/$\,=\,2\,$PT_1$. \label{fig4}} \end{figure} \begin{figure} \psfig{file=fig6.ps,height=12cm,width=15cm,rheight=9.7cm,bbllx=-1cm,bblly=-5cm,bburx=20cm,bbury=25.8cm} \caption{Compared line spectrum of a coupled {\sl TIGA\/} and a {\sl PHC\/} resonator layout in an ideally spherical system. The weakly coupled central frequency in the {\sl TIGA\/} is drawn dashed. The frequency pair is 5-fold degenerate for this layout, while the two outer pairs of the {\sl PHC\/} are doubly degenerate each, and the inner pair is non-degenerate. Units in abscissas are $\eta^{1/2}\Omega$, and the central value, labeled 0.0, corresponds to $\Omega$. \label{fig5}} \end{figure} Based on these facts one may next ask which is a suitable transducer distribution with an axis of pentagonal symmetry. In Figure \ref{fig3} we give a plot of the eigenvalues (\ref{6.24.a}) as a function of $\alpha\/$, the angular distance of the resonator set from the symmetry axis. Several criteria may be adopted to select a specific choice in view of this graph. An interesting one was proposed by us in reference \cite{ls} with the following argument. If for ease of mounting, stability, etc., it is desirable to have the detector milled into a close-to-spherical {\it polyhedric\/} shape\footnote{ This is the philosophy suggested and experimentally implemented by Merkowitz and Johnson at {\sl LSU\/}.} then polyhedra with axes of pentagonal symmetry must be searched. The number of quasi regular {\it convex\/} polyhedra is of course finite ---there actually are only 18 of them \cite{pacoM,tsvi}---, and we found a particularly appealing one in the so called {\it pentagonal hexacontahedron\/} ({\sl PHC\/}), which we see in Figure \ref{fig4}, left. This is a 60 face polyhedron, whose faces are the identical {\it irregular pentagons\/} of Figure \ref{fig4}, right. The {\sl PHC\/} admits an {\it inscribed sphere\/} which is tangent to each face at the central point marked in the Figure. It is clearly to this point that a resonator should be attached so as to simulate an as perfect as possible spherical distribution. The {\sl PHC\/} is considerably spherical: the ratio of its volume to that of the inscribed sphere is 1.057, which quite favourably compares to the value of 1.153 for the ratio of the circumscribed sphere to the TI volume. If we now request that the frequency pairs $\omega_{m\pm}$ be as {\it evenly spaced\/} as possible, compatible with the {\sl PHC\/} face orientations, then we must choose $\alpha\/$\,=\,67.617$^\circ$, whence \begin{equation} \omega_{0\pm} = \omega_{12}\,\left(1\pm 0.5756\,\eta^{1/2}\right)\ ,\ \ \ \omega_{1\pm} = \omega_{12}\,\left(1\pm 0.8787\,\eta^{1/2}\right)\ ,\ \ \ \omega_{2\pm} = \omega_{12}\,\left(1\pm 1.0668\,\eta^{1/2}\right) \label{6.29} \end{equation} for instance for $\Omega$\,=\,$\omega_{12}$, the first quadrupole harmonic. In Figure \ref{fig5} we display this frequency spectrum together with the multiply degenerate {\it TIGA\/} for comparison. The criterion leading to the {\sl PHC\/} proposal is of course not unique, and alternatives can be considered. For example, if the 5 faces of a regular icosahedron are selected for sensor mounting ($\alpha\/$\,=\,63.45$^\circ$) then a four-fold degenerate pair plus a single non-degenerate pair is obtained; if the resonator parallel is 50$^\circ$ or 22.6$^\circ$ away from the ``north pole'' then the three frequencies $\omega_{0+}$, $\omega_{1+}$, and $\omega_{2+}$ are equally spaced; etc. The number of choices is virtually infinite if the sphere is not milled into a polyhedric shape \cite{ts,grg}. Let us finally recall that the complete {\sl PHC\/} proposal \cite{ls} was made with the idea of building an as complete as possible spherical GW antenna, which amounts to making it sensitive at the first {\it two\/} quadrupole frequencies {\it and\/} at the first monopole one. This would take advantage of the good sphere cross section at the second quadrupole harmonic \cite{clo}, and would enable measuring (or thresholding) eventual monopole GW radiation. Now, the system {\it pattern matrix\/} $\hat\Lambda_a^{(lm)}(s;\Omega)$ has {\it identical structure\/} for all the harmonics of a given $l\/$ series ---see (\ref{6.12}) and (\ref{6.15})---, and so too identical criteria for resonator layout design apply to either set of transducers, respectively tuned to $\omega_{12}$ and $\omega_{22}$. The {\sl PHC\/} proposal is best described graphically in Figure \ref{fig4}, left: a {\it second\/} set of resonators, tuned to the second quadrupole harmonic $\omega_{22}$ can be placed in an equivalent position in the ``southern hemisphere'', and an eleventh resonator tuned to the first monopole frequency $\omega_{10}$ is added at an arbitrary position. It is not difficult to see, by the general methods outlined earlier on in this paper, that cross interaction between these three sets of resonators is only {\it second order\/} in $\eta^{1/2}\/$, therefore weak. A spherical GW detector with such a set of altogether 11 transducers would be a very complete multi-mode multi-frequency device with an unprecedented capacity as an individual antenna. Amongst other it would practically enable monitoring of coalescing binary {\it chirp\/} signals by means of a rather robust double passage method \cite{cf}, a prospect which was considered so far possible only with broadband long baseline laser interferometers \cite{klm1,klm2}, and is almost unthinkable with currently operating cylindrical bars. \section{A calibration signal: hammer stroke} \label{sec:hs} This section is a brief digression from the main streamline of the paper. We propose to assess now the system response to a particular, but useful, calibration signal: a perpendicular {\it hammer stroke\/}. We first go back to equation (\ref{3.16}) and replace $\hat u_a^{\rm external}(s)$ in its rhs with that corresponding to a hammer stroke, which is easily calculated ---cf.\ appendix~\ref{app:a}: \begin{equation} \hat u_a^{\rm stroke}(s) = -\sum_{nl}\,\frac{f_0}{s^2+\omega_{nl}^2}\, \left|A_{nl}(R)\right|^2\,P_l({\bf n}_a\!\cdot\!{\bf n}_0)\ ,\qquad a=1,\ldots,J \label{7.1} \end{equation} where ${\bf n}_0$ are the spherical coordinates of the hit point on the sphere, and $f_0$\,$\equiv$\,${\bf n}_0\!\cdot\!{\bf f}_0/{\cal M\/}$. Clearly, the hammer stroke excites {\it all\/} of the sphere's vibration eigenmodes, as it has a completly flat spectrum. The coupled system resonances are again those calculated in appendix~\ref{app:b}. The same procedures described in section~\ref{sec:srgw} for a GW excitation can now be pursued to obtain \begin{eqnarray} \hat q_a(s) & = & \eta^{-1/2}\,(-1)^{J-1}\,\sqrt{\frac{2l+1}{4\pi}} \,f_0\,\left|A_{nl}(R)\right|\,\times \nonumber \\ & \times & \sum_{b=1}^J\,\left\{\sum_{\zeta_c\neq 0}\,\frac{1}{2}\left[ \left(s^2+\omega_{c+}^2\right)^{-1}-\left(s^2+\omega_{c-}^2\right)^{-1} \right]\,\frac{v_a^{(c)}v_b^{(c)*}}{\zeta_c}\right\}\, P_l({\bf n}_b\!\cdot\!{\bf n}_0) + O(0)\ , \qquad a=1\,\ldots,J \label{7.2} \end{eqnarray} when the system is tuned to the $nl\/$-th spheroidal harmonic, i.e., $\Omega$\,=\,$\omega_{nl}$. It is immediately seen from here that the system response to this signal when the resonators are tuned to a {\it monopole\/} frequency is given by \begin{equation} \hat q_a(s) = \eta^{-1/2}\,(-1)^{J-1}\,\frac{f_0}{\sqrt{4\pi J}}\, \left|A_{n0}(R)\right|\,\frac{1}{2}\left[\left( s^2+\omega_+^2\right)^{-1}-\left(s^2+\omega_-^2\right)^{-1}\right] \ ,\qquad \Omega=\omega_{n0} \label{7.3} \end{equation} an expression which holds for all $a\/$, and is independent of either the resonator layout or the hit point, which in particular prevents any determination of the latter, as obviously expected. The frequencies $\omega_\pm$ are those of (\ref{6.11}), and we find here again a global factor $J^{-1/2}$, as also expected. We consider next the situation when quadrupole tuning is implemented, $\Omega$\,=\,$\omega_{n2}$. We shall however do so only for the {\sl PHC\/} and {\sl TIGA\/} configurations, as more general considerations are not quite as interesting at this point. \begin{figure} \psfig{file=fig8.ps,height=22cm,width=15cm,rheight=23cm,bbllx=1.1cm,bblly=3.3cm,bburx=17.5cm,bbury=27cm} \caption{Simulated response of a {\sl PHC\/} to a hammer stroke: the time series and their respective spectra, both for direct resonator readouts and mode channels. Note that while the former are {\it not\/} simple beats, the latter are. \label{fig7}} \end{figure} \subsection{{\sl PHC\/} and {\sl TIGA\/} response to a hammer stroke} Expanding equation \ref{7.2} by substitution of the eigenvalues $\zeta_m\/$ and eigenvectors $v^{(m)}_a\/$ of the {\sl PHC\/}, one readily finds that the system response is given by \begin{equation} \hat q_a(s) = \eta^{-1/2}\,f_0\,\sqrt{\frac{4\pi}{5}}\, \left|A_{n2}(R)\right|\,\sum_{m=-2}^2\,\frac{1}{2}\left[\left( s^2+\omega_{m+}^2\right)^{-1}-\left(s^2+\omega_{m-}^2\right)^{-1}\right] \,\zeta_m^{-1}\,Y_{2m}({\bf n}_a)\,Y_{2m}^*({\bf n}_0) \ ,\qquad a=1,\ldots,5 \label{7.7} \end{equation} and the mode channels by \begin{equation} \hat y^{(m)}(s) = \eta^{-1/2}\,f_0\,\left|A_{n2}(R)\right| \frac{1}{2}\left[\left(s^2+\omega_{m+}^2\right)^{-1} - \left(s^2+\omega_{m-}^2\right)^{-1}\right]\,Y_{2m}^*({\bf n}_0) \ ,\ \ m=-2,\ldots,2 \label{7.8} \end{equation} These equations indicate that the system response $q_a(t)$ is a {\it superposition of three different beats\/}\footnote{ A {\it beat\/} is a modulated oscillation of the form $\sin\frac{1}{2}(\omega_+-\omega_-)t\,\cos\Omega t$, where $\omega_+$ and $\omega_-$ are nearby frequencies, and $\omega_+$\,$+$\,$\omega_-$\,=\,2\,$\Omega$. The Laplace transform of such function of time is precisely $(\Omega/2)$\,$\left[\left( s^2+\omega_+^2\right)^{-1}-\left(s^2+\omega_-^2\right)^{-1}\right]$, up to higher order terms in the difference $\omega_+$\,$-$\,$\omega_-$, which in our case is proportional to $\eta^{1/2}$.}, while the mode channels are {\it single\/} beats each, but with {\it differing modulation frequencies\/}. This is represented graphically in Figure \ref{fig7}, where we see the result of a numerical simulation of the {\sl PHC\/} response to a hammer stroke, delivered to the solid at a given location. The readouts $q_a(t)$ are somewhat complex time series, whose frequency spectrum shows {\it three pairs of peaks\/} ---in fact, the {\it lines\/} in the ideal spectrum of Figure \ref{fig5}. The mode channels on the other hand are {\it pure beats\/}, whose spectra consist of the {\it individually separate\/} pairs of the just mentioned peaks. The response of the {\sl TIGA\/} layout to a hammer stroke has been described in detail by Merkowitz and Johnson ---see e.g.\ reference \cite{jm97}. Our formalism does of course recover the results obtained by those authors; in the notation of this paper, we have \begin{eqnarray} \hat q_a(s) & = & -\eta^{-1/2}\,\frac{5}{\sqrt{24\pi}}\, f_0\,\left|A_{n2}(R)\right|\,\frac{1}{2}\left[ \left(s^2+\omega_+^2\right)^{-1}-\left(s^2+\omega_-^2\right)^{-1}\right] \,P_2({\bf n}_a\!\cdot\!{\bf n}_0)\ ,\qquad {\sl TIGA} \label{7.4.a} \\ \hat y^{(m)}(s) & = & -\eta^{-1/2}\,f_0\,\left|A_{n2}(R)\right| \frac{1}{2}\left[\left(s^2+\omega_+^2\right)^{-1} - \left(s^2+\omega_-^2\right)^{-1}\right]\,Y_{2m}^*({\bf n}_0) \ ,\ \ m=-2,\ldots,2 \label{7.5} \end{eqnarray} for the system response and the mode channels, respectively, where \begin{equation} \omega_\pm^2 = \omega_{n2}^2\,\left(1\pm\sqrt{\frac{3}{2\pi}}\, \left|A_{n2}(R)\right|\eta^{1/2}\right) + O(\eta)\ , \qquad a=1,\ldots,6 \label{6.19} \end{equation} are the five-fold degenerate frequency pairs corresponding to the {\sl TIGA\/} distribution. Comparison of the mode channels shows that they are identical for {\sl PHC\/} and {\sl TIGA\/}, except that the former come at different frequencies depending on the index $m\/$. One might perhaps say that the {\sl PHC\/} gives rise to a sort of ``Zeeman splitting'' of the {\sl TIGA\/} degenerate frequencies, which can be attributed to an {\it axial symmetry breaking\/} of that resonator distribution: the {\sl PHC\/} mode channels partly split up the otherwise degenerate multiplet into its components. \section{Symmetry defects} \label{sec:symdef} So far we have made the assumption that the sphere is perfectly symmetric, that the resonators are identical, that their locations on the sphere's surface are ideally accurate, etc. This is of course unrealistic. So we propose to address now how such departures from ideality affect the system behaviour. As we shall see, the system is rather {\it robust\/}, in a sense to be made precise shortly, against a number of small defects. In order to {\it quantitatively\/} assess ideality failures we shall adopt a philosophy which is naturally suggested by the results already obtained in an ideal system. It is as follows. As we have seen in previous sections, the solution to the general equations (\ref{3.16}) must be given as a {\it perturbative\/} series expansion in ascending powers of the small quantity $\eta^{1/2}$. This is clearly a fact {\it not\/} related to the system's symmetries, so it will survive symmetry breakings. It is therefore appropriate to {\it parametrize\/} deviations from ideality in terms of suitable powers of $\eta^{1/2}$, in order to address them {\it consistently with the order of accuracy of the series solution to the equations of motion\/}. An example will better illustrate the situation. In a {\it perfectly ideal\/} spherical detector the system frequencies are given by equations (\ref{5.2}). Now, if a small departure from e.g. spherical symmetry is present in the system then we expect that a correspondingly small correction to those equations will be required. Which specific correction to the formula will actually happen can be {\it qualitatively\/} assessed by a {\it consistency\/} argument: if symmetry defects are of order $\eta^{1/2}$ then equations (\ref{5.2}) will be significantly altered in their $\eta^{1/2}$ terms; if on the other hand such defects are of order $\eta\/$ or smaller then any modifications to equations (\ref{5.2}) will be swallowed into the $O(0)$ terms, and the more important $\eta^{1/2}$ terms will remain unaffected by the symmetry failure. We will say in the latter case that the system is {\it robust\/} to that ideality breaking. More generally, this argument can be extended to see that the only system defects standing a chance to have any influences on lowest order ideal system behaviour are defects of order $\eta^{1/2}$ relative to an ideal configuration. Defects of such order are however {\it not necessarily guaranteed\/} to be significant, and a specific analysis is required for each specific parameter in order to see whether or not the system response is {\it robust\/} against the considered parameter deviations. We therefore proceed as follows. Let $P\/$ be one of the system parameters, e.g. a sphere frequency, or a resonator mass or location, etc. Let $P_{\rm ideal}$ be the {\it numerical value\/} this parameter has in an ideal detector, and let $P_{\rm real}$ be its value in the real case. These two will be assumed to differ by terms of order $\eta^{1/2}$, or \begin{equation} P_{\rm real} = P_{\rm ideal}\,(1+p\,\eta^{1/2}) \label{8.1} \end{equation} For a given system, $p\/$ is readily determined adopting (\ref{8.1}) as the {\it definition\/} of $P_{\rm real}$, once a suitable {\it hypothesis\/} has been made as to which is the value of $P_{\rm ideal}$. In order for the following procedure to make sensible sense it is clearly required that $p\/$ be of order 1 or, at least, appreciably larger than $\eta^{1/2}$. Should $p\/$ thus calculated from (\ref{8.1}) happen to be too small, i.e., of order $\eta^{1/2}$ itself or smaller, then the system will be considered {\it robust\/} as regards the affected parameter. We now apply this criterion to various departures from ideality. \subsection{The suspended sphere \label{s8.1}} An earth based observatory obviously requires a {\it suspension mechanism\/} for the large sphere. If a {\it nodal point\/} suspension is e.g.\ selected then a diametral {\it bore\/} has to be drilled across the sphere \cite{phd}. The most immediate consequence of this is that spherical symmetry is broken, what in turn results in {\it degeneracy lifting\/} of the free spectral frequencies $\omega_{nl}\/$, which now {\it split\/} up into multiplets $\omega_{nlm}\/$ ($m\/$\,=\,$-l\/$,...,$l\/$). The resonators' frequency $\Omega$ {\it cannot\/} therefore be matched to {\it the\/} frequency $\omega_{n_0l_0}$, but at most to {\it one\/} of the $\omega_{n_0l_0m}\/$'s. In this subsection we keep the hypothesis that all the resonators are identical ---we shall relax it later---, and assume that $\Omega$ falls {\it within\/} the span of the multiplet of the $\omega_{n_0l_0m}\/$'s. Then we write \begin{equation} \omega_{n_0l_0m}^2 = \Omega^2\,(1+p_m\,\eta^{1/2})\ ,\qquad m=-l_0,\ldots,l_0 \label{8.2} \end{equation} We now search for the coupled frequencies, i.e., the roots of equation (\ref{3.18}). The kernel matrix $\hat K_{ab}(s)$ is however no longer given by (\ref{4.2}), due the removed degeneracy of $\omega_{nl}\/$, and we must stick to its general expression (\ref{3.17}), or \begin{equation} \hat K_{ab}(s) = \sum_{nlm}\,\frac{\Omega_b^2}{s^2+\omega_{nlm}^2}\, \left|A_{nl}(R)\right|^2\,\frac{2l+1}{4\pi}\, Y_{lm}^*({\bf n}_a)\,Y_{lm}({\bf n}_b) \equiv \sum_{nlm}\,\frac{\Omega_b^2}{s^2+\omega_{nlm}^2}\,\chi_{ab}^{(nlm)} \label{8.3} \end{equation} Following the steps of appendix \ref{app:a} we now seek the roots of the equation \begin{equation} \det\,\left[\delta_{ab} + \eta\,\sum_{m=-l_0}^{l_0}\, \frac{\Omega^2s^2}{(s^2+\Omega^2)(s^2+\omega_{n_0l_0m}^2)} \,\chi_{ab}^{(n_0l_0m)} + \eta\,\sum_{nl\neq n_0l_0,m}\, \frac{\Omega^2s^2}{(s^2+\Omega^2)(s^2+\omega_{nlm}^2)}\, \chi_{ab}^{(nlm)}\right] = 0 \label{8.4} \end{equation} Since $\Omega$ relates to $\omega_{n_0l_0m}\/$ through equation (\ref{8.2}) we see that the roots of (\ref{8.4}) fall again into either of the two categories (\ref{4.11.a})-(\ref{4.11.b})(see Appendix~\ref{app:b}), i.e., roots close to $\pm i\Omega$ and roots close to $\pm i\omega_{nlm}\/$ ($nl\/$\,$\neq$\,$n_0l_0$). We shall exclusively concentrate on the former now. Direct substitution of the series (\ref{4.11.a}) into (\ref{8.4}) yields the following equation for the coefficient $\chi_{\frac{1}{2}}$: \begin{equation} \det\left[\delta_{ab} - \frac{1}{\chi_\frac{1}{2}}\,\sum_{m=-l_0}^{l_0} \,\frac{\chi_{ab}^{(n_0l_0m)}}{\chi_\frac{1}{2}-p_m}\right] = 0 \label{8.5} \end{equation} This is a variation of (\ref{5.1}), to which it reduces when $p_m\/$\,=\,0, i.e., when there is full degeneracy. The solutions to (\ref{8.5}) no longer come in symmetric pairs, like (\ref{5.2}). Rather, there are 2$l_0$+1+$J\/$ of them, with a {\it maximum\/} number of 2(2$l_0$+1) non-identically zero roots if $J\/$\,$\geq$\,2$l_0$+1\footnote{ This is a {\it mathematical fact\/}, whose proof is relatively cumbersome, and will be omitted here; we just mention that it has its origin in the linear dependence of more than 2$l_0$+1 spherical harmonics of order $l_0$.}. For example, if we choose to select the resonators' frequency close to a quadrupole multiplet ($l_0$\,=\,2) then (\ref{8.5}) has at most 5+$J\/$ non-null roots, {\it with a maximum ten\/} no matter how many resonators in excess of 5 we attach to the sphere. Modes associated to null roots of (\ref{8.5}) can be seen to be {\it weakly coupled\/}, just like in a free sphere, i.e., their amplitudes are smaller than those of the strongly coupled ones by factors of order $\eta^{1/2}$. In order to assess the reliability of this method we have applied it to see what are its predictions for a {\it real system\/}. To this end, data taken with the {\sl TIGA\/} prototype at {\sl LSU\/}\footnote{ These data are contained in reference \protect\cite{phd}, and we want to express our gratitude to Stephen Merkowitz for kindly handing them to us.} were used to confront with. The {\sl TIGA\/} was drilled and suspended from its centre, so its first quadrupole frequency split up into a multiplet of five frequencies. Their reportedly measured values are \begin{equation} \omega_{120} = 3249\ {\rm Hz}\ ,\ \ \omega_{121} = 3238\ {\rm Hz}\ ,\ \ \omega_{12\,-1} = 3236\ {\rm Hz}\ ,\ \ \omega_{122} = 3224\ {\rm Hz}\ ,\ \ \omega_{12\,-2} = 3223\ {\rm Hz}\ ,\ \ \label{8.6} \end{equation} All 6 resonators were equal, and had the following characteristic frequency and mass, respectively: \begin{equation} \Omega = 3241\ {\rm Hz}\ ,\qquad\eta = \frac{1}{1762.45} \label{8.7} \end{equation} Substituting these values into (\ref{8.2}) it is seen that \begin{equation} p_0=0.2075\ ,\ \ p_1=-0.0777\ ,\ \ p_{-1}=-0.1036\ ,\ \ p_2=-0.4393\ ,\ \ p_{-2}=-0.4650 \label{8.8} \end{equation} \begin{table} \label{t1} \caption{Numerical values of measured and theoretically predicted frequencies (in Hz) for the {\sl TIGA\/} prototype with varying number of resonators. Percent differences are also shown. The {\it calculated\/} values of the tuning and free multiplet frequencies are taken {\it by definition\/} equal to the measured ones, and quoted in brackets. In square brackets the frequency of the {\it weakly coupled\/} sixth mode in the full, 6~resonator {\sl TIGA\/} layout. These data are plotted in Figure \protect\ref{fig8}.} \begin{center} \begin{tabular}{lccc||lccc} Item & Measured & Calculated & \% difference & Item & Measured & Calculated & \% difference \\ \hline Tuning & 3241 & (3241) & (0.00) & 4 resonators & 3159 & 3155 & $-0.12$ \\ Free multiplet & 3223 & (3223) & (0.00) & & 3160 & 3156 & $-0.11$ \\ & 3224 & (3224) & (0.00) & & 3168 & 3165 & $-0.12$ \\ & 3236 & (3236) & (0.00) & & 3199 & 3198 & $-0.05$ \\ & 3238 & (3238) & (0.00) & & 3236 & 3236 & $ 0.00$ \\ & 3249 & (3249) & (0.00) & & 3285 & 3286 & $ 0.03$ \\ 1 resonator & 3167 & 3164 & $-0.08$ & & 3310 & 3310 & $ 0.00$ \\ & 3223 & 3223 & $ 0.00$ & & 3311 & 3311 & $ 0.00$ \\ & 3236 & 3235 & $-0.02$ & & 3319 & 3319 & $ 0.00$ \\ & 3238 & 3237 & $-0.02$ & 5 resonators & 3152 & 3154 & $ 0.08$ \\ & 3245 & 3245 & $ 0.00$ & & 3160 & 3156 & $-0.14$ \\ & 3305 & 3307 & $ 0.06$ & & 3163 & 3162 & $-0.03$ \\ 2 resonators & 3160 & 3156 & $-0.13$ & & 3169 & 3167 & $-0.08$ \\ & 3177 & 3175 & $-0.07$ & & 3209 & 3208 & $-0.02$ \\ & 3233 & 3233 & $ 0.00$ & & 3268 & 3271 & $ 0.10$ \\ & 3236 & 3236 & $ 0.00$ & & 3304 & 3310 & $ 0.17$ \\ & 3240 & 3240 & $ 0.00$ & & 3310 & 3311 & $ 0.03$ \\ & 3302 & 3303 & $ 0.03$ & & 3313 & 3316 & $ 0.10$ \\ & 3311 & 3311 & $ 0.00$ & & 3319 & 3321 & $ 0.06$ \\ 3 resonators & 3160 & 3155 & $-0.15$ & 6 resonators & 3151 & 3154 & $ 0.11$ \\ & 3160 & 3156 & $-0.13$ & & 3156 & 3155 & $-0.03$ \\ & 3191 & 3190 & $-0.02$ & & 3162 & 3162 & $ 0.00$ \\ & 3236 & 3235 & $-0.02$ & & 3167 & 3162 & $-0.14$ \\ & 3236 & 3236 & $ 0.00$ & & 3170 & 3168 & $-0.07$ \\ & 3297 & 3299 & $ 0.08$ & & [3239] & [3241] & [0.06] \\ & 3310 & 3311 & $ 0.02$ & & 3302 & 3309 & $ 0.23$ \\ & 3311 & 3311 & $ 0.00$ & & 3308 & 3310 & $ 0.06$ \\ & & & & & 3312 & 3316 & $ 0.12$ \\ & & & & & 3316 & 3317 & $ 0.02$ \\ & & & & & 3319 & 3322 & $ 0.10$ \end{tabular} \end{center} \end{table} Equation (\ref{8.5}) can now be readily solved once the resonator positions are fed into the matrices $\chi_{ab}^{(12m)}$. Such positions correspond to the pentagonal faces of a truncated icosahedron. Merkowitz \cite{phd} gives a complete account of all the measured system frequencies as resonators are progressively attached to the selected faces, beginning with one and ending with six. In Figure \ref{fig8} we present a graphical display of the experimentally reported frequencies along with those calculated theoretically by solving equation (\ref{8.5}). In Table 1 we give the numerical values. As can be seen, coincidence between our theoretical predictions and the experimental data is remarkable: the worst error is 0.2\%, while for the most part it is below 0.1\%. Thus {\it discrepancies between our theoretical predictions and experiment are of order $\eta$\/}, as indeed expected ---see (\ref{8.7}). In addition, it is also reported in reference \cite{jm97} that the 11-th, weakly coupled mode of the {\sl TIGA\/} (enclosed in square brackets in Table 1) has a practically zero amplitude, again in excellent agreement with our general theoretical predictions about modes beyond the tenth ---see paragraph after equation~(\ref{8.5}). \begin{figure} \psfig{file=fig9.ps,height=22cm,width=15cm,rheight=22cm \caption{The frequency spectrum of the {\sl TIGA\/} distribution as resonators are progressively added from none to 6. Continuous lines correspond to measured values, and dashed lines correspond to their $\eta^{1/2}$ theoretical estimates with equation~(\protect\ref{8.5}). \label{fig8}} \end{figure} This is an encouraging result which motivated us to try a better fit by estimating the {\it next order\/} corrections, i.e., $\chi_1$ of (\ref{4.11.a}). As it turned out, however, matching between theory and experiment did not consistently improve. This is not really that surprising, though, as M\&J explicitly state \cite{jm97} that control of the general experimental conditions in which data were obtained had a certain degree of tolerance, and they actually show satisfaction that $\sim$1\% coincidence between theory and measurement is comfortably accomplished. But 1\% is {\it two orders of magnitude larger than $\eta\/$} ---cf. equation (\ref{8.7})---, so failure to refine our frequency estimates to order $\eta\/$ is again fully consistent with the accuracy of available real data. A word on a technical issue is in order. Merkowitz and Johnson's equations for the {\sl TIGA\/} \cite{jm93,jm95} are identical to ours to lowest order in $\eta\/$. Remarkably, though, their reported theoretical estimates of the system frequencies are not quite as accurate as ours \cite{jm97}. The reason for this is probably the following: in M\&J's model these frequencies appear within an algebraic system of 5+$J\/$ linear equations with as many unknowns which has to be solved; in our model the algebraic system has only $J\/$ equations and unknowns, actually equations (\ref{3.16}). This is a very appreciable difference for the range of values of $J\/$ under consideration. While the roots for the frequencies can be seen to {\it mathematically\/} coincide in both approaches, in actual practice these roots are {\it estimated\/}, generally by means of computer programmes. It is here that problems most likely arise, for the numerical reliability of an algorithm to solve matrix equations normally decreases as the rank of the matrix increases. The significant algebraic simplification of our model's equations should therefore be considered one of real practical value. \subsection{Other mismatched parameters} We now assess the system sensitivity to small mismatches in resonators' masses, locations and frequencies. \subsubsection{Resonator mass mismatches} If the {\it masses\/} are slightly non-equal then we can write \begin{equation} M_a = \eta{\cal M}\,(1+\mu_a\,\eta^{1/2})\ ,\qquad a=1,\ldots,J \label{8.9} \end{equation} where $\eta\/$ can be defined e.g. as the ratio of the {\it average\/} resonator mass to the sphere's mass. It is immediately obvious from equation (\ref{8.9}) that mass non-uniformities of the resonators only affect our equations in {\it second order\/}, since resonator mass non-uniformities result, as we see, in corrections of order $\eta^{1/2}$ to $\eta^{1/2}$ itself, which is the very parameter of the perturbative expansions. The system is thus clearly {\it robust\/} to mismatches in the resonator masses of the type (\ref{8.9}). \subsubsection{Errors in resonator locations} The same happens if the {\it locations\/} of the resonators have tolerances relative to a {\it pre-selected\/} distribution. For let ${\bf n}_a\/$ be a set of resonator locations, for example the {\sl TIGA\/} or the {\sl PHC\/} positions, and let ${\bf n'}_a\/$ be the real ones, close to the former: \begin{equation} {\bf n'}_a = {\bf n}_a + {\bf v}_a\,\eta^{1/2}\ ,\qquad a=1,\ldots,J \label{8.10} \end{equation} The values ${\bf n}_a$ determine the eigenvalues $\zeta_a\/$ in equation (\ref{5.2}), and also they appear as arguments to the spherical harmonics in the system response functions of sections~\ref{sec:srgw}--\ref{sec:hs}. It follows from~(\ref{8.10}) by continuity arguments that \begin{eqnarray} Y_{lm}({\bf n'}_a) & = & Y_{lm}({\bf n}_a) + O(\eta^{1/2}) \label{8.105.a} \\ \zeta'_a & = & \zeta_a + O(\eta^{1/2}) \label{8.105.b} \end{eqnarray} Inspection of the equations of sections~\ref{sec:srgw}--\ref{sec:hs} shows that both $\zeta_a\/$ and $Y_{lm}({\bf n}_a)$ {\it always\/} appear within lowest order terms, and hence that corrections to them of the type~(\ref{8.105.a})-(\ref{8.105.b}) will affect those terms in {\it second order\/} again. We thus conclude that the system is also {\it robust\/} to small misalignments of the resonators relative to pre-established positions. \subsubsection{Resonator frequency mistunings} The resonator {\it frequencies\/} may also differ amongst them, so let \begin{equation} \Omega_a = \Omega\,(1+\rho_a\,\eta^{1/2})\ ,\qquad a=1,\ldots,J \label{8.11} \end{equation} To assess the consequences of this, however, we must go back to equation (\ref{3.18}) and see what the coefficients in its series solutions of the type (\ref{4.11.a}) are. The procedure is very similar to that of section \ref{s8.1}, and will not be repeated here; the lowest order coefficient $\chi_\frac{1}{2}$ is seen to satisfy the algebraic equation \begin{equation} \det\left[\delta_{ab} - \frac{1}{\chi_\frac{1}{2}}\,\sum_{c=0}^{J}\, \frac{\chi_{ac}^{(n_0l_0)}\,\delta_{cb}}{\chi_\frac{1}{2}-\rho_c}\right] = 0 \label{8.12} \end{equation} which reduces to (\ref{5.1}) when all the $\rho\/$'s vanish, as expected. This appears to potentially have significant effects on our results to lowest order in $\eta^{1/2}$, but a more careful consideration of the facts shows that it is probably unrealistic to think of such large tolerances in resonator manufacturing as implied by equation (\ref{8.11}) in the first place. In the {\sl TIGA\/} experiment, for example \cite{phd}, an error of order $\eta^{1/2}$ would amount to around 50 Hz of mistuning between resonators, an absurd figure by all means. In a full scale sphere ($\sim$40 tons, $\sim$3 metres in diameter, $\sim$800 Hz fundamental quadrupole frequency, $\eta\/$\,$\sim$\,10$^{-5}$) the same error would amount to between 5 Hz and 10 Hz in resonator mistunings for the lowest frequency. This is probably excessive for a capacitive transducer, but may be realistic for an inductive one. With this exception, it is thus more appropriate to consider that resonator mistunings are at least of order $\eta\/$. If this is the case, though, we see once more that the system is quite insensitive to such mistunings. Summing up the results of this section, we can say that the resonator system dynamics is quite {\it robust\/} to small (of order $\eta^{1/2}$) changes in its various parameters. The important exception is of course the effect of suspension drillings, which do result in significant changes relative to the ideally perfect device, but which can be relatively easily calculated. This theoretical picture is fully supported by experiment, as {\it robustness\/} in the parameters here considered has been reported in reference \cite{jm97}. \section{Conclusions} A spherical GW antenna is a natural multimode device with very rich potential capabilities to detect GWs on earth. But such detector is not just a bare sphere, it requires a set of {\it motion sensors\/} to be practically useful. It appears that transducers of the {\it resonant\/} type are the best suited ones for an efficient performance of the detector. Resonators however significantly interact with the sphere, and they affect in particular its frequency spectrum and vibration modes in a specific fashion, which has to be properly understood before reliable conclusions can be drawn from the system readout. The main objective of this paper has been the construction and development of an elaborate theoretical model to describe the joint dynamics of a solid elastic sphere and a set of {\it radial motion\/} resonators attached to its surface at arbitrary locations, with the purpose to make predictions of the system characteristics and response, in principle with arbitrary mathematical precision. We have shown that the solutions to our equations of motion are expressible as an ascending series in powers of the small ``coupling constant'' $\eta\/$, the ratio of the average resonator mass to the mass of the large sphere. The {\it lowest order\/} approximation corresponds to terms of order $\eta^{1/2}$ and, to this order, we recover, and widely generalise, other authors' results \cite{jm97,ts,grg}, obtained by them on the basis of certain simplifying assumptions. This has in particular enabled us to assess the system response for arbitrary resonator layouts, and to search the equations for configurations other than the highly symmetric {\sl TIGA\/}. This search has led us to make a specific proposal, the {\sl PHC\/}, which is based on a pentagonally symmetric set of 5 rather than 6 resonators per quadruopole mode sensed. The {\sl PHC\/} distribution has the very interesting property that {\it mode channels\/} can be constructed from the resonators' readouts, much in the same way as in the {\it TIGA\/} \cite{jm95}. In the {\sl PHC\/} however a new and distinctive characteristic is present: different {\it wave amplitudes\/} selectively couple to different {\it detector modes\/} having different frequencies, so that the antenna's mode channels come at different rather than equal frequencies. The {\sl PHC\/} philosophy can be extended to make a {\it multifrequency\/} system by using resonators tuned to the first two quadrupole harmonics of the sphere {\it and\/} to the first monopole, an altogether 11 transducer set \cite{ls}. The assessment of {\it symmetry failure\/} effects, as well as other parameter departures form ideality, has also interested us here. This is seen to receive a particularly clear treatment in our general scheme: the theory transparently shows that the system is {\it robust\/} against relative disturbances of order $\eta\/$ or smaller in any system parameters, and provides a systematic procedure to assess larger tolerances ---up to order $\eta^{1/2}$. The system is shown to still be robust to tolerances of this order in some of its parameters, whilst it is not to others. Included in the latter group is the effect of spherical symmetry breaking due to system suspension in the laboratory, which causes {\it degeneracy lifting\/} of the sphere's eigenfrequencies, now split up into multiplets. By using our algorithms we have succeeded in numerically reproducing the reportedly measured frequencies of the {\sl LSU\/} prototype antenna \cite{phd} with a fully satisfactory precision of four decimal places. The experimentally reported robustness of the system to resonator mislocations \cite{jm97} is also in full agreement with our theoretical predictions. The perturbative approach we have adopted is naturally open to refined analysis of the system response in higher orders in $\eta\/$. For example, we can systematically address the weaker coupling of non-quadrupole modes. It appears however that such refinements will be largely masked by {\it noise\/} in a real system, as shown by Merkowitz and Johnson \cite{jm98}, and this must therefore be considered first. So our next step is to include noise in the model and see its effect. Stevenson \cite{ts} has already made some progress in this direction, and partly assessed the characteristics of {\sl TIGA\/} and {\sl PHC\/}, but more needs to be done since not too high signal-to-noise ratios should realistically be be considered in an actual GW detector. In particular, the discovery of {\it mode channels\/} also for the {\sl PHC\/} distribution opens the possibility of analysis of noise correlations and dependencies, as well as the errors in GW parameter estimation. These are natural extensions of this research, and some of them are currently underway \cite{lms}. \section*{Acknowledgments} We are indebted with Stephen Merkowitz for his kind supply of the {\sl TIGA\/} prototype data, without which a significant part of this work would have been speculative. Fruitful discussions with him are also gratefully acknowledged. We thank Eugenio Coccia for interaction and encouragement throughout the development of this research, and also Curt Cutler for pointing out to us an initial error in equation~(\ref{2.1.b}). We have received financial support from the Spanish Ministry of Education through contract number PB96-0384, and from the Institut d'Estudis Catalans.
\section{Introduction} Oscillations are ubiquitous in neural systems and have been the focus of several recent studies (for reviews see e.g.~Gray 1994, Singer and Gray 1995, Buzs{\'a}ki and Chrobak 1995, Ritz and Sejnowski 1997). In particular, fast global oscillations in the gamma frequency range ($> 30$ Hz) have been reported in the visual cortex (Gray et al 1989, Eckhorn et al 1993, Kreiter and Singer 1996), in the olfactory cortex (Laurent and Davidowitz 1994) and in the hippocampus (Bragin et al 1995). Even faster oscillations (200Hz) occur in the hippocampus of the rat (Buzs{\'a}ki et al 1992, Ylinen et al 1995). In some experimental data, (see e.g.~Eckhorn et al 1993, Csicsvari et al 1998, Fisahn et al 1998) individual neuron recordings show irregular spike emission, at a rate which is low compared to the global oscillation frequency\footnote{ Fast oscillations may be due in some cases to a synchronized subset of cells with high firing rates. The observation of cells with the required property has been recently reported in (Gray and McCormick 1996).}. This raises the question of whether a network composed of neurons firing irregularly at low rates can exhibit fast collective oscillations, which theoretical analyses and modelling studies may help to answer. Previous studies of networks of spiking neurons have mostly analyzed, or simulated, synchronized oscillations in regimes in which neurons behave themselves as oscillators, with interspike intervals strongly peaked around their average value (see e.g.~Mirollo and Strogatz 1990, Abbott and van Vreeswijk 1993, van Vreeswijk et al 1994, Gerstner 1995, Hansel et al 1995, Gerstner et al 1996, Wang and Buzs{\'a}ki 1996, Traub et al 1996). Several oscillatory regimes have been found with either full or partial synchronization. A regime particular to globally coupled systems has been described where the network breaks into a few fully synchronized clusters (Golomb and Rinzel 1994, van Vreeswijk 1996). In some simulations of networks with detailed biophysical characteristics, cells fire sparsely and irregularly during a global oscillation (Traub et al 1989, Kopell and LeMasson 1994, Wang et al 1995), but the complexity of individual neurons in these models makes it difficult to clearly understand of the origin of the phenomenon. The possible appearance of fast oscillations in a network where all neurons fire irregularly with an average frequency which is much lower than the population frequency therefore remains an intriguing question. It is the focus of the present work. Recurrent inhibition plays an important role in the generation of synchronized oscillations as shown by in vivo (McLeod and Laurent 1996) and in vitro experiments (Whittington et al 1995) in different systems. This has been confirmed by several modelling studies (van Vreeswijk et al 1994, Gerstner et al 1996, Wang and Buzs{\'a}ki 1996, Traub et al 1996). It has also been recently shown using simple models that networks in which inhibition balance excitation (Tsodyks and Sejnowski 1995, Amit and Brunel 1997a, van Vreeswijk and Sompolinsky 1996) are naturally composed of neurons with low and irregular firing. Simulations (Amit and Brunel 1997b) have shown that, in one such model composed of sparsely connected integrate-and-fire (IF) neurons, the highly irregular single neuron activity is accompanied by damped fast oscillations of the global activity. In order to study the coexistence of individual neurons with low firing rates and fast collective oscillations in its simplest setting, we analyze in the present paper a sparsely connected network entirely composed of identical inhibitory IF neurons. Our aim is to provide a clear understanding of this type of synchrony and to precisely determine :\\ - i) under which conditions collective excitations of high frequencies arise in such networks\\ - ii) what controls the different characteristics (amplitude, frequency, coherence time,...) of the global oscillation. Simulation results are presented first which shows that the essence of the phenomenon is present even in this simple system. Both the neurons firing rates and the auto-correlation of the global activity are very similar to those reported in (Amit and Brunel 1997b). We begin by presenting simple arguments which give an estimation of the firing rate of individual neurons and the frequency of the global oscillation and which lead to think that the global oscillation only appears above a well-defined parameter threshold. In order to make the analysis more precise and complete, we then generalize the analytic approach of Amit and Brunel (1997a) which was restricted to the computation of firing rates in stationary states. The sparse random network connectivity leads the firing patterns of different neurons to be only weakly correlated. As a consequence, the network state can be described by the instantaneous distribution of membrane potentials of the neuronal population, together with the firing probability in this population. We obtain the coupled temporal evolution equations for these quantities, the time-independent solution of which coincides with the stationary solution of (Amit and Brunel 1997a). A linear stability analysis shows that this time-independent solution becomes unstable only when the strength of recurrent inhibition exceeds a critical level, in agreement with our simple arguments. When this critical level is reached, the stationary solution becomes unstable and an oscillatory solution develops (via a Hopf bifurcation). The time scale of the period of the corresponding global oscillations is set by a synaptic time, independently of the firing rate of individual neurons, but the period precise value also depends on the characteristics of the external input. The analysis is then pushed to higher orders. We obtain a reduced evolution equation describing the network collective dynamics. The effects coming from the finite size of the network are also discussed. We show that having a large but finite number of neurons gives a small stochastic component to the collective evolution equation. As a result, it is shown that cross-correlations in a finite network present damped oscillations both above and below the critical inhibition level. Below the critical level, the noise controls the oscillation amplitude which decreases as the number of neurons is increased (at a fixed number of connections per neuron). Above the critical level, the main effect of the noise is to produce a phase diffusion of the global oscillation. An increase in the number of neurons results in an increase of the global oscillation coherence time and in a reduced damping in average cross-correlations. Finally, the effect of some of our simplifying assumptions is studied. We shortly discuss the effect of allowing variability in synaptic times and number of synaptic connections from neuron to neuron. We also consider the effect of introducing a more detailed description of postsynaptic currents into the model. The technical aspects of our computations are detailed in several appendices. \section{Description of the network and simulations} We analyse the dynamics of a network composed of $N$ identical inhibitory single compartment integrate-and-fire (IF) neurons. Each neuron receives $C$ randomly chosen connections from other neurons in the network. It also receives $C_{ext}$ connections from excitatory neurons outside the network (see Fig.~\ref{networkmap}). We consider a sparsely connected case with $\epsilon=C/N \ll 1$. \begin{figure} \setlength{\unitlength}{1cm} \begin{picture}(14,7) \put(-2,-8){\special{psfile="testfig.ps"}} \end{picture} \caption{Schematic diagram of the connections in the network of $N$ neurons; each neuron (indicated as an open disk) receives $C$ inhibitory connections (indicated as black) from within the network and $C_{ext}$ excitatory connections (indicated as grey) from neurons outside the network.} \label{networkmap} \end{figure} \begin{figure} \setlength{\unitlength}{1cm} \begin{picture}(14,6) \put(0,-2){\special{psfile="IPSP.ps"}} \put(0,5){Presynaptic spike} \put(0,3.1){PSC ($RI(t)$)} \put(0,1.3){PSP ($V(t)$)} \end{picture} \caption{Comparison of the synaptic response characteristics in our model and in a more realistic model. The top trace shows the presynaptic spike. The middle trace shows the corresponding postsynaptic current (PSC). The bottom trace shows the corresponding postsynaptic potential (PSP) for a neuron initially at resting potential. Full lines: our model, in which the synaptic current is described by a delta function a time $\delta$ after the presynaptic spike. Dashed lines: a more realistic synaptic response, in which the PSC is described by an $\alpha$-function with latency (transmission delay) $\tau_L$ and synaptic time constant $\tau_S$ $(t-\tau_L)\exp(-(t-\tau_L)/\tau_S)/\tau_S$. Our synaptic characteristic time $\delta$ can roughly be identified with the sum of latency and synaptic decay time, $\tau_L+\tau_S$. See the discussion in Section \ref{section:synaptic}.} \label{IPSP} \end{figure} Each neuron is simply described by its membrane potential. Let us suppose that neuron $i$ receives an inhibitory (excitatory) connection from neuron $j$. When the presynaptic neuron $j$ emits a spike at time $t$, the potential of the postsynaptic neuron $i$ is decreased (increased) by $J$ at time $t+\delta$ and returns exponentially to the resting potential in a time $\tau$ which represents the integration time constant of the membrane. In this simple model, the single time $\delta$ is meant to represent the transmission delays but also and most importantly, the longer time needed to obtain the full hyperpolarization of the post-synaptic neuron corresponding to a given presynaptic spike. Therefore, finding the correspondence between $\delta$ and the different synaptic time scales of a more realistic description needs some care. As pictorially shown in Fig.~\ref{IPSP}, $\delta$ should roughly be identified to the characteristic duration of the synaptic currents. In the following, we thus refer to $\delta$, which plays a crucial role in the generation of global oscillations, as the "synaptic time". The correspondence between $\delta$ and the different synaptic time scales of a more realistic description is further elaborated in Section \ref{section:synaptic} where synaptic current of finite duration are considered. Mathematically, the depolarization $V_i(t)$ of neuron $i$ ($i=1,\ldots,N$) at its soma obeys the equation, \begin{equation} \tau \dot{V}_i(t) = -V_i(t) + R I_i(t) \label{potdyn} \end{equation} where $I_i(t)$ are the synaptic currents arriving at the soma. These synaptic currents are the sum of the contributions of spikes arriving at different synapses (both local and external). These spike contributions are modelled as delta functions in our basic IF model: \begin{equation} \label{ispikes} R I_i(t)= \tau \sum_j J_{ij} \sum_k \delta(t - t_j^k-\delta) \end{equation} where the first sum on the r.h.s is a sum on different synapses ($j=1,\ldots,C+C_{ext}$), with postsynaptic potential (PSP) amplitude (or efficacy) $J_{ij}$, while the second sum represents a sum on different spikes arriving at synapse $j$, at time $t=t_j^k +\delta$, where $t_j^k$ is the emission time of $k$-th spike at neuron $j$. For simplicity, we take PSP amplitudes equal at each synapse, i.e.~$J_{ij} = J_{ext}>0$ for excitatory synapses and $J_{ij} =- J$ for inhibitory ones. External synapses are activated by independent Poisson processes with rate $\nu_{ext}$. A firing threshold $\theta$, completes the description of the IF neuron : when $V_i(t)$ reaches $\theta$, an action potential is emitted by neuron $i$, and the depolarization is reset to $V_r<\theta$ after a refractory period $\tau_{rp}$ during which the potential is insensitive to stimulation. A typical value would be $\tau_{rp}\sim 2$ms. We are interested here in network states in which the frequency is much lower than the corresponding maximal frequency $1/\tau_{rp}\sim 500$Hz. In this regime, we have checked that the exact value of $\tau_{rp}$ does not play any role. Thus in the following we set $\tau_{rp}$ to zero, for the sake of simplicity. \begin{figure} \setlength{\unitlength}{1cm} \begin{picture}(15,20) \put(1.,0){ \put(-3,-1.5){\special{psfile="lfp1.ps"}}\put(1,3){LFP} \put(-3,5.75){\special{psfile="lfp25.ps"}}\put(1,10.25){LFP} \put(-3,13.){\special{psfile="lfp5.ps"}}\put(1,17.5){LFP} \put(-3,1.3){\special{psfile="raster10L.ps"}} \put(-3,8.55){\special{psfile="raster25L.ps"}} \put(-3,15.8){\special{psfile="raster50L.ps"}} \put(7,-1.5){\special{psfile="cc10.ps"}}\put(12.5,3){AC} \put(7,5.75){\special{psfile="cc25.ps"}}\put(12.5,10.25){AC} \put(7,13.){\special{psfile="cc50.ps"}}\put(12.5,17.5){AC} \put(7,2.2){\special{psfile="isi10.ps"}}\put(12.5,6.7){ISI} \put(7,9.45){\special{psfile="isi25.ps"}}\put(12.5,13.95){ISI} \put(7,16.7){\special{psfile="isi50.ps"}}\put(12.5,21.2){ISI} \put(4.,-0.1){time(ms)} \put(11.2,-0.1){time(ms)} } \put(0,21){A} \put(0,13.75){B} \put(0,6.5){C} \end{picture} \caption{Left: Time evolution of the global activity (LFP) during a 100ms interval of the dynamics of a network of 5,000 neurons (total number of firing neurons in 0.4ms bins), together with spike rasters of 50 neurons, for different values of the external noise: $\sigma_{ext}=5$mV (A), 2.5mV (B), and 1 mV (C). Right: autocorrelation of the global activity (AC) and inter-spike interval (ISI) histogram averaged over 1000 neurons, corresponding to the left pictures. Note the different time scales of AC and ISI in abscissa. Parameters: $\theta=20$mV, $V_r=10$mV, $\tau=20$ms, $\delta=2$ms, $C=1000$, $J=0.1$mV, $\mu _{ext}=25$mV.} \label{figuresim1} \end{figure} The outcome of a typical simulation is shown in Figs.~\ref{figuresim1}. Neurons are driven by the random external excitatory input above threshold; however, since feedback interactions are inhibitory, the global activity stays at rather low levels (about 5Hz for the parameters indicated in Fig.~\ref{figuresim1}). For weak external noise levels ($\sigma_{ext}=1$mV), the global activity (total number of firing neurons in 0.4ms bins) is strongly oscillatory with a period of about 7 ms, as testified by Fig.~\ref{figuresim1}C. On the other hand, increasing the external noise level strongly damps and decreases the amplitude of the global oscillation. Note that the global activity should roughly correspond to the local field potential (LFP) often recorded in neurophysiological experiments. On the other hand, even when the global activity is strongly oscillatory, individual firing is extremely irregular as shown in the rasterfile of 50 neurons, Fig.~\ref{figuresim1}C (above the LFP), and in the inter-spike interval histogram (to the right of the spike rasters). In each oscillatory event only a small fraction of the neurons fire. This oscillatory collective behavior is also shown by fast oscillations in the temporal autocorrelation (AC) of the global activity which are damped on a longer time scale (Fig.~\ref{figuresim1}, to the right of the LFP). It is also reflected in the cross-correlations (CC) between the spike trains of a pair of neurons, which are typically equal to the AC of the global activity. These simulation results raise several questions on the origin and characteristics of the observed oscillations. What is the mechanism of the fast oscillation? In which parameter region is the network oscillating? What are the network parameters which control the amplitude and the different time scales (frequency, damping time constant) of the global oscillation? How do they scale with the network size? The model is simple enough and an analytical study gives precise answers to these questions as shown in the following sections. \section{An analysis of the network dynamics} \label{sec:an:analysis} Several features simplify the analysis as noted in a previous study (Amit and Brunel 1997a) of the neuron mean firing rates. First, as a consequence of the network sparse random connectivity ($C\ll N$), two neurons share a small number of common inputs and pair correlations can be neglected in the limit $C/N \rightarrow 0$. Second, we consider a regime where individual neurons have a firing rate $\nu$ low compared to their inverse integration time $1/\tau$ and receive a large number of inputs per integration time $\tau$, each input making a contribution small compared to the firing threshold ($J \ll \theta$)\footnote{Typical numbers in cortex are $C=5000$, $\tau=20$ms, $\nu=5$Hz, $J=0.1$mV, $\theta=20$mV so that $C\nu\tau$ is typically several hundreds while $\theta/J$ is of order 100 (Abeles 1991, Braitenberg and Shutz 1991). In the simulation shown in Fig.~\ref{figuresim1} $C\nu\tau\sim 100$, $\theta/J\sim 200$.}. In this situation, the synaptic current of a neuron can be approximated by an average part plus a fluctuating gaussian part, and the spike trains of all neurons in the network can be self consistently described by Poisson processes with a common instantaneous firing rate $\nu(t)$ but otherwise uncorrelated from neuron to neuron (that is, between $t$ and $t+dt$, a spike emission has a probability $\nu(t) dt$ of occurring for each neuron but these events occur statistically independently in different neurons) The synaptic current at the soma of a neuron (neuron $i$) can thus be written as, \begin{equation} R I_i(t) = \mu (t) + \sigma\sqrt{\tau} \eta_i(t) \label{idiffusion} \end{equation} The average part $\mu (t)$ is related to the firing rate at time $t-\delta$ and is a sum of local and external inputs \begin{equation} \mu = \mu _l + \mu _{ext} \,\,\mbox{ with }\,\, \mu _l = - CJ\nu(t-\delta)\tau, \;\;\;\; \mu _{ext}= C_{ext}J_{ext}\nu_{ext}\tau \label{mu} \end{equation} Similarly the fluctuating part, $\sigma\sqrt{\tau} \eta_i(t)$, is given by the fluctuation in the sum of internal and external poissonian inputs of rate $C\nu$ and $C_{ext}\nu_{ext}$. Its magnitude is given by \begin{equation} \sigma^2 = \sigma^2_l+\sigma^2_{ext} \,\,\mbox{ with }\,\, \sigma_l = J\sqrt{C\nu(t-\delta)\tau}, \;\;\;\; \sigma_{ext}= J_{ext}\sqrt{C_{ext}\nu_{ext}\tau} \label{sigma} \end{equation} and $\eta_i(t)$ is a gaussian white noise uncorrelated from neuron to neuron, $\langle \eta_i(t)\rangle =0$ and $\langle \eta_i(t) \eta_j(t')\rangle = \delta_{i,j}\delta(t-t')$. Before describing our precise results, it may be useful to give simple estimates which show how the neuron firing rates, the collective oscillation frequency and the oscillatory threshold can be obtained from Eqs.(\ref{idiffusion}-\ref{sigma}). Let us first consider the stationary case. The case of interest corresponds to $\mu <\theta$. When expression (\ref{idiffusion}) is used for the synaptic current, the dynamics of the neuron depolarization (\ref{potdyn}) is a stochastic motion in the harmonic potential $(V-\mu )^2$ truncated at the firing threshold $V=\theta$. The neuron firing rate $\nu_0$ is the escape rate from this potential. For a weak noise, it is given by the inverse of the time scale of the motion $1/\tau$ diminished by an Arrhenius activation factor. So, one obtains the simple estimate (up to an algebraic prefactor), \begin{equation} \nu_0\sim\frac{1}{\tau}\exp\left(-\frac{(\theta-\mu )^2}{\sigma^2}\right) \label{simpstatrate} \end{equation} This becomes a self-consistent equation for $\nu_0$ once $\mu $ and $\sigma$ are expressed in terms of $\nu_0$ using Eq.~(\ref{mu},\ref{sigma}). The simple estimate (\ref{simpstatrate}) is made precise below by following Kramers's classic treatment of the thermal escape over a potential barrier (Chandrasekhar 1943). The origin of the collective oscillation can also be simply understood. An increase of activity in the network due to a fluctuation provokes an increase in the average feedback inhibitory input. Thus after a period of about one synaptic time the activity should decrease due to the increase of the inhibitory input. This decrease will itself provoke a decrease in the inhibitory input, and a corresponding increase in the activity after a new period equal to the synaptic time. This simple argument predicts a global oscillation period of about a couple of times the synaptic time $\delta$, not too far from the period observed in the simulations. However, it does not seem to have been noted previously that a global oscillation of period $\delta$ can in fact occur only if it is not masked by the intrinsic noise in the system. The resulting oscillation threshold can be simply estimated in the limit where $\delta$ is short compared to the time scale of the depolarization dynamics. During a short time interval $\delta$, a neuron membrane potential receives from the local network an average input of magnitude $C\nu_0 \delta J$. The fluctuation in its membrane potential in the same time interval (due to intrinsic fluctuations in the total incoming current) is $\sigma \sqrt{\delta/\tau}$. The change in the average local input can be detected only if it is larger than the intrinsic potential fluctuations. A global oscillation can therefore occur only when $$\frac{CJ\nu_0\tau}{\sigma} = -\frac{\mu _l}{\sigma}\,\stackrel{>}{\sim}\, \sqrt{\frac{\tau}{\delta}}.$$ These simple estimations are confirmed by the analysis presented below and replaced by precise formulas. \subsection{Dynamics of the distribution of neuron potentials} When pair correlations are neglected, the system can be described by the distribution of the neuron depolarization $P(V,t)$, i.e.~the probability of finding the depolarization of a randomly chosen neuron at $V$ at time $t$. This distribution is the (normalized) histogram of the depolarization of all neurons at time $t$ in the large $N$ limit $N\rightarrow\infty$. The stochastic equation (\ref{potdyn},\ref{idiffusion}) for the dynamics of a neuron depolarization can be transformed into a Fokker-Planck equation describing the evolution of their probability distribution (Chandrasekhar 1943) \begin{equation} \label{fp} \tau\frac{\partial P(V,t)}{\partial t} = \frac{\sigma^2(t)}{2} \frac{\partial^2 P(V,t)}{\partial V^2}+ \frac{\partial}{\partial V}\left[ (V-\mu (t)) P(V,t)\right] \end{equation} The two terms in the r.h.s.~of (\ref{fp}) correspond respectively to a diffusion term coming from the current fluctuations and a drift term coming from the average part of the synaptic input. $\sigma(t)$ and $\mu (t)$ are related to $\nu(t-\delta)$, the probability per unit time of spike emission at time $t-\delta$, by Eq.~(\ref{mu},\ref{sigma}). Note that the Fokker-Planck equation has been used previously in studies of globally coupled oscillators (Sakaguchi et al 1988, Strogatz and Mirollo 1991, Abbott and van Vreeswijk 1993, Treves 1993). The resetting of the potential at the firing threshold ($V=\theta$) imposes the absorbing boundary condition $P(\theta,t)=0$. Moreover, the probability current through $\theta$ gives the probability of spike emission at $t$, \begin{equation} \frac{\partial P}{\partial V}(\theta,t) = -\frac{2\nu(t)\tau}{\sigma^2(t)} \label{bcth} \end{equation} At the reset potential $V=V_r$, $P(V,t)$ is continuous but the entering probability current imposes the following derivative discontinuity, \begin{equation} \frac{\partial P}{\partial V}(V_r^+,t) - \frac{\partial P}{\partial V}(V_r^-,t) = -\frac{2\nu(t)\tau}{\sigma^2(t)} \label{bcv} \end{equation} At $V= -\infty$, $P$ should tend sufficiently quickly toward zero to be integrable, i.e. \begin{equation} \lim_{V\rightarrow -\infty} P(V,t)=0 \;\;\; \lim_{V\rightarrow -\infty} V P(V,t)=0. \label{bci} \end{equation} Last, $P(V,t)$ is a probability distribution and should satisfy the normalization condition \begin{equation} \int_{-\infty}^{\theta} P(V,t) dV = 1 \label{norm} \end{equation} \subsection{Stationary states} We first consider stationary solutions $P(V,t)=P_0(V)$. Time independent solutions of Eq.~(\ref{fp}) satisfying the boundary conditions (\ref{bcth},\ref{bcv},\ref{bci}) are given by \begin{equation} \label{statdistr} P_0(V) = 2\frac{\nu_0\tau}{\sigma_0} \exp\left(-\frac{(V-\mu _0)^2}{\sigma_0^2} \right)\int_{\frac{V-\mu _0}{\sigma_0}}^{\frac{\theta-\mu _0}{\sigma_0}} \Theta\left(u - \frac{V_r-\mu _0}{\sigma_0}\right) e^{u^2}du \label{p0} \end{equation} with \begin{equation} \mu _0 =-CJ\nu_0 \tau +\mu _{ext}, \;\;\;\;\sigma_0^2 = CJ^2\nu_0\tau + \sigma_{ext}^2 \end{equation} (in (\ref{p0}), $\Theta(x)$ denotes the Heaviside function, $\Theta(x)=1\,$ for $x>0$ and $\Theta(x)=0$ otherwise). The normalization condition (\ref{norm}) provides the self-consistent condition which determines $\nu_0$ \begin{eqnarray} \frac{1}{\nu_0\tau} &=& 2 \int_{\frac{V_r-\mu _0}{\sigma_0}}^{\frac{\theta-\mu _0}{\sigma_0}} du e^{u^2}\int_{-\infty}^{u}dv e^{-v^2} \nonumber\\ &=& \int_0^{+\infty} du e^{-u^2}\left[\frac{e^{2y_{\theta} u} -e^{2y_r u}}{u} \right] \label{kraex} \end{eqnarray} with $y_{\theta}=\frac{\theta-\mu _0}{\sigma_0}, y_r= \frac{V_r-\mu _0}{\sigma_0}$. In the regime $(\theta-\mu _0)\gg \sigma_0$, Eq.~(\ref{kraex}) becomes \begin{equation} \nu_0\tau\simeq\frac{(\theta-\mu _0)}{\sigma_0\sqrt{\pi}} \exp\left(-\frac{(\theta-\mu _0)^2}{\sigma_0^2}\right) \label{kraas} \end{equation} In Fig.~(\ref{nurates}), the firing rates obtained by solving Eq.~(\ref{kraex}) and (\ref{kraas}) are compared with those obtained from simulations of the network. It shows an almost linear increase in the rates as a function of $\sigma_{ext}$ in the range 3-6Hz and a good agreement between Eq.~(\ref{kraex}) and the results of simulations. The asymptotic expression (\ref{kraas}) is also rather close to the simulation results in this range of $\sigma$. \begin{figure} \setlength{\unitlength}{1cm} \begin{picture}(14,6) \put(0,-1){\special{psfile="figurev.ps"}} \end{picture} \caption[]{The neuron firing rate vs $\sigma_{ext}$: simulation ($\diamond$); solution of Eq.~(\ref{kraex})(full line); solution of the approximate asymptotic form (\ref{kraas}) (dashed line). Others parameters are fixed as in Fig.~2 : $\tau=20$ms, $J=0.1$mV, $C=1000$, $N=5000$, $\theta=20$mV, $V_r=10$mV, $\mu _{ext}=25$mV, $\delta=2$ms.} \label{nurates} \end{figure} \subsection{Linear stability of the stationary states} We can now investigate in which parameter regime the time independent solution $(P_0(V),\nu_0)$ is stable. To simplify the study of the Fokker-Planck equation (\ref{fp}), it is convenient to rescale $P$, $V$ and $\nu$ by \begin{equation} P=\frac{2\tau\nu_0}{\sigma_0}Q,\,\,\, y=\frac{V-\mu _0}{\sigma_0},\,\,\, \nu=\nu_0(1+n(t)) \label{pytrntext} \end{equation} $y$ is the difference between the membrane potential and the average input in the stationary state, in units of the average fluctuation of the input in the stationary state. $n(t)$ corresponds to the relative variation of the instantaneous frequency around the stationary frequency. After these rescalings, Eq.(\ref{fp}) becomes \begin{equation} \tau \frac{\partial Q}{\partial t}=\frac{1}{2} \frac{\partial^2 Q}{\partial y^2} + \frac{\partial}{\partial y}(yQ) + n(t-\delta) \left(G \frac{\partial Q}{\partial y} +\frac{H}{2} \frac{\partial^2 Q}{\partial y^2}\right), \label{Eq} \end{equation} where $G$ is the ratio between the mean local inhibitory inputs and $\sigma_0$, and $H$ is the ratio between the variance of the local inputs and the total variance (local plus external): \begin{equation} G=\frac{C J\tau\nu_0}{\sigma_0}=\frac{-\mu _{0,l}}{\sigma_0},\,\,\, H=\frac{C J^2\tau\nu_0}{\sigma_0^2}=\frac{\sigma_{0,l}^2}{\sigma_0^2}, \label{ghtext} \end{equation} These parameters are a measure of the relative strength of the recurrent inhibitory interactions. Eq.~(\ref{Eq}) holds on the two intervals $-\infty<y<y_r$ and $y_r<y<y_{\theta}$. The boundary conditions on $Q$ are imposed at $y_{\theta}=\frac{\theta-\mu _0}{\sigma_0}$ and $y_r=\frac{V_r-\mu _0}{\sigma_0}$. Those on the derivatives of $Q$ read, \begin{equation} \frac{\partial Q}{\partial y}(y_{\theta},t)= \frac{\partial Q}{\partial y}(y_r^+,t)-\frac{\partial Q}{\partial y}(y_r^-,t) =-\,\frac{1+n(t)}{1+H n(t-\delta)} \end{equation} The linear stability of the stationary solution is studied in detail in Appendix \ref{app:linear:stability}. This can be done in a standard way (Hirsch and Smale, 1974) by expanding $Q=Q_0+Q_1+\ldots$ and $n=n_1+\ldots$ around the steady state solution. The linear equation obtained at first order has solutions which are exponential in time, $Q_1=\exp(w t/\tau)\hat Q_1$, $n_1\sim \exp(w/\tau) \hat{n}_1$, where $w$ is a solution of the eigenvalue equation (\ref{eigeneq}) of the Appendix. The stationary solution becomes unstable when the real part of $w$ becomes positive. When the synaptic time $\delta$ becomes much smaller than $\tau$, the roots $w$ of this equation become large. We consider the regime $\delta/\tau\ll 1$ but $\delta/\tau\gg 1/C$, which is the relevant case in simulations and correspond to the realistic regime. $\delta/\tau\gg 1/C$ is needed because otherwise the equations giving $G$ and $H$ become inconsistent with the condition $\tau\nu_0\ll 1$. At the oscillatory instability onset, $w$ is purely imaginary $w=i\omega_c$, where $\omega_c/\tau$ is the frequency of the oscillation which develops. The eigenvalue equation takes in the limit $\delta/\tau\rightarrow 0$, $\omega\rightarrow \infty$ the form \begin{equation} [\frac{G}{\sqrt{\omega_c}}(i-1)+H]\, \exp(-i \omega_c \delta/\tau) =1. \end{equation} In this limit, the instability line in the parameter space $(G,H)$ is obtained parametrically as \begin{eqnarray*} G&=&\sqrt{\omega_c}\sin\left(\frac{\omega_c\delta}{\tau}\right) \\ H&=&\sin\left(\frac{\omega_c\delta}{\tau}\right) + \cos \left(\frac{\omega_c\delta}{\tau}\right) \end{eqnarray*} $H$ is by definition constrained to be between 0 and 1 (it is the ratio between local and total variances): $H=0$ corresponds to the limit of very large external fluctuations, $\sigma_{ext}\gg \sigma_l$, while $H=1$ corresponds to $\sigma_{ext}=0$. We find that the frequency of the oscillation varies from \begin{eqnarray} \frac{\omega_c}{\tau} & = & \frac{3\pi}{4\delta}\quad\mbox{ when } H=0, \mbox{ to } \nonumber \\ \frac{\omega_c}{\tau} & = & \frac{\pi}{2\delta}\quad\mbox{ when } H=1. \label{omega:limit} \end{eqnarray} This corresponds to an oscillation with a period between $8\delta/3$ and $4\delta$, not too far from the value $2\delta$ obtained by simple arguments. At the same time the critical value of $G$ goes from \begin{eqnarray*} G_c & = & \sqrt{\frac{3\pi\tau}{8\delta}}\quad\mbox{ when } H=0, \mbox{ to } \\ G_c & = & \sqrt{\frac{\pi\tau}{2\delta}}\quad\mbox{ when } H=1. \end{eqnarray*} Again we find that it is proportional to $\sqrt{\tau/\delta}$ as anticipated. \begin{figure} \setlength{\unitlength}{1cm} \begin{picture}(14,4.5) \put(6.5,0){\put(-2,-1.5){\special{psfile="musi.ps"}} \put(0.5,3.3){$\mu _{ext}$} \put(4.5,-0.2){$\sigma_{ext}$}} \put(-0.5,0){\put(-2,-1.5){\special{psfile="geha.ps"}} \put(0.2,3.3){$H$} \put(4.,-0.2){$G\sqrt{\delta/\tau}$}} \put(2.5,3.5){SS}\put(5.7,3){OS}\put(9,4.5){OS}\put(12,2.5){SS} \end{picture} \caption{Left: instability line in the plane $(H,G\sqrt{\delta/\tau})$. Full line: instability line for parameters of Fig.\ref{figuresim1}, and $\delta=0.1\tau$. Long-dashed line: $\delta=0.05\tau$. Short-dashed line: asymptotic limit $\delta/\tau\rightarrow 0$. The stationary state (SS) is unstable to the right of the instability line, where an oscillatory instability develops (OS). Right: instability line in the plane $(\mu _{ext},\sigma_{ext})$. Full line: parameters of Fig.2, and $\delta=0.1\tau$. Short-dashed line is constructed taking the asymptotic instability line in the plane $(H,G\sqrt{\delta/\tau})$, and calculating the corresponding instability line in $(\mu _{ext},\sigma_{ext})$ with $\delta=0.1\tau$. The stationary state (SS) becomes unstable above the instability line. The long dashed line shows the average ($\mu _{ext}$) and the fluctuations ($\sigma_{ext}$) of the external inputs when the frequency of a Poissonian external input through synapses of strength $J_{ext}=0.1$mV is varied. For low external frequencies the network is in its stationary state. When the external frequency increases the network goes to its oscillatory state (OS).} \label{musi} \end{figure} This instability line can be translated in terms of the parameters $\mu _{ext}$, $\sigma_{ext}$, and calculated numerically using Eq.~(\ref{eigeneq}) for any value of the network parameters. This line of instability in the plane $(\mu _{ext},\sigma_{ext})$ is shown in the right part of Fig.~\ref{musi}. The stationary solution is unstable above the full line. Thus, if the external input is Poissonian, an increase in the frequency of external stimulation will typically bring the network from the stationary to the oscillatory regime, as indicated by the dashed line in Fig.~\ref{musi}, which represents the average ($\mu _{ext}$) and the fluctuations ($\sigma_{ext}$) of the external inputs when the frequency of a Poissonian external input through synapses of strength $J_{ext}=0.1$mV is varied. \subsection{Weakly non-linear analysis} \label{sec:weakly:nl} The linear stability analysis of the previous section shows that a small oscillation grows when one crosses the instability line in the plane $\mu _{ext}$, $\sigma_{ext}$. But it does not say much on the resulting characteristics of the resulting finite amplitude oscillation. In order to describe it and to be able to quantitatively compare analytic results to simulation data, one needs to compute the non linear terms which saturate the instability growth. This can be done in a standard manner (Bender and Orszag, 1987) by computing terms beyond the linear order in an expansion around the stationary state. The explicit computation is detailed in Appendix \ref{app:weakly:nl}. The collective oscillation is determined by the deviation $n_1$ of the neuron firing rate from its stationary value: $$ n_1(t) = \hat n_1(t) \exp(i\omega_c t/\tau) + \hat n_1^{\star}(t) \exp(-i\omega_c t/\tau) $$ $\hat n_1$ determines the amplitude of the collective oscillation as well as the nonlinear contribution to its frequency in the vicinity of the instability line. The analysis shows that the dynamics of the (small) deviation around the stationary firing rate can be described by the reduced equation \begin{equation} \label{eqmotiontext} \tau \frac{d\hat n_1}{dt} = A \hat n_1 - B |\hat n_1|^2 \hat n _1 \end{equation} in which $A$ and $B$ are complex numbers. The value of $A$ comes from the linear stability analysis. If Re$(A) <0$ a small initial value of $n_1$ decays and the stationary state is stable. On the contrary, if $Re(A)>0$ a global oscillation develops. When $|\hat n_1|$ grows, the second nonlinear term on the r.h.s. of (\ref{eqmotiontext}) becomes important. It is found here that Re$(B)>0$ (a "normal" or "supercritical" Hopf bifurcation) so that the nonlinear term saturates the linear growth. The characteristics of the oscillatory final state comes from the balance between the two terms. The explicit expression of $A$ and $B$ is given in Eqs.~(\ref{a},\ref{b}) as a ratio of hypergeometric functions of the network parameters. $A$ depends linearly on the deviation of the parameters $G$ and $H$ from their critical values, i.e.~$G-G_c$, $H-H_c$. In the limit $\delta/\tau\rightarrow 0$, the expressions of $A$ and $B$ simplify. For example, when $H=0$ (large external fluctuations), we find in the limit $\delta/\tau\rightarrow 0$ \begin{eqnarray} A & = & \frac{\tau}{\delta}\frac{\left(1+2i/3\pi\right)}{(1+4/9\pi^2)} \frac{G-G_c}{G_c} \simeq \frac{\tau}{\delta} \left(1.35 + 0.29i\right)\frac{G-G_c}{G_c} \nonumber \\ B & = & \frac{\tau}{\delta}\left(\frac{9\pi^2}{4+9\pi^2}\right)\left[\frac{13-5\sqrt{2}}{10}-\frac{9-5\sqrt{2}}{15\pi} +i\left(\frac{13-5\sqrt{2}}{15\pi}+\frac{9-5\sqrt{2}}{10}\right)\right] \nonumber \\ & \simeq & \frac{\tau}{\delta} (0.53+0.30 i) \label{ABsimp} \end{eqnarray} Generally, the complex numbers $A$ and $B$ can be written in terms of their real and imaginary parts, $A=A_r +iA_i$, $B=B_r +iB_i$. On the critical line, i.e. for $G=G_c$, $H=H_c$, $A_r=A_i=0$; above the critical line an instability develops, $A_r>0$, proportionnally to $G-G_c$ and $H-H_c$. The amplitude of this instability is controlled by the cubic term. The stable limit cycle solution of Eq.~(\ref{eqmotiontext}), above the critical line, is \begin{equation} \label{noiselessnu1text} \hat n_1(t) = R \exp\left(i\Delta\omega \frac{t}{\tau}\right) \end{equation} where $$ R = \sqrt{\frac{A_r}{B_r}}\quad\mbox{ and } \quad\Delta\omega = A_i - B_i\frac{A_r}{B_r} $$ The autocorrelation (AC) of the global activity, normalized by $\nu_0$, is, when $A_r>0$, \begin{eqnarray} C(s) & = & \lim_{T\rightarrow\infty} \frac{1}{T-s} \int_0^{T-s} (1+n_1(t))(1+n_1(t+s)) dt \\ \nonumber & = & 1+ 2R^2 \cos\left[(\omega_c+\Delta\omega)s/\tau\right]\nonumber \end{eqnarray} The AC is a cosine function of frequency $(\omega_c+\Delta\omega)/\tau$ and amplitude $R^2$. Compared with the AC function observed in the simulation, Fig.~\ref{figuresim1}C, we see a qualitative difference: there is no damping of the oscillation. The next Section shows that the damping is due to finite size effects. We analyze them before comparing quantitatively the analytical results with simulations. \subsection{Finite size effects and phase diffusion of the collective oscillation} We discuss the effect of having a large but only finite number of neurons in the network. It is well-known that for stochastic dynamics, a sharp transition can only occur in the limit $N\rightarrow\infty$ and that it will be smoothened by finite size effects. In the sparse connectivity limit, which allows to treat the quenched random geometry of the lattice in an annealed fashion\footnote{Here we do not consider the correlations due to the quenched connectivity for finite $\epsilon$. These correlations would give small corrections to the parameters calculated in the limit $\epsilon \rightarrow $0, but do not give rise to qualitatively new effects for the global activity such as the phase diffusion phenomenon discussed in this section.} the fluctuations in the input of a given neuron $i$ can be seen as the result of the randomness of two different processes: the first is the spike emission process $S(t)$ of the whole network; and the second, for each spike emitted by the network, is the presence or absence of a synapse between the neuron that emitted the spike and the considered neuron: if a spike is emitted at time $t$, $\rho_i(t)=1$ with probability $C/N$, and 0 otherwise. The input to the network is then $$ RI_i(t) = -J\tau \rho_i(t) S(t-\delta) $$ Both processes can be decomposed between their mean and their fluctuation, $$ \rho_i(t)=\frac{C}{N}+\delta \rho_i(t),\quad S(t)= N\nu(t) + \delta S(t) $$ Thus the input becomes $$ RI_i(t) = \mu (t) - J\tau N\nu(t) \delta \rho_i(t) -J\tau\frac{C}{N} \delta S(t) $$ in which $\mu (t)$ is given by Eq.~(\ref{mu}). The input is the sum of a constant part $\mu $, and of two distinct random processes superimposed on $\mu $: the first is uncorrelated from neuron to neuron, and we have already seen in Section \ref{sec:an:analysis} that it can be described by $N$ uncorrelated Gaussian white noises $\sigma\sqrt{\tau}\eta_i(t)$, $i=1,\ldots, N$ where $<\eta_i(t)\eta_j(t')> =\delta_{ij}\delta(t-t')$. The second part is independent of $i$: it comes from the intrinsic fluctuations in the spike train of the whole network which are seen by all neurons. This part becomes negligible when $\epsilon=C/N\rightarrow 0$, but can play a role as we will see when $C/N$ is finite. The global activity in the network is essentially a Poisson process with instantaneous frequency $N\nu(t)$. Such a Poisson process has mean $N\nu(t)$, which is taken into account in $\mu $, and variance $N\nu(t)\delta(t-t')$. The fluctuating part of this process is well approximated by a Gaussian white noise $\sqrt{N\nu_0}\xi(t)$, where $\xi(t)$ satisfies $<\xi(t)>=0$, $<\xi(t)\xi(t')>=\delta(t-t')$. Note that for simplicity we take the variance of this noise to be independent of time, which is the case for $ n_1(t)\ll 1$. These fluctuations are global and perceived by all neurons in the network. Thus, the mean synaptic input received by the neurons becomes $$ CJ\tau\nu(t) + J\sqrt{\epsilon C\nu_0\tau}\sqrt{\tau}\xi(t)+\mu _{ext} $$ Inserting this mean synaptic input in the drift term of the Fokker-Planck equation, we can rewrite Eq.~(\ref{Eq}) as \begin{equation} \tau \frac{\partial Q}{\partial t}=\frac{\partial}{\partial y} \{[y+ G n(t-\delta) + \eta \sqrt{\tau}\xi(t)] Q\} +\frac{1}{2} \frac{\partial^2 Q}{\partial y^2} \end{equation} where $\eta$ denotes the intensity of the noise stemming from these global fluctuations. $\eta$ tends to zero as the network size increases \begin{equation} \label{eta} \eta=\sqrt{\epsilon}\frac{\sigma_0^l}{\sigma_0} \end{equation} Taking into account this global noise term in the derivation of the reduced equation, we obtain, after some calculations described in Appendix \ref{app:noise}, \begin{equation} \label{noisymotiontext} \tau\frac{d\hat n_1}{dt} = A \hat n_1 - B |\hat n_1|^2 \hat n_1 + D \sqrt{\tau} \zeta(t) \end{equation} in which $A$, $B$ and $D$ are given by Eqs.~(\ref{a},\ref{b},\ref{c}), and $\zeta$ is a complex white noise such that $<\zeta(t)\zeta^{\star}(t')>=\delta(t-t')$. $D$ is proportional to $\eta$, i.e.~to both the square root of the connection probability and to the ratio between local and total fluctuations. Thus, the effect of the finite size of the network is to add a small stochastic component to the evolution equation of $n_1$, Eq.~(\ref{noisymotiontext}). Its main effect is to produce a phase diffusion of the collective oscillation {\footnote This global phase diffusion in a network of finite size is well-known (see e.g. (Rappel and Karma, 1996) for a simple example)} which leads to the damping of the oscillation in the autocorrelation function. \subsubsection*{Amplitude of the autocorrelation} From the reduced Eq.~(\ref{noisymotiontext}), one can compute exactly the autocorrelation at zero time $C(0)$ as shown in Appendix \ref{app:noise}. This gives : \begin{itemize} \item In the stationary regime far from the critical line, $A_r <0,|D|/|A_r|\ll 1 $: \begin{equation} C(0) -1 \sim \frac{|D|^2}{|A_r|} \sim O\left(\frac{C}{N}\right) \label{c01} \end{equation} The amplitude of the fluctuations in the global activity are proportional to $C/N$ and thus vanish when the connection probability goes to zero. \item On the critical line, $A_r=0$ \begin{equation} C(0)-1 = \frac{2 |D|}{\sqrt{\pi B_r}}\sim O\left(\sqrt{\frac{C}{N}}\right) \label{c02} \end{equation} The amplitude of the fluctuations are proportional to the square root of the connection probability. \item In the oscillatory regime far from the critical line, $A_r >0, |D|/A_r \ll 1$ : \begin{equation} C(0)-1 \sim \frac{2A_r}{B_r}\sim O\left(1\right) \label{c03} \end{equation} In this regime the amplitude of the oscillation is to leading order independent of the noise amplitude. \end{itemize} \subsubsection*{Oscillations below the critical line} In the stationary regime far from the critical line, the fluctuations of activity $n_1$ provoked by the noise term can be considered small and thus we can neglect the cubic term. It is then easy to calculate the autocorrelation (AC) of the activity, \begin{equation} \label{ACbelow} C(s) = 1+ \frac{|D|^2}{|A_r|} \exp\left(-\frac{|A_r|s}{\tau}\right) \cos\left([\omega_c + A_i]\frac{s}{\tau}\right) \end{equation} It is a damped cosine function. The damped oscillation has frequency $(\omega_c +A_i)/\tau$ and damping time constant proportional to $\tau/|A_r|$. The amplitude of the autocorrelation function is proportional to $C/N$. \subsubsection*{Oscillations above the critical line} In the oscillatory regime far from the critical line, we find in Appendix \ref{app:noise} an AC function of the form \begin{equation} \label{ACabove} C(s) = 1 + 2\frac{A_r}{B_r} \cos\left((\omega_c+\Delta \omega) s/\tau\right)\exp\left(-\frac{\gamma^2(s)}{2} \right) \end{equation} It is again a damped cosine function. The damping factor $\exp\left(-\gamma^2(s)/2 \right)$ is different from an exponential only at short times $s \sim \delta$. At longer times, $s\gg \delta$, we obtain again an exponential $$ \exp\left(-\frac{\gamma^2(s)}{2}\right) = \exp\left(-\frac{|D|^2}{4 R^2}\left(1+ \frac{B_i^2}{B_r^2}\right)\frac{s}{\tau}\left[1+\frac{|D|^2}{2A_r}+ O\left(|D|^4\right)\right] \right) $$ The damping time constant is proportional to leading order in $|D|$ to $1/|D|^2\sim N/C$, i.e.~to the inverse of the connection probability. When $N$ goes to infinity at $C$ fixed the `coherence time' of the oscillation increases linearly with $N$. This `phase diffusion' effect is the main finite size effect above the critical line. Both the amplitude and frequency of the oscillation are essentially unaffected by these finite size effects. \subsection{Comparison between simulations and theory} The autocorrelation (AC) of the global activity was computed for each set of parameters from a simulation of 20 seconds. Few longer simulations were performed as a check. The autocorrelation obtained in the longer simulations are essentially identical to the one obtained in the 20s simulation. Since the analysis predicts AC functions described by damped cosine functions, a least square fit of all AC functions was performed with such functions. Thus the full AC is reduced to three parameters, its amplitude at zero lag $C_0$, its frequency $\omega$, and its damping time constant (or coherence time) $\tau_c$ $$ C(s) = 1 + C_0 \exp\left(-\frac{|s|}{\tau_c}\right) \cos(\omega s) $$ We then compared the result of the fitting procedure with the analytical expressions. We have varied the magnitude of the external noise $\sigma_{ext}$ from 0 to 5 mV. This brings the network from the `oscillatory' to the `stationary' state. \begin{figure} \setlength{\unitlength}{1cm} \begin{picture}(14,20) \put(0,14){\put(0,6.5){A}\put(-1,-1.5){\special{psfile="siamp.ps"}} \put(1,4){$C_0$} } \put(0,7){\put(0,6.5){B} \put(-1,-1.5){\special{psfile="sifre.ps"}} \put(0.5,4){$\frac{\omega}{2\pi}$(Hz)} } \put(0,6.5){C}\put(-1,-1.5){\special{psfile="sidam.ps"}} \put(6,0){$\sigma_{ext}$(mV)} \put(0.5,4){$\tau_c$(ms)} \end{picture} \caption[]{Parameters of the AC function vs $\sigma_{ext}$. A. Amplitude of the AC at zero lag. B. Frequency. C. Damping time constant. Diamonds: simulation of the full network. Crosses : simulation of the reduced equation. Dashed lines: theory. In A, the short-dashed line represents the amplitude in the limit $N\rightarrow\infty$ Parameters: $\tau=20$ms, $J=-0.1$mV, $C=1000$, $N=5000$, $\theta=20$mV, $V_r=10$mV, $\mu _{ext}=25$mV, $\delta=2$ms.} \label{siparams} \end{figure} In Fig.~\ref{siparams} we plot together the results of simulations and theory. In these figures the diamonds are the simulation results; the dashed lines, the analytical results. In A, the short-dashed line indicates the amplitude in the limit $N\rightarrow\infty$, while the long-dashed line indicates the amplitude calculated analytically taking into account finite size effects. Last, the crosses are obtained simulating numerically the reduced equation, Eq.~\ref{noisymotiontext}. We find that, in the `stationary' regime as well as in the oscillatory regime close to the bifurcation point, the amplitude of the oscillation obtained in the simulation is in very good agreement with the calculation (Fig.~\ref{siparams}.A). On the other hand, as the amplitude of the oscillation becomes of the same order as the average frequency, $C_0\sim 1$, higher order effects become important and the calculation overestimates the amplitude of the AC. For the frequency of the oscillation (Fig.~\ref{siparams}.B), the calculation reproduces quite well the results of the simulations, except for very low noise levels, for which we are rather far from the bifurcation point. Note that the frequency ranges for this set of parameters from 70 to 180Hz, depending on the level of external noise. Thus, without varying the time constants $\tau$ and $\delta$, we find that the same network is able to sustain a collective oscillation at quite different frequencies. Last, the approximate analytical expressions for the damping time constant agree well with the simulation away from the bifurcation point, as expected (Fig.~\ref{siparams}.C). On the other hand, the simulation of the reduced equation is in good agreement with the network simulations in the whole range of $\sigma_{ext}$. \begin{figure} \setlength{\unitlength}{1cm} \begin{picture}(14,20) \put(0,14){\put(0,6.5){A}\put(-1,-1.5){\special{psfile="cc2.ps"}} \put(1,4){$C(t)$} } \put(0,7){\put(0,6.5){B}\put(-1,-1.5){\special{psfile="cc3.ps"}} \put(1,4){$C(t)$} } \put(0,6.5){C}\put(-1,-1.5){\special{psfile="cc4.ps"}} \put(6,0){$t$ (ms)} \put(1,4){$C(t)$} \end{picture} \caption[]{AC for: A. $\sigma_{ext}=2$mV. B. $\sigma_{ext}=3$mV. C. $\sigma_{ext}=4$mV. Parameters as in Fig.~\ref{siparams}. Full lines: network simulation. Dashed lines: theory (simulation of the reduced equation).} \label{ccs} \end{figure} In Fig.~\ref{ccs} we compare the full AC functions from theory (simulation of the reduced equation) and network simulations in three regimes, to show the good agreement between both. \section{Extensions} In the previous sections a very simple network has been analyzed and the question of the effect of some of our simplifying assumptions legitimately arises. In particular, we have chosen exactly identical neurons. It can be wondered how the results are modified when some variations in neuron properties are taken into account. In order to address this question, we show how the previous analysis can be generalized in two cases. Since we have seen that the oscillation frequency is tightly linked to synaptic times, the effect of a fluctuation in synaptic times is investigated first. We then consider the effect of a fluctuation in the number of connections per neuron which has been found to result in a wide spectrum of neuron steady discharge rates (Amit and Brunel, 1997b). In both cases, it is reassuring to find that the picture obtained from the simple model analysis remains accurate. We finally consider a model with synaptic currents of finite duration to analyse more precisely which time scale plays the role of our "synaptic time" in this more realistic case. \subsection{Effect of inhomogeneous synaptic times} \label{section:delays} The analysis can easily be extended to the case in which time constants at each synaptic site are drawn randomly and independently from an arbitrary probability density function (pdf) $\Pr(\delta)$ (see Appendix \ref{app:delays}). In the following we consider the case of a uniform pdf between 0 and $2\delta$. \begin{figure} \setlength{\unitlength}{1cm} \begin{picture}(14,6) \put(-2,-1.5){\special{psfile="musidel.ps"}} \put(0,3.7){$\mu _{ext}$} \put(5.,0){$\sigma_{ext}$} \put(2.2,5.5){OS}\put(6,3){SS} \end{picture} \caption{Instability line in the plane $(\mu _{ext},\sigma_{ext})$ for $\tau=20$ms, $J=0.1$mV, $C=1000$, $\theta=20$mV, $V_r=10$mV, $\delta=2$ms. Full line: all synaptic times equal to $\delta$. Dashed line: synaptic times drawn from a uniform distribution from 0 to 2$\delta$.} \label{musidel} \end{figure} Fig.~\ref{musidel} shows how the instability line is modified by random synaptic times. The region where the oscillatory instability appears shrinks to the area above the dashed line. As the distribution of synaptic times widens, the stationary state becomes more stable. The introduction of random synaptic times also slightly reduces the frequency of the oscillation. The critical line is thus quite sensitive to the distribution of synaptic times. In fact, distributions of synaptic times can be found such that the stationary state is always stable (e.g.~for an exponential distribution $\Pr(\delta) = \exp(-\delta/\delta_0)/\delta$). \subsection{Effect of inhomogeneous connectivity} \begin{figure} \setlength{\unitlength}{1cm} \begin{picture}(14,6) \put(-2,-1.5){\special{psfile="inmusi.ps"}} \put(0,3.7){$\mu _{ext}$} \put(5.,0){$\sigma_{ext}$} \put(3.5,5){OS}\put(6.5,3){SS} \end{picture} \caption{Effect of inhomogeneity in the connections on the instability line in the plane $(\mu _{ext},\sigma_{ext})$ for $\tau=20$ms, $J=-0.1$mV, $C=1000$, $\theta=20$mV, $V_r=10$mV, $\delta=2$ms. Full line: all neurons receive $C$ connections. Dashed line: connections are drawn randomly and independently at each synaptic site with probability $C/N$.} \label{musiinhom} \end{figure} \begin{figure} \setlength{\unitlength}{1cm} \begin{picture}(14,5.5) \put(7.5,0){\put(-2,-1.5){\special{psfile="ccvsrate.ps"}} \put(0.,3.2){$C(\nu)$} \put(4.,0){$\nu$ (Hz)}} \put(0,0){\put(-2,-1.5){\special{psfile="distri.ps"}} \put(0.2,3.2){$\Pr(\nu)$} \put(4.,0){$\nu$ (Hz)}} \end{picture} \caption{Left: Distribution of spike rates (Histogram: simulation. Dashed line: theory). The distribution is similar to a Gaussian, unlike the distributions observed in (Amit and Brunel 1997b), which are much wider, due to the balance between excitation and inhibition. Right: Relative amplitude of CC between individual neurons and the global activity vs neuronal firing rate (Diamonds: simulation. Full line: theory). $\tau=20$ms, $J=-0.1$mV, $C=1000$, $\theta=20$mV, $V_r=10$mV, $\delta=2$ms, $\mu _{ext}=25$mV, $\sigma_{ext}=2.58$mV. } \label{ccvsrate} \end{figure} The analysis can also be extended to the case when the number of connections impinging on a neuron is no longer fixed at $C$, but rather connections are drawn at random independently at each site. In that case the number of connections received by a neuron is a random variable with mean $C$ and standart deviation $\sim\sqrt{C}$. This inhomogeneity in the connectivity provokes a significant inhomogeneity in the individual spike rates even for $C$ large, because differences between the average input received by two neurons are of the same order as the SD of the synaptic input. The distribution of frequencies for an arbitrary network of excitatory and inhibitory neurons has been obtained in (Amit and Brunel 1997b). The main steps leading to this distribution are described in appendix \ref{app:inhomogeneous}. Next we study how inhomogeneity affects the dynamical properties of the network. Fig.~\ref{musiinhom} shows that the instability line is almost unaffected by the inhomogeneity. The frequency of the global oscillation is also very close to the one of the homogeneous case. Amit and Brunel (1997b) had shown by simulations that the degree of synchronization of a neuron with the global activity is strongly affected by its spike rate: neurons with low firing frequencies tend to be more synchronized with the global activity than neurons with high frequencies. In appendix \ref{app:inhomogeneous} we calculate analytically the degree of synchronization of individual neurons as a function of their frequency. The result is shown in Fig.~\ref{ccvsrate} in which the relative amplitude $C(\nu)$ of the cross-correlation between neurons firing at frequency $\nu$ and the global activity obtained analytically is compared with the result of simulations. It shows indeed that low-rate neurons are more synchronized with the global activity than high-rate neurons. The relative amplitude of the cross-correlation between two neurons of frequency $\nu_1$ and $\nu_2$ is given by the product of the two amplitudes, $C(\nu_1)C(\nu_2)$. Note that the heterogeneity in rates and cross-correlations is not very pronounced here, because near the critical line the fluctuations in the external input dominate the local fluctuations, which tends to suppress this heterogeneity. In a network with both excitatory and inhibitory neurons with an external excitatory input of the same order than the internal excitatory contribution, this heterogeneity is much more pronounced (Amit and Brunel 1997b). \subsection{Effect of more realistic synaptic responses} \label{section:synaptic} Our analysis has been carried out for synaptic currents which are described by a delta pulse. One may wonder how the analysis generalizes for more realistic postsynaptic currents. We consider a function $f(t)$ describing the shape of the postsynaptic current when a spike is emitted at time $t=0$ (see e.g.~Gerstner 1995 for a review of different types of synaptic responses). $f(t)$ is chosen such as $$ \int dt f(t) = 1 $$ An example often used in modelling studies and shown in Fig.~\ref{IPSP} is the $\alpha$-function with a latency $\tau_L$ and a characteristic synaptic time $\tau_S$: \begin{equation} f(t)= \left\{ \begin{array}{ll} \frac{t-\tau_L}{\tau_S^2}\exp\left(-\frac{t-\tau_L}{\tau_S}\right) &\mbox{for $t>\tau_L$} \\ 0 & \mbox{otherwise.} \end{array} \right. \end{equation} The total synaptic current arriving at neuron $i$ is now $$ R I_i(t) = \tau \sum_j J_{ij} \sum_k f\left(t-t_j^k\right) $$ In the diffusion approximation the synaptic current becomes $$ R I_i(t) = \mu (t) + \Xi_i(t) $$ in which the average part is given as a function of the frequency $\nu$ and the synaptic response function $f$ by $$ \mu (t) = \mu _{ext} -CJ\int dt' \nu(t')f(t-t') \tau. $$ On the other hand, the fluctuating part $\Xi_i(t)$ can no longer be approximated by a pure white noise and exhibits temporal correlations at the scale of the width of the PSC function $f(t)$. These temporal correlations in the currents complicate significantly the analysis, since the evolution of the distribution of the membrane potentials is no longer given by a simple one-dimensional Fokker-Planck equation. For the case of the $\alpha$-function, we would need to solve the problem described by a three dimensional Fokker-Planck equation. Such an analysis is beyond the scope of the present paper. Here, we choose to ignore, as a first approximation, these temporal correlations. Thus we consider only the effect of the PSC function on the average synaptic currents. In this approximation, the effect of the PSC function becomes equivalent to that of a distribution of synaptic times in the delta pulse PSC case considered in section \ref{section:delays}. For example, in the limit in which $\tau_S$ and $\tau_L$ are small compared to the integration time constant, the equations for the bifurcation point are \begin{eqnarray} G & = & \sqrt{\omega} \left[2\frac{\tau_S}{\tau} \omega \cos \left(\omega\frac{\tau_L}{\tau_S} \right) + \left(1-\frac{\tau_S^2}{\tau^2}\omega^2 \right) \sin \left(\omega\frac{\tau_L}{\tau} \right) \right] \nonumber \\ H & = & \left(1-\frac{\tau_S^2}{\tau^2}\omega^2\right) \left[\cos\left(\omega\frac{ \tau_L}{\tau}\right) + \sin\left(\omega\frac{\tau_L}{\tau}\right)\right] \nonumber\\ & &+2\frac{\tau_S}{\tau}\omega \left[\cos\left(\omega\frac{\tau_L}{\tau}\right) -\sin\left(\omega\frac{\tau_L}{\tau}\right)\right] \label{freqalpha} \end{eqnarray} In the case $\tau_L=0$ (zero latency) the equations simplify to \begin{eqnarray} G & = & 2\sqrt{\omega} \frac{\tau_S}{\tau} \omega \\ H & = & 1-\frac{\tau_S^2}{\tau^2}\omega^2 +2\frac{\tau_S}{\tau}\omega \end{eqnarray} In the case $H=1$, the frequency of the oscillation near the bifurcation point is equal to $1/(\pi\tau_S)$. Note that the dependence of the frequency on $\tau_S$ in the $\alpha$ function PSC case is similar to the dependence on $\delta$ in the delta pulse PSC case, Eq.~(\ref{omega:limit}). \begin{figure} \setlength{\unitlength}{1cm} \begin{picture}(14,6) \put(-1,-1.5){\special{psfile="synaptic.ps"}} \put(5.5,0){$\tau_S$} \put(1,3.5){$\frac{\omega}{2\pi}$} \end{picture} \caption{Dependence of the frequency of the oscillation near the bifurcation threshold on the synaptic decay time constant $\tau_S$, for $\tau_L=$ 2ms. Network parameters as in Fig.~\ref{figuresim1}. External inputs have $\mu _{ext}=$ 25mV, $\sigma_{ext}=$ 2mV. This point is near the bifurcation line in the whole range of $\tau_S$. $\diamond$: simulations. Full lines: frequency given by the approximate analysis, Eq.~\ref{freqalpha}, for $H=1$ (lower curve), and $H=0$ (upper curve).} \label{omtaus} \end{figure} To check the validity of this approximation, we have performed numerical simulations with fixed latency $\tau_L=$ 2ms, varying the decay time constant of the inhibitory post synaptic currents (IPSC) $\tau_S$. The results are shown in Fig.~\ref{omtaus}. The approximate analysis predicts the frequency is in the region between the two full lines (corresponding to $H=0$ and $H=1$). Simulation results deviate from the approximate analysis already at rather small values of $\tau_S$, because of the effect of temporal correlations in the synaptic currents, which have the same scale as the period of the oscillation. Nonetheless the approximation gives a good qualitative picture of the dependence of the frequency on $\tau_S$. Note that the frequencies obtained in this way can be directly compared to the data of (Whittington et al 1995, Traub et al 1996) since the decay time constant of the PSCs can be identified with their parameter $\tau_{GABA}$. The frequencies obtained in the simulations are very close to the ones obtained in that study. For example, we obtain a frequency of about 40Hz when $\tau_S=$ 10ms, in agreement with the in vitro recordings and the simulations of the more complex model of (Whittington et al 1995, Traub et al 1996). However, one has to be careful with such a comparison, since in that {\em in vitro} study, interneurons seem to fire at population frequency. \section{Conclusion} We have studied the existence of fast global oscillation in networks where individual neurons show irregular spiking at a low rate. We have first shown that the phenomenon can be observed in a sparsely connected network composed of basic integrate and fire neurons. In this very simplified setting, the phenomenon has been precisely analyzed. At the simplest level, it differs from other modes of synchronisation which lead to global oscillation in that recording at the individual neuron level shows a stochastic spike emission with nearly Poissonian interspike intervals and little indication of the collective behavior (see the ISI histograms in Fig.~\ref{figuresim1}). This oscillation regime has some similarity with that obtained in Wang et al (1995) where a hyperpolarization-activated cation current seems to play the role of our random external inputs in generating intermittent activity in the network. This type of weak synchronization has sometimes been rationalized as coming from filtering of external noise by recurrent inhibition (Traub et al 1989 and refs.~therein). Our analysis leads to a somewhat different picture. We have found that, in the limit of an infinite network, the global oscillation is due to an oscillatory instability (a supercritical Hopf bifurcation) of the steady state. This instability occurs at a well defined threshold and arises from the competition between the recurrent inhibition which favors oscillations and the intrinsic noise in the system which tends to suppress it. We have found that the global oscillation period is controlled by the synaptic time. This appears to agree with previous experimental findings on slices of the rat hippocampus and with simulations results (Whittington et al 1995, Traub et al 1996) where it is however assumed that neurons fire at population frequency, unlike those of our model. A similar decrease in population frequency when the GABA characteristic time is varied is also observed in a recent {\em in vitro} experiment in which neurons fire sparsely (Fisahn et al 1998). More work is necessary to clarify the relative roles of the different time constants (latency, IPSC rise time, IPSC decay time) that are commonly used to describe the synaptic response. The oscillation period also depends on the characteristics of the external input, and particularly on the magnitude of the external noise, as shown by Fig.~\ref{siparams}. The initial rise in the frequency when one increases $\sigma_{ext}$ followed by a saturation at sufficiently large $\sigma_{ext}$ looks in fact similar to the dependence of the frequency on the amount of glutamate applied to hippocampal CA1 region {\em in vitro} (Traub et al 1996). Our network is in a stationary state when external inputs are low and switches to an oscillatory regime when the magnitude of the external inputs is increased. This phenomenon resembles the induction of a gamma rhythm in the hippocampal slice mediated by carbachol (Fisahn et al 1998), and the induction of faster 200Hz rhythms, believed to be provoked by a massive excitation of CA1 cells through Schaeffer collaterals (see e.g.~Buzs{\'a}ki et al 1992). It is also interesting to note that a single network, with its internal parameters fixed, is able to sustain collective oscillations in different frequency ranges, when the characteristics of the external input are varied. In a finite network, the sharp transition is smoothened but the global oscillation has different characteristics above and below the critical threshold. Below threshold, its amplitude decreases as the network size is increased. Above threshold, an increase in the neuron number does not greatly modify the oscillation amplitude but increases its coherence time. It has been shown that the whole picture of a Hopf bifurcation with a well-defined threshold remains accurate when some of our simplifying assumptions are relaxed. It would be interesting to extend this finding to more realistic descriptions. Our analysis also raises the important question of the synchronisation mode used in real neural systems. Do neocortical or hippocampal neurons behave as oscillators with a frequency equal to the population frequency, or irregularly with firing rates lower than the population frequency? In hippocampus, pyramidal cells seems clearly to be in a irregular, low rate regime, during in vivo gamma (Bragin et al 1995), in vivo 200Hz (Buzs{\'a}ki et al 1992) and in vitro gamma oscillations (Fisahn et al 1998). More recent experimental data indicates that interneurons also typically fire at a lower frequency than the population frequency during 200Hz oscillations in CA1 (Csicsvari 1998). Further experimental work is needed in order to clarify this important issue. We have obtained a reduced description of the collective dynamics. The analysis can certainly be extended to more complicated networks, composed of neurons of different types or that are spatially extended. This reduced description will hopefully prove useful in clarifying the mechanisms of long range synchrony and in studying propagation phenomena (Delaney et al 1996, Prechtl et al 1997). Finally, and most importantly, the exact roles of fast oscillations remain, at present, unclear. Are they useful for putting in resonance different neuronal populations as it has been suggested? Can they serve to build a fast detector with slowly firing neurons? Are they used as a clock mechanism? Or do they reflect the usefulness of having a network where different neuronal populations fire in succession on a short time scale, to code spatial information in the temporal domain? Recent experiments (MacLeod and Laurent 1996, Stopfer et al 1997) make us hope that elucidating the real meaning of these collective oscillations, at least in some neural systems, is now an attainable goal. This is a question to which we hope to return in the future. {\bf Acknowledgments}. We are grateful to A. Karma for discussions and for his very stimulating role at the beginning of this work, and to T. Bal, R. Gervais and P. Salin for informing us on real neural networks. N.B. is grateful to S. Fusi for useful discussions. V.H. is glad to thank at last A. Babloyantz for an invitation to a stimulating ESF workshop in Lanzarote which was a nice opportunity to first learn about fast neuronal oscillations. We thank D.~Amit and anonymous referees for their helpful comments on the manuscript.
\section{Introduction} The gamma-ray burst (GRB) 990123 was an extraordinary event. It was the brightest burst yet detected with the Wide Field Camera on the BeppoSAX satellite (Feroci et al. 1999), and had a total gamma-ray fluence of $\sim 5\times 10^{-4}\,{\rm erg}\,{\rm cm}^{-2}$, which is in the top 0.3\% of all bursts. It was the first burst to be simultaneous detected in the optical band. Optical emission with peak magnitude of $V\sim 9$ was discovered by the Robotic Optical Transient Search Experiment (ROTSE) during the burst and was found to have rapidly faded down immediately after the gamma-ray emission (Akerlof et al. 1999). The detection of the redshift showed that the burst appears at $z\ge 1.6$ (Andersen et al. 1999; Kulkarni et al. 1999a). This implies that if the GRB emission was directed isotropically, the inferred energy release is $\ge 1.6\times 10^{54}\, {\rm ergs}$ (Kulkarni et al. 1999a; Briggs et al. 1999). The burst's afterglow was detected and monitored at X-ray, optical and radio bands. It was the brightest of all GRB X-ray afterglows observed until now. The BeppoSAX detected the flux of the afterglow at 2-10 keV six hours after the gamma-ray trigger to be $1.1\times 10^{-11}\,{\rm erg}\,{\rm cm}^{-2}\,{\rm s}^{-1}$ and the subsequent temporal decay index to be $\alpha_X=-1.44\pm 0.07$ (Heise et al. 1999a, b). The R-band optical afterglow about 3.5 hours after the burst showed a power-law decay with index $\alpha_{1R}=-1.1\pm 0.03$ (Kulkarni et al. 1999a; Castro-Tirado et al. 1999; Fruchter et al. 1999). This law continued until about $2.04\pm 0.46$ days after the burst. Then the optical emission began to decline based on another power law with index $\alpha_{2R}=-1.65\pm 0.06$ (Kulkarni et al. 1999a) or $-1.75\pm 0.11$ (Castro-Tirado et al. 1999) or $-1.8$ (Fruchter et al. 1999). In addition, a radio flare was also detected about 1 day after the burst (Kulkarni et al. 1999b; Galama et al. 1999). A scenario has been proposed to explain these observations. If the burst is assumed to be produced from a jet, the steepening of the late optical afterglow decay is due to the possibility that this jet has undergone the transition from a spherical-like phase to a sideways-expansion phase (Rhoads 1997, 1999; Kulkarni et al. 1999a; Fruchter et al. 1999; Sari, Piran \& Halpern 1999) or that we have observed the edge of the jet (Panaitescu \& M\'esz\'aros 1998; M\'esz\'aros \& Rees 1999). In this {\em Letter} we propose another possible scenario, in which the steepening of the late optical afterglow decay is due to the shock which has evolved from a relativistic phase to a nonrelativistic phase in a dense medium. According to the standard afterglow shock model (for a review see Piran 1998), the afterglow is produced by synchrotron radiation or inverse Compton scattering in the external forward wave (blast wave) of the GRB fireball expanding in a homogeneous medium. The external reverse shock of the fireball may lead to a prompt optical flash (Sari \& Piran 1999). As more and more ambient matter is swept up, the forward shock gradually decelerates and eventually enters a nonrelativistic phase. In the meantime, the emission from such a shock fades down, dominating at the beginning in X-rays and progressively at optical to radio energy band. There are two limiting cases (adiabatic and highly radiative) for the hydrodynamical evolution of the shock. These cases have been well studied both analytically (e.g., M\'esz\'aros \& Rees 1997; Wijers, Rees \& M\'esz\'aros 1997; Waxman 1997a, b; Reichart 1997; Sari 1997; Vietri 1997; Katz \& Piran 1997; M\'esz\'aros, Rees \& Wijers 1998; Dai \& Lu 1998a; Sari, Piran \& Narayan 1998; etc) and numerically (e.g., Panaitescu, M\'esz\'aros \& Rees 1998; Huang et al. 1998; Huang, Dai \& Lu 1998). A partially radiative (intermediate) case has been investigated (Chiang \& Dermer 1998; Cohen, Piran \& Sari 1998; Dai, Huang \& Lu 1999). Here we only consider the limiting cases. In the highly radiative model, since all shock-heated electrons cool faster than the age of the shock, the optical afterglow should have the same temporal decay index as the X-ray afterglow (Sari et al. 1998), incompatible with the observations (Kulkarni et al. 1999a). In the adiabatic model, however, the difference in the decay index between optical and X-ray afterglows is found to be likely $1/4$, which is consistent with the observational result $\Delta\alpha =\alpha_{1R}- \alpha_X\approx 0.3$. This implies that the shock producing the afterglow of GRB 990123 has evolved adiabatically. This is the starting point of our analysis. For an adiabatic shock, the time at which it enters a nonrelativistic phase $\propto n^{-1/3}$, where $n$ is the baryon number density of the medium. Therefore, this time for a shock expanding in a dense medium with density of $n\sim 10^6\,{\rm cm}^{-3}$ is two orders of magnitude smaller than that for a shock with the same energy in a thin medium with density of $n\sim 1\,{\rm cm}^{-3}$. Furthermore, as given in Section 2, the afterglow at the nonrelativistic phase decays faster than at the relativistic phase. It is natural to expect that this effect can provide an explanation for the steepening feature of the afterglow from GRB 990123. Dense media have been discussed in the context of GRBs. First, Katz (1994) suggested collisions of relativistic nucleons with a dense cloud as an explanation of the delayed hard photons from GRB 940217. Second, to explain the radio flare of GRB 990123, Shi \& Gyuk (1999) speculated that a relativistic shock may have ploughed into a dense medium off the line of sight. Third, Piro et al. (1999) and Yoshida et al. (1999) have reported an iron emission line in the X-ray afterglow spectrum of GRB 970508 and GRB 970828 respectively. The observed line intensity requires a dense medium with a large iron mass concentrated in the vicinity of the burst (Lazzati, Campana \& Ghisellini 1999). Finally, dense media (e.g., clouds or ejecta) may appear in the context of some energy source models, e.g., failed supernovae (Woosley 1993), hypernovae (Paczy\'nski 1998), supranovae (Vietri \& Stella 1998), phase transition of neutron stars to strange stars (Dai \& Lu 1998b), baryon decay of neutron stars (Pen \& Loeb 1998), etc. \section{The Evolution of a Shock in a Dense Medium} \subsection{Relativistic Phase} Now we consider an adiabatic relativistic shock expanding in a dense medium. The Blandford-McKee (1976) self similar solution gives the Lorentz factor of the shock, \begin{eqnarray} \gamma & = & \frac{1}{4}\left[ \frac{17E(1+z)^3}{\pi nm_pc^5t_\oplus^3} \right]^{1/8} \nonumber \\ & = & 2E_{54}^{1/8}n_5^{-1/8}t_{\rm day}^{-3/8}[(1+z)/2.6]^{3/8}, \end{eqnarray} where $E=E_{54}\times 10^{54}{\rm ergs}$ is the total isotropic energy, $n_{5}=n/10^5\,{\rm cm}^{-3}$, $t_\oplus=t_{\rm day}\times 1\,{\rm day}$ is the observer's time since the gamma-ray trigger, $z$ is the the redshift of the source generating this shock, and $m_p$ is the proton mass. In analyzing the spectrum and light curve of synchrotron radiation from the shock, one needs to know two crucial frequencies: the synchrotron radiation peak frequency ($\nu_m$) and the cooling frequency ($\nu_c$). In the standard afterglow shock picture, the electrons heated by the shock are assumed to have a power-law distribution: $dN_e/d\gamma_e\propto \gamma_e^{-p}$ for $\gamma_e\ge\gamma_{em}$, where $\gamma_e$ is the electron Lorentz factor and the minimum Lorentz factor $\gamma_{em}=610\epsilon_e\gamma$. The power-law index $p\approx 2.56$ by fitting the spectrum and light curve of the observed afterglow of GRB 990123 (see below). We further assume that $\epsilon_e$ and $\epsilon_B$ are ratios of the electron and magnetic energy densities to the thermal energy density of the shocked medium respectively. Based on these assumptions, the synchrotron radiation peak frequency in the observer's frame can be written as \begin{eqnarray} \nu_m & = & \frac{\gamma\gamma_{em}^2}{1+z}\frac{eB'}{2\pi m_ec} \nonumber \\ & = & 8.0\times 10^{11}\epsilon_e^2\epsilon_{B,-6}^{1/2} E_{54}^{1/2}t_{\rm day}^{-3/2} \nonumber \\ & & \times [(1+z)/2.6]^{1/2}\,\,{\rm Hz}, \end{eqnarray} where $\epsilon_{B,-6}=\epsilon_B/10^{-6}$ and $B'=(32\pi \epsilon_B\gamma^2 nm_pc^2)^{1/2}$ is the internal magnetic field strength of the shocked medium. According to Sari et al. (1998), the cooling frequency, the frequency of electrons with Lorentz factor of $\gamma_c$ that cool on the dynamical time of the shock, is given by \begin{eqnarray} \nu_c & = & \frac{\gamma\gamma_c^2}{1+z}\frac{eB'}{2\pi m_ec} = \frac{18\pi em_ec(1+z)}{\sigma_T^2B'^3\gamma t_\oplus^2} \nonumber \\ & = & 1.9\times 10^{16}\epsilon_{B,-6}^{-3/2}E_{54}^{-1/2} n_5^{-1}t_{\rm day}^{-1/2}\nonumber \\ & & \times [(1+z)/2.6]^{-1/2}\,\,{\rm Hz}, \end{eqnarray} where $\sigma_T$ is the Thompson scattering cross section. From equations (2) and (3), Sari et al. (1998) have further defined two critical times, when the breaking frequencies $\nu_m$ and $\nu_c$ cross the observed frequency $\nu=\nu_{15}\times 10^{15}\,{\rm Hz}$: $t_m=8.6\times 10^{-3}\epsilon_e^{4/3}\epsilon_{B,-6}^{1/3} E_{54}^{1/3}[(1+z)/2.6]^{1/3}\nu_{15}^{-2/3}\,{\rm days}$, and $t_c=380\epsilon_{B,-6}^{-3}E_{54}^{-1} n_5^{-2}[(1+z)/2.6]^{-1}\nu_{15}^{-2}\,{\rm days}$. Therefore we see that for $E_{54}\sim 1.6$, $\epsilon_e\sim 0.1$, $\epsilon_{B,-6}\sim 0.02$, and $n_5\sim 30$ inferred in the next section, the optical afterglow in several days after the burst should result from those slowly-cooling electrons and the X-ray afterglow from those fastly-cooling electrons. The observed synchrotron radiation peak flux can be obtained by \begin{eqnarray} F_{\nu_m} & = & \frac{N_e\gamma P'_{\nu_m}(1+z)}{4\pi D_L^2} \nonumber \\ & = & 4.2\epsilon_{B,-6}^{1/2}E_{54}n_5^{1/2}[(1+z)/2.6] D_{L,28}^{-2}\,\,{\rm Jy}, \end{eqnarray} where $N_e$ is the total number of swept-up electrons, $P'_{\nu_m} =m_ec^2\sigma_TB'/(3e)$ is the radiated power per electron per unit frequency in the frame comoving with the shocked medium, and $D_L=D_{L,28}\times 10^{28}\,{\rm cm}$ is the distance to the source. In the light of equations (2)-(4), one can easily find the spectrum and light curve of the afterglow, \begin{equation} F_\nu=\left \{ \begin{array}{llll} (\nu/\nu_m)^{-(p-1)/2}F_{\nu_m}\\ \,\,\,\,\,\,\,\, \propto \nu^{-(p-1)/2} t_\oplus^{3(1-p)/4}\,\,\,\,\,\, {\rm if}\,\, \nu_m<\nu<\nu_c; \\ (\nu_c/\nu_m)^{-(p-1)/2}(\nu/\nu_c)^{-p/2}F_{\nu_m} \\ \,\,\,\,\,\,\,\, \propto\nu^{-p/2}t_\oplus^{(2-3p)/4} \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, {\rm if}\,\, \nu>\nu_c, \end{array} \right. \end{equation} where the low-frequency radiation component has not been considered (Sari et al. 1998). In the GRB 990123 case, we require $\nu_m<\nu<\nu_c$ for the optical afterglow and $\nu>\nu_c$ for the X-ray afterglow. Thus, the R-band afterglow decay index $\alpha_R=3(1-p)/4$ and the X-ray decay index $\alpha_X= (2-3p)/4$, which are well consistent with the observational results $\alpha_{1R}=1.1\pm 0.03$ and $\alpha_X=-1.44\pm 0.07$ if $p\approx 2.56$. \subsection{Nonrelativistic Phase} As it sweeps up sufficient ambient matter, the shock will eventually go into a nonrelativistic phase. During such a phase, the shock's velocity $v\propto t_\oplus^{-3/5}$, its radius $r\propto t_\oplus^{2/5}$, the internal field strength $B'\propto t_\oplus^{-3/5}$ and the typical electron Lorentz factor $\gamma_{em}\propto t_\oplus^{-6/5}$. Thus, we obtain the synchrotron peak frequency $\nu_m\propto \gamma_{em}^2B' \propto t_\oplus^{-3}$, the cooling frequency $\nu_c\propto B'^{-3} t_\oplus^{-2}\propto t_\oplus^{-1/5}$ and the peak flux $F_{\nu_m}\propto N_eP'_{\nu_m}\propto r^3B'\propto t_\oplus^{3/5}$. According to these scaling laws, we further derive the spectrum and light curve at the nonrelativistic stage: \begin{equation} F_\nu=\left \{ \begin{array}{llll} (\nu/\nu_m)^{-(p-1)/2}F_{\nu_m} \\ \,\,\,\,\, \propto \nu^{-(p-1)/2} t_\oplus^{(21-15p)/10}\,\,\,\, {\rm if}\,\, \nu_m<\nu<\nu_c; \\ (\nu_c/\nu_m)^{-(p-1)/2}(\nu/\nu_c)^{-p/2}F_{\nu_m} \\ \,\,\,\,\, \propto\nu^{-p/2}t_\oplus^{(4-3p)/2} \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, {\rm if}\,\, \nu>\nu_c. \end{array} \right. \end{equation} From equation (6), we can see the R-band decay index $\alpha_R=(21-15p)/10$ for radiation from slowly-cooling electrons or $\alpha_R=(4-3p)/2$ for radiation from rapidly-cooling electrons. If $p\approx 2.56$, then $\alpha_R \approx -1.74$ or $-1.84$, in excellent agreement with the observations in the time interval of 2.5 days to 20 days after the burst (Kulkarni et al. 1999a; Fruchter et al. 1999; Castro-Tirado et al. 1999). \section{Constraints on Parameters} In the above section, we show that as an adiabatic shock expands in a dense medium from an ultrarelativistic phase to a nonrelativistic phase, the decay of the radiation from such a shock will steepen. This effect may fit the observed steepening better than the alternative interpretation --- jet sideways expansion. In the latter interpretation, the temporal decay of a late afterglow is very likely to be $\propto t_\oplus^{-p}$ (Rhoads 1997, 1999; Sari et al. 1999). We further analyze our effect and infer some parameters of the model. According to the analysis on the R-band light curve of the GRB 990123 afterglow (Kulkarni et al. 1999a; Fruchter et al. 1999; Castro-Tirado et al. 1999), the observed break occurred at $t_\oplus = 2.04\pm 0.46\,$days. This implies $\gamma\sim 1$ at $t_{\rm day}\approx 2.5$. From equation (1), therefore, we find $n_5\sim 16E_{54}$, where the redshift $z=1.6$ has been used. We now continue to consider two observational results. First, on January 23.577 UT, the Palomar 60-inch telescope detected the R-band magnitude $R=18.65\pm 0.04$, corresponding to the flux $F_R\sim 100\,\mu{\rm Jy}$ at $t_{\rm day}\approx 0.17$ (Kulkarni et al. 1999a). Considering this result in equation (5) together with equations (2) and (4), we can derive \begin{equation} \epsilon_e^{p-1}\epsilon_{B,-6}^{(p+1)/4}E_{54}^{(p+3)/4} n_5^{1/2} \sim 0.01, \end{equation} where the right number has been obtained by taking $p\approx 2.56$ and $D_{L,28}\sim 3.7$. Second, on January 24.65 UT, the BeppoSAX observed the X-ray (2-10\,keV) flux $F_X\sim 5\times 10^{-2}\,\mu$Jy (Heise et al. 1999a, b). Combining this result with equations (2)-(5), we can also derive \begin{equation} \epsilon_e^{p-1}\epsilon_{B,-6}^{(p-2)/4}E_{54}^{(p+2)/4}\sim 0.03. \end{equation} Since $E_{54}\sim 1.6$ (Briggs et al. 1999; Kulkarni et al. 1999a), the medium density $n_5 \sim 30$ and the solution of equations (7) and (8) is $\epsilon_e\sim 0.1$ and $\epsilon_{B,-6}\sim 0.02$. Our inferred value of $\epsilon_e$ is near the equipartition value, in agreement with the result of Wijers \& Galama (1998) and Granot, Piran \& Sari (1998), while our $\epsilon_B$ is about six orders of magnitude smaller than the value inferred from the afterglow of GRB 970508. Of course, the field density for GRB 971214 has been estimated to be less than $10^{-5}$ times the equipartition value (Wijers \& Galama 1998). As suggested by Galama et al. (1999), such differences in field strength may reflect differences in energy flow from the central engine. \section{Discussion and Conclusion} In the above section, we find the medium density $n\sim 3\times 10^6\,{\rm cm}^{-3}$ for our model to fit the observed optical and X-ray afterglow of GRB 990123. Now we show that even in the the presence of such a dense medium, the optical and X-ray radiations from the forward shock were neither self absorbed in the shocked medium nor scattered in the unshocked medium. First, the self-absorption frequency of the shocked medium is (Wijers \& Galama 1998; Granot et al. 1998) $\nu_a\sim 10^3\,{\rm GHz}(\epsilon_e/0.1)^{-1}(\epsilon_{B,-6} /0.01)^{1/5}E_{54}^{1/5}(n_5/10)^{3/5}$. This estimate should be the upper limit because of the presence of a possible low-energy electron population (Waxman 1997b). Clearly, $\nu_a$ is much less than the optical frequency, implying that the self absorption in the shocked medium didn't affect the optical and X-ray afterglow. In fact, this estimate is valid only for $\nu_a < \nu_m$. When $\nu_a>\nu_m$, $\nu_a$ must have decayed. As a result, the flux at 8.46 GHz first increased as $t_\oplus^{1.25}$ and then declined as $t_\oplus^{-1.74}$ for $\nu_a<8.46$ GHz during the nonrelativistic phase. This might provide an explanation for the observed radio flare. Second, a photon emitted from the shock may be scattered by the electrons in the unshocked medium. The scattering optical depth $\tau\sim\sigma_TnR$ (where $R$ is the typical radius of the medium). If the medium was distributed isotropically and homogeneously and its mass $M\sim 10M_\odot$ (the typical mass of a supernova ejecta), then $\tau\sim 0.05(M/10M_\odot)^{1/3}(n_5/10)^{2/3}\ll 1$. This implies that the afterglow from the shock was hardly affected by the medium. For other well-studied afterglows, e.g., GRB 970228 and GRB 970508, their ambient densities must be very low for three reasons: (i) In these bursts there was no observed break in the optical light curve as long as the afterglow could be observed (Fruchter et al. 1998; Zharikov et al. 1998). (ii) The fluctuation appearing in the radio afterglow light curve of GRB 970508 requires the shock had been relativistic for several weeks (Waxman, Kulkarni \& Frail 1998). (iii) The analysis of the afterglow spectrum of GRB 970508 leads to a low ambient density $n<10\,{\rm cm}^{-3}$ (Wijers \& Galama 1998; Granot et al. 1998). However, the observed iron emission line in the X-ray afterglow spectrum of GRB 970508 indeed requires a dense medium with density $\sim 10^9\,{\rm cm}^{-3}$ (Lazzati et al. 1999). The only way to reconcile a monthly lasting power-law afterglow with iron line emission is through a particular geometry, in which the line of sight is devoid of the dense medium. In contrast to this idea, we suggest that for GRB 990123 a dense medium of $n\sim 3\times 10^6\,{\rm cm}^{-3}$ appears at least at the line of sight or perhaps isotropically. How was the dense medium produced? One possibility was a cloud and another possibility was an ejecta from the GRB site. There have been several source models (mentioned in Introduction) in the literature which may lead to massive ejecta. Here we want to discuss one of them in detail. Timmes, Woosley \& Weaver (1996) showed that Type II supernovae may produce a kind of neutron star with $\sim 1.73M_\odot$. If these massive neutron stars have very short periods at birth, they may subsequently convert into strange stars due to rapid loss of angular momenta (Cheng \& Dai 1998), and perhaps the strange stars are differentially rotating (Dai \& Lu 1998b). Even though this model is somewhat similar to the supranova model of Vietri \& Stella (1998), resultant compact objects are strange stars in our model and black holes in the supranova model. We further discuss implications of our model. First, the model leads to low-mass loading matter because of thin baryonic crusts of the strange stars. Second, such stars result in GRBs with spiky light curves, being consistent with the analytical result from the observed data of GRB 990123 (Fenimore, Ramirez-Ruiz \& Wu 1999). The third advantage of this model is to be able to explain well the property of the early afterglow of GRB 970508 by considering energy injection from the central pulsar (Dai \& Lu 1998b, c). Finally, a dense medium, the supernova ejecta, appears naturally. Our scenario proposed in this {\em Letter} requires a dense medium with density $\sim 3\times 10^6\,{\rm cm}^{-3}$ to explain the steepening in the temporal decay of the R-band afterglow about 2.5 days after GRB 990123. We also suggest that this medium could be a supernova/supranova/hypernova ejecta. Thus, if the mass of the medium is assumed to be $M\sim 10M_\odot$, its radius can be estimated to be $R\sim 3\times 10^{17}\,{\rm cm} (M/10M_\odot)^{1/3}(n_5/10)^{-1/3}$. According to equation (1), we can integrate $dr=2\gamma^2 cdt_\oplus$ and thus find that the postburst 2.5-day time in the observer's frame corresponds to about 20 days in the unshocked medium's frame. This implies that the radius at which the shock entered a nonrelativistic phase is about $5\times 10^{16}$ cm. This radius is much less than that of the medium. Therefore, the medium discussed here was so wide and dense that the ultrarelativistic shock must have become nonrelativistic about 2.5 days after the burst. In summary, a simple explanation for the ``steepening'' observed in the temporal decay of the late R-band afterglow of GRB 990123 is that a shock expanding in a dense medium with density of $\sim 3\times 10^6\, {\rm cm}^{-3}$ has evolved from a relativistic phase to a nonrelativistic phase. We find that this scenario not only explains well the optical afterglow but also accounts for the observed X-ray afterglow quantatitively. \acknowledgments We would like to thank J. I. Katz, S. R. Kulkarni, A. Mitra and the anonymous referee for invaluable suggestions, and Y. F. Huang and D. M. Wei for helpful discussions. This work was supported by the National Natural Science Foundation of China (grants 19825109 and 19773007).
\section{Introduction} In the Standard Model the charged current interactions of the tau lepton are mediated by the $W$ boson with a pure $V\!-\!A$ coupling. We consider new derivative couplings in the Hamiltonian which are parametrised by the parameters $\kappa$ and ${\tilde{\kappa}}$, the (CP-conserving) magnetic and (CP-violating) electric dipole form factors respectively~\cite{RIZZO97A,CHIZHOV96A}. These are the charged current analogues of the weak neutral current dipole moments, measured using $Z\rightarrow\tau^+\tau^-$ events~\cite{PICH97A}, and the electromagnetic dipole moments~\cite{BIEBEL96A,TTGNUCPHYSB}, measured using $Z\rightarrow\tau^+\tau^-\gamma$ events~\cite{OPALTTG,L3TTG,TAYLOR_TAU98}. The only limits so far obtained for $\kappa$ and $\tilde{\kappa}$ are derived from analyses of the partial widths for $\tau^-\rightarrow \ell^- \bar{\nu}_{\ell}\nu_{\tau}$, for $\ell={\mathrm{e}}, \mu$~\cite{RIZZO97A,ANOMALOUS_COUPLINGS,DOVA_SWAIN_TAYLOR_TAU98}. In this paper we consider for the first time the effects of anomalous charged current dipole moments on tau decays involving hadrons. We analyse the $\tnpp$ process which has largest branching fraction of all the tau decay modes. This process is particularly topical due to a recently reported difference of $2.2\sigma$ between the measured $\tnpp$ branching fraction and the (lower) value predicted using $e^+e^- \rightarrow \pi^+\pi^-$ data in the neighbourhood of the $\rho$ meson resonances and the Conserved Vector Current (CVC) hypothesis~\cite{TAU98_EIDELMAN}. While this could be attributed to a fluctuation, we note that non-zero values of $\kappa$ and $\tilde{\kappa}$ would also yield a higher measured value for $\BR(\tnpp)$. We present predictions for the differential $\tnpp$ decay distributions and the partial width, $\Gamma(\tnpp)$, as functions of $\kappa$ and ${\tilde{\kappa}}$. The sensitivity of the differential distributions is analysed for typical samples of $\tnpp$ decays in ${\mathrm{e^+e^-\rightarrow\tau^+\tau^-}}$ events which are reconstructed by the LEP and SLC experiments. The partial width is compared to the experimental measurements of $\BR(\tnpp)$ to yield quantitative constraints on $\kappa$ and ${\tilde{\kappa}}$. \section{Parametrisation of anomalous couplings in ${\mathbf{\tnpp}}$ decays} The matrix element for the decay $\tnpp$ is given by \begin{equation} M = \frac {\GF}{\sqrt{2}} V_{ud} J^\mu H_\mu, \end{equation} where $\GF = \GFMUval$ is the Fermi constant, $V_{ud} = \VUDval$ is the appropriate CKM matrix element~\cite{PDG98}, and $J^\mu$ and $H_\mu$ are the leptonic and hadronic currents respectively. The effects of anomalous weak charged current dipole moment couplings at the $\tau\nu_\tau W$ vertex are parametrised by augmenting the usual $V-A$ charged current such that $J^\mu$ is given by \begin{equation} J^\mu = \bar{u}_\nu \left( \gamma^\mu \left( 1-\gamma^5 \right) - \frac{i\sigma^{\mu\nu} q_\nu}{2m_\tau} (\kappa - i\tilde{\kappa}\gamma_5) \right) u_\tau, \label{current}\\ \end{equation} where $\sigma^{\mu\nu} = i /2[\gamma^\mu,\gamma^\nu]$, $q^\mu$ is the four-momentum transfer, and $m_\tau = \MTval$~\cite{PDG98} is the tau mass. The parameters $\kappa$ and ${\tilde{\kappa}}$ are in general complex but henceforth we assume that ${\tilde{\kappa}}$ is real, as required by $CPT$ invariance. The hadronic current is parametrised as \begin{equation} H^\mu = \sqrt{2} F(q^2) (q_1 - q_2)^\mu, \end{equation} where $q_i$ denote the four-momenta of the two final-state \mbox{pions} and $F(q^2)$ is a form factor. A convenient choice for the kinematic observables, following K{\"{u}}hn and Mirkes~\cite{KUHN92A}, is: $q^2$, the invariant mass-squared of the hadronic system; $\cos\theta$, the cosine of the angle between the tau spin-vector and the hadronic centre-of-mass direction as seen in the tau rest frame; and $\cos\beta$, the cosine of the angle between the charged pion and the axis pointing in the direction of the laboratory viewed from the hadronic centre-of-mass frame (henceforth referred to as the $z$-axis). \subsection{Differential Decay Rate, ${\mathbf{d\Gamma(\tnpp)/dq^2 d\cos\theta d\cos\beta}}$} After integration over the unobservable neutrino direction and the azimuthal angle of the charged pion, and neglecting the mass difference between the charged and neutral pions, the differential decay rate is given by \begin{eqnarray} d\Gamma & = & \frac{1}{(4\pi)^3} \frac{G_F^2}{4 m_\tau^3} |V_{ud}|^2 |F(q^2)|^2 S_{EW} \nonumber \\ & & \times (q^2-4 m_\pi^2)^{3/2} (m_\tau^2-q^2)^2 \nonumber \\ & & \times \left\{ \left[ f_0+Re(\kappa) f_1+(|\kappa|^2+\tilde{\kappa}^2) f_2 \right] \right. \nonumber \\ & & \left. \mbox{~~~}+ P_\tau \left[ g_0+Re(\kappa) g_1+\tilde{\kappa} Im(\kappa) g_2 \right] \right\} \nonumber \\ & & \times \frac{dq^2}{\sqrt{q^2}} \frac{d\cos\beta}{2} \frac{d\cos\theta}{2} \;, \label{E1} \end{eqnarray} where $P_\tau$ is the tau polarisation and the factor of $S_{EW} = 1.0194$ accounts for electroweak corrections to leading logarithm~\cite{MARCIANO88A}. The functions $f_i$ and $g_i$ $(i=0,1,2)$ are given by \begin{eqnarray} f_0 & = & 2 \left[ 1+\frac{m_\tau^2-q^2}{q^2} Y \right], \\ f_1 & = & 1, \\ f_2 & = & 1/4 \left[ 1-\frac{m_\tau^2-q^2}{m_\tau^2} Y \right], \\ g_0 & = & 2 \left[ \frac{2 m_\tau}{\sqrt{q^2}} X - \cos\theta \right], \\ g_1 & = & \frac{m_\tau^2+q^2}{m_\tau \sqrt{q^2}} X - \left[ 1+\frac{(m_\tau^2-q^2)^2}{2 m_\tau^2 q^2} Y \right] \cos\theta, \\ g_2 & = & \frac{\cos\theta}{2} - \frac{\sqrt{q^2}}{m_\tau} X, \end{eqnarray} where \begin{eqnarray} Y & = & \frac{1}{3} \left[ 1+(3\cos^2\psi-1) \frac{3\cos^2\beta-1}{2} \right], \\ X & = & \frac{m_\tau^2+q^2}{2 m_\tau \sqrt{q^2}} Y \cos\theta + \sin\theta \frac{\sin 2\psi}{2} \frac{3\cos^2\beta-1}{2}, \end{eqnarray} and $\psi$ is the angle between the tau direction of flight in the hadronic rest frame and the $z$-axis. At LEP energies the following approximation is valid: \begin{eqnarray} \cos\psi & = & \frac{\eta + \cos\theta}{1+\eta \cos\theta},~~~{\mathrm{with}} \\ \eta & = & \frac{m_\tau^2-q^2}{m_\tau^2+q^2}. \end{eqnarray} $F(q^2)$ describes the resonant structure of the two-pion invariant mass and the model used to describe it is discussed in more detail in the following section. \subsection{Dependence of Apparent Polarisation on Anomalous Couplings} The usual determination of tau polarisation from energy and angular distributions of decay products of the tau depends crucially on the assumed $V-A$ structure of the charged current to serve as a polarimeter. Additional couplings in $\tau\rightarrow\rho\nu$ decays will produce measured values of polarisation which differ from analyses of other $\tau$ decay modes and the predictions from global fits to Electroweak parameters in the context of the Standard Model. The observed agreement of polarisation measured in $\tnpp$ decays with other determinations may be used to constrain $\kappa$ and $\tilde\kappa$. We first integrate the differential width presented above with respect to $q^2$ and $\cos\beta$. The $q^2$ dependence of $F$ must be explicitly considered prior to this integration. $F(q^2)$ for the $\tnpp$ channel is dominated by the $\rho(770)$ vector meson with a small admixture of $\rho^\prime(1450)$ and a negligible contribution from the $\rho^{\prime\prime}(1700)$, as verified by the ALEPH experiment~\cite{ALEPHRHOFIT}. We work within the context of the K{\"{u}}hn and Santamaria model~\cite{KUHN90A} in which the $\rho$ and $\rho^\prime$ resonances are each described by a Breit-Wigner propagator with an energy-dependent width~\cite{DECKER93A} \begin{equation} B_x(q^2) = \frac{m_x^2}{m_x^2-q^2-i \sqrt{s} \Gamma_x(q^2)}, \end{equation} where: \begin{eqnarray} \Gamma_x(q^2) & = & \Gamma^0_x \frac{m_x^2}{q^2} \left( \frac{p(q^2)} {p(m_x^2)} \right)^3; \qquad {\mathrm{and}} \nonumber\\ p(s) & = & 1/2 \sqrt{s-4 m_\pi^2}. \end{eqnarray} where $\Gamma^0_x$ is a constant. The normalisation is fixed by chiral symmetry constraints in the limit of soft meson momenta, such that the form factor is given by \begin{equation} F(q^2) = \frac{B_\rho(q^2)+\beta B_{\rho^\prime}(q^2)}{1+\beta} \;. \label{fks} \end{equation} where $\beta = -0.145$ \cite{KUHN90A}. The differential width, retaining only the $\theta$ dependence, is of the form \begin{equation} d\Gamma = ( A + B P_\tau \cos\theta ) \frac{d\cos\theta}{2}. \label{dcostheta} \end{equation} The coefficients $A$ and $B$ depend on $\kappa$ and $\tilde{\kappa}$ and are given by \begin{eqnarray} A & = & f^\dag_0+Re(\kappa) f^\dag_1+(|\kappa|^2+\tilde{\kappa}^2) f^\dag_2 \nonumber\\ B & = & g^\dag_0+Re(\kappa) g^\dag_1+\tilde{\kappa} Im(\kappa) g^\dag_2 \end{eqnarray} where, for $\beta = -0.145$, we obtain the following numerical values \begin{eqnarray} f^\dag_0 & = (518.6 \pm 6.4 ) \cdot 10^{-15}\,{\mathrm{GeV}} \\ f^\dag_1 & = (111.3 \pm 1.9 ) \cdot 10^{-15}\,{\mathrm{GeV}} \\ f^\dag_2 & = (20.74 \pm 0.38 ) \cdot 10^{-15}\,{\mathrm{GeV}} \\ g^\dag_0 & = (221.6 \pm 2.04 ) \cdot 10^{-15}\,{\mathrm{GeV}} \\ g^\dag_1 & = (-37.11 \pm .63 ) \cdot 10^{-15}\,{\mathrm{GeV}} \\ g^\dag_2 & = (32.73 \pm 0.50 ) \cdot 10^{-15}\,{\mathrm{GeV}} \end{eqnarray} Comparison of the results for the measured ``apparent polarisation'' from $\tnpp$ channel with other channels or the results of Electroweak fits would permit constraints to be placed on $\kappa$ and $\tilde{\kappa}$. \subsection{Dependence of the Total Width, ${\mathbf{\Gamma(\tnpp)}}$ on Anomalous Couplings} Integration of Eq.~\ref{dcostheta} over $\cos\theta$ yields the effect of the anomalous couplings on the total rate \begin{equation} \Gamma(\tnpp) = \Gamma^0 \left[1 + a_1 Re(\kappa) + a_2 (|\kappa|^2+\tilde{\kappa}^2) \right], \label{gammarho} \end{equation} which naturally is independent of the polarisation (the polarisation term is proportional to $\cos\theta$ and therefore integrates to zero). The parameters $a_1$ and $a_2$ are given by \begin{eqnarray} a_1 & \equiv f_1^\dag/f_0^\dag & = 0.202; \qquad {\mathrm{and}} \\ a_2 & \equiv f_2^\dag/f_0^\dag & = 0.037. \end{eqnarray} $\Gamma^0 (\equiv f_0^\dag)$ represents the partial width in the absence of anomalous couplings, {\em{i.e.}} $\kappa = \tilde{\kappa} = 0$. \section{Sensitivity of the differential decay rate to $\kappa$ and ${\tilde{\kappa}}$} The sensitivity of Eq.~\ref{E1} to the dipole moment couplings was studied for the case of CP-conserving interaction, i.e. $\kappa$ real and $\tilde{\kappa}=0$. We consider the quantity \begin{equation} \sigma \sqrt{N} = \left[ \int d\Omega \frac{1}{f} \left(\frac{\partial f}{\partial \kappa}\right)^2\right]^{-1/2} \end{equation} as a function of the $\tau$ polarisation, where $\sigma$ is the expected error one standard deviation on $\kappa$, $N$ is the number of $\tnpp$ decays, and $d\Omega$ is the elemental phase space volume. The choice of the quantity $\sigma \sqrt{N}$ simply reflects the $1/\sqrt{N}$ dependence of the statistical error $\sigma$. The distribution $f$ is given by Eq.~\ref{E1} for the $\tnpp$ channel and Eq.~11 of Rizzo~\cite{RIZZO97A} for $\tnnl$. Figure~\ref{F3} shows $\sigma \sqrt{N}$ as a function of $P_\tau$ for the particular case $\kappa\approx 0$ for (a) the $\tnpp$ decay mode, and (b) the $\tnnl$ decay mode ($\ell = {\mathrm{e}}$ or $\mu$, not both combined). \begin{figure}[!htb] \begin{center} \epsfig{file=sensi.eps,width=0.95\linewidth} \end{center} \caption{The $\kappa$ sensitivity quantity, $\sigma \sqrt{N}$, as a function of $P_\tau$ for $\kappa\approx 0$, for (a) the $\tnpp$ decay mode and (b) the $\tnnl$ decay mode.} \label{F3} \end{figure} The $\tnpp$ mode is intrinsically more sensitive than the leptonic decay modes and is less dependent on the $\tau$ polarisation. In addition, the branching fraction for $\tnpp$ is larger than each leptonic channel. For example, at the Z peak ($P_\tau \approx -0.15$) with a typical sample of reconstructed decays for each LEP experiment ($\sim 45\,000$ $\tnpp$ decays and $\sim 30\,000$ $\tnnl$ decays) the predicted statistical errors are: \begin{eqnarray} \sigma_\rho & \sim & 0.020 , {\mathrm{and}} \\ \sigma_\ell & \sim & 0.065 , \end{eqnarray} where detector effects are neglected, apart from their influence on the reconstruction efficiency which is reflected in the number of decays assumed. The corresponding statistical error for the combined e and $\mu$ channels is $\sim$0.046 which is more than a factor of two less precise than from the $\tnpp$ channel alone. A practical disadvantage of the semileptonic decay is the multi-dimensional character of the distribution function. In the analysis of the tau polarisation, this problem has been overcome using a single ``optimal variable''~\cite{DAVIER93A}. Although the optimal variable for the tau polarisation is not the optimal variable for $\kappa$ (nor for $\tilde{\kappa}$), we find it still provides distinguishing power. We fit hypothetical distributions of the optimal variable for simulated samples of 45\,000 $\tnpp$ decays each, generated with $P_\tau=-0.15$ to represent $\tau$'s produced at the Z peak. Typical errors are $\sigma_\rho \sim 0.038$ which is degraded compared to the full multi-dimensional fit but is still statistically more sensitive than the combined e and $\mu$ channels. The apparent disadvantage of the leptonic channels is, however, mitigated by the need to know $F(q^2)$ for the $\tnpp$ channel which has a non-negligible systematic error, as discussed below. In this paper we cannot derive constraints on $\kappa$ and ${\tilde{\kappa}}$ from fits to the differential decay distributions due to a lack of the necessary experimental information. We can, however, determine constraints from the (intrinsically less sensitive) measurements of $\BR(\tnpp)$, as described below. \section{Constraints on $\kappa$ and ${\tilde{\kappa}}$ from $\BR(\tnpp)$} We derive quantitative constraints on $\kappa$ and ${\tilde{\kappa}}$ by considering the likelihood for the theoretical prediction for $\BR(\tnpp)$ to agree with the experimentally determined average value of \begin{equation} \BR(\tnpp) = \BRTRval, \label{brexpt} \end{equation} as a function of $\kappa$ and $\tilde{\kappa}$. $\Gamma^0$ of Eq.~\ref{gammarho} is determined from $e^+e^- \rightarrow \pi^+\pi^-$ data using CVC. A combined analysis of all data, allowing for radiative corrections and $\rho-\omega$ interference, yields the CVC prediction of~\cite{TAU98_EIDELMAN} \begin{equation} \BR(\tnpp) = (24.52 \pm 0.33)\%, \label{brcvc} \end{equation} where the error includes statistical and systematic uncertainties and conservatively allows for a possible systematic discrepancy of the DM1 data compared to CMD, CMD-2, and OLYA. The experimental value of $\BR(\tnpp)$ is higher than the CVC prediction by $2.2$ standard deviations of the combined error. We fix $\Gamma^0$ to the CVC prediction of $\BR(\tnpp)$ so that only $a_1$ and $a_2$ depend on the description of hadronic spectral function. This reduces the sensitivity of the results to the details of the hadronic modelling. We construct likelihoods as a function of $\kappa$ and $\tilde{\kappa}$ conservatively assuming in each case that the other parameter is zero. The errors are propagated numerically~\cite{NIM_LIKELIHOOD_PAPER} taking into account the error on the experimental measurement of $\BR(\tnpp)$ (Eq.~\ref{brexpt}), the uncertainty on the CVC prediction (Eq.~\ref{brcvc}), a systematic error of $0.5$\% for radiative corrections not included in $S_{EW}$~\cite{MARCIANO88A}, and a systematic error of 0.3\% for the effect of $\rho-\omega$ interference~\cite{TAU96_EIDELMAN}. Figure~\ref{fig:probs} shows the likelihood distributions for (a) $\kappa$ and (b) $\tilde{\kappa}$. The distribution for $\kappa$ has a single peak due to the dominance of the term linear in $\kappa$ in Eq.~\ref{gammarho}. The distribution for $\tilde{\kappa}$ has two symmetric peaks due to the lack of a term linear in $\tilde{\kappa}$ in Eq.~\ref{gammarho}, therefore it is more appropriate to constrain the quantity $|\tilde\kappa|$. We determine \begin{eqnarray} \kappa & = & \RHOKAPPAMval ,~~~{\mathrm{and}} \\ |\tilde\kappa| & = & \RHOKAPPAEval , \end{eqnarray} where the errors correspond to the 68\% confidence level. At the 95\% confidence level we constrain $\kappa$ and $\tilde\kappa$ to the ranges: \begin{eqnarray} & \RHOKAPPAMlim &;~~~{\mathrm{and}} \\ & \RHOKAPPAElim &. \end{eqnarray} These results are slightly more than two standard deviations from the SM expectations of zero which, though intriguing, cannot be considered statistically compelling evidence of new physics. \begin{figure}[!htb] \begin{center} \epsfig{file=probs.eps,width=0.9\linewidth} \end{center} \caption{The likelihood distributions for (a) $\kappa$ and (b) $\tilde{\kappa}$.} \label{fig:probs} \end{figure} The results may be compared to complementary results previously obtained from purely leptonic tau decays~\cite{ANOMALOUS_COUPLINGS}, which are $\kappa = {\KAPPAMval}$ and $\tilde\kappa = {\KAPPAEval}$ or ${\KAPPAMlim}$ and ${\KAPPAElim}$ at the 95\% C.L.~\cite{ANOMALOUS_COUPLINGS}. These are more restrictive than those we obtain from $\tnpp$ decays. This is partly due to the larger uncertainties in the theoretical and experimental values of the $\tnpp$ branching fractions. In principle these effects could be counteracted by a higher intrinsic sensitivity of the $\tnpp$ channel due to larger values of $a_1$ and $a_2$ relative to the leptonic modes. From our calculations, however, we see in retrospect that the numerical values for $a_1$ and $a_2$ are smaller than their leptonic counterparts (0.5 and 0.1 respectively). Therefore, if only the total rate information is used the $\tnpp$ channel is less sensitive than the leptonic channels, in contrast to the higher statistical sensitivity of the $\tnpp$ channel when the differential decay distribution is analysed. \section{Summary} We present calculations of the differential and total decay rates for the process $\tnpp$, allowing for anomalous charged current magnetic and electric dipole moments, $\kappa$ and $\tilde\kappa$ respectively. This constitutes the first such analysis of a hadronic tau decay mode. The analysis of the differential distributions for the $\tnpp$ decay mode is found to be statistically more sensitive than the corresponding analysis of purely leptonic modes, $\tnnl$, irrespective of the tau polarisation. The branching fraction, $\BR(\tnpp)$, is also sensitive to $\kappa$ and $\tilde\kappa$ although less so than for the leptonic branching fractions. By comparison of the measured value of $\BR(\tnpp)$ with the predictions of CVC, we determine $\kappa = \RHOKAPPAMval$ and $|\tilde\kappa| = \RHOKAPPAEval$. which differ from the SM expectations by approximately two standard deviations. The values for $\kappa$ and $\tilde\kappa$ obtained from $\BR(\tnnl)$ are consistent with zero. This could mean that the measured result for $\BR(\tnpp)$ and CVC differ only due to a fluctuation, or that there is a theoretical or experimental uncertainty which is not correctly taken into account. The new $e^+e^- \rightarrow \pi^+\pi^-$ data in the neighbourhood of the $\rho$ meson resonances should reduce the experimental uncertainty in the CVC prediction by a factor of almost two in 1999~\cite{TAU98_EIDELMAN}. Hopefully these data will clarify whether this is a statistical or systematic effect or the first indication of some new physics which affects hadronic tau decays but not purely leptonic decays. \section*{Acknowledgements} We acknowledge gratefully the financial support of CONICET and the Fundaci{\'{o}}n Antorchas, Argentina (M.T.D. and P.L.) and the NSF, USA (J.S. and L.T.).
\section{Introduction} Besides the recent high precision measurements of the $W$ mass~\cite{Karlen98,Dorigo98}, $M_W$, the most important input into precision tests of electroweak theory continues to come from the $Z$ factories LEP~1~\cite{Karlen98} and SLC~\cite{Baird98}. The vanguard of the physics program at LEP~1 is the analysis of the $Z$ lineshape. Its parameters are the $Z$ mass, $M_Z$, the total $Z$ width, $\Gamma_Z$, the hadronic peak cross section, $\sigma_{\rm had}$, and the ratios of hadronic to leptonic decay widths, $R_\ell = {\Gamma({\rm had})\over \Gamma(\ell^+\ell^-)}$, where $\ell = e$, $\mu$, or $\tau$. They are determined in a common fit with the leptonic forward-backward (FB) asymmetries, $A_{FB} (\ell) = {3\over 4} A_e A_\ell$. With $f$ denoting the fermion index, \begin{equation} A_f = {2 v_f a_f\over v_f^2 + a_f^2} \end{equation} is defined in terms of the vector ($v_f = I_{3,f} - 2 Q_f \sin^2 \theta_f^{\rm eff}$) and axial-vector ($a_f = I_{3,f}$) $Zf\bar{f}$ coupling; $Q_f$ and $I_{3,f}$ are the electric charge and third component of isospin, respectively, and $\sin^2 \theta_f^{\rm eff} \equiv \bar{s}^2_f$ is an effective mixing angle. The polarization of the electron beam at the SLC allows for competitive and complementary measurements with a much smaller number of $Z$'s than at LEP. In particular, the left-right (LR) cross section asymmetry, $A_{LR} = A_e$, represents the most precise determination of the weak mixing angle by a single experiment (SLD).~\cite{Baird98} Mixed FB-LR asymmetries, $A^{FB}_{LR} (f) = {3\over 4} A_f$, single out the final state coupling of the $Z$ boson. For several years there has been an experimental discrepancy at the $2 \sigma$ level between $A_\ell$ from LEP and the SLC. With the 1997/98 high statistics run at the SLC, and a revised value for the FB asymmetry of the $\tau$ polarization, ${\cal P}^{FB}_\tau$, the two determinations are now consistent with each other, \begin{equation} \begin{array}{l} \label{aell} A_\ell ({\rm LEP}) = 0.1470 \pm 0.0027, \\ A_\ell ({\rm SLD}) = 0.1503 \pm 0.0023. \end{array} \end{equation} \begin{table}[p] \caption{Principal precision observables from CERN, FNAL, SLAC, and elsewhere. Shown are the experimental results, the SM predictions, and the pulls. The SM errors are from the uncertainties in $M_Z$, $\ln M_H$, $m_t$, $\alpha (M_Z)$, and $\alpha_s$. They have been treated as Gaussian and their correlations have been taken into account. $\bar{s}_\ell^2 (Q_{FB} (q))$ is the weak mixing angle from the hadronic charge asymmetry; $R^-$ and $R^\nu$ are cross section ratios from deep inelastic $\nu$-hadron scattering; $g_{V,A}^{\nu e}$ are effective four-Fermi coefficients in $\nu$-e scattering; and the $Q_W$ are the weak charges from parity violation measurements in atoms. The uncertainty in the $b\rightarrow s\gamma$ observable includes theoretical errors from the physics model, the finite photon energy cut-off, and from uncalculated higher order effects. There are other precision observables which are not shown but included in the fits. Very good agreement with the SM is observed. Only $A_{LR}$ and the two measurements sensitive to $A_b$ discussed in the text, show some deviation, but even those are below $2\sigma$. \label{zpole}} \vspace{0.2cm} \begin{center} \footnotesize \begin{tabular}{|lcccr|} \hline Quantity & Group(s) & Value & Standard Model & pull \\ \hline $M_Z$ \hspace{0pt} [GeV]& LEP &$ 91.1867 \pm 0.0021 $&$ 91.1865 \pm 0.0021 $&$ 0.1$ \\ $\Gamma_Z$ \hspace{3pt} [GeV]& LEP &$ 2.4939 \pm 0.0024 $&$ 2.4957 \pm 0.0017 $&$-0.8$ \\ $\sigma_{\rm had}$ [nb] & LEP &$ 41.491 \pm 0.058 $&$ 41.473 \pm 0.015 $&$ 0.3$ \\ $R_e$ & LEP &$ 20.783 \pm 0.052 $&$ 20.748 \pm 0.019 $&$ 0.7$ \\ $R_\mu$ & LEP &$ 20.789 \pm 0.034 $&$ 20.749 \pm 0.019 $&$ 1.2$ \\ $R_\tau$ & LEP &$ 20.764 \pm 0.045 $&$ 20.794 \pm 0.019 $&$-0.7$ \\ $A_{FB} (e)$ & LEP &$ 0.0153 \pm 0.0025 $&$ 0.0161 \pm 0.0003 $&$-0.3$ \\ $A_{FB} (\mu)$ & LEP &$ 0.0164 \pm 0.0013 $&$ $&$ 0.2$ \\ $A_{FB} (\tau)$ & LEP &$ 0.0183 \pm 0.0017 $&$ $&$ 1.3$ \\ \hline $R_b$ & LEP + SLD &$ 0.21656\pm 0.00074$&$ 0.2158 \pm 0.0002 $&$ 1.0$ \\ $R_c$ & LEP + SLD &$ 0.1735 \pm 0.0044 $&$ 0.1723 \pm 0.0001 $&$ 0.3$ \\ $A_{FB} (b)$ & LEP &$ 0.0990 \pm 0.0021 $&$ 0.1028 \pm 0.0010 $&$-1.8$ \\ $A_{FB} (c)$ & LEP &$ 0.0709 \pm 0.0044 $&$ 0.0734 \pm 0.0008 $&$-0.6$ \\ $A_b$ & SLD &$ 0.867 \pm 0.035 $&$ 0.9347 \pm 0.0001 $&$-1.9$ \\ $A_c$ & SLD &$ 0.647 \pm 0.040 $&$ 0.6676 \pm 0.0006 $&$-0.5$ \\ \hline $A_{LR} + A_\ell$ & SLD &$ 0.1503 \pm 0.0023 $&$ 0.1466 \pm 0.0015 $&$ 1.6$ \\ ${\cal P}_\tau: A_e+A_\tau$ & LEP &$ 0.1452 \pm 0.0034 $&$ $&$-0.4$ \\ $\bar{s}_\ell^2 (Q_{FB})$ & LEP &$ 0.2321 \pm 0.0010 $&$ 0.2316 \pm 0.0002 $&$ 0.5$ \\ \hline $m_t$ \hspace{6pt} [GeV]& Tevatron &$173.8 \pm 5.0 $&$171.4 \pm 4.8 $&$ 0.5$ \\ $M_W$ \hspace{0pt} [GeV]& all &$ 80.388 \pm 0.063 $&$ 80.362 \pm 0.023 $&$ 0.4$ \\ \hline $R^-$ & NuTeV &$ 0.2277 \pm 0.0021 \pm 0.0007 $&$ 0.2297 \pm 0.0003 $&$-0.9$\\ $R^\nu$ & CCFR &$ 0.5820 \pm 0.0027 \pm 0.0031 $&$ 0.5827 \pm 0.0005 $&$-0.2$\\ $R^\nu$ & CDHS &$ 0.3096 \pm 0.0033 \pm 0.0028 $&$ 0.3089 \pm 0.0003 $&$ 0.2$\\ $R^\nu$ & CHARM &$ 0.3021 \pm 0.0031 \pm 0.0026 $&$ $&$-1.7$\\ \hline $g_V^{\nu e}$ & all &$ -0.041 \pm 0.015 $&$ -0.0395 \pm 0.0004 $&$-0.1$\\ $g_A^{\nu e}$ & all &$ -0.507 \pm 0.014 $&$ -0.5063 \pm 0.0002 $&$-0.1$\\ \hline $Q_W({\rm Cs})$& Boulder &$ -72.41 \pm 0.25\pm 0.80 $&$ -73.10 \pm 0.04 $&$ 0.8$\\ $Q_W({\rm Tl})$& all &$-114.8 \pm 1.2 \pm 3.4 $&$-116.7 \pm 0.1 $&$ 0.5$\\ \hline ${\Gamma (b\rightarrow s\gamma)\over \Gamma (b\rightarrow c e\nu)}$& CLEO &$ 3.26^{+0.75}_{-0.68} \times 10^{-3} $&$ 3.14^{+0.19}_{-0.18} \times 10^{-3} $&$ 0.1$\\ \hline \end{tabular} \end{center} \end{table} \noindent The LEP value is from $A_{FB}(\ell)$, ${\cal P}_\tau$, and ${\cal P}^{FB}_\tau$, while the SLD value is from $A_{LR}$ and $A^{FB}_{LR} (\ell)$. The data is consistent with lepton universality, which is assumed here. There remains a $2.5 \sigma$ discrepancy between the two most precise determinations of $\bar{s}^2_\ell$, i.e.\ $A_{LR}$ and $A_{FB} (b)$ (assuming no new physics in $A_b$). Of particular interest are the results on the heavy flavor sector~\cite{Karlen98} including $R_q = {\Gamma (q\bar{q}) \over \Gamma ({\rm had})}$, $A_{FB} (q)$, and $A^{FB}_{LR} (q)$, with $q = b$ or $c$. At present, there is some discrepancy in $A^{FB}_{LR} (b) = {3\over 4} A_b$ and $A_{FB} (b) = {3\over 4} A_e A_b$, both at the $2 \sigma$ level. Using the average of Eqs.~(\ref{aell}), $A_\ell = 0.1489 \pm 0.0018$, both can be interpreted as measurements of $A_b$. From $A_{FB} (b)$ one would obtain $A_b = 0.887 \pm 0.022$, and the combination with $A^{FB}_{LR} (b) = {3\over 4} (0.867 \pm 0.035)$ would yield $A_b = 0.881 \pm 0.019$, which is almost $3 \sigma$ below the SM prediction. Alternatively, one could use $A_\ell ({\rm LEP})$ above (which is closer to the SM prediction) to determine $A_b ({\rm LEP}) = 0.898 \pm 0.025$, and $A_b = 0.888 \pm 0.020$ after combination with $A^{FB}_{LR} (b)$, i.e., still a $2.3 \sigma$ discrepancy. An explanation of the 5--6\% deviation in $A_b$ in terms of new physics in loops, would need a 25--30\% radiative correction to $\hat\kappa_b$, defined by $\bar{s}^2_b \equiv \hat\kappa_b\sin^2\hat\theta_{\overline{\rm MS}} (M_Z)$. Only a new type of physics which couples at the tree level preferentially to the third generation~\cite{Erler95}, and which does not contradict $R_b$ (including the off-peak measurements by DELPHI~\cite{Abreu96}), can conceivably account for a low $A_b$. Given this and that none of the observables deviates by $2 \sigma$ or more, we can presently conclude that there is no compelling evidence for new physics in the precision observables, some of which are listed in Table~\ref{zpole}. \section{Bayesian Higgs mass inference} The data show a strong preference for a low $M_H \sim {\cal O} (M_Z)$, \begin{equation} \label{mh_fit} M_H = 107^{+67}_{-45} \mbox{ GeV}, \end{equation} where the central value (of the global fit to all precision data, including $m_t$) maximizes the likelihood, $N e^{-\chi^2 (M_H)/2}$. Correlations with other parameters, $\xi^i$, are accounted for, since minimization w.r.t. these is understood, $\chi^2 \equiv \chi^2_{\rm min}$. Bayesian methods, on the other hand, are based on Bayes theorem~\cite{Bayes63}, \begin{equation} \label{Bayes} p(M_H | {\rm data}) = \frac{p({\rm data}| M_H) p(M_H)}{p({\rm data})}, \end{equation} which must be satisfied once the {\em likelihood\/}, $p({\rm data}| M_H)$, and {\em prior\/} distribution, $p(M_H)$, are specified. $p(data) \equiv \int p({\rm data}| M_H) p(M_H) d M_H$ in the denominator provides for the proper normalization of the {\em posterior\/} distribution on the l.h.s. The prior can contain additional information not included in the likelihood model, or chosen to be {\em non-informative}. Occasionally, the Bayesian method is criticized for the need of a prior, which would introduce unnecessary subjectivity into the analysis. Indeed, care and good judgement is needed, but the same is true for the likelihood model, which has to be specified in both approaches. Moreover, it is appreciated among Bayesian practitioners, that the explicit presence of the prior can be advantageous: it manifests model assumptions and allows for sensitivity checks. From the theorem~(\ref{Bayes}) it is also clear that the maximum likelihood method corresponds, mathematically, to a particular choice of prior. Thus Bayesian methods differ rather in attitude: by their strong emphasis on the entire posterior distribution and by their first principles setup. Given extra parameters, $\xi^i$, the distribution function of $M_H$ is defined as the marginal distribution, $p(M_H|{\rm data}) = \int p(M_H, \xi^i | {\rm data}) \prod_i p(\xi^i) d \xi^i$. If the posterior factorizes, $p(M_H, \xi^i) = p(M_H) p(\xi^i)$, the $\xi^i$ dependence can be ignored. If not, but $p(\xi^i | M_H)$ is (approximately) multivariate normal, then \begin{equation} \chi^2 (M_H,\xi^i) = \chi^2_{\rm min} (M_H) + {1\over 2} \frac{\partial^2 \chi^2 (M_H)} {\partial \xi_i \partial \xi_j} (\xi^i - \xi^i_{\rm min} (M_H)) (\xi^j - \xi^j_{\rm min} (M_H)). \end{equation} The latter applies to our case, where $\xi^i = (m_t,\alpha_s,\alpha(M_Z))$. Integration yields, \begin{equation} p(M_H | {\rm data}) \sim \sqrt{\det E}\, e^{- \chi^2_{\rm min} (M_H)/2}, \end{equation} where the $\xi^i$ error matrix, $E = (\frac{\partial^2 \chi^2 (M_H)} {\partial \xi_i \partial \xi_j})^{-1}$, introduces a correction factor with a mild $M_H$ dependence. It corresponds to a shift relative to the standard likelihood model, $\chi^2 (M_H) = \chi^2_{\rm min}(M_H) + \Delta \chi^2 (M_H)$, where \begin{equation} \Delta \chi^2 (M_H) \equiv \ln \frac{\det E (M_H)}{\det E (M_Z)}. \end{equation} For example, $\Delta \chi^2 (300 \mbox{ GeV}) \sim 0.1$, which would {\em tighten} the $M_H$ upper limit by at most a few GeV. At present, we neglect this effect. We choose $p(M_H)$ as the product of $M_H^{-1}$, corresponding to a uniform (non-informative) distribution in $\log M_H$, times the exclusion curve from LEP~2.~\cite{McNamara98} This curve is from Higgs searches at center of mass energies up to 183 GeV. We find the 90 (95, 99)\% confidence upper limits, \begin{equation} \label{mh_limits} M_H < 220 \mbox{ (255, 335) GeV}. \end{equation} Theory uncertainties from uncalculated higher orders increase the 95\% CL by about 5~GeV. These limits are robust within the SM, but we caution that the results on $M_H$ are strongly correlated with certain new physics parameters~\cite{Erler99}. The one-sided confidence interval~(\ref{mh_limits}) is not an exclusion limit. For example, the 95\% upper limit of the standard uniform distribution, $x \in [0,1]$, is at $x = 0.95$, but all values of $x$ are equally likely, and $x > 0.95$ cannot be excluded. If there is a discrete set of competing hypotheses, $H_i$, one can use Bayes factors, $p({\rm data} | H_i)/p({\rm data} | H_j)$, for comparison. For example, LEP~2 rejects a standard Higgs boson with $M_H < 90$~GeV at the 95\% CL, because \begin{equation} \frac{p({\rm data} | M_H = M_0)}{p({\rm data} | M_H \neq M_0)} < 0.05 \hspace{20pt} \forall\; M_0 < 90 \hbox{ GeV}. \end{equation} On the other hand, the probability for $M_H < 90$~GeV is only $5\times 10^{-4}$. One could similarly note, that $p(M_H = M_0) < 0.05\, p(M_H = 107 \hbox{ GeV})$ for $M_0 > 334$ GeV; but the (arbitrary) choice of the best fit $M_H$ value as reference hypothesis is hardly justifiable. This affirms that variables continuously connecting a set of hypotheses should be treated in a fully Bayesian analysis. \section*{Acknowledgement} I would like to thank the organizers of WIN 99 for a very pleasant and memorable meeting and Paul Langacker for collaboration. \section*{References}
\section{Introduction} \label{sec:intro} According to the fluctuation-dissipation theorem, the attenuation or amplification of a signal always adds noise. In optical amplifiers, this fact is usually phrased as ``one noise photon'' added to the signal from the spontaneous emission processes in the reservoir. This assumption about the noise gives rise to the phase diffusion responsible for the Schawlow-Townes linewidth \cite{schawlow,scully,lax} of lasers. However, this is not true generally. In particular cases, the noise can exceed the intensity of one photon by the so-called excess-noise factor or Petermann $K$-factor \cite{peter}. Experimentally this phenomenon was first confirmed in a laser cavity with large output coupling leading to an enhancement of a few times \cite{output}. Later, even a factor of a few hundreds was achieved for solid state lasers \cite{semicond} and gas lasers \cite{gas}. Also experiments with a coupling of the polarizations \cite{polar} and an inserted small aperture \cite{aperture} have demonstrated large excess noise. A recent experiment has shown that excess noise can be colored due to saturation effects \cite{color}. After the prediction of excess noise by Petermann for the case of gain-guided semiconductor lasers \cite{peter}, the first more general theory of excess noise was given by Siegman using a semi-classical description \cite{semi}. Until recently, only a few simple systems have been discussed from a quantum mechanical point of view \cite{squant}. In a previous paper \cite{bardroff} we introduced a master equation describing a multi-mode field interacting with a reservoir describing the general linear amplifier or attenuator in a strictly quantum mechanical formulation. We find that under certain conditions, the reservoirs create couplings between the undamped modes of the system. Such dissipative couplings lead to a non-Hermitian eigenvalue problem, which introduces non-orthogonal quasi modes in a natural manner. The amplitudes of these modes are then found to display the expected excess noise, which we here ascribe to the reservoir-induced mode-mode coupling. In this paper we derive the quantum Langevin formalism following from our theory in Ref.~\cite{bardroff}. Whereas the dynamic variable of the master equation is the quantum state, the Langevin equations are for the field operators. This allows us a direct comparison of our approach with the well-known semi-classical treatment introduced by Siegman \cite{semi}. As this has provided the physical understanding and the mathematical expressions for the excess noise, we are pleased that we can essentially derive his starting equations from our fully quantum mechanical treatment. We are also able to generalize the semi-classical analysis of excess noise to cases beyond the paraxial approximation. \section{Master equation} \label{sec:master} In this section we briefly review the results of the quantum derivation of the excess-noise factor based upon the master equation. We use orthonormal real mode functions $u_n(x)$ of the electromagnetic field with frequency $\omega_n$ which fulfill the boundary conditions for the given configuration in the whole ``universe'' and satisfy the orthonormality relation \begin{equation} \frac{1}{V}\int d^3x\,u_n(x)u_m(x)=\delta_{nm}, \label{eq:orthu} \end{equation} where $V$ is the volume of the whole space. Note that the mode function $u_n(x)$ is a vector including the polarization orientation and that we choose them to be real for convenience. The electric field operator then reads \begin{equation} \label{eq:efield} \hat E(x)=\sum_n \varepsilon_nu_n(x)\left(\hat a_n +\hat a_n^\dagger\right), \end{equation} where $\hat a_n$ and $\hat a_n^\dagger$ are the usual creation and annihilation operators of the field excitations and the so-called vacuum field amplitude is \begin{equation} \varepsilon_n=\sqrt{\frac{\hbar \omega_n}{2\epsilon_0 V}}. \label{eq:vacfield} \end{equation} We start from the multi-mode master equation \cite{bardroff} \begin{eqnarray} \frac{d}{dt}\hat\rho(t)&=&\frac{1}{2}\sum_{n,m}L_{m,n}\left\{ 2\hat a_n^\dagger\hat\rho(t)\hat a_m - \hat a_m\hat a_n^\dagger\hat\rho(t) - \hat\rho(t)\hat a_m\hat a_n^\dagger \right\}\nonumber\\ &&+\frac{1}{2}\sum_{n,m}\Gamma_{m,n}\left\{ 2\hat a_n\hat\rho(t)\hat a_m^\dagger - \hat a_m^\dagger\hat a_n\hat\rho(t) - \hat\rho(t)\hat a_m^\dagger\hat a_n \right\} -i\sum_n\omega_n[\hat a_n^\dagger\hat a_n,\hat\rho(t)]. \label{eq:masterda} \end{eqnarray} with the two symmetric matrices $\Gamma_{m,n}$ and $L_{m,n}$ given by \begin{mathletters} \begin{equation} \label{eq:Gamma} \Gamma_{m,n}=\frac{\tau^2}{\hbar^2} \varepsilon_n\varepsilon_m \frac{1}{V}\int d^3x\, r_\Gamma(x)[u_n(x)d][u_m(x)d] \end{equation} and \begin{equation} L_{m,n}=\frac{\tau^2}{\hbar^2} \varepsilon_n\varepsilon_m \frac{1}{V}\int d^3x\, r_L(x)[u_n(x)d][u_m(x)d]. \label{eq:L} \end{equation} \label{eq:GaLa} \end{mathletters} The former describes losses and the latter amplification due to the interaction with the reservoirs. The two reservoirs for amplification and attenuation are assumed to consist of two-level atoms injected in the upper or lower state, respectively \cite{scully}. They are completely characterized by the position dependent injection rates $r_L(x)$ and $r_\Gamma(x)$, the interaction time $\tau$ of the individual atoms with the field and the orientation of the atomic dipole moment $d$. In principle, the dipole orientation could be different for damping and attenuation and it may depend on position. This treatment of the damping can describe spatially localized absorption due to an inserted aperture or due to a detector placed outside the cavity. Assuming a perfect absorber (or detector) surrounding our cavity, the reservoir can also model the damping due to output coupling. Taking the limit of a infinitely large ``universe'' ($V\rightarrow\infty$), and hence using a continuum of modes, would be another way of including losses due to output coupling in our model as shown in Ref.~\cite{lang}. Because of the interaction through the reservoir, the time evolution of the mean values \begin{equation} \label{eq:meana} \frac{d}{dt}\langle\hat a_n\rangle=\frac{1}{2}\sum_m\left( L_{m,n}-\Gamma_{m,n}\right)\langle\hat a_m\rangle -i\omega_n\langle\hat a_n\rangle \end{equation} exhibits coupling between different modes. The definition of the quasi modes operator $\hat A$ follows from imposing the condition \begin{equation} \label{eq:qmodeA} \frac{d}{dt}\langle \hat A\rangle =\Big\{\frac{1}{2}(\lambda-\gamma)-i\Omega\Big\}\langle \hat A \rangle, \end{equation} where $\Omega$ is the frequency, $\lambda$ is the amplification rate and $\gamma$ is the attenuation rate. For later convenience, we split the net-amplification rate $(\lambda-\gamma)$ into the two separate contributions $\lambda$ and $\gamma$. Note that $\Omega$, $\lambda$ and $\gamma$ are real. We write this mode operator in terms of the free field mode operators as \begin{equation} \label{eq:decomp} {\cal E}\hat A=\sum_n\varepsilon_n c_n \hat a_n \end{equation} with the expansion coefficients $c_n$. This transformation includes the vacuum-field amplitudes $\varepsilon_n$ and we define ${\cal E}=\sqrt{\frac{\hbar\Omega}{2\epsilon_0 V}}$, because then the classical field amplitudes $\varepsilon_n\langle \hat a_n\rangle$ obey the same transformation as the operators. Inserting Eq.~(\ref{eq:qmodeA}) into (\ref{eq:decomp}) we get an eigenvalue equation \begin{equation} \label{eq:eigenvalue} \sum_n\left\{ \frac{1}{2}(L_{m,n}-\Gamma_{m,n}) -i\delta_{n,m}\omega_n \right\} \frac{\varepsilon_n}{\varepsilon_m} c_n= \left\{\frac{1}{2}(\lambda-\gamma)-i\Omega\right\} c_m \end{equation} for the non-Hermitian matrix $\{\frac{1}{2}(L_{m,n}-\Gamma_{m,n}) -i\delta_{n,m}\omega_n\} \frac{\varepsilon_n}{\varepsilon_m}$. Here $c_n^{(\nu)}$ is the right eigenvector; the corresponding left eigenvector is $\varepsilon^2_n c_n^{(\nu)}$ \cite{fn:ev}. The superscript $\nu$ distinguishes the different eigenvectors. The detailed properties of the quasi modes are summarized in the Appendix. We can now calculate the noise of the quadrature operator \begin{equation} \label{eq:X} \hat X_\nu(x)= {\cal E}_\nu \left[U_\nu(x)\hat A_\nu+U^*_\nu(x)\hat A_\nu^\dagger\right] \end{equation} with the definition of the quasi-mode function $U_\nu(x)$ given by Eq.~(\ref{eq:rmode}) in the Appendix. Taking the noise averaged over position and comparing to the usual single mode master equation with the same frequency $\Omega_\nu$, damping rate $\gamma_\nu$ and amplification rate $\lambda_\nu$ we find an enhancement by the factor \begin{equation} \label{eq:Kqm} K_\nu=\left|\frac{\sum\varepsilon_n^2|c^{(\nu)}_n|^2}{ \sum_m \varepsilon_m^2 {c^{(\nu)}_m}^2}\right|^2 \end{equation} for the noise added by the reservoir; cf.\ Ref.~\cite{bardroff}. The excess noise is large when the matrices $L_{m,n}$ and $\Gamma_{m,n}$, defined in Eqs.~(\ref{eq:GaLa}), have large off-diagonal terms and when they are not identical. The former follows when the injection rates $r_L(x)$ and $r_\Gamma(x)$ are not spatially constant whereas the latter when damping and amplification are spatially separated. \section{Quantum Langevin equation} \label{sec:qlangevin} Following the usual treatment \cite{lax}, we replace the time evolution described by the master equation (\ref{eq:masterda}) by an equivalent quantum Langevin equation. This contains non-commuting noise forces which are designed such as to give the same moments as those derived from the master equation. The quantum Langevin equation is written \begin{equation} \label{eq:qla} \frac{d}{dt}\hat a(t)_n=\frac{1}{2}\sum_m\left( L_{m,n}-\Gamma_{m,n}\right)\hat a_m(t) -i\omega_n\hat a_n(t)+\hat f_n(t), \end{equation} where the Langevin noise sources $\hat f_n(t)$ obey the correlation relations \begin{mathletters} \begin{eqnarray} \label{eq:corrdef} \langle \hat f_m(t) \hat f_n^\dagger(t')\rangle&=&2\langle\hat D_{\hat a_m \hat a_n^\dagger} \rangle\delta(t-t'),\\ \langle \hat f_n^\dagger(t)\hat f_m(t')\rangle&=&2\langle\hat D_{ \hat a_n^\dagger\hat a_m} \rangle\delta(t-t') \end{eqnarray} \end{mathletters} and \begin{equation} \langle \hat f_m(t)\rangle= \langle \hat f_m(t) \hat f_n(t')\rangle=\langle \hat f_m^\dagger(t) \hat f_n^\dagger(t')\rangle=0. \label{eq:corrzero} \end{equation} It then follows from Eq.~(\ref{eq:qla}) that the expectation values obey Eq.~(\ref{eq:meana}). The diffusion coefficients $\langle\hat D_{\hat a_m \hat a_n^\dagger}\rangle$ and $\langle\hat D_{\hat a_n^\dagger\hat a_m} \rangle$ have to be determined to give the correct noise correlations $\langle\hat a_n^\dagger \hat a_m\rangle$ and $\langle\hat a_m\hat a_n^\dagger\rangle$. We compare the time evolution of the noise correlations derived from the master equation, given by Eqs.~(\ref{eq:acorr}), to the one derived from the Langevin equation (\ref{eq:qla}) to find the relations \begin{mathletters} \label{eq:corrcorr} \begin{equation} \langle\hat f_n^\dagger(t)\hat a_m(t) +\hat a_n^\dagger(t)\hat f_m(t)\rangle=\Gamma_{m,n} \end{equation} and \begin{equation} \langle\hat f_n(t)\hat a_m^\dagger(t) +\hat a_n(t)\hat f_m^\dagger(t)\rangle=L_{m,n}. \end{equation} \end{mathletters} With the help of the Einstein relations \begin{mathletters} \label{eq:einstein} \begin{eqnarray} 2\langle\hat D_{\hat a_m \hat a_n^\dagger}\rangle&=&\frac{d}{dt} \langle\hat a_m(t)\hat a_n^\dagger(t)\rangle- \langle\hat a_m(t) (\frac{d}{dt}\hat a_n^\dagger(t)-\hat f_n^\dagger(t))\rangle- \langle(\frac{d}{dt}\hat a_m(t)-\hat f_m(t)) \hat a_n^\dagger(t)\rangle, \\ 2\langle\hat D_{\hat a_n^\dagger\hat a_m}\rangle&=&\frac{d}{dt} \langle\hat a_n^\dagger(t)\hat a_m(t)\rangle- \langle(\frac{d}{dt}\hat a_n^\dagger(t)-\hat f_n^\dagger(t)) \hat a_m(t)\rangle- \langle\hat a_n^\dagger(t) (\frac{d}{dt}\hat a_m(t)-\hat f_m(t))\rangle, \end{eqnarray} \end{mathletters} and Eqs.~(\ref{eq:corrcorr}), we determine the diffusion coefficients to be \begin{mathletters} \label{eq:noisecorr} \begin{equation} 2\langle\hat D_{\hat a_m \hat a_n^\dagger}\rangle= \langle\hat f_n^\dagger(t)\hat a_m(t) +\hat a_n^\dagger(t)\hat f_m(t)\rangle=\Gamma_{m,n} \end{equation} and \begin{equation} 2\langle\hat D_{\hat a_n^\dagger\hat a_m} \rangle= \langle\hat f_n(t)\hat a_m^\dagger(t) +\hat a_n(t)\hat f_m^\dagger(t)\rangle=L_{m,n}. \end{equation} \end{mathletters} These relations clearly show the mode correlations due to the reservoir. In the following, we derive a wave equation for the propagation of the electric field operator including amplification, damping and the corresponding noise source. Starting from Eqs.~(\ref{eq:qla}) and (\ref{eq:efield}) we can find the exact equation \begin{eqnarray} \label{eq:exact} \left\{\frac{\partial^2}{\partial t^2} -c^2\nabla^2 \right\} \hat E(x,t)-\sum_n\varepsilon_n u_n(x)\Big\{ \sum_k(L_{k,n}-\Gamma_{k,n})\frac{d}{dt}\hat a_k+h.c.\Big\} =\nonumber\\ \sum_n\varepsilon_n u_n(x)\Big\{ -\frac{1}{4}\sum_{k,l}(L_{k,n}- \Gamma_{k,n})(L_{l,k}-\Gamma_{l,k})\hat a_l\nonumber\\ +\frac{i}{2}\sum_k (L_{k,n}-\Gamma_{k,n})(\omega_k-\omega_n)\hat a_k\nonumber\\ -\frac{1}{2}\sum_k(L_{k,n}-\Gamma_{k,n})\hat f_k-i\omega_n\hat f_n +\frac{d}{dt}\hat f_n \Big\} +h.c. \end{eqnarray} with the mode functions $u_n(x)$ fulfilling the Helmholtz equation \begin{equation} \label{eq:hh} (c^2\nabla^2 +\omega_n^2)u_n(x)=0 \end{equation} together with the appropriate boundary conditions. At this point we introduce a number of approximations based on the assumption that the average oscillation frequency $\bar\omega$ of the electric field is much higher than the decay or amplification rates, e.g.\ $|L_{n,m}|$ or $|\Gamma_{n,m}|$. This is well justified in the optical regime where the former is at least six orders of magnitude larger than the latter. The spectral width $\Delta\omega$ of the relevant frequencies $\omega_n$ is assumed to be of the order of the decay or amplification rate. To be more specific, we will neglect terms of the order ${\cal O}(\lambda_\nu/\bar\omega)^2$, ${\cal O}(\lambda_\nu \Delta\omega/\bar\omega^2)$ and ${\cal O}(\Delta\omega/\bar\omega)^2$ or smaller, and we assume ${\cal O}(\lambda_\nu)\approx {\cal O}(\gamma_\mu)$. On the LHS of Eq.~(\ref{eq:exact}) the two terms $\frac{\partial^2}{\partial t^2}\hat E(x,t)$ and $c^2\nabla^2 \hat E(x,t)$ are of the order ${\cal O}(\bar\omega)^2$. Since the remaining term on the LHS of Eq.~(\ref{eq:exact}) is proportional to the damping and amplification rate, we can approximate the frequencies by the mean frequency $\bar \omega$. Hence inserting the definitions of $L_{n,m}$ and $\Gamma_{n,m}$, Eqs.~(\ref{eq:GaLa}), and of $\varepsilon_n$, Eq.~(\ref{eq:vacfield}), the remaining term on the LHS of Eq.~(\ref{eq:exact}) yields \begin{eqnarray} \label{eq:approxLG} &&\sum_n\varepsilon_n u_n(x) \sum_k(L_{k,n}-\Gamma_{k,n})\frac{d}{dt}\hat a_k+h.c.=\nonumber\\ &&\sum_n\varepsilon_n u_n(x)\sum_k\frac{\tau^2}{\hbar^2} \varepsilon_n\varepsilon_k \frac{1}{V}\int d^3x'(r_L(x')-r_\Gamma(x'))[u_n(x')d][u_k(x')d] \frac{d}{dt}\hat a_k+h.c.=\nonumber\\ &&\frac{\tau^2}{\hbar^2}\sum_n\varepsilon_n^2 u_n(x) \frac{1}{V}\int d^3x'(r_L(x')-r_\Gamma(x'))[u_n(x')d]d^T \frac{\partial}{\partial t}\hat E(x',t) \approx\nonumber\\ &&\frac{\tau^2 \bar\omega}{2\epsilon_0\hbar}\int d^3x'(r_L(x')-r_\Gamma(x'))\delta_T(x-x')d\otimes d^T \frac{\partial}{\partial t} \hat E(x',t)= (R_L(x)-R_\Gamma(x)) \frac{\partial}{\partial t}\hat E(x,t). \end{eqnarray} Here the matrices $L_{n,m}$ and $\Gamma_{n,m}$ occur in their position representations \begin{mathletters} \label{eq:posrep} \begin{equation} \frac{1}{V} \sum_{n,m}u_n(x)\otimes u_m^T(x')L_{n,m}\approx \frac{\tau^2 \bar\omega}{2\epsilon_0\hbar} r_L(x)\delta_T(x-x')d\otimes d^T\equiv R_L(x)\delta_T(x-x') \end{equation} and \begin{equation} \frac{1}{V} \sum_{n,m}u_n(x)\otimes u_m^T(x')\Gamma_{n,m}\approx \frac{\tau^2 \bar\omega}{2\epsilon_0\hbar} r_\Gamma(x)\delta_T(x-x')d\otimes d^T\equiv R_\Gamma(x)\delta_T(x-x'). \end{equation} \end{mathletters} Note that $R_L(x)$, $R_\Gamma(x)$, $d\otimes d^T$ and the transverse $\delta$-function \begin{equation} \delta_T(x-x')=\frac{1}{V}\sum_n u_n(x)\otimes u_n^T(x') \label{eq:delta} \end{equation} are tensors. We can neglect the terms on the RHS of Eq.~(\ref{eq:exact}) containing the field operator since they are of the order ${\cal O}(\lambda_\nu)^2$ or ${\cal O}(\lambda_\nu \Delta\omega)$, respectively. For the noise we only take terms of lowest order. Therefore we may neglect the first of the noise terms in Eq.~(\ref{eq:exact}) and we approximate $\frac{d}{dt}\hat f_n\approx-i\bar\omega\hat f_n$. Introducing the position representation \begin{equation} \label{eq:fpos} \hat f(x,t)=\sum_n \varepsilon_n u_n(x) \hat f_n(t) \end{equation} of the noise source we find \begin{equation} \label{eq:tele} \left\{\frac{\partial^2}{\partial t^2} -c^2\nabla^2-(R_L(x)-R_\Gamma(x)) \frac{\partial}{\partial t} \right\} \hat E(x,t)=-2i\bar\omega\hat f(x,t) +h.c. \end{equation} The correlations of the noise operators are \begin{mathletters} \label{eq:corrpos} \begin{equation} \langle \hat f(x,t) \hat f^\dagger(x',t')\rangle= \frac{\hbar\bar\omega}{2\epsilon_0}R_\Gamma(x) \delta_T(x-x')\delta(t-t') \end{equation} and \begin{equation} \langle \hat f^\dagger(x,t) \hat f(x',t')\rangle= \frac{\hbar\bar\omega}{2\epsilon_0}R_L(x) \delta_T(x-x')\delta(t-t'). \end{equation} \end{mathletters} Consequently the total noise on the RHS of Eq.~(\ref{eq:tele}) obeys \begin{equation} \label{eq:corrpostot} \langle(-2i\bar\omega\hat f(x,t)+h.c.)^2\rangle= \frac{\hbar\bar\omega^3}{\epsilon_0}(R_L(x)+R_\Gamma(x)) \delta_T(x-x')\delta(t-t'). \end{equation} As expected, the effects of amplification and damping add for the noise whereas they subtract for the amplification. \section{Semi-classical treatment} \label{sec:semi} Starting from Eq.~(\ref{eq:tele}) we can now perform the transition to the semi-classical treatment replacing operators with $c$-numbers. The solution of Eq.~(\ref{eq:tele}) can conveniently be written using the positive frequency part $E^{(+)}$ of the electromagnetic field. The real part is the electric field and the imaginary part relates to the magnetic field. With the help of the Green function \begin{equation} \label{eq:green} G^{(+)}(x,x',t)=\sum_\nu U_\nu(x)\bar U_\nu(x') e^{\frac{1}{2}(\lambda_\nu-\gamma_\nu)t-i\Omega_\nu t} \end{equation} and the accumulated noise \begin{equation} \label{eq:addnoise} N^{(+)}(x,t)=-2i\bar\omega \int\limits_0^t\! dt' \sum_\nu e^{\frac{1}{2}(\lambda_\nu-\gamma_\nu)(t-t')-i\Omega_\nu (t-t')} U_\nu(x)\frac{1}{V}\int d^3x'\bar U_\nu(x') f(x',t'), \end{equation} we find the field to be given by \begin{equation} \label{eq:evol} E^{(+)}(x,t)=\frac{1}{V}\int d^3x' G^{(+)}(x,x',t) E^{(+)}(x',0) +N^{(+)}(x,t) \end{equation} starting from the initial field $E^{(+)}(x',t=0)$. Within the approximations made, the quasi-mode functions $U_\nu(x)$ and $\bar U_\nu(x)$ and their eigenvalues $\lambda_\nu$, $\gamma_\nu$ and $\Omega_\nu$ are the same as defined using the master equation, Eqs.~(\ref{eq:rmode}) and (\ref{eq:lmode}). When we now calculate the variance of the electric field $E(x,t)$ averaged over position and compare with damping and amplification processes described by the usual single mode master equation, we recover the same $K$-factor, Eq.~(\ref{eq:Kqm}), as before. We find for the noise term \begin{eqnarray} \label{eq:Nsq} \frac{1}{V}\int d^3x \langle (N^{(+)}(x,t)+N^{(-)}(x,t))^2\rangle =\nonumber\\ \frac{\hbar\bar\omega^3}{\epsilon_0}\sum_{\nu,\mu}\frac{ \exp\{\frac{1}{2}(\lambda_\nu+\lambda_\mu-\gamma_\nu-\gamma_\mu)t- i(\Omega_\nu-\Omega_\mu)t\}-1}{\frac{1}{2}(\lambda_\nu+ \lambda_\mu-\gamma_\nu-\gamma_\mu)-i(\Omega_\nu-\Omega_\mu)} \nonumber\\ \times \frac{1}{V}\int d^3x\,U_\nu(x)U_\mu^*(x)\frac{1}{V} \int d^3x\,\bar U_\nu(x) (R_L(x)+R_\Gamma(x))\bar U_\mu^*(x). \end{eqnarray} Considering only one quasi mode, the noise simplifies to \begin{eqnarray} \label{eq:Nsq2} \frac{1}{V}\int d^3x \langle (N_\nu^{(+)}(x,t)+N_\nu^{(-)}(x,t))^2\rangle \approx\nonumber\\ \frac{\hbar\bar\omega^3}{\epsilon_0}\frac{ \exp\{(\lambda_\nu-\gamma_\nu)t\}-1}{\lambda_\nu-\gamma_\nu} (\lambda_\nu+\gamma_\nu)K_\nu \end{eqnarray} with the enhancement factor in the commonly used form \begin{equation} \label{eq:KK} K_\nu=\frac{\int d^3x\,U_\nu(x)U_\nu^*(x) \int d^3x\,\bar U_\nu(x) \bar U_\nu^*(x)}{\left|\int d^3x\,U_\nu(x)\bar U_\nu(x)\right|^2}. \end{equation} Note that our choice of the normalization for the quasi-mode functions is given as in Eq.~(\ref{eq:ortho}). We have shown in Ref.~\cite{bardroff} that Eq.~(\ref{eq:KK}) agrees with Eq.~(\ref{eq:Kqm}) up to the order ${\cal O}(\Delta\omega/\bar\omega)^2$. Siegman \cite{semi} used an equation analogous to Eq.~(\ref{eq:tele}) as the starting point for his derivation of the excess-noise factor. However, there are two interesting differences in the details of the noise source correlations, Eq.~(\ref{eq:corrpostot}). The first difference in Ref.~\cite{semi} is that the spatial transverse $\delta$-function is replaced by a usual $\delta$-function and that the temporal $\delta$-function is replaced by the Hertzian bandwidth $\Delta\omega/(2\pi)$ of the reservoir. The latter circumstance is explained by using the Fourier representation of our noise correlation in Eq.~(\ref{eq:corrpostot}); this leads to the same equation with $\delta(t-t')$ replaced by $\delta(\omega-\omega')/2\pi$. We then integrate with respect to $\omega$ and $\omega'$ over the frequency bandwidth $\Delta\omega$ to obtain \begin{eqnarray} \frac{1}{2}\left[\,\,\, \int\!\!\!\!\!\!\!\! \int\limits_{\bar\omega-\Delta\omega/2}^{\bar\omega+\Delta\omega/2}\! \frac{d\omega d\omega'}{2\pi} \delta(\omega-\omega') e^{-i(\omega t-\omega' t')}+c.c.\right] &=&\frac{\sin(\Delta\omega(t-t')/2)}{2\pi(t-t')}e^{-i\bar\omega (t-t')}+c.c. \nonumber\\ &\approx&\frac{1}{2}\left[ \frac{\Delta\omega}{2\pi} e^{-i\bar\omega (t-t')}+c.c.\right] \label{eq:herzb} \end{eqnarray} for $|t-t'|<(\Delta\omega)^{-1}$. This approximation is reasonable when the mean frequency $\bar\omega$ of the noise is much larger then the bandwidth $\Delta\omega$. The second difference in Ref.~\cite{semi} is that $(R_L(x)+R_\Gamma(x))$ is replaced by $2\left(\frac{R_L}{R_L-R_\Gamma}\right)(R_L-R_\Gamma)=2R_L$ with spatially constant $R_L$ and $R_\Gamma$. This simplification is justified only when averaging over the whole volume V and when amplification and damping are balanced. For the derivation of the $K$-factor which involves an average over position, this is a valid replacement. However, one has to be aware of the subtlety that only non-constant $R_L(x)$ and $R_\Gamma(x)$ with $R_L(x)\neq R_\Gamma(x)$ lead to non-orthogonal quasi modes and hence can give $K>1$. It is interesting to note that within the paraxial approximation, we obtain an equation analogous to the position representation of Eq.~(\ref{eq:qla})---the starting point of our semi-classical analysis. Making the ansatz \begin{equation} \label{eq:ansatz} E^{(+)}(x,t)=e^{i\bar\omega(z/c-t)}\tilde E^{(+)}(x) \end{equation} with $\tilde E^{(+)}(x)$ slowly varying with respect to the longitudinal coordinate $z$, we get from Eq.~(\ref{eq:tele}) \begin{equation} \label{eq:paraxial} c \frac{\partial}{\partial z}\tilde E^{(+)}(x)= \left\{\frac{1}{2}(R_L(x)-R_\Gamma(x)) +\frac{i c^2}{2\bar\omega}\nabla_T^2\right\}\tilde E^{(+)}(x) +\tilde f(x) \end{equation} where $f(x,t)\approx e^{i\bar\omega(z/c-t)}\tilde f(x)$. The time derivative $d/dt$ of Eq.~(\ref{eq:qla}) is replaced by the derivative with respect the longitudinal coordinate $c \partial/\partial z$ which is equivalent in a frame moving with the electromagnetic wave. The frequency part of Eq.~(\ref{eq:qla}) is replaced by the transverse Laplacian. Frequently, Eq.~(\ref{eq:paraxial}) is solved with mode functions of the transverse Laplace equation, depending only parametrically on the longitudinal coordinate. This distinction between longitudinal and transverse coordinates leads to a factorization of the $K$-factor into a longitudinal and a transverse part. \section{Discussion} \label{sec:discussion} In an earlier paper \cite{bardroff}, we derived the master equation for a set of modes coupled to amplifying and attenuating reservoirs. This introduces couplings between the undamped modes of the total ``universe'' and leads directly to the introduction of quasi modes, which are found to exhibit the excess noise described originally by Petermann \cite{peter}. Our treatment has been carried out only in the linear regime so far. This describes an amplifier or an attenuator, where the treatment is most straightforward and the results display the most transparent physical insight. However, the saturation in an operating laser will need to be considered, and we are for the moment carrying out such calculations, which show the influence of the excess noise in the strong field situation. The best physical picture of this noise was provided by Siegman \cite{semi}, who also supplied the quasi-mode expression for the excess-noise factor. This has then been used successfully to describe the experimental findings \cite{output,semicond,gas,polar,aperture}. In \cite{bardroff} we showed that our quantum mechanical approach naturally provides an expression which is essentially identical with Siegman's results. Siegman, however, utilized a semi-classical Langevin approach, where the noise forces were added {\em ad hoc} to the classical equations for the amplitudes; the noise forces were then supplied with properly chosen correlation properties, which was shown to imply the presence of excess noise. Because this approach has been found to give both a physically attractive and theoretically justified description of the situation, we find it interesting to connect that treatment to our quantum approach in some detail. In this paper we derive the quantum Langevin equations following from our general master equation. Here we utilize techniques known from quantum noise theory, and obtain results that can be directly compared with the treatment of Siegman's, when the semi-classical limit is taken. Except for some minor differences, our resulting equations are identical with those used by Siegman. We thus claim that we have justified his formulation of the problem from a more fundamental quantum mechanical point of view. The differences found are either based on natural approximations or obvious qualifications of the results as e.g.\ the introduction of the transverse delta function in the noise correlations. In addition, we have been able to generalize the theory to situations outside the paraxial approximation. The results of our treatment, however, have bearings beyond the problem of excess noise in highly lossy cavities. The approach is quite general, and in addition to the Markov approximation we only need the rotating wave approximation for the interaction with the reservoirs. The master equation is then derived from first principles, and the nonorthogonal quasi modes emerge in a natural manner. The theory is fully general and may well be applicable to other high loss physical systems as well. For the moment we know of no observation that would show the equivalent of the laser excess noise, but novel situations may soon turn up. The lively research activity in quantum information processing, atom optics and novel measurement situations may provide potential applications of the present theory. The physics of our approach resides in the coupling of the undamped modes through the reservoirs. In such a situation, the only essential assumption in our derivation is the Markovian approximation. In highly damped systems, this may not necessarily hold, and the introduction of memory effects in our theory has not been considered so far. Some features are, however, expected to survive, but also unexpected complications may appear. These questions remain to be investigated.
\section{Introduction} It is well known that Ashtekar's formulation of gravity admits degenerate triads and hence degenerate metrics [1]. Various kinds of degenerate solutions to the Ashtekar's equations have been investigated [2-8]. Using a ``covariant approach'', Bengtsson and Jacobson [6] obtained a few 4-dimensional spacetimes containing a ``phase boundary'' separating a degenerate region from a nondegenerate one. According to Ref.[6], the covariant approach starts from a nondegenerate metric which solves Einstein's equations, and then reparametrize one of the coordinates. This reparametrization is chosen so that it is not a diffeomorphism at some particular value of the coordinate. Adopting the new coordinate, the solution can be smoothly matched to a solution to the Ashtekar equations with a degenerate metric at the surface where the transformation misbehaves. To make things clearer we reformulate this procedure as follows. Let $M$ be a 4-dimensional manifold and $M_1$ a 4-dimensional submanifold with a 3-dimensional boundary $\partial M_1$. Suppose $\hat{M}$ is a 4-dimensional manifold with a nondegenerate metric $% \hat{g}_{\mu \nu }$ which solves the Einstein's equations, and $\phi $ is a diffeomorphism from $M_1$ to some open set $\hat{M}_1\subset \hat{M}$. Extend the domain of $\phi $ to the whole of $M$ so that $\phi :M\rightarrow \hat{M}$ is smooth with $M-M_1$ being mapped onto $\phi [\partial M_1]$, and the pushforward $\phi _{*}$ restricted to the tangent bundle of $\partial $$% M_1$ to that of $\phi [\partial M_1]$ is nondegenerate. (It is assumed that $% \phi :M_1\rightarrow \hat{M}_1$ has been chosen so that such an extension is possible.) Then the pullback $g_{\mu \nu }\equiv \phi ^{*}\hat{g}_{\mu \nu }$ is nondegenerate on $M_1$ and degenerate on $M-M_1$. One therefore has a spacetime $(M,g_{\mu \nu })$ with a ``phase boundary'' separating a nondegenerate region from a degenerate one. It is clear that the ``reparametrization procedure'' mentioned above is a special case of this treatment. The authors of Ref.[6] viewed $\phi [\partial M_1]$ as the phase boundary and raised an interesting question: Is the phase boundary always null? They conjectured that the answer is ``yes'' provided that the metric is a ``regular'' solution to Ashtekar's evolution equations, that is, solutions in which the canonical variables $(A_a^i,E_i^a)$, the shift vector $N^i$, and the lapse density $\underline{N}$ all take finite values which, except for $\underline{N}$, are allowed to vanish. (Since $\underline{N}=N/% \sqrt{q}$, where $N$ is the usual lapse scalar and $q$ the determinant of the spatial metric, the requirement that $\underline{N}$ should stay finite is non-trivial when the spatial metric becomes degenerate.) Having reformulated the ``reparametrization procedure'' in the mapping language as stated above, in our opinion it seems reasonable that the ``phase boundary'' should refer to $\partial M_1$ rather than $\phi [\partial M_1]$ since the latter is not at all a boundary between a degenerate region and a nondegenerate one, although Ref.[6] took a different view. We will first show in Sec.2 that $\phi [\partial M_1]$ has to be a null hypersurface. In Sec.3 we will argue that under certain circumstances $% \partial M_1$ could be nonnull as judged by the criterion similar to that of Ref.[6]. Some discussions about the criterion are given in Sec.4. \section{On the boundary $\phi [\partial M_1]$} We now show that the hypersurface $\phi [\partial M_1]$, which is viewed as the degenerate phase boundary in Ref.[6], must be null if the pullback metric $g_{\mu \nu }$ on $M$ is a regular solution to Ashtekar's equations. Consider a ``3+1 decomposition'' of the metric : \begin{equation} ds^2=g_{00}dt^2+2g_{0i}dtdx^i+g_{ij}dx^idx^j=(-N^2+N^iN_i)dt^2+2N_idtdx^i+g_{ij}dx^idx^j, \end{equation} where $N$ is the lapse scalar and $N^i$ the shift vector which relates to the metric components via \begin{equation} g_{ij}N^j=g_{0i},\qquad i=1,2,3. \end{equation} Since $q\equiv \det (g_{ij})=0$ in the degenerate region of $M$, there exists a non-vanishing 3-vector $\lambda ^i$ such that $g_{ij}\lambda ^i=0$, and Eq.(2) then implies that $g_{0i}\lambda ^i=0$. Hence there exists a 4-vector \[ T^\nu =\left( \begin{array}{c} 0 \\ \lambda ^i \end{array} \right) \] at each point of $M-M_1$ such that $g_{\mu \nu }T^\nu =0$. Furthermore, in the degenerate region the lapse scalar $N$ must vanish in order to keep the lapse density $\underline{N}$ finite, hence it follows from Eq.(1) that \begin{equation} -g_{00}+g_{0i}N^i=N^2=0. \end{equation} Eq.(3) together with Eq.(2) provides another 4-vector \[ S^\nu =\left( \begin{array}{c} 1 \\ -N^i \end{array} \right) \] at each point of $M-M_1$ such that $g_{\mu \nu }S^\nu =0$. It is obvious that $T^\nu $ and $S^\nu $ are linearly independent of each other, and hence represent two independent degenerate directions of $g_{\mu \nu }$. That is to say, the degenerate subspace of the tangent space at each point of $M-M_1$ is at least 2-dimensional. Since $\partial M_1$ is 3-dimensional, there must be some degenerate vector field, $W^\nu $, that is tangent to $\partial M_1$% . It then follows from the nondegeneracy of the pushforward $\phi _{*}$ (restricted to $\partial M_1$) that there is a vector field, $\phi _{*}W^\nu $, on $\phi [\partial M_1]$ such that (i) $\phi _{*}W^\nu \neq 0$; (ii) $% \phi _{*}W^\nu $ is tangent to $\phi [\partial M_1]$; (iii) $\phi _{*}W^\nu $ is orthogonal to all vector fields tangent to $\phi [\partial M_1]$ due to $% g_{\mu \nu }=\phi ^{*}\hat{g}_{\mu \nu }$. We therefore conclude that $\phi [\partial M_1]$ is a null hypersurface with null normal $\phi _{*}W^\nu $, and hence the conjecture in Ref.[6] has been proved. Note, however, that the Ashtekar's equations are not at all needed in our proof. It turns out that these equations being necessary for the validity of the conjecture as claimed in Ref.[6] is simply caused by an error, i.e., a superfluous term, $H_R\dot{R}^2$, in Eq.(55) of it. \section{On the degenerate phase boundary $\partial M_1$} In this section we will argue through an example that, although $\phi [\partial M_1]$ is always null, it is not the case for $\partial M_1$ according to the criterion similar to that of Ref.[6]. Since the metric $% g_{\mu \nu }$ is degenerate on $\partial M_1$, it is a delicate issue what definition of nullness of $\partial M_1$ is used. Noticing the criterion for the nullness of $\phi [\partial M_1]$ used in Ref.[6], we define $\partial M_1$ to be null if it is null ``when viewed from the nondegenerate side''. More precisely, suppose $\partial M_1$ is given by $f=0$, where $f$ is a smooth function with $\nabla _af|_{\partial M_1}\neq 0$, then $\partial M_1$ is said to be null if $g^{\mu \nu }\nabla _\mu f\nabla _\nu f\rightarrow 0$ as $\partial M_1$ is approached. Let $(U,X^i)\,(i=1,2,3)$ be a coordinate system on $\hat{M}$ with $U=0$ representing the null hypersurface $\phi [\partial M_1]$ (assuming that $% \phi [\partial M_1]$ can be covered by a single 4-dimensional coordinate patch), and the line element of $\hat{g}_{\mu \nu }$ in this coordinate system reads \begin{equation} d\hat{s}=\hat{g}_{UU}dU^2+2\hat{g}_{Ui}dUdX^i+\hat{q}_{ij}dX^idX^j. \end{equation} The nullness of $\phi [\partial M_1]$ then implies $\lim {_{U\rightarrow 0}}% \hat{q}=0$ where $\hat{q}\equiv \det (\hat{q}_{ij})$. The mapping $\phi :M\rightarrow \hat{M}$ induces four functions $\phi ^{*}U,\phi ^{*}X^i\,(i=1,2,3)$ on $M$ with $\phi ^{*}U|_{M-M_1}=0$. Let $(u,x^i)$ be a coordinate system on $M$ with $u|_{\partial M_1}=0$ and $x^i=\phi ^{*}X^i$, then one has a function $U(u)$ [short for $(\phi ^{*}U)(u)$ ] with $% U^{\prime }(u)|_{M-M_1}\equiv \frac{dU}{du}|_{M-M_1}=0$. The line element of $g_{\mu \nu }$ in this coordinate system is as follows: \begin{equation} ds^2=U^{\prime }\hat{g}_{UU}du^2+2U^{\prime }\hat{g}_{Ui}dudx^i+\hat{q}% _{ij}dx^idx^j. \end{equation} This is exactly the procedure of ``reparametrization of one of the coordinates'' mentioned in Sec.1. Now the key quantity needed for judging whether $\partial M_1$ is null is \begin{equation} g^{\mu \nu }\nabla _\mu u\nabla _\nu u=\hat{q}/g, \end{equation} where $g$ denotes the determinant of the line element (5) and can be expressed as \begin{equation} g=(U^{\prime })^2\hat{g} \end{equation} with $\hat{g}$ the determinant of the line element (4), which does not vanish since $\hat{g}_{\mu \nu }$ is nondegenerate. It then follows from Eqs.(6) and (7) that \begin{equation} \lim_{u\rightarrow 0^{+}}g^{\mu \nu }\nabla _\mu u\nabla _\nu u=\lim_{u\rightarrow 0^{+}}\frac{\hat{q}}{(U^{\prime })^2\hat{g}}% =\lim_{u\rightarrow 0^{+}}\frac 1{2\hat{g}U^{\prime \prime }}\frac{\partial \hat{q}}{\partial U}, \end{equation} where $U^{\prime \prime }\equiv dU^{\prime }/du$, and we assume $u>0$ in $% M_1 $ for convenience. Since $\hat{g}$ is finite and $U^{\prime \prime }$ approaches zero as $u\rightarrow 0^{+}$, it is quite probable to construct an example in which the hypersurface $u=0$ is nonnull by requiring $% \lim_{u\rightarrow 0^{+}}\partial \hat{q}/\partial U\neq 0$ or the rate of approaching zero of $\partial \hat{q}/\partial U$ is equal to or less than that of $U^{\prime \prime }$. The following is a concrete example. Let $(\hat{M},\hat{g}_{\mu \nu })$ be the Minkowski spacetime and the line element in double null coordinates $(\bar{U},\bar{V},Y,Z)$ reads \begin{equation} d\hat{s}^2=-d\bar{U}d\bar{V}+dY^2+dZ^2. \end{equation} A simple coordinate transformation \[ \bar{U}=Ue^{-V},\ \bar{V}=V \] turns it to \begin{equation} d\hat{s}^2=-e^{-V}(dU-UdV)dV+dY^2+dZ^2. \end{equation} It is obvious that $U=0$ is a null hypersurface which serves as $\phi [\partial M_1]$ of the previous discussion. Define $u$ on $M_1$ such that \begin{equation} U(u)=u^3e^u \end{equation} in $M_1$. [The fact $U(u)=0$ in $M-M_1$ follows automatically from the mapping $\phi $ that requires $\phi (M-M_1)=\phi [\partial M_1]$.] the metric $g_{\mu \nu }\equiv \phi ^{*}\hat{g}_{\mu \nu }$ now reads \begin{equation} ds^2=-e^{-V}(U^{\prime }du-UdV)dV+dY^2+dZ^2, \end{equation} where \begin{equation} U^{\prime }\equiv \frac{dU}{du}=\left\{ \begin{array}{ll} u^2(u+3)e^u & \mbox{ in }M_1 \\ 0 & \mbox{ in }M-M_1 \end{array} \right. . \end{equation} It is obvious from Eq.(10) that $\hat{q}=Ue^{-V}$ and hence \[ \lim_{u\rightarrow 0^{+}}\frac{\partial \hat{q}}{\partial U}=e^{-V}\neq 0, \] therefore $g^{\mu \nu }\nabla _\mu u\nabla _\nu u$ approaches infinity rather than zero as $u\rightarrow 0^{+}$, and consequently the phase boundary $\partial M_1$ is nonnull in the sense above. To check that the example is really a regular solution to Ashtekar's equations, we make a simple coordinate transformation \[ u=t-x,\quad V=t+x \] and obtain from Eq.(12) that \begin{equation} ds^2=e^{-V}[(U-U^{\prime })dt^2+2Udtdx+(U+U^{\prime })dx^2]+dY^2+dZ^2. \end{equation} Eqs.(11) and (13) imply \[ U-U^{\prime }\leq 0,\quad U+U^{\prime }\geq 0, \] hence Eq.(14) is a standard formulation of the spacetime metric that can be regarded as a regular solution to Ashtekar's constraint and evolution equations \[ {\cal D}_aE_i^a=0,\ E_i^aF_{ab}^i=0,\ E_i^aE_j^bF_{abk}\epsilon ^{ijk}=0, \] \[ \dot{E}_i^b=-i{\cal D}_a(\underline{N}E_j^aE_k^b)\epsilon ^{ijk}+2{\cal D}% _a(N^{[a}E_i^{b]}), \] \[ \dot{A}_b^i=i\underline{N}E_j^aF_{abk}\epsilon ^{ijk}+N^aF_{ab}^i, \] where \begin{eqnarray} E_1^a &=&\left( \begin{array}{c} 0 \\ -\frac 12[e^{-V}(U+U^{\prime })+1] \\ \frac i2[e^{-V}(U+U^{\prime })-1] \end{array} \right) , \nonumber \\ E_2^a &=&\left( \begin{array}{c} 0 \\ -\frac i2[e^{-V}(U+U^{\prime })-1] \\ -\frac 12[e^{-V}(U+U^{\prime })+1] \end{array} \right) , \nonumber \\ E_3^a &=&\left( \begin{array}{c} 1 \\ 0 \\ 0 \end{array} \right) , \nonumber \end{eqnarray} and \[ A_a^i=0, \] with lapse density and shift vector \[ \underline{N}=\frac{U^{\prime }}{U+U^{\prime }},\ \ N^1=\frac U{U+U^{\prime }% },\ N^2=N^3=0. \] Note that \[ \lim_{u\rightarrow 0^{+}}\underline{N}=1,\quad \lim_{u\rightarrow 0^{+}}N^1=0 \] as the degenerate region is approached. It should be noted that the third derivative of the function $U(u)$ is not continuous at the phase boundary $u=0$. However, the power 3 of $u$ in Eq.(11) can be replaced by any real number greater than 3 to obtain the desired differentiability. \section{Discussions} The idea that the phase boundary, $\partial M_1$, is always null is so attractive that it seems intriguing to attribute the existence of the counterexample presented in the previous section simply to the inappropriate definition of the nullness of the phase boundary. As a matter of fact, the definition used above has a fatal drawback: whether the phase boundary is null depends upon the choice of the function which vanishes on the boundary. Let $f$ and $\bar{f}$ be two distinct functions with $f=\bar{f}=0$ on the phase boundary and $\nabla _\mu f|_{f=0}\neq 0$ and $\nabla _\mu \bar{f}|_{% \bar{f}=0}\neq 0$, then there exists a function $\lambda $ on the boundary such that $\nabla _\mu f|_{f=0}=\lambda \nabla _\mu \bar{f}|_{\bar{f}=0}$. If $g^{\mu \nu }|_{f=0}$ were finite, then $g^{\mu \nu }\nabla _\mu f\nabla _\nu f|_{f=0}$ would be equal to $\lambda ^2g^{\mu \nu }\nabla _\mu \bar{f}% \nabla _\nu \bar{f}|_{\bar{f}=0}$, and hence $g^{\mu \nu }\nabla _\mu f\nabla _\nu f|_{f=0}=0$ if and only if $g^{\mu \nu }\nabla _\mu \bar{f}% \nabla _\nu \bar{f}|_{\bar{f}=0}=0$. However, since $g^{\mu \nu }|_{f=0}$ is infinite, to judge the nullness of the phase boundary one has to calculate $% g^{\mu \nu }\nabla _\mu f\nabla _\nu f$ and $g^{\mu \nu }\nabla _\mu \bar{f}% \nabla _\nu \bar{f}$ in the nondegenerate side and then take the limit. Since $g^{\mu \nu }\nabla _\mu f\nabla _\nu f\neq g^{\mu \nu }\nabla _\mu \bar{f}\nabla _\nu \bar{f}$ in general, there is no guarantee for the equivalence of $\lim_{f\rightarrow 0}g^{\mu \nu }\nabla _\mu f\nabla _\nu f=0 $ and $\lim_{\bar{f}\rightarrow 0}g^{\mu \nu }\nabla _\mu \bar{f}\nabla _\nu \bar{f}=0$. If one chooses $f$ so that all $f=const.$ hypersurfaces in the nondegenerate side are null, then $g^{\mu \nu }\nabla _\mu f\nabla _\nu f=0$ everywhere in the nondegenerate side, hence the limit vanishes, implying the nullness of the phase boundary. For instance, we could choose $% f=ue^{(u-V)/3} $ in our example of Sec.3 to obtain this result. However, the function $u$ in Sec.3 (playing the same role as $f$ here) was chosen intentionally so that the hypersurfaces $u=const.$ are nonnull except for $% u=0$, leading to the conclusion that $\lim_{u\rightarrow 0^{+}}g^{\mu \nu }\nabla _\mu u\nabla _\nu u\neq 0$ and hence the same boundary becomes nonnull. To save the attractive idea that the phase boundary, $\partial M_1$% , is always null just as its image, $\phi [\partial M_1]$, it is necessary to look for a suitable definition of the nullness of the boundary. (This open question has been solved by the time when this paper is accepted [9].) \acknowledgments The authors would like to thank Prof. Ted Jacobson for his comments on the original manuscript and his valuable suggestions. This work has been supported by the National Science Foundation of China.
\section{Introduction} Jeans's theory fails because it is linear. Linear theories typically fail when applied to nonlinear processes; for example, applications of the linearized Navier-Stokes equations to problems of fluid mechanics give laminar solutions, but observations show that actual flows are turbulent when the Reynolds number is large. Cosmology (e.g.; \cite{pbl93}, \cite{pad93}) relies exclusively on Jeans's theory in modelling the formation of structure by gravitational instability. Because $L_J$ for baryonic (ordinary) matter in the hot plasma epoch following the Big Bang is larger than the Hubble scale of causality $L_H \equiv ct$, where $c$ is the speed of light and $t$ is the time, no baryonic structures can form until the cooling plasma forms neutral gas. Star and galaxy formation models invented to accommodate both Jeans's theory and the observations of early structure formation have employed a variety of innovative maneuvers and concepts. Nonbaryonic ``cold dark matter'' was invented to permit nonbaryonic condensations in the plasma epoch with gravitational potential wells that could guide the early formation of baryonic galaxy masses. Fragmentation theories were proposed to produce $M_{\sun}$-stars rather than Jeans-superstars at the $10^{5} M_{\sun}$ proto-globular-cluster Jeans mass of the hot primordial gas. Both of these concepts have severe fluid mechanical difficulties according to the \cite{gib96} theory. Cold dark matter cannot condense at the galactic scales needed because its nonbaryonic, virtually collisionless, nature requires it to have an enormous diffusivity with supergalactic $L_{SD}$ length scales. Fragmentation theories (e.g., \cite{low76}) are based on a faulty condensation premise that implies large velocities and a powerful turbulence regime that would produce a first generation of large stars with minimum mass determined by the turbulent Schwarz scale $L_{ST} \equiv \varepsilon ^{1/2} /(\rho G)^{3/4}$, where $\varepsilon$ is the viscous dissipation rate of the turbulence, with a flurry of starbursts, supernovas, and metal production that is not observed in globular star clusters. The population of small, long-lived, metal-free, globular cluster stars observed is strong evidence of a quiet, weakly turbulent formation regime. \section{Jeans's acoustic theory} Jeans considered the problem of gravitational condensation in a large body of nearly constant density, nearly motionless gas. Viscosity and diffusivity were ignored. The density and momentum conservation equations were linearized by dropping second order terms after substituting mean plus fluctuating values for the density, pressure, gravitational potential, and velocity. Details of the derivation are given in many cosmological texts (e.g.; \cite{kol94}, p342) so they need not be repeated here. The mean gravitational force $\nabla \phi$ is assumed to be zero, violating the Poisson equation \begin{equation} \nabla ^2 \phi = 4 \pi G \rho , \end{equation} where $\phi$ is the gravitational potential, in what is known as the Jeans swindle. Cross-differentiating the linearized perturbation equations produces a single, second order differential equation satisfied by Fourier modes propagating at the speed of sound $V_s$. From the dispersion equation \begin{equation} \omega^2 = V_s^2 k^2 - 4 \pi G \rho , \end{equation} where $\omega$ is the frequency and $k$ is the wavenumber, a critical wavenumber $k_J = (4 \pi G \rho / V_s^2)^{1/2}$ exists, called the $\emph{Jeans wavenumber}$. For $k$ less than $k_J$, $\omega$ is imaginary and the mode grows exponentially with time. For $k$ larger than $k_J$, the mode is a propagating sound wave. Density was assumed to be a function only of pressure (the barotropic assumption). Either the barotropic assumption or the linearization of the momentum and density equations are sufficient to reduce the problem to one of acoustics. Physically, sound waves provide density nuclei at wavecrests that can trigger gravitational condensation if their time of propagation $\lambda/V_s$ for wavelength $\lambda$ is longer than the \emph{gravitational free fall time} $\tau_g \equiv (\rho G)^{-1/2}$. Setting the two times equal gives the Jeans gravitational instability criterion: gravitational condensation occurs only for $\lambda \ge L_J$. Jeans's analysis fails to account for the effects of gravity, diffusivity, or fluid mechanical forces upon nonacoustic density maxima and density minima; that is, points surrounded on all sides by either lower or higher density. These move approximately with the fluid velocity, not $V_s$, (\cite{gib68}). The evolution of such \emph{zero gradient points} and associated \emph{minimal gradient surfaces} is critical to turbulent mixing theory (\cite{gak88}). Turbulence scrambles passive scalar fields such as temperature, chemical species concentration and density to produce nonacoustic extrema, saddle points, doublets, saddle lines and minimal gradient surfaces. A quasi-equilibrium develops between convection and diffusion at such zero gradient points and minimal gradient surfaces that is the basis of a universal similarity theory of turbulent mixing (\cite{gib91}) analogous to the universal similarity theory of Kolmogorov for turbulence. Just as turbulent velocity fields are damped by viscosity at the Kolmogorov length scale $L_K \equiv (\nu/\gamma)^{1/2}$, where $\nu$ is the kinematic viscosity and $\gamma$ is the rate-of-strain, scalar fields like temperature are damped by diffusivity at the Batchelor length scale $L_B \equiv (D/\gamma)^{1/2}$, where $D$ is the molecular diffusivity. This prediction has been confirmed by laboratory experiments and numerical simulations (\cite{gak88}) for the range $10^{-2} \le Pr \le 10^5$, where the Prandtl number $Pr \equiv \nu/D$. On cosmological length scales, density fields scrambled by turbulence are not necessarily dynamically passive but may respond to gravitational forces. In the density conservation equation \begin{equation} \partial \rho/\partial t + v_i (\partial \rho/\partial x_i) = D_\mathrm{eff}\partial ^2 \rho/\partial x_j \partial x_j \end{equation} the effective diffusivity of density $D_\mathrm{eff} \equiv D - L^2 /\tau_g$ is affected by gravitation in the vicinity of minimal density gradient features, and reverses its sign to negative if the feature size $L$ is larger than the diffusive Schwarz scale $L_{SD}$ (\cite{gib98}). $L_{SD} \equiv (D^2/\rho G)^{1/4}$ is derived by setting the diffusive velocity $v_D \approx D/L$ of an isodensity surface a distance $L$ from a minimal gradient configuration equal to the gravitational velocity $v_g \approx L/\tau_g$. Thus, nonacoustic density maxima in a quiescent, otherwise homogeneous, fluid are absolutely unstable to gravitational condensation, and nonacoustic density minima are absolutely unstable to void formation, on scales larger than $L_{SD}$. Jeans believed from his analysis (\cite{jns29}) that sound waves with $\lambda \ge L_J$ would grow in amplitude indefinitely, producing unlimited kinetic energy from his gravitational instability. This is clearly incorrect, since any wavecrest that collects a finite quantity of mass from the ambient fluid will also collect its zero momentum and become a nonacoustic density nucleus. From the enormous Jeans mass values indicated at high temperature, he believed he had proved his speculation that the cores of galaxies consisted of hot gas (emerging from other Universes!) and not stars, which could only form in the cooler (smaller $L_J$) spiral arms, thrown into cold outer space by centrifugal forces of the spinning core. The concepts of \emph{pressure support} and \emph{thermal support} often used to justify Jeans's theory are good examples of bad dimensional analysis, lacking any proper physical basis. \section{Fluid mechanical theory} Gravitational condensation on a nonacoustic density maximum is limited by either diffusion or by viscous, magnetic or turbulent forces at diffusive, viscous, magnetic, or turbulent Schwarz scales $L_{SX}$, whichever is largest, where $X$ is $D,V,M,T$, respectively (\cite{gib96, gib98}). Magnetic forces are assumed to be unimportant for the cosmological conditions of interest. Gravitational forces $F_g \approx \rho^2 G L^4$ equal viscous forces $F_V \approx \rho \nu \gamma L^2$ at $L_{SV} \equiv (\nu \gamma/\rho G)^{1/2}$, and turbulent forces $F_T \approx \rho (\varepsilon)^{2/3}L^{8/3}$ at $L_{ST} \equiv \varepsilon ^{1/2} /(\rho G)^{3/4}$. Kolmogorov's theory is used to estimate the turbulent forces as a function of length scale $L$. The criterion (Gibson 1996, 1997a, 1997b; Gibson and Schild 1998a, 1998b) for gravitational condensation or void formation at scale $L$ is therefore \begin{equation} L \ge \left( L_{SX} \right)_{\mathrm {max}}; X = D,V,M,T . \end{equation} \section{Structures in the plasma epoch} Without the Jeans constraint, structure formation begins in the early stages of the hot plasma epoch after the Big Bang when decreasing viscous forces first permit gravitational decelerations and sufficient time has elapsed for the information about density variations to propagate; that is, the decreasing viscous Schwarz scale $L_{SV}$ becomes smaller than the increasing Hubble scale $L_H \equiv ct$, where $c$ is the velocity of light. Low levels of temperature fluctuations of the primordial gas indicated by the COsmic microwave Background Experiment (COBE) satellite ($\delta T/T \approx 10^{-5}$) constrain the velocity fluctuations $\delta v /c \ll 10^{-5}$ to levels of very weak turbulence. Setting the observed mass of superclusters $\approx 10^{46}$ kg equal to the Hubble mass $\rho L_H^3$ computed from Einstein's equations (\cite{win72}, Table 15.4) indicates the time of first structure was $\approx 10^{12}$ s, or $30\,000$ y (\cite{gib97b}). Setting $L_H \approx 3 \, 10^{20}$ m (10 kpc) = $L_{SV}$ gives $\nu \approx 6 \, 10^{27}$ $ \mathrm{m^2 \, s^{-1}}$ with $\rho \approx 10^{-15}$ $ \mathrm{kg \, m^{-3}}$ and $\gamma \approx 1/t = 10^{-12}$ $ \mathrm{rad \, s^{-1}}$. Such a large viscosity suggests a neutrino-electron-proton coupling mechanism, presumably through the Mikheyev-Smirnov-Wolfenstein (MSW) effect (\cite{bak97}), supporting the Neutrino-98 claim that neutrinos have mass. The viscous condensation mass $\rho L_{SV}^3$ decreases to about $10^{42}$ kg (\cite{gib96}) as the Universe expands and cools to the plasma-gas transition at $t \approx 10^{13}$ s, or $300\,000$ y, based on Einstein's equations to determine $T$ and $\rho$ and assuming the usual dependence of viscosity $\nu$ on temperature $T$ (\cite{win72}). Assuming gravitational decelerations that are possible always occur, we see that protosupercluster, protocluster, and protogalaxy structure formation should be well underway before the emergence of the primordial gas. \section{Primordial fog formation} The first condensation scales of the primordial gas mixture of hydrogen and helium are the maximum size Schwarz scale, and an initial length scale $L_{IC} \equiv (RT/\rho G)^{1/2}$ equal to the Jeans scale $L_J$ but independent of Jeans's linear perturbation stability analysis, and acoustics, where $R$ is the gas constant of the mixture. From the ideal gas law $p/\rho = RT$ we see that density increases can be compensated by pressure increases with no change in temperature in a uniform temperature gas, and that gravitational forces $F_g \approx \rho ^2 G L^4$ will dominate the resulting pressure gradient forces $F_p \approx p L^2 = \rho R T L^2$ for length scales $L \ge L_{IC}$. Taking $R \approx 5\,000$ $\mathrm{m^2 \, s^{-2} \, K^{-1}}$, $\rho \approx 10^{-18}$ $\mathrm{kg \, m^{-3}}$ m (\cite{win72}) and $T \approx 3\,000$ K gives a condensation mass $\rho L_{IC}^3 \approx 10^5 M_{\sun}$, the mass of a globular cluster of stars. Because the temperature of the primordial gas was observed to be quite uniform by COBE, we can expect the protogalaxy masses of primordial gas emerging from the plasma epoch to immediately fragment into proto-globular-cluster (PGC) gas objects on $L_{IC} \approx 3 \, 10^{17}$ m (10 pc) scales, with subfragments at $\left( L_{SX} \right)_{\mathrm {max}}$. The kinematic viscosity $\nu$ of the primordial gas mixture decreased by a factor of about a trillion from plasma values at transition, to $\nu \approx 2.4 \, 10^{12}$ $\mathrm{m^2 \, s^{-1}}$ assuming the density within the PGC objects are about $10^{-17}$ $\rm kg \> m^{-3}$. Therefore, the viscous Schwarz scale $L_{SV} \approx (2.4 \, 10^{12} \, 10^{-13} / 10^{-17} \, 6.7 \, 10^{-11})^{1/2} = 1.9 \, 10^{13}$ m, so the viscous Schwarz mass $M_{SV} \approx L_{SV}^3 \rho = 6.8 \, 10^{22}$ kg, or $M_{SV} = 6.8 \, 10^{24}$ kg using $\rho = 10^{-18}$. The turbulent Schwarz mass $M_{ST} \approx 8.8 \, 10^{22}$ kg assuming $10 \%$ of the COBE temperature fluctuations are due to turbulent red shifts ($[(\delta v / c ) / ( \delta T/T)] = 10^{-1}$) as a best estimate. We see that the entire universe of primordial H-He gas turned to fog soon after the plasma-gas transition, with primordial fog particle (PFP) mass values in the range $10^{23}$ to $10^{25}$ kg depending on the estimated density and turbulence levels of the gas. The time required to form a PFP is set by the time required for void regions to grow from minimum density points and maximum density saddle points to surround and isolate the condensing PFP objects (\cite{gib98}). Voids grow as rarefaction waves with a limiting maximum velocity $V_s$ set by the second law of thermodynamics, so the minimum PFP formation time is $\tau_{PFP} \le (L_{SX})_{\mathrm{max}}/V_s$, or about $10^3$ y. Full condensation of the PFP to form a dense core near hydrostatic equilibrium requires a much longer time, near the gravitational free fall time $\tau_g \approx 2 \, 10^6$ y. Radiation heat transport during the PFP condensation period before the creation of dust should have permitted cooling to temperatures near those of the expanding universe. After about a billion years hydrogen dew point and freezing point temperatures (20-13 K) would be reached, forming the micro-brown-dwarf conditions expected for these widely separated ($10^3-10^4$ AU) small planetary objects that comprise most of the baryonic dark matter of the present universe and the materials of construction for the stars and heavy elements. Because such frozen objects occupy an angle of less than a micro-arcsecond viewed from their average separation distance, they are invisible to most observations except by gravitational microlensing, or if a nearby hot star brings these volatile comets out of cold storage. \section{Observations} A variety of observations confirm the new theory that fluid mechanical forces and diffusion limit gravitational condensation (\cite{gib96}), and confute Jeans's (1902 \& 1929) acoustic criterion: \begin{itemize} \item quasar microlensing at micro-brown-dwarf frequencies (\cite{sch96}), \item tomography of dense galaxy clusters indicating diffuse (nonbaryonic) superhalo dark matter at $L_{SD}$ scales with $D_{nb} \approx 10^{28} \mathrm{m^2 \, s^{-1}}$ (\cite{tyf95}), \item the Gunn-Peterson missing gas sequestered as PFPs, \item the dissipation of `gas clouds' in the $Ly-\alpha$ forest, \item extreme scattering events, cometary globules, FLIERS, ansae, Herbig-Haro `chunks', etc. \end{itemize} Evidence that the dark matter of galaxies is dominated by small planetary mass objects has been accumulating from reports of many observers that the multiple images of lensed quasars twinkle at corresponding high frequencies. After several years spent resolving a controversy about the time delay between images Q0957+561A,B, to permit correction for any intrinsic fluctuations of the light intensity of the source by subtraction of the properly phased images, \cite{sch96} announced that the lensing galaxy mass comprises $\approx 10^{-6} M_{\sun}$ ``rogue planets'' that are ``likely to be the missing mass.'' Star-microlensing studies from the Large Magellanic Cloud have failed to detect lensing at small planetary mass frequencies, thus excluding this quasar-microlensing population as the Galaxy halo missing mass (\cite{alc98}, \cite{ren98}). However, the exclusion is based on the unlikely assumption that the number density of such small objects is uniform. The population must have mostly primordial gas composition since no cosmological model predicts this much baryonic mass of any other material, and must be primordial since it constitutes the material of construction, and an important stage, in the condensation of the gas to form stars. Gravitational aggregation is a nonlinear, self-similar, cascade process likely to produce an extremely intermittent lognormal spatial distribution of the PFP number density, with mode value orders of magnitude smaller than the mean. Since star-microlensing from a small solid angle produces a small number of independent samples, the observations estimate the mode rather than the mean, resolving the observational conflict (\cite{gs98}). \section{Conclusions} \begin{enumerate} \item Jeans's gravitational instability criterion $L \ge L_J$ is irrelevant to gravitational structure formation in cosmology and astrophysics, and is egregiously misleading in all of its applications. \item The correct criterion for gravitational structure formation is that $L$ must be larger than the largest Schwarz scale; that is, $L \ge \left( L_{SX} \right)_{\mathrm {max}}$, where $X$ is $D,V,M,T$, depending on whether diffusion or viscous, magnetic or turbulent forces limit the gravitational effects. \item Structure formation began in the plasma epoch with protosupercluster to protogalaxy decelerations. \item Gravitational condensations began soon after the plasma-gas transition, forming micro-brown-dwarfs, clustered in PGCs, that persist as the dominant dark matter component of inner galactic halos (50 kpc). \item The present fluid mechanical theory and its cosmological consequences regarding the forms of baryonic and nonbaryonic dark matter (\cite{gib96}) is well supported by observations, especially the quasar-microlensing of \cite{sch96} and his inference that the lens galaxy mass of Q0957+561A,B is dominated by small rogue planets (interpreted here as PFPs). \item Star-microlensing studies that rule out MBDs as the Galaxy missing mass (\cite{alc98}, \cite{ren98}), contrary to the quasar-microlensing evidence and the present theory, are subject to extreme undersampling errors from their unwarranted assumption of a uniform number density distribution, rather than extremely intermittent lognormal distributions expected from nonlinear aggregational cascades of such small objects as they form nested clusters and stars (\cite{gs98}). \end{enumerate} \begin{acknowledgements} Numerous helpful suggestions were provided by Rudy Schild. \end{acknowledgements}
\section{Introduction} Years ago, Berends and Gastmans \cite{berends} calculated the one-loop contribution of virtual massless gravitons to $\ifmmode a_{\ell}\else $a_{\ell}$\fi=\tfrac{1}{2} (g-2)_{\ell},$ the anomalous magnetic moment of a lepton $\ell,$ and obtained a finite result. Recently, Graesser \cite{graesser} has redone this calculation in the context of a revised picture of gravity \cite{add} which is of much current interest: in this version, the gravitational sector lives in an expanded $D=(4+n)$-dimensional spacetime, with the extra $n$ dimensions compactified on surfaces whose characteristic sizes may be as large as a millimeter. Graesser has considered the contribution to \ifmmode a_{\ell}\else $a_{\ell}$\fi\ of the resulting Kaluza-Klein tower of spin-2 gravitons (and spin-0 partners). The contribution of each KK mode is again finite, as is the total if the sum on modes is appropriately cut off at or near the $D$-dimensional fundamental scale $\ifmmode{M_D} \else $M_D$\fi$ \cite{garousi}. The finiteness of these results raises an interesting question: is a well-known sum rule for \ifmmode a_{\ell}\else $a_{\ell}$\fi$^2,$ the Drell-Hearn-Gerasimov (DHG) sum rule \cite{dhg} satisfied in quantum gravity, at least at the one-loop level? In accord with a general argument by Brodsky and Schmidt \cite{brodsky}, it will be seen that, at the one-loop level, this sum rule requires the vanishing of a certain integral involving tree-level contributions to a difference of polarized cross sections for the photoproduction of gravitons. As a result of a tedious but straightforward computation, it will be shown that this sum rule is not satisfied, neither in 4-dimensional nor in $(4+n)$-dimensional gravity. At one level, this may be a plausible result: gravity, being a non-renormalizable theory, does not satisfy the finiteness criteria (to be detailed below) necessary for validity of the sum rule. {}From a different perspective, the finiteness of the gravitational contribution to \ifmmode a_{\ell}\else $a_{\ell}$\fi\ at the one-loop level, combined with a string-based belief in Reggeization of amplitudes (including gravitational) at very high energies, suggest that perhaps the sum rule would regain validity upon inclusion of all the high energy string excitations. In conformance with the original $s$- and $t$-channel string duality, this suggested an examination of the constraints on the Regge behavior in order that high energy contributions at at least be consistent with restoring the validity of the sum rule. This could only be done in a rough and speculative fashion, and a discussion of the problems associated with imposing the resulting constraints will be discussed in the context of Type I$^{\, \prime}$ string theories. Speculations aside, the principal concrete result remains that, to lowest order, the DHG sum rule is not obeyed by the tree-level contributions involving photoproduction of gravitons. \section{The DHG Sum Rule and Quantum Gravity} Under certain conditions on the high energy behavior of both the real and imaginary part of the forward spin-difference Compton amplitudes, there exists a sum rule for $\ifmmode a_{\ell}\else $a_{\ell}$\fi^2.$ For a spin-1/2 target, it reads \cite{dhg} \begin{equation} \ifmmode a_{\ell}\else $a_{\ell}$\fi^2\ =\ \frac{m_{\ell}^2}{2\pi^2\alpha}\ \int_{\nu_{th}}^{\infty}\ \frac{d\nu}{\nu}\ \Delta\sigma(\nu)\ \ , \labeq{dhg} \end{equation} where $\Delta\sigma(\nu)\equiv \sigma_P(\nu)-\sigma_A(\nu)$ is the difference between photoabsorption cross sections for the scattering of a photon of lab energy $\nu$ from a target lepton in the cases where the initial photon and lepton spin components along the incident photon direction are parallel and anti-parallel, respectively. The validity of the sum rule is predicated on both the vanishing of $\Delta\sigma$ at high energies and on the absence of polynomial terms in the real part of the forward Compton spin-difference amplitude $f_2$. Both conditions would obtain if the full Compton amplitude were to Reggeize \cite{jaffe}; the details of this Reggeization at string energies will be critical in resolving the problem raised by the calculation which follows. In the Standard Model, the validity of the sum rule requires the vanishing of the integral on the RHS of \eq{dhg} when the cross section is calculated from the lowest order tree graphs, namely $\gamma+\ell\rightarrow\gamma+\ell,\; W^-+\nu_{\ell}$ \cite{altarelli}. This follows from the observation that the LHS of \req{dhg} when calculated in the Standard Model is of \order{\alpha^2} (for fixed $\sin\theta_W),$ whereas in Born approximation the RHS is of \order{\alpha}, and hence must vanish. This is indeed found to be the case, as shown in the explicit calculation of Altarelli, Cabibbo, and Maiani \cite{altarelli}. Using loop expansion techniques, the result has been generalized by Brodsky and Schmidt \cite{brodsky}: the validity of the DHG sum rule requires the vanishing of the integral on the RHS for the sum of $2\rightarrow 2$ processes $\gamma\ell\rightarrow bc$ in the Born approximation. This then raises the interesting question: since the one-loop quantum gravity contributions to $\ifmmode a_{\ell}\else $a_{\ell}$\fi$ are finite and calculable, is the DHG sum rule satisfied to the appropriate order in this case? If not, can we learn anything from the failure to satisfy? In order to simplify the discussion, I will limit the physics to QED and gravity (the rest of the Standard Model can be included with little complication). In this case, at the one loop level, the corrections to $\ifmmode a_{\ell}\else $a_{\ell}$\fi$ {}from these sources are additive: \begin{equation} \ifmmode a_{\ell}\else $a_{\ell}$\fi=\al^{\rm QED}+\al^{\rm QG}\qquad \mbox{\em(1 loop)}\ \ . \labeq{oneloop} \end{equation} Thus, to one-loop level, the sum rule may be written as \begin{equation} \left({\al^{\rm QED}}\right)^{\ 2}+2\ \al^{\rm QED}\ \al^{\rm QG}\ + \left({\al^{\rm QG}}\right)^{\ 2}\ = \frac{m_{\ell}^2}{2\pi^2\alpha}\ \int_{\nu_{th}}^{\infty}\ \frac{d\nu}{\nu}\ \left(\left.\Delta\sigma(\nu)\right|_{\rm QED}+\Delta\sigma^{\, \prime}(\nu)\right)\ \ . \labeq{add} \end{equation} The first term under the integral is the pure QED contribution, up to the appropriate order, and the second is the mixed gravity-QED cross section. Since QED by itself obeys the DHG sum rule, the first terms on the LHS and RHS will cancel. With this cancellation, and to lowest order in $m_{\ell}^2,$ the sum rule reads \begin{equation} 2\ \al^{\rm QED}\ \al^{\rm QG}\ + \left({\al^{\rm QG}}\right)^{\ 2}\ = \frac{m_{\ell}^2}{2\pi^2\alpha}\ \int_{s_{th}}^{\infty}\ \frac{ds}{s}\ \Delta\sigma^{\, \prime}(s)\ \ . \labeq{adda} \end{equation} I will now consider separately the cases with gravity propagating in $n>0 \ (D>4)$ and $n=0\ (D=4)$ dimensions, respectively.\bigskip \noindent\underline{\boldmath {$n>0$}}\bigskip \noindent I first briefly review the properties of $4+n$-dimensional gravity I will need for the discussion which follows. For compactification on an $n$-torus, the $D$-dimensional Planck scale $\ifmmode{M_D} \else $M_D$\fi$ is related to the large radius $R_n$ and the reduced Planck mass via \cite{add,addone,giud} \begin{equation} (\ifmmode{{\bar M}_P} \else ${\bar M}_P$\fi/\ifmmode{M_D} \else $M_D$\fi)^2 = (\ifmmode{M_D} \else $M_D$\fi R_n)^{n}\ \ , \labeq{msr} \end{equation} so that $\ifmmode{M_D} \else $M_D$\fi$ can range from $\sim \ \mbox{TeV}$ to $M_P$ for $R\le$1 mm as long as $n\ge 2.$ The resulting Kaluza-Klein tower of spin-2 gravitons (and spin-0 partners, the radions) can have very small mass splitting $\Delta m^2=(1/R_n)^2.$ This scale also characterizes the surface tension of the soliton (say a D-brane) to which is tied the open string containing ordinary matter. Feynman rules have been developed \cite{giud,lykk} for these couplings, and a large number of authors have explored the phenomenological implications of this view of gravity. Thus, on the RHS of \req{adda}, the lowest order contributions are the tree processes $\gamma\ell\rightarrow G \ell,$ $\gamma\ell\rightarrow \Phi\ell$ where $G$ is one of the spin-2 graviton, and $\Phi$ is its scalar partner (the radion). The cross section for the latter process is suppressed by a factor of $m_{\ell}^2,$ so it does not enter the present consideration.\footnote{In accordance with the scaling argument in \cite{brodsky}, the radion contribution should cancel when combined with \order{m_{\ell}^2}\ corrections to the tree-level $\Delta\sigma(\gamma \ell\ra G\ell).$} The calculation of $\Delta\sigma(\gamma \ell\ra G\ell)$ is straightforward \cite{peskin}. I fix the kinematics so that the photon is incident along the $+z$-axis with momentum $q,$ helicity +1, the lepton with momentum $p_1$ along the $-z$-axis. For massless leptons, the amplitude \req{ampl} is helicity conserving, and the polarization amplitudes for $\gamma\ell\rightarrow G\ell$ are then written as \begin{equation} {\cal M}_{P(A)}=\bar u_{L(R)}(p_2)\ {\cal O}^{\mu\nu\rho}\ u_{L(R)}(p_1) \ \epsilon_{\rho}^{(+1)}(q)\ \left(\epsilon_{\mu\nu}^{\Lambda}(k)\right)^*\ \ , \labeq{em} \end{equation} Squaring and summing over the final state graviton helicity, one finds (for all particles massless except for the graviton, with mass $m$) the differential cross sections in the $c.m.$ \begin{equation} d\sigma_{P(A)}/d\cos\theta=\displaystyle\sum_{\Lambda}\left|{\cal M}_{P(A)}\right|^2/(8\pi\sqrt{s})^2\ \cdot (1-x)\ . \labeq{diffa} \end{equation} where $x\equiv m^2/s,$ and \begin{equation} \displaystyle\sum_{\Lambda}\left|{\cal M}_{P(A)}\right|^2 = (\epsilon_{\rho\prime}^{(+1)}(q))^*\mbox{Tr}\left(\widetilde{\cal O}^{\mu^{\, \prime}\nu^{\, \prime}\rho^{\, \prime}}p_2\hspace{-10pt}/\ \ {\cal O}^{\mu\nu\rho}p_1\hspace{-10pt}/\ \ (1+(-)\gamma_5)/2\right)\ \epsilon_{\rho}^{(+1)}(q)\ {\cal P}_{\mu^{\, \prime}\nu^{\, \prime};\mu\nu}(k)\ \ , \labeq{diffb} \end{equation} The graviton spin-2 projection operator ${\cal P}$ is given in Refs.\cite{giud,lykk}, and $\widetilde {\cal O}=\gamma^{0\dagger}{\cal O}^{\dagger}\gamma^0.$ Finally, from \req{diffa} and \req{diffb}, \begin{eqnarray} \frac{d\sigma_P}{d\cos\theta}-\frac{d\sigma_A}{d\cos\theta} &=&(\epsilon_{\rho\prime}^{(+1)}(q))^* \ \mbox{Tr}\left(\widetilde{\cal O}^{\mu^{\, \prime}\nu^{\, \prime}\rho^{\, \prime}}p_2\hspace{-10pt}/\ \ {\cal O}^{\mu\nu\rho}p_1\hspace{-10pt}/\ \ \gamma_5\right)\ \epsilon_{\rho}^{(+1)}(q) \ \cdot\nonumber\\[.1in] && \quad{\cal P}_{\mu^{\, \prime}\nu^{\, \prime};\mu\nu}(k)\ \cdot (1-x)\ /\ (8\pi\sqrt{s})^2 \ \ . \labeq{pol} \end{eqnarray} The contributing diagrams are given in Refs.\cite{giud,lykk} and are shown in Fig.1. \bigskip \begin{center} \SetScale{.7} \begin{picture}(540,160)(0,0) \SetOffset(80,0) \ArrowLine(0,0)(90,0) \Vertex(90,0){1.5} \ArrowLine(90,0)(180,0) \Photon(0,60)(90,60){1.5}{5} \Vertex(90,60){1.5} \Photon(90,0)(90,60){1.5}{5} \Photon(90,60)(180,60){2}{5} \Photon(90,60)(180,60){-2}{5} \SetOffset(250,0) \ArrowLine(0,0)(90,0) \Vertex(90,0){1.5} \ArrowLine(90,0)(180,0) \Photon(0,60)(90,0){1.5}{8} \Photon(90,0)(180,60){2}{8} \Photon(90,0)(180,60){-2}{8} \SetOffset(80,110) \ArrowLine(0,0)(60,0)\Text(0,-5)[lt]{$p_1$} \Vertex(60,0){1.5} \Line(60,0)(120,0) \Vertex(120,0){1.5} \ArrowLine(120,0)(180,0)\Text(130,-5)[rt]{$p_2$} \Photon(0,60)(60,0){1.5}{6}\Text(0,45)[lb]{ $q$} \Photon(120,0)(180,60){2}{6} \Photon(120,0)(180,60){-2}{6} \Text(125,45)[lb]{$k$} \SetOffset(250,110) \ArrowLine(0,0)(60,0) \Vertex(60,0){1.5} \Line(60,0)(120,0) \Vertex(120,0){1.5} \ArrowLine(120,0)(180,0) \Photon(0,60)(120,0){1.5}{10} \Photon(60,0)(180,60){2}{10} \Photon(60,0)(180,60){-2}{10} \end{picture} \end{center}\bigskip \centerline{ Figure 1: {\em Graphs contributing to} $\gamma\ell\rightarrow G\ell$.}\bigskip \noindent The amplitude can be obtained from these references, and I write it here for completeness (kinematics in Fig.1): \begin{eqnarray} i{\cal O}_{\mu\nu\rho}&=& \left(\frac{ieQ_{\ell}}{4\ifmmode{{\bar M}_P} \else ${\bar M}_P$\fi}\right) \left[ \gamma_{\mu}(P+p_2)_{\nu} \ (P\hspace{-7pt}/\ /s)\ \gamma_{\rho}+\gamma_{\rho}\ (K\hspace{-7pt}/\ /u) \ \gamma_{\mu}(p_1+K)_{\nu}\right]\nonumber\\[.1in] &&- \left(\frac{ieQ_{\ell}}{\ifmmode{{\bar M}_P} \else ${\bar M}_P$\fi}\right)\left[ q_{\mu}Q_{\nu}\gamma_{\rho}+(q\cdot Q) \eta_{\mu\rho}\gamma_{\nu} -\eta_{\mu\rho}Q_{\nu}q\hspace{-6pt}/\ -\gamma_{\mu}q_{\nu}Q_{\rho}\right]/t\nonumber\\[.1in] &&-\left(\frac{ieQ_{\ell}}{2\ifmmode{{\bar M}_P} \else ${\bar M}_P$\fi}\right)\gamma_{\mu}\eta_{\nu\rho}\quad + \mu\leftrightarrow\nu\ \ . \labeq{ampl} \end{eqnarray} In \eq{ampl}, $Q_{\ell}$ is the charge on the lepton, $\ifmmode{{\bar M}_P} \else ${\bar M}_P$\fi=(8\pi G_N)^{-1/2},\ P=p_1+q,\ Q=k-q,\ \mbox{and}\ K=p_1-k$ (see Fig. 1). {}From \req{pol} and \req{ampl} I find \begin{eqnarray} \frac{d\sigma_P}{d\cos\theta}-\frac{d\sigma_A}{d\cos\theta}&=&\frac{1}{16}\ \frac{\alpha} {\ifmmode{{\bar M}_P} \else ${\bar M}_P$\fi^2}\ (1-x) \ \left[ \left(\frac{s}{t}\right)( 4x^2 - 8x ) + \left(\frac{s}{u}\right)(x- 2x^2)\right.\nonumber\\[.1in] &&\left. +\ (2x^2-4x-4 ) + \left(\frac{u}{s}\right)(4-x )\right]\ \ , \labeq{dsig} \end{eqnarray} where $t=Q^2,\ u=K^2.$ The cross section has a smooth massless limit $(x=0)$ which coincides with that obtained from starting with massless gravitons. With the substitutions $(t,u)=-\tfrac{1}{2} s(1-x)(1\mp \cos\theta),$ and the imposition of a collinear cutoff $-1+\delta\le \cos\theta\le 1-\delta$, the total polarization cross section is obtained by integrating \req{dsig} over $\cos\theta:$ \begin{equation} \Delta\sigma_{\gamma\ell\ra G \ell}(s,m^2)=\frac{1}{16}\ \frac{\alpha} {\ifmmode{{\bar M}_P} \else ${\bar M}_P$\fi^2}\ \left[\log\left(\frac{2}{\delta}-1\right)( 14x - 4x^2 ) - (12 -9x-6x^2+3x^3)\right]\ \ . \labeq{cs} \end{equation} The contribution to $\Delta\sigma_{\gamma\ell\ra G \ell}(s)$ from an entire KK tower of gravitons is given (approximately) by integrating over the density of states\ \cite{giud,lykk,peskin} up to the kinematic limit \begin{eqnarray} \Delta\sigma_{\gamma\ell\ra G \ell}(s)&=&\frac{2\pi^{n/2}}{\Gamma(n/2)}\ R^n\ \int_0^{\sqrt{s}\ }\Delta\sigma_{\gamma\ell\ra G \ell}\ (s,m^2)\ m^{n-1}\ dm\nonumber\\[.1in] &=&\frac{\pi^{n/2}}{\Gamma(n/2)}\ \frac{\ifmmode{{\bar M}_P} \else ${\bar M}_P$\fi^2}{\ifmmode{M_D} \else $M_D$\fi^2}\left(\frac{s}{\ifmmode{M_D} \else $M_D$\fi^2}\right)^{n/2}\ \int_0^1\Delta\sigma_{\gamma\ell\ra G \ell}\ (x)\ x^{n/2-1}\ dx\nonumber\\[.1in] &=&\frac{1}{16}\ \frac{\alpha}{\ifmmode{M_D} \else $M_D$\fi^2}\cdot\ \left(\frac{s}{\ifmmode{M_D} \else $M_D$\fi^2}\right)^{n/2}\ A_n\ \ , \labeq{intm} \end{eqnarray} where, {\em e.g.,} \begin{eqnarray*} A_2&=&\pi\left(-25/4+(17/3)\log(2/\delta-1)\right)\\ A_4&=&\pi^2\left(-21/10+(11/3)\log(2/\delta-1)\right)\end{eqnarray*} Finally, the contribution of \req{intm}, integrated up to some upper limit $\bar s\le \ifmmode{M_D} \else $M_D$\fi^2$ to the RHS of the sum rule \req{adda} is \begin{equation} \frac{m_{\ell}^2}{2\pi^2\alpha}\ \int_{s_{th}}^{\bar s}\ \frac{ds}{s}\ \Delta\sigma_{\gamma\ell\ra G \ell}(s)= \frac{1}{16\pi^2}\ \left(\frac{m_{\ell}^2}{\ifmmode{M_D} \else $M_D$\fi^2}\right)\ \left(\frac{A_n}{n}\right)\ \kappa^{n/2}\ \ , \labeq{dhgb}\end{equation} where $\kappa\equiv \bar s/\ifmmode{M_D} \else $M_D$\fi^2.$ \noindent On the LHS of \req{adda}, $\al^{\rm QED}\sim\order{\alpha}$ and (using the same cutoff at $\bar s)$ $\al^{\rm QG}\sim\kappa^{n/2}\order{\ml^2/\md^2}$ \cite{graesser}; thus \req{dhgb} by itself violates the DHG sum rule at the one-loop level. Before discussing this result, I will present the analogous result for the case with no extra dimensions. \noindent\underline{\boldmath $n=0$}\bigskip \noindent As mentioned after \eq{dsig}, the $D=4$ case is obtained by setting $x=0$ in that equation, or in \eq{cs}: \begin{equation} \Delta\sigma_{\gamma\ell\ra G \ell}(s)=\Delta\sigma_{\gamma\ell\ra G \ell}(s,0)=-\frac{3}{4}\ \frac{\alpha} {\ifmmode{{\bar M}_P} \else ${\bar M}_P$\fi^2}=-\frac{6\pi\alpha}{\ifmmode{M_P} \else $M_P$\fi^2}\ \ . \labeq{csfour} \end{equation} Then, integration up to $\bar s\le \ifmmode{M_P} \else $M_P$\fi^2$ gives a contribution to the RHS of the sum rule \begin{equation} \frac{m_{\ell}^2}{2\pi^2\alpha}\ \int_{s_{th}}^{\bar s}\ \frac{ds}{s}\ \Delta\sigma_{\gamma\ell\ra G \ell}(s)= -\frac{3}{\pi}\ \frac{m_{\ell}^2}{\ifmmode{M_P} \else $M_P$\fi^2} \cdot\ \log\left(\frac{\bar s}{s_{th}}\right) \ \ . \labeq{dhgbfour} \end{equation} Again, since $\al^{\rm QED}\sim\order{\alpha}$ and $\al^{\rm QG}\sim\order{m_{\ell}^2/\ifmmode{M_P} \else $M_P$\fi^2}$ \cite{berends}, this contribution by itself violates the sum rule \req{adda}. I now turn to discuss these results. \section{Discussion} The failure of the gravitational contributions integrated to a large scale $\bar s$ to satisfy the DHG sum rule at the one-loop level suggests at least two possibilities: (1) the sum rule is not valid for processes involving quantum qravity (2) there are contributions from $s>\bar s$ which cancel the low energy contribution and render the sum rule valid. As mentioned in the introduction, and discussed at length in Ref. \cite{jaffe}, the sum rule can fail because the spin-difference forward Compton amplitude $f_2(s)$ has fixed poles ({\em i.e.,} a polynomial piece to the real part), or an imaginary part whose asymptotic behavior requires a subtraction for the dispersion relation. One or both of these is certainly possible: for example, in the $n=0$ case, there may be at order $\alpha/\ifmmode{M_P} \else $M_P$\fi^2$ a gravity-induced ``seagull'' term in $f_2.$ It is not immediately apparent, however, how this can cancel against the logarithmic cutoff in \req{dhgbfour}. As far as the imaginary part is concerned, one may certainly conceive of spin-dependent gravitational contributions ({\em e.g.,} spinning black holes) whose contributions vitiate the sum rule. In such cases, within the present state of knowledge about non-perturbative quantum gravity, there is not much more to say. It is then interesting to speculate about the second possibility above, the possibility of cancellation.\bigskip \noindent {\large {\bf Possible Role of String Theory}}\medskip \noindent String theory suggests that $f_2$ Reggeizes for $s$ larger than the string scale $\ifmmode{M_s} \else $M_s$\fi^2,$ at once eliminating the possibility of the fixed poles and the non-convergence. Moreover, the required Reggeization takes place at the tree-level (the Veneziano amplitude being equivalent to a sum of poles), so that according to the loop-counting criteria discussed in \cite{brodsky}, the additive form of \req{adda} is valid. I will discuss in turn the two cases of large extra dimensions $(n>0)$ and no large extra dimensions $(n=0)$.\medskip \noindent {\boldmath $n>0$:} In this case, I will adopt the string description presented in \cite{ant}, with reference to previous work in \cite{lykken} and \cite{polchwitt}, and detailed in \cite{addone,tye}. This is based on a type I theory of open and closed strings, with a $T$-duality transformation allowing a large radius in the $n$ extra dimensions, as well as a weak coupling description. Matter resides in open strings tied to D3-branes and $\lambda=2\alpha$ serves as the string coupling expansion parameter \cite{ant}. The 4-point open string Compton amplitude is then written as in \cite{green}, with the appropriate kinematic factors assuring the behavior (for $s\gg\ifmmode{M_s} \else $M_s$\fi^2)$ \begin{equation} \mbox{Im} f_2(s)\ \sim\ \alpha\ \cdot\ const,\quad \Delta\sigma^{\, \prime}\sim \alpha/s \labeq{asym} \end{equation} which allows convergence. The contribution to the DHG sum rule from the Regge region is \begin{equation} \frac{m_{\ell}^2}{2\pi^2\alpha}\ \int_{\bar s}^{\infty}\ \frac{ds}{s}\ \Delta\sigma^{\, \prime}(s)\sim \left(\frac{m_{\ell}^2}{\ifmmode{M_D} \else $M_D$\fi^2}\right)\ \left(\frac{1}{\kappa}\right)\ \ . \labeq{regge} \end{equation} A necessary (but certainly not sufficent!) condition that \req{regge} could cancel the lower energy integral \req{dhgb} is that $\kappa\sim \order{1}$, {\em i.e.,} the perturbative treatment of $\gamma \ell\rightarrow G\ell$ is reliable for $\sqrt{s}\ \raisebox{-0.5ex}{$\stackrel{<}{\sim}$}\ \ifmmode{M_D} \else $M_D$\fi.$ Some discussion of this point is given in \cite{giud}. An important ingredient is that the D-brane surface tension, which controls the ``soft'' scale for gravitons emitted transverse to the brane, is larger than the string scale \cite{garousi}, so that $\ifmmode{M_D} \else $M_D$\fi$ may be reasonable as an energy cutoff for the perturbative treatment of graviton emission. The requirement $\bar s\sim \ifmmode{M_D} \else $M_D$\fi^2$ {\em vs.} the expected $\bar s\sim \ifmmode{M_s} \else $M_s$\fi^2$ is perhaps disconcerting, and some discussion will be presented in a following section. All of this is a long way from claiming that cancellation occurs: for example, the the collinear cutoff dependence (which occurs only in this extra-dimension case) in \req{intm} has no counterpart in \req{regge}. One is assured by the Lee-Nauenberg theorem \cite{lee} that including all channels will regulate such singularities to measurable quantities, but how (and whether) this happens smoothly over the whole energy range in the present case certainly remains to be shown.\medskip \noindent {\boldmath $n=0$:} I use the same model here as above, except with no large dimensions. In addition to $\lambda=2 \alpha,$ one also has (after the T-duality transformation on all six compact coordinates)\cite{tye} \begin{equation} \ifmmode{M_s} \else $M_s$\fi=(1/\sqrt{2})\alpha\ifmmode{M_P} \else $M_P$\fi/r^3\ \ , \labeq{msmp} \end{equation} where $r\ge 1,$ the compactification radius in units of $\ifmmode{M_s} \else $M_s$\fi^{-1},$ can be used to adjust $\ifmmode{M_s} \else $M_s$\fi=M_{GUT}.$ The Compton amplitude in the Regge region is the same as in \req{asym}, so that the contribution from $s\ge \bar s$ to the DHG integral is given by \begin{equation} \frac{m_{\ell}^2}{2\pi^2\alpha}\ \int_{\bar s}^{\infty}\ \frac{ds}{s}\ \Delta\sigma^{\, \prime}(s) \sim\frac{\lambda}{\alpha}\ \frac{m_{\ell}^2}{\bar s} \sim\left(\frac{m_{\ell}^2}{\ifmmode{M_P} \else $M_P$\fi^2}\right)\left( \frac{\ifmmode{M_P} \else $M_P$\fi^2}{\bar s}\right)\ \ . \labeq{reggefour} \end{equation} In this case, cancellation of the lower energy contribution \req{dhgbfour} is possible only for $\bar s\sim\ifmmode{M_P} \else $M_P$\fi^2\ \raisebox{-0.5ex}{$\stackrel{>}{\sim}$}\ \ifmmode{M_s} \else $M_s$\fi^2/\alpha^2.$ Again, this is merely a necessary condition -- there is no demonstration of cancellation, and problems attached to it as well: for example, the logarithmic cutoff factor in \req{dhgbfour} does not appear in the purported cancelling term \req{reggefour}.\bigskip \noindent {\large {\boldmath $\ifmmode{M_D} \else $M_D$\fi$ (or $\ifmmode{M_P} \else $M_P$\fi$) {\em vs.} $\ifmmode{M_s} \else $M_s$\fi$}}\medskip \noindent To pursue the possibility of cancellation, one must reconcile the appearance of $\bar s=\ifmmode{M_D} \else $M_D$\fi^2 (\ifmmode{M_P} \else $M_P$\fi^2)$ with the expected $\bar s= \ifmmode{M_s} \else $M_s$\fi^2$ as the cutoff. As discussed above, there may be plausible arguments why graviton emission into the bulk may be adequately described by perturbation theory for energies up to $\ifmmode{M_D} \else $M_D$\fi;$ however, it is difficult to understand why the open string contribution to Compton scattering would be delayed for energies above $\ifmmode{M_s} \else $M_s$\fi,$ until $\ifmmode{M_D} \else $M_D$\fi.$ It might be thought that the difference is not significant, since in the string theory model $\ifmmode{M_s} \else $M_s$\fi$ and $\ifmmode{M_D} \else $M_D$\fi$ are numerically close \cite{ant,addone}; however, this is not the case. The insertion of $\bar s=\ifmmode{M_s} \else $M_s$\fi^2$ as the cutoff would cause a mismatch of a factor of $\alpha^2$ between the contributions below and above $\bar s:$ in the case $n>0,$ this occurs when the relation $\ifmmode{M_D} \else $M_D$\fi\sim\alpha^{-2/(n+2)}\ifmmode{M_s} \else $M_s$\fi$ \cite{ant,addone} (along with $\bar s=\ifmmode{M_s} \else $M_s$\fi^2)$ is inserted into \req{dhgb} and $\bar s=\ifmmode{M_s} \else $M_s$\fi^2$ in \req{regge}; for $n=0,$ one has a similar situation, with $\ifmmode{M_P} \else $M_P$\fi\sim \ifmmode{M_s} \else $M_s$\fi/\alpha$ in \req{dhgbfour} and $\bar s=\ifmmode{M_s} \else $M_s$\fi^2$ in \req{reggefour}. There is only one parameter available to play with, namely the compactification radius $R_{6-n}$ of the $6-n$ remaining dimensions in the bulk. Can this be of a form such as to remove this parametric disparity between $\ifmmode{M_D} \else $M_D$\fi(\ifmmode{M_P} \else $M_P$\fi)$ and $\ifmmode{M_s} \else $M_s$\fi?$ Consider again the Type I$^{\, \prime}$ theory \cite{ant,addone,tye}, with the standard model fields residing in open strings tied to a D3-brane, and gravity propagating in the 10-dimensional bulk. Allowing (as before) $n$ of the compact dimensions to be large, of equivalent toroidal radius $R_n$, and $6-n$ to be small (of radius $R_{6-n}\sim \ifmmode{M_s} \else $M_s$\fi^{-1}$), one obtains the relation \cite{ant,addone} \begin{equation} (\ifmmode{M_D} \else $M_D$\fi/\ifmmode{M_s} \else $M_s$\fi)^{n+2}=\frac{1}{4\pi\alpha^2}\ (R_{6-n}\ifmmode{M_s} \else $M_s$\fi)^{6-n}\ \ . \labeq{trms} \end{equation} Thus, for $\alpha\sim 0.1$ it is possible to simultaneously have $R_{6-n}\ifmmode{M_s} \else $M_s$\fi>1$ and $\ifmmode{M_D} \else $M_D$\fi\sim \order{\alpha^0}\ifmmode{M_s} \else $M_s$\fi$ if \begin{equation} (R_{6-n}\ifmmode{M_s} \else $M_s$\fi)^{6-n}\ \raisebox{-0.5ex}{$\stackrel{>}{\sim}$}\ 100\ \alpha^2\ =\ 25\ \lambda_{I^{\, \prime}}^2\ \ . \labeq{six} \end{equation} where $\lambda_{I^{\, \prime}}$ is the Type I$^{\, \prime}$ string coupling constant. This is a statement that the compactification scale is parametrically tied to the dilaton expectation value \cite{dine}. It is not meant to be a perturbative statement, but as a constraint on the non-perturbative minima in the $S-T$ modular space. Needless to say, this is very {\em ad hoc} -- I only present it as a hypothetical way of reconciling the $\ifmmode{M_s} \else $M_s$\fi/\ifmmode{M_D} \else $M_D$\fi$ problem. This scenario is more circumscribed when there are no large extra dimensions. In that case one has, in analogy to \req{trms}, $(\ifmmode{M_P} \else $M_P$\fi/\ifmmode{M_s} \else $M_s$\fi)^2=(2/\alpha^2)(R_6\ifmmode{M_s} \else $M_s$\fi)^6$\ \cite{tye}. However, now $\ifmmode{M_s} \else $M_s$\fi$ is identified with $M_{GUT},$ so that $\ifmmode{M_P} \else $M_P$\fi/\ifmmode{M_s} \else $M_s$\fi\simeq 500.$ If this be due to modular geometry rather than a factor of $\alpha,$ we must impose $(R_6\ifmmode{M_s} \else $M_s$\fi)^6\simeq 30,000\ \lambda_{I^{\, \prime}}^2$ on the compactification volume, or $R_6\ifmmode{M_s} \else $M_s$\fi\simeq 6\ \lambda_{I^{\, \prime}}^{\tfrac{1}{3}},$ on the compactification radius. \section{Concluding Remarks} The finiteness of the one-loop gravitation contribution to the anomalous magnetic moment \ifmmode a_{\ell}\else $a_{\ell}$\fi\ led naturally to the question of whether the DHG sum rule for \ifmmode a_{\ell}\else $a_{\ell}$\fi$^2$ is satisfied. The failure to satisfy the sum rule for perturbative contributions below an arbitrary scale $\bar s$ could be ascribed to the failure to understand quantum gravity in the strong coupling (high energy) region. Alternatively, there were presented some speculations concerning possibly compensating contributions above the string scale. In both cases $(n>0\ \mbox{and}\ n=0)$ this possibility was beset with uncertainty concerning various logarithmic cutoff factors, not to speak of the absence of an exact calculation. In the simplest (Regge) approximation, the possibility of a string `fix' for the validity of the sum rule imposed a {\em necessary} parametric condition: namely, that the ordinary tree level perturbative contribution be included all the way to the respective Planck scales $\ifmmode{M_D} \else $M_D$\fi,\ \ifmmode{M_P} \else $M_P$\fi$ (rather than the expected string scale $\ifmmode{M_s} \else $M_s$\fi),$ and that the string Compton amplitude used thereafter. Identifying the string and Planck scales is possible through a certain dependence of the compactification volume for the $6-n$ `small' dimensions on the dilaton expectation value. The simple analysis presented here, using only an open string Compton amplitude, ignores specifically non-perturbative gravitational contributions (such as black hole formation \cite{nussinov}) which may set in at $\ifmmode{M_D} \else $M_D$\fi\ (\ifmmode{M_P} \else $M_P$\fi)$. Perhaps these do not contribute to the forward spin-difference amplitude $f_2$ (the leading graviton trajectory considered in \cite{nussinov} does not contribute to $f_2(0)$); however, no such statement is possible in the presence of spinning black holes. It is certainly an open question whether these have been eliminated in the strong-weak duality transformation particular to the model considered. In sum, the principal finding is that the DHG sum rule is not satisfied at the one-loop level. Although one may speculate along certain stringy fixes to this situation, there is at present no compelling reason to adopt these in preference to simply pleading ignorance about the convergence properties of $f_2$ in the non-perturbative regime of quantum gravity. \subsection*{Acknowledgement} I would like to thank Zurab Kakushadze and Tom Taylor for helpful conversations. This research was supported in part by the National Science Foundation through Grant No. PHY-9722044.
\section*{Introduction} The convexity theorem of Atiyah \cite{at} and Guillemin-Sternberg \cite{gs} says that if $T$ is a torus acting in a Hamiltonian fashion on a compact, connected symplectic manifold $M$, then the image of the corresponding moment mapping is a {\em rational} convex polytope. One of the most interesting applications of this theorem is a classification theorem of Delzant \cite{d}, which states that if $\dim{M}=2\dim{T}$ and the action is effective, then the space is completely characterized by the image of the moment mapping, which is a simple rational convex polytope satisfying a special integrality condition. One of the features of Delzant's result is that it provides an explicit construction for associating to each polytope the corresponding space; this construction involves the technique of symplectic reduction. The results of Atiyah, Guillemin-Sternberg and Delzant have subsequently been extended by Lerman-Tolman \cite{lt} to the case of torus actions on symplectic orbifolds; the image of the moment mapping in this case is still a rational polytope, and the extension of Delzant's theorem involves simple rational polytopes. However, it is very natural to ask oneself whether a simple convex polytope that is {\em not} rational can also be viewed as the image of the moment mapping for a suitable symplectic space. Answering affirmatively to this question amounts to being able to perform symplectic reduction under rather general assumptions, thus allowing the resulting space to be pathological. This has lead us to consider a new class of spaces which we call {\em quasifolds}. Roughly speaking, a quasifold of dimension $k$ is a space that is locally modeled on orbit spaces of discrete group actions on open subsets of the space $\R^k$. Manifolds and orbifolds are special cases of quasifolds, but quasifolds in general are not Hausdorff topological spaces. Just as for orbifolds, geometric objects on quasifolds may be thought of as collections of objects on the open sets of the space $\R^k$ that are invariant under the discrete group actions, and that behave correctly under coordinate changes. The natural analogue of a torus in this geometry is a {\em quasitorus}, which is the quotient of a vector space by a {\em quasilattice}. It is then possible to define Hamiltonian quasitorus actions on symplectic quasifolds and to extend the Delzant construction to show that every simple convex polytope $\Delta$ is the image of the moment mapping for quasitorus actions on a family, ${\cal M}_{\Delta}$, of quasifolds. We remark that the initial motivation for this article came from a discussion with Traynor on the role of non-rational polytopes in the study of symplectic packings \cite{mp, t}. Orbit spaces of discrete group actions have been studied by Connes in the context of {\em noncommutative geometry} \cite[chapter II]{c}; our approach is different and we do not fully understand the connection. Quasitori of dimension one have been studied by Donato, Iglesias and Lachaud \cite{di, i, il} in the framework of the theory of {\em diffeological spaces}; on this occasion Iglesias introduced the terminology {\em irrational tori}. On the other hand Weinstein considered quasitori of dimension one to prequantize arbitrary symplectic manifolds \cite{w1, w2}; he introduced the term {\em infracircles}. The subject of this article is also related to the {\em geometry of quasicrystals} \cite{ar, se}; for example the regular pentagon is not only a celebrated quasicrystal but is also a simple non-rational convex polytope. This is the reason underlying our choice of the terms quasifold and quasitorus; the term quasilattice on the other hand had already been introduced by quasicrystallographers. The article is structured as follows: in Section~\ref{folds} we define quasifolds and the essentials of their geometry, in Section~\ref{tori} we define quasitori and Hamiltonian actions, in Section~\ref{delzant} we prove a symplectic reduction theorem and the extension of Delzant's construction to this setting. A brief appendix recalls the definitions of rational and simple convex polyhedral sets. All definitions and results are illustrated by examples. The contents of this article have been announced in \cite{p}. In the sequel we will give a more thorough treatment of the convexity theorem and of the failure of the uniqueness part of Delzant's theorem. In an article in collaboration with Battaglia \cite{bp} we introduce complex and K\"ahler structures on quasifolds, and see how the spaces in the family ${\cal M}_{\Delta}$ can be viewed as natural generalizations of the toric varieties that are usually associated to those simple convex polytopes that are rational. We wish to thank Ana Cannas da Silva, Patrick Iglesias, Reyer Sjamaar and the Referee for their helpful remarks. We are also very grateful to Fiamma Battaglia for her crucial help on several aspects of this work. \section{Quasifolds}\label{folds} We begin by introducing the local model for quasifolds. \begin{defn}[Model]\label{model}{\rm Let $\tilde{U}$ be a connected, simply connected manifold of dimension $k$ and let $\Gamma$ be a discrete group acting smoothly on the manifold $\tilde{U}$ so that the set of points, $\tilde{U}_0$, where the action is free, is connected and dense. Consider the space of orbits, $\tilde{U}/\Gamma$, of the action of the group $\Gamma$ on the manifold $\tilde{U}$, endowed with the quotient topology, and the canonical projection $p\,\colon\,\tilde{U}\rightarrow\tilde{U}/\Gamma$. A {\em model} of dimension $k$ is the triple $(\tilde{U}/\Gamma, p, \tilde{U})$, shortly $\tilde{U}/\Gamma$.}\nolinebreak\hfill{$\triangle$}\par\vspace{0.5\parskip} \end{defn} \begin{remark}\label{simplyc}{\rm We remark that the assumption in Definition~\ref{model} that the manifold $\tilde{U}$ be simply connected could be omitted, at the expense of the definitions of smooth mapping, diffeomorphism, vector field and form, which would then become more complicated. This assumption happens to be very natural in our setting and, in practice, is not as strong as one may think. Assume in fact that the manifold $\tilde{U}$ is connected, but not simply connected. Consider its universal cover, $\pi\,\colon\,U^{\#}\rightarrow\tilde{U}$, and its fundamental group, $\Pi$; the manifold $U^{\#}$ is connected and simply connected, the mapping $\pi$ is smooth, the discrete group $\Pi$ acts smoothly, freely and properly on the manifold $\tilde{U}$ and $\tilde{U}=U^{\#}/\Pi$. Consider the extension of the group $\Gamma$ by the group $\Pi$, $1\longrightarrow \Pi\longrightarrow\Lambda \longrightarrow\Gamma\longrightarrow 1$, defined as follows $$\Lambda=\left\{\;\lambda\in\mbox{Diff}(U^{\#})\;|\;\exists\; \gamma\in\Gamma\;\mbox{s. t.}\;\pi(\lambda(u^{\#}))=\gamma\cdot \pi(u^{\#})\;\forall\; u^{\#}\inU^{\#}\;\right\}.$$ It is easy to verify that $\Lambda$ is a discrete group, that it acts on the manifold $U^{\#}$ according to the assumptions of Definition~\ref{model} and that $\tilde{U}/\Gamma=U^{\#}/\Lambda$.} \nolinebreak\hfill{$\bigtriangledown$}\par\vspace{0.5\parskip}\end{remark} \begin{defn}[Tangent space]{\rm Consider a model $(\tilde{U}/\Gamma, p, \tilde{U})$. The group $\Gamma_{\tilde{u}}=\mbox{Stab}(\tilde{u},\Gamma)$ acts on the vector space $T_{\tilde{u}}\tilde{U}$ for any point $\tilde{u}$ in $\tilde{U}$. We define the {\em tangent space} of the model $\tilde{U}/\Gamma$ at the point $u=p(\tilde{u})$, denoted $T_u(\tilde{U}/\Gamma)$, to be the space of orbits $(T_{\tilde{u}}\tilde{U})/\Gamma_{\tilde u}$.\nolinebreak\hfill{$\triangle$}\par\vspace{0.5\parskip} }\end{defn} \begin{remark}{\rm We remark that $T_u(\tilde{U}/\Gamma)$ itself defines a model and that it is a true vector space for all points $u$ in $p(\tilde{U}_0)$.} \nolinebreak\hfill{$\bigtriangledown$}\par\vspace{0.5\parskip}\end{remark} \begin{defn}[Smooth mapping, diffeomorphism of models]\label{smloc}{\rm A {\em smooth mapping} of the models $(\tilde{U}/\Gamma, p, \tilde{U})$ and $(\tilde{V}/\Delta, q, \tilde{V})$ is a mapping $f\,\colon\, \tilde{U}/\Gamma\longrightarrow \tilde{V}/\Delta$ having the property that there exists a smooth mapping $\tilde{f}\,\colon\,\tilde{U}\longrightarrow\tilde{V}$ such that $q\circ \tilde{f}=f\circ p$; we will say that $\tilde{f}$ is a {\em lift} of $f$. We will say that the smooth mapping $f$ is a {\em diffeomorphism of models} if it is bijective and if the lift $\tilde{f}$ is a diffeomorphism.}\nolinebreak\hfill{$\triangle$}\par\vspace{0.5\parskip} \end{defn} If the mapping $\tilde{f}$ is a lift of a smooth mapping of models $f\,\colon\,\tilde{U}/\Gamma\longrightarrow \tilde{V}/\Delta$ so are the mappings $\tilde{f}^{\gamma}(-)=\tilde{f}(\gamma\cdot -)$, for all elements $\gamma$ in $\Gamma$ and $^{\delta}\tilde{f}(-)=\delta\cdot\tilde{f}(-)$, for all elements $\delta$ in $\Delta$. We are about to show that if the mapping $f$ is a diffeomorphism, then these are the only other possible lifts. \begin{lemma}[The orange lemma] Consider two models, $\tilde{U}/\Gamma$ and $\tilde{V}/\Delta$, and let $f\,\colon\,\tilde{U}/\Gamma\longrightarrow\tilde{V}/\Delta$ be a diffeomorphism of models. For any two lifts, $\tilde{f}$ and $\bar{f}$, of the diffeomorphism $f$ there exists a unique element $\delta$ in $\Delta$ such that $\bar{f}={}^\delta\tilde{f}$. \end{lemma} \mbox{\bf Proof.\ \ } Let $\tilde{V}_0$ be the connected and dense set of points in the manifold $\tilde{V}$ where the action of the group $\Delta$ is free, and consider a point $\tilde{v}$ in $\tilde{V}_0$, and the corresponding point $\tilde{u}=\tilde{f}^{-1}(\tilde{v})$. Then there is a unique element $\delta(\tilde{v})$ in $\Delta$ such that $\bar{f}(\tilde{u})= \delta(\tilde{v})\cdot\tilde{f}(\tilde{u})$. Since the group $\Delta$ is discrete, and the set $\tilde{V}_0$ is connected and dense, there exists a unique element $\delta$ in $\Delta$ such that $\bar{f}={}^\delta\tilde{f}$. \nolinebreak\hfill{$\Box$}\par\vspace{0.5\parskip} \begin{lemma}[The green lemma]\label{green} Consider two models, $\tilde{U}/\Gamma$ and $\tilde{V}/\Delta$, and a diffeomorphism $f\,\colon\,\tilde{U}/\Gamma\longrightarrow\tilde{V}/\Delta$. Then, for a given lift, $\tilde{f}$, of the diffeomorphism $f$, there exists a group isomorphism $F\,\colon\,\Gamma\longrightarrow\Delta$ such that $\tilde{f}^{\gamma}={}^{F(\gamma)}\tilde{f}$, for all elements $\gamma$ in $\Gamma$. \end{lemma} \mbox{\bf Proof.\ \ } Take an element $\gamma$ in $\Gamma$. Apply the orange lemma to the lifts $\tilde{f}$, $\bar{f}=\tilde{f}^\gamma$, and define $F(\gamma)=\delta$. Repeat for all elements $\gamma$ in $\Gamma$ and check that $F$ is an isomorphism with the required property. \nolinebreak\hfill{$\Box$}\par\vspace{0.5\parskip} \begin{defn}[Vector field, $h$-form on a model]{\rm A {\em vector field}, $\mbox{X}$, [respectively {\em $h$-form}, $\omega$,] on a model $\tilde{U}/\Gamma$ is the assignment of a $\Gamma$-invariant vector field, $\tilde{\mbox{X}}$, [respectively $h$-form, $\tilde{\omega}$,] on the manifold $\tilde{U}$.\nolinebreak\hfill{$\triangle$}\par\vspace{0.5\parskip}}\end{defn} \begin{defn}[Pushforward of a vector field]{\rm Consider two models, $\tilde{U}/\Gamma$ and $\tilde{V}/\Delta$, and a diffeomorphism $f\,\colon\,\tilde{U}/\Gamma\longrightarrow\tilde{V}/\Delta$. Let $\mbox{X}$ be a smooth vector field on the model $\tilde{U}/\Gamma$; we define the {\em pushforward} of $\mbox{X}$ via $f$, denoted $f_*\mbox{X}$, to be the vector field on the model $\tilde{V}/\Delta$ that corresponds to the assignment of the $\Delta$-invariant vector field $\tilde{f}_*\tilde{\mbox{X}}$, for any lift $\tilde{f}$ of the diffeomorphism $f$.\nolinebreak\hfill{$\triangle$}\par\vspace{0.5\parskip}}\end{defn} The notions of differential and pullback of a form, and the notion of interior product of a form with a vector field are defined in an analogous way. \begin{defn}[Symplectic form on a model]{\rm A {\em symplectic form}, $\omega$, on a model $\tilde{U}/\Gamma$ is the assignment of a $\Gamma$-invariant symplectic form, $\tilde{\omega}$, on the manifold $\tilde{U}$.\nolinebreak\hfill{$\triangle$}\par\vspace{0.5\parskip}}\end{defn} We are now ready to define quasifolds. \begin{defn}[Quasifold] {\rm A dimension $k$ {\em quasifold structure} on a topological space $M$ is the assignment of an {\em atlas}, or collection of {\em charts}, ${\cal A}= \{\; (U_{\alpha},\phi_{\alpha},\ut_{\alpha}/\Gamma_{\alpha})\;|\;\alpha\in A\;\}$ having the following properties: \begin{enumerate} \item The collection $\{\;U_{\alpha}\;|\;\alpha\in A\;\}$ is a cover of $M$. \item For each index $\alpha$ in $\cal A$, the set $U_{\alpha}$ is open, the space $\ut_{\alpha}/\Gamma_{\alpha}$ defines a model, where the set $\ut_{\alpha}$ is an open, connected and simply connected subset of the space $\R^k$, and the mapping $\phi_{\alpha}$ is a homeomorphism of the space $\ut_{\alpha}/\Gamma_{\alpha}$ onto the set $U_{\alpha}$. \item For all indices $\alpha, \beta$ in $A$ such that $U_{\alpha}\capU_{\beta}\neq\emptyset$, the sets $\phi_{\alpha}^{-1}(U_{\alpha}\capU_{\beta})$ and $\phi_{\beta}^{-1}(U_{\alpha}\capU_{\beta})$ define models and the mapping $$g_{\alpha\beta}=\phi_{\beta}^{-1}\circ\phi_{\alpha}\,\colon\phi_{\alpha}^{-1}(U_{\alpha}\capU_{\beta}) \longrightarrow\phi_{\beta}^{-1}(U_{\alpha}\capU_{\beta})$$ is a diffeomorphism. We will then say that the mapping $g_{\alpha\beta}$ is a {\em change of charts} and that the corresponding charts are {\em compatible}. \item The atlas ${\cal A}$ is {\em maximal}, that is: if the triple $(U,\phi,\tilde{U}/\Gamma)$ satisfies property 2. and is compatible with all the charts in ${\cal A}$, then $(U,\phi,\tilde{U}/\Gamma)$ belongs to ${\cal A}$. \end{enumerate} We will say that a space $M$ with a quasifold structure is a {\em quasifold}. }\nolinebreak\hfill{$\triangle$}\par\vspace{0.5\parskip} \end{defn} \begin{remark}{\rm A quasifold where all the groups $\Gamma_{\alpha}$ are trivial is a manifold, one where all the groups $\Gamma_{\alpha}$ are finite is an orbifold. \nolinebreak\hfill{$\bigtriangledown$}\par\vspace{0.5\parskip}}\end{remark} \begin{ex}[The quasisphere]\label{qsphere1}{\rm Let $s,t$ be two positive real numbers such that $s/t\notin\mbox{\bbb{Q}}$. Consider the space $\C^2$ with the standard symplectic form $\omega_0=\frac{1}{2\pi i}(dz_1\wedge d\bar{z}_1 + dz_2\wedge d\bar{z}_2)$ and with the $\mbox{\bbb{R}}$-action: $(\theta, (z_1,z_2)) = ( \e z_1, e^{2\pi i\theta\frac{s}{t}} z_2)$ of moment mapping \begin{eqnarray*} \Psi\,\colon&\C^2 &\longrightarrow \mbox{\bbb{R}} \\ &(z_1,z_2)&\longmapsto |z_1|^2 +\frac{s}{t}|z_2|^2-s. \end{eqnarray*} Consider the level set $\Psi ^{-1} (0)$; this space is an ellipsoid of dimension $3$ with center the origin and radii $(\sqrt{s},\sqrt{t})$. Consider now the space of orbits $M=\Psi^{-1}(0)/\mbox{\bbb{R}}$. We want to show that it is a quasifold of dimension $2$. We cover it with two open sets, $U_S=\{\,[z_1:z_2]\in M\;|\;z_2\neq 0\,\}$ and $U_N=\{\,[z_1:z_2] \in M\;|\;z_1\neq 0\,\}$. Denote by $B(r)$, for any $r>0$, the open ball in the space $\mbox{\bbb{C}}$ of center the origin and radius $\sqrt{r}$. Then the discrete group $\Gamma_S=\mbox{\bbb{Z}}$ acts on the open set $\tilde{U}_S=B(s)$ by the rule $(k,z)\mapsto e^{2\pi ik\frac{t}{s}}\cdot z$; this action is free on the connected, dense subset $\tilde{U}_S-\{0\}$ and the mapping \begin{eqnarray*} \phi_S\,\colon\tilde{U}_S/\Gamma_S &\longrightarrow &U_S\\ \left[z\right]&\longmapsto &\left[z:\sqrt{t-\frac{t}{s}|z|^2}\right] \end{eqnarray*} is a homeomorphism. Similarly the group $\Gamma_N=\mbox{\bbb{Z}}$ acts on the open set $\tilde{U}_N=B(t)$ by the rule $(m,w)\mapsto e^{2\pi im\frac{s}{t}}\cdot w$; this action is free on the connected, dense subset $\tilde{U}_N-\{0\}$ and the mapping \begin{eqnarray*} \phi_N\,\colon\tilde{U}_N/\Gamma_N &\longrightarrow &U_N\\ \left[w\right]&\longmapsto &\left[\sqrt{s-\frac{s}{t}|w|^2}:w\right] \end{eqnarray*} is a homeomorphism. Let us check that these two charts are compatible. The set $\phi_S^{-1}(U_S\cap U_N)$ defines a model: it is the quotient of $\mbox{\bbb{R}}\times \left(0,\sqrt{s}\right)$ by the following action of $\Z^2$: $\left(\left( h,k\right) ,\left( \sigma, \rho\right)\right)\mapsto \left(\sigma+h+k\frac{t}{s},\rho\right).$ Similarly the set $\phi_N^{-1}(U_S\cap U_N)$ is the quotient of $\mbox{\bbb{R}}\times (0,\sqrt{t})$ by the following action of $\Z^2$: $\left(\left( l,m\right),\left( \tau, \upsilon \right)\right) \mapsto \left(\tau+l+m\frac{s}{t},\upsilon\right)$. Remark that \begin{eqnarray*} g_{SN}=\phi_N^{-1}\circ\phi_S\,\colon \phi_S^{-1}(U_S\cap U_N) &\longrightarrow &\phi_N^{-1}(U_S\cap U_N)\\ \left[z=e^{2\pi i\sigma}\rho\right]&\longmapsto &\left[w=e^{-2\pi i \sigma\frac{s}{t}}\sqrt{t-\frac{t}{s}\rho^2} \right] \end{eqnarray*} is a diffeomorphism of models: its lift is given by $(\sigma,\rho)\longmapsto \left( -\sigma\frac{s}{t},\sqrt{t-\frac{t}{s}\rho^2}\right)$. Now complete this collection with all other compatible charts. }\nolinebreak\hfill{$\Diamond$}\par\vspace{0.5\parskip}\end{ex} We now proceed to give quasifolds all the necessary geometrical structure. \begin{defn}[Smooth mapping, diffeomorphism of quasifolds]\label{smglob} {\rm Let $M$ and $N$ be two quasifolds. A continuous mapping $f\,\colon\, M\rightarrow N$ is said to be a {\em smooth mapping of quasifolds} if there exists a chart $(U_{\alpha}, \phi_{\alpha}, \ut_{\alpha}/\Gamma_{\alpha})$ around each point $m$ in the space $M$, a chart $(V_{\alpha}, \psi_{\alpha}, \vt_{\alpha}/\Delta_{\alpha})$ around the point $f(m)$, and a smooth mapping of models $f_{\alpha}\,\colon\,\ut_{\alpha}/\Gamma_{\alpha}\rightarrow \vt_{\alpha}/\Delta_{\alpha}$ such that $\psi_{\alpha}\circf_{\alpha}=f\circ\phi_{\alpha}$. If the smooth mapping $f$ is bijective, and if its inverse is smooth, we will say that it is a {\em diffeomorphism of quasifolds}. }\end{defn} Let us say a word about the definition of smooth mapping. Consider Definition \ref{smglob} and denote by $\tilde{\fa}$ a lift of the smooth mapping of models $f_{\alpha}$, by $p_{\alpha}$ the canonical projection $\ut_{\alpha} \rightarrow \ut_{\alpha}/\Gamma_{\alpha}$, and by $q_{\alpha}$ the canonical projection $\vt_{\alpha} \rightarrow \vt_{\alpha}/\Delta_{\alpha}$. Then, by combining Definitions \ref{smloc} and \ref{smglob}, we get that the following diagram commutes $$ \begin{array}{ccl} \ut_{\alpha} & \stackrel{\tilde{\fa}}{\mbox{\LARGE $\longrightarrow$}} &\vt_{\alpha}\\ {}\!\!\!\!\!\stackrel{p_{\alpha}}{}\,\stackrel{}{\mbox{\LARGE $\downarrow$}} & & \stackrel{}{\mbox{\LARGE $\downarrow$}}\,\stackrel{q_{\alpha}}{}\\ \ut_{\alpha}/\Gamma_{\alpha} &\stackrel{f_{\alpha}}{\mbox{\LARGE $\longrightarrow$}} & \vt_{\alpha}/\Delta_{\alpha}\\ {}\!\!\!\!\!\stackrel{\phi_{\alpha}}{}\,\stackrel{}{\mbox{\LARGE $\downarrow$}} & & \stackrel{}{\mbox{\LARGE $\downarrow$}}\,\stackrel{\psi_{\alpha}}{}\\ U_{\alpha} & \stackrel{f}{\mbox{\LARGE $\longrightarrow$}} & V_{\alpha}. \end{array} $$ Let us look at the special case $N=V$, a vector space (this includes all moment maps; see Definition~\ref{mmap}). The space $V$ is a smooth quasifold of one chart so a mapping $f\,\colon\, M\longrightarrow V$ is smooth if, and only if, there exists a chart $\phi_{\alpha}\,\colon\,\ut_{\alpha}/\Gamma_{\alpha}\longrightarrowU_{\alpha}$ around each point $m$ in the space $M$, such that the mapping $\tilde{\fa}=f\circ\phi_{\alpha}\circp_{\alpha}\,\colon\,\ut_{\alpha}\longrightarrow V$ is smooth (here $p_{\alpha}$ still denotes the canonical projection $\ut_{\alpha} \rightarrow \ut_{\alpha}/\Gamma_{\alpha}$). \begin{defn}[Vector field, $h$-form on a quasifold]\label{vff}{\rm A {\em vector field}, $X$, [respectively {\em $h$-form}, $\omega$], on a quasifold $M$ is the assignment of a chart $(U_{\alpha}, \phi_{\alpha},\ut_{\alpha}/\Gamma_{\alpha})$ around each point $m$ in the space $M$ and of a vector field, $\mbox{X}_{\alpha}$, [respectively $h$-form, $\omega_\alpha$,] on the model $\ut_{\alpha}/\Gamma_{\alpha}$. We require that whenever we have two such charts, $(U_{\alpha}, \phi_{\alpha},\ut_{\alpha}/\Gamma_{\alpha})$ and $(U_{\beta}, \phi_{\beta},\ut_{\beta}/\Gamma_{\beta})$, with the property that $U_{\alpha}\capU_{\beta}\neq\emptyset$, then $(g_{\alpha\beta})_*\mbox{X}_{\alpha} = \mbox{X}_{\beta}$ [respectively $(g_{\alpha\beta})^*\omega_\beta = \omega_\alpha$] for the corresponding change of charts $g_{\alpha\beta}$. }\nolinebreak\hfill{$\triangle$}\par\vspace{0.5\parskip}\end{defn} \begin{defn}[Pushforward of a vector field]{\rm Let $M$ and $N$ be two quasifolds, let $X$ be a vector field on the quasifold $M$, and let $f\,\colon\, M\rightarrow N$ be a diffeomorphism; then there exists a chart $(U_{\alpha}, \phi_{\alpha},\ut_{\alpha}/\Gamma_{\alpha})$ around any given point $m$ in the space $M$, a chart $(V_{\alpha}, \psi_{\alpha},\vt_{\alpha}/\Delta_{\alpha})$ around the point $n=f(m)$, a vector field $\mbox{X}_{\alpha}$ on the model $\ut_{\alpha}/\Gamma_{\alpha}$, and a smooth mapping $f_{\alpha}\,\colon\, U_{\alpha}\rightarrowV_{\alpha}$ such that $\psi_{\alpha}\circf_{\alpha}=f\circ\phi_{\alpha}$. We define the {\em pushforward} of $\mbox{X}$ via $f$, denoted $f_*\mbox{X}$, to be the vector field on the quasifold $N$ given by the assignment of the chart $(V_{\alpha}, \psi_{\alpha},\vt_{\alpha}/\Delta_{\alpha})$ around the point $n$ and of the vector field ${f_{\alpha}}_*\mbox{X}_{\alpha}$ on the model $\vt_{\alpha}/\Delta_{\alpha}$.}\nolinebreak\hfill{$\triangle$}\par\vspace{0.5\parskip}\end{defn} Completely analogous definitions hold for the notions of differential and pullback of a form, and for the notion of interior product of a form with a vector field. \begin{defn}[Symplectic form-structure-quasifold, symplectomorphism]{\rm A {\em symplectic form} on a quasifold $M$ is a $2$-form, $\omega$, such that each form $\omega_\alpha$ (see Definition~\ref{vff}) is symplectic. A {\em symplectic structure} on a quasifold $M$ is the assignment of a symplectic form $\omega$, and we will say that $(M,\omega)$, or shortly $M$, is a {\em symplectic quasifold}. A {\em symplectomorphism} between two symplectic quasifolds $(M,\omega)$ and $(N,\sigma)$ is a diffeomorphism $f\,\colon\, M\longrightarrow N$ such that $f^*\sigma=\omega$.\nolinebreak\hfill{$\triangle$}\par\vspace{0.5\parskip}}\end{defn} \begin{ex}[Quasilinear model]\label{darboux1}{\rm Let $V$ be a symplectic vector space with a linear, effective and symplectic action of a torus $T$. Take any discrete subgroup $\Gamma\subset T$, and consider its induced action on the space $V$. The group $\Gamma$ acts freely on a connected, dense subset of the space $V$, thus the space of orbits $V_{\G}=V/\Gamma$ is a symplectic quasifold of dimension $2l=\dim{V}$. }\nolinebreak\hfill{$\Diamond$}\par\vspace{0.5\parskip}\end{ex} The quasisphere in Example~\ref{qsphere1} can also be endowed with a symplectic structure. \begin{ex}[Quasisphere]\label{qsphere2}{\rm Consider the quasisphere of Example~\ref{qsphere1} and define a symplectic form by assigning the $\Gamma_S$-invariant symplectic form $\tilde{\omega}_S=\frac{1}{2\pi i}\, dz\wedge d\bar{z}$ to the set $\tilde{U}_S$ and the $\Gamma_N$-invariant symplectic form $\tilde{\omega}_N=\frac{1}{2\pi i}\, dw\wedge d\bar{w}$ to the set $\tilde{U}_N$. }\nolinebreak\hfill{$\Diamond$}\par\vspace{0.5\parskip}\end{ex} \section{Quasitori and their actions on quasifolds}\label{tori} We devote this section to quasitori and their Hamiltonian actions on symplectic quasifolds. We start with a number of definitions and properties and we end with some crucial examples. \begin{defn}[Quasilattice, quasitorus]{\rm Let $\mbox{\frak d}$ be a vector space of dimension $n$. A {\em quasilattice} in $\mbox{\frak d}$ is the $\mbox{\bbb{Z}}$-span, $Q$, of a set of $\mbox{\bbb{R}}$-spanning vectors $X_1,\ldots,X_d$ in $\mbox{\frak d}$. We call {\em quasitorus} of dimension $n$ the group and quasifold of one chart $D=\mbox{\frak d}/Q$.}\nolinebreak\hfill{$\triangle$}\par\vspace{0.5\parskip} \end{defn} Notice that in the previous definition $d\geq n$ and that if $d=n$, then the quasilattice $Q$ is a lattice and the quasitorus $D$ is a honest torus. A quasitorus is compact, connected and abelian, and the group operations of multiplication and inversion are smooth quasifold mappings. \begin{ex}[A quasicircle]\label{quasicircle}{\rm The first example of a (non-smooth) quasitorus is the quasitorus of dimension $1$ ({\em quasicircle}) $D^1=\mbox{\bbb{R}}/Q$, where $Q=s\mbox{\bbb{Z}}+t\mbox{\bbb{Z}}$, $s/t\notin\mbox{\bbb{Q}}$. To discover everything about this innocuous-looking group we refer the reader to \cite{di,i,il}. }\nolinebreak\hfill{$\Diamond$}\par\vspace{0.5\parskip}\end{ex} The quasifold tangent space at the identity of a quasitorus $D=\mbox{\frak d}/Q$ is always the vector space $\mbox{\frak d}$. By analogy with the smooth case we make the following \begin{defn}[Quasi-Lie algebra, exponential mapping] {\rm Let $D=\mbox{\frak d}/Q$ be a quasitorus. We define the {\em quasi-Lie algebra} of $D$ to be the vector space $\mbox{\frak d}$. The natural projection of $\mbox{\frak d}$ onto $D$ is called {\em exponential mapping}, denoted $\exp_{D}$, or simply $\exp$.}\nolinebreak\hfill{$\triangle$}\par\vspace{0.5\parskip} \end{defn} \begin{defn}[Quasitorus homomorphism, isomorphism and epimorphism] {\rm A group homomorphism [respectively epimorphism and isomorphism] between quasitori that is a smooth quasifold map is called {\em quasitorus homomorphism} [respectively {\em epimorphism} and {\em isomorphism}].}\nolinebreak\hfill{$\triangle$}\par\vspace{0.5\parskip} \end{defn} Given two quasitori, $D_1=\mbox{\frak d}_1/Q_1$ and $D_2=\mbox{\frak d}_2/Q_2$, and a quasitorus homomorphism $f\,\colon\, D_1\rightarrow D_2$, it is easy to check that the unique lift, $\tilde{f}$, of the homomorphism $f$ satisfying $\tilde{f} (0)=0$ is a linear mapping $\tilde{f}\,\colon\, (\mbox{\frak d}_1,Q_1)\rightarrow (\mbox{\frak d}_2,Q_2)$, and is an epimorphism, respectively isomorphism, whenever the homomorphism $f$ is. Again by analogy with honest tori, we will call this lift the quasi-Lie algebra homomorphism associated to the quasitorus homomorphism $f$. The following proposition explains why we are interested in quasitori. \begin{prop}\label{groupquotient} Let $T$ be a torus and $N$ a Lie subgroup \footnote{We allow and actually prefer immersed subgroups.}. Then $T/N$ is a quasitorus of dimension $n=\dim{T}-\dim{N}$. \end{prop} \mbox{\bf Proof.\ \ } Choose a complement, $\mbox{\frak d}$, of the vector subspace $\mbox{\frak n}=\mbox{Lie}(N)$ in the vector space $\mbox{\frak t}=\mbox{Lie}(T)$; consider the surjective mapping $p_{\mbox{\sfrak d}}=\Pi\circ{\exp_T}|_{\mbox{\frak d}} \,\colon\,\mbox{\frak d}\longrightarrow T/N$ where $\Pi\,\colon\, T\longrightarrow T/N$ denotes the canonical projection. Then the set $Q=\ker{p_{\mbox{\sfrak d}}}$ is a quasilattice (a lattice if the group $N$ is compact) and the mapping $p_{\mbox{\sfrak d}}$ induces a group isomorphism $\mbox{\frak d}/Q\simeq T/N$. Notice that two different choices of a complement $\mbox{\frak d}$ yield isomorphic quasitori; the group $T/N$ thus inherits a well defined structure of quasitorus. \nolinebreak\hfill{$\Box$}\par\vspace{0.5\parskip} \vspace{.4cm} \noindent We remark that the subspace $\mbox{\frak d}$ of the preceding proof is the quasi-Lie algebra of the quasitorus $D\simeq T/N$ and that $p_{\mbox{\sfrak d}}=\exp_D$. One important special case is the quotient of a torus $T$ by any of its discrete subgroups, $\Gamma$. In this case we have $T/\Gamma=\mbox{\frak d}/Q$, where $\mbox{\frak d}\simeq\mbox{\frak t}$. Another example is the quotient of a two-dimensional torus by an immersed line of slope $s/t\notin\mbox{\bbb{Q}}$ ({\em Kronecker foliation}); the corresponding quasitorus is the quasicircle of Example~\ref{quasicircle}. \begin{defn}[Smooth action] {\rm A {\em smooth action} of a quasitorus $D$ on a quasifold $M$ is a smooth mapping $\tau\,\colon\, D\times M\longrightarrow M$ such that $\tau(d_1\cdot d_2,m)=\tau(d_1,\tau(d_2,m))$ and $\tau(1_{D},m)=m$ for all elements $d_1, d_2$ in the quasitorus $D$ and for each point $m$ in the space $M$.} \nolinebreak\hfill{$\triangle$}\par\vspace{0.5\parskip}\end{defn} According to this definition, there exist charts $(U_{\alpha},\phi_{\alpha},\ut_{\alpha}/\Gamma_{\alpha})$ and $(V_{\alpha},\psi_{\alpha},\vt_{\alpha}/\Delta_{\alpha})$ around each point $m$ in the space $M$, and smooth mappings $\tilde{\tau}_{\alpha},\tau_{\alpha}$ such that the following diagram commutes $$ \begin{array}{ccl} \mbox{\frak d}\times \ut_{\alpha} & \stackrel{\tilde{\tau}_{\alpha}}{\mbox{\LARGE $\longrightarrow$}} & \vt_{\alpha} \\ {}\!\!\!\!\!\stackrel{}{}\,\stackrel{}{\mbox{\LARGE $\downarrow$}} & & \stackrel{}{\mbox{\LARGE $\downarrow$}}\,\stackrel{}{}\\ D\times(\ut_{\alpha}/\Gamma_{\alpha}) & \stackrel{\tau_{\alpha}}{\mbox{\LARGE $\longrightarrow$}} & \vt_{\alpha}/\Delta_{\alpha}\\ {}\!\!\!\!\!\stackrel{}{}\,\stackrel{}{\mbox{\LARGE $\downarrow$}} & & \stackrel{}{\mbox{\LARGE $\downarrow$}}\,\stackrel{}{}\\ D\timesU_{\alpha} & \stackrel{\tau}{\mbox{\LARGE $\longrightarrow$}} & V_{\alpha} \end{array} $$ Notice that, since $\tau(1_{D},p)=p$ for each point $p$ in the space $M$, we have that the set $U_{\alpha}$ is contained in the set $V_{\alpha}$; it is therefore possible to assume that $\tilde{\tau}_{\alpha}(0,\tilde{u})= \tilde{u}$ for each point $\tilde{u}$ in the set $\ut_{\alpha}$, and that the set $\ut_{\alpha}$ is contained in the set $\vt_{\alpha}$. Now fix an element $X$ in the space $\mbox{\frak d}$; then, for small enough real numbers $t$, the points $\tilde{\tau}_{\alpha}(tX,\tilde{u})$ belong to the set $\ut_{\alpha}$ whenever the point $\tilde{u}$ does. These data allow us to define the fundamental vector field of the smooth action $\tau$. \begin{defn}[Fundamental vector field]{\rm Consider a smooth action, $\tau$, of a quasitorus $D=\mbox{\frak d}/Q$ on a quasifold $M$. For any element $X$ in the space $\mbox{\frak d}$ we define a vector field $\mbox{X}_M$ on the space $M$, called {\em fundamental vector field of the action} corresponding to $X$, which is given by the assignment, for each point $m$ in the space $M$, of the chart $(U_{\alpha},\phi_{\alpha},\ut_{\alpha}/\Gamma_{\alpha})$ (see discussion above) and of the $\Gamma_{\alpha}$-invariant vector field on the set $\ut_{\alpha}$ given by $$ \tilde{\mbox{X}}_M(\tilde{u})=\frac{d}{dt}|_0 \,\tilde{\tau}_{\alpha}(tX,\tilde{u}),\quad \tilde{u}\in\ut_{\alpha}. $$ }\nolinebreak\hfill{$\triangle$}\par\vspace{0.5\parskip}\end{defn} Notice that, for a fixed element $d$ in the quasitorus $D$, the mapping $\tau_{d}(-)=\tau(d,-)$ is a diffeomorphism of the quasifold $M$. \begin{defn}[Hamiltonian action, moment mapping]\label{mmap}{\rm A smooth action, $\tau$, of a quasitorus $D=\mbox{\frak d}/Q$ on a symplectic quasifold $(M,\omega)$ is {\em Hamiltonian} if it preserves the symplectic form ($\tau_{d}^*\omega=\omega$ for all $d$ in the quasitorus $D$) and if there exists a smooth $D$-invariant mapping $\Phi\,\colon\, M\rightarrow\d^*$, which we call {\em moment mapping}, such that $\imath(\mbox{X}_M)\omega= d<\Phi,X>$, for each element $X$ in the space $\mbox{\frak d}$.}\nolinebreak\hfill{$\triangle$}\par\vspace{0.5\parskip} \end{defn} \begin{ex}[The quasilinear model]\label{darboux3}{\rm Consider the quasilinear model $V_{\G}$ of Example~\ref{darboux1}. The linear, effective and symplectic action of the torus $T$ on the space $V$ is Hamiltonian and it can be described as follows. Write $T=\mbox{\frak t}/L$, where $\mbox{\frak t}$ denotes the Lie algebra of the torus $T$ and $L$ is the lattice $\ker\exp_T$, and consider the corresponding weight lattice $$L^*=\{\;\mu\in\t^*\;|\;\mu(X)\in\mbox{\bbb{Z}}\quad\forall X\in L\;\}.$$ The space $V$ decomposes into $l$ complex $1$-dimensional $T$-invariant subspaces $V_j$ and there exist weights $\alpha_j$ in the lattice $L^*$, $j=1,\ldots, l$, such that the action is given by $$ \begin{array}{cccccl} \hat{\tau}\,\colon& T&\times&V&\longrightarrow& V\\ &(\exp_T(X)&,&v)&\longmapsto&(e^{2\pi i \alpha_1(X)} v_1,\ldots, e^{2\pi i \alpha_l(X)} v_l), \end{array} $$ and the moment mapping is given by $$ \begin{array}{cccl} \hat{\Phi}\,\colon&V&\longrightarrow &\t^*\\ &v&\longmapsto & \sum_{j=1}^l |v_j|^2 \alpha_j. \end{array} $$ The image of $\hat{\Phi}$ is the rational convex polyhedral cone $\hat{\cal C}$ of vertex $O$ and spanned by the weights $\alpha_j$. Denote by $p$ the projection $V\rightarrowV_{\G}$, by $D$ the quasitorus $\mbox{\frak d}/Q\simeq T/\Gamma$, by $\Pi$ the projection $T\rightarrow D$, and by $\pi\,\colon(\mbox{\frak t},L)\rightarrow(\mbox{\frak d},Q)$ the corresponding quasi-Lie algebra isomorphism. The action of the torus $T$ on the vector space $V$ induces an action, $\tau$, of the quasitorus $D$ on the space $V_{\G}$ as follows $$ \begin{array}{ccl} T\times V & \stackrel{\hat{\tau}}{\mbox{\LARGE $\longrightarrow$}} & V \\ {}\!\!\!\!\!\stackrel{\Pi\times p}{}\,\stackrel{}{\mbox{\LARGE $\downarrow$}} & & \stackrel{}{\mbox{\LARGE $\downarrow$}}\,\stackrel{p}{}\\ D\timesV_{\G} & \stackrel{\tau}{\mbox{\LARGE $\longrightarrow$}} & V_{\G}. \end{array} $$ This action is Hamiltonian and the corresponding moment mapping $\Phi$ is given by $$ \begin{array}{ccl} V & \stackrel{\hat{\Phi}}{\mbox{\LARGE $\longrightarrow$}} & \t^*\\ {}\!\!\!\!\!\stackrel{p}{}\,\stackrel{}{\mbox{\LARGE $\downarrow$}} & & \stackrel{}{\mbox{\LARGE $\uparrow$}}\,\stackrel{\pi^*}{}\\ V_{\G} & \stackrel{\Phi}{\mbox{\LARGE $\longrightarrow$}} &\d^*. \end{array} $$ Notice that the image of the mapping $\Phi$ is the convex polyhedral cone ${\cal C}=(\pi^*)^{-1}(\hat{\cal C})$, which is spanned by the elements $\beta_j=(\pi^*)^{-1}(\alpha_j)$ in the space $\d^*$. }\nolinebreak\hfill{$\Diamond$}\par\vspace{0.5\parskip}\end{ex} \begin{ex}[The quasisphere]\label{qsphere3}{\rm Let us go back to the quasisphere $M$ of Examples~\ref{qsphere1} and \ref{qsphere2}. Consider now the quasilattice $Q=s\mbox{\bbb{Z}}+t\mbox{\bbb{Z}}$ and the quasicircle $D^1=\mbox{\bbb{R}}/Q$. The mapping $$ \begin{array}{cccl} \tau\,\colon&D^1\times M&\longrightarrow &M\\ &([\theta],[z:w])&\longmapsto &[\eones z:w] \end{array} $$ defines a Hamiltonian action of the quasicircle $D^1$ (a quasirotation) on the quasifold $M$, with moment mapping $$ \begin{array}{cccl} \Phi\,\colon&M&\longrightarrow &\R^*\\ &[z:w]&\longmapsto &\frac{|z|^2}{s}=1-\frac{|w|^2}{t}. \end{array} $$ Notice finally that $\Phi(M)=[0,1]$ just like for truly rotating spheres, teardrops, or rugby balls. }\nolinebreak\hfill{$\Diamond$}\par\vspace{0.5\parskip}\end{ex} We conclude with an example of a honest torus acting on a quasifold. This example has a different flavor than all the others that we treat. \begin{ex}[The horocycle foliation]{\rm Let us consider the upper half-plane ${\cal H}=\{\,(x,y)\in\R^2\;|\;\mbox{s.t.}\quad y>0\}$ with the standard symplectic form $dx\wedge dy$. We let the group $\mbox{\bbb{Z}}$ act on the space ${\cal H}$ as follows: $(k,(x,y))\longmapsto(x+ky,y)$. This action is free and symplectic. We now consider the following free and Hamiltonian $S^1$-action on the quotient space ${\cal H}/\mbox{\bbb{Z}}$: $(\e, [x:y])\longmapsto [x+\theta y:y]$; the moment mapping is given by $[x:y]\longmapsto \frac{1}{2}y^2$.}\nolinebreak\hfill{$\Diamond$}\par\vspace{0.5\parskip}\end{ex} \section{From simple polytopes to symplectic quasifolds} The aim of this section is to extend Delzant's construction and to show that any simple convex polytope is the image of the moment mapping for a family of effective Hamiltonian quasitorus actions on symplectic quasifolds of the appropriate dimension. This is a consequence of the following symplectic reduction theorem. \label{delzant} \begin{thm}\label{reduction} Let $T$ be a torus of Lie algebra $\mbox{\frak t}$, let $T\times X \longrightarrow X$ be a Hamiltonian action of the torus $T$ on a symplectic manifold $X$ and assume that the moment mapping $J\,\colon\, X \longrightarrow \t^*$ is proper. Consider the induced action of any Lie subgroup $N$ of $T$ and suppose that 0 is a regular value of the corresponding moment mapping $\Psi\,\colon \,X \longrightarrow \n^*$ ($\mbox{\frak n}$ denotes the Lie algebra of $N$). Then $M=\Psi^{-1}(0)/N$ is a symplectic quasifold of dimension $\dim{X}-2\dim{N}$ and the induced ($T/N$)-action on the quasifold $M$ is Hamiltonian. \end{thm} \mbox{\bf Proof.\ \ } The slice theorem (see \cite{k}) applied to the $T$-action on the manifold $\Psi^{-1}(0)$ gives invariant neighborhoods of the orbits $T\cdot x$ that are of the form $T\times_{T_x} B_x$, where $T_x=\mbox{Stab}(x,T)$, and $B_x$ is an open ball in the space $T_x(\psi^{-1}(0))/T_x(T\cdot x)$. The quotient $(T\times_{T_x} B_x)/N$ is a ($T/N$)-invariant neighborhood of the orbit $(T/N)\cdot [x]$ in the space $M$. Let us check that this neighborhood is a quasifold chart; the argument is quite similar to the one in the proof of Proposition~\ref{groupquotient}. Denote by $\mbox{\frak t}_x$ the Lie algebra of the group $T_x$. Since the value $0$ is regular for the mapping $\Psi$, we have that $\mbox{\frak t}_x\cap\mbox{\frak n}=\{0\}$; choose a complement $\mbox{\frak d}_x$ of the vector subspace $\mbox{\frak t}_x\oplus\mbox{\frak n}$ in the space $\mbox{\frak t}$. Denote by $\Pi_x$ the projection $T\times_{T_x} B_x\longrightarrow (T\times_{T_x} B_x)/N$ and define a surjective mapping $p_x\,\colon\,\mbox{\frak d}_x\times B_x\longrightarrow (T\times_{T_x} B_x)/N$ according to the following rule: $p_x(Y,b)=\Pi_x([\exp_T{Y}:b])$, $(Y,b)\in\mbox{\frak d}_x\times B_x$. Now consider the quasilattice $Q$ of the proof of Proposition~\ref{groupquotient} chosen relatively to the complement $\mbox{\frak d}=\mbox{\frak d}_x\oplus\mbox{\frak t}_x$ of the subspace $\mbox{\frak n}$ in the space $\mbox{\frak t}$. It is easy to check that the discrete group $$\Lambda_x=\left\{\;(Y_Q,\exp_TT_Q)\in\mbox{\frak d}_x\times T_x\;|\; Y_Q+T_Q\in Q\;\right\}$$ acts on the connected, simply connected open set $\mbox{\frak d}_x\times B_x$ as follows $$ \begin{array}{cccccc} &\Lambda_x&\times&(\mbox{\frak d}_x\times B_x)&\longrightarrow& \mbox{\frak d}_x\times B_x\\ &((Y_Q,\exp_TT_Q)&,&(Y,b))&\longmapsto&(Y+Y_Q,\exp_TT_Q\cdot b), \end{array} $$ and that the mapping $p_x$ induces a homeomorphism $(\mbox{\frak d}_x\times B_x)/\Lambda_x\simeq (T\times_{T_x} B_x)/N$. The remainder of the proof proceeds like the proof of the classical symplectic reduction theorem. The symplectic form on the manifold $X$ induces a $\Lambda_x$-invariant symplectic form on the open set $\mbox{\frak d}_x\times B_x$, thus a symplectic form on each chart $(T\times_{T_x} B_x)/N$; similarly the action of the torus $T$ on the manifold $X$ induces a Hamiltonian action of the quasitorus $T/N$ on each chart, the corresponding moment mapping being induced by the one for the $T$-action on the manifold $X$. The required compatibility properties are satisfied. \nolinebreak\hfill{$\Box$}\par\vspace{0.5\parskip} \begin{remark}[Quasi-universal covers]{\rm We like to think of the manifolds $U^{\#}$ in Remark~\ref{simplyc}, $\mbox{\frak d}$ in the proof of Proposition~\ref{groupquotient}, and $\mbox{\frak d}_x\times B_x$ in the proof of Theorem~\ref{reduction}, as the {\em quasi-universal covers} of the quasifolds $\tilde{U}/\Gamma$, $T/N$ and $(T\times_{T_x} B_x)/N$, respectively; the discrete groups $\Lambda$, $Q$ and $\Lambda_x$ would then be the corresponding fundamental groups. If the group $\Gamma$ were finite and the group $N$ were compact this would be in agreement with Thurston's notion of orbifold universal cover. \nolinebreak\hfill{$\bigtriangledown$}\par\vspace{0.5\parskip}}\end{remark} Let us now apply Theorem~\ref{reduction} to extend Delzant's construction. Let $\mbox{\frak d}$ be a vector space of dimension $n$. The key idea is the observation that any simple convex polytope in the dual space $\d^*$ can be obtained by slicing a translate of the positive ortant of the space $(\rd)^*$ with an appropriate subspace. \begin{thm}\label{del} Let $\mbox{\frak d}$ be a vector space of dimension $n$. For any simple convex polytope $\Delta\subset\d^*$ there exists an $n$-dimensional quasitorus $D$ of quasi-Lie algebra $\mbox{\frak d}$, a $2n$-dimensional compact symplectic quasifold $M$, and an effective Hamiltonian action of the quasitorus $D$ on the quasifold $M$ such that the image of the corresponding moment mapping is the polytope $\Delta$. \end{thm} \mbox{\bf Proof.\ \ } Consider the space $\C^d$ endowed with the standard symplectic form $\omega_0=\frac{1}{2\pi i}\sum_{j=1}^d dz_j\wedge d\bar{z}_j$ and the standard action of the torus $T^d=\R^d/\Z^d$: $$ \begin{array}{cccccl} \tau\,\colon& T^d&\times&\C^d&\longrightarrow& \C^d\\ &((\et1,\ldots,e^{2\pi i\theta_d})&,&\underline{z})&\longmapsto&(\et1 z_1,\ldots, e^{2\pi i\theta_d} z_d). \end{array} $$ This action is effective and Hamiltonian and its moment mapping is given by $$ \begin{array}{cccl} J\,\colon&\C^d&\longrightarrow &(\rd)^*\\ &\underline{z}&\longmapsto & \sum_{j=1}^d |z_j|^2 e_j^*+\lambda,\quad\lambda\in(\rd)^* \;\mbox{constant}. \end{array} $$ The mapping $J$ is proper and its image is the cone ${\cal C}_\lambda=\lambda+{\cal C}_0$, where ${\cal C}_0$ denotes the positive ortant in the space $(\rd)^*$. Write the polytope $\Delta$ as in the appendix, formula~(\ref{polydecomp}) and consider the surjective linear mapping \begin{eqnarray*} \pi\,\colon &\R^d \longrightarrow \mbox{\frak d},\\ &e_j \longmapsto X_j. \end{eqnarray*} Let $Q$ be any quasilattice in the vector space $\mbox{\frak d}$ containing the vectors $X_1,\ldots,X_d$ (for example $Q=\sum_{j=1}^d X_j\mbox{\bbb{Z}}$), and consider the dimension $n$ quasitorus $D=\mbox{\frak d}/Q$. Then the linear mapping $\pi$ induces a quasitorus epimorphism $\Pi\,\colon\,T^d \longrightarrow D$. Define now $N$ to be the kernel of the mapping $\Pi$ and choose $\lambda=\sum_{j=1}^d \lambda_j e_j^*$. Then, according to Theorem~\ref{reduction}, the quasitorus $T^d/N$ acts in a Hamiltonian fashion on the symplectic quasifold $M=\Psi^{-1}(0)/N$. Denote by $i$ the Lie algebra inclusion $\mbox{Lie}(N)\rightarrow\R^d$. If we identify the quasitori $D$ and $T^d/N$ using the epimorphism $\Pi$, we get a Hamiltonian action of the quasitorus $D$ whose moment mapping has image equal to ${(\pi^*)}^{-1}({\cal C}_\lambda\cap\ker{i^*})= {(\pi^*)}^{-1}({\cal C}_\lambda\cap\mbox{im}\,\pi^*)= {(\pi^*)}^{-1}(\pi^*(\Delta))=\Delta$. This action is effective since the level set $\Psi^{-1}(0)$ contains points of the form $\underline{z}\in\C^d$, $z_j\neq0$, $j=1,\ldots,d$, where the $T^d$-action is free. Notice finally that $\dim{M}=2d-2\dim{N}= 2d-2(d-n)=2n=2\dim{D}$. \nolinebreak\hfill{$\Box$}\par\vspace{0.5\parskip} \begin{remark}[Uniqueness?] {\rm Notice that we had many choices in this construction. To begin with, the pairs $(X_j,\lambda_j)$ in (\ref{polydecomp}) are far from being unique; moreover there are infinitely many quasilattices that contain a fixed choice of the vectors $X_j$. As a consequence, the quasitorus, quasifold and action are far from being unique (see Example~\ref{segment} below), but we will return to this matter in future work. For the moment we just point out that if the polytope $\Delta$ is rational relatively to a lattice $L$, by choosing the elements $X_j$ to be in the lattice $L$, and the quasilattice $Q$ to be equal to the lattice $L$ itself, we distinguish among our spaces a family of orbifolds, in accordance with \cite{lt}; if the polytope $\Delta$ also satisfies Delzant's integrality condition, by taking the elements $X_j$ to be primitive in the lattice $L$, we obtain a manifold, in accordance with \cite{d}. \nolinebreak\hfill{$\bigtriangledown$}\par\vspace{0.5\parskip}}\end{remark} We conclude this section with three telling examples, where we apply the construction described in Theorem~\ref{del} to three different polytopes. \begin{ex}[The unit interval]\label{segment} {\rm As a first example we consider the unit interval $[0,1]\subset \R^*$. We apply the construction with the choice of vectors $X_1=s, X_2=-t$, $s, t\in\R_+^*$, and with the corresponding quasilattice $Q=X_1\mbox{\bbb{Z}}+X_2\mbox{\bbb{Z}}$. We leave it as an exercise to show that if $s/t\notin\mbox{\bbb{Q}}$ we obtain the quasisphere of Examples~\ref{qsphere2} and \ref{qsphere3}, while in the remaining cases we get the standard sphere, and its orbifold cousins, the teardrop and rugby ball. }\nolinebreak\hfill{$\Diamond$}\par\vspace{0.5\parskip}\end{ex} \begin{ex}[The right triangle] {\rm As a second example we consider the right triangle in $(\rtwo)^*$ of vertices $(0,0)$, $(s,0)$ and $(0,t)$, where $s,t$ are two positive real numbers such that $s/t\notin\mbox{\bbb{Q}}$. We apply the construction with the choice of vectors $X_1=(1,0)$, $X_2=(0,1)$, $X_3=(-t,-s)$ and with the corresponding quasilattice $Q=X_1\mbox{\bbb{Z}}+X_2\mbox{\bbb{Z}}+X_3 \mbox{\bbb{Z}}$. Then we have $\lambda_1=\lambda_2=0$, $\lambda_3= -st$ and a linear mapping \begin{eqnarray*} \pi\,\colon &(\R^3,\Z^3) &\longrightarrow (\R^2,Q)\\ &(x,y,z) &\longmapsto (x-tz,y-sz) \end{eqnarray*} that induces a quasitorus homomorphism $\Pi\,\colon\,T^3\rightarrow D^2=\R^2/Q$ whose kernel is given by $$N=\{\,(e^{2\pi i\sigma t},e^{2\pi i\sigma s},e^{2\pi i\sigma})\;|\; \sigma\in\mbox{\bbb{R}}\,\}.$$ Consider now the standard action $\tau\,\colon\,T^3\times\C^3\longrightarrow\C^3$ with moment mapping given by $$ \begin{array}{cccl} J\,\colon&\C^3&\longrightarrow &(\rthree)^*,\\ &\underline{z}&\longmapsto &(|z_1|^2,|z_2|^2,|z_3|^2-st). \end{array} $$ Then the $N$-moment mapping is given by $$ \begin{array}{cccl} \Psi\,\colon&\C^3&\longrightarrow &\R^*\\ &\underline{z}&\longmapsto &t|z_1|^2+s|z_2|^2+|z_3|^2-st, \end{array} $$ and $$\Psi^{-1}(0)=\{\,\underline{z}\in\C^3\quad|\quad t|z_1|^2+ s|z_2|^2+|z_3|^2=st\,\}$$ is the dimension $5$ ellipsoid of center the origin and of radii $(\sqrt{s},\sqrt{t},\sqrt{st})$. The quasitorus $D^2$ acts on the quasifold $M=\Psi^{-1}(0)/N$ with moment mapping $$ \begin{array}{cccl} \Phi\,\colon&M&\longrightarrow &(\rtwo)^*\\ &[\underline{z}]&\longmapsto &(|z_1|^2,|z_2|^2), \end{array} $$ and $\Phi(M)=\Delta$. We call the quasifold $M$ projective quasispace, by analogy with the case of the rational right triangle ($s/t\in\mbox{\bbb{Q}}$), which gives either a weighted or an ordinary projective space.}\nolinebreak\hfill{$\Diamond$}\par\vspace{0.5\parskip} \end{ex} The unit interval and the right triangle are actually rational (with respect to the appropriate choice of lattices). Here comes finally an example of a polytope that is not. \begin{ex}[The regular pentagon] {\rm Let us take the regular pentagon in $(\rtwo)^*$. We choose the vectors $X_1=(1,0), X_2=(a,b)$, $X_3=(c,d)$, $X_4=(c,-d)$, $X_5=(a,-b)$ and the corresponding quasilattice $Q=\sum_{j=1}^{5}X_j\mbox{\bbb{Z}}$, where $a=\cos{\frac{2\pi}{5}}$, $b=\sin{\frac{2\pi}{5}}$, $c=\cos{\frac{4\pi}{5}}$, $d=\sin{\frac{4\pi}{5}}$. Then we have $\lambda_1=\lambda_2= \lambda_3=\lambda_4=\lambda_5=c$ and a linear mapping \begin{eqnarray*} \pi\,\colon &(\R^5,\Z^5) &\longrightarrow (\R^2,Q)\\ &(x_1,x_2,x_3,x_4,x_5) &\longmapsto (x_1+a(x_2+x_5)+c(x_3+x_4), b(x_2-x_5)+d(x_3-x_4)). \end{eqnarray*} that induces a quasitorus homomorphism $\Pi\,\colon\,T^5\rightarrow D^2=\R^2/Q$ whose kernel is given by $$N=\{\,(e^{2\pi i\phi},\e,e^{2\pi i\sigma},e^{2\pi i[2a(\theta-\sigma)+\phi)]}, e^{2\pi i[2a(\theta-\phi)+\sigma)]})\;|\; (\phi,\theta,\sigma) \in\R^3 \,\}.$$ Consider now the standard action $\tau\,\colon\,T^5\times\C^5\longrightarrow\C^5$ with moment mapping given by $$ \begin{array}{cccl} J\,\colon&\C^5&\longrightarrow &(\rfive)^*,\\ &\underline{z}&\longmapsto &(|z_1|^2+c,|z_2|^2+c,|z_3|^2+c,|z_4|^2+c,|z_5|^2+c). \end{array} $$ Then the $N$-moment mapping is given, for $\underline{z}\in\C^5$, by $\Psi(\underline{z})=-\left(\frac{\sqrt{5}}{2},\sqrt{5}c, \frac{\sqrt{5}}{2}\right)+$ $$ \left(|z_1|^2+|z_4|^2-2a|z_5|^2, |z_2|^2+2a(|z_4|^2+|z_5|^2),|z_3|^2+|z_5|^2-2a|z_4|^2\right), $$ and $\Psi^{-1}(0)=$ $$\left\{|z_1|^2+|z_4|^2-2a|z_5|^2= |z_3|^2+|z_5|^2-2a|z_4|^2=\frac{\sqrt{5}}{2}, |z_2|^2+2a(|z_4|^2+|z_5|^2)=\sqrt{5}c\right\}.$$ The quasitorus $D^2$ acts on the quasifold $M=\Psi^{-1}(0)/N$ and $\Phi(M)=\Delta$. }\nolinebreak\hfill{$\Diamond$}\par\vspace{0.5\parskip} \end{ex} \setcounter{sect}{1} \setcounter{thm}{0}\def\Alph{sect}.\arabic{equation}{\Alph{sect}.\arabic{equation}} \setcounter{equation}{0}\def\Alph{sect}.\arabic{thm}{\Alph{sect}.\arabic{thm}} \vspace{4.5ex}
\section{Introduction} \label{intro:ch3} In van Hoof \& Van de Steene (1999, Paper~I) we presented and tested a new method to derive simultaneously and self-consistently all physical parameters of a planetary nebula from a set of observed quantities. A modified version of the photo-ionization code {\sc cloudy}\ (Ferland 1993) is used to calculate various models, searching for a best fit of the predictions to the observables in an automated way. This method uses emission line ratios, the angular diameter, the radio and the infrared flux to constrain the model. It also takes dust into account in the radiative transport. With this method we are able to determine the stellar temperature and luminosity, the inner, Str\"omgren\ and outer radius of the nebula, the density, the dust-to-gas mass ratio and the abundances. We investigated the accuracy of the determination of the physical parameters by applying this method to an artificial set of observables. First we proved that this method can pass a formal convergence test. Subsequently we introduced either measurement errors in the observables or changed the model assumptions, and investigated how this affects the best-fit model. In this way we gained an understanding of the robustness of our method and hence of the reliability of the physical parameters. Our method was also compared with classical methods to determine the electron temperature and density and nebular abundances. It was shown that our method suffers less from noise in the spectrum than the classical line diagnostics. However, this advantage may be lost if the model assumptions are not appropriate for the nebula being studied. The weakest points are currently the use of a blackbody approximation, the assumption that the inner dust radius coincides with the inner gas radius and the assumption of spherical symmetry. Distance determinations of Planetary Nebulae (PNe) are still very problematic. Various methods are in use, but the range in distances obtained is often very large and no method has found general acceptance. Reviews of the current status can be found in Pottasch \cite{c2:rev}, Terzian \cite{c2:terzian} and Pottasch \cite{c2:pot:dens}. The lack of a reliable, generally applicable method to determine distances to PNe poses a problem when using photo-ionization models. To circumvent this problem we applied the method to a small sample of galactic bulge nebulae, which can be assumed to be all at the same well-known distance. Our aim is to study the accuracy of the determination of physical parameters by comparing our results with other literature values. A summary of the method and model assumptions is given in Section~\ref{method}. The sample selection is presented in Section~\ref{sample} and the modelling results in Section~\ref{results}. Each PN in the sample is discussed individually, with special emphasis on the problems encountered during the modelling in Section~\ref{remarks}. The resulting physical parameters are discussed by comparing them with results from other studies in the literature in Section~\ref{discussion}. Our conclusions are given in Section~\ref{conclusions}. \section{Summary of the model assumptions and the method} \label{method} The model assumptions and the method were extensively described and discussed in Paper~I. This section presents only a brief summary. To model the planetary nebula, we use a modified version of the photo-ionization code {\sc cloudy}\ 84.06 (Ferland 1993). The model for the PN is quite simple, and comprises the following assumptions: \begin{enumerate} \renewcommand{\theenumi}{(\arabic{enumi})} \item{The central star has a blackbody spectrum.} \item{The nebula is spherically symmetric.} \item{The density is constant inside the Str\"omgren\ radius of the nebula, and varies as $1/r^2$ outside.} \item{Dust grains are intermixed with the gas at a constant dust-to-gas mass ratio; if no information on the composition is available they are assumed to be a mixture of graphite and silicates.} \item{The filling factor, describing the small scale clumpiness of the gas, can be fixed at any value. If no information is available it is taken to be unity.} \item{The distance to the nebula is fixed by an independent individual or statistical method.} \end{enumerate} The above assumptions leave the following free parameters: the stellar temperature, the luminosity of the central star, the hydrogen density in the ionized region, the inner radius of the nebula, the dust to gas ratio, and the abundances in the nebula. \\ The outer radius of the nebula is not fixed as an input parameter, but calculated from the long wavelength end of the dust emission, or, as a fail-safe, when the electron density drops below 0.1~cm$^{-3}$. Adopting certain values for the input parameters, it is possible to calculate a model for the nebula with {\sc cloudy}, predicting the continuum and line fluxes, photometric magnitudes (including the contribution of line emission) and the Str\"omgren\ radius. To compare the model predictions with the observed quantities, a goodness-of-fit estimator is calculated. This estimator is minimized by varying all the input parameters of the model, using the algorithm {\sc amoeba}\ (Press et~al. 1986). It is assumed that there exists a unique set of input parameters, for which the resulting model predictions give the best fit to a given set of observables. These input parameters are then considered the best estimate for the physical properties of the PN. The full set of observed quantities necessary to derive the physical parameters of a PN are: \begin{enumerate} \renewcommand{\theenumi}{(\arabic{enumi})} \item The emission line spectrum of the nebula. Usually this is an optical spectrum, but might also be an ultraviolet and/or infrared spectrum. The line ratios make it possible to constrain the stellar temperature, the density and the electron temperature in the nebula. They are also required to determine the abundances. For elements for which no lines are available we assume standard abundances (Aller \& Czyzak 1983). \item Since dust is included in the model we also need information on the mid- and far-infrared continuum. For this the {\it IRAS}\ fluxes are used. \item To constrain the emission measure, either an optically thin radio continuum measurement (e.g. at 6~cm) is needed, or the absolute flux value of some hydrogen recombination line (usually \relax\ifmmode{\rm H\beta}\else{\rm H$\beta$}\fi). \item An accurate angular diameter $\Theta_{\rm d}$ of the nebula is needed, which we define as $\Theta_{\rm d} = 2r_{\rm str}/D$. Here $r_{\rm str}$ stands for the Str\"omgren\ radius of the nebula and $D$ is the distance to the nebula. \end{enumerate} \section{The sample of galactic bulge PNe} \label{sample} We selected a small sample of galactic bulge nebulae from Ratag et al.\ (1997, RPDM). Galactic bulge nebulae can be assumed to be all at a distance of approximately 7.8~kpc (Feast 1987). We chose the nebulae from RPDM since they publish good quality spectra and also carried out their own photo-ionization analysis of the data which we can use for comparison. The radio observations for these PNe are described in Gathier et al.\ \cite{c2:gathier}. The following selection criteria were used: \begin{enumerate} \renewcommand{\theenumi}{(\arabic{enumi})} \item The PNe should have a quality 2 or 3 {\it IRAS}\ 12~\relax\ifmmode{\mu{\rm m}}\else{$\mu$m}\fi\ flux and quality 3 {\it IRAS}\ 25~\relax\ifmmode{\mu{\rm m}}\else{$\mu$m}\fi\ and 60~\relax\ifmmode{\mu{\rm m}}\else{$\mu$m}\fi\ fluxes. \item The absolute value for the radial velocity should be larger than 100~\relax\ifmmode{\rm km\,s^{-1}}\else{km\,s$^{-1}$}\fi. \item The excitation class should not be labelled peculiar. \end{enumerate} The resulting five PNe are presented in Table~\ref{ratag:sam}. All nebulae except M~2$-$4 are indicated by Acker et al.\ \cite{c2:ack:cat} as likely bulge PNe. In view of the large radial velocity of M~2$-$4, $v_{\rm LSR}$ = $-$175.8~\relax\ifmmode{\rm km\,s^{-1}}\else{km\,s$^{-1}$}\fi\ (Gathier et al. 1983) it is unlikely to be a foreground object. \begin{table} \caption{Our sample of PNe selected from Ratag et al.\ \protect\cite{c2:ratag}.} \label{ratag:sam} \begin{tabular}{ll@{\hspace{6mm}}r@{\hspace{2mm}}r@{\hspace{2mm}}r@{}lr@{\hspace{2mm}}r@{\hspace{2mm}}r} \hline & & \ctr{4}{\hspace{-3mm}$\alpha$(2000)}{0} & \ctr{3}{$\delta$(2000)}{-3} \\ \ctr{1}{name}{0} & \ctr{1}{PN G}{6} & h & m & s && \degr & \arcmin & \arcsec \\ \\ H~1$-$40 & 359.7$-$02.6 & 17 & 55 & 36&.0 & $-30$ & 33 & 33 \\ M~1$-$20 & 006.1$+$08.3 & 17 & 28 & 57&.5 & $-19$ & 15 & 53 \\ M~2$-$4 & 349.8$+$04.4 & 17 & 01 & 06&.2 & $-34$ & 49 & 39 \\ M~2$-$23 & 002.2$-$02.7 & 18 & 01 & 42&.6 & $-28$ & 25 & 44 \\ M~3$-$15 & 006.8$+$04.1 & 17 & 45 & 31&.6 & $-20$ & 58 & 02 \\ \hline \end{tabular} \end{table} \section{Modelling results} \label{results} In Table~\ref{ratag:inp} we give the input values for the observables used for the modelling, together with the resulting model predictions. As can be seen from this table, not all the lines present in the spectra are predicted by {\sc cloudy}, most notably the higher Balmer lines of hydrogen and several helium lines. Also the element chlorine is not included in the code. The resulting physical parameters for the nebulae are given in Table~\ref{ratag:phys}. The hydrogen density shown in this table is the constant density within the Str\"omgren\ sphere. \begin{table*} \small \caption{Comparison of the observed quantities (mostly taken from Ratag et al. 1997) and the model fit for our sample of PNe. The strength of the emission lines is given relative to \relax\ifmmode{\rm H\beta}\else{\rm H$\beta$}\fi\ = 100. The measured line fluxes have been dereddened. The entries \fb{O}{ii} $\lambda$ 3727, $\lambda$ 7325 and \fb{S}{ii} $\lambda$ 4071 all stand for the entire multiplet. All observables for which entries in both columns obs.\ and model are present, have been weighted in the goodness-of-fit estimator, except where indicated.} \label{ratag:inp} \begin{tabular}{l@{}r @{\hspace{6mm}} r@{.}l@{\x{1.5}}r@{.}l@{}r@{.}l@{\x{1.5}}r@{.}l@{}r@{.}l@{\x{1.5}}r@{.}l@{}r@{.}l@{\x{1.5}}r@{.}l@{}r@{.}l@{\x{1.5}}r@{.}l} \hline \ctr{1}{ion}{4} & \ctr{1}{$\lambda$}{6} & \ctr{4}{H~1$-$40}{8} & \ctr{4}{M~1$-$20}{7} & \ctr{4}{M~2$-$4}{6} & \ctr{4}{M~2$-$23}{6} & \ctr{4}{M~3$-$15}{2} \\ & \ctr{1}{\AA}{6} & \ctr{2}{\x{-0.5}obs.}{0} & \ctr{2}{\x{-1}model}{8} & \ctr{2}{\x{-1.5}obs.}{0} & \ctr{2}{\x{-1}model}{7} & \ctr{2}{\x{-1}obs.}{0} & \ctr{2}{\x{-1.5}model}{6} & \ctr{2}{\x{0.5}obs.}{0} & \ctr{2}{\x{-1}model}{7} & \ctr{2}{\x{-1}obs.}{0} & \ctr{2}{\x{-3}model}{0} \\ \hline \fb{O}{ii} & 3727 & 32&4\rlap{:}& 38&5 & 55&5 & 83&1 & 97&1 & 125&2 & 14&2 & 20&4 & 48&6 & 101&5 \\ H\,12 & 3750 & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & 3&4 & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} \\ H\,11,\al{O}{iii} & 3771 & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & 4&2 & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} \\ H\,10 & 3798 & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & 5&5 & \multicolumn{2}{c}{} & 4&0 & \multicolumn{2}{c}{} & 5&1 & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} \\ H\,9 & 3835 & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & 6&7 & \multicolumn{2}{c}{} & 7&1 & \multicolumn{2}{c}{} & 6&5 & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} \\ \fb{Ne}{iii} & 3869 & 79&1 & 79&4 & 69&1 & 67&6 & 64&7 & 67&6 & 82&4 & 80&0 & 89&9 & 90&1 \\ H\,8,\al{He}{i} & 3889 & 12&9 & \multicolumn{2}{c}{} & 18&1 & \multicolumn{2}{c}{} & 17&4 & \multicolumn{2}{c}{} & 13&9 & \multicolumn{2}{c}{} & 16&6 & \multicolumn{2}{c}{} \\ \fb{Ne}{iii},\relax\ifmmode{\rm H\epsilon}\else{\rm H$\epsilon$}\fi & 3969 & 19&9 & \multicolumn{2}{c}{} & 25&5 & \multicolumn{2}{c}{} & 33&6 & \multicolumn{2}{c}{} & 38&3 & \multicolumn{2}{c}{} & 25&4 & \multicolumn{2}{c}{} \\ \al{He}{i} & 4026 & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & 2&7 & \multicolumn{2}{c}{} & 2&7 & \multicolumn{2}{c}{} & 1&82 & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} \\ \fb{S}{ii} & 4071 & 3&3 & 2&5 & \multicolumn{2}{c}{} & 2&3 & 3&6 & 3&0 & 2&5 & 3&8 & \multicolumn{2}{c}{} & 1&8 \\ \relax\ifmmode{\rm H\delta}\else{\rm H$\delta$}\fi,\al{N}{iii} & 4102 & 25&8 & 29&4 & 25&6 & 30&3 & 24&7 & 30&5 & 24&0 & 30&2 & 26&1 & 29&5 \\ \al{C}{ii} & 4267 \\ \relax\ifmmode{\rm H\gamma}\else{\rm H$\gamma$}\fi & 4340 & 47&1 & 50&2 & 49&1 & 51&1 & 45&9 & 51&2 & 48&2 & 50&9 & 49&0 & 50&2 \\ \fb{O}{iii} & 4363 & 4&6\rlap{$^\ddagger$} & 12&0 & 7&5 & 5&4 & 2&9 & 2&1 & 13&9 & 9&9 & 3&3\rlap{:} & 7&5 \\ \al{He}{i} & 4472 & 6&5 & 4&4 & 5&8 & 5&2 & \multicolumn{2}{c}{} & 4&5 & 5&1 & 5&5 & 5&0 & 5&1 \\ \al{N}{iii} & 4641 & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & 20&3 & \multicolumn{2}{c}{} \\ \al{He}{ii} & 4686 & 17&1?\rlap{$^\ddagger$}& 1&9 & \multicolumn{2}{c}{} & 0&4 & \multicolumn{2}{c}{} & 0&14 & \multicolumn{2}{c}{} & 0&6 & 4&1 & 3&5 \\ \fb{Ar}{iv},\al{He}{i} & 4712 & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & 1&33 & \multicolumn{2}{c}{} & 0&91 & \multicolumn{2}{c}{} & 1&09 & \multicolumn{2}{c}{} & 1&94 & \multicolumn{2}{c}{} \\ \fb{Ar}{iv} & 4740 & 4&45 & 4&70 & 0&71 & 1&16 & 0&33\rlap{:}& 0&69 & 0&80 & 1&42 & \multicolumn{2}{c}{} & 6&67 \\ \relax\ifmmode{\rm H\beta}\else{\rm H$\beta$}\fi & 4861 & 100& & 100& & 100& & 100& & 100& & 100& & 100& & 100& & 100& & 100& \\ \al{He}{i} & 4922 & 1&36\rlap{:}& \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & 1&42 & \multicolumn{2}{c}{} & 0&87 & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} \\ \fb{O}{iii} & 4959 & 307& & 276& & 336&\rlap{$^\dagger$}& 315& & 272&\rlap{$^\dagger$} & 213& & 304& & 350& & 328&\rlap{$^\dagger$}& 219& \\ \fb{O}{iii} & 5007 & 915& & 827& & 1009&\rlap{$^\dagger$}& 946& & 815&\rlap{$^\dagger$} & 640& & 1006& & 1051& & 983&\rlap{$^\dagger$}& 658& \\ \fb{N}{i} & 5201 & \multicolumn{2}{c}{} & 0&20 & \multicolumn{2}{c}{} & 0&07 & \multicolumn{2}{c}{} & 0&22 & \multicolumn{2}{c}{} & 0&00 & \multicolumn{2}{c}{} & 0&59 \\ \fb{Cl}{iii} & 5517 & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & 0&21 & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} \\ \fb{Cl}{iii} & 5538 & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & 0&47 & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & 0&25 & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} \\ \fb{N}{ii} & 5755 & 1&92\rlap{:}& 2&16 & 0&92 & 1&05 & 1&39 & 1&52 & 1&19 & 1&06 & 1&0\rlap{:} & 1&4 \\ \al{He}{i} & 5876 & 15&5 & 15&9 & 15&7 & 16&0 & 13&8 & 13&8 & 17&2 & 17&8 & 16&3 & 16&3 \\ \fb{O}{i} & 6300 & 2&8 & 2&7 & 4&7 & 4&1 & 2&9 & 3&7 & 4&2 & 3&3 & 2&8 & 3&6 \\ \fb{S}{iii} & 6312 & 1&70 & 1&48 & 0&79\rlap{:}& 1&12 & 1&59 & 1&33 & 2&4 & 2&3 & 1&24 & 1&00 \\ \fb{O}{i} & 6364 & 0&88 & 0&88 & 1&46 & 1&35 & 1&00 & 1&23 & 1&42 & 1&09 & 0&79 & 1&19 \\ \fb{N}{ii} & 6548 & \multicolumn{2}{c}{} & 19&9 & \multicolumn{2}{c}{} & 14&6 & \multicolumn{2}{c}{} & 32&1 & \multicolumn{2}{c}{} & 6&4 & \multicolumn{2}{c}{} & 18&8 \\ \relax\ifmmode{\rm H\alpha}\else{\rm H$\alpha$}\fi & 6563 & 280& & 278& & 303& & 269& & 275& & 268& & 283& & 269& & 305& & 278& \\ \fb{N}{ii} & 6584 & 61&4 & 59&7 & 45&4 & 43&8 & 85&6 & 96&2 & 18&9 & 19&2 & 57&6 & 56&3 \\ \al{He}{i} & 6678 & 3&8 & \multicolumn{2}{c}{} & 4&0 & \multicolumn{2}{c}{} & 3&2 & \multicolumn{2}{c}{} & 4&7 & \multicolumn{2}{c}{} & 4&6 & \multicolumn{2}{c}{} \\ \fb{S}{ii} & 6716 & 0&99 & 0&96 & 1&17 & 1&45 & 2&72 & 3&16 & 0&78 & 0&47 & 2&65 & 3&38 \\ \fb{S}{ii} & 6731 & 1&71 & 1&99 & 2&32 & 2&91 & 5&0 & 6&0 & 1&55 & 1&05 & 5&4 & 5&2 \\ \al{He}{i} & 7065 & 7&9 & 11&4 & 10&5 & 9&7 & 5&6 & 7&1 & 14&4 & 12&0 & 7&5 & 9&2 \\ \fb{Ar}{iii} & 7136 & 13&6 & 13&3 & 9&1 & 7&5 & 15&2 & 13&0 & 14&0 & 11&2 & 19&2 & 19&3 \\ \al{He}{i} & 7281 & \multicolumn{2}{c}{} & \multicolumn{2}{c}{} & 0&83 & \multicolumn{2}{c}{} & 0&3\rlap{:} & \multicolumn{2}{c}{} & 0&94 & \multicolumn{2}{c}{} & 0&59 & \multicolumn{2}{c}{} \\ \fb{O}{ii} & 7325 & 9&2 & 14&2 & 14&9 & 17&1 & 8&3 & 14&0 & 19&3 & 21&4 & 6&6 & 8&8 \\ \\ \ctr{1}{obs.}{2} & unit \\ \\ $F_\nu$(12 \relax\ifmmode{\mu{\rm m}}\else{$\mu$m}\fi) & Jy & 2&38 & 2&38 & 1&13 & 1&00 & 0&56 & 0&53 & 1&93 & 2&10 & $<0$&53\rlap{$^\S$} & 0&19 \\ $F_\nu$(25 \relax\ifmmode{\mu{\rm m}}\else{$\mu$m}\fi) & Jy & 18&45 & 19&11 & 3&94 & 4&44 & 5&00 & 5&83 & 9&31 & 6&54 & 5&66 & 6&02 \\ $F_\nu$(60 \relax\ifmmode{\mu{\rm m}}\else{$\mu$m}\fi) & Jy & 11&91 & 11&42 & 2&38 & 2&30 & 5&77 & 5&18 & 1&64 & 1&64 & 8&02 & 7&77 \\ $F_\nu$(100 \relax\ifmmode{\mu{\rm m}}\else{$\mu$m}\fi) & Jy &$<73$&48 & 3&22 & $<4$&59 & 0&66 & $<12$&59 & 1&73 & $<126$&70 & 0&24 & $<10$&39& 2&72 \\ $F_\nu$(6 cm) & mJy & 31& & 31&0 & 47& & 47&7 & 32& & 32&2 & 41& & 41&5 & 65& & 65&4 \\ $\Theta_{\rm d}$ & arcsec & 1&26 & 1&27 & 1&98 & 1&81 & 2&16 & 2&13 & 0&72 & 0&67 & 5&4 & 5&19 \\ \\ $\chi^2$ & & \multicolumn{2}{c}{} & 0&63 & \multicolumn{2}{c}{} & 1&68 & \multicolumn{2}{c}{} & 2&28 & \multicolumn{2}{c}{} & 4&17 & \multicolumn{2}{c}{} & 2&75 \\ \hline \end{tabular} \ntd{\rlap{:}\x{4}A colon indicates that the value is uncertain.} \ntd{\rlap{$^\dagger$}\x{4}The sum of the intensities of the doublet was split using the ratio 3:1.} \ntd{\rlap{$^\ddagger$}\x{4}This line was not weighted in the goodness-of-fit estimator $\chi^2$, see also Section~\ref{remarks}.} \ntd{\rlap{$^\S$}\x{4}This flux is not listed as an upper limit in the {\it IRAS}\ Point Source Catalog, see also Section~\ref{remarks}.} \normalsize \end{table*} \section{Individual remarks} \label{remarks} The PNe in our sample all have nearly the same medium excitation class. This probably is partially a result of our selection criterion that the nebulae should have been detected by {\it IRAS}\ in the 12~\relax\ifmmode{\mu{\rm m}}\else{$\mu$m}\fi\ band (criterion 1). Old bulge PNe, having a high excitation class and cool dust, might have insufficient 12~\relax\ifmmode{\mu{\rm m}}\else{$\mu$m}\fi\ flux to be detected by {\it IRAS}. In the rest of this section each of the PNe in our sample will be discussed individually, with special emphasis on the problems encountered during the modelling. \subsection{H~1$-$40} \label{pni:ind} Two lines were omitted from the list of observables because of the following reasons. First the \al{He}{ii} $\lambda$ 4686 line was omitted, because the flux ratio given by RPDM is quite high, indicative of a high stellar temperature. However, the rest of the observational data are not consistent with such a high stellar temperature. Also, this line is listed in Table~3 of RPDM, but is not present in their Table~1. Webster (1988, W88) took a spectrum of this PN, and she didn't report the detection of this line. She should however have detected a line of the strength mentioned by RPDM. Tylenda et al. \cite{ty94} list an upper limit of 5 for the intensity of this line. In view of these uncertainties we decided to omit this line. Since RPDM included this line in their modelling, this probably explains the higher stellar temperature they obtain. The fitting of the \fb{O}{iii} $\lambda$ 4363 line was also problematic. The observed flux was far too low to be consistent with the electron temperature predicted by our model. Since the electron temperature derived from the \fb{N}{ii} line ratio is much higher (and more consistent with the value determined by our model), and also because the \fb{O}{iii} $\lambda$ 4363 line is much stronger in the spectrum of W88 (however not as strong as predicted by our model), we decided that its value was too uncertain and omitted it from the input. \begin{table} \footnotesize \caption{The physical parameters of the galactic bulge PNe in our sample determined with {\sc cloudy}. Abundances of elements for which only one line was observed are marked uncertain. Since we only model the core region of M~2$-$23, no values for the outer radius and total shell mass are entered} \label{ratag:phys} \begin{tabular}{l@{\x{2.0}}r@{\x{2.0}}r@{\x{2.0}}r@{\x{2.0}}r@{\x{2.0}}r} \hline & H~1$-$40 & M~1$-$20 & M~2$-$4 & M~2$-$23 & M~3$-$15 \\ \\ ${\rm log}(T_\ast$/K) & 4.800 & 4.774 & 4.705 & 4.782 & 4.916 \\ ${\rm log}(L_\ast$/\relax\ifmmode{\rm L_\odot}\else{L$_\odot$}\fi) & 3.798 & 3.607 & 3.555 & 3.639 & 3.663 \\ ${\rm log}(n_{\rm H}$/cm$^{-3}$) & 4.321 & 4.124 & 3.923 & 4.855 & 3.527 \\ $r_{\rm in}$/mpc & 13 & 0.21 & 11 & 9 & 33 \\ $r_{\rm str}$/mpc & 24 & 34 & 40 & 13 & 98 \\ $r_{\rm out}$/mpc & 280\rlap{:} & 360\rlap{:} & 350\rlap{:} & & 520\rlap{:} \\ $M_{\rm ion}$/\relax\ifmmode{\rm M_\odot}\else{M$_\odot$}\fi & 0.042 & 0.092 & 0.088 & 0.015 & 0.47 \\ $M_{\rm sh}$/\relax\ifmmode{\rm M_\odot}\else{M$_\odot$}\fi & 1.3\rlap{:} & 2.3\rlap{:} & 1.9\rlap{:} & & 6.5\rlap{:} \\ ${\rm log}\Gamma$ & $-1.70$ & $-3.11$ & $-2.50$ & $-2.46$ & $-2.60$ \\ $\epsilon$(He) & 10.96 & 11.02 & 10.96 & 11.05 & 11.03 \\ $\epsilon$(N) & 7.78 & 7.81 & 8.13 & 7.67 & 7.60 \\ $\epsilon$(O) & 8.23 & 8.72 & 8.84 & 8.67 & 8.22 \\ $\epsilon$(Ne) & 7.39\rlap{:} & 7.82\rlap{:} & 8.15\rlap{:} & 7.75\rlap{:} & 7.53\rlap{:} \\ $\epsilon$(S) & 6.37 & 6.66 & 6.90 & 6.79 & 6.31 \\ $\epsilon$(Ar) & 5.98 & 5.99 & 6.36 & 6.08 & 6.20\rlap{:} \\ $T_{\rm e}$/kK & 12.7 & 9.5 & 8.3 & 10.2 & 12.0 \\ ${\rm log} U$ & $-1.40$ & $+2.14$ & $-1.16$ & $-1.80$ & $-1.58$ \\ \hline \end{tabular} \nt{\rlap{:}\x{4}A colon indicates that the value is uncertain.} \normalsize \end{table} \subsection{M~1$-$20} \label{pnii:ind} The intensity of the \relax\ifmmode{\rm H\alpha}\else{\rm H$\alpha$}\fi\ line seems quite high, and is not fitted well. The discrepancy is too large to be attributed to measurement errors, hence this might indicate that the spectrum has not been sufficiently dereddened. There is however no evidence from the fits to the other lines to support this suspicion. Our model gives a very small inner radius, also resulting in a very high ionization parameter. This is caused by the high {\it IRAS}\ 12~\relax\ifmmode{\mu{\rm m}}\else{$\mu$m}\fi\ over 25~\relax\ifmmode{\mu{\rm m}}\else{$\mu$m}\fi\ flux ratio, which might indicate the presence of hot dust. See also the discussion in Paper I. \subsection{M~2$-$4} \label{pniii:ind} The spectrum is fitted quite well, but there is slight discrepancy for the \fb{O}{iii} $\lambda$ 4959 and $\lambda$ 5007 lines. This is caused by the \fb{O}{ii} $\lambda$ 3727 doublet, which is not fitted well. The latter doublet usually has a larger uncertainty due to extinction and detector insensitivity. \subsection{M~2$-$23} \label{pniv:ind} This PN has the highest $\chi^2$ of all PNe in our sample. This is mainly caused by the weak lines, which might indicate that this spectrum has a lower signal-to-noise when compared to the other spectra. RPDM do not list error margins for their line flux ratios, so we had to assume reasonable values. The model is not able to fit the {\it IRAS}\ 25~\relax\ifmmode{\mu{\rm m}}\else{$\mu$m}\fi\ flux, which is very high compared both to the 12~\relax\ifmmode{\mu{\rm m}}\else{$\mu$m}\fi\ and 60~\relax\ifmmode{\mu{\rm m}}\else{$\mu$m}\fi\ flux. A possible explanation could be the presence of a 30~\relax\ifmmode{\mu{\rm m}}\else{$\mu$m}\fi\ dust feature in the spectrum (Hoare 1990). This would imply that the nebula is carbon-rich, since this feature has only been observed in carbon-rich nebulae. The central star has spectral type Of (Aller \& Keyes 1987, AK87). The large difference between the optical diameter of 8.5\arcsec\ (Acker et al. 1992) and the radio diameter of 0.72\arcsec\ (Gathier et al. 1983) suggests that this nebula might be a core-halo nebula. All other evidence gathered in this paper also is consistent with this assumption and we will adopt it throughout the paper. Since we used the radio diameter for the modelling, our model is only valid for the core region. The fact that our model is density bounded and gives a low ionized mass is consistent with the fact that we are only modelling the core region. \subsection{M~3$-$15} \label{pnv:ind} There is a suggestion of a systematic trend when comparing the observed and the modelled line flux as a function of wavelength. Also the observed intensity of the \relax\ifmmode{\rm H\alpha}\else{\rm H$\alpha$}\fi\ line seems quite high. This might indicate that the spectrum has not been sufficiently dereddened. This PN has a [WC]-type central star (AK87). The central star temperature, the excitation class 5.5 (taken from RPDM) and the low {\it IRAS}\ 12~\relax\ifmmode{\mu{\rm m}}\else{$\mu$m}\fi\ to 25~\relax\ifmmode{\mu{\rm m}}\else{$\mu$m}\fi\ flux ratio all are consistent with an early spectral type: [WC3-4] (cf.\ Kaler 1989, M\'{e}ndez \& Niemela 1982, and Zijlstra et al. 1994, respectively). The {\it IRAS}\ 12~\relax\ifmmode{\mu{\rm m}}\else{$\mu$m}\fi\ flux is not listed as an upper limit in the Point Source Catalogue. However, when we used this value for the modelling, the resulting model was unrealistic. We therefore assume that the 12~\relax\ifmmode{\mu{\rm m}}\else{$\mu$m}\fi\ flux suffers from confusion and took the quoted value to be an upper limit. See also the discussion in Paper I. \section{Discussion} \label{discussion} In this section the modelling results are discussed by comparing them with results from other studies in the literature. Since distance dependent parameters are usually not given by other authors, we will restrict ourselves to the distance independent parameters of PNe. \subsection{Stellar temperatures} In Table~\ref{temp:tab} we present a comparison with the stellar temperatures given in the literature. These temperatures were derived using the Zanstra method and photo-ionization modelling. Results using the energy balance or Stoy method are not listed since we consider this method, or at least the data for the nebulae being studied here, to be unreliable (Pottasch, private communication). One can see that the derived values agree quite well with only a few outliers. The temperatures determined by our method agree well with the hydrogen Zanstra temperatures, with the single exception of the temperature for M~2$-$23 given by Tylenda et al.\ \cite{c2:tylenda}. Since the other three determinations using the Zanstra method agree quite well, we assume the value given by Tylenda et al.\ \cite{c2:tylenda} to be erroneous. \begin{table} \caption{Comparison of the stellar temperatures for the PNe in our sample. The temperatures are given in kilokelvin. The abbreviations for the methods have the following meaning: \al{H}{i}~-- hydrogen Zanstra method, \al{He}{ii}~-- helium Zanstra method, AM~-- photo-ionization modelling using model atmospheres, BB~-- photo-ionization modelling using blackbody approximation.} \label{temp:tab} \begin{tabular}{r@{\x{5}}r@{\x{5}}r@{\x{5}}r@{\x{5}}r@{\x{5}}r@{\x{2}}r} \hline H~1$-$40\hspace*{-2.5mm} & M~1$-$20\hspace*{-2.5mm} & M~2$-$4\hspace*{-2.5mm} & M~2$-$23\hspace*{-2.5mm} & M~3$-$15\hspace*{-2.5mm} & ref. & meth.\\ \\ & & 55.\phantom{0}& 64.\phantom{0}& & 3 & \al{H}{i} \\ & 53.\phantom{0}& & 56.\phantom{0}& & 4 & \al{H}{i} \\ & & 51.5 & 65.0 & & 6 & \al{H}{i} \\ & 65.\phantom{0}& & 85.\phantom{0}& & 8 & \al{H}{i} \\ & 49.5 & 49.9 & & & 5 & \al{H}{i} \\ & 59.9 & & & & 5 &\al{He}{ii}\\ & & & 50.\phantom{0}& 62.5 & 1 & AM \\ 80.0\rlap{?}& 50.0 & 50.0 & 57.5 & 72.5 & 7 & AM \\ 64.\phantom{0} & & & & & 2 & BB \\ 63.1 & 59.4 & 50.7 & 60.5 & 82.4 & 9 & BB \\ \hline \end{tabular} \nt{References --- 1. Aller \& Keyes \cite{c2:aller:keyes} using model atmospheres by Husfeld et al.\ \cite{c2:husfeld} 2. Dopita et al.\ \cite{c2:dopita} 3. Gleizes, Acker \& Stenholm \cite{c2:gleizes} 4. Kaler \& Jacoby \cite{c2:kej} 5. Mal'kov \cite{malkov} 6. Pottasch \& Acker \cite{c2:pot:ack} 7. Ratag et al.\ \cite{c2:ratag} using model atmospheres by Clegg \& Middlemass \cite{c2:cem} and Husfeld et al.\ \cite{c2:husfeld} 8. Tylenda et al.\ \cite{c2:tylenda} 9. This work} \end{table} To derive stellar temperatures for photo-ionization modelling, sometimes certain line-ratios are used as temperature indicators (e.g.\ \al{He}{ii} $\lambda$ 4686 over \relax\ifmmode{\rm H\beta}\else{\rm H$\beta$}\fi). Especially for cooler central stars, where few temperature sensitive lines are available, this makes the determination dependent on one or two lines. Nevertheless, the results from the photo-ionization models usually are in good agreement. Exceptions are the temperature for H~1$-$40 derived by RPDM, and the temperatures for M~3$-$15. The deviating value for H~1$-$40 given by RPDM can probably be attributed to the \al{He}{ii} $\lambda$ 4686 line, which they used as a temperature indicator. We refer to the discussion in Section~\ref{pni:ind}. For M~3$-$15 we find a higher stellar temperature than other authors. The largest discrepancy is with the value from AK87. This can probably be attributed to the fact that AK87 did not report a detection of the \al{He}{ii} $\lambda$ 4686 line in their spectrum (although a detection of roughly the same strength as RPDM was reported in Aller \& Keyes 1985). Since M~3$-$15 has a [WC]-type central star, part of the \al{He}{ii} $\lambda$ 4686 flux may originate from the central star. Unfortunately, no detection of the \fb{Ar}{iv} $\lambda$ 4740 line has ever been reported, so that no alternative temperature sensitive line is available. In view of this, the central star temperature for M~3$-$15 should be viewed with some caution. We conclude that the temperature determination for the central stars in this sample is fairly reliable, although the situation for M~3$-$15 is not completely clear. This confirms our results from Paper I in which we found the temperature determination to be robust. \subsection{Electron temperatures} \label{te:litt} In Table~\ref{te:tab} the electron temperatures derived by different authors are compared. The electron temperature determined by {\sc cloudy}\ is a weighted mean of the temperature in the nebula: $\overline{T_{\rm e}}$ = $\int n_{\rm e}^2 T_{\rm e} \mbox{\rm d} V / \int n_{\rm e}^2 \mbox{\rm d} V$. The observational material shows a large spread in most cases, even when the same method is used. This indicates that the electron temperature determination, at least in those cases where diagnostic lines have been used, is not very reliable. This is in agreement with our results in Paper I. Note the large difference between the \fb{N}{ii} and \fb{O}{iii} temperatures in the case of M~2$-$23. This difference is not caused by measurement error. For this particular object, the temperature derived from the \fb{N}{ii} lines has no physical meaning (Liu, private communication). \begin{table} \caption{Various determinations of the electron temperature for the nebulae in our sample. The temperatures are given in kilokelvin. The abbreviations for the methods have the following meaning: ave.~-- average of \fb{N}{ii} and \fb{O}{iii}, model~-- average model prediction (see text).} \label{te:tab} \begin{tabular}{r@{\x{5}}r@{\x{5}}r@{\x{5}}r@{\x{5}}r@{\x{5}}r@{\x{2}}r} \hline H~1$-$40\hspace*{-2.5mm} & M~1$-$20\hspace*{-2.5mm} & M~2$-$4\hspace*{-2.5mm} & M~2$-$23\hspace*{-2.5mm} & M~3$-$15\hspace*{-2.5mm} & ref. & meth.\\ \\ & & & 18.0 & & 1 & \fb{N}{ii} \\ & 17.1 & & & & 2 & \fb{N}{ii} \\ & 12.5 & 12.5 & & & 4 & \fb{N}{ii} \\ & & 10.2 & & & 7 & \fb{N}{ii} \\ & 10.4 & & & & 8 & \fb{N}{ii} \\ 13.1 & 10.2 & 9.7 & 19.2 & 9.4\rlap{:}& 9 & \fb{N}{ii} \\ 10.1 & & & & & 11 & \fb{N}{ii} \\ & & & 11.0 & & 1 & \fb{O}{iii} \\ & 10.8 & & & & 2 & \fb{O}{iii} \\ & 13.1 & 11.6 & & & 4 & \fb{O}{iii} \\ & & 8.7 & & & 5 & \fb{O}{iii} \\ & & & 12.9 & & 6 & \fb{O}{iii} \\ & 9.9 & & & & 8 & \fb{O}{iii} \\ 9.3 & 10.4 & 8.5 & 13.0 & 8.4\rlap{:}& 9 & \fb{O}{iii} \\ 9.7 & & & & & 11 & \fb{O}{iii} \\ & & & 12.6 &11.2\rlap{:}& 3 & ave. \\ & & 11.1 & & & 10 & ave. \\ 12.7 & 9.5 & 8.3 & 10.2 & 12.0 & 12 & model \\ \hline \end{tabular} \nt{\rlap{:}\x{4}A colon indicates that the value is uncertain.} \nt{References --- 1. Acker et~al.\ \cite{c2:ackko} 2. Acker et~al. \cite{c2:ackko2} 3. Aller \& Keyes \cite{c2:aller:keyes} 4. Costa et al.\ \cite{c2:costa} 5. Cuisinier, Acker \& K\"oppen \cite{c2:cuisinier} 6. Kaler \cite{c2:kaler:i} 7. Kaler et al. \cite{c2:kalerea} 8. Kaler et al. \cite{c2:kalerea2} 9. Ratag et al.\ \cite{c2:ratag} 10. Tylenda et al. \cite{c2:tylenda2} 11. Webster \cite{c2:webster} 12. This work} \end{table} The electron temperatures derived from our method are in most cases just outside the range of values found with line diagnostics; three times at the low end and twice at the high end. The results from Paper~I indicate that the electron temperature determination with our method should be robust. It is not apparent to us why the average values of the electron temperature derived from line diagnostics do not coincide with our results. This might indicate a problem, although the fact that we find both higher and lower results is not indicative of a systematic effect. Nevertheless, this issue should be investigated further in future research, using a larger sample. \subsection{Electron densities} \label{eden:disc} In Table~\ref{ne:tab} the electron densities derived by different authors are compared. The electron density determined by {\sc cloudy}\ is a weighted mean of the density in the nebula: $\overline{n_{\rm e}}$ = $\int n_{\rm e}^3 \mbox{\rm d} V / \int n_{\rm e}^2 \mbox{\rm d} V$. There are enormous differences between the various determinations in the literature, even when the same method has been applied. This indicates that the determination of densities with line diagnostics is unreliable, which confirms our results in Paper~I. Also note the enormous differences between the \fb{S}{ii}, \fb{Cl}{iii} and \fb{Ar}{iv} densities for M~1$-$20 derived by Kaler et al. \cite{c2:kalerea2}. \begin{table} \caption{Various determinations of the electron density for the nebulae in our sample. The densities are given in $10^3$~cm$^{-3}$. The abbreviations for the methods have the following meaning: radio~-- density determined from the radio flux, model~-- average model prediction (see text).} \label{ne:tab} \begin{tabular}{r@{\x{5}}r@{\x{5}}r@{\x{5}}r@{\x{5}}r@{\x{5}}r@{\x{2}}r} \hline H~1$-$40\hspace*{-2.5mm} & M~1$-$20\hspace*{-2.5mm} & M~2$-$4\hspace*{-2.5mm} & M~2$-$23\hspace*{-2.5mm} & M~3$-$15\hspace*{-2.5mm} & ref. & meth.\\ \\ & & & 13.6 & & 1 & \fb{S}{ii} \\ 10.9 & 4.7 & & & 24.3 & 2 & \fb{S}{ii} \\ & & & 3.0 & 2.5 & 3 & \fb{S}{ii} \\ & 17.8 & 4.5 & & & 5 & \fb{S}{ii} \\ & & 7.0 & & & 6 & \fb{S}{ii} \\ 15.0 & & & & & 7 & \fb{S}{ii} \\ & & 4.2 & & & 9 & \fb{S}{ii} \\ & 85.\phantom{0}& & & & 10 & \fb{S}{ii} \\ 4.4 & 9.2 & 5.6 & 11.5 & 10.6 & 11 & \fb{S}{ii} \\ & & & & 4.2 & 13 & \fb{S}{ii} \\ & & 3.6 & & & 14 & \fb{S}{ii} \\ 35.1 & & & & & 16 & \fb{S}{ii} \\ & & 5.7 & & & 9 & \fb{Cl}{iii} \\ & 7.8 & & & & 10 & \fb{Cl}{iii} \\ & & & 79.\phantom{0}& & 4 & \fb{Ar}{iv} \\ & 1.0 & & & & 10 & \fb{Ar}{iv} \\ & & & 63.\phantom{0}& & 15 & \fb{Ar}{iv} \\ 13.5 & & & & & 7 & radio \\ & & & 20.\phantom{0}& & 8 & radio \\ & 10.\phantom{0}& & & & 12 & radio \\ 22.7 & 14.6 & 9.1 & 79.3 & 3.7 & 17 & model \\ \hline \end{tabular} \nt{References --- 1. Acker et al.\ \cite{c2:ackko} 2. Acker et al. \cite{c2:ackko2} 3. Aller \& Keyes \cite{c2:aller:keyes} 4. Boffi \& Stanghellini \cite{c2:boffi} 5. Costa~et al.\ \cite{c2:costa} 6. Cuisinier et al. \cite{c2:cuisinier} 7. Dopita et al.\ \cite{c2:dopita} 8. Kaler \cite{c2:kaler:i} 9. Kaler et al. \cite{c2:kalerea} 10. Kaler et al. \cite{c2:kalerea2} 11. Ratag et al.\ \cite{c2:ratag} 12. Shaw \& Kaler \cite{c2:shaw2} 13. Stanghellini \& Kaler \cite{c2:stang} 14. Tylenda et al. \cite{c2:tylenda2} 15. Webster \cite{c2:web76} 16. Webster \cite{c2:webster} 17. This work} \end{table} Our values differ substantially from the values given by RPDM, although they are based on the same observational data. This is because we use a completely different method to determine the density. For three out of five nebulae we find results which are within the range of values found with other methods. For M~2$-$4 we find a value which is a bit larger. The results in Paper~I indicate that our determination of the density is somewhat susceptible to measurement errors and errors in the model assumptions. This might provide an explanation for the discrepancy. The fact that we model only the core region of M~2$-$23 provides an explanation for the very high density we find for this nebula. Webster \cite{c2:web76} and Boffi \& Stanghellini (1994, using the same spectrum) also find a high value using the \fb{Ar}{iv} line ratio. The \fb{Ar}{iv} lines are expected to be formed predominantly in the core region and hence this would confirm our results. On the other hand, the excitation in the core is too high to form large amounts of S$^+$. Hence the \fb{S}{ii} lines can be expected to originate predominantly from the halo and should therefore indicate lower densities. This of course also depends on the exact position of the slit over the nebula. All of this might be an explanation for the extremely large spread of values found for this nebula. The quality of the data in Table~\ref{ne:tab} makes a comparison with our results meaningless. However, the data are at least consistent with the assumption that our results are more accurate than the results from line diagnostics. \subsection{Nebular abundances} \begin{table*} \small \caption{Comparison of the abundance determinations of the PNe in our sample.} \label{abun:lit} \begin{tabular}{lrrrrrrrrrrrrrr} \hline & \ctr{4}{H~1$-$40}{0} & & \ctr{4}{M~1$-$20}{0} & & \ctr{4}{M~2$-$4}{0} \\ ref:\x{5} &\sctr{4}&\sctr{9}&\sctr{12}&\sctr{13} & &\sctr{2}&\sctr{7}&\sctr{9}&\sctr{13} & &\sctr{2}&\sctr{3}&\sctr{9}&\sctr{13}\\ \noalign{\vskip2pt} \\ \noalign{\vskip2pt} $\epsilon$(He) & 11.03 & 11.06 & 11.04 & 10.96 & & 11.07 & 10.94 & 11.02 & 11.02 & & 11.11 &10.96\rlap{:}& 10.99 & 10.96 \\ $\epsilon$(N) & 7.72 & 8.08 & & 7.78 & & 7.39 & 7.80 & 7.75 & 7.81 & & 7.65 & 8.17 & 8.09 & 8.13 \\ $\epsilon$(O) & 8.52 & 8.70 & 8.53 & 8.23 & & 8.30 & 8.65 & 8.62 & 8.72 & & 8.30 & 8.77 & 8.80 & 8.84 \\ $\epsilon$(Ne) & 7.89 & 7.69 & & 7.39\rlap{:}& & & & 7.79 & 7.82\rlap{:}& & & & 7.90 & 8.15\rlap{:}\\ $\epsilon$(S) & 6.88 & 6.77 & & 6.37 & & 6.43\rlap{:}& & 6.52 & 6.66 & & 6.64 & 7.03 & 6.96 & 6.90 \\ $\epsilon$(Ar) & 6.6\phantom{0}& 6.43 & & 5.98 & & & & 6.05 & 5.99 & & 6.31 & 6.36 & 6.25 & 6.36 \\ \\ & \ctr{8}{M~2$-$23}{0} & & \ctr{4}{M~3$-$15}{0} \\ ref:\x{5} &\sctr{1}&\sctr{5}&\sctr{6}&\sctr{8}&\sctr{9}&\sctr{10}&\sctr{11}&\sctr{13} & &\sctr{1}&\sctr{5}&\sctr{9}&\sctr{13}\\ \noalign{\vskip2pt} \\ \noalign{\vskip2pt} $\epsilon$(He) & 11.00 & 10.93 & &10.88\rlap{:}& 10.98 & 10.92 & 10.96 & 11.05 & & 11.03 & 11.01 & 11.03 & 11.03 \\ $\epsilon$(N) & 7.68 & 8.20 & 7.70 & 7.55 & 7.40 & 8.13 & & 7.67 & & 8.08 & & 8.14 & 7.60 \\ $\epsilon$(O) & 8.40 & 8.42 & 8.18 & 8.34 & 8.22 & 8.47 & 8.11 & 8.67 & & 8.41 & 8.51 & 8.74 & 8.22 \\ $\epsilon$(Ne) & 7.60 & 7.62 & & 6.46\rlap{:}& 7.15 & 7.65 & & 7.75\rlap{:}& & 7.48 & 7.62 & 7.86 & 7.53\rlap{:}\\ $\epsilon$(S) & 6.6\phantom{0}& & & 6.29 & 6.30 & & & 6.79 & &6.7\rlap{:}\phantom{0}& & 6.86 & 6.31 \\ $\epsilon$(Ar) & 5.75 & & & 5.86 & 5.81 & & & 6.08 & & 6.5\phantom{0}& & 6.53 & 6.20\rlap{:}\\ \hline \end{tabular} \ntd{\rlap{:}\x{4}A colon indicates that the value is uncertain.} \ntd{References --- 1. Aller \& Keyes \cite{c2:aller:keyes} 2. Costa et al. \cite{c2:costa} 3. Cuisinier et al. \cite{c2:cuisinier} 4. Dopita et al.\ \cite{c2:dopita} 5. Henry \cite{c2:henry} 6. Kaler \cite{c2:kaler:ii} 7. Kaler et al. \cite{c2:kalerea2} 8. K\"oppen (private communication) 9. Ratag et al. \cite{c2:ratag} 10. Walton, Barlow \& Clegg \cite{c2:wal:bar} 11. Webster \cite{c2:web76} 12. Webster \cite{c2:webster} 13. This work} \normalsize \end{table*} In Table~\ref{abun:lit} we give a comparison of the abundances we determined with other literature values. We did not include the nitrogen abundance for M~3$-$15 from Henry \cite{c2:henry}. After a discussion with Dr. Henry it was established that this abundance was flawed by an error in the analysis (as is also the case for the nitrogen abundances of M~4$-$3 and H~1$-$23 listed in the same paper; all other results are not affected). We also did not include the abundances for M~2$-$23 listed in K\"oppen, Acker \& Stenholm \cite{c2:koppen}. It was established that this analysis was flawed by an error as well, and Dr. K\"oppen kindly provided us with a re-analysis of his data. The higher nitrogen abundance given by Walton et al.\ \cite{c2:wal:bar} for M~2$-$23 might be a result of the inclusion of {\it IUE}\ data in their analysis. They systematically find higher nitrogen abundances for bulge PNe than other authors. One can see that large differences can be found between the various abundance determinations in the literature. If we exclude our own results, we find the following statistics. For elements heavier than helium, we find a difference between the lowest and highest abundance determination larger than or equal to 0.3~dex in 12 out of 22 cases, and larger than or equal to 0.5~dex in 4 out of 22 cases. For the helium abundances we find a spread larger than 0.1~dex in 2 out of 5 cases. Especially the abundances for M~2$-$23 show a large spread and should be considered uncertain. From this we draw the conclusion that, at least for the sample studied here, abundance determinations can not be considered very accurate. Uncertainties exceeding 0.2~dex to 0.3~dex are not uncommon. When one compares the abundances for the individual PNe with the values from RPDM, one can see that for the two objects where the electron temperature is in good agreement (M~1$-$20 and M~2$-$4), the abundances also agree very well. For the other objects the abundance determinations differ. We attribute this to the difference in the determination of the electron temperature. When we compare our abundance determinations with the other values found in the literature, we see that our results often are slightly outside the range of values found by other authors. This behaviour is well correlated: either all outliers are at the low end or at the high end. This behaviour is also well correlated with our electron temperature determination. When our electron temperature determination is at the low end, our abundances are at the high end, and the reverse is also the case (see also Section~\ref{te:litt}). This indicates that the main source of uncertainty in the abundance determination is the electron temperature. Hence the discussion given in Section~\ref{te:litt} applies here as well. \subsection{Stellar broadband fluxes} Since {\sc cloudy}\ calculates the attenuation of the stellar continuum separately from the transport of the diffuse nebular continuum, we are able to predict broadband photometric fluxes for the central star as they would appear through the nebula. In this way we could calculate a prediction for the Johnson~{\rm\sl B}\ and {\rm\sl V}\ magnitudes. However, observed stellar magnitudes will be reddened due to interstellar extinction as well, and we have to take this into account in our predictions. To calculate the total extinction towards the nebula, we averaged all the measurements we could find in the literature. Since the continuum fluxes predicted by {\sc cloudy}\ already take the internal extinction into account, we only have to correct the stellar magnitudes for the external extinction. Hence we used the internal extinction from our model, and subtracted it from the total extinction. Then we used this value for the external extinction to predict the reddened {\rm\sl B}\ and {\rm\sl V}\ magnitudes of the central star. Where necessary, we applied the interstellar reddening law given by Pottasch \cite{c2:book:pot}. A comparison of the calculated values with the literature values taken from Acker et al.\ \cite{c2:ack:cat} is given in Table~\ref{mag:tab}. \begin{table*} \small \caption{For the PNe in our sample we give in column 2 and 3 the Johnson~{\rm\sl B}\ and {\rm\sl V}\ magnitudes resp.\ predicted by our model, in column 4 the internal extinction in the Johnson~{\rm\sl V}\ band derived from our model, in column 5 the average total extinction derived from the Balmer decrement and the radio flux, in columns 6 and 7 the predicted reddened values for the Johnson~{\rm\sl B}\ and {\rm\sl V}\ magnitudes and in column 8 and 9 the measured magnitudes given in Acker et al.\ \protect\cite{c2:ack:cat}.} \label{mag:tab} \begin{tabular}{lcccccccc} \hline name & {\rm\sl B}$_{\rm mod}$ & {\rm\sl V}$_{\rm mod}$ & $A_{\rm\sl V}^{\rm int}$ & $A_{\rm\sl V}^{\rm tot}$ & {\rm\sl B}$_{\rm pred}$ & {\rm\sl V}$_{\rm pred}$ & {\rm\sl B} & {\rm\sl V} \\ & {\rm mag} & {\rm mag} & {\rm mag} & {\rm mag} & {\rm mag} & {\rm mag} & {\rm mag} & {\rm mag} \\ \\ H~1$-$40 & 15.70 & 15.68 & 1.11 & 5.05$\pm$0.26 & 20.86$\pm$0.34 & 19.62$\pm$0.26 & & \\ M~1$-$20 & 14.62 & 14.90 & 0.06 & 2.61$\pm$0.10 & 17.96$\pm$0.13 & 17.45$\pm$0.10 & 17.7 & 17.1 \\ M~2$-$4 & 14.41 & 14.64 & 0.14 & 2.78$\pm$0.15 & 17.87$\pm$0.19 & 17.28$\pm$0.15 & 17.6 & 17.0 \\ M~2$-$23 & 14.63 & 14.90 & 0.08 & 1.81$\pm$0.22 & 16.90$\pm$0.28 & 16.63$\pm$0.22 & 16.7 & \\ M~3$-$15 & 15.52 & 15.80 & 0.10 & 4.69$\pm$0.25 & 21.53$\pm$0.33 & 20.39$\pm$0.25 & & \\ \hline \end{tabular} \normalsize \end{table*} The predicted magnitudes are slightly fainter than observed, but still in remarkable good agreement, considering the fact that we use a blackbody approximation to determine these values. Given the fact that a blackbody of a given temperature has more ionizing photons than a realistic spectrum with the same effective temperature, one can expect that in the best-fit model the total luminosity will be underestimated to compensate for this effect. However, we find that this effect is only very modest and this can be understood from the fact that we include the dust emission in the modelling. Grains can be heated very efficiently by Balmer continuum photons, as well as by Lyman continuum photons. Therefore, the {\it IRAS}\ fluxes give a good constraint on the Balmer continuum flux. This counteracts the previously mentioned underestimation of the total luminosity and explains the remarkable accuracy of our stellar broadband fluxes. \subsection{Distances} \label{dist:discussion} In our model assumptions we assume the distance to be a fixed number. However, our method can easily be changed in such a way that the distance would be a free parameter. When this is done, the best-fit model would also give an estimate for the distance. We have investigated the possibility to determine the distance this way (van Hoof \& Van de Steene 1996). We found that, though possible in principle, the spread in the resulting distance determinations is large. The distance determination is vulnerable to various observational errors, but especially to the error in the determination of the angular diameter. Since the angular diameter is notoriously hard to measure, this sensitivity makes the results very uncertain. When we determined the distances to the bulge nebulae in our sample with this method, we found the spread in the values to be larger than what is obtained from a statistical method (Van de Steene \& Zijlstra 1995). Closer investigation reveals that this method of determining distances is in essence identical to the method described in Phillips \& Pottasch \cite{c2:phillips}. They already concluded that this method is unreliable. The use of a wrong value for the distance not only influences the distance dependent parameters but also some distance independent parameters, as was already discussed in Paper~I. We therefore do not recommend this method, and advise the use of separately determined distances. \section{Conclusions} \label{conclusions} We applied our method which enables a fully self-consistent determination of the physical parameters of a PN, using the spectrum, the {\it IRAS}\ and radio fluxes and the angular diameter of the nebula, to a sample of five galactic bulge PNe. Comparison of the distance independent physical parameters with published data shows that the stellar temperatures generally are in good agreement and can be considered reliable. The literature data for the electron temperature, electron density and also for the abundances show a large spread, indicating that the use of line diagnostics is not reliable. Comparison of the various abundance determinations indicates that the uncertainty in the electron temperature is the main source of uncertainty in the abundance determination. The large spread in the literature data makes a comparison with our results meaningless. The stellar magnitudes predicted by the photo-ionization models are in good agreement with observed values. \section*{Acknowledgments} We thank Drs.\ S.R.\ Pottasch, L.B.F.M.\ Waters and J.M.\ van der Hulst for helpful comments. Drs.\ R.D.\ Oudmaijer, T. de Jong, G.J. Ferland and the anonymous referee are thanked for critically reading the manuscript. Drs. R.B.C. Henry and J. K\"oppen are thanked for re-examining their data. The authors acknowledge support from the Netherlands Foundation for Research in Astronomy (ASTRON) through grant no.\ 782--372--033 and 782--372--035. PvH wishes to thank ESO for their hospitality and financial support during his stay in Santiago where part of this paper was written. PvH is currently supported by the NSF through grant no.\ AST 96--17083. The photo-ionization code {\sc cloudy}\ was used, written by Gary Ferland and obtained from the University of Kentucky, USA. We thank Gary Ferland for his invaluable help in debugging and modifying the code.
\section{Introduction} One of the frontiers of current astronomical research is the observation of supernovae at high-redshift. In fact it is expected that by using SNe~Ia as distance indicators it will be possible to constrain the geometry of the Universe within a few years. Equally important and difficult, is to determine the rate of the different types of SNe as a function of redshift. The rationale is that the various types of SNe have progenitors of different ages; in particular core-collapse SN~II+Ib/c result from young, massive stars and SN~Ia originate from intermediate to old population stars (eg. Branch et al. \cite{bnf}). Therefore, the evolution of the relative SN rates with redshift can be used to probe the average SFR rate history in galaxies and, in turn, constrain scenarios for galaxy formation and evolution. So far, there have been only exploratory attempts in this direction (J{\o}rgensen et al. \cite{jorg}; Sadat et al. \cite{sadat}; Madau et al. \cite{madau}), which however have demonstrated the potential of this approach and motivated new observational efforts. The accurate determination of the present time SN rates is the benchmark, crucial to exploiting these efforts to the full. Another important factor is to compare the rates of various SN types with different indicators of the stellar population content of galaxies in the local Universe. One problem with these rates is that local SNe are rare and therefore it requires several years or decades to collect sufficient statistics. In addition, in order to obtain accurate estimates of the SN rate, it is necessary to know: $i)$ the sample of galaxies which have been searched for SNe, $ii)$ the frequency and limiting magnitude of observations and $iii)$ the instruments/techniques which are used for detection in order to assess search biases. Very few groups of professional astronomers have had the perseverance and force to carry out a SN search program long enough to be really useful for this purpose ({\em cf} Cappellaro et al. \cite{stat95} hereafter C97). Among the amateurs in this field, an outstanding case is the visual SN search which has been conducted by Evans since 1980 (Evans \cite{ev:97}). Indeed, estimates of the SN rate based on the first 10 years of Evans' SN search have already been published (Evans et al. \cite{ev:89}; van den Bergh \& Mc~Clure \cite{vdbmc}). In this paper we will analyze the updated log of this survey which doubles the statistics with respect to previously published estimates (Sec.~\ref{evans}). Following the protocol described in a previous paper (C97), we pooled together Evans' log and those of photographic searches and used the improved statistical basis to test a different approach for the correction of selection effects (Sec.~\ref{sec_hatano}). We used this combined database, which is the largest ever built for such a purpose, to obtain updated estimates of the SN rates (Sec.~\ref{update}). Finally, we compared the SN rates with other SFR indicators such as the far infrared luminosity, the integrated color and the activity of the galaxy. \section{Evans' visual SN search}\label{evans} Evans began his search of SNe in 1980 using a 25~cm telescope. The observations were conducted visually, which has the advantage of being very fast and inexpensive, but the limiting magnitude for SN discovery is not very deep ($m_{\rm lim}=14.5$ mag). At the end of 1985 the telescope was replaced with a 41~cm telescope ($m_{\rm lim}= 15.0$ mag), and most recently complemented by a 100~cm telescope at Siding Spring Observatory ($m_{\rm lim}=16.0$). \begin{table*} \caption{SN rates in spirals of different inclination (not corrected for the inclination bias.)}\label{inc} \begin{tabular}{cccccccc} \hline & \multicolumn{3}{c}{Evans search} && \multicolumn{3}{c}{photographic searches (from C97)} \\ \cline{2-4}\cline{6-8} inc & N. & N. & rate && N. & N. & rate \\ ${\rm [deg]}$& galaxies & SNe & [SNu]$^*$ &&galaxies &SNe&[SNu]$^*$\\ \hline 0-45 & 616 & 18 & $0.88\pm0.21$ && 1581 & 36 & $0.94\pm0.16$ \\ 45-65 & 666 & 13 & $0.49\pm0.14$ && 1702 & 25 & $0.59\pm0.12$ \\ 65-90 & 592 & 11 & $0.34\pm0.10$ && 1818 & 15 & $0.32\pm0.08$\\ \hline \end{tabular} $*$ $1 \,{\rm SNu} = 1\, {\rm SN}\, (100 {\rm yr})^{-1}\, (10^{10} L_\odot^{\rm B})^{-1}$. \\ \end{table*} During the almost two decades of the search, Evans collected over 200,000 individual observations, surveying a sample of more than 3000 nearby galaxies. The search resulted in the discovery of 32 SNe, and another 22 SNe, which first had been discovered by others were also detected, for a total sample of 54 SNe. These numbers qualify Evans' search as the most successful amateur SN search ever and make it very competitive even to professional searches. For the calculation of SN rates based on Evans' search log, we used the control time method and the protocol described in C97. The essential galaxy data, that is distances, morphological types, luminosities and axial ratios, have been retrieved from the updated version of the RC3 catalog (de Vaucouleurs et al. \cite{rc3}) distributed by the ``Centre de Donn{\'e}es Astronomiques de Strasbourg''. With the improved SN statistics, the Poissonian errors on SN events were reduced and other sources of errors became dominant, in particular in the correction of search biases. All SN searches suffer from specific biases and therefore it is important to compare the results of different kinds of searches. Because it is unique, Evans' visual search is especially useful for comparisons with traditional photographic surveys (C97). Specific biases in SN searches occur because SNe appear embedded in their parent galaxies. Since the detection efficiency of new objects depends on the contrast against the background, obviously it is more difficult to discover a SN in the (luminous) inner regions of a galaxy than in its outskirts. This bias is at its most pronounced if one uses a wide field/small scale telescope and a detector with a small dynamic range. Also the fraction of SNe lost is greater in more distant galaxies because of the small angular size of the galaxy image. By comparing the radial distributions of SNe in galaxies at different distances, we estimated that up to 50\% of the SNe exploding in the more distant galaxies are lost in photographic searches using Schmidt telescopes. The nuclear bias seems negligible only in CCD and visual searches of nearby galaxies (C97). A precursory examination of the general list of SNe reveals another severe bias against SN detection: that in inclined spirals. In the past, it has been claimed that this bias does not affect visual and CCD searches (Evans et al. \cite{ev:89}; Muller et al. \cite{muller}) but so far the evidence has been inconclusive (C97). It is thus of interest to exploit the improved statistics of Evans' visual search to address this question. We began by calculating the overall SN rate in spiral galaxies of different inclination both for Evans' visual search and for the combined photographic search sample constructed by C97, including the SN searches of Asiago (Cappellaro et al. \cite{PI}), Crimea (Tsvetkov \cite{tsv:83}), OCA (Pollas \cite{pol}) and Cal\'an/Tololo (Hamuy et al. \cite{mario:93}). For this particular calculation we turned off the correction for parent galaxy inclination which was included in the recipe of C97. The results are reported in Table~\ref{inc}, where the bins were chosen to give roughly the same number of galaxies in each inclination bin. It would appear that based only on Evans' log, the SN rate in edge-on spirals is 2.6 fold less than in face-on ones. This clearly demonstrates that the bias in inclined spirals also affects visual searches and is almost as severe as in photographic searches (2.9). The natural interpretation of this bias is that SNe occurring in the disk of inclined spirals appear on the average dimmer than those in face-on spirals because of the increased optical depth through the dust layer. As a consequence, the probability of SN discovery in inclined spirals is reduced. In line with this interpretation, it is expected that the bias is more severe in searches carried out in the blue band, (e.g. photographic searches), than in visual ones (CCD searches in the red would be even less affected). As a first order approach to correct for the inclination bias, we can assume a plane parallel geometry for the distribution of dust in the disk of spirals. In this case the average extinction of the SN population scales with $\sec i$, where $i$ is the inclination of the galaxy disk with respect to the line of sight. We showed in C97 that this assumption results in an over-correction of the SN rate in edge-on galaxies. It was argued that this is evidence that dust is not uniformly distributed in the disk of spirals but is instead in discrete clouds. Until more evidence is available we have adopted, in analogy to C97, a conservative extinction law that is intermediate between the $\sec i$ and an empirical relation, derived from the assumption that the SN rate is the same in face-on and edge-on galaxies (see Section ~\ref{sec_hatano} for an alternative model). When such a correction is included, we can compute the SN rates for the complete Evans search and compare it with those obtained from the first ten years of the search and with those derived from the combined photographic search sample. The results are reported in Table~\ref{evares} where the number of galaxies and SNe for each sample is indicated in the first two rows. We must stress that, since we adopt the same galaxy catalog, input parameters, bias corrections and numerical recipe, the differences between the columns in Table~\ref{evares} are only due to the different logs of observations. The new Evans rates are consistent with the earlier results but, at least with respect to the average value, in much better agreement with the rate from photographic searches. Looking more in detail, it appears that for early type galaxies (E-S0 and S0a-Sb) the updated Evans value is in better agreement with the photographic search rate, whereas for late spirals (Sbc-Sd) the old value was closer. We attribute these fluctuations to the small statistics of individual SN searches, which become wider when the sample is divided into bins. \begin{table} \caption{Comparison of the SN rate [SNu] obtained from the Evans' updated statistics (1980-1998), the first 10 years of the search (1980-1989) and the combined photographic search sample} \label{evares} \begin{tabular}{lccc} \hline & Evans 80-98 & Evans 80-89 & ph. search \\ N. galaxies & 3068 & 1377 & 7319 \\ N. SNe & 54 & 24 & 94 \\ \hline \\ E-S0 & $0.17\pm0.06$ & $0.13\pm0.08$ & $0.18\pm0.05$ \\ S0a-Sb & $0.83\pm0.20$ & $1.22\pm0.41$ & $0.66\pm0.13$ \\ Sbc-Sd & $0.96\pm0.13$ & $1.20\pm0.38$ & $1.34\pm0.20$ \\ \\ All$^*$& $0.66\pm0.09$ & $0.81\pm0.17$ & $0.67\pm0.07$ \\ \hline \end{tabular} $*$ Including Sm, irregulars and peculiars. \end{table} \section{An alternative model for bias corrections}\label{sec_hatano} As mentioned before, corrections for search biases are the most controversial step in the calculation of SN rates. In our approach (cf. C97), the correction factors are tuned to cancel the sign of the biases from the calculated SN rates regardless of their physical causes. The ideal would be to build a model for the dust and SN distribution which is consistent with the present data on galaxies and SN progenitor populations and derive from it estimates of the biases. In a recent paper, Hatano et al. (\cite{hatano}) described a simple model based on an assumed distribution of the dust and SN populations which predicts that, because of extinction, SNe in inclined spirals appear on average dimmer and shoe a much wider magnitude scatter than those in face-on spirals. Moreover, because the dust distribution peaks in the central regions of galaxies, this effect is more pronounced for SNe occurring in those regions. This provides an alternative to the classical explanation of the selection effect in the central region of galaxies which would thus derive from the enhanced extinction of SNe instead of the reduced luminosity contrast. According to this scenario, the bias would be most severe in dust, inclined spirals and, because of the small scale height, for core collapse SNe. Indeed, all these features were found in the observed SN sample (Table~1 of Cappellaro \& Turatto \cite{ct:97}). A special characteristic of the Hatano et al. model is that core collapse SNe do not occur within 3 Kpc of the center of the galaxy. Therefore SN~II or Ib/c do not appear in the central regions of face-on galaxies although an increasing number of core collapse SNe appears to be projected on the centers of the more inclined spirals due to projection effects. In any case, type Ia SNe are more highly concentrated than type II or Ib/c. Though Hatano et al. claim that the observations confirm their model, we should mention that van den Bergh (\cite{vdb97}) and Wang et al. (\cite{wang}) reached the opposite conclusion on the basis of similar data. Even if the Hatano et al. model should be considered as exploratory given the above controversy, it is of interest to test how adopting it can change the SN rate estimates. We thus have replaced the empirical bias corrections mentioned in Section~\ref{evans} with the observed SN luminosity distribution for each SN type in spirals of different inclination derived from the Hatano et al. (1998) model. Then we computed the control times for each bin of the luminosity function for each galaxy and SN type. The total control time was obtained as the weighted average according to the observed luminosity distribution . The results of this calculation are shown in Table~\ref{hattab}. Taken at face value and compared with the empirical bias corrections (Tab.~\ref{final}), we found that by using the Hatano et al. model, the SN~Ia rate in the entire sample of spirals results 10-20\% higher and the SN~II+Ib/c rate 15\% smaller. These differences all fall within the errors and should not be considered significant. \begin{table} \caption{The SN rate corrected using the Hatano et al. (1998) model}\label{hattab} \begin{tabular}{lccc} \hline galaxy &\multicolumn{3}{c}{rate [SNu]} \\ type & Ia & II+Ib/c & All\\ \hline S0a-Sb & $0.27\pm0.08$ & $0.63\pm0.24$ & $0.91\pm0.26$\\ Sbc-Sd & $0.24\pm0.10$ & $0.86\pm0.31$ & $1.10\pm0.32$\\ \\ Spirals$^*$ & $0.25\pm0.09$ & $0.76\pm0.27$ & $1.01\pm0.29$\\ \hline \end{tabular} $*$ Includes types from Sm, irregulars and peculiars. \end{table} We notice, however, that by adopting the Hatano et al. model the SN rate in edge-on spirals remains 1.5 times smaller than in face-on spirals, and that the rate in intermediate inclination spirals is even smaller. Slightly increasing the optical depth of the dust layer helps but does not solve the problem. A more perplexing feature of the bias corrections based on the Hatano et al. model is that it produces SN rates which increase with galaxy distances: the rate in galaxies with $v>3000$ km s$^{-1}$ results almost twice that in galaxies with $v<3000$ km s$^{-1}$. This could be resolved by reducing the average reddening but which is the opposite of the previous recommendation. This apparent contradiction may simply indicate that, as already suggested, the SN progenitors and dust distributions in real galaxies are more complex than in this simple exploratory model. In conclusion, though the approach seems promising, the Hatano model needs further refinement and in the meanwhile we decided to maintain the {\em empirical} bias corrections of C97. \section{SN rates and indicators of the galactic SFR }\label{update} Once we had verified that the SN rates derived from Evans' visual search are similar to those obtained from photographic searches and decided the correction for search biases, we merged all search logs in a single database. In this way we obtained an improvement in the statistics compared to C97 (from 110 to 137 SNe) and equally important we balanced the weights of different types of searches (over one third of the SNe in the new sample were discovered in Evans' visual search). The SN rates computed using these updated statistics are reported in Table~\ref{final} \footnote{Through this paper we assumed H$_0$=75 km s$^{-1}$ Mpc$^{-1}$. SN rates reported in this paper can be transformed to other values of the Hubble constant multiplying by (H$_0/75)^2$} where errors include not only event statistics but also uncertainties in the input parameters and in the bias corrections. The differences with C97 are small ($<15$\%) and well within the errors. We notice that in C97 the SN~Ia rate appeared to increase when progressing from early to late type galaxies whereas this effect had now nearly vanished. The relatively low rates of SN~Ib/c compared with SNII were, instead, confirmed. \begin{table*} \caption{SN rate( in SNu) from the combined search sample.}\label{final} \begin{tabular}{lrrrccccc} \hline galaxy & \multicolumn{3}{c}{N. SNe$^*$} &&\multicolumn{3}{c}{rate [SNu]} \\ \cline{2-4}\cline{6-9} type & Ia & Ib/c & II && Ia & Ib/c & II & All\\ \hline E-S0 & 22.0 & & && $0.18\pm0.06$ & $<0.01$ & $<0.02$ & $0.18\pm0.06$\\ S0a-Sb & 18.5 & 5.5 & 16.0 && $0.18\pm0.07$ & $0.11\pm0.06$ & $0.42\pm0.19$ & $0.72\pm0.21$\\ Sbc-Sd & 22.4 & 7.1 & 31.5 && $0.21\pm0.08$ & $0.14\pm0.07$ & $0.86\pm0.35$ & $1.21\pm0.37$\\ Others$^\#$ & 6.8 & 2.2 & 5.0 && $0.40\pm0.16$ & $0.22\pm0.16$ & $0.65\pm0.39$ & $1.26\pm0.45$\\ \\ All & 69.6 & 14.9& 52.5&& $0.20\pm0.06$ & $0.08\pm0.04$ & $0.40\pm0.19$ & $0.68\pm0.20$\\ \hline \end{tabular} $*$ Similar to C97, 10 unclassified SNe have been redistributed among the three basic SN types according to the observed distribution that is 100\% Ia in E-S0, in spirals: type Ia 35\%, type Ib 15\%, type II 50\%. $\#$ Others includes types Sm, Irregulars and Peculiars \end{table*} The normalization of the SN rate to the galaxy blue luminosity has been introduced after the demonstration that the former scales with the latter. This is convenient because $i)$ integrated $B$ magnitudes are available for a large number of galaxies and $ii)$ the B luminosity for a given galaxy type scales with the total mass at a first approximation. Physically, the blue luminosity is a good tracer of the young stellar population in starburst galaxies, but not in normal galaxies where a considerable fraction of the continuum luminosities is produced by old stars also in the blue (Sage \& Solomon \cite{sage}; Kennicutt \cite{kenni}). In principle, by using different photometric bands should be possible to sample selected stellar populations and hence to obtain useful information for progenitor scenarios. For instance, van den Bergh (\cite{vdb:90}) and Della Valle \& Livio (\cite{mdv:94}) normalized the rate of SN~Ia to H and K luminosity: in these bands the role of old stars in all galaxy types is dominant. If all SN~Ia result from low mass stars we would expect the SN~Ia rate per unit of H and K luminosities not to be correlated to galaxy type. The fact that the rate in these units increases considerably when moving from ellipticals to late spirals was taken to indicate that a significant fraction of SN~Ia result from intermediate age stars. Even if their conclusion is probably correct, it must be stressed that these estimates were not direct measurements but a simple scaling of the SN rates in unit blue luminosity based on the assumption of an average B-H and B-K color per galaxy type. Because H and K photometry is available only for a small fraction of the galaxies in our sample, unfortunately, the SN rate in these units cannot be directly measured. It would also be of interest to estimate the rates of SNe, in particular of core collapse SNe, in galaxies with different star formation rates (SFR). The different diagnostic methods which are used to probe SFR in galaxies have been reviewed in a recent paper by Kennicutt (\cite{kenni}). Because we are limited by the SN statistics we need tracers that are available for large samples of galaxies. In this respect, integrated colors and far infrared (FIR) luminosities are particularly appealing. \subsection{SN rates and galaxy integrated colors}\label{sec_color} Integrated broad band colors are very useful for statistical purposes, as they are reliable indicators of the galaxy stellar population with bluer galaxies expected to host stars that are younger and more massive than redder ones. Colors are most interesting because, by using evolutionary synthesis models, it is possible to estimate the SFR per unit mass or luminosity required to produce a given integrated color for a given stellar population. It is well known that along the Hubble sequence the galaxy color becomes bluer moving from early to late types and that this corresponds to a sequence in SFR which is virtually zero in ellipticals and maximum in late spirals. However, especially in spirals, there is a significant dispersion in the average color from galaxy to galaxy, indicating that SFR can vary significantly even for a given Hubble type. Conveniently, $(B-V)_T^0$ and $(U-B)_T^0$ colors, corrected for galactic and internal extinction are listed in the RC3 catalog for a fair percentage of the galaxies of our sample (24\% and 19\% respectively). From these we derived also $(U-V)_T^0$ colors which, allowing for the extended wavelength baseline, are more sensitive SFR indicators. For each bin of galaxy morphological types we divided the galaxies into subsamples, containing galaxies bluer and redder than the global average. We then computed separately the SN rates for each of these subsamples. The results are reported in Table~\ref{color}, where the galaxy types are in col~1, the average colors for the galaxies of the specific subset are in cols~2 and 5, the SN rates in SNu for SN~Ia and for core-collapse SNII+Ib/c in cols~3-4 and 6-7. \begin{table*} \caption{SN rates in SNu for galaxies with integrated colors bluer and redder than the average. }\label{color} \begin{tabular}{lccc|ccc} \hline galaxy & \multicolumn{3}{c|}{blue galaxies} &\multicolumn{3}{c}{red galaxies} \\ type & $<(B-V)_T^0>$ & Ia & II+Ib & $<(B-V)_T^0>$ & Ia & II+Ib \\ \hline E-S0 & 0.86 & $0.3\pm0.1$ & & 0.95 & $0.2\pm0.1$ \\ S0a-Sb & 0.60 & $0.2\pm0.1$ & $0.6\pm0.2$ & 0.80 & $0.2\pm0.1$ & $0.5\pm0.2$\\ Sbc-Sd & 0.45 & $0.1\pm0.1$ & $1.5\pm0.3$ & 0.62 & $0.3\pm0.1$ & $0.9\pm0.2$\\ & & & & & &\\ All$^*$ & 0.56 & $0.2\pm0.1$ & $1.0\pm0.2$ & 0.88 & $0.2\pm0.1$ & $0.1\pm0.1$\\ \hline \end{tabular} \begin{tabular}{lccc|ccc} \hline galaxy & \multicolumn{3}{c|}{blue galaxies} &\multicolumn{3}{c}{red galaxies} \\ type & $<(U-V)_T^0>$ & Ia & II+Ib & $<(U-V)_T^0>$ & Ia & II+Ib \\ \hline E-S0 & 1.26 & $0.3\pm0.1$ & & 1.48 & $0.2\pm0.1$ \\ S0a-Sb & 0.67 & $0.2\pm0.1$ & $0.9\pm0.3$ & 1.12 & $0.2\pm0.1$ & $0.4\pm0.2$\\ Sbc-Sd & 0.28 & $0.2\pm0.1$ & $1.7\pm0.5$ & 0.62 & $0.3\pm0.1$ & $0.8\pm0.3$\\ & & & & & &\\ All$^*$ & 0.54 & $0.2\pm0.1$ & $1.1\pm0.2$ & 1.32 & $0.2\pm0.1$ & $0.1\pm0.1$\\ \hline \end{tabular} $*$ Including Sm, irregulars and peculiars. \end{table*} As expected, the rate of core collapse SNe (II+Ib/c) is higher in the bluer spirals. By using B-V color this effect is seen only for late spirals (the rate is higher by a factor of 1.7 for Sbc-Sd), but becomes clear for all spirals when using U-V color (over a factor of 2). Instead the rate of SN~Ia is, within the uncertainties, independent on galaxy colors. With regards to the rows labeled ``All'' (which includes galaxies of all types) we should note that dividing galaxies into bluer and redder colors to a large extent corresponds to separating them into early and late type galaxies. Therefore the great difference in the core collapse SN rates in bluer and redder galaxies simply reflects the fact that core collapse SNe are not found in early type galaxies. We can compare the observed SN rates with the predicted SFR in galaxies of different colors. This is done in Fig.\ref{sfr} where the dots represent the SN rates in SNu (left-hand scale) in galaxies of different $U-V$ integrated colors and the line is the SFR per unit of blue luminosity (right-hand scale) taken from the evolutionary synthesis models of Kennicutt (\cite{kenni}). In general, for a galaxy of luminosity L$_B$, because of the short life of progenitor evolution, the number of core collapse SNe per century corresponds to the number of new born stars within the appropriate mass range, namely: $$ SN\,rate [SNu] \times L_B \simeq {SFR \times f_{M_L}^{M_U} \over <M_{SN}>} \times 100 $$ where $f_{M_L}^{M_U}$ is the mass fraction of stars which are born with mass in the range $M_L$ to $M_U$, the lower and upper limit for core-collapse SN progenitors, and $<M_{SN}>$ is the average mass of SN progenitors. According to the standard scenarios, $M_L \simeq 8 M_\odot$ and $M_U \simeq 40 M_\odot$. Adopting a Salpeter mass function, $f_{M_L}^{M_U} \simeq 10^{-1}$ and $<M_{SN}> \simeq 10 M_\odot$ which compensate the factor 100 which accounts for the difference in the time scale. In conclusion, even if the exact coincidence of the two scales in our figure is to some degree fortuitous, the nice agreement of the SFR measured through core collapse SN rates and that deduced by synthesis modeling for ``average'' spiral galaxies, lends support to the general scenario for stellar population evolution. Conversely, the fact that the rate of SN~Ia shows no dependence on the galaxy U-V color requires a significant delay between the SFR episodes and the onset of SN~Ia events. \begin{figure} \resizebox{\hsize}{!}{\includegraphics*{sfr.eps}} \caption{SN rates in SNu (left-hand scale) in spirals with different U-V color (for this plot early and late spirals have been considered together). Filled symbols are core-collapse SN~II+Ib/c and open circle are SN~Ia. Error-bars only accounts for the SN statistics. The line gives the SFR rate per unit B luminosity (right-hand scale) for galaxies with different colors as predicted by the evolutionary synthesis models of Kennicutt (1998). The surprising coincidence of the two scales can be partially understood by simple units conversion (see text).}\label{sfr} \end{figure} The relation between core collapse SN rates and colors provides a useful tool for the comparison of local and high-$z$ SN rates. Indeed, for galaxies at high-$z$ integrated colors can be measured relatively easily, whereas morphological types, requiring superb imaging, are not generally available. Conversely, it is clear that reporting the average SN rates for uncharacterized galaxy samples may turn out to be pointless for constraining galaxy evolution models. \subsection{SN rates and galaxy FIR luminosities}\label{sec_fir} The interest in deriving the SN rate in units of the FIR luminosities was stressed by J{\o}rgensen (\cite{jorg90}) who made a first attempt based on the general SN catalog. Here we report our calculations based on the control time technique. The near infrared emission of spiral galaxies shows at least two components: a warm component associated with dust around young stars and a cool component associated with more extended dusty heated by the general stellar radiation field, including radiation from old stars. The warm emission gives a direct measurement of the SFR, but in normal galaxies it is heavily contaminated by the cool component. In addition to the extended star formation in the disk, many spiral galaxies show an enhanced SFR in the nuclear region. The observations show that the nuclear and extended components are mostly decoupled. In these ``starburst'' galaxies the nuclear SFR could reach 1-1000 M$_\odot yr^{-1}$ and the integrated infrared emission is largely dominated by the nuclear component (Kennicutt \cite{kenni}). At the same time we expect a very high rate of core collapse SNe, which are however difficult to detect in the optical. This is because in the nuclear starburst regions one expects several magnitudes of extinction and a severe bias for optical SN searches. These considerations must be kept in mind when interpreting the results. FIR fluxes have been measured by the IRAS survey for over 30000 galaxies in the range $10-100 \mu{\rm m}$ and FIR magnitudes are reported in the RC3 catalog for $\sim 30$\% of the galaxies of our sample. They have been converted to units of solar FIR luminosites using the relation: $$ {L_{\rm FIR} \over L_{\rm FIR,\odot}} = 10^{-0.4 m_{\rm FIR}}\, 3.1 \times 10^{11}\, d^2 $$ where $d$ is the galaxy distance in Mpc. First we computed the SN rates for unit infrared luminosity $L_{\odot,{\rm FIR}}$. \begin{table} \caption{SN rates per unit FIR luminosity. $ 1\,{\rm SNuIR} = 1\, {\rm SN} (100 yr)^{-1} (10^{10} L_{{\rm FIR},\odot})^{-1}. $ } \label{lfir} \begin{tabular}{lcccc} \hline galaxy &\multicolumn{3}{c}{SN rate [SNuIR]} \\ type & Ia & II+Ib/c & All\\ \hline E-S0 & $1.8\pm0.8$ & & $1.8\pm0.8$\\ S0a-Sb & $0.6\pm0.2$ & $2.0\pm0.5$ & $2.7\pm0.5$\\ Sbc-Sd & $0.6\pm0.1$ & $3.5\pm0.6$ & $4.1\pm0.6$\\ \\ All$^*$ & $0.7\pm0.1$ & $2.5\pm0.3$ & $3.2\pm0.3$\\ \hline \end{tabular} $*$ Includes types Sm, Irregulars and Peculiars. \end{table} If the FIR luminosity is a direct measure of the SFR in spirals, as is often assumed, we would expect the rate of core collapse SNe per unit FIR luminosity to be constant through all galaxy types. Instead, the results reported in Table~\ref{lfir} show that the rate of core collapse SNe in FIR units increases almost 2 fold moving from early to late spirals, whereas the rate of SN~Ia remains constant (in Table~\ref{lfir} we report the SN rate in SNuIR also for E-S0 galaxies, though we do not expect there to be any relation between L$_{\rm FIR}$ and SFR) . We have already stressed that there are different contributing factors to FIR luminosities, and in particular early spiral galaxies often exhibit low temperature, relatively high FIR luminosities attributable to dust heating from the general stellar radiation field, and not directly related to SFR (Kennicutt \cite{kenni}). It has been claimed that a more reliable discriminant of the SFR is the infrared excess L$_{\rm FIR}$/L$_{\rm B}$ (eg. Tomita et al. \cite{tomita}). This is because by normalizing to the blue luminosity we partially remove the effect of the general radiation field. In Table~\ref{lfirlb} we report the SN rates in SNu for galaxies with different infrared excess, along with that of galaxies not detected by IRAS. Though in general we cannot translate ``not detected'' into a precise upper limit, it is reasonable to assume that, for our RC3 galaxy sample, the average FIR luminosity of the undetected sample is smaller than that of the detected sample. Support for this belief comes from the fact that the distance distributions of the detected and not detected RC3 galaxy samples are similar. \begin{table*} \caption{SN rates in SNu for galaxies with different infrared excess }\label{lfirlb} \begin{tabular}{lcccccc} \hline galaxy & \multicolumn{2}{c}{not detected by IRAS} & \multicolumn{2}{c}{$L_{FIR}/L_{B}\le 0.35$} & \multicolumn{2}{c}{$L_{FIR}/L_{B}>0.35$} \\ type & Ia & II+Ib/c & Ia & II+Ib/c & Ia & II+Ib/c \\ \hline E-S0 &$0.2\pm0.1$& & $0.4\pm0.2$ & &$<0.5$ &\\ S0a-Sb &$0.2\pm0.1$& $0.3\pm0.2$& $0.2\pm0.1$ & $0.5\pm0.2$ &$0.3\pm0.1$ &$1.1\pm0.4$\\ Scd-Sd &$0.3\pm0.1$& $0.7\pm0.3$& $0.2\pm0.1$ & $1.0\pm0.2$ &$0.2\pm0.1$ &$1.2\pm0.3$\\ \\ All$^*$ &$0.2\pm0.1$& $0.2\pm0.1$& $0.2\pm0.1$ & $0.9\pm0.1$ &$0.3\pm0.1$ &$1.1\pm0.2$\\ \hline \end{tabular} $*$ Including Sm, irregulars and peculiars. \end{table*} The rate of core collapse SNe is higher in the IR detected galaxies compared with the not detected sample, whereas this is not the case for SN~Ia (Table~\ref{lfirlb}) whilst there are no significant differences between galaxies with small and large infrared excess. This again supports the idea that, whereas a fraction of the FIR luminosity originates from SF regions, the other contributing factors to the IR emission of galaxies eliminate, at least in normal galaxies, the relation between L$_{\rm FIR}$ and SFR. \subsection{SN rates in active galaxies}\label{AGN} It is generally believed that nuclear activity stimulates the SF (Rodriguez-Espinoza et al. \cite{rodri}) and therefore that the rate of core-collapse SNe in AGN must be higher than in normal galaxies. An open issue is whether the SF is stimulated throughout the whole AGN host galaxy or only in the circumnuclear region. From the observational point of view, in the first case we would expect an enhanced detection rate, whereas due to the high extinction in the nuclear starburst regions this may not occur in the latter case. To address this question we crossed our RC3 galaxy list with the Catalog of Quasars and Active Galactic Nuclei of V\'eron-Cetty \& V\'eron (\cite{veron}) (distributed by the CDS). This catalog contains a list of almost 15000 quasars and AGN most of which are too distant for normal SN searches (only $\sim 1100$ have recession velocities smaller than 15000 km s$^{-1}$). We found that 283 galaxies out of our combined RC3 sample, ($\sim$ 3\%) are also listed in the V\`eron-Cetty \& V\`eron catalog (this simply reflects the relative occurrence of AGN in the local Universe). Most of them ($\sim$ 88\%) are Seyfert and the rest HII galaxies. In these galaxies our searches have discovered 17 SNe, that is 12\% of the total SN sample. This could be taken as evidence that the SN rate in AGN is enhanced compared with the general sample. However, when the control time method is applied, the average SN rate in the AGN sample is $0.6\pm0.1$ SNu ($0.4\pm0.1$ SNu for core-collapse SNe), identical to that of the general sample ($0.7\pm0.1$ SNu for all SNe and $0.5\pm0.1$ SNu for core-collapses). We note that the AGN sample shows roughly the same distribution of morphological types as the general sample. There are two reasons why the high detection rate in our AGN galaxy sample does not reflect in higher SN rates in SNu. First of all, the average control time for a galaxy of the AGN galaxy sample (5.08 yr) is almost twice that of the general galaxy sample (2.67 yr). Secondly, the galaxies of the AGN sample are over 2 times more luminous ($<L_B> = 3.1^{10} L_\odot$) than the average ``normal'' galaxies ($<L_B> = 1.4^{10} L_\odot$). This stresses the risks of interpreting statistics derived from general SN samples and not from actual search logs. A similar conclusion was reached by Richmond et al. (\cite{richm}) as the result of a dedicated SN search in 142 nearby starburst galaxies. They obtained 1.1 SNu (scaled to $H_0=75$) for the total rate and 0.7 for core-collapse only, somewhat larger than our corresponding estimates. However, their statistical error is also quite large (they had a sample of only 5 SNe) and allowing also for the different computational protocols, the difference should not be regarded as significant. The conclusion is that the SN rate in active galaxies is the same as in normal ones (cf Petrosian \& Turatto \cite{petro}). More precisely, this finding only applies to the AGN host galaxies and not to the AGNs themselves, which because of the high extinction rate cannot be probed by optical SN searches. Therefore our claim is that the nuclear engine does not significantly stimulate the SFR outside the nuclear region of the host galaxy. \section{Conclusions} We have presented new estimates of the SN rates in galaxies, obtained by including the updated log of the Evans' visual SN search in our database. In this way we have obtained a sample of 137 SNe in a reference sample of about $10^4$ galaxies. Based on the comparison between visual and photographic surveys we tested the effectiveness of the bias corrections and verified that they are consistent with our understanding of galaxies and SN progenitors. In particular, we show that the Hatano et al. (\cite{hatano}) simple model for the SN and dust distributions in galaxies explains, at least to the first order, both the bias in the nuclear region and in inclined spirals, though actually some refinement is needed before it can be used to correct SN rates. The new rates have been compared with other tracers of the average SFR in galaxies. We found that the rates of core-collapse SNe are higher in bluer spirals, while the same is not true for SNIa. This was expected, since bluer galaxies host stars that are younger and more massive than redder ones. In particular, we find that the correlation between galaxy colors and core-collapse SN rates is similar to that predicted by the evolutionary models of Kennicutt (\cite{kenni}). We have found that there is not a direct relation between core-collapse SN rates and FIR luminosities confirming that FIR luminosity is not a universal measurement of SFR. This can be explained by considering that FIR emission in galaxies at least in the normal ones, is made up of different components and not exclusively related to young stars. Finally, our data confirms previous findings that the SN rates in AGN host galaxies are not enhanced. This conclusion does not apply to the nuclear starburst regions which cannot be probed by current SN searches. \begin{acknowledgements} R.E. wishes to thank Mr. R. McDonell of Ark Angles Business and Personal Computer Systems, Hazelbrook (Australia) for designing the computer program used to record his observations. We wish also to thank Janet Clench for revising the manuscript. \end{acknowledgements} \bibliographystyle{astron}
\section{Nonlinear Modifications of Quantum Mechanics} Many authors\footnote{Let us mention just some of them: T.W.B.~Kibble\,\cite{Kibble}, A.~Ashtekar and T.A.~Schilling\,\cite{Ashtekar}, P.~Bona\,\cite{Bona}, R.~Haag and U.~Bannier\,\cite{Haag-Bannier}, Mielnik\,\cite{Mielnik}, S.~Weinberg\,\cite{Weinberg}, G.~Auberson and P.C.~Sabatier\,\cite{Sabatier}, M.D.~Kostin\,\cite{Kostin}, H.D.~Doebner and Goldin\,\cite{dg1,dg2}, I.~Bialynicki-Birula and J.~Mycielski\,\cite{bbm}.} considered nonlinear Schr\"odinger equations of the form \begin{equation} \label{NLSE} i\hbar\frac{\partial}{\partial t} \Psi_t({\vec x}) = \left(-\frac{\hbar^2}{2m}\Delta+V({\vec x},t)\right) \Psi_t({\vec x}) + R[\Psi_t]({\vec x}) \Psi_t({\vec x})\,, \end{equation} where $F[\Psi] = R[\Psi] \Psi$ is a local\footnote{Here locality means $\Psi({\vec x})=0\Longrightarrow \left(F[\Psi]\right)({\vec x}) =0$ in the distribution theoretical sense.} nonlinear Functional of $\Psi\,$. The essential point is that (\ref{NLSE}) is not interpreted as an equation for some field operator but as a classical evolution equation for the quantum mechanical wave function: \begin{equation} \label{StInt} \int_{{\cal B}} \modulus{\Psi_t({\vec x})}^2\,{\rm d}{\vec x} = \parbox{6cm}{probability for location within ${\cal \region B}$} \end{equation} (at time $t$ for normalized $\Psi_t$). Special cases were even tested experimentally.\,\cite{tests} \vskip 5mm \noindent All these efforts seemed useless according to N.~Gisin's claim\,\cite{gis1,gis2}: \begin{equation} \label{gis-cl} \mbox{``All deterministic nonlinear Schr\"odinger equations are irrelevant.''} \end{equation} By this Gisin meant the following: \begin{quote} Consider a Bell-like situation as sketched in Figure \ref{fig:gisin}. \begin{figure}[ht $$ \font\thinlinefont=cmr5 \mbox{\beginpicture \setcoordinatesystem units <0.9cm,0.9cm> \unitlength=1.04987cm \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) \setshadesymbol ({\thinlinefont .}) \setlinear \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \circulararc 134.333 degrees from 4.572 24.575 center at 4.006 25.121 \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \circulararc 65.361 degrees from 4.572 24.575 center at 3.247 24.796 \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \put{\makebox(0,0)[l]{\circle*{ 0.047}}} at 0.948 24.213 \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \put{\makebox(0,0)[l]{\circle*{ 0.047}}} at 0.961 23.995 \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \putrule from 7.144 24.289 to 3.715 24.289 \plot 3.969 24.352 3.715 24.289 3.969 24.225 / \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \putrectangle corners at 2.857 24.575 and 3.715 24.098 \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \plot 2.762 24.384 1.194 25.290 / \plot 1.446 25.218 1.194 25.290 1.382 25.108 / \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \plot 1.321 25.453 1.238 25.502 / \plot 1.238 25.502 1.048 25.171 / \plot 1.048 25.171 1.130 25.125 / \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \plot 2.762 24.289 0.978 24.604 / \plot 1.239 24.622 0.978 24.604 1.217 24.497 / \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \plot 0.991 24.807 0.897 24.824 / \plot 0.897 24.824 0.830 24.448 / \plot 0.830 24.448 0.923 24.431 / \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \plot 2.766 24.196 1.065 23.575 / \plot 1.282 23.722 1.065 23.575 1.325 23.603 / \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \plot 1.016 23.779 0.929 23.747 / \plot 0.929 23.747 1.056 23.387 / \plot 1.056 23.387 1.145 23.421 / \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \putrule from 7.239 24.289 to 7.429 24.289 \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \putrule from 7.334 24.384 to 7.334 24.194 \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \plot 7.262 24.361 7.398 24.225 / \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \plot 7.404 24.369 7.269 24.234 / \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \putrule from 7.525 24.289 to 9.906 24.289 \plot 9.652 24.225 9.906 24.289 9.652 24.352 / \put{$A$} [lB] at 3.143 24.194 \put{$\alpha_n$} [lB] at 0.45 23.461 \put{$\alpha_1$} [lB] at 0.622 25.364 \put{$\alpha_2$} [lB] at 0.32 24.558 \put{$\rho_{\rm right}$} [lB] at 10.097 24.194 \put{Source} [lB] at 6.858 23.812 \put{\scriptsize Moon} [lB] at 4.6 25.75 \linethickness=0pt \putrectangle corners at 0.434 25.929 and 10.097 23.362 \endpicture} $$ \caption{Gisin's Gedanken experiment. \label{fig:gisin}} \end{figure} Then, if the source produces entangled 2-particle states, there is always a physical observable for the particle sent to the right,\footnote{Note that, contrary to Bell nonlocality,\cite{cl-sh} the observable effect does not refer to the resulting correlations between both particles.} the probability distribution of which is instantaneously (substantially) changed by suitable measurements (involving only low energy transfer) on the other particle `behind the moon'. \end{quote} \vskip 5mm \noindent Actually, Gisin\,\cite{gis1} assumed the following:\footnote{His justification: ``So far we have only used linear quantum mechanics''.} \begin{quote} \begin{itemize} \item[(G1):] If the observable $A$ resp.~$B$ of the particle `behind the moon' is measured at time $t=0$ the partial state of the other particle at times $t\geq0$ is given by a density matrix $\rho_{\rm right}(t)$ of the form\footnote{This especially implies $\sum_{\alpha} x_\alpha\, P_\Psi(\alpha,t) = \sum_{\beta} x_\beta\, P_\Psi(\beta,t)$ for $t=0$ but --- as realized by Gisin --- not generally for $t>0\,$.} $$ \sum_{\alpha} x_\alpha\, P_\Psi(\alpha,t)\quad \mbox{resp.}\quad \sum_{\beta} x_\beta\, P_\Psi(\beta,t)\,, $$ where the $P_\Psi(\alpha,t)$ resp.~$P_\Psi(\beta,t)$ are pure states evolving according to the corresponding 1-particle equation. \item[(G2):] {\bf All} self-adjoint \aside{bounded} operators correspond to observables. \end{itemize} \end{quote} \vskip 5mm That Gisin's claim (\ref{gis-cl}) is wrong since Gisin's assumption (G1), the projection postulate, is unjustified in nonlinear quantum mechanics has already been pointed out by Polchinski\,\cite{pol}, who determined conditions which are sufficient for the absence of essential nonlocality.\footnote{See also \cite{czach1} for examples.} \vskip 5mm Accepting Gisin's assumption (G2), Polchinski concluded that his conditions -- violated for prominent examples of nonlinear Schr\"odinger equations -- are also necessary to avoid essential nonlocality. However, as explained already in \cite{lu1}, also assumption (G2) is unjustified in nonlinear quantum mechanics and definitely wrong for the situation reconsidered in the next section. This is why valid proofs of essential nonlocality have to be more involved. \medskip \section{What we can Learn from Nonlinear Gauge Transformations} \label{section:CE} Consider the well-defined special case $$ \left(\Op N_{\! D}(\Psi)\right)({\vec x}) \stackrel{\rm def}{=} e^{i\frac{2mD}{\hbar}\ln\modulus{\Psi({\vec x})}}\Psi({\vec x})\;,\quad D\in{\rm I\!R}\,. $$ of the `nonlinear gauge transformations' exploited by H.-D.~Doebner et al.\,\cite{nlgt} If $\Psi_t'({\vec x})$ is a solution of (\ref{NLSE}) for $R=0$ then straightforward calculation shows that $$ \Psi_t({\vec x}) = \left(\Op N_{\! D}(\Psi'_t)\right)({\vec x}) $$ is a solution of (\ref{NLSE}) for \begin{equation} \label{GDGE} R[\Psi_t] = \hbar D \Bigl(\frac{i}{2} \frac{\Delta\rho_t}{\rho_t}\Psi_t + c_1 \frac{{\vec \nabla}\cdot \vec {J}_t}{\rho_t} + c_2 \frac{\Delta\rho_t}{\rho_t} + c_3 \frac{\vec{J}_t}{\rho_t}+ +c_4 \frac{\vec{J}_t\cdot {\vec \nabla}\rho_t}{(\rho_t)^2} + c_5 \frac{({\vec \nabla}\rho_t)^2}{(\rho_t)^2}\Bigr) \end{equation} with\footnote{The {\em general Doebner-Goldin equation\/} is given by (\ref{NLSE}) and (\ref{GDGE}) without the restriction (\ref{el}) on the parameters $c_\nu\in{\rm I\!R}\,$.} \begin{equation} \label{el} c_1=1\;, \quad c_2=-2c_5=-mD/\hbar\;,\quad c_3=0\;,\quad c_4=-1\,, \end{equation} where $$ \rho_t\stackrel{\rm def}{=}\modulus{\Psi_t}^2\,,\; \vec{J}_t \stackrel{\rm def}{=} \frac{1}{2i} \left( \overline{\Psi_t} {\vec \nabla} \Psi_t - \Psi_t {\vec \nabla} \overline{\Psi_t} \right)\,. $$ This way we get a deterministic nonlinear Schr\"odinger equation which, interpreted by (\ref{StInt}), describes the same physics as the corresponding ($R=0$) linear Schr\"odinger equation, since $$ \modulus{\Psi_t({\vec x})} = \modulus{\Psi'_t({\vec x})}\,. $$ That, contrary to Gisin's as well as Polchinski's claim, there is no real problem with locality is no surprise since now nonlinear projection operators $$ \Op E = \Op N_D \circ \Op E' \circ \Op N_D^{-1}\,, $$ instead of linear projectors $\Op E'\,$, have to be used\,\cite{lu1} to get the correct probabilities $$ \norm{\Op E(\Psi_t)}^2 = \norm{\Op E'(\Psi'_t)}^2 = \left\langle \Psi'_t \mid \Op E'\,\Psi'_t \right\rangle\,. $$ Hence assumption (G2) is obviously wrong in nonlinear quantum mechanics. Moreover, w.r.t.~the nonlinear $\Op E\,$, density matrices are inadequate for the description of classical mixtures: $$ \sum_{\alpha} x_\alpha \norm{\Op E(\Psi_\alpha)}^2 = {\rm trace}\left(\Op E'\, \sum_\alpha x_\alpha \Op P_{\Psi'_\alpha}\right) \stackrel{\rm i.g.}{\ne} {\rm trace}\left(\Op E\, \sum_\alpha x_\alpha \Op P_{\Psi_\alpha}\right)\,. $$ Therefore also assumption (G1) turns out to be quite inadequate for nonlinear modifications of quantum mechanics. The simple example (\ref{NLSE})/(\ref{GDGE})/(\ref{el}) tells us that essential nonlocality should to be checked by using nothing else than the evolution equation together with its basic interpretation (\ref{StInt}). \medskip \section{The Doebner-Goldin Equation Interpreted as 2-Particle Equation} \label{section:dg} Let us interpret (\ref{NLSE}) as a two-particle equation: $$ {\vec x}=({\vec x}_1,{\vec x}_2)\;,\quad {\vec x}_j = \mbox{position of particle } j\,. $$ Moreover assume that the potential is of the form $$ V({\vec x}) = V_2({\vec x}_2-{\vec x}_0)\;,\quad {\vec x}_0 \mbox{ fixed}\,, $$ and that $\Psi_t({\vec x})$ is the solution of (\ref{NLSE}) fulfilling the initial condition $$ \Psi_0({\vec x}_1,{\vec x}_2) = f({\vec x}_1,{\vec x}_2-{\vec x}_0)\,. $$ Then, obviously, we have an unacceptable nonlocality, if the position probability density\footnote{Note that (\ref{pp1}) does not depend on ${\vec x}_0\,$. Therefore the effect on particle 1, if any, can be produced by acting on particle 2 as far away as one likes.} \begin{equation} \label{pp1} \rho_1({\vec x}_1,t) \stackrel{\rm def}{=} \int \modulus{\Psi_t({\vec x}_1,{\vec x}_2)}^2{\rm d}{\vec x}_2 \end{equation} for particle 1 depends on $V_2\,$. That the latter happens for certain cases of the general Doebner-Goldin equation, if interpreted as 2-particle equation, was first proved by Werner.\,\cite{werner1} Inspired by E.~Nelson\,\cite{nelson} he considered pairs of 1-dimensional particles, i.e.~${\vec x}=(x_1,x_2)\in{\rm I\!R}^2\,$, with \begin{equation} \label{osz} V_2(x_2) = \lambda\,(x_2)^2\;,\quad \lambda\in{\rm I\!R}\,, \end{equation} and entangled initial conditions of Gaussian type. The corresponding solutions are of the form $$ \Psi_t(x_1,x_2)= e^{\gamma(t) -\sum_{j,k=1}^2 C_{jk}(t)x_jx_k/2}\;,\quad C_{jk}=C_{kj}\,, $$ where the $C_{jk}(t)$ fulfill a simple system of first order ordinary differential equations that can be used to determine their time derivatives at $t=0$ and thus \begin{equation} \label{test1} \left({\partial_t}^n\int (x_1)^2 \rho_1(x_1,t)\,{\rm d}x_1\right)_{|_{t=0}} \end{equation} for arbitrary $n\in{\mathchoice {\hbox{$\sans\textstyle Z\kern-0.4em Z$}_+\,$. Werner found that (\ref{test1}) depends on $\lambda$ for $n=3$ and suitable $C_{jk}(0)$ unless \begin{equation} \label{gi} c_3=c_1+c_4=0\,. \end{equation} In principle, using (\ref{NLSE}) directly,\,\cite{luna} one may calculate \begin{equation} \label{test2} \partial_\lambda \left(\partial_t^n \int\modulus{\Psi_t(x,y)}^2\,{\rm d}y\right)_{|_{t=0}} \end{equation} as a functional of $\Psi_0$ and $V$ for given $R\,$. This shows that (\ref{test1}) varies also with strictly localized changes of $V_2\,$. However, for $n>3$ the calculation becomes too involved and could not even managed by use of computer algebra. On the other hand, Werner's Ansatz turned out to be too special to uncover essential nonlocality of the Galilei covariant cases of the Doebner-Goldin 2-particle equation.\footnote{Note that (\ref{gi}) is equivalent to Galilei covariance of the Doebner Goldin equation.\,\cite{dg2}} This is why \begin{equation} \label{test3} \left({\partial_t}^n\int x_1 \rho_1(x_1,t)\,{\rm d}x_1\right)_{|_{t=0}} \end{equation} was checked for $n=4$ in \cite{lu2} showing that the Doebner-Goldin equation is essentially nonlocal, when interpreted as a 2-particle equation, unless the parameters $c_\nu$ are chosen such that (\ref{NLSE}) is (formally) linearizable by some nonlinear gauge transformation. \medskip \section{Inconsistency of Bialynicki-Birula's and Mycielski's Theory} \label{section:bbm} For the nonlinear Schr\"odinger equation of Bialynicki-Birula and Mycielski, given by\footnote{Equation (\ref{bbm}) was already considered by H.~Ko\v s\v t\'al\,\cite{kostal} with $b<0\,$.} \begin{equation} \label{bbm} \left(R[\Psi]\right)({\vec x}) = -2b\,\ln\modulus{\Psi({\vec x})}\;,\quad b\in{\rm I\!R}\,, \end{equation} testing (\ref{test3}) is of no use, since the Ehrenfest relations hold.\,\cite{bbm} In such cases one should check \begin{equation} \label{test4} S_{k,n}[\Psi_0,V_2] \stackrel{\rm def}{=} \partial_\lambda \left(\partial_t^n \int e^{ikx}\,\modulus{\Psi_t(x,y)}^2 {\rm d}x\,{\rm d}y\right)_{|_{t=0}} \end{equation} for $k\in{\rm I\!R}$ and $n\in{\mathchoice {\hbox{$\sans\textstyle Z\kern-0.4em Z$}_+\,$. In principle this is equivalent to testing (\ref{test2}) but has the advantage of testing (\ref{test3}): The integral over $x$ allows for partial integrations which simplify the resulting expressions considerably. \vskip 5mm \noindent For simplicity, let $\frac{\hbar^2}{2m}=1$ and assume $R$ to be real valued,\footnote{For the Doebner-Goldin case (\ref{GDGE}) the latter can always be achieved\,\cite{NaDip} by some nonlinear gauge transformation of the type considered in Section \ref{section:CE}.} as in (\ref{bbm}): \begin{equation} \label{rR} \left(R[\Psi]\right)({\vec x}) = \overline{\left(R[\Psi]\right)({\vec x})}\,. \end{equation} Then, defining $$ \begin{array}[c]{rcl} T_{k,\nu}(t) &\stackrel{\rm def}{=}&\displaystyle \int e^{ikx}\,\overline{\Psi_t}\,\partial_x^\nu \Psi_t\,{\rm d}x\,{\rm d}y\,,\\ \displaystyle D_{k,\mu,\nu}(t) &\stackrel{\rm def}{=}&\displaystyle \int e^{ikx}\,\left(\partial_x^\mu R[\Psi_t]\right)\overline{\Psi_t}\, \partial_x^\nu \Psi_t\,{\rm d}x\,{\rm d}y\,, \end{array} $$ we get from (\ref{NLSE}) by partial integration $$ i\frac{{\rm d}}{{\rm d} t} T_{k,\nu}= -k^2\, T_{k,\nu} + 2ik\, T_{k,\nu+1} + \sum_{\mu=1}^{\nu}{\nu\choose \mu} D_{k,\mu,\nu-\mu} $$ for ${\vec x}=(x_1,x_2)\in{\rm I\!R}^2$ and $V({\vec x})=V_2(x_2)\,$. Iteration of this gives $$ \begin{array}[c]{rcl} \displaystyle \left(i \frac{{\rm d}}{{\rm d} t}\right)^3 T_{k,0} &=& -k^6\, T_{k,0} + 6ik^5\, T_{k,1} + 12k^4 T_{k,2} - 8 ik^3 T_{k,3}\\ && -\left(4ik^3\,D_{k,1,0} +8k^2 D_{k,1,1} + 4k^2 D_{k,2,0} + 2k\partial_t D_{k,1,0}\right)\,. \end{array} $$ Hence, e.g., $$ \begin{array}[c]{rcl} \displaystyle \left(\partial_t^{n+3}\!\int x^2\modulus{\Psi_t(x,y)}^2{\rm d}x\,{\rm d}y \right)_{|_{t=0}} &=&\displaystyle -\left(\partial_t^{n+3}\,(\partial_k)^2 T_{k,0}(t)\right)_{|_{t=k=0}}\\[4mm] &=& 8i\,\partial_t^n\left(2\,D_{0,1,1} + D_{0,2,0}\right). \end{array} $$ For the special case $$ m=1\;,\quad V_2(y) = \lambda\,y^2\;,\quad \Psi_0(x,y) = \frac{e^{-x^2 -y^2 -xy}}{\int e^{-2x^2 -2y^2 -2xy}\, {\rm d}x\,{\rm d}y}\,, $$ running a simple computer algebra program (see appendix) shows that (\ref{NLSE}) and (\ref{bbm}) imply\footnote{This result was confirmed by R.~Werner\,\cite{werner2} using his method described in Section \ref{section:dg}.} $$ \partial_\lambda\left((i\partial_t)^3\left(2\,D_{0,1,1}(t) + D_{0,2,0}(t)\right)\right)_{|_{t=0}} = 32b\,. $$ This means that (\ref{NLSE})/(\ref{bbm}) is essential nonlocal --- against the basic philosophy of Bialynicki-Birula's and J.~Mycielski's theory.\cite{bbm} \medskip \section{Identical Particles} Up to now we tacitly assumed that the two particles (with equal masses) considered in Figure \ref{fig:gisin} can be distinguished. Therefore one might still hope that Bialynicki-Birula's and J.~Mycielski's theory is consistent for identical particles. However, even for 2-particles states which are symmetric or antisymmetric w.r.t.~exchange of the particles essential nonlocality is unavoidable. To show this denote by $\Psi_t^{g,U}$ the solution of $$ i\partial_t \Psi_t^{g,U} = \left(-\Delta +\lambda U - 2b \ln\modulus{\Psi_t^{g,U}}\right)\Psi_t^{g,U} $$ fulfilling the initial condition $$ \Psi_0^{g,U}(x,y) = g(x,y)\,. $$ For fixed $f(x,y)$ and $\sigma\in\left\{+1,-1\right\}$ define $$ \begin{array}[c]{c} \begin{array}[c]{rcl} U^{(d)}(x,y) &\stackrel{\rm def}{=}& V(y-d) + V(x-d)\,,\\ \Psi_0^{(d)}(x,y) &\stackrel{\rm def}{=}& f(x,y-d) +\sigma f(y,x-d)\,, \end{array}\\ \chi_0^{(d)}(x,y) \stackrel{\rm def}{=} f(x,y-d)\;,\quad \phi_0^{(d)}(x,y) \stackrel{\rm def}{=} f(y,x-d)\,. \end{array} $$ and $$ \chi_t^{(d)} \stackrel{\rm def}{=} \Psi_t^{\chi_0^{(d)},U^{(d)}}\;,\quad \phi_t^{(d)} \stackrel{\rm def}{=} \Psi_t^{\phi_0^{(d)},U^{(d)}}\;,\quad \Psi_t^{(d)} \stackrel{\rm def}{=} \Psi_t^{\Psi_0^{(d)},U^{(d)}}\,. $$ Obviously, if $V\in\testspace D({\rm I\!R})$ and $f\in\testspace S({\rm I\!R}^2)\,$, $$ \partial_\lambda\left(\partial_t^6 \int_{\cal G}\left( \modulus{\Psi^{(d)}_t(x,y)}^2 - \modulus{\chi^{(d)}_t(x,y)}^2 - \modulus{\phi^{(d)}_t(x,y)}^2\right){\rm d}x{\rm d}y\right)_{|_{t=0}} \stackscript{\longrightarrow}_{d \to \infty} 0 $$ holds for every region ${\cal G}\subset{\rm I\!R}^3\times{\rm I\!R}^3\,$. \noindent Moreover, $$ \begin{array}[c]{l} \displaystyle \lim_{d\to\infty} \partial_\lambda\left( \partial_t^6 \int_{\cal G} \modulus{\phi^{(d)}_t(x,y)}^2\,{\rm d}x{\rm d}y\right)_{|t=0}\\[8mm] =\displaystyle \lim_{d\to\infty} \partial_\lambda\left( \partial_t^6 \int_{\cal G} \modulus{\chi^{(d)}_t(x,y)}^2\,{\rm d}x{\rm d}y\right)_{|t=0}\\[8mm] = c^{f,U}_{\lambda} \stackrel{\rm def}{=}\displaystyle \partial_\lambda\left( \partial_t^6 \int_{x\in\cal O} \modulus{\Psi_t^{f,U}(x,y)}^2 {\rm d}x\,{\rm d}y\right)_{|t=0} \end{array} $$ holds for $$ \region G=\left\{(x,y)\in{\rm I\!R}^2: x\in\region O\,\lor\,y\in\region O\right\}\;,\quad \region O \mbox{ bounded}\;,\quad U(x,y)=V(y)\,. $$ Therefore, under these conditions, $$ \partial_\lambda\left(\partial_t^6 \int_{\cal G}\left( \modulus{\Psi^{(d)}_t(x,y)}^2\,{\rm d}x\,{\rm d}y\right) \right)_{|_{t=0}} \stackscript{\longrightarrow}_{d \to \infty} = c^{f,U}_\lambda\,. $$ Since, as shown in Section \ref{section:bbm}, $c^{f,U}_{\lambda}$ can be arranged to be nonzero we conclude: \begin{quote} The postulate of symmetry or antisymmetry of the wave function w.r.t.~to exchange of particles does not prevent essential nonlocality. \end{quote} \section*{Acknowledgments} I am grateful to M.~Czachor, H.-D.~Doebner, G.A.~Goldin, and R.~Werner for stimulating discussions. \section*{Appendix: Maple V (Release 4) Session} \scriptsize \begin{verbatim} PROCEDURES: > del := proc(f) > global x,y,t; > option operator; > unapply( diff(f(x,y,t), x$2) + diff(f(x,y,t), y$2), (x,y,t)); > end: > pot := proc(f) > global x,y,t; > option operator; > unapply(V(y) * f(x,y,t),(x,y,t)); > end: > Idot := proc(f) > global x,y,t; > option operator; > unapply(simplify(subs(diff(P(x,y,t),t)=(-del(P)(x,y,t)+ pot(P)(x,y,t)), diff(PB(x,y,t),t)=(del(PB)(x,y,t)-pot(PB)(x,y,t)), diff(f(x,y,t), t))), (x,y,t)); > end: EVALUATION: > term0 := (x,y,t) -> 2 * diff(ln( PB(x,y,t) * P(x,y,t)), x) * PB(x,y,t) * diff(P(x,y,t), x) + diff(ln( PB(x,y,t) * P(x,y,t)), x$2) * PB(x,y,t) * P(x,y,t): > term1 := (x,y,t) -> simplify(Idot(term0)(x,y,t)): > term2 := (x,y,t) -> simplify(Idot(term1)(x,y,t)): > term3 := (x,y,t) -> simplify(Idot(term2)(x,y,t)): SPECIAL CASE: > spec := proc(f) > global x,y,t; > option operator; > unapply( subs(V(y)=lambda * y^2, P(x,y,t)=exp(-x^2 -y^2 -x*y), PB(x,y,t)=exp(-x^2 -y^2 -x*y), f(x,y,t)), (x,y,t)); > end: RESULT: > int(int(simplify(diff(spec(term3)(x,y,t),lambda)), x=-infinity..infinity), y=-infinity..infinity); 1/2 - 32/3 Pi 3 > int(int(exp(-2*x^2 -2*y^2 -2*x*y), x=-infinity..infinity), y=-infinity..infinity); 1/2 1/3 Pi 3 \end{verbatim} \normalsize
\section{Introduction} If string theory is to be a model of supersymmetric particle physics, one should be able to find four-dimensional vacua preserving $\mathcal{N}=1$ supersymmetry. The first examples of such backgrounds arose as geometrical vacua, from the compactification of the heterotic string on a Calabi--Yau threefold. Recent advances in string duality have provided new ways of obtaining geometrical $\mathcal{N}=1$ vacua, by compactifying other limits of string theory and, in particular, by including branes in the background. Of particular interest is the strongly coupled limit of the $E_8\times E_8$ heterotic string. The low-energy effective theory is eleven-dimensional supergravity compactified on an $S^1/Z_2$ orbifold interval, with a set of $E_8$ gauge fields on each of the ten-dimensional orbifold fixed planes~\cite{hw1,hw2}. To construct a theory with $\mathcal{N}=1$ supersymmetry one further compactifies on a Calabi--Yau threefold $X$~\cite{w:cy}. One is then free to choose general $E_8$ bundles which satisfy the hermitian Yang--Mills equations. Furthermore, one can include some number of five-branes. The requirements of four-dimensional Lorentz invariance and supersymmetry mean that these branes must span the four-dimensional Minkowski space, while the remaining two dimensions wrap a holomorphic curve within the Calabi--Yau threefold~\cite{w:cy,bbs,bvs}. In the low-energy four-dimensional effective theory, there is an array of moduli. There are geometrical moduli describing the dimensions of the Calabi--Yau manifold and orbifold interval. There are bundle moduli describing the two $E_8$ gauge bundles. And finally, there are moduli describing the positions of the fivebranes. It is this last set of moduli, together with the generic low-energy gauge fields on the fivebranes, upon which we shall focus in this paper. We note that, although we consider this problem in the specific case of M~fivebranes, the moduli spaces are quite general and should have applications to other supersymmetric brane configurations. We shall look at this question in the particular case of an elliptically fibered Calabi--Yau threefold with section. This expands on the discussion of we gave in a recent letter~\cite{dlow1}. There we used this special class of manifolds to construct explicitly an array of new particle physics vacua. The general structure of the constructions was given in detail in a second paper~\cite{dlow2}. This is the companion paper which explains the structure of the five-brane moduli space as well as the nature of gauge enhancement on the five-branes. The constructions in~\cite{dlow1} and~\cite{dlow2} used the analysis of gauge bundles on elliptic Calabi--Yau threefolds given by Friedman, Morgan and Witten~\cite{fmw}, Donagi~\cite{ron} and Bershadsky \textit{et al.}~\cite{bjps}. The vacua preserved, for example, $SU(5)$ or $SO(10)$ gauge symmetry with three families. The presence of the five-branes allowed a much larger class of possible backgrounds. The number of families is given by an index first calculated in this context by Andreas~\cite{andreas} and Curio~\cite{curio}. Curio also gave explicit examples of bundles where this index was three. Subsequently, the case of non-simply connected elliptic Calabi--Yau threefolds with bisections has been considered in~\cite{ack}. We note that, in the M~theory context, another class of explicit models with non-standard gauge bundles but with orbifold Calabi--Yau spaces, were first contructed in~\cite{stie}, while the generic form of the effective four- and five-dimensional theories, including fivebranes, is given in~\cite{nse}. In constructing these vacua the fivebranes cannot be chosen arbitrarily~\cite{w:cy}. The boundaries of $S^1/Z_2$ and the fivebranes are all magnetic sources for the four-form field strength $G$ of eleven-dimensional supergravity. The fact that there can be no net magnetic charge for $G$ in the Calabi--Yau threefold fixes the homology class of the fivebranes in terms of the gauge bundles and $X$ itself. As discussed in~\cite{dlow1,dlow2}, to describe real fivebranes this homology class must be ``effective'' in $X$. Mathematically, then, the problem we wish to solve is to find the moduli space of holomorphic curves in $X$ in a given effective homology class. Specific to the M~theory case, we will also include the moduli corresponding to moving the fivebranes in $S^1/Z_2$ and, in addition, their axionic moduli partners~\cite{nse,w:hw,ketal}. These latter fields are compact scalars which arise as zero modes of the self-dual three-form $h$ on the fivebranes. The other zero modes of $h$ lead to low-energy gauge fields. Generically the gauge group is $U(1)^g$, where $g$ is the genus of the holomorphic curve~\cite{nse,w:hw,ketal}. Since we will be able to calculate $g$ at each point in the moduli space, we will also be able to identify the low-energy gauge multiplets. One consequence of considering elliptically fibered Calabi--Yau threefolds with section is that there is a dual F~theory description~\cite{vafa,vm1,vm2}. For the fivebranes, those wrapping purely on the fiber of $X$ correspond to threebranes on the F~theory side~\cite{fmw}. Rajesh~\cite{rajesh} and Diaconescu and Rajesh~\cite{dr} have recently argued that fivebranes lying wholly in the base of $X$ correspond to blow-ups of the corresponding curve in the F~theory vacuum. We will not comment in detail on this interesting correspondence. However, we will show that, locally, the moduli spaces match those expected from duality to F~theory. We will also comment on how the global structure is encoded on the M~theory side through a twisting of the axion modulus. This will be discussed further in~\cite{atwist}. An additional point we will only touch on is the structure of additional low-energy fields which can appear when fivebranes intersect. We will, however, clearly identify these points in moduli space in our analysis. Finally, we will also ignore any non-perturbative corrections. In general, since the low-energy theory is only $\mathcal{N}=1$, one expects that some of the directions in moduli space are lifted by non-perturbative effects, in particular by instantonic membranes stretching between fivebranes. Specifically, we do the following. In section~2, we briefly review the anomaly cancellation condition in heterotic M-theory, discuss how that constraint leads to non-perturbative vacua with fivebranes and review some aspects of homology and cohomology theory required in our analysis. Properties of elliptically fibered Calabi--Yau threefolds and a discussion of their algebraic and effective classes are presented in section~3. Section~4 is devoted to studying the simple case of the the moduli spaces of fivebranes wrapped purely on the elliptic fiber. We also comment on global structure of the moduli space and the relation to F~theory. In section~5, we present two examples of fivebranes wrapping curves with a component in the base. We analyze, in detail, the moduli space of these two examples, including the generic low-energy gauge groups on and possible intersections of the fivebrane. Techniques developed in sections~4 and~5 are generalized in section~6, where we give a procedure for the analysis of the moduli spaces of fivebranes wrapped on any holomorphic curve, generically with both a fiber and a base component. We note a particular exceptional case which occurs when the fivebrane wraps an exceptional divisor in the base. Finally, in sections~7,~8 and~9 we make these methods concrete by presenting three specific examples, two with a del Pezzo base and one with a Hirzebruch base. Two of these examples correspond realistic three-family, non-perturbative vacua in Ho\v rava-Witten theory. \section{Heterotic M~theory vacua with fivebranes} \subsection{Conditions for a supersymmetric background} As we have discussed in the introduction, the standard way to obtain heterotic vacua in four dimensions with $\mathcal{N}=1$ supersymmetry, is to compactify eleven-dimensional M~theory on the manifold $X\tsS^1/Z_2\times M_4$~\cite{hw1,hw2}. Here $X$ is a Calabi--Yau threefold, $S^1/Z_2$ is an orbifold interval, while $M_4$ is four-dimensional Minkowski space. This background is not an exact solution, but is good to first order in an expansion in the eleven-dimensional Planck length. To match to the low-energy particle physics parameters, the Calabi--Yau threefold is chosen to be the size of the Grand Unified scale, while the orbifold is somewhat larger~\cite{w:cy,bd}. In general, there is a moduli space of different compactifications. There are the familiar moduli corresponding to varying the complex structure and the K\"ahler metric on the Calabi--Yau threefold. Similarly, one can vary the size of the orbifold. These parameters all appear as massless fields in the low-energy four-dimensional effective action. In general, there are additional low-energy scalar fields coming from zero-modes of the eleven-dimensional three-form field $C$. The second ingredient required to specify a supersymmetric background is to choose the gauge bundle on the two orbifold planes. In general, one can turn on background gauge fields in the compact Calabi--Yau space. Supersymmetry implies that these fields cannot be arbitrary. Instead, they are required to satisfy the hermitian Yang--Mills equations \begin{equation} F_{ab} = F_{{\bar a}{\bar b}} = 0 \qquad g^{a{\bar b}} F_{a{\bar b}} = 0 \label{hYM} \end{equation} Here $a$ and $b$ are holomorphic indices in $X$, ${\bar a}$ and ${\bar b}$ are antiholomorphic indices, while $g_{a{\bar b}}$ is the K\"ahler metric on $X$. Having fixed the topology of the gauge bundle, that is, how the bundle is patched over the Calabi--Yau manifold, there is then a set of different solutions to these equations. There are additional low-energy moduli which interpolate between these different solutions. In general, the full moduli space of bundles is hard to analyze. However, when the Calabi--Yau threefold is elliptically fibered, the generic structure of this moduli space can be calculated and has been discussed in~\cite{fmw,ron} and also in~\cite{gmod}. The final ingredient to the background is that one can include fivebranes~\cite{w:cy,nse}. In order to preserve supersymmetry and four-dimensional Lorentz invariance, the fivebranes must span the four-dimensional Minkowski space while the remaining two dimensions wrap a holomorphic curve within the Calabi--Yau threefold~\cite{w:cy,bbs,bvs}. In addition, each brane must be parallel to the orbifold fixed planes. Thus it is localized at a single point in $S^1/Z_2$. Again, there are a set of moduli giving the positions of the five-branes within the Calabi--Yau manifold as well as in the orbifold interval. As we will discuss below, there are also extra moduli coming from the self-dual tensor fields on the fivebranes~\cite{nse,w:hw,ketal}. These fields generically give some effective $\mathcal{N}=1$ gauge theory in four-dimensions. Finding the moduli space of the fivebranes, and some information about the effective gauge theory which arises on the fivebrane worldvolumes will be the goal of this paper. In summary, the M~theory background is determined by choosing \begin{itemize} \item a spacetime manifold of the form $X\tsS^1/Z_2 \times M_4$, where $X$ is a Calabi--Yau threefold \item two $E_8$ gauge bundles, $V_1$ and $V_2$, satisfying the hermitian Yang--Mills equations~\eqref{hYM} on $X$ \item a set of fivebranes parallel to the orbifold fixed planes and wrapped on holomorphic curves within $X$. \end{itemize} This ensures that we preserve $\mathcal{N}=1$ supersymmetry in the low-energy four-dimensional effective theory. \subsection{Cohomology condition} The above conditions are not sufficient to ensure that one has a consistent background. Anomaly cancellation on both the ten-dimensional orbifold fixed planes and the six-dimensional fivebranes is possible only because each is a magnetic source for the supergravity four-form field strength $G=dC$~\cite{w:cy}. This provides an inflow mechanism to cancel the anomaly on the lower dimensional space. In general, the magnetic sources for $G$ are five-forms. Explicitly, if $0\leq x^{11}\leq\pi\rho$ parameterizes the orbifold interval, one has~\cite{nse} \begin{equation} dG = J_1 \wedge \delta(x^{11}) + J_2 \wedge \delta(x^{11}-\pi\rho) + \sum_i J_5^{(i)} \wedge \delta(x^{11}-x^{(i)}) \label{Bid} \end{equation} where $J_1$ and $J_2$ are four-form sources on the two fixed planes and $J_5^{(i)}$ is a delta-function four-form source localized at the position of the $i$-th five-brane in $X$. The explicit one-form delta functions give the positions of the orbifold fixed planes at $x^{11}=0$ and $x^{11}=\pi\rho$ and the five-branes at $x^{11}=x^{(i)}$ in $S^1/Z_2$. Compactifying on $X\tsS^1/Z_2$, we have the requirement that the net charge in the internal space must vanish, since there is nowhere for flux to escape. Equivalently, the integral of $dG$ over any five-cycle in $X\tsS^1/Z_2$ must be zero since $dG$ is exact. Integrating over the orbifold interval then implies that the integral of $J_1+J_2+\sum_i J_5^{(i)}$ over any four cycle in $X$ must vanish. Alternatively, this means that the sum of these four-forms must be zero up to an exact form, that is, they must vanish cohomologically. Explicitly, the source on each orbifold plane is proportional to \begin{equation} J_{n} \sim \mathrm{tr} F_n\wedge F_n - \frac{1}{2} \mathrm{tr} R\wedge R \end{equation} where $F_n$ for $n=1,2$ is the $E_8$ field strength on the $n$-th fixed plane, while $R$ is the spacetime curvature. The full cohomology condition can then be written as \begin{equation} \lambda(TX) = w(V_1) + w(V_2) + [W] \label{cohocond} \end{equation} with \begin{equation} \begin{aligned} w(V) &= - \frac{1}{60\cdot 8\pi^2} \mathrm{tr}_{\bf 248} F \wedge F \\ \lambda(TX) &= \frac{1}{2}p_1(TX) = - \frac{1}{2\cdot 8\pi^2} \mathrm{tr}_{\bf 6} R \wedge R \end{aligned} \end{equation} where the right-hand sides of these expressions really represent cohomology classes, rather than the forms themselves. The traces are in the adjoint ${\bf 248}$ of $E_8$ and the vector representation of $SO(6)$. $[W]$ represents the total cohomology class of the five-branes, which we will discuss in a moment. Note that $\lambda$ is half the first Pontrjagin class. It is, in fact, an integer class because we are on a spin manifold. On a Calabi--Yau threefold it is equal to the second Chern class $c_2(TX)$, where the tangent bundle $TX$ is viewed as an $SU(3)$ bundle and the trace is in the fundamental representation. Thus, the cohomology condition simplifies to \begin{equation} [W] = c_2(TX) - w(V_1) - w(V_2) \label{cocond} \end{equation} What do we mean by the cohomology class $[W]$? We recall that we associated four-form delta function sources to the five-branes in $X$. The class $[W]$ is then the cohomology class of the sum of all these sources. Recall that the five-branes wrap on holomorphic curves within the Calabi--Yau threefold. The sum of the five-branes thus represents an integer homology class in $H_2(X,\mathbf{Z})$. In general, one can then use Poincar\'e duality to associate an integral cohomology class in $H^4(X,\mathbf{Z})$ to the homology class of the fivebranes, or also a de Rham class in $H_{\mathrm{DR}}^4(X,\mathbf{R})$. This is the class $[W]$ which enters the cohomology condition, though we will throughout use the same expression $[W]$ for the integral homology class in $H_2(X,\mathbf{Z})$, the integral cohomology class in $H^4(X,\mathbf{Z})$, and the de Rham cohomology class in $H_{\mathrm{DR}}^4(X,\mathbf{R})$. \subsection{Homology classes and effective curves} Let us now turn to analyzing the cohomology condition~\eqref{cocond} in more detail. One finds that the requirement that $[W]$ correspond to the homology class of a set of supersymmetric fivebranes puts a constraint on the allowed bundle classes~\cite{dlow1,dlow2}. Since the sources are all four-forms, equation~\eqref{cocond} is clearly a relation between de~Rahm cohomology classes $H_{\mathrm{DR}}^4(X,\mathbf{R})$. However, in fact, the sources are more restricted than this. In general, they are all in integral cohomology classes. By this we mean that their integral over any four-cycle in the Calabi--Yau threefold gives an integer. (As noted above, this is even true when we no longer have a Calabi--Yau threefold but only a spin manifold, and $c_2(TX)$ is replaced by $\frac{1}{2} p_1(X)$.) The class $[W]$ is integral because it is Poincar\'e dual to an integer sum of fivebranes, an element of $H_2(X,\mathbf{Z})$. Note that there is a general notion of the integer cohomology group $H^p(X,\mathbf{Z})$ which, in general, includes discrete torsion groups such as $\mathbf{Z}_2$. This maps naturally to de~Rahm cohomology $H_{\mathrm{DR}}^p(X,\mathbf{R})$. However, it is important to note that the map is not injective. Torsion elements in $H^p(X,\mathbf{Z})$ are lost. The integral classes to which we refer in this paper are to be identified with the images of $H^p(X,\mathbf{Z})$ in $H_{\mathrm{DR}}^p(X,\mathbf{R})$. In general, $[W]$ cannot be just any integral class. We have seen that supersymmetry implies that fivebranes are wrapped on holomorphic curves within $X$. Thus $[W]$ must correspond to the homology class of holomorphic curves. Furthermore, $[W]$ must refer to some physical collection of fivebranes. Included in $H_2(X,\mathbf{Z})$ are negative classes like $-[C]$ where $C$ is, for example, a holomorphic curve in $X$. These have cohomology representatives which would correspond to the ``absence'' of a five-brane, contributing a negative magnetic charge to the Bianchi identity for $G$ and negative stress-energy. Such states are physically not allowed. The condition that $[W]$ describes physical, holomorphic fivebranes further constrains $c_2(TX)$, $w(V_1)$ and $w(V_2)$ in the cohomology condition~\eqref{cocond}. In order to formalize these constraints, we need to introduce some definitions. We will use the following terminology. \begin{itemize} \item A \textbf{curve} is a holomorphic complex curve in the Calabi--Yau manifold. A curve is \textbf{reducible} if it can be written as the union of two curves. \item A \textbf{class} is a homology class in $H_2(X,\mathbf{Z})$ (or the Poincar\'e dual cohomology class in $H^4(X,\mathbf{Z})$). In general, it may or may not have a representative which is a holomorphic curve. If it does, then a class is \textbf{irreducible} if it has an irreducible representative. Note that there may be other curves in the class which are reducible, but the class is irreducible if there is at least one irreducible representative. \item A class which can be written as a sum of irreducible classes with arbitrary integer coefficients is called \textbf{algebraic}. \item A class is \textbf{effective} if it can be written as the sum of irreducible classes with positive integer coefficients. \end{itemize} Note that we will occasionally use analogous terminology to refer to surfaces (or divisors) in $X$. These are holomorphic complex surfaces in the Calabi--Yau threefold, so they have four real dimensions, and their classes lie in $H_4(X,\mathbf{Z})$. Physically, the above definitions correspond to the following. A curve $W$ describes a collection of supersymmetric fivebranes wrapped on holomorphic two-cycles in the Calabi--Yau space. A reducible curve is the union of two or more separate five-branes. A general class in $H_2(X,\mathbf{Z})$ has representatives which are a general collection of five-branes, perhaps supersymmetric, perhaps not, and maybe including ``negative'' fivebranes of the form mentioned above. An algebraic class, on the other hand, has representatives which are a collection of only five-branes wrapped on holomorphic curves and so supersymmetric, but again includes the possibility of negative fivebranes. Finally, an effective class has representatives which are collections of supersymmetric fivebranes but exclude the possibility of non-physical negative fivebrane states. From these conditions, we see that the constraint on $[W]$ is that we must choose the Calabi--Yau threefold and the gauge bundles $V_1$ and $V_2$ such that \begin{equation} \text{$[W]$ must be effective} \end{equation} As it stands, it is not clear that $[W]=c_2(TX)-w(V_1)-w(V_2)$ is algebraic, let alone effective. However, supersymmetry implies that both the tangent bundle and the gauge bundles are holomorphic. There is then a useful theorem that the classes of holomorphic bundles are algebraic\footnote{This is a familiar result for Chern classes (see~\cite{griffharr}). For $E_8$, or other groups, it can be seen by taking any matrix representation of the group and treating it as a vector bundle, that is, by embedding $E_8$ in $GL(n,\mathbf{C})$. The second Chern class of the vector bundle is then algebraic and is some integer multiple $p$ of the class $w(V)$, where the factor is related to the quadratic Casimir of the representation. We conclude that $w(V)$ is \textit{rationally algebraic}: it is integral, and a further integral multiple of it is algebraic.}, and so $[W]$ is in fact necessarily algebraic. However, there remains the condition that $[W]$ must be effective which does indeed constrain the allowed gauge bundles on a given Calabi--Yau threefold. \subsection{The theory on the fivebranes, $\mathcal{N}=1$ gauge theories and the fivebrane moduli space} While two of the fivebrane dimensions are wrapped on a curve within the Calabi--Yau manifold, the remaining four dimensions span uncompactified Minkowski space. The low-energy massless degrees of freedom on a given fivebrane consequently fall into four-dimensional $\mathcal{N}=1$ multiplets. At a general point in moduli space there are a set of complex moduli $(m_i,\bar{m}_i)$ describing how the fivebrane curve can be deformed within the Calabi--Yau three-fold. These form a set of chiral multiplets. In addition, there is a single real modulus $x^{11}$ describing the position of the fivebrane in the orbifold interval. This is paired under supersymmetry with an axion $a$ which comes from the reduction of the self-dual three-form degree of freedom, $h$, on the fivebrane to form a further chiral multiplet. When the fivebrane is non-singular, that is, does not intersect itself, touch another fivebrane, or pinch, at any point, the remaining degrees of freedom are a set of $U(1)$ gauge multiplets, where the gauge fields also arise from the reduction of the self-dual three-form. The number of $U(1)$ fields is given by the genus $g$ of the curve. In summary, generically, we have \begin{equation} \begin{aligned} \text{chiral multiplets:}& \quad (x^{11},a),\, (m_i,\bar{m}_i) \\ \text{vector multiplets:}& \quad \text{$g$ multiplets with $U(1)^g$ gauge group} \end{aligned} \end{equation} for each distinct fivebrane. When the fivebrane becomes singular, new degrees of freedom can appear. These correspond to membranes stretched between parts of the same fivebrane, or the fivebrane and other fivebranes, which shrink and become massless when the fivebrane becomes singular. They may be new chiral or vector multiplets. In the following, we will not generally identify all the massless degrees of freedom at singular configurations but, rather, concentrate on describing the degrees of freedom on the smooth parts of the moduli space. In conclusion, we have seen that fixing the Calabi--Yau manifold and gauge bundles, in general, fixes an element $[W]$ of $H_2(X,\mathbf{Z})$ describing the homology class of the holomorphic curve in $X$ on which the fivebranes are wrapped. In order to describe an actual set of fivebranes, $[W]$ must be effective, which puts a constraint on the choice of gauge bundles. In general, there are a great many different arrangements of fivebranes in the same homology class. The fivebranes could move about within the Calabi--Yau threefold and also in the orbifold interval. In addition, there can be transitions where branes split and join. The net effect is that there is, in general, a complicated moduli space of five-branes parameterizing all the different possible combinations. In the low-energy effective theory on the fivebranes, the moduli space is described by a set of chiral multiplets. In order to describe the structure of this moduli space, it is clear that we need to analyze the moduli space of all the holomorphic curves in the class $[W]$, including the possibility that each fivebrane can move in $S^1/Z_2$ and can have a different value of the axionic scalar $a$. \section{Elliptically fibered Calabi--Yau manifolds} The moduli spaces we will investigate in detail in this paper are those for five-branes wrapped on smooth elliptically fibered Calabi--Yau threefolds $X$. Consequently, in this section we will briefly summarize the structure of $X$, then identify the generic algebraic classes and finally understand the conditions for these classes to be effective. \subsection{Properties of elliptically fibered Calabi--Yau threefolds} An elliptically fibered Calabi--Yau threefold $X$ consists of a base $B$, which is a complex surface, and an analytic map \begin{equation} \pi : X \to B \label{XtoB} \end{equation} with the property that for a generic point $b \in B$, the fiber $E_{b}=\pi^{-1}(b)$ is an elliptic curve. That is, $E_{b}$ is a Riemann surface of genus one with a particular point, the origin $p$, identified. In particular, we will require that there exist a global section, denoted $\sigma$, defined to be an analytic map \begin{equation} \sigma : B \to X \label{section} \end{equation} that assigns to every point $b\in B$ the origin $\sigma(b)=p\in E_{b}$. We will sometimes refer to this as the zero section. The requirement that the elliptic fibration have a section is crucial for duality to F~theory. However, one notes that from the M~theory point of view it is not necessary. In order to be a Calabi--Yau threefold, the canonical bundle of $X$ must be trivial. From the adjunction formula, this implies that the normal bundle to the section, $N_{B/X}$, which is a line bundle over $B$ and tells us how the elliptic fiber twists as one moves around the base, must be related to the canonical bundle of the base, $K_B$. In fact, \begin{equation} N_{B/X} = K_B. \label{cL} \end{equation} Further conditions appear if one requires that the Calabi--Yau threefold be smooth. The canonical bundle $K_B$ is then constrained so that the only possibilities for the base manifold are as follows~\cite{vm2,grassi}: \begin{itemize} \item for a smooth Calabi--Yau manifold the base $B$ can be a del Pezzo ($dP_r$), Hirzebruch ($F_r$) or Enriques surface, or a blow-up of a Hirzebruch surface. \end{itemize} These are the only possibilities we will consider. The structure of these surfaces is discussed in detail in an appendix to~\cite{dlow2}. In the following, we will adopt the notation used there. It will be useful to recall that, in general, there is a set of points in the base at which the fibration becomes singular. These form a curve, the discriminant curve $\Delta$, which is in the homology class $-12K_B$, as can be shown explicitly by considering the Weierstrass form of the fibration. \subsection{Algebraic classes on $X$} \label{algclasses} Since $[W]$ is algebraic, we need to identify the set of algebraic classes on our elliptically fibered Calabi--Yau manifold. This was discussed in~\cite{dlow1,dlow2}, but here we will be more explicit. It will be useful to identify these classes both in $H_2(X,\mathbf{Z})$ and $H_4(X,\mathbf{Z})$. In general, the full set of classes will depend on the particular fibration in question. However, there is a generic set of classes which are always present, independent of the fibration, and this is what we will concentrate on. Simply because we have an elliptic fibration, the fiber at any given point is a holomorphic curve in $X$. Consequently, one algebraic class in $H_2(X,\mathbf{Z})$ which is always present is the class of the fiber, which we will call $F$. The existence of a section means there is also a holomorphic surface in $X$. Thus the class of the section, which we will call $D$, defines an algebraic class in $H_4(X,\mathbf{Z})$. Some additional algebraic classes may be inherited from the base $B$. In general, $B$ has a set of algebraic classes in $H_2(B,\mathbf{Z})$. One useful fact is that for all the bases which lead to smooth Calabi--Yau manifolds, one finds that every class in $H_2(B,\mathbf{Z})$ is algebraic. This follows from the Lefschetz theorem~\cite{griffharr} which tells us that we can identify algebraic classes on a surface $S$ with the image of integer classes in the Dolbeault cohomology $H^{1,1}(S)$. One then has the following picture. In general, the image of $H^2(S,\mathbf{Z})$ is a lattice of points in $H^2(S,\mathbf{R})$. Choosing a complex structure on $S$ corresponds to fixing an $h^{1,1}$-dimensional subspace within $H^2(S,\mathbf{R})$ describing the space $H^{1,1}(S)$. Generically, no lattice points will intersect the subspace and so there are no algebraic classes on $S$. The exception is when $h^{2,0}=0$, which is the case for all the possible bases $B$. Then the subspace is the whole space $H^2(S,\mathbf{R})$ so all classes in $H^2(S,\mathbf{Z})$ are algebraic. If $\Omega$ is an algebraic class in $H_2(B,\mathbf{Z})$, there are two ways it can lead to a class in $X$. First, one can use the section $\sigma$ to form a class in $H_2(X,\mathbf{Z})$. If $C$ is some representative of $\Omega$, then the inclusion map $\sigma$ gives a curve $\sigma(C)$ in $X$. The homology class of this curve in $H_2(X,\mathbf{Z})$ is denoted by $\sigma_*\Omega$. Second, we can use the projection map $\pi$ to pull $\Omega$ back to a class in $H_4(X,\mathbf{Z})$. For a given representative $C$, one forms the fibered surface $\pi^{-1}(C)$ over $C$. The homology class of this surface in $H_4(X,\mathbf{Z})$ is then denoted by $\pi^*\Omega$. This structure is indicated in Figure~\ref{algclass}. \begin{figure} \centerline{\psfig{figure=algclass.eps,height=3in}} \caption{Generic algebraic classes on $X$, where $[C]=\sigma_*\Omega$ and the fiber class $[E_p]=F$ are in $H_2(X,\mathbf{Z})$, and $[\pi^{-1}(C)]=\pi^*\Omega$ and base class $[B]=D$ are in $H_4(X,\mathbf{Z})$.} \label{algclass} \end{figure} In general, these maps may have kernels. For instance, two curves which are non-homologous in $B$, might be homologous once one embeds them in the full Calabi--Yau threefold. In fact, we will see that this is not the case. One way to show this, which will be useful in the following, is to calculate the intersection numbers between the classes in $H_2$ and $H_4$. We find \begin{equation} \begin{array}{rrccc} & & &\multicolumn{2}{c}{H_2(X,\mathbf{Z})} \\ & & \vline & \sigma_*\Omega' & F \\ \cline{2-5} H_4(X,\mathbf{Z}) & \begin{array}{c} \pi^*\Omega \\ D \end{array} & \vline & \begin{array}{c} \Omega\cdot \Omega' \\ K_B \cdot \Omega' \end{array} & \begin{array}{c} 0 \\ 1 \end{array} \end{array} \label{inters} \end{equation} where the entries in the first column are the intersections of classes in $B$. The intersection of $\sigma_*\Omega'$ with $D$ is derived by adjunction, recalling that the normal bundle to $B$ is $N_{B/X}=K_B^{-1}$. Two classes are equivalent if they have the same intersection numbers. If we take a set of classes $\Omega_i$ which from a basis of $H_2(B,\mathbf{Z})$, we see that the matrix of intersection numbers of the form given in~\eqref{inters} is non-degenerate. Thus, for each nonzero $\Omega\in H_2(B,\mathbf{Z})$, we get nonzero classes $\sigma_*\Omega$ and $\pi^*\Omega$ in $H_2(X,\mathbf{Z})$ and $H_4(X,\mathbf{Z})$. As we mentioned, the algebraic classes we have identified so far are generic, always present independently of the exact form of the fibration. There are two obvious sources of additional classes. Consider $H_4(X,\mathbf{Z})$. First, we could have additional sections non-homologous to the zero-section $\sigma$. Second, the pull-backs of irreducible classes on $B$ could split so that $\pi^*(\Omega)=\Sigma_1+\Sigma_2$. This splitting comes from the fact that there can be curves on the base over which the elliptic curve degenerates, for example, into a pair of spheres. New classes appear from wrapping the four-cycle over either one sphere or the other. Now consider $H_2(X,\mathbf{Z})$. We see that the possibility of degeneration of the fiber means that the fiber class $F$ can similarly split, with representatives wrapped, for instance, on one sphere or the other. Finally, the presence of new sections means there is a new way to map curves from $B$ into $X$ and, in general, classes in $H_2(B,\mathbf{Z})$ will map under the new section to new classes in $H_2(X,\mathbf{Z})$. In all our discussions in this paper, we will ignore these additional classes. This will mean that our moduli space discussion is not in general complete. However, this restriction will allow us to analyze generic properties of the moduli space. In summary, we have identified the generic algebraic classes $\Omega$ in $H_2(X,\mathbf{Z})$ as classes in $B$ (since these are all algebraic for the bases in question) mapped via the section into $X$, together with the fiber class $F$; while in $H_4(X,\mathbf{Z})$ the generic algebraic classes are the pull-backs $\pi^*(\Omega)$ of classes in $B$, together with the class $D$ of the section. Furthermore, distinct algebraic classes in $B$ lead to distinct algebraic classes in $X$. \subsection{Effective classes on $X$} We argued in the previous section that a generic algebraic class $[W]$ in $H_2(X,\mathbf{Z})$ on a general $X$ can be written as \begin{equation} [W] = \sigma_*\Omega + f F \label{Wdecomp} \end{equation} where $\Omega$ is an algebraic class in $B$ (which is then mapped to $X$ via the section) and $F$ is the fiber class, while $f$ is some integer. If $[W]$ is to be the class of a set of five-branes it must be effective. What are the conditions, then, on $\Omega$ and $f$ such that $[W]$ is effective? We showed in~\cite{dlow2} that the following is true. First, for a base which is any del Pezzo or Enriques surface, $[W]$ is effective if and only if $\Omega$ is an effective class in $B$ and $f\geq 0$. Second, this is also true for a Hirzebruch surface $F_r$, with the exception of when $\Omega$ happens to contain the negative section $\mathcal{S}$ and $r\geq3$. Here, following the notation of~\cite{dlow2}, we write a basis of algebraic classes on $F_r$ as the negative section $\mathcal{S}$ and the fiber $\mathcal{E}$. In this case, there is a single additional irreducible class $\sigma_*\mathcal{S}-(r-2)\sigma_*\mathcal{E}$. In this paper, for simplicity, we will not consider these exceptional cases for which the statement is untrue. Thus, under this restriction, we have that \begin{equation} W = \sigma_*\Omega + f F \text{ is effective in } X \Longleftrightarrow \Omega \text{ is effective in } B \text{ and } f\geq 0 \label{thrm} \end{equation} This reduces the question of finding the effective curves in $X$ to knowing the generating set of effective curves in the base $B$. For the set of base surfaces $B$ we are considering, finding such generators is always possible (see for instance~\cite{dlow2}). The derivation of this result goes as follows. Clearly, if $\Omega$ is effective in $B$ and $f$ is non-negative then since effective curves in $B$ must map under the section to effective curves in $X$, we can conclude that $[W]$ is effective. One can also prove that the converse is true in almost all cases. One sees this as follows. First, unless a curve is purely in the fiber, in which case $\Omega=0$, the fact that $X$ is elliptically fibered means that all curves $W$ project to curves in the base. The class $[W]$ similarly projects to the class $\Omega$. The projection of an effective class must be effective, thus if $[W]$ is effective in $X$ then so is $\Omega$ in $B$. The only question then is whether there are effective, irreducible, curves in $X$ with negative $f$. To address this, we use the fact that any irreducible class in $H_2(X,\mathbf{Z})$ must have non-negative intersection with any effective class in $H_4(X,\mathbf{Z})$ unless all the representative curves are contained within the representative surfaces. We start by noting that if $\Omega$ is an effective class in $B$ then $\pi^*\Omega$ must be an effective class in $H_4(X,\mathbf{Z})$. This can be seen by considering any given representative of $B$ and its inverse image in $X$. From the intersection numbers given in~\eqref{inters} and the generic form of $[W]$~\eqref{Wdecomp} we see that, if $\Omega'$ is an effective class in $B$, then \begin{equation} \begin{aligned} \pi^*\Omega' \cdot [W] &= \Omega' \cdot \Omega \\ D \cdot [W] &= K_B \cdot \Omega + f \end{aligned} \label{effinters} \end{equation} From the first intersection one simply deduces again that if $[W]$ is effective then so is $\Omega$. Now suppose that $f$ is non-zero. Then $W$ cannot be contained within $B$ and so, from the second expression, we have $f\geq -K_B\cdot \Omega$. We recall that for del Pezzo and Enriques surfaces $-K_B$ is nef, so that its intersection with any effective class $\Omega$ is non-negative. Thus, we must have $f\geq 0$ for $[W]$ to be effective. The exception is a Hirzebruch surface $F_r$ for $r\geq 3$. We then have $-K_B\cdot \mathcal{E}=2>0$ but $-K_B\cdot \mathcal{S}=2-r<0$. This allows the existence of effective classes of the form $\sigma_*\mathcal{S}+fF$ with $f$ negative. Indeed, the existence of such a class can be seen as follows. Consider a representative curve $C$ of $\mathcal{S}$ in $F_r$ with $r\geq 3$ (in fact, the representative is unique). It is easy to see that $C$ is topologically $\mathbf{P}^1$ (see equation~\eqref{genuscS} below). The surface $\pi^{-1}(C)$ above $C$ should thus be an elliptic fibration over $\mathbf{P}^1$. However, as shown in equation~\eqref{pcS} below, in fact, $C$ is contained within the discriminant curve $\Delta$ of the Calabi--Yau fibration. Thus all the fibers over $C$ are singular. The generic singular fiber is a $\mathbf{P}^1$, suggesting that $S_C$ is a $\mathbf{P}^1$ fibration over $\mathbf{P}^1$. In fact it can be shown that $S_C$ is indeed itself the Hirzebruch surface $F_{r-2}$ (or a blow-up of such a surface). What class is our original curve $C$ in the new surface $F_{r-2}$? If we write the classes of $F_{r-2}$ as $\mathcal{S}'$ and $\mathcal{E}'$, we identify $\mathcal{E}'=F$ since this is just the fiber class of the $F_{r-2}$. In addition, one can show that $\mathcal{S}=\mathcal{S}'+(r-2)\mathcal{E}'$. However, we know that $\mathcal{S}'$ itself is an irreducible class, so $\mathcal{S}'=\mathcal{S}+(2-r)F$ is irreducible in $H_2(X,\mathbf{Z})$. Thus we see there is one new irreducible class with negative $f$ which saturates the condition that $f\geq -(r-2)$. \section{The moduli space for fivebranes wrapping the elliptic fiber and the role of the axion} Probably the simplest example of a fivebrane moduli space is the case where the fivebranes wrap only the elliptic fiber of the Calabi--Yau threefold. By way of introduction to calculating moduli spaces, in this section, we will consider this case, first for a single fivebrane and then for a collection of fivebranes. These configurations are well understood in the dual F-theory picture as collections of D3-branes~\cite{fmw}. We end the section with a discussion of the connection between our results and the F-theory description. \subsection{$[W]=F$} If it wraps a fiber only once, the class of the fivebrane curve is simply given by \begin{equation} [W] = F \end{equation} A fivebrane wrapping any of the elliptic fibers will be in this class. One might imagine that there are other fivebranes in this class, where not all the fivebrane lies at the same point in the Calabi--Yau threefold. Instead, as one moves along the fivebrane in the fiber direction, the fivebrane could have a component in the base directions. However, if the curve is to be holomorphic, every point in the fivebrane curve must lie over the same point in the base. Similarly, in order to preserve $\mathcal{N}=1$ supersymmetry, the brane must be parallel to the orbifold fixed planes, so it is also at a fixed point in the orbifold. Since these position moduli are independent, the moduli space appears to be $B\tsS^1/Z_2$. The two complex coordinates on $B$ form a pair of chiral $\mathcal{N}=1$ superfields. The metric on this part of the moduli space should simply come from the K\"ahler metric on the base $B$. However, we have, thus far, ignored the axionic scalar, $a$, on the fivebrane world volume. We have argued that this is in a chiral multiplet with the orbifold modulus $x^{11}$. Furthermore, it is compact, describing an $S^1$. However, at the edges of the orbifold this changes. It has been argued in~\cite{gh,sw} that there is a transition when a fivebrane reaches the boundary. At the boundary, the brane can be described by a point-like $E_8$ instanton. New low-energy fields then appear corresponding to moving in the instanton moduli space. Similarly, some of the fivebrane moduli disappear. Throughout this transition the low-energy theory remains $\mathcal{N}=1$. Thus, since the $x^{11}$ degree of freedom disappears in the transition, so must the axionic degree of freedom. Consequently the axionic $S^1$ moduli space must collapse to a point at the boundary. We see that the full $(x^{11},a)$ moduli space is just the fibration of $S^1$ over the interval $S^1/Z_2$, where the $S^1$ is singular at the boundaries, that is, the orbifold and axion part of the moduli space is simply $S^2=\mathbf{P}^1$. The fact that the axionic degree of freedom disappears on the boundary can be seen in another way. In the fivebrane equation of motion, one can write the self-dual three-form field strength $h$ in terms of a two-form potential, $b$, in combination with the pull-back onto the fivebrane worldvolume of the eleven-dimensional three-form potential $C$ as~\cite{fbraneeom} \begin{equation} h = db - C \label{hdef} \end{equation} Under the $Z_2$ orbifold symmetry $C$ is odd unless it has a component in the direction of the orbifold. Since the fivebrane must be parallel to the orbifold fixed-planes this is not the case. This implies that $h$ must also be odd. Consequently, $h$ must be zero on the orbifold fixed planes implying that the axion $a$ also disappears on the boundary. In summary, the full moduli space is given locally by \begin{equation} \mathcal{M}(F) = B \times \PP^1_a. \label{Fmod} \end{equation} where the subscript on $\PP^1_a$ denotes that this part of moduli space describes the axion multiplet. Globally, this $\mathbf{P}^1$ could twist as we move in $B$; so $\mathcal{M}(F)$ is really a $\mathbf{P}^1$ bundle over $B$. We will return to this point below. What about the vector degrees of freedom? Since the fiber is elliptic, the fivebrane curve must be topologically a torus. Thus we have \begin{equation} g=\genus{W}=1 \end{equation} and there is a single $U(1)$ vector multiplet in the low-energy theory. \subsection{$[W]=nF$} \label{sec:fF} The generalization to the case where the fivebrane class is a number of elliptic fibers is straightforward. The class \begin{equation} [W] = fF \end{equation} where $f\geq 1$, means we have a collection of curves which wrap the fiber $f$ times. In general, we could have one component which wraps $f$ times or two or more components each wrapping a fewer number of times. In the limiting case, there are $f$ distinct components each wrapping only once. A single component must wrap all at the same point in the base. In addition, it must be at a fixed point in the orbifold interval and must have a single value of the axionic scalar. Two or more distinct components can wrap at different points in the base and have different values of $x^{11}$ and $a$. As homology cycles, there is no distinction between the case where some number $n$ of singly wrapped components overlap, lying at the same point in the base, and the case where there is a single component wrapping $n$ times. Both cases represent the same two-cycle in the Calabi--Yau manifold. Physically, they could be distinguished if the $n$ singly wrapped components were at different points in $S^1/Z_2$ or had different values of the axion. However, if the values of $x^{11}$ and $a$ were also the same, by analogy with D~branes, we would expect that we could not then distinguish, in terms of the scalar fields on the branes, the $n$ singly wrapped fivebranes from a single brane wrapped $n$ times. From the discussion in the last section, each singly wrapped fivebrane has a moduli space given locally by $B\times\PP^1_a$. Thus for $f$ components, we expect the full scalar field moduli space locally has the form \begin{equation} \mathcal{M}(fF) = \left(B\times\PP^1_a\right)^f/\mathbf{Z}_f \label{fFmod} \end{equation} where we have divided out by permutations since the fivebranes are indistinguishable. The ambiguous points in moduli space, which could correspond to a number of singly wrapped fivebranes or a single multiply wrapped fivebrane, are then the places where two or more of the points in the $f$ $B\times\PP^1_a$ factors coincide. Note, in addition, that this is again only the local structure of $\mathcal{M}(fF)$. We do not know how the $\PP^1_a$ factors twist as we move the fivebranes in the base. Thus, globally, $\mathcal{M}(fF)$ is the quotient of a $(\mathbf{P}^1)^f$ bundle over $B^f$. In a similar way, the gauge symmetry on the fivebranes also follows by analogy with D-branes. At a general point in the moduli space, we have $n$ distinct fivebranes each wrapping a torus and so, as in the previous section, each with a single $U(1)$ gauge field. When two branes collide in the Calabi--Yau threefold, and are at the same point in the orbifold and have the same value of the axion, we expect the symmetry enhances to $U(2)$. The new massless states come from membranes stretched between the fivebranes. The maximal enhancement is when all the fivebranes collide and the group becomes $U(n)$. \subsection{Duality to F~theory and twisting the axion} \label{axiontwist} The results of the last two sections are extremely natural from the F~theory point of view. It has been argued that fivebranes wrapping an elliptic fiber of $X$ correspond to threebranes spanning the flat $M_4$ space on the type IIB side~\cite{fmw}. To understand the correspondence, we first very briefly review the relation between M and F~theory~\cite{vafa,vm1,vm2}. The duality states that the heterotic string on an elliptically fibered Calabi--Yau threefold $X$ is dual to F~theory on a Calabi-Yau fourfold $X'$ fibered by K3 over the same base $B$. The M~theory limit of the heterotic string we consider here is consequently also dual to the same F~theory configuration. In addition, the duality requires that the K3 fibers should themselves be elliptically fibered. This means that the fourfold $X'$ also has a description as an elliptic fibration over a threefold base $B'$. Since the base of an elliptically fibered K3 manifold is simply $\mathbf{P}^1$, this implies that $B'$ must be a $\mathbf{P}^1$ fibration over $B$. As a type IIB background, the spacetime is $B'\times M_4$, where $M_4$ is flat Minkowski space. The complex structure of the elliptic fibers of $X'$ then encode how the IIB scalar doublet, the dilaton and the Ramond-Ramond scalar, vary as one moves over the ten-dimensional manifold $B'\times M_4$. As such, they describe some configuration of seven-branes in type IIB. M~theory fivebranes which wrap the elliptic fiber in $X$, map to threebranes spanning $M_4$ in the dual F~theory vacuum. As such, the three-brane is free to move in the remaining six compact dimensions. Thus we expect that the threebrane moduli space is simply $B'$. However, we have noted that $B'$ is a $\mathbf{P}^1$ fibration over $B$. Thus we see that locally the moduli space as calculated on the F~theory side exactly coincides with the moduli space of the fivebrane given in~\eqref{Fmod} above. The $\mathbf{P}^1$ fiber in $B'$ is precisely the orbifold coordinate $x^{11}$ together with the axion $a$. For a collection of $f$ threebranes, we expect the moduli space is simply promoted to the symmetric product ${B'}^f/Z_f$. Again, locally, this agrees with the moduli space~\eqref{fFmod} of the corresponding M theory fivebranes. Similarly, it is well known that a threebrane carries a single $U(1)$ gauge field, as does the M~theory fivebrane. For a collection of $f$ threebranes this is promoted to $U(f)$, which was really the motivation for our claim for the vector multiplet structure calculated in the M~theory picture. In general, the arguments given in the previous two sections were only sufficient to give the local structure of the axion multiplet part of the fivebrane moduli space. We did not determine how the axion fiber $\PP^1_a$ twisted as one moved the fivebrane in the Calabi--Yau manifold. From duality with F~theory, we have seen that, in general, we expect this twisting is non-trivial. In fact, it can also be calculated from the M~theory side. We will not give the details here but simply comment on the mechanism. A full description will be given elsewhere~\cite{atwist}. The key is to recall that the self-dual three-form on the fivebrane~\eqref{hdef} depends on the pull-back of the supergravity three-form potential $C$. This leads to holonomy for the axion degree of freedom as one moves the fivebrane within the Calabi--Yau threefold. The holonomy can be non-trivial if the field strength $G$ is non-trivial. However, from the modified Bianchi identity~\eqref{Bid}, we see that this is precisely the case when there are non-zero sources from the boundaries of $S^1/Z_2$ and also from the fivebranes in the bulk. In general, one can calculate how the axion twists, and hence how $\PP^1_a$ twists, in terms of the different sources. This phenomenon is interesting but not central to the structure of the fivebrane moduli spaces, such as the dimension of the space, how its different branches intersect, or waht is the the genus of the fivebrane curve. Thus, for simplicity, in the rest of this paper we will ignore the issue of how $a$ twists as one moves a given collection of fivebranes within the Calabi--Yau manifold. Consequently, the moduli spaces we quote will strictly only be locally correct for the axion degrees of freedom. So that it is clear where the extra global structure can appear, we will always label the $\mathbf{P}^1$ degrees of freedom associated with the axions as $\PP^1_a$. \section{Two examples with fivebranes wrapping curves in the base} The discussion of the moduli space becomes somewhat more complicated once one includes classes where the fivebrane wraps a curve in the base manifold. Again, we will take two simple examples to illustrate the type of analysis one uses. In both cases, we will assume, for specificity, that the base manifold is a $dP_8$ surface, though the methods of our analysis would apply to any base $B$. Throughout, we will use the notation and results of~\cite{dlow2}. A $dP_8$ surface is a $\mathbf{P}^2$ surface blown up at eight points, $p_1,\dots,p_8$. In general there are nine algebraic classes in the base: the class $l$ inherited from the class of lines in the $\mathbf{P}^2$ and the eight classes of the blown-up points $E_1,\dots,E_8$. In the following, we will often describe curves in $dP_8$ in terms of the corresponding plane curve in $\mathbf{P}^2$. \subsection{$[W]=\sigma_*l-\sigma_*E_1$} \label{exbase1} We first take an example where the fivebrane class includes no fiber components \begin{equation} [W] = \sigma_*l - \sigma_*E_1 \end{equation} where $\sigma_*l$ and $\sigma_*E_1$ are the images in the Calabi--Yau manifold of the corresponding classes in the base. Since $\Omega=l-E_1$ is an effective class in the base, (see~\cite{dlow2}), from~\eqref{thrm} we see that $[W]$ is effective in $X$, as required. If we knew that the curve lay only in the base, the moduli space would then simply be the space of curves in a $dP_8$ surface in the class $l-E_1$, which is relatively easy to calculate. In general, however, $W$ lies somewhere in the full Calabi--Yau threefold. The fact that its homology class is the image of a homology class in the base does not imply that $W$ is stuck in $B$. Nonetheless, we do know that, under the projection map $\pi$ from $X$ to $B$, the curve $W$ must project onto a curve $C$ in the base as shown in Figure~\ref{piC}. Furthermore the class of $C$ must be $\Omega=l-E_1$ in $B$. What we can do is find the moduli space of such curves $C$ in the base and then ask, for each such $C$, what set of curves $W$ in the full Calabi--Yau manifold would project onto $C$. That is to say, the full moduli space should have a fibered structure. The base of this space will be the moduli space of curves $C$ in $B$, while the fiber above a given curve is the class of $W$ in $X$ which projects onto the given $C$. \begin{figure} \centerline{\psfig{figure=piC.eps,height=3in}} \caption{The curve $W$ and its image in the base $C$} \label{piC} \end{figure} In our example, the moduli space of curves $C$ in the class $\Omega=l-E_1$ is relatively easy to analyze. In $\mathbf{P}^2$, $\Omega$ describes the class of lines through one point, $p_1$. A generic line in $\mathbf{P}^2$ is a homogeneous polynomial of degree one, \begin{equation} ax + by + cz = 0 \end{equation} where $[x,y,z]$ are homogeneous coordinates on $\mathbf{P}^2$. Since the overall coefficient is irrelevant, a given line is fixed by giving $[a,b,c]$ up to an overall scaling. Thus the moduli space of lines is itself $\mathbf{P}^2$. Furthermore, we see that a given point in the line is specified by fixing, for instance, $x$ and $y$ up to an overall scaling. Consequently, we see that, topologically, a line in $\mathbf{P}^2$ is just a sphere $\mathbf{P}^1$. For the class $\Omega$, we further require that the line pass through a given point $p_1=[x_1,y_1,z_1]$. This provides a single linear constraint on $a$, $b$ and $c$, \begin{equation} ax_1 + by_1 + cz_1 = 0 \end{equation} We now have only the set of lines radiating from $p_1$ and the moduli space is reduced to $\mathbf{P}^1$. Topologically, the line in $\mathbf{P}^2$ is still just a sphere and, generically, its image in $dP_8$ will also be a sphere. There are, however, seven special points in the moduli space. A general line passing through $p_1$ will not intersect any other blown-up point. However, there are seven special lines radiating from $p_1$ which also pass through a second blown-up point. (To be a $dP_8$ manifold, the eight blow-up points must be in general position, so no three are ever in a line.) This is shown in Figure~\ref{lE1}. Let us consider one of these seven lines, say the one which passes through $p_2$. The transform of such a line to $dP_8$ splits into two curves \begin{equation} C = C_1 + C_2 \end{equation} The first component $C_1$ projects back to the line in $\mathbf{P}^2$. The second component corresponds to a curve wrapping the blown up $\mathbf{P}^1$ at $p_2$ and so has no analog in $\mathbf{P}^2$. Specifically, the classes of the two curves are \begin{equation} [C_1] = l - E_1 - E_2, \qquad [C_2] = E_2 \label{Csplit} \end{equation} Using the results in the Appendix to~\cite{dlow2}, we see that \begin{equation} [C_1] \cdot [C_1] = [C_2] \cdot [C_2] = -1 \end{equation} It follows that both curves are in exceptional classes in $dP_8$ and so cannot be deformed within the base. Hence, no new moduli for moving in the base appear when the curve splits. \begin{figure} \centerline{\psfig{figure=lE1.eps,height=2.5in}} \caption{The moduli space of lines in the class $l-E_1$. The solid lines represent two examples of the special case where the proper transform of the line splits into two components} \label{lE1} \end{figure} From the form of~\eqref{Csplit}, we see that, when the curve splits, $C_1$ remains a line in $\mathbf{P}^2$ so is topologically still a sphere, while $C_2$ wraps the blown up $\mathbf{P}^1$ and, so, is also topologically a sphere. Furthermore, the intersection number \begin{equation} [C_1]\cdot [C_2]=1 \end{equation} implies that the two spheres intersect at one point. What has happened is that the single sphere $C$ has pinched off into a pair of spheres as shown in Figure~\ref{Cpinch}. In summary, for the moduli space of curves $C$ in the base, in the homology class $\Omega=l-E_1$, we have, \begin{equation} \begin{array}{c|cc} [C] & \text{genus} & \text{moduli space} \\ \hline l-E_1 & 0 & \mathbf{P}^1 - 7\text{ pts.} \\ (l-E_1-E_i) + (E_i) & 0+0 & \text{single pt.} \end{array} \label{Cmodspace} \end{equation} where in the second line $i=2,\dots,8$. \begin{figure} \centerline{\psfig{figure=Cpinch.eps,height=1in}} \caption{Splitting a single sphere into a pair of spheres} \label{Cpinch} \end{figure} The next step is to find, for a given curve $C$, how many curves $W$ there are in the full Calabi--Yau space which project onto $C$. Furthermore, $W$ must be in the homology class $\sigma_*l-\sigma_*E_1$. Let us start with a curve $C$ at a generic point in the moduli space~\eqref{Cmodspace}, that is, a point in the first line of the table where the curve has not split. Any curve $W$ which projects onto $C$ must lie somewhere in the space of the elliptic fibration over $C$. Thus, we are interested in studying the complex surface \begin{equation} S_C = \pi^{-1}(C) \end{equation} This structure is shown in Figure~\ref{piC}. By definition, this surface is an elliptic fibration over $C$ which, means it is a fibration over $\mathbf{P}^1$. In general, the surface will have some number of singular fibers. This is equal to the intersection number between the discriminant curve $\Delta$, which gives the position of all the singular fibers on $B$, and the base curve $C$. Recall that $[\Delta]=-12K_B$. Using the results summarized in the Appendix to~\cite{dlow2}, since the base is a $dP_8$ surface and intersection numbers depend only on homology classes, we have \begin{equation} [\Delta] \cdot [C] = 12\left(3l-E_1-\dots-E_8\right) \cdot \left(l-E_1\right) = 24 \end{equation} Thus we see that, generically, $S_C$ is an elliptic fibration over $\mathbf{P}^1$ with 24 singular fibers. This implies ~\cite{griffharr} that \begin{equation} S_C \text{ is a K3 surface} \end{equation} The curve $C$ is the zero section of the fibration. Further, projection gives us a map from the actual curve $W$ to its image $C$ in the base. The projection only wraps $C$ once, so, since $C$ is not singular, the map is invertible and $W$ must also be a section of $S_C$. Our question then simplifies to asking: what is the moduli space of sections of $S_C$ in the class $\sigma_*l-\sigma_*E_1$? To answer this question, we start by identifying the algebraic classes in $S_C$. We know that we have at least two classes inherited from the Calabi--Yau threefold: the class of the zero section $C$, which we write as $D_C$, and the class of the elliptic fiber, $F_C$. Specifically, under the inclusion map \begin{equation} i_C : S_C \to X \end{equation} $D_C$ and $F_C$ map into the corresponding classes in $X$ \begin{equation} \begin{aligned} i_{C*}D_C &= [C] = \sigma_*l - \sigma_*E_1 \\ i_{C*}F_C &= F \end{aligned} \label{K3inclmap} \end{equation} where $i_{C*}$ is the map between classes \begin{equation} i_{C*} : H_2(S_C,\mathbf{Z}) \to H_2(X,\mathbf{Z}) \end{equation} These are the only relevant generic classes in $X$. However, there may be additional classes on $S_C$ which map to the same class in $X$ so that the map $i_{C*}$ has a kernel. That is to say, two curves which are homologous in $X$ may not be homologous in $S_C$. However, we note that a generic K3 surface would have no algebraic classes since $h^{2,0}\neq 0$ (see the discussion in section~\ref{algclasses}). Given that in our case of an elliptically fibered K3 with section we have at least two algebraic classes, the choice of complex structure on $S_C$ cannot be completely general. However, generically, we have no reason to believe that there are any further algebraic classes. For particular choices of complex structure additional classes may appear but, since here we are considering the generic properties of the moduli space, we will ignore this possibility. Now, we require that $W$, like $C$, is also in the class $\sigma_*l-\sigma_*E_1$ in the full Calabi--Yau space. This immediately implies, given the map~\eqref{K3inclmap}, that $W$ is also in the class $D_C$ of the zero section $C$ in $S_C$. Furthermore, we can calculate the self-intersection number of this class within $S_C$. This can be done as follows. Recall that the Riemann--Hurwitz formula~\cite{griffharr} applied to the curve $C$ states that \begin{equation} 2g - 2 = \deg K_C \label{RHformula} \end{equation} where $g$ is the genus and $K_C$ is cohomology class of the canonical bundle of $C$. The adjunction formula~\cite{griffharr} then gives \begin{equation} \deg K_C = \left(K_{S_C} + D_C \right) \cdot D_C \label{adj} \end{equation} where $K_{S_C}$ is the canonical class of the K3 surface $S_C$. Using the fact that the canonical class of a K3 surface is zero, $K_{S_C}=0$, and that $C$ is a sphere so $g=0$, it follows from~\eqref{RHformula} and~\eqref{adj} that \begin{equation} D_C \cdot D_C = - 2 \label{Omegasq1} \end{equation} This implies that the section cannot be deformed at all within the surface $S_C$. In conclusion, we see that there is, generically, no moduli space of curves $W$ which project onto $C$. Rather, the only curve in $S_C$ in the class $\sigma_*l-\sigma_*E_1$ is the section $C$ itself. We see that, generically, the curve $W$ can only move in the base of $X$ and cannot be deformed in a fiber direction. Recall that a fivebrane wrapped on $W$ also has a modulus describing its position in $S^1/Z_2$, as well as the axionic modulus. Together, as was discussed in the previous section, these form a $\mathbf{P}^1$ moduli space. Thus, we conclude that the moduli space associated with a generic curve in~\eqref{Cmodspace} is locally simply \begin{equation} \mathcal{M}_{\text{generic}} = \left(\mathbf{P}^1 - \text{7 pts.}\right) \times \PP^1_a \label{genericmod} \end{equation} As discussed above, since the axion can be twisted, globally, this extends to a $\PP^1_a$ bundle over $\mathbf{P}^1$. Physically, we have a single fivebrane wrapping an irreducible curve in the Calabi--Yau threefold, which lies entirely within the base $B$. The curve can be deformed in the base, which gives the first factor in the moduli space, but cannot be deformed in the fiber direction. It can also move in the orbifold interval and have different values for the axionic modulus, which gives the second factor in~\eqref{genericmod}. Since the curve has genus zero, there are no vector fields in the low-energy theory. Thus far we have discussed the generic part of the moduli space. The full moduli space will have the form \begin{equation} \mathcal{M}(\sigma_*l-\sigma_*E_1) = \left(\mathbf{P}^1 - \text{7 pts.}\right) \times \PP^1_a \cup \mathcal{M}_{\text{non-generic}} \end{equation} where the additional piece $\mathcal{M}_{\text{non-generic}}$ describes the moduli space at each of the special 7 points where the curve $C$ splits into two components. To analyze this part of the moduli space, we must consider each component separately, but we can use the same procedure we used above. The fact that the image $C$ splits, means that the original curve $W$ must also split in $X$ \begin{equation} W = W_1 + W_2 \label{simplesplit} \end{equation} with $C_1$ being the projection onto the base of $W_1$ and $C_2$ the projection of $W_2$. Let us consider the case where the line in $\mathbf{P}^2$ also intersects $p_2$. Then the homology classes of $W_1$ and $W_2$ must split as \begin{equation} [W_1] = \sigma_*l - \sigma_*E_1 - \sigma_*E_2 \qquad [W_2] = \sigma_*E_2 \label{Wsplit} \end{equation} Note that one might imagine adding $nF$ to $[W_1]$ and $-nF$ to $[W_2]$, still leaving the total $[W]$ unchanged and having the correct projection onto the base. However, from~\eqref{thrm}, we see that one class would not then be effective and so, since $W_1$ and $W_2$ must each correspond to a physical fivebrane, such a splitting is not allowed. If we start with $W_1$, to find the curves in $X$ which project onto $C_1$ and are in the homology class $\sigma_*l-\sigma_*E_1-\sigma_*E_2$, we begin, as above, with the surface $S_{C_1}=\pi^{-1}(C_1)$ above $C_1$. Calculating the number of singular fibers, we find \begin{equation} \Delta \cdot [C_1] = 12\left(3l-E_1-\dots-E_8\right) \cdot \left(l-E_1-E_2\right) = 12 \end{equation} Since $C_1$ is a sphere, we have an elliptic fibration over $\mathbf{P}^1$ with 12 singular fibers, which implies that~\cite{griffharr} \begin{equation} S_{C_1} \text{ is a $dP_9$ surface} \end{equation} Similarly, if we consider $[C_2]=E_2$, \begin{equation} \Delta \cdot [C_2] = 12\left(3l-E_1-\dots-E_8\right) \cdot E_2 = 12 \end{equation} Since the curve $C_2$ is also a sphere, it follows that we again have an elliptic fibration over $\mathbf{P}^1$ with 12 singular fibers, and hence we also have \begin{equation} S_{C_2} \text{ is a $dP_9$ surface} \end{equation} Thus, we are considering the degeneration of the K3 surface $S_C$, which had 24 singular fibers, into a pair of $dP_9$ surfaces $S_{C_1}$ and $S_{C_2}$, each with 12 singular fibers. On a given $dP_9$ surface, say $S_{C_1}$, we are guaranteed, as for the K3 surface, that there are at least two algebraic classes, the section class $D_{C_1}$ and the fiber class $F_{C_1}$. However, the $dP_9$ case is more interesting than the case of a K3 surface since there are always other additional algebraic classes. On a $dP_9$ surface, $h^{2,0}=0$. Consequently, as was discussed in section~\ref{algclasses}, whatever complex structure one chooses, all classes in $H_2(dP_9,\mathbf{Z})$ are algebraic. Thus, one finds that the algebraic classes on $dP_9$ form a 10-dimensional lattice. Since there are only two distinguished classes on the Calabi--Yau threefold (namely $\sigma_*l-\sigma_*E_1$ and the fiber class $F$), this implies that distinct classes in $S_{C_1}$ must map to the same class in $X$. That is to say, curves which are not homologous in $S_{C_1}$ are homologous once one considers the full threefold $X$. The full analysis of the extra classes on $S_{C_1}$ will be considered in section~\ref{sec:dP9}. In our particular case, it will turn out that, for $W_1$ to be in the same class as $C_1$ in the full Calabi--Yau threefold, it must also be in the same class within $S_{C_1}$. Thus, we are again interested in the moduli space of the section class $D_{C_1}$ in $S_{C_1}$. Now, we recall (see, for instance, the Appendix to~\cite{dlow2}) that the canonical class for a $dP_9$ is simply $K_{S_{C_1}}=-F_D$. So by the analogous calculation to~\eqref{Omegasq1}, using the fact that $D_{C_1}\cdot F_D=1$ since $C_1$ is a section, we have that \begin{equation} D_{C_1} \cdot D_{C_1} = -1 \label{dP9sec} \end{equation} This means that the curve $C_1$ cannot be deformed within $S_{C_1}$. Thus, as in the K3 case, the only possible $W_1$ is the section $C_1$ itself. An identical calculation goes through for the other component $W_2$. Furthermore, the analysis is the same at each of the other six exceptional points in moduli space. Given that the curve has split into two components at each of these points, we have two separate moduli describing the position of each component in $S^1/Z_2$ as well as two moduli describing the axionic degree of freedom for each component. It follows that \begin{equation} \mathcal{M}_{\text{non-generic}} = 7\left(\PP^1_a \times \PP^1_a\right) \label{nongeneric} \end{equation} where $7(\mathbf{P}^1\times\mathbf{P}^1)=\mathbf{P}^1\times\mathbf{P}^1\cup\dots\cup\mathbf{P}^1\times\mathbf{P}^1$. We find, then, that the full moduli space has a branched structure, \begin{equation} \mathcal{M}(\sigma_*l-\sigma_*E_1) = \left(\mathbf{P}^1 - \text{7 pts.}\right) \times \PP^1_a \cup 7\left(\PP^1_a \times \PP^1_a\right) \label{lE1mod} \end{equation} where, globally, the first component, $\mathcal{M}_{\text{generic}}$, in fact, extends to a $\mathbf{P}^1$ bundle over $\mathbf{P}^1$. We can also describe the way each copy of $\PP^1_a\times\PP^1_a$ is attached to $\mathcal{M}_{\text{generic}}$: the diagonal of $\PP^1_a\times\PP^1_a$, the set of points where the two components intersect, is glued to a fiber of the $\PP^1_a$ bundle $\mathcal{M}_{\text{generic}}$. Physically, as we discussed above, at a generic point in the moduli space we have a single fivebrane wrapping a curve which lies solely in the base of the Calabi--Yau threefold and is topologically a sphere. The curve can be moved in the base and in $S^1/Z_2$ but not in the fiber direction. In moving around the base there are seven special points where the fivebrane splits into two curves intersecting at one point, as in Figure~\ref{Cpinch}. These are each fixed in both the base and the fiber of the Calabi--Yau manifold, but can now each move independently in $S^1/Z_2$. The two fivebranes can then be separated so that they no longer intersect. In making the transition from one of these branches of the moduli space to the case where there is a single fivebrane, the two fivebranes must be at the same point in $S^1/Z_2$ and have the same value of the axionic scalar $a$. They can then combine and be deformed away within the base as a single curve. This structure is shown in Figure~\ref{lE1modfig}. Note that, unlike the pure fiber case~\eqref{fFmod}, the two curves $W_1$ and $W_2$ are distinguishable, since they wrap different cycles in the base, so we do not have to be concerned with modding out by discrete symmetries. \begin{figure} \centerline{\psfig{figure=lE1mod.eps,height=2.8in}} \caption{The moduli space $\mathcal{M}(\sigma_*l-\sigma_*E_1)$.} \label{lE1modfig} \end{figure} Since in our example all the curves are topologically spheres, there are generically no vector fields in the low-energy theory. However, at the points where there is a transition between the two-fivebrane branch and the single fivebrane branch, additional low-energy fields can appear. These correspond to membranes which stretch between the two fivebranes becoming massless as the fivebranes intersect. \subsection{$[W]=\sigma_*l-\sigma_*E_1+F$} \label{exbase2} We can generalize the previous example by including a fiber component in the class of $W$, so that \begin{equation} [W] = \sigma_*l - \sigma_*E_1 + F \end{equation} Note that from~\eqref{thrm} this class is effective. We immediately see that one simple possibility is that $W$ splits into two curves \begin{equation} W = W_0 + W_F \end{equation} where \begin{equation} [W_0] = \sigma_*l - \sigma_*E_1, \qquad [W_F] = F \end{equation} The moduli space of the $W_0$ component will be exactly the same as our previous example, while for the pure fiber component, as given in equation~\eqref{Fmod}, the moduli space is locally $B\times \PP^1_a$. Since the base in this example is $dP_8$ here we conclude that, when the curve splits, this part of the moduli space is just the product of the moduli spaces for $W_0$ and $W_F$, that is \begin{equation} \mathcal{M}(\sigma_*l-\sigma_*E_1) \times \left(dP_8 \times \PP^1_a\right) \label{largecomp} \end{equation} where $\mathcal{M}(\sigma_*l-\sigma_*E_1)$ was given above in~\eqref{lE1mod}. Physically we have two fivebranes, one wrapped on the fiber and one on the base, which can each move independently. As discussed above, for the curve wrapped on the base there are certain special points in moduli space where it splits into a pair of fivebranes, so that, at these special points, we have a total of three independent fivebranes. Since the curves of the base are all topologically spheres, their genus is zero. Hence, the only vector multiplets come from the fivebrane wrapping the fiber which, being topologically a torus with $g=1$, gives a $U(1)$ theory. Generically, the five brane wrapping the fiber $W_F$ does not intersect the fivebranes in the base $W_0$. However, there is a curve of points in the moduli space of $W_F$ where both fivebranes are in the same position in $S^1/Z_2$, with the same value of $a$, and $W_0$ lies above $W_B$ in the Calabi--Yau fibration. Generically, this gives a single intersection. However, there is a special point, when the base curves splits, as in~\eqref{simplesplit} and~\eqref{Wsplit}, and the fiber component intersects exactly the point where the two base curves intersect. At such a point in moduli space, we have three fivebranes intersecting at a single point in the Calabi--Yau threefold. These different possible intersections are shown in Figure~\ref{enhance}. Generically, we expect there to be additional multiplets in the low-energy theory at such points. \begin{figure} \centerline{\psfig{figure=enhance.eps,height=2in}} \caption{Possible enhancements in the moduli space of $[W]=\sigma_*l-\sigma_*E_1+F$. The first figure is where the curve in the base $W_0$ does not split. The second gives two possible cases when $W_0$ splits.} \label{enhance} \end{figure} We might expect that there is also a component of the moduli space where the curve $W$ does not split at all, that is, where we have a single fivebrane in the class $\sigma_*l-\sigma_*E_1+F$. To analyze this second possibility we simply follow the analysis given above, where we first consider the moduli space of the image $C$ of $W$ in the base and then find the moduli space of curves which project down to the same given curve $C$. Since the projection of $F$ onto the base is zero, the image of $[W]$ in the base is $\Omega=l-E_1$, as above. Thus many of the results of the previous discussion carry over to this situation. The moduli space of $C$ is given by~\eqref{Cmodspace}. At a generic point in moduli space $S_C=\pi^{-1}(C)$ is a K3 surface, while at the seven special points where $C$ splits, the surface above each of the two components of $C$ is a $dP_9$ surface. If we consider first a generic point in the moduli space, $W$ is again a section of the K3 surface, but must now be in the class $\sigma_*l-\sigma_*E_1+F$. How many such sections are there? It turns out that for generic K3 there are none. We can see this as follows. By adjunction, since the genus of $C$ was zero, we showed in equation~\eqref{Omegasq1} that its class in $S_C$ satisfies $D_C\cdot D_C=-2$. The identical calculation applies to any section, since all sections have genus zero. Thus, in particular, we have \begin{equation} [W]_C \cdot [W]_C = - 2 \label{contr1} \end{equation} where by $[W]_C$, we mean the class of $W$ in $S_C$. However, from the map between classes~\eqref{K3inclmap}, it is clear that $[W]_C=D_C+F_C$. Since $D_C$ is the class of the zero section, we have $D_C\cdot F_C=1$. For the fiber class we always have $F_C\cdot F_C=0$. Hence, we must also have \begin{equation} [W]_C \cdot [W]_C = 0 \label{contr2} \end{equation} This contradiction implies that there can be no sections of $S_C$ in the class $\sigma_*l-\sigma_*E_1+F$. In other words, we have shown that, generically, we cannot have just a single fivebrane in the class $\sigma_*l-\sigma_*E_1+F$. Rather, the fivebrane always splits into a pure fiber component and a pure base component, as described above. What about the special points in the moduli space where the curve $C$ splits into two? Do we still have to have a separate pure fiber component? The answer is no, for the reason that, as discussed above, the space above each component is a $dP_9$ surface and, unlike the K3 case, there are many more algebraic classes on $dP_9$ than just the zero section and the fiber. Specifically, suppose there is no separate pure fiber component in $W$ and consider the point where $C$ splits into $C_1+C_2$ with $[C_1]=l-E_1-E_2$ and $[C_2]=E_2$. The actual curve $W$ must also split into $W_1$ and $W_2$. Given that each component must be effective, we then have two possibilities, depending on which component includes the fiber class \begin{equation} \begin{gathered} {}[W_1] = \sigma_*l - \sigma_*E_1 - \sigma_*E_2 + F, \qquad [W_2] = \sigma_*E_2 \\ \text{or} \\ {}[W_1] = \sigma_*l - \sigma_*E_1 - \sigma_*E_2, \qquad [W_2] = \sigma_*E_2 + F \end{gathered} \label{dP9Fdecomp} \end{equation} Let us concentrate on the first case, although a completely analogous analysis holds in the second example. Above, we calculated the number of curves in the case where the class contains no $F$. We found, for instance, that if $[W_2]=\sigma_*E_2$ then $W_2$ is required to be precisely the section $C_2$ and there is no moduli space for moving the curve in the fiber direction. The situation is richer, however, for the class with an $F$ component. We will discuss this is more detail in section~\ref{sec:dP9} below, but it turns out that there are 240 different sections of $dP_9$ in the class $[W_1]=\sigma_*l-\sigma_*E_1-\sigma_*E_2+F$. It is a general result, just repeating the calculation that led to~\eqref{dP9sec}, that any section of the $dP_9$ has self-intersection $-1$. Consequently none of the 240 different sections in the class $\sigma_*l-\sigma_*E_1-\sigma_*E_2+F$ can be deformed in the fiber direction and, hence, they simply provide a discrete set of different $W_1$ which all map to the same $C_1$. Furthermore, one can show that none of these sections intersect the base of the Calabi--Yau manifold. Thus, since, in the case we are considering, $W_2$ lies solely in the base, we find that $W_1$ and $W_2$ can never overlap. Their relative positions within the Calabi--Yau threefold are shown in Figure~\ref{discrete}. \begin{figure} \centerline{\psfig{figure=discrete.eps,height=3in}} \caption{$W=W_1+W_2$ in the case where $[W_1]=\sigma_*l-\sigma_*E_1-\sigma_*E_2+F$ and $[W_2]=\sigma_*E_2$.} \label{discrete} \end{figure} It is important to note that these curves are completely stuck within the Calabi--Yau threefold. They cannot combine into a single curve and move away from the exceptional point in the moduli space of $C$ (the projection of $W$ into the $dP_8$ base) where $C$ splits into two curves. Furthermore, we have argued that they cannot move in the fiber. Thus, the only moduli for this component of the moduli space are the positions of the two fivebranes in the orbifold interval and the values of their axions, giving a moduli space of \begin{equation} \PP^1_a \times \PP^1_a \label{smallcomp} \end{equation} Furthermore, since all the sections of $dP_9$ are topologically spheres, there are no vector multiplets in this part of the moduli space. As we noted above, the fivebranes cannot overlap. Hence, there is no possibility of additional multiplets appearing. Finally, we note that there were seven ways $C$ could split into $C_1+C_2$, and, for each splitting, $W$ can decompose one of two ways~\eqref{dP9Fdecomp}. Since for each decomposition there are 240 distinct sections, we see that there is a grand total of 3360 ways of making the analogous decomposition to that we have just discussed. In conclusion, we see that the full moduli space for $[W]=\sigma_*l-\sigma_*E_1+F$ has a relatively rich structure. It splits into a large number of disconnected components. The largest component is where $W$ splits into separate fiber and base components, $W=W_0+W_F$. The moduli space is then given by~\eqref{largecomp}, which includes the possibility of the base component splitting. There are then 3360 disconnected components where $W$ splits into two irreducible components, one of which includes the fiber class $F$. We can summarize this structure in a table \begin{equation} \begin{array}{c|cc} [W] & \text{genus} & \text{moduli space} \\ \hline\hline (\sigma_*l-\sigma_*E_1) + (F) & 0+1 & \left(\left[\mathbf{P}^1 - 7\text{ pts.}\right]\times\PP^1_a\right) \times \left(dP_8\times\PP^1_a\right) \\ (\sigma_*l-\sigma_*E_1-\sigma_*E_2) + (\sigma_*E_2) + (F) & 0+0+1 & \PP^1_a\times\PP^1_a \times \left(dP_8\times\PP^1_a\right) \\ \hline (\sigma_*l-\sigma_*E_1-\sigma_*E_2+F) + (\sigma_*E_2) & 0+0 & \PP^1_a\times\PP^1_a \end{array} \label{lE1Fmodspace} \end{equation} Here, the first column gives the homology classes in $X$ of the different components of $W$. The first two rows describe the moduli space~\eqref{largecomp} where $W$ splits into $W_0+W_F$, first for a generic component in the base and then at one of seven points where the base component splits. The final row describes one of the 3360 disconnected components of the moduli space~\eqref{smallcomp}. From the genus count, we see that in the first two cases we expect a $U(1)$ gauge field on the fivebrane which wraps the fiber, while in the last case there are no vector multiplets. As we have noted, the component of the moduli given in the first two rows has the possibility that the fivebranes intersect leading to additional low-energy fields, as depicted in Figure~\ref{enhance}. The disconnected components, have no such enhancement mechanism. Furthermore, their moduli space is severely restricted since neither fivebrane can move within the Calabi--Yau manifold. \section{General procedure for analysis of the moduli space} \label{gen} From the examples above, we can distill a general procedure for the analysis of the generic moduli space. We start with a general fivebrane curve in the class \begin{equation} [W] = \sigma_*\Omega + f F \label{initialW} \end{equation} Furthermore, $[W]$ is assumed to be effective, so that, by~\eqref{thrm}, $\Omega$ is some effective class in $H_2(B,\mathbf{Z})$ and $f\geq 0$. We need to first find the moduli space of the projection $C$ of $W$ onto the base. One then finds all the curves in the full Calabi--Yau threefold in the correct homology class which project onto $C$. In general, we will find that the above case, where the space $S_C=\pi^{-1}(C)$ above $C$ was a K3 surface, is the typical example. There, we found that no irreducible curve which projects onto $C$ could include a fiber component in its homology class. Hence, $[W]$ splits into a pure base and a pure fiber component. One exception, as we saw above, is when $S_C$ is a $dP_9$ surface. In the following, we will start by analysing the generic case and then give a separate discussion for the case where $dP_9$ appears. \subsection{Decomposition of the moduli space} \label{decomp} Any curve $W$ in the Calabi--Yau threefold can be projected to the base using the map $\pi$. In general, $W$ may have one or many components. Typically, a component will project to a curve in the base. However, there may also be components which are simply curves wrapping the fiber at different points in the base. These curves will all project to points in the base rather than curves. Thus, the first step in analyzing the moduli space is to separate out all such curves. We therefore write $W$ as the sum of two components, each of which may be reducible, % \begin{equation} W = W_0 + W_F \label{decompF} \end{equation} but with the assumption that none of the components of $W_0$ are pure fiber components. For general $[W]$ given in~\eqref{initialW}, the classes of these components are \begin{equation} [W_0] = \sigma_*\Omega + nF \qquad [W_F] = (f-n)F \end{equation} with $0\leq n \leq f$, since each class must be separately effective. Note that, although $W_0$ has no components which are pure fiber, its class may still involve $F$ since, in general, components of $W_0$ can wrap around the fiber as they wrap around a curve in the base. Except when $n=f$, so $W_F=0$, $W$ has at least two components in this decomposition and so there are at least two five-branes. In general, the decomposition~\eqref{decompF} splits the moduli space into $f+1$ different components depending on how we partition the $f$ fiber classes between $[W_0]$ and $[W_F]$. Within a particular component, the moduli space is a product of the moduli space of $W_0$ and $W_F$. If $\mathcal{M}_F((f-n)F)$ is the moduli space of $W_F$ and $\mathcal{M}_0(\sigma_*\Omega+nF)$ is the moduli space of $W_0$, we can then write the full moduli space as \begin{equation} \mathcal{M}(\sigma_*\Omega + fF) = \bigcup_{n=0}^f \mathcal{M}_0(\sigma_*\Omega+nF) \times \mathcal{M}((f-n)F) \label{moddecomp} \end{equation} The problem is then reduced to finding the form of the moduli spaces for $W_0$ and $W_F$. The latter moduli space has already been analyzed in section~\ref{sec:fF}. We found that, locally, \begin{equation} \mathcal{M}((f-n)F) = \left(B\times\PP^1_a\right)^{f-n}/\mathbf{Z}_{f-n} \label{WFmod} \end{equation} Thus we are left with $\mathcal{M}_0(\sigma_*\Omega+nF)$, which can be analyzed by the projection techniques we used in the preceding examples. Projecting $W_0$ onto the base, we get a curve $C$ in the class $\Omega$ in $B$. Let us call the moduli space of such curves in the base $\mathcal{M}_B(\Omega)$. This space is relatively easy to analyze since we know the form of $B$ explicitly. In general, it can be quite complicated with different components and branches as curves degenerate and split. To find the full moduli space $\mathcal{M}_0(\sigma_*\Omega+nF)$, we fix a point in $\mathcal{M}_B(\Omega)$ giving a particular curve $C$ in $B$ which is the projection of the original curve $W_0$ in the Calabi--Yau threefold. In the following, we will assume that $C$ is not singular. By this we mean that it does not, for instance, cross itself or have a cusp in $B$. If it is singular, it is harder to analyze the space of curves $W_0$ which project onto $C$. In general, $C$ splits into $k$ components, so that \begin{equation} C = C_1 + \dots + C_k \label{genCdecomp} \end{equation} We then also have \begin{equation} \Omega = \Omega_1 + \dots + \Omega_k \end{equation} where $\Omega_i=[C_i]$ and, so, must be an effective class on the base for each $i$. Clearly if the curve in the base has more than one component then so does the original curve $W_0$, so that \begin{equation} W_0 = W_1 + \dots + W_k \label{Widecomp} \end{equation} with $\pi(W_i)=C_i$. In general, the class of $W_0$ will be partitioned into a sum of classes of the form \begin{equation} [W_i] = \sigma_*\Omega_i + n_i F \label{Wiclasses} \end{equation} where, for each curve to be effective, $n_i\geq 0$ and $n_1+\dots+n_k=n$, leading to a number of different possible partitions. One now considers a particular component $C_i$. To find the moduli space over $C_i$, one needs to find all the curves $W_i$ in $X$ in the cohomology class $\sigma_*\Omega_i+n_iF$ which project on $C_i$. Recall, in addition, that we have assumed in our original partition~\eqref{decompF} that $[W_i]$ contains no pure fiber components. Repeating this procedure for each component and for each partition of the $n$ fibers into $\{n_i\}$, gives the moduli space over a given point $C$ in $\mathcal{M}_B(\Omega)$ and, hence, the full moduli space. Consequently, we have reduced the problem of finding the full moduli space $\mathcal{M}_0(\sigma_*\Omega+nF)$ to the following question. To simplify notation, let $R$ be the given irreducible curve $C_i$ in the base, and let $V$ be the corresponding curve $W_i$ in the full Calabi--Yau threefold. Let us further write the class $\Omega_i$ of $C_i$ as $\Lambda$ and write $m$ for $n_i$. Our general problem is then to find, for the given irreducible curve $R$ in the effective homology class $\Lambda$, what are all the curves $V$ in the Calabi--Yau threefold in the class $\sigma_*\Lambda+mF$, where $m\geq 0$ which project onto $R$. Necessarily, all the curves $V$ which project into $R$ lie in the surface $S_R=\pi^{-1}(R)$. By construction, $S_R$ is elliptically fibered over the base curve $R$. Furthermore, typically, the map from $V$ to $R$ wraps $R$ only once. It is possible that $R$ is some number $q$ of completely overlapping curves in $B$, so that $[R]=q\Gamma$ for some effective class $\Gamma$ in $B$. Then the map from $V$ to $R$ wraps the base curve $q$ times. This will occur in one of the examples we give later in the paper, but here, since we are discussing generic properties, let us ignore this possibility. Then, assuming $R$ is not singular, the map is invertible and we see that $V$ must be a section of the fibered surface $S_R$. Furthermore, we note that $S_R$ can be characterized by the genus $g$ of the base curve $R$ and the number of singular fibers. The former is, by adjunction, given by \begin{equation} 2g - 2 = \left(K_B + \Lambda\right) \cdot \Lambda \label{genus} \end{equation} The latter is also a function only of the class $\Lambda$ of $R$ and can be found by intersecting the discriminant class $[\Delta]$ with $\Lambda$. Since $[\Delta]=-12K_B$, the number of singular fibers must be of the form $12p$ with \begin{equation} p = - K_B \cdot \Lambda \label{pdef} \end{equation} where $p$ is an integer. If $p$ is negative, the curve $R$ lies completely within the discriminant curve of the elliptically fibered Calabi--Yau manifold. This means that every fiber above $R$ is singular. The form of $S_R$ then depends on the structure of the particular fibration of the Calabi--Yau threefold. Since we want to consider generic properties of the moduli space, we will ignore this possibility and restrict ourselves to the case where $p$ is non-negative. We note that this is not very restrictive. For del Pezzo and Enriques surfaces $-K_B$ is nef, meaning that its intersection with any effective class in the base is non-negative. Hence, since $\Lambda$ must be effective, we necessarily have $p\geq 0$. The only exceptions, are Hirzebruch surfaces $F_r$ with $r\geq 3$ and where $\Lambda$ includes the negative section $\mathcal{S}$. Finally, then, finding the full moduli space $\mathcal{M}_0(\sigma_*\Omega+nF)$ has been reduced to the following problem \begin{itemize} \item For a given irreducible curve $R$ in the base $B$ with homology class $\Lambda$, find the moduli space of sections $V$ of the surface $S_R=\pi^{-1}(R)$ in the homology class $\sigma_*\Lambda+mF$ in the full Calabi--Yau threefold $X$, where $m\geq 0$. $S_R$ is characterized by $g=\genus{C}$, as given in~\eqref{genus}, and $p$, where $12p$ is the number of singular elliptic fibers, as given in~\eqref{pdef}. Consequently, we write this moduli space as $\mathcal{M}(g,p;m)$. \end{itemize} We will assume that $p\geq 0$. This is necessarily true, except when $B$ is an $F_r$ surface with $r\geq 3$ and $\Lambda$ contains the class of the section at infinity $\mathcal{S}$. \subsection{The generic form of $\mathcal{M}(g,p;m)$} \label{generic} To understand the sections of $S_R$, we start by finding the algebraic classes on $S_R$. As for the K3 surface, a generic surface $S$ has no algebraic classes since $h^{2,0}\neq 0$. For $S_R$, we know that two classes are necessarily present, the class of the zero section $D_R$ and the fiber class $F_R$. However, generically, there need not be any other classes. Additional classes may appear for special choices of complex structure, but here we will consider only the generic case. The obvious exception is the case where $g=0$ and $p=1$. From equations~\eqref{genus} and~\eqref{pdef}, this implies that $R$ is an exceptional curve in $B$. The surface $S_R$ is then an elliptic fibration over $\mathbf{P}^1$ with 12 singular fibers, which is a $dP_9$ surface. In this case $h^{2,0}=0$ and every class in $H_2(S_R,\mathbf{Z})$ is algebraic. We will return to this case in the next section. The inclusion map $i_R:S_R\to X$, gives a natural map between classes in $S_R$ and in $X$ \begin{equation} i_{R*} : H_2(S_R,\mathbf{Z}) \to H_2(X,\mathbf{Z}) \end{equation} In general, with only two classes the map is simple. By construction, we have \begin{equation} \begin{aligned} i_{R*}D_R &= \sigma_*\Lambda \\ i_{R*}F_R &= F \end{aligned} \label{geninclmap} \end{equation} Just as in the K3 example, we will find that the existence of only two classes strongly constrains the moduli space $\mathcal{M}(g,p;m)$. We can find the analog of the contradiction of equations~\eqref{contr1} and~\eqref{contr2} as follows. Let $K_{S_R}$ be the cohomology class of the canonical bundle of the surface. Since both $V$ and $R$ are sections, they have the same genus $g$. By the Riemann--Hurwitz formula and adjunction, we have \begin{equation} 2g - 2 = \left(K_{S_R} + D_R\right) \cdot D_R = K_{S_R} \cdot D_R + D_R \cdot D_R \label{Radj} \end{equation} where $D_R$ is the class of the zero section, and \begin{equation} 2g - 2 = \left(K_{S_R} + [V]_R\right) \cdot [V]_R = K_{S_R} \cdot [V]_R + [V]_R \cdot [V]_R \label{Vadj} \end{equation} where $[V]_R$ the class of $V$ in $S_R$. We also have, for the fiber, by a similar calculation, since it has genus one and, since two generic fibers do not intersect, $F_R\cdot F_R=0$, that \begin{equation} 0 = \left(K_{S_R} + F_R\right) \cdot F_R = K_{S_R} \cdot F_R \end{equation} Finally, since we require in $X$ that $[V]=\sigma_*\Lambda+mF$ and since we assume that no additional classes exist on $S_R$, % \begin{equation} [V]_R = D_R + mF_R \label{Vclass} \end{equation} Substituting this expression into~\eqref{Vadj} and using~\eqref{Radj}, we find \begin{equation} [V]_R \cdot [V]_R = D_R \cdot D_R \label{Vsqr1} \end{equation} On the other hand, given that $F_R\cdot D_R=1$ because the fiber intersects a section at one point, we can compute the self-intersection of $[V]_R=D_R+mF_R$, explicitly, yielding \begin{equation} [V]_R \cdot [V]_R = D_R \cdot D_R + 2m \label{Vsqr2} \end{equation} Comparing~\eqref{Vsqr1} with~\eqref{Vsqr2}, we are left with the important conclusion that we must have $m=0$. This implies that, generically, no component of $W_0$ can contain any fibers in its homology class. We conclude that \begin{equation} \mathcal{M}(g,p;m) = \emptyset \qquad \text{unless} \quad m=0 \end{equation} In fact, we can go further. Since we require $m=0$, we see from~\eqref{Vclass} that $V$ is in the same class $D_R$ as the zero section $R$. Using the Riemann--Roch formula and Kodaira's description of elliptically fibered surfaces, one can show that the canonical class in the cohomology of $S_R$ is given by \begin{equation} K_{S_R} = \left(2g - 2 + p\right) F_R \label{KSR} \end{equation} Then, substituting this expression into~\eqref{Radj}, we see that \begin{equation} D_R \cdot D_R = - p \end{equation} Thus we see that for $p>0$, $R$ is an exceptional divisor and cannot be deformed in $S_R$. Consequently, all the fivebrane can do is to move in the orbifold direction and change its value of $a$, so we have a moduli space of $\PP^1_a$. If $p=0$, the fibration is locally trivial. It may or may not be globally trivial. However, it is always globally trivial when pulled back to some finite cover of the base. Every section is in the class $D_R$ and the moduli space simply corresponds to moving the fivebrane in the fiber direction and in $S^1/Z_2$ and $a$. If the original fibration was globally trivial, this yields a moduli space of $E\times\PP^1_a$, where $E$ is an elliptic curve describing motion of $V$ in the fiber direction. If it was only locally trivial, then these deformations in the fiber directions still make sense over the cover, but only a finite subset of them happens to descend, so the actual moduli space consists in this case of some finite number of copies of $\PP^1_a$. In summary, we see that, generically, if we exclude the case $g=0$, $p=1$ where $S_R$ is a $dP_9$ surface, \begin{equation} \mathcal{M}(g,p;m) = \emptyset \qquad \text{for} \quad m > 0 \label{gencMR1} \end{equation} while if $m=0$ we have, \begin{equation} \mathcal{M}(g,p;0) = \begin{cases} N \PP^1_a & \text{if } p = 0 \text{ and the fibration is not globally trivial} \\ E\times\PP^1_a & \text{if } p = 0 \text{ and the fibration is globally trivial}\\ \PP^1_a & \text{if } p > 0 \end{cases} \label{gencMR2} \end{equation} where $N$ is some integer depending on the global structure of $S_C$. In each case, the gauge group on the fivebrane is given by $U(1)^g$ where $g$ is the genus of the curve $R$ in the base $B$. Since both $R$ and $V$ are sections of $S_R$, this is equal to the genus of the curve $V$ in the space $X$. \subsection{The $dP_9$ exception and ${\cal{M}}(0,1;m)$} \label{sec:dP9} As we have mentioned, the obvious exception to the above analysis is when $S_R$ is a $dP_9$ surface. This occurs when the base curve $R$ is topologically $\mathbf{P}^1$ and there are 12 singular elliptic fibers in $S_R$, that is if $[R]=\Lambda$ in the base $B$, \begin{equation} 2g - 2 = \left(K_B+\Lambda\right)\cdot\Lambda = -2, \qquad p = - K_B \cdot \Lambda = 1 \label{dP9cond} \end{equation} As we mentioned above, this implies that $R$ is an exceptional curve in the base. In this case, $D_R$ and $F_R$ are not the only generic algebraic classes on $S_R$. Rather, since $h^{2,0}=0$, every integer class in $dP_9$ is algebraic. The surface $dP_9$ can be described as the plane $\mathbf{P}^2$ blown up at nine points which are at the intersection of two cubic curves. Consequently, there are ten independent algebraic classes on $dP_9$, the image $l'$ of the class of a line in $\mathbf{P}^2$ and the nine exceptional divisors, $E'_i$ for $i=1,\dots,9$, corresponding to the blown up points. Here we use primes to distinguish these classes from classes in the base $B$ of the Calabi--Yau threefold (specifically, the $E_i$ classes in $B$ when the base is a del Pezzo surface). The point here is that, in the full Calabi--Yau manifold $X$, there are only two independent classes associated with $S_R$, namely, the class of the base curve $\sigma_*\Lambda$ and the fiber class $F$. Consequently, if $i_R:S_R\to X$ is the inclusion map from the $dP_9$ surface into the Calabi--Yau threefold, the corresponding map between classes \begin{equation} i_{R*} : H_2(S_R,\mathbf{Z}) \to H_2(X,\mathbf{Z}) \label{iRmap} \end{equation} must have a non-empty kernel, since it maps the ten independent classes on $S_R$ mentioned above into only two in $X$. Recall that our goal is to find the set of sections $V$ of $S_R$ which are in the class $\sigma_*\Lambda+mF$ in $X$. Previously, we found that generically there were no such sections unless $m=0$. This followed from equations~\eqref{Radj} to~\eqref{Vsqr2}. Now, the appearance of a kernel in the map~\eqref{iRmap}, means that $[V]_R$ is no longer necessarily of the form $D_R+mF_R$, as in equation~\eqref{Vclass}. Consequently we can no longer conclude that we must have $m=0$. In fact, as we will see, there are several different sections $V$ in the same class $\sigma_*\Lambda+mF$ in $X$ with $m>0$. Let us start by recalling why a $dP_9$ surface is also an elliptic fibration. Viewed as $\mathbf{P}^2$ blown up at nine points, we have the constraint that the nine points lie at the intersection of two cubics. This is represented in Figure~\ref{dP9}. The cubics can be written as two third-order homogeneous polynomials in homogeneous coordinates $[x,y,z]$ on $\mathbf{P}^2$. Let us call these polynomials $f$ and $g$. Clearly any linear combination $af+bg$ defines a new cubic polynomial. By construction, this polynomial also passes through the same nine points. Since the overall scale does not change the cubic, the set of cubics is given by specifying $a$ and $b$ up to overall scaling. Thus, we have a $\mathbf{P}^1$ of cubics passing through the nine points. Since each cubic defines an elliptic curve, we can think of this as an elliptic fibration over $\mathbf{P}^1$. Furthermore, the cubics cannot intersect anywhere else in $\mathbf{P}^2$. The space of cubics spans the whole of the plane and, further, it blows up each intersection point into a $\mathbf{P}^1$ of distinct points, one point in $\mathbf{P}^1$ for each cubic passing through the intersection. Thus the space of cubics is giving an alternative description of the $dP_9$ surface. In addition, we note that each of the exceptional divisors $E'_i$, the blow-ups of the intersection points, is a $\mathbf{P}^1$ surface which intersects each fiber at a single point, and so corresponds to section of the fibration. Furthermore, the anti-canonical class of $dP_9$ is given by \begin{equation} - K_{S_R} = 3l' - E'_1 - \dots - E'_9 = F_R \label{dP9FR} \end{equation} and is precisely the class of the cubics passing through the nine intersection points. It follows that \begin{equation} K_{S_R} = - F_R \end{equation} which, since $g=0$ and $p=1$, agrees with the general expression~\eqref{KSR}. \begin{figure} \centerline{\psfig{figure=dP9.eps,height=2.5in}} \caption{The space $dP_9$} \label{dP9} \end{figure} It is natural to ask if there are other sections of $dP_9$ aside from the exceptional curves $E'_i$. We first note that any section of $dP_9$ is exceptional, with $[V]_R\cdot[V]_R=-1$. This follows from~\eqref{Vadj} and~\eqref{dP9FR}, recalling that any section will have genus zero and intersects the fiber (the anti-canonical class) only once. Next, we recall that there is a notion of addition of points on elliptic curves. Consequently, we can add sections point-wise to get a new section. Thus, we see that the set of sections forms an infinite Abelian group containing all the exceptional curves on the $dP_9$ surface. From the point of view of curves in $\mathbf{P}^2$, these additional exceptional classes correspond to curves of higher degree passing through the nine intersection points with some multiplicity (see for instance~\cite{dlow2}). In general, we can write an exceptional class (except the classes $E'_i$) as \begin{equation} Q = q l' - \sum q_i E'_i \label{gensec} \end{equation} such that \begin{equation} q, q_i \geq 0, \quad q^2 - \sum q_i^2 = -1, \quad 3q - \sum q_i = 1 \label{genseccond} \end{equation} where the first condition is required in order to describe an effective curve in $\mathbf{P}^2$, the second condition gives $Q\cdot Q=-1$ and the third condition gives $Q\cdot F_R=1$. The appearance of an infinite number of exceptional classes means that these equations have an infinite number of solutions for $i=1,\dots,9$. Having identified all the relevant classes in $dP_9$, we can now turn to a description of the map~\eqref{iRmap}. First, we note that the group structure of the set of sections means that we can choose any section as part of the basis of classes. In our case, we have singled out one section as the class of the base curve $R$. Thus, without loss of generality, we can identify this class with one of the $E'_i$, for example, $E'_9$. Thus we set \begin{equation} D_R = E'_9 \end{equation} By construction, we know that $D_R$ maps to $\sigma_*\Lambda$ and $F_R$ maps to $F$ under $i_{R*}$. Thus, in terms of $l'$ and $E'_i$, we have, using~\eqref{dP9FR} and the generic result~\eqref{geninclmap}, \begin{equation} \begin{gathered} i_{R*} E'_9 = \sigma_*\Lambda \\ i_{R*} \left(3l'-E'_1-\dots-E'_9\right) = F \end{gathered} \label{genincl} \end{equation} We now need to understand how the remaining independent classes $E'_1,\dots,E'_8$ map under $i_{R*}$. We first note that, since all these classes are sections, they must project onto the class $\Lambda$ in the base. The only ambiguity is how many multiples of the fiber class $F$ each contains. Thus, we know \begin{equation} i_{R*}E'_i = \sigma_*\Lambda + c_i F \end{equation} for some $c_i$. The easiest way to calculate $c_i$ is to recall that a class in $H_2(X,\mathbf{Z})$ is uniquely determined by its intersection numbers with a basis of classes in $H_4(X,\mathbf{Z})$. A suitable basis was given in section~\ref{algclasses}. In particular, we can consider the intersection of $E'_i$ with the class $D$ of base $B$ of the full Calabi--Yau threefold. We note that $E'_i\cdot E'_9=0$ for $i=1,\dots,8$, so the extra sections $E'_i$ do not intersect the base $R$ of the $dP_9$. However, since $R$ describes the intersection of $B$ with the $dP_9$ surface $S_R$, we see that the extra classes cannot intersect $B$. Thus, we must have \begin{equation} i_{R*} E'_i \cdot D = 0 \qquad \text{for} \quad i = 1,\dots,8 \end{equation} From the table of intersections~\eqref{inters}, using~\eqref{dP9cond}, we see that we must have $c_i=1$ for $i=1,\dots,8$. Together with the result~\eqref{genincl} we find, in conclusion, that the map~\eqref{iRmap} is given by \begin{equation} \begin{aligned} i_{R*} l' &= 3 \sigma_*\Lambda + 3 F \\ i_{R*} E'_i &= \sigma_*\Lambda + F \qquad \text{for} \quad i = 1,\dots,8 \\ i_{R*} E'_9 &= \sigma_*\Lambda \end{aligned} \label{fulliR} \end{equation} demonstrating explicitly that the map has a kernel. Having identified the map, we would now like to return to our original question, which was how many sections are there in the class $\sigma_*\Lambda+mF$ and what is their moduli space. The second part is easy to answer. We have noted that all sections are exceptional with self-intersection $-1$. This implies that they cannot be moved within the $dP_9$ surface. However, they can move in the orbifold and have different values of the axion. Consequently, each section has a moduli space of $\PP^1_a$. If there are a total of $N(m)$ sections for a given $m$, then the total moduli space $\mathcal{M}(0,1;m)$ has the form \begin{equation} \mathcal{M}(0,1;m)=N(m)\PP^1_a \label{cMm} \end{equation} The value of $N(m)$ is a problem in discrete mathematics. Recall that the class of a general section had the form given in~\eqref{gensec}. Under the map $i_{R*}$ this maps into \begin{equation} i_{R*}Q = \sigma_*\Lambda + \left(q_9+1\right) F \end{equation} where we have used the condition $3q-\sum q_i=1$. Thus, we can summarize the problem as finding the number of solutions to \begin{equation} \begin{aligned} q^2 - \sum q_i^2 + 1 &= 0 \\ 3q - \sum q_i &= 1 \end{aligned} \label{sectcond} \end{equation} with \begin{equation} q \geq 0, \quad q_i \geq 0, \quad q_9 = m - 1 \end{equation} We will not solve this problem in general, but just note the solution of $m=0$ and $m=1$. The former case is not actually included in the above form, since it implies $q_9=-1$. However, this case is easy to analyze since, with $m=0$, we see, from~\eqref{inters} and~\eqref{dP9cond}, that the intersection of $[V]=\sigma_*\Lambda$ with the base class $D$ of the Calabi--Yau manifold is $-1$. Consequently, a component of $V$ must lie in the base $B$. Since, by assumption, $V$ is irreducible, this means that the whole of $V$ lies in $B$. Thus, $V$ can only be the base section $R$ in the $dP_9$ surface. Thus we conclude that, for a $dP_9$ surface $S_R$, $N(0)=1$. That is, \begin{equation} \mathcal{M}(0,1;0) = \PP^1_a \end{equation} corresponding to moving the single curve $R$ in $S^1/Z_2$. This result was used in analyzing the example in section~\ref{exbase1}. For $m=1$, there are 240 solutions to the equations~\eqref{sectcond}, that is $N(1)=240$. The easiest way to see this is to note that $m=1$ implies $q_9=0$. Thus, we can ignore $E'_9$. Effectively, one is then trying to count the number of exceptional curves on a $dP_8$ surface. This is known to be a finite number, $240$~\cite{dPref}. Explicitly, they are of the following forms \begin{equation} \begin{aligned} {}& E'_i \\ {}& l' - E'_i - E'_j \\ {}& 2l' - E'_{i_1} - \dots - E'_{i_5} \\ {}& 3l' - 2E'_{i} - E'_{i_1} - \dots - E'_{i_6} \\ {}& 4l' - 2E'_{i_1} - \dots - 2E'_{i_3} - E'_{i_4} - \dots - E'_{i_8} \\ {}& 5l' - 2E'_{i_1} - \dots - 2E'_{i_6} - E'_{i_7} - E'_{i_8} \\ {}& 6l' - 2E'_{i_1} - \dots - 2E'_{i_7} - 3E'_{i_8} \end{aligned} \end{equation} where all the indices run from $1$ to $8$. One can see, using~\eqref{fulliR}, that all these classes map under $i_{R*}$ to $\sigma_*\Lambda+F$. Again, since none of these sections can move in the $dP_9$ surface, the moduli space is simply 240 copies of $\PP^1_a$, \begin{equation} \mathcal{M}(0,1;1) = 240\, \PP^1_a \end{equation} This result was used in analyzing the example in section~\ref{exbase2}. We also note that none of these sections intersects the base section $R$. \section{A three-family $dP_8$ example} In the previous section we have given a general procedure for analyzing the moduli space of fivebranes wrapped on a holomorphic curve in a particular effective class in the Calabi--Yau manifold. The problem is of particular importance because it has been shown~\cite{nse,dlow1,dlow2} that including fivebranes in supersymmetric M~theory compactifications greatly enlarges the number of vacua with reasonable grand unified gauge groups and three families of matter. In the remaining sections, we apply this procedure to some specific examples which arise in the construction of phenomenological models. We will find that the moduli spaces are very rich. Nonetheless, we will find that there are isolated parts of moduli space where the number of moduli is greatly reduced. We will not describe the full moduli spaces here, but rather consider various characteristic components. Let us start with the example given in~\cite{dlow1} and expanded upon in~\cite{dlow2}. There, the base $B$ was a $dP_8$ surface and the class of the fivebranes was given by \begin{equation} [W]= 2\sigma_*E_1 + \sigma_*E_2 + \sigma_*E_3 + 17F \label{eq:y1} \end{equation} where $l$ and $E_i$ for $i=1,\dots,8$ are the line class and the exceptional blow-up classes in the $dP_8$. Since $2E_1+E_2+E_3$ is effective in $dP_8$, this describes an effective class in the Calabi--Yau manifold. This choice of fivebrane class, together with a non-trivial $E_8$ bundle $V_1$, led to a low-energy $SU(5)$ theory with three families. \subsection{General decomposition} Let us follow exactly the procedure laid down in the previous section. First, we separate from $W$ all the pure fiber components, writing it as the sum of $W_0$ and $W_F$ as in~\eqref{decompF}. We write \begin{equation} [W_0] = 2E_1 + E_2 + E_3 + nF, \qquad [W_F] = (17-n) F \end{equation} with $0\leq n\leq 17$. Unless $n=17$, we have at least two separate fivebranes. This splits the moduli space into several different components depending on the partition of 17 into $n$ and $17-n$, as given in~\eqref{moddecomp}. The moduli space for the pure fiber component $W_F$ is just the familiar form, read off from~\eqref{WFmod} \begin{equation} \mathcal{M}((17-n)F) = \left(dP_8 \times \PP^1_a\right)^{17-n}/\mathbf{Z}_{17-n} \end{equation} More interesting is the analysis of the $W_0$ moduli space. As described above, the first step in the analysis is to project $W_0$ onto the base. This gives the curve $C$ which is in the homology class \begin{equation} [C] = 2E_1 + E_2 + E_3 \end{equation} We then need to find the moduli space $\mathcal{M}_B(2E_1+E_2+E_3)$ of $C$ in the base. We recall that the del Pezzo surface $dP_8$ can be viewed as $\mathbf{P}^2$ blown up at eight points. The exceptional classes $E_i$ each have a unique representative, namely the exceptional curve $\mathbf{P}^1$ at the $i$-th blown-up point. Furthermore, we have the intersection numbers $E_i\cdot E_j=-\delta_{ij}$. Thus $[C]$ has a negative intersection number with each of $E_1$, $E_2$ and $E_3$. This implies that it must have a component contained completely within each of the exceptional curves described by $E_1$, $E_2$ and $E_3$. Since these are all distinct, $C$ must be reducible into three components \begin{equation} C = C_1 + C_2 + C_3 \end{equation} where \begin{equation} [C_1] = 2E_1 \qquad [C_2] = E_2 \qquad [C_3] = E_3 \end{equation} None of these components can be moved in the base, since they all have negative self-intersection number. Consequently, we have \begin{equation} \mathcal{M}_B(2E_1+E_2+E_3) = \text{single pt.} \end{equation} corresponding to three fivebranes, each wrapping a different exceptional curve. Since the projection $C$ splits, so must the curve $W_0$ itself. We must have \begin{equation} W_0 = W_1 + W_2 + W_3 \end{equation} We can partition the fiber class in $[W_0]$ in different ways. In general we write \begin{equation} [W_1] = 2\sigma_*E_1 + n_1 F \qquad [W_2] = \sigma_*E_2 + n_2 F \qquad [W_3] = \sigma_*E_3 + n_3 F \label{split1} \end{equation} with $n_1+n_2+n_3=n$ and $n_i\geq0$ since each curve must be separately effective. The problem of finding the full moduli space has now been reduced to finding the moduli space of $W_1$, $W_2$ and $W_3$ separately. We see that, unless $n=17$, we have at least four separate five-branes, one wrapping the pure fiber curve $W_F$ and one wrapping each of the three curves $W_1$, $W_2$ and $W_3$. As discussed in section~\ref{sec:fF}, the pure fiber component can move in the base $B$, as well as in the orbifold. Furthermore it has transitions where it separates into more than one fivebrane. The components $W_i$, meanwhile, are stuck above fixed exceptional curves in the base. They are free to move in the orbifold and may be free to move in the fiber direction (this will be discussed in the following sections). The $W_i$ components cannot intersect since the exceptional curves in the base over which they are stuck cannot intersect. However, the pure fiber components can intersect the $W_i$, leading to the possibility of additional low-energy fields appearing. \subsection{The $W_2$ and $W_3$ components} \label{W2W3} We start by analyzing the $W_2$ and $W_3$ components. As usual, we are interested in the number of sections of the surface $S_{C_i}$ which are in the class $\sigma_*E_i+n_iF$ in the full Calabi--Yau manifold for $i=2,3$. In each case, the base curve $C_i$ wraps an exceptional curve $\mathbf{P}^1$ of one of the blown-up points in $dP_8$. Such a case is familiar from the examples given in sections~\ref{exbase1} and~\ref{exbase2}. The corresponding surface $S_{C_i}$, as expected, since $C_i$ wraps an exceptional curve, is a $dP_9$ manifold. Explicitly, for both $C_2$ and $C_3$ we have \begin{equation} g_i = \text{genus}(C_i) = 0 \end{equation} and the number of singular fibers in the fibration is given by $12p_i$, where \begin{equation} \begin{split} p_i &= - K_{dP_8} \cdot E_i \\ &= \left(3l-E_1-\dots-E_8\right)\cdot E_i = 1 \end{split} \end{equation} Thus, we see that \begin{equation} S_{C_i} \text{ is a $dP_9$ surface for } i = 2, 3 \end{equation} The moduli space for each $W_{i}$, for $i=2,3$, is then the moduli space of sections of the $dP_{9}$ surface in the homology class $\sigma_*E_i + n_i F$ in the full Calabi--Yau space $X$. Since in this case $g_{i}=0$ and $p_{i}=1$, it follows that we are interested in the moduli spaces $\mathcal{M}(0,1;n_i)$ discussed in section~\ref{sec:dP9}. Let us concentrate on $W_2$, since the moduli space of $W_3$ is completely analogous. We recall from section~\ref{sec:dP9} that, for a given $n_2$, there are a finite number of sections of the $dP_9$ in the class $\sigma_*E_2 + n_2 F$ in $X$. Furthermore, all these sections are exceptional with self-intersection $-1$ and so there is no moduli space for moving these curves within $dP_{9}$. Since the curve $C_2$ is also fixed in the base, we see that there are no moduli for moving $W_2$ in the Calabi--Yau threefold. All we are left with are the moduli for moving in $S^1/Z_2$ and the axion modulus. We showed in section~\ref{sec:dP9} that there was precisely one section in the class $\sigma_*E_2$ and 240 in $\sigma_*E_2+F$. Rather than do a general analysis, let us consider some examples for which $n_2=2$. That is \begin{equation} [W_2]=\sigma_*E_2 + 2F \label{eq:hello1} \end{equation} Consequently, we will be interested in the moduli space $\mathcal{M}(0,1;2)$. We know from the previous discussion that \begin{equation} \mathcal{M}(0,1;2)=N(2)\PP^1_a \label{eq:hi} \end{equation} Here, we will not evaluate $N(2)$ but content ourselves with several specific examples. From the general map~\eqref{fulliR}, with $\Lambda=E_2$, we see that we can, for instance, write the class of $W_2$ as \begin{equation} [W_2] = i_{C_2*}\left(l'-E'_1-E'_9\right) \label{eq:hello2} \end{equation} or \begin{equation} [W_2]= i_{C_2*}\left(2l'-E'_1-E'_2-E'_3-E'_4-E'_9\right) \label{eq:hello3} \end{equation} or many other decompositions which we will not discuss here. We might be tempted to include the case where \begin{equation} [W_2] = i_{C_2*}\left(3l' - E'_2 - \dots - E'_9\right) \label{badex} \end{equation} However, this is not a section. This can be seen directly from the fact that it fails to satisfy the equations~\eqref{sectcond}. Alternatively, we note that the nine blown up points in $dP_9$ are not in general position. If a cubic passes through eight of them, then it also passes through the ninth. Consequently, the class $3l'-E'_2-\dots-E'_9$, which is the class of a cubic through eight of the nine points, always splits into two classes: the fiber $F_{C_2}=3l'-E'_1-\dots-E'_9$ and the base $E'_1$. Consequently, the example~\eqref{badex} is always reducible, splitting into a pure fiber component and the section in the class $E'_1$. Since all the sections are genus zero like the base, we have a simple table for the cases given in~\eqref{eq:hello2} and~\eqref{eq:hello3} \begin{equation} \begin{array}{c|cc} [W_2] \text{ in $dP_9$} & \text{genus} & \text{moduli space} \\ \hline l'-E'_1-E'_9 & 0 & \PP^1_a \\ 2l'-E'_1-E'_2-E'_3-E'_4-E'_9 & 0 & \PP^1_a \end{array} \end{equation} All the other possible sections in the class $\sigma_*E_2+2F$ will have the same genus and moduli space. \subsection{The $W_1$ Component} The remaining component $W_1$ is considerably more interesting. We start by noting that the projection $C_1=2R$, where $R$ is the exceptional curve corresponding to $E_1$. This implies that the projection map from $W_1$ to $R$ is a double cover. For the same reason as in the previous section, we have that the fibered space $S_R$ above $R$ satisfies \begin{equation} S_R \text{ is a $dP_9$ surface} \end{equation} However, the fact that $W_1$ is a double cover implies that, unlike all the cases we have considered thus far, $W_1$ is not a section. Hence, the $W_{1}$ moduli space is not given by $\mathcal{M}(0,1;n_{1})$. However, we can still analyze its moduli space using the general map between classes~\eqref{fulliR}. Suppose, for simplicity, we choose $n_1=2$, that is \begin{equation} [W_1] = 2E_1 + 2 F \label{eq:today1} \end{equation} since other cases can be analyzed in an analogous way. As above, we will not consider all possibilities for the class of $W_1$ in the $dP_9$. Instead, we will restrict our discussion to a few interesting examples. For instance, using~\eqref{fulliR} with $\Lambda=E_1$, some possibilities are \begin{subequations} \label{W1ex} \begin{align} {} [W_{1}]&= i_{R*}\left(2E'_1\right) \label{W1ex1}\\ {} [W_{1}]&= i_{R*}\left(E'_1+E'_2\right) \label{W1ex2}\\ {} [W_{1}]&= i_{R*}\left(l'-E'_1\right) \label{W1ex3}\\ {} [W_{1}]&= i_{R*}\left(3l'-E'_1-\dots-E'_7\right) \label{W1ex4} \end{align} \end{subequations} We can analyze the moduli spaces in the $dP_9$ surface of each of these different cases by considering the right-hand sides of~\eqref{W1ex} as classes of curves through some number of points in $\mathbf{P}^2$. In examples \eqref{W1ex1} and \eqref{W1ex2}, the curves are always reducible to just two copies of an exceptional curve in $dP_8$ and, hence, have no moduli for moving in $dP_{9}$. Since the curve $R$ is also fixed in the base, these cases have no moduli for moving in the Calabi--Yau manifold at all. The third case, \eqref{W1ex3}, corresponds to a line (topologically $\mathbf{P}^1$) in $\mathbf{P}^2$ through one point. As such, as discussed in section~\ref{exbase1}, its moduli space is $\mathbf{P}^1$. Furthermore, there are special points in the moduli space where the line passes through one of the other blown-up points in $dP_9$ and the curve becomes reducible (in analogy to the process described in Figure~\ref{lE1}). We can have, for instance, \begin{equation} W_1 = U_1 + U_2 \label{W1split} \end{equation} where the classes in $dP_9$ are \begin{equation} [U_1]_R = l' - E'_1 - E'_2 \qquad [U_2]_R = E'_2 \end{equation} What was previously a single sphere has now been reduced to a pair of spheres, each an exceptional curve, which intersect at a single point (as in Figure~\ref{Cpinch}). The last case, \eqref{W1ex4}, is even more interesting. This class corresponds to a cubic in $\mathbf{P}^2$ passing through 7 points. A cubic is an elliptic curve and so has genus one. A general cubic has a moduli space of $\mathbf{P}^9$. However, by being restricted to pass through seven points, the remaining moduli space is simply $\mathbf{P}^2$. At special points in the moduli space the cubic can degenerate. First, we can have a double point. This occurs when the discriminant of the curve vanishes. It corresponds to one of the cycles of the torus pinching, as shown in Figure~\ref{toruspinch}. The vanishing of the discriminant is a single additional condition on the parameters and so gives a curve, which we will call $\Delta_{W_1}$, in $\mathbf{P}^2$. When the cubic degenerates, the blown up curve is a sphere. Thus, it has changed genus. We can now go one step further. At certain places, the discriminant curve $\Delta_{W_1}$ in $\mathbf{P}^2$ has a double point. This corresponds to places where the curve becomes reducible. The curve $W_1$ splits into two \begin{equation} W_1 = U_1 + U_2 \label{eq:today2} \end{equation} with, for instance, $U_1$ and $U_2$ describing a line and a conic, \begin{equation} [U_1]_R = l' - E'_1 - E'_2 \qquad [U_2]_R = 2l' - E'_3 - \dots - E'_7 \end{equation} Note that each of the resulting curves is exceptional and so has no moduli space in the $dP_9$. However, we note that this splitting can happen $\binom{7}{2}=21$ different ways. The two curves intersect at two points, corresponding to a double pinching of the torus into a pair of spheres, as in Figure~\ref{toruspinch}. For completeness, we note that there is one further singularity possible. That is where the cubic develops a cusp. However, the topology remains that of a sphere, so we will not distinguish these points. \begin{figure} \centerline{\psfig{figure=toruspinch.eps,height=0.8in}} \caption{Pinching a torus at one point and at a pair of points} \label{toruspinch} \end{figure} We can summarize these branches of moduli space in the following table \begin{equation} \begin{array}{c|cc} [W_1] \text{ in $dP_9$} & \text{genus} & \text{moduli space} \\ \hline \hline 2E'_1 & 0+0 & \PP^1_a\times\PP^1_a/\mathbf{Z}_2 \\ \hline E'_1+E'_2 & 0+0 & \PP^1_a \times \PP^1_a \\ \hline l'-E'_1 & 0 & \left(\mathbf{P}^1-8\text{ pts.}\right) \times \PP^1_a \\ \left(l'-E'_1-E'_2\right) + \left(E'_2\right) & 0+0 & \PP^1_a \times \PP^1_a \\ \hline 3l'-E'_1-\dots-E'_7 & 1 & \left(\mathbf{P}^2-\Delta_{W_1}\right) \times \PP^1_a \\ 3l'-E'_1-\dots-E'_7 & 0 & \left(\Delta_{W_1}-21\text{ pts.}\right) \times \PP^1_a \\ \left(l'-E'_1-E'_2\right)+\left(2l'-E'_3-\dots-E'_7\right) & 0+0 & \PP^1_a \times \PP^1_a \end{array} \label{W1mod} \end{equation} Note that in the first line, we mod out by $\mathbf{Z}_2$ since the fivebranes wrap the same curve in the Calabi--Yau manifold and so are indistinguishable. This is not the case in the second example. We note that there is another type of splitting possible for the fourth example. The cubic could pass though one of the additional blown up points. It would then pass through eight of the nine points. However, as we have discussed above, it must then also pass through the ninth point. The curve $W_1$ would then decompose into a pure fiber component plus two sections, namely \begin{equation} W_1 = U_1 + U_2 + U_3 \end{equation} with \begin{equation} [U_1]_R = F_R = 3l'-E'_1-\dots-E'_9, \qquad [U_2]_R = E'_8, \qquad [U_3]_R = E'_9 \end{equation} Thus we see that, unlike the case for $W_2$ and $W_3$, it is also possible to have a transition where a fiber component splits off from $W_1$. Let us end this section with an important observation. We have not discussed the full moduli space of $W$ here but only various characteristic branches. There is, however, a certain type of branch that we would like to emphasize. Consider a branch where we choose $n=17$, so that $W_F=0$, and split the curve $W$ as follows \begin{equation} W = W_0 = W_1 + W_2 + W_3 \end{equation} with, for example \begin{equation} [W_1] = 2\sigma_*E_1 + 2F, \qquad [W_2] = \sigma_*E_2 + 5F, \qquad [W_3] = \sigma_*E_3 + 10F \end{equation} From our previous discussion, $W_2$ and $W_3$ are required to be sections of the $dP_9$ surface above the exceptional curve in the class $E_2$ and $E_3$ respectively. As such, they have no moduli to move within the Calabi--Yau threefold. Furthermore, we have seen an example in the second line of the table~\eqref{W1mod} where $W_1$ splits into two components, neither of which can move in the Calabi--Yau space. Consequently, we see that there is a component of moduli space where we simply have four fivebranes, each wrapping a fixed curve within the Calabi--Yau threefold. Furthermore, none of the fivebranes can intersect. We then have a very simple moduli space. It is \begin{equation} \PP^1_a \times \PP^1_a \times \PP^1_a \times \PP^1_a \end{equation} corresponding to moving each fivebrane in $S^1/Z_2$ and changing the value of the axions. Thus, even when the fivebrane class is relatively complicated, we see that there are components of the moduli space with very few moduli. \section{A second three-family $dP_8$ example} In this section, we will briefly discuss a second example of a realistic fivebrane moduli space. Again, we will take the base \begin{equation} B \text{ is a $dP_8$ surface} \end{equation} and choose the fivebrane class \begin{equation} [W] = \sigma_*l - \sigma_*E_1 + \sigma_*E_2 + \sigma_*E_3 + 27F \end{equation} This gives a three family model with an unbroken $SU(5)$ gauge group, as can be seen explicitly using the rules given in~\cite{dlow1} and~\cite{dlow2} (taking $\lambda=\frac{1}{2}$ in the equations given there). From the condition~\eqref{thrm}, since $l-E_1+E_2+E_3$ is effective in the base, $[W]$ is an effective class as required. To calculate the moduli space, first, one separates the pure fiber components, partitioning $W$ as $W=W_0+W_F$, as in~\eqref{decompF}, with \begin{equation} [W_0] = \sigma_*l - \sigma_*E_1 + \sigma_*E_2 + \sigma_*E_3 + nF, \qquad [W_F] = \left(27-n\right) F \end{equation} where $0\leq n\leq 27$. As usual, unless $n=27$, this implies that we have at least two distinct fivebranes. This partition splits the moduli space into 28 components, as in~\eqref{moddecomp}. The moduli space of the $W_F$ component is the usual symmetric product given in~\eqref{WFmod} \begin{equation} \mathcal{M}((27-n)F) = \left( F_r \times \PP^1_a\right)^{27-n}/\mathbf{Z}_{27-n} \label{WFex2} \end{equation} Next, we analyze the moduli space of $W_0$ for a given $n$. If we project $W_0$ onto the base, we get the curve $C$ in the class \begin{equation} [C] = l - E_1 + E_2 + E_3 \end{equation} We then need to find the moduli space $\mathcal{M}_B(l-E_1+E_2+E_3)$ of such curves in the base. The last two classes correspond to curves wrapping exceptional blow-ups and so, as in the previous section, $C$ must be reducible. We expect that there is always one component wrapping the exceptional curve in $E_2$ and another component wrapping the exceptional curve in $E_3$. In general, there are eight distinct parts of $\mathcal{M}_B(l-E_1+E_2+E_3)$, with different numbers of curves in the base. In general, $C$ decomposes into $k$ curves as \begin{equation} C = C_1 + \dots + C_k \end{equation} with \begin{equation} [C_i] = \Omega_i, \qquad \Omega_1 + \dots + \Omega_k = l - E_1 + E_2 + E_3 \end{equation} As $C$ splits, so does $W_0$. In general we can partition the $n$ fiber components in different ways, so that \begin{equation} W = W_1 + \dots + W_k, \qquad [W_i] = \sigma_*\Omega_i + n_iF \end{equation} where $n_i\geq 0$ and $n_1+\dots+n_k=n$. The eight different parts of the moduli space of $C$ can be summarized as follows \begin{equation} \begin{array}{cc|c} [C_i] & \genus{C_i} & \text{moduli space} \\ \hline\hline \begin{array}{c} {}[C_1] = l - E_1 \\ {}[C_2] = E_2 \\ {}[C_3] = E_3 \end{array} & \begin{array}{c} 0 \\ 0 \\ 0 \end{array} & \mathbf{P}^1 - 7\text{ pts.} \\ \hline \begin{array}{c} {}[C_1] = l - E_1 - E_2 \\ {}[C_2] = 2E_2 \\ {}[C_3] = E_3 \end{array} & \begin{array}{c} 0 \\ 0 \\ 0 \end{array} & \text{single pt.} \\ \hline \begin{array}{c} {}[C_1] = l - E_1 - E_3 \\ {}[C_2] = E_2 \\ {}[C_3] = 2E_3 \end{array} & \begin{array}{c} 0 \\ 0 \\ 0 \end{array} & \text{single pt.} \\ \hline \begin{array}{c} {}[C_1] = l - E_1 - E_i \\ {}[C_2] = E_2 \\ {}[C_3] = E_3 \\ {}[C_4] = E_i \end{array} & \begin{array}{c} 0 \\ 0 \\ 0 \\ 0 \end{array} & \begin{array}{c} \text{single pt.} \\ i=4,\dots,8 \end{array} \end{array} \label{modCtable} \end{equation} The analysis is very similar to that in section~\ref{exbase1}. The $C_2$ and $C_3$ components are always stuck on the exceptional curves in $E_2$ and $E_3$ and so have no moduli. In the first row in the table, $C_1$ is a curve in the $dP_8$ corresponding to a line through the blown-up point $p_1$. As such it has a moduli of $\mathbf{P}^1$, except for the seven special points where it also passes through one of the other seven blown-up points. If it passes through one of the points $p_4,\dots,p_8$, the curve splits into two curves and we have a total of four distinct fivebranes. This case is given in the last row in~\eqref{modCtable}. If it passes through $p_2$ or $p_3$, generically, we still only have three components, but now $C_2$ or $C_3$ is in the class $2E_2$ or $2E_3$ respectively. These are the second and third rows in~\eqref{modCtable}. All the curves are spheres so have genus zero. In conclusion, we have \begin{equation} \mathcal{M}_B(l-E_1+E_2+E_3) = \mathbf{P}^1 \end{equation} Where there are seven special points in the moduli space: the five cases given in the last row in~\eqref{modCtable}, where $C$ splits into four curves, and the two further points where the class of $C_2$ and $C_3$ changes as given in the second and third rows of~\eqref{modCtable}. We see that, unlike the previous example, $\mathcal{M}_B$ is more than just a single point. Let us concentrate on a generic point in moduli space, as in the first row of~\eqref{modCtable}. As we noted above, $W_0$ then splits into three distinct components, \begin{equation} W_0 = W_1 + W_2 + W_3 \end{equation} with \begin{equation} [W_1] = \sigma_*l - \sigma_*E_1 + n_1 F, \qquad [W_2] = \sigma_*E_2 + n_2 F, \qquad [W_3] = \sigma_*E_3 + n_3 F \label{ex2split1} \end{equation} where $n_1+n_2+n_3=n$ and $n_i\geq 0$. We note that $W_2$ and $W_3$ are exactly of the form we analyzed in section~\ref{W2W3} above. Recall that each component is a curve stuck above the exceptional curve $C_2$ or $C_3$ in the base. The space $S_{C_i}=\pi^{-1}(C_i)$, for $i=2,3$, above each exceptional curve was given by \begin{equation} S_{C_i} \text{ is a $dP_9$ surface for $i=2,3$} \end{equation} This means we have $g_i=0$ and $p_i=1$ and the corresponding moduli spaces are given by $\mathcal{M}(0,1;n_i)$. These spaces were given in~\eqref{cMm} and are a discrete number $N(n_i)$ of copies of $\PP^1_a$, corresponding to different sections of a $dP_9$ surface. We have \begin{equation} \mathcal{M}(0,1;n_i) = N(n_i) \PP^1_a \qquad \text{for $i=2,3$} \label{modW2W3} \end{equation} There are no moduli for moving the curves within either the base or the fiber of the Calabi--Yau manifold. Since $g_i=0$ there are no vector multiplets on the fivebranes. We now turn to $W_1$. This is essentially of the form we considered section~\ref{exbase2}. There, we showed that the surface $S_{C_1}$ is given by \begin{equation} S_{C_1} \text{ is a K3 surface} \end{equation} That is to say, the genus $g_1$ of $C_1$ is zero and $p_1=2$, so there are 24 singular fibers. Thus we are interested in the moduli space $\mathcal{M}(0,2;n_1)$. From the general discussion in section~\ref{generic}, we find from equation~\eqref{gencMR2} that the moduli space for $W_1$ is empty unless $n_1=0$. One then has \begin{equation} \mathcal{M}(0,2;0) = \PP^1_a \end{equation} Putting this together with the moduli space for $W_2$ and $W_3$ given in~\eqref{modW2W3}, and recalling the form of the $C$ moduli space~\eqref{modCtable}, we can write, for this generic part of the moduli space of $W_0$, \begin{multline} \mathcal{M}_0(\sigma_*l-\sigma_*E_1+\sigma_*E_2+\sigma_*E_3+nF) \\ = \left(\left[\mathbf{P}^1 - 7\text{ pts.}\right]\times\PP^1_a\right) \times N(n_2)\PP^1_a \times N(n_3)\PP^1_a + \mathcal{M}_{\text{non-generic}} \end{multline} where we must have $n_1=0$, and $n_2+n_3=n$ in the decomposition~\eqref{ex2split1}. We have three distinct fivebranes, two of which, $W_2$ and $W_3$ are stuck in the Calabi--Yau threefold. The third fivebrane can move within the base of the Calabi--Yau. Each fivebrane wraps a curve of genus zero and so there are no vector multiplets. The full $W$ moduli space is constructed, for each $n$, from the product of this space together with the $W_B$ moduli space given in~\eqref{WFex2}. We will not consider all the seven possible exceptional cases in the moduli space of $C$, listed in~\eqref{modCtable}. Rather, consider just one of the cases where $C$ splits into four components, \begin{equation} [C_1] = l - E_1 - E_4, \qquad [C_2] = E_2, \qquad [C_3] = E_3, \qquad [C_4] = E_4 \end{equation} with the corresponding split of $W_0$ as \begin{equation} \begin{aligned} {}[W_1] &= \sigma_*l - \sigma_*E_1 - \sigma_*E_4 + n_1 F, \quad & {}[W_2] &= \sigma_*E_2 + n_2 F, \\ {}[W_3] &= \sigma_*E_3 + n_3 F, & {}[W_4] &= \sigma_*E_4 + n_4 F \end{aligned} \end{equation} with $n_i\geq 0$ and $n_1+n_2+n_3+n_4=n$. Let us further assume that $n=27$ so that $W_F=0$ and there are no pure fiber components. Each of the $C_i$ is an exceptional curve in the base. Consequently, we have \begin{equation} S_{C_i} \text{ is a $dP_9$ surface for $i=1,2,3,4$} \end{equation} (For $C_1$ the calculation is just as in section~\ref{exbase2}). Thus we have $g_i=0$ and $p_i=1$ for each curve and so we have the moduli spaces \begin{equation} \mathcal{M}(0,1;n_i) = N(n_i)\PP^1_a \qquad \text{for } i = 1,2,3,4 \end{equation} where, as usual, $N(n_i)$ counts the number of distinct sections in each $dP_9$. In particular, we are no longer required to take $n_1=0$. We now have a total of four distinct fivebranes, wrapping $W_1$, $W_2$, $W_3$ and $W_4$. All these curves are stuck in the Calabi--Yau threefold, so that a given connected part of the moduli space has the form \begin{equation} \PP^1_a \times \PP^1_a \times \PP^1_a \times \PP^1_a \end{equation} (Here we have assumed that either $n_1$ or $n_4$ is non-zero so that this branch of moduli space if not connected to the generic branch discussed above.) As in the previous section, we see that there are disconnected components of the moduli space with very few moduli. Each curve has genus zero, so there are no vector multiplet degrees of freedom. \section{Two simple Hirzebruch examples} In this section, we will briefly discuss two simple examples where the base is a Hirzebruch surface $F_r$. These are $\mathbf{P}^1$ fibrations over $\mathbf{P}^1$, characterized by a non-negative integer $r$. Following the notation of~\cite{dlow2}, they have two independent algebraic classes, the class $\mathcal{S}$ of the section of the fibration at infinity and the fiber class $\mathcal{E}$. These have the following intersection numbers \begin{equation} \mathcal{S} \cdot \mathcal{S} = -r, \qquad \mathcal{S} \cdot \mathcal{E} = 1, \qquad \mathcal{E} \cdot \mathcal{E} = 0 \end{equation} The canonical bundle is given by \begin{equation} K_B = -2 \mathcal{S} - (2+r) \mathcal{E} \end{equation} Effective classes in $F_r$ are of the form $\Omega=a\mathcal{S}+b\mathcal{E}$ with $a$ and $b$ non-negative. Thus, from~\eqref{thrm}, a general class of effective curves in the Calabi-Yau threefold can be written as \begin{equation} [W] = a \sigma_*\mathcal{S} + b \sigma_*\mathcal{E} + f F \end{equation} where $a$, $b$ and $f$ are all non-negative. (Note, however, that, as discussed above equation~\eqref{thrm}, there is actually an additional effective class for $r\geq 3$, which we will ignore here.) For any realistic model with three families of matter and realistic gauge groups, if the hidden $E_8$ group is unbroken, then the coefficients $a$ and $b$ are typically large~\cite{dlow2}. The moduli space is then relatively complicated to analyze. Thus, for simplicity, we will consider only two very simple cases with either a single $\mathcal{S}$ class or a single $\mathcal{E}$ class. \subsection{$[W]=\sigma_*\mathcal{S}+fF$} We consider first \begin{equation} B \text{ is an $F_r$ surface with $r\geq 2$} \end{equation} and \begin{equation} [W] = \sigma_*\mathcal{S} + f F \end{equation} using the general procedure we outlined above. Note that we require $r\geq 2$ to exclude the trivial case of $F_0$, which is just the product $\mathbf{P}^1\times\mathbf{P}^1$, and $F_1$, which is actually the del Pezzo surface $dP_1$. We recall that the first step is to split off any pure fiber components from $W$, writing it as a sum of $W_0$ and $W_F$ as in~\eqref{decompF} \begin{equation} [W_0] = \sigma_*\mathcal{S} + n F, \qquad [W_F] = (f-n) F \label{cSpartition} \end{equation} with $0\leq n \leq f$. As usual, unless $n=f$, this implies we have at least two distinct fivebranes. This splits the moduli space into $f+1$ components as in~\eqref{moddecomp}. The moduli space of $W_F$ has the familiar form, from~\eqref{WFmod}, \begin{equation} \mathcal{M}((f-n)F) = \left( F_r \times \PP^1_a\right)^{f-n}/\mathbf{Z}_{f-n} \end{equation} Next we turn to analyzing the $W_0$ moduli space for a given $n$. Projecting $W_0$ onto the base gives the curve $C$ in the homology class \begin{equation} [C] = \mathcal{S} \end{equation} Our first step is then to find the moduli space $\mathcal{M}_B(\mathcal{S})$ of $C$ in the base. This, however, is very simple. Since the self-intersection of $\mathcal{S}$ is negative, there is a unique representative of the class $\mathcal{S}$, namely the section at infinity. Thus \begin{equation} \mathcal{M}_B(\mathcal{S}) = \text{single point} \end{equation} The second step is the to find all the curves $W_0$ in the class $\pi_*\mathcal{S}+nF$ in the Calabi--Yau threefold which project onto $C$. To answer this, we need to characterize the surface $S_C=\pi^{-1}(C)$. We recall that this will be an elliptic fibration over $C$ and is characterized by the genus $g$ of $C$ and the number $12p$ of singular fibers. Since $C$ is a section of the Hirzebruch surface, it must be topologically $\mathbf{P}^1$. Consequently \begin{equation} g = \text{genus}(C) = 0 \label{genuscS} \end{equation} The expected number number of singular fibers is given by $12p$, where \begin{equation} \begin{aligned} p &= - K_B \cdot [C] \\ &= \left(2\mathcal{S} + (2+r)\mathcal{E}\right) \cdot \mathcal{S} = 2 - r \end{aligned} \label{pcS} \end{equation} For $r\geq 3$ this seems to predict that we have a negative number of singular fibers. Actually, this reflects the fact that the curve $C$ is contained within the discriminant curve of the Calabi--Yau elliptic fibration. Thus the elliptic curve over $C$ is singular everywhere. This is a non-generic case we wish to avoid in discussing the moduli space. Thus, we will assume $r<3$. Given that $F_0$ and $F_1$ do not give new surfaces, we are left with restricting to $r=2$ and so \begin{equation} B \text{ is $F_2$} \end{equation} Then we have $p=0$ which implies the elliptic fibration is locally trivial. If we assume that the surface $S_C$ is also a globally trivial fibration, it is then simply the product \begin{equation} S_C = \mathbf{P}^1 \times E \end{equation} where $E$ is an elliptic curve. Comparing with the notation of section~\ref{gen}, we have $g=p=0$ and so we are interested in the moduli space $\mathcal{M}(0,0;n)$. However, we have argued,~\eqref{gencMR2}, that these spaces are empty unless $n=0$. In this case $\mathcal{M}(0,0;0)=E\times\PP^1_a$, where the first factor comes from moving the fivebrane within $S^1/Z_2$ and the second factor comes from moving it within $S_C$. (Again, here we are assuming that $S_C$ is a globally as well as locally trivial fibration.) In particular, the curve is a section of $S_C$ and, so, can lie at any point on the elliptic curve $E$. Thus, we see that the only partition~\eqref{cSpartition} allowed is where $n=0$. In that case the moduli space of the curve $W_0$ is simply \begin{equation} \mathcal{M}_0(\sigma_*\mathcal{S}) = E \times \PP^1_a \end{equation} The full moduli space is then the product \begin{equation} \mathcal{M}(\sigma_*\mathcal{S}+fF) = \left(E \times \PP^1_a\right) \times \frac{\left( F_2 \times \PP^1_a\right)^f}{\mathbf{Z}_f} \end{equation} Generically, we have $f+1$ fivebranes. One is $W_0$, which lies over the section at infinity in $F_2$ and can be at any position in the elliptic fiber over the base as well as in $S^1/Z_2$. The other $f$ fivebranes are pure fiber components, which can each be at arbitrary points in the base and in $S^1/Z_2$. Since the genus of $C$ is zero, so is the genus of $W_0$. Thus, generically, the only vector multiplets come from $W_F$. At a generic point in moduli space, we therefore have $U(1)^f$ gauge symmetry. It is possible for the fiber components to intersect $W_0$, in which case we might expect new massless fields to appear in the low-energy theory. \subsection{$[W]=\sigma_*\mathcal{E}+fF$} Now consider the case where again \begin{equation} B \text{ is a $F_r$ surface with $r\geq 2$} \end{equation} but we take \begin{equation} [W] = \sigma_*\mathcal{E} + f F \end{equation} As always, we first split off the pure fiber components in $W$, writing $W=W_0+W_F$ with $f+1$ different partitions \begin{equation} [W_0] = \sigma_*\mathcal{S} + n F, \qquad [W_F] = (f-n) F \label{cEpartition} \end{equation} The moduli space of $W_F$ is the familiar form \begin{equation} \mathcal{M}((f-n)F) = \left( F_r \times \PP^1_a\right)^{f-n}/\mathbf{Z}_{f-n} \end{equation} To analyze the the moduli space of $W_0$, we project onto the base, giving the curve $C$ with \begin{equation} [C] = \mathcal{E} \end{equation} Thus $C$ is in the fiber class of the $F_n$ base. Since $F_n$ is a $\mathbf{P}^1$ fibration over $\mathbf{P}^1$, the fiber can lie at any point in the base $\mathbf{P}^1$ so we have that the moduli space of $C$ is given by \begin{equation} \mathcal{M}_B(\mathcal{E}) = \mathbf{P}^1 \label{cEbase} \end{equation} Next we need to find all the curves $W_0$ which project onto a given $C$. We first characterize the surface $S_C=\pi^{-1}(C)$. Since $C$ is a fiber of $F_r$, it must be topologically $\mathbf{P}^1$. Hence \begin{equation} g = \text{genus}(C) = 0 \end{equation} The number of singular fibers, $12p$, in the elliptic fibration $S_C$ given in terms of \begin{equation} \begin{aligned} p &= -K_B \cdot [C] \\ &= \left(2\mathcal{S}+(2+r)\mathcal{E}\right) \cdot \mathcal{E} = 2 \end{aligned} \end{equation} for any $r$. This implies that $S_C$ is an elliptic fibration over $\mathbf{P}^1$ with 24 singular fibers, that is \begin{equation} S_C \text{ is a K3 surface} \end{equation} In the notation of section~\ref{gen}, we have $g=0$ and $p=2$ and so we are interested in the moduli space $\mathcal{M}(0,2;n)$. However, we have argued,~\eqref{gencMR2}, that these spaces are empty unless $n=0$, in which case $\mathcal{M}(0,2;0)=\PP^1_a$. That is, the curve is completely stuck within $S_C$, but is free to move within $S^1/Z_2$. Thus the only partition~\eqref{cEpartition} allowed is where $n=0$. In this case, the full moduli space of $W_0$ is then given by \begin{equation} \mathcal{M}_0(\sigma_*\mathcal{E}) = \mathbf{P}^1 \times \PP^1_a \end{equation} The first factor of $\mathbf{P}^1$ reflects the moduli space of curves $C$ in the base~\eqref{cEbase}. The second factor reflects the fact that, for a given $C$, the moduli space of $W_0$ is $\mathcal{M}(0,2;0)=\PP^1_a$. We note that there are no moduli for moving $W_0$ in the direction of the elliptic fiber. The full moduli space is then given by the product \begin{equation} \mathcal{M}(\sigma_*\mathcal{E}+fF) = \left(\mathbf{P}^1 \times \PP^1_a\right) \times \frac{\left( F_r \times \PP^1_a\right)^f}{\mathbf{Z}_f} \end{equation} As in the previous example, generically we have $f+1$ fivebranes. One is $W_0$ which, in the base, can be deformed in the base and can move in $S^1/Z_2$, but has no moduli for moving in the elliptic fiber. The other $f$ fivebranes are pure fiber components, which can each be at arbitrary points in the base and in $S^1/Z_2$. Again, since the genus of $C$ is zero, so is the genus of $W_0$. Thus, generically, the only vector multiplets come from $W_F$. At a generic point in moduli space we have, therefore, $U(1)^f$ gauge symmetry. It is possible for the fiber components to intersect $W_0$, in which case we might expect new massless fields to appear in the low-energy theory. \subsection*{Acknowledgments} B.A.O. and D.W. would like to thank Angel Uranga for helpful discussions. R.D. is supported in part by an NSF grant DMS-9802456 as well as by grants from the University of Pennsylvania Research Foundation and Hebrew University. B.A.O. is supported in part by a Senior Alexander von Humboldt Award, by the DOE under contract No. DE-AC02-76-ER-03071 and by a University of Pennsylvania Research Foundation Grant. D.W. is supported in part by the DOE under contract No. DE-FG02-91ER40671.
\section{Introduction} The unusual and peculiar SN1998bw was (\cite{K98}) the most luminous radio supernova ever observed. Its visible properties were also extraordinary (\cite{G98}); a reported expansion speed approaching 60,000 km/s is larger than that of other SN. Modeling (\cite{I98}) indicates a kinetic energy of $\sim 3 \times 10^{52}$ ergs, more than an order of magnitude greater than that of most supernovae. Perhaps most remarkable was the apparent association of SN1998bw with the unusual GRB980425, leading to the suggestion (\cite{B98}) of a new class of supernova-gamma-ray burst. The statistical significance and reality of this association have been disputed (\cite{GLM99}), but the supernova is extraordinary even without an association with a GRB. \cite{K98} have argued on the basis of the absence of interstellar scintillation and bounds on the radio brightness temperature that the radio source of SN1998bw expanded semi-relativistically, with a Lorentz factor $\Gamma$ in the range 1.6--2. This value may be a clue to understanding SN1998bw and its radio emission. \cite{WL98} argued that the expansion was in fact much slower, but this has not been generally accepted, and I will adopt, at least approximately, the estimates of \cite{K98}. A possible origin of the semi-relativistic expansion as a result of electrons Compton-scattered by $^{56}$Ni and $^{56}$Co gamma-rays is suggested in \S 2. In \S 3 I estimate the magnetic field in the source region implied by the synchrotron self-absorption interpretation of the radio spectrum of SN1998bw and discuss the origin of the field. In \S 4 I mention possible mechanisms for achieving electron-ion and particle-field equipartition. In \S 5 I ask if the radio spectrum may in fact be the result of inverse bremsstrahlung by a thermal gas; although possible, this does not lead to drastically lower estimates of the field than those presented in \S 3. \S 6 questions the association of SN1998bw with GRB980425. \S 7 contains a summary discussion and predictions. \section{Compton Electrons} \cite{I98} inferred, on the basis of its visible light curve, that SN1998bw contained about $0.7 M_\odot$ of $^{56}$Ni, or $N_{56} = 1.5 \times 10^{55}$ atoms. The 6.10 d half life of $^{56}$Ni (an $e$-folding decay time of 8.8 d) is comparable to the duration of the first peak of radio intensity found by \cite{K98}. The daughter nucleus $^{56}$Co decays with a half life of 77 d (an $e$-folding decay time of 111 d), comparable to the rise and decay time scale of the second, lower frequency, peak of radio emission in SN1998bw. The two characteristic time scales of radio emission and the double-peaked intensity observed at some wavelengths suggest two distinct, but analogous, processes. In addition, the observation of expansion with Lorentz factor $\Gamma$ in the approximate range 1.6--2 calls for explanation. Is there is a natural explanation of the observation of mildly relativistic expansion with this particular value of $\Gamma$? The decay of $^{56}$Ni (which is by electron capture) produces gamma-rays of 0.788 MeV, 0.812 MeV and 1.56 MeV, which are emitted in 48\%, 85\% and 14\% of the decays, respectively (\cite{LHP67}). The decay of $^{56}$Co, 80\% by electron capture, produces gamma-rays of several energies, of which the most important are 0.847 MeV (100\%), 1.04 MeV (15\%), 1.24 MeV (66\%), 1.76 MeV (15\%), 2.02 MeV (11\%), 2.60 MeV (17\%) and 3.26 MeV (13\%); 20\% of its decays are by e$^+$ emission, with an energy distribution extending up to 1.5 MeV. This copious production of gamma-rays offers a possible explanation of the existence and expansion speed of the radio source and perhaps of its magnetic field. It is necessary to assume that the $^{56}$Ni is mixed to the surface of the debris; perhaps a jet or plume of $^{56}$Ni-rich material penetrates any envelope and is expelled. The absence of H or He in the spectrum of SN1998bw establishes the absence at least of a massive envelope cloaking the high-Z material. After $10^6$ s at the observed photospheric velocity of $6 \times 10^9$ cm/s this matter is in a shell of column density $\approx 3$ gm/cm$^2$. If mixed with an equal quantity of other material, to make up a Chandrasekhar mass of debris, the total column density is $\approx 6$ gm/cm$^2$, which is transparent (optical depth $\approx 0.5$) to 0.812 MeV gamma-rays, the most important emission of $^{56}$Ni. After the gamma-rays emerge from the supernova debris they enter any surrounding medium, moving outward at the speed of light. Their source gradually becomes transparent to them as it expands, and their emergent intensity increases with a rise time of a few days, until the competition between increasing transparency and radioactive decay leads to a peak at $\approx 10$ d, corresponding to the first peak in radio intensity. Averaging over the Klein-Nishina cross-section, the mean electron produced by Compton scattering of a 0.812 MeV gamma-ray has $\gamma = 1.65$ and the most energetic has $\gamma = 2.12$. The stopping column density of a $\gamma = 1.65$ (kinetic energy $E_C = 330$ KeV) electron in dilute ($n_e \sim 10^4$ cm$^{-3}$; the dependence on $n_e$ is only through the Coulomb logarithm and is very weak) ionized plasma is $\ell_s \approx 8 \times 10^{21}$ cm$^{-2}$ (\cite{L80}). Comparing this to the Klein-Nishina cross-section $\sigma_{KN} = 2.6 \times 10^{-25}$ cm$^{-2}$ yields an estimated efficiency of conversion of gamma-rays to Compton electrons $\epsilon = \ell_s \sigma_{KN} \approx 2 \times 10^{-3}$; the remaining energy appears as thermal heating of the plasma by Compton electrons which stop within it. If there is less than a stopping length of low-Z medium surrounding the $^{56}$Ni then the efficiency will be reduced because in traversing the high-Z debris the Compton electrons suffer Coulomb scattering by the nuclei as well as energy losses to the electrons; the scattering length in pure $^{56}$Ni is only $\approx 10^{19}$ cm$^2$ (\cite{S62}). A hydrogen-rich shell of column density $10^{22}$ cm$^{-2}$ (requiring only $4 \times 10^{-3} M_\odot$ at a radius of $6 \times 10^{15}$ cm, the distance traveled by the massive debris in $10^6$ s) would be sufficient to regenerate the full flux of Compton electrons and to restore the efficiency to the value of the preceding paragraph. As discussed above, supernovae are expected to be surrounded by the remains of the winds of their progenitors. These Compton electrons move outward with a mean speed $v_e \approx 0.8 c$ and current density ${\vec j}_{Compt}$. Once they have expanded to a radius $> 2$ times that at which they were born they move essentially radially outward, which may be shown either by considering them as an adiabatically expanding gas (non-radial components of motion are, in effect, random thermal velocities, soon converted to ordered outward motion) or as individual ballistic particles. This free expansion requires a denser ambient medium in order to provide a return (counter-) current density ${\vec j}_{cc}$ to maintain electrical neutrality. The free-streaming density of Compton electrons $$n_{Ce} = {\epsilon N_{56} \nu_{56} \exp{(-\nu_{56} \Delta t)} \over 4 \pi r^2 v_e}, \eqno(1)$$ where $\nu_{56}$ is the radioactive decay rate ($1.3 \times 10^{-6}$ s$^{-1}$ for $^{56}$Ni) and $\Delta t$ the time between creation of the $^{56}$Ni and the decays whose Compton-scattered electrons are observed; the nonrelativistic motion of the $^{56}$Ni and the gamma rays' path between production and Compton scattering are neglected. Adopting $\Delta t = 7$ d and $r = 3 \times 10^{16}$ cm yields $n_{Ce} \approx 3 \times 10^4$ cm$^{-3}$, corresponding to $\approx 3 \times 10^{-3} M_\odot$ of hydrogen, similar to values estimated elsewhere in this paper. A medium even slightly denser than this, such as plausibly produced by loss of the progenitor's envelope over $\sim 10^3$ y, would supply the countercurrent required by charge neutrality with only a small cost in energy to the Compton electrons (the potential required to drive the countercurrent retards the Compton electrons, but only slightly if the countercurrent density is significantly larger and velocity significantly less). After the $^{56}$Ni decays, $^{56}$Co produces its own gamma-rays, Compton electrons and positrons. The combination of positrons and Compton electrons permits a neutral mildly relativistic wind even in the absence of a background plasma to provide a countercurrent, should there be none. In other respects, the effects of $^{56}$Co decays are similar to those of $^{56}$Ni decays, although the $^{56}$Co gamma-rays and Compton electrons are more energetic (the abundant 1.24 MeV gamma-ray produces a mean Compton $\gamma_e = 2.14$ and the gamma-ray spectrum extends up to 3.26 MeV). This new wave of Compton electrons will eventually outrun the $^{56}$Ni Compton electrons and produce a second peak of radio emission, as observed. \section{Magnetic Field} \subsection{Self-Absorbed Spectrum} Inspection of the radio spectrum of SN1998bw (\cite{K98}) shows evidence of self-absorption. The characteristic self-absorption frequency is $\approx 5$ GHz 10 d after the supernova, declines to $\approx 2$ GHz after 30 d, and continues to decline as the source fades thereafter. This accords with expectations for an expanding self-absorbed synchrotron source. The low frequency (self-absorbed) flux first rises as the radiating area expands. The higher frequency flux declines as a consequence of declining magnetic field or electron energies. It is possible to fit simple models to the data, but the observed (\cite{K98}) double-peaked time dependence of the flux at most frequencies implies that simple models will not be satisfactory. The brightness temperature $T_B^\prime$ in the emitting frame may be obtained directly from the observed intensity, if the size of the radiating region is known, and \cite{K98} used such brightness temperatures, combined with energetic arguments, to infer the rate of expansion of the self-absorbed source: $$k T_B^\prime(\nu^\prime) = {S_\nu d^2 \over 2 \pi b^2 t^2 \nu^2 D}, \eqno(2)$$ where $k$ is Boltzmann's constant, $S_\nu$ is the observed flux density, $d = 38$ Mpc is the distance (assuming a Hubble constant of 65 km/s/Mpc), $\nu$ is the frequency of observation, $t$ is the elapsed time since expansion began, $bc$ is the apparent velocity of expansion ($b = \Gamma \beta$ where $\Gamma$ and $\beta c$ are the Lorentz factor and expansion speed), $D = [\Gamma(1 - \beta \cos\theta)]^{-1} \sim \Gamma$ and $\nu^\prime = \nu/D$ where $\nu$ is the observed frequency and $\nu^\prime$ the frequency in the comoving frame. The brightness temperature of a self-absorbed source will generally approximate the energy of the radiating particles (in thermal equilibrium it equals the particle temperature). Hence the radiating electrons have a Lorentz factor in the co-moving frame $$\gamma_e^\prime \approx {S_\nu d^2 \over 2 \pi b^2 m_e c^2 t^2 \nu^2 D}. \eqno(3)$$ \subsection{Required Field} The co-moving magnetic field $B^\prime$ may be estimated from (3) using the synchrotron radiation relation $B^\prime \approx 2 \pi m_e c \nu^\prime/e \gamma_e^{\prime\,2}$: $$B^\prime = {8 \pi^3 m_e^3 c^5 \nu^5 b^4 t^4 D^2 \over e S_\nu^2 d^4}. \eqno(4)$$ The values of $B$ inferred for SN1998bw are remarkably high. For mildly relativistic motion approximate $D \approx 1$, $b \approx \beta$ and $B \approx B^\prime$. The numerical values (\cite{K98}) $\nu = 2.49$ GHz (nominally $\lambda = 13$ cm), $t = 1.01 \times 10^6$ s (11.7 d), $S_\nu = 19.7$ mJy and $d = 38$ Mpc yield $$B \approx 0.13 \beta^4\ {\rm gauss}. \eqno(5)$$ These large inferred fields are a consequence of the low $T_b^\prime$ (and $T_b$) implied by a rapidly expanding source of large linear size. The low brightness temperature combined with the interpretation of self-absorption implies radiation by electrons of comparatively low energy, and hence a high magnetic field is required to produce radiation of the observed frequency. \cite{WL98} suggest $\beta \approx 0.3$. This would imply a substantially lower value of $B$ than for mildly relativistic motion (their numerical estimates are somewhat larger than that of (5) because of differences in various details). \cite{K98} dispute such low values of $\beta$, and this paper does not attempt to resolve this issue. Is the magnetic field (5) plausible, extended over a region of size $r \approx \Gamma \beta c t \approx 3 \times 10^{16} b$ cm$^2$? The implied magnetic energy is $${\cal E}_B = {D^2 B^{\prime\,2} r^3 \over 6} \sim 8 \times 10^{46} \beta^{11} \Gamma^{11} D^4\ {\rm ergs}. \eqno(6)$$ This is modest if the expansion is not relativistic ($\Gamma \approx D \approx 1$). In that case \cite{WL98} point out that ${\cal E}_B$ is several orders of magnitude less than the kinetic energy of SN1998bw and the energy of the radiating electrons. Only if $\Gamma$ exceeds 2 does ${\cal E}_B$ approach the kinetic energy of the supernova debris. \subsection{Sources of Field} \subsubsection{Flux} Energy is not the only relevant criterion. If a magnetic field's dependence on distance can be described by a power law $B \propto r^{-\alpha}$ three simple models may be considered. A static dipole field has $\alpha = 3$; this is clearly inadequate. If the field is frozen into a conducting outflow either from the progenitor or from the SN itself, then flux is conserved and $\alpha = 2$. In this case the apparent inferred flux $$\Phi_m \approx 10^{32} \beta^6\ {\rm gauss\,cm}^2; \eqno(7)$$ this flux is only apparent because it is $\int \vert {\vec B} \vert\,dS$ rather than $\int {\vec B} \cdot {\vec {dS}}$. For comparison, the flux of a typical pulsar is $\sim 10^{24}$ gauss-cm$^2$, that of a (perhaps-hypothetical) ``magnetar'' is $\sim 10^{27}$ gauss-cm$^2$ and that of the Sun is $\sim 10^{23}$ gauss-cm$^2$. The magnetic fields of SN progenitors are not directly measured but it is clear that the estimate (7) is excessive, and that such a flux cannot be produced by a flux-conserving flow. If this model is to be salvaged there must be another source of field. \subsubsection{Hydrodynamic Dynamo Fields} When supernova debris runs into a surrounding medium the contact discontinuity between the two shocked fluids is generally hydrodynamically unstable and may amplify pre-existing fields by a turbulent dynamo mechanism, but this is not readily quantified. Such a medium and shocks are not necessary parts of a supernova model, but will occur if the supernova is surrounded by the remains of a wind produced by its progenitor (\cite{B99}), as is assumed elsewhere in this paper. This is the only mechanism in which the field is powered by the hydrodynamic energy of the supernova, although only a small fraction of this energy is available unless the surrounding medium is as massive as the fast debris. \subsubsection{Radiation Fields} Another possibility is the field of a propagating electromagnetic wave, for which $\alpha = 1$. Because the field alternates in direction with a short wavelength the actual flux is not large, even though the field and apparent flux may be large. The magnetic field at a distance $r \gg c/\omega$ from a source of rotational frequency $\omega$ and dipole moment $\mu$ is $$B \sim {\mu \omega^2 \over c^2 r}. \eqno(8)$$ For a magnetic neutron star ($\mu \sim 10^{30}$ gauss-cm$^3$) rotating at breakup ($\omega \sim 1.5 \times 10^4$ s$^{-1}$) $B \sim 10$ gauss at $r = 3 \times 10^{16}$ cm. This is certainly more than sufficient, and allows for smaller $\omega$ or $\mu$. An electromagnetic wind from a neutron star (or magnetized accretion disc) would be required to emerge through the dense debris implied by the visible light curve of the supernova. This is in contrast to a GRB in which the presence of relativistic outflow implies the absence of dense debris, at least in directions in which gamma-rays are observed. If the supernova debris is confined to dense filaments then the electromagnetic wind may penetrate between these filaments, as in the much older Crab Nebula. If the magnetic field has its origin in the radiation of a new pulsar, then (8) determines $\mu \omega^2$. The spindown power is $$P = {2 \over 3}{(\mu \omega^2)^2 \over c^3} \sim {2 \over 3}{B^2 r^2 c} \sim 10^{40}\ {\rm erg/s}. \eqno(9)$$ In principle, such a pulsar may be observed once the nonrelativistic debris becomes transparent. Eq.~(8) may be applied to other astronomical objects. For example, a similar estimate has been used to explain the field of the Crab nebula, in which there is a pulsar of known properties, and no dense gas intervening between it and the synchrotron emitting nebula. Accretion discs around black holes in AGN and extragalactic double radio sources are another application; the rotating magnetized disc implies an oscillating magnetic dipole moment and radiation. If the radiation is beamed into an angle $\Omega$ then (8) should be replaced by \begin{eqnarray} B \sim {\mu \omega^2 \over c^2 r} \left({4 \pi \over \Omega} \right)^{1/2} &\sim& \left({B_d \over 10^4\,{\rm gauss}}\right) \left({M \over 10^8 M_\odot}\right) \left({r \over 300\,{\rm Kpc}}\right)^{-1} \left({\Omega \over 10^{-2}\,{\rm sterad}}\right)^{-1/2}\ 3\,\mu{\rm gauss}\\ &\sim& \left({L \over 10^{46}\,{\rm erg/s}}\right)^{1/2} \left({r \over 300\, {\rm Kpc}}\right)^{-1} \left({\Omega \over 10^{-2}\,{\rm sterad}} \right)^{-1/2}\ 3\,\mu{\rm gauss}, \end{eqnarray} where $B_d$ is the disc field and the last relation is obtained assuming only that the disc viscosity is magnetic (\cite{K91}) and $L$ is the accretional power. In double radio sources equipartition is also suggested if the pressure of the relativistic wind is balanced by that of the electrons accelerated as its fields reconnect. \subsubsection{Compton Current Fields} Electrostatic neutrality requires ${\vec \nabla}\cdot{\vec j}_{Compt} = - {\vec \nabla}\cdot{\vec j}_{cc}$ to high accuracy. It does not require ${\vec j}_{Compt} = - {\vec j}_{cc}$, and if the flow is not spherically symmetric the latter equality is not likely to hold. As a result, the Compton electrons may create closed loops of net current, with resulting magnetic fields. One characteristic value of the field is obtained from Amp\`ere's Law: $B \sim \epsilon N_{56} \nu_{56} \exp{(-\nu_{56} \Delta t)} e/(rc) \sim 10^{10}$ gauss! It is evident on energetic grounds that fields of this magnitude cannot be created; magnetic forces will induce countercurrents which will cancel the divergence-free part of the Compton current (as well as its divergence) to high accuracy. However, this cancellation will not be exact. Just as electrostatically-driven countercurrents leave a net potential (which drives them) of some fraction of the Compton electron energy, magnetic countercurrents may leave a net current sufficient to produce a field whose energy approaches equipartition with that of the Compton electrons. This energy is $\sim \epsilon N_{56} \exp{(-\nu_{56} \Delta t)} E_c \sim 10^{46}$ ergs, corresponding to $B \sim 0.04$ gauss for $r \sim 3 \times 10^{16}$ cm. This is comparable to that suggested by the synchrotron self-absorption argument, and offers a possible explanation of the apparent deviation of the magnetic energy from energetic equipartition---equipartition is, in fact, achieved, but with the Compton electrons' energy rather than with the bulk hydrodynamic energy. \subsubsection{Plasma Instability Fields} A related hypothesis attributes the magnetic field to electromagnetic plasma instabilities resulting form the interpenetration of Compton- and counter-currents. In this case the field is chaotic on fine scales rather than ordered, but the possible field energy is again comparable to (but somewhat less than) the energy of the Compton electrons. \section{Equipartition?} \subsection{Particle-field Equipartition} \cite{K98} assume equipartition between magnetic and particle energies in their theoretical argument for mildly relativistic expansion. There are two classical arguments for this assumption. The first is based on attempts to calculate the generation of magnetic fields by dynamo mechanisms. This argument comes in as many forms as there are theories of dynamos, but it generally concludes (or assumes) that when the magnetic energy density becomes a significant fraction of the kinetic energy density its back-reaction on the motion will suppress further dynamo activity. If the particle energy density is similarly limited by the hydrodynamic energy density (\cite{K91}) then rough particle-magnetic equipartition will be achieved. Clearly, without detailed understanding of the dynamo and acceleration processes in any particular configuration this argument is very approximate, or perhaps only suggestive, but it does appear to be crudely correct for the interstellar medium and (excluding the energetic particles) for the Solar convective motion. The second classical argument for particle-field equipartition notes that for a given total energy the synchrotron power radiated will be maximized if equipartition obtains. Astronomical surveys are generally flux-limited, implying that most detected sources will be fairly close to equipartition, if it can be achieved. It is unclear if either of these arguments applies to SN1998bw. There may be opportunity for Rayleigh-Taylor instability at the contact discontinuity between the forward and reverse shocks when the debris encounters the surrounding medium (likely a wind ejected by the SN progenitor; {\it cf.} {\cite{B99}), but this depends on the details of the hydrodynamics. Departures from spherical symmetry may lead to Kelvin-Helmholtz instability. It is unknown if, or how effectively, these processes may produce dynamo field amplification. The radio counterpart of SN1998bw was detected by a search targeted on this unusual visible and gamma-ray object, and was radio flux-limited only in the implicit sense that its detection was limited by the radio sensitivity. \subsection{Electron-ion Equipartition} Electron-ion equipartition presents a question entirely distinct from that of particle-magnetic field equipartition. All GRB models require at least an approximation to electron-ion equipartition in order to couple the ion kinetic energy to the electrons which radiate. \cite{WL98} require this too, and in fact their assumed velocity fairly approximates that required (assuming a composition of helium or heavier elements) to produce their estimated electron energies if equipartition occurs. The mechanism of electron-ion equipartition in a collisionless shock, relativistic or non-relativistic may be as simple as an electrostatic double-layer with a potential sufficient to slow the ions as implied by the shock jump condition (\cite{K94}). If cold electron and ion streams penetrate an orthogonal magnetic field their differing gyroradii would lead to a charge separation which can only be avoided by the presence of a potential sufficient to equalize the gyroradii, which (in the relativistic limit) equally divides the kinetic energy between electrons and ions (in the nonrelativistic limit the electron kinetic energy exceeds that of the ions by the ion/electron mass ratio). This is an oversimplification because it assumes the magnetic stress greatly exceeds the hydrodynamic stress, but the qualitative justification for equating gyroradii (and energies, in the relativistic regime)---that only if the gyroradii are equal will electrostatic neutrality be maintained when shocks occur in magnetic fields---may be valid. If the gyroradii are unequal an electrostatic potential will develop which will equalize them. \section{Inverse Bremsstrahlung?} The interpretation of the low-frequency turnover in the radio spectrum of SN1998bw as synchrotron self-absorption led to the inference of a magnetic field too large to be easily explained. An alternative explanation for this turnover is absorption by inverse bremsstrahlung, as suggested by \cite{C82}. Inverse bremsstrahlung absorption will not, in general, produce the quantitative $F_\nu \propto \nu^2$ or $F_\nu \propto \nu^{5/2}$ spectra of synchrotron self-absorption, but when the low frequency turnover is only defined by two spectral points of moderate accuracy, as was the case for SN1998bw (\cite{K98}) it is impossible to distinguish between these two explanations on spectral grounds alone. If the low frequency turnover is the result of inverse bremsstrahlung absorption then the observed brightness temperature is only a lower bound on the brightness temperature in the (optically thin) emitting region. This, in turn, is only a lower bound on the energy of the emitting electrons. Hence the magnetic field is only bounded from above, and may be as small as required by a theory of field generation. The actual field value depends on the electron energy (or {\it vice versa}); in the absence of a detailed theory of particle acceleration it is generally impossible to reject an inferred value of the electron energy. High electron energies need not, however, imply excessive brightness temperatures because the source is optically thin. Hence, as argued by \cite{K98}, the synchrotron self-Compton catastrophe may be avoided if the source is expanding relativistically, which reduces the inferred brightness temperature, as first pointed out by \cite{W66}. If the magnetic field is too small then the ratio of inverse Compton luminosity $L_{IC}$ to synchrotron $L_S$ becomes excessive. The value of $L_{IC}$ permitted by the observations depends on its (unknown) frequency, but it probably safe to require it to be less than the visible luminosity $L_V \sim 10^{43}$ erg/s, roughly $3 \times 10^4$ times the radio power at $10^6$ s (\cite{K98}). This sets a lower bound on $B$: $$B \approx \left({2 L_V \over r^2 c}{L_S \over L_{IC}}\right)^{1/2} > 0.04\ {\rm gauss}, \eqno(11)$$ where the numerical value has assumed only mildly relativistic expansion. The numerical result is close enough to the estimates derived for the self-absorbed synchrotron model that discarding this assumption has not materially reduced the difficulty of explaining the required field. If the synchrotron source is expanding relativistically then $r \approx c t \Gamma^2$ and the visible radiation intensity in the frame of the synchrotron source is $\sim L_V/(4 \pi r^2 \Gamma^2) \propto \Gamma^{-6}$, where the two extra powers of $\Gamma$ come from the redshift of the visible radiation, and the lower bound on the comoving $B^\prime$ is reduced $\propto \Gamma^{-3}$. The bound on the laboratory frame $B$ is only reduced $\propto \Gamma^{-2}$. For mildly relativistic expansion, such as inferred for SN1998bw, the required $B$ remains large. In addition, because of the larger inferred $r$, the various scaling laws yield estimates of of $B$ which are reduced by factors $\propto \Gamma^{-2\alpha}$. The difficulty is not resolved, and, in fact, is worsened for $\alpha > 1$. The inverse bremsstrahlung opacity of hydrogen, including the effects of stimulated emission, is $$\kappa_{ff} = 3.69 \times 10^8 {n^2 h \over k T^{3/2} \nu^2} g_{ff}\ {\rm cm}^{-1} \eqno(12)$$ (\cite{S62}). Taking $T = 10^{4\,\circ}$K and $\nu = 2.49 \times 10^9$ Hz (for which $g_{ff} \approx 5$) yields $\kappa_{ff} \approx 1.2 \times 10^{-26} n^2$ cm$^{-1}$. Effective absorption will be obtained in $3 \times 10^{16}$ cm if $n = 10^5$ cm$^{-3}$, requiring $\sim 10^{-2}$ M$_\odot$ of gas. Such a circum-SN evelope, produced by a wind from the progenitor star, is possible; if expelled at 10 km/s (as appropriate to a red supergiant) its lifetime is $\sim 10^3$ yr. Much more massive winds are plausible. Such winds have been widely discussed since the discovery of ring structures around SN1987A, and have been inferred in other supernovae ({\it cf.} \cite{B99}). It is, of course, necessary that the envelope be ionized in order for inverse bremsstrahlung to occur. The envelope consists of $\sim 10^{55}$ atoms, which require $\sim 2 \times 10^{44}$ ergs of ionizing ultraviolet radiation to ionize. The flux of SN1998bw in the Lyman continuum is not directly observable, but the required energy is only $\sim 10^{-5}$ of the visible radiation emitted, which is certainly plausible and will be found for a Planck spectrum with $T > 9000^\circ$ K. The resulting temperature of the photoionized gas depends on the shape of the ultraviolet spectrum, but for a comparatively cool spectrum it will be $\approx 10^{4\,\circ}$K, as assumed. A dense and comparatively cool ionized gas is a source of recombination line radiation. A straightforward estimate using the preceding parameters and standard theory (\cite{O74}) leads to unobservably small ($\sim 10^{-5}$\AA) equivalent widths, even for the Balmer lines, so that it is not possible to test the hypothesis of inverse bremsstrahlung absorption in this manner. As the ionized envelope is dispersed by the supernova debris (once shock-heated its absorption coefficient decreases greatly because of the temperature dependence of $\kappa_{ff}$) the inverse bremsstrahlung absorption will decrease. This can qualitatively account for the evolution of the radio spectrum, whose early deficiency of longer wavelength flux gradually disappears. Quantitative modeling might, in principle, distinguish these predictions from those of self-absorption models, but all models contain several free parameters; together with the complexity of the observed radio spectral behavior (\cite{K98}) this makes an unambiguous discrimination between the models difficult. \section{GRB980425?} GRB980425, if produced by SN1998bw, poses another problem. The early states of a stellar explosion would be expected to be very optically thick, and to produce roughly a black body spectrum. The observed spectrum (\cite{B98}), if fitted to a black body, suggests a temperature $\sim 30$ KeV. Combining this with the inferred power (\cite{G98}) of $5.5 \times 10^{46}$ erg/s implies an emitting radius (assumed spherical) of $7 \times 10^7$ cm. However, the observed expansion speed of $6 \times 10^9$ cm/s (\cite{K98}) means that this radius will be exceeded about 0.01 s after the explosion begins, inconsistent the observed GRB duration of about 10 s. The power and duration are consistent if the temperature is $\sim 1$ KeV, but this is completely inconsistent with the observed spectrum. Matter expelled from a supernova core with temperature $\sim 30$ MeV at $r \sim 6 \times 10^6$ cm will cool adiabatically to $\sim 3$ KeV, much less than the value estimated from the observed spectrum, by the time it reaches $6 \times 10^{10}$ cm, the radius at the period of gamma-ray emission. In addition, a radiating photosphere will be much cooler than the temperature deep within the outflow, for which adiabatic expansion is the only cooling process. The difficulty of explaining the observed properties of GRB980425 in the context of a supernova may best be resolved, despite the apparent coincidence, if they are in fact unrelated events, as suggested on statistical grounds (\cite{GLM99}). This is supported by the observation (\cite{P99}) of two transient X-ray sources, one long-lived and coincident with SN1998bw and the other, briefer and not coincident with SN1998bw, resembling a GRB X-ray afterglow. \section{Discussion} The most difficult part of the problem of the radio emission of SN1998bw is explaining how the kinetic energy of the Compton electrons is converted to the acceleration of the electrons which produced the observed radio emission. These must have $\gamma_e^\prime \sim 100$, making them many times more energetic than the Compton electrons. The interpenetration of Compton- and counter-currents is unstable to electrostatic and electromagnetic plasma instabilities, which may explain both the electron acceleration and the magnetic field (although previous sections have also discussed Compton current, hydrodynamic dynamo and pulsar radiation zone models of the field). The absence of radio polarization (\cite{K98}) suggests that the magnetic field is disordered, which would be consistent with the dynamo, pulsar and plasma instability models of the field, but might also be explained by differential Faraday rotation (which may be estimated, using the previous parameters, as $\sim 10^4$ radians even at $\lambda = 3$ cm) within the Compton current model. Compton electron models predict the presence of the nuclear gamma-rays. At distances of many Mpc these gamma-rays are unlikely to be detectable directly, but their effects on the visible light curve are evident (and, in fact, led to the inference of $A = 56$ radioactive decay). In these models the rise of the radio flux in the first week is dependent on the escape of the gamma-rays, which will only occur so soon for supernovae (such as SN1998bw) with very fast debris. It is therefore predicted that strong early radio emission will be correlated with the debris expansion speed. It is also predicted that strong early radio emission requires the presence of a surrounding medium, presumably the result of mass loss by the progenitor, of density $> 10^4$ cm$^{-3}$, far in excess of typical interstellar densities. Such a medium may be independently confirmed or disproved because of its effects on the SN visible light curve, pulse dispersion of any newly born pulsar, or X-ray emission resulting from its interaction with the debris. As the debris collides with the radio-absorbing gas cloud a forward and reverse shock are produced. The forward shock has a post-shock temperature $\sim 1$ MeV because of the high supernova debris speed, and emits comparatively little radiation. However, the reverse shock is cooler and propagates in denser matter (the debris having a mass of several $M_\odot$, compared to the mass $\ll M_\odot$ of the radio-absorbing envelope) and may well produce the X-ray source S1 discovered (\cite{P99}) to be associated with SN1998bw and to decay over $\sim 6$ months. The decay is attributable to the forward shock reaching the outer extent of the radio-absorbing cloud, after which a rarefaction reflects and erodes the reverse shock. Supernovae in which Compton electrons produce radio emission should have several similar properties: They will all have approximately the same Lorentz factors (1.6--2) for their expanding radio photospheres, because this is determined by nuclear physics. They will all show similar double-peaked radio intensity curves, again because the nuclear physics is the same. Finally, their debris will show a nearly complete absence of a low-Z envelope (as distinct from a surrounding low-Z medium) because such an envelope would prevent the escape of the Compton electrons. It is also worth noting that the energy of explosive C or O burning (about 1 MeV/nucleon) is insufficient to produce the debris velocity observed in SN1998bw; gravitational energy from collapse to nearly neutron star density (and probably neutrino transport) must be appealed to. After the submission of the original version of this paper \cite{NP99} reported that SN1987A was accompanied by two visible sources asymmetrically located on opposite sides of the supernova. Assuming these sources to represent jets directed outward from the supernova at the same speed, they found that their speed of ejection was $0.80 c$. This is the speed of the mean Compton electrons produced by the abundant 0.812 MeV gamma-rays of $^{56}$Ni (\S 2), and is consistent with the radio source expansion speed found by \cite{K98} for SN1998bw. Although relativistic expansion is observed at very different frequencies in these two objects, perhaps because of observational limitations, it may be that both the spots near SN1987A and the radio emission of SN1998bw are powered by similar Compton electrons. \acknowledgements I thank D. A. Frail and S. R. Kulkarni for useful discussions.
\section{\sc Preliminaries} If $(a_n)$ and $(b_n)$ are scalar sequences we write \mbox{$a_n \prec b_n$} whenever there is some $c \ge 0$ such that $a_n \le c \cdot b_n$ for all $n$, and \mbox{$a_n \asymp b_n$} whenever \mbox{$a_n \prec b_n$} and \mbox{$b_n \prec a_n$}. For $1 \le p \le \infty$ the number $p'$ is defined by $1/p + 1/{p'} =1$. \par We shall use standard notation and notions from Banach space theory, as presented e.\,g. in \cite{djt}, \cite{lt} and \cite{tj}. If $E$ is a Banach space, then $B_E$ is its (closed) unit ball and $E'$ its dual. We consider complex Banach spaces only (but note that most of the consequences of our abstract interpolation results can be formulated also for real spaces). As usual ${\cal L}(E,F)$ denotes the Banach space of all (bounded and linear) operators from $E$ into $F$ endowed with the operator norm $\norm{\cdot}$. With $\mathbf{T_2}(E)$ and $\mathbf{C_2}(E)$ we denote the (Gaussian) type~2 and cotype~2 constant of a Banach space $E$ with this property, respectively, and with $\mathbf{M_{(u)}}(E)$ and $\mathbf{M^{(u)}}(E)$ the $u$-concavity and $u$-convexity constant of a Banach lattice $E$. \par We call a Banach space $E \subset c_0$ (the space of all zero sequences) a symmetric Banach sequence space if the $i$-th standard unit vectors $e_i$ form a symmetric basis, i.\,e. the $e_i$'s form a Schauder basis such that $\norm{x}_E = \norm{\sum_{i=1}^\infty \varepsilon_i x_{\pi(i)} e_i}_E$ for each $x \in E$, each permutation $\pi$ of $\Bbb{N}$ and each choice of scalars $\varepsilon_i$ with $|\varepsilon_i|=1$. Moreover, denote for each $n$ the subspace $\text{{\rm span}} \{e_i \, | \, 1 \le i \le n\}$ of $E$ by $E_n$. Together with its natural order a symmetric Banach sequence space $E$ forms a Banach lattice, and clearly its basis is $1$-unconditional. The associated unitary ideal $\mathcal{S}_E$ is the Banach space of all compact operators $T\in {\cal L}(\ell_2 ,\ell_2)$ with singular numbers $(s_i(T))_i$ in $E$ endowed with the norm $\norm{T}_{\mathcal{S}_E}:= \norm{(s_i(T))_i}_E$; with $\mathcal{S}_E^n$ we denote ${\cal L}(\ell_2^n,\ell_2^n)$ together with the norm $\norm{T}_{\mathcal{S}_E^n}:= \norm{(s_i(T))_{i=1}^n}_{E_n}$. For $E=\ell_u$ ($1 \le u < \infty$) one gets the well-known Schatten-$u$-class $\mathcal{S}_u$; for simplicity put $\mathcal{S}_\infty:= {\cal L}(\ell_2,\ell_2)$. \par For all information on Banach operator ideals, in particular on summing and mixing operators, see e.\,g. \cite{df}, \cite{djt} and \cite{pietsch}. An operator $T \in {\cal L}(E,F)$ is called absolutely $(r,p)$-summing $(1 \le p \le r \le \infty)$ if there is a constant $\rho \ge 0$ such that $$ \bigl ( \sum_{i=1}^n \norm{Tx_i}^r \bigr )^{1/r} \le \rho \cdot \sup \biggl \{ \bigl ( \sum_{i=1}^n |\langle x',x_i \rangle |^p \bigr )^{1/p} \, | \, x' \in B_{E'} \biggr \} $$ for all finite sets of elements $x_1, \dots, x_n \in E$ (with the obvious modifications for $p$ or $r$ $=\infty$). In this case, the infimum over all possible $\rho \ge 0$ is denoted by $\pi_{r,p}(T)$, and the Banach operator ideal of all absolutely $(r,p)$-summing operators by $(\Pi_{r,p},\pi_{r,p})$; the special case $r=p$ gives the ideal $(\Pi_p, \pi_p)$ of all absolutely $p$-summing operators. \par An operator $T \in {\cal L}(E,F)$ is called $(s,p)$-mixing $(1 \le p \le s \le \infty)$ whenever its composition with an arbitrary operator $S \in \Pi_s(F,Y)$ is absolutely $p$-summing; with the norm $$ \mu_{s,p}(T) := \sup\{\pi_p(ST) \, | \, \pi_s(S) \le 1 \} $$ the class ${\cal M}_{s,p}$ of all $(s,p)$-mixing operators forms again a Banach operator ideal. Obviously, $({\cal M}_{p,p}, \mu_{p,p}) = ({\cal L}, \norm{\cdot})$ and $({\cal M}_{\infty,p}, \mu_{\infty,p}) = (\Pi_p, \pi_p)$. Recall that due to \cite{maurey} (see also \cite[32.10--11]{df}) summing and mixing operators are closely related: $$ ({\cal M}_{s,p}, \mu_{s,p}) \subset (\Pi_{r,p}, \pi_{r,p}) \qquad \text{for } 1/s + 1/r = 1/p, $$ and ``conversely'' $$ (\Pi_{r,p}, \pi_{r,p}) \subset ({\cal M}_{s_0,p}, \mu_{s_0,p}) \qquad \text{for } 1 \le p \le s_0 < s \le \infty \text{ and } 1/s + 1/r = 1/p. $$ Moreover, it is known that each $(s,2)$-mixing operator on a cotype~2 space is even $(s,1)$-mixing (see again \cite{maurey} and \cite[32.2]{df}). \par For an operator $T \in {\cal L}(E,F)$ the $n$-th Weyl number $x_n(T)$ of $T$ is defined by $$ x_n(T) := \sup \{a_n(TS) \,| \, S \in {\cal L}(\ell_2, E) \text{ with } \norm{S}=1 \},$$ where $a_n(TS)$ denotes the $n$-th approximation number of $TS$. We will use the following important inequality of K\"onig in order to obtain lower estimates: \begin{equation} \label{weyl} n^{1/r} \cdot x_n(T) \le \pi_{r,2}(T), \qquad T \in \Pi_{r,2} \end{equation} (for all details on $s$-numbers and this inequality see \cite[2.a.3]{koenig} or \cite{pietsch2}). \section{\sc Complex Interpolation of Mixing Operators} The aim of this section is to prove a complex interpolation formula for the mixing norm of a fixed operator acting between two complex interpolation spaces. \par For all information on complex interpolation we refer to \cite{BL}. A couple $[E_0,E_1]$ of Banach spaces is called an interpolation couple if there exists a topological Hausdorff vector space $E$ in which $E_0$ and $E_1$ can be continuously embedded. We speak of a finite-dimensional interpolation couple $[E_0, E_1]$, if $E_0$ and $E_1$ are finite-dimensional Banach spaces with the same dimensions. For an interpolation couple $[E_0,E_1]$ of Banach spaces and $0 \le \theta\le 1$ we denote by $[E_0,E_1]_\theta = E_\theta$ the interpolation space obtained by the complex interpolation method of Calder\'on. Well-known examples of complex interpolation spaces are the Minkowski spaces $\ell_p^n(E)$ and Schatten classes $\mathcal{S}_p^n$: For $1\le p_0,p_1 \le \infty $, $0 < \theta < 1$ and an interpolation couple $[E_0,E_1]$ $$ [\ell_{p_0}^n(E_0), \ell_{p_1}^n(E_1)]_\theta = \ell_p^n(E_\theta) \qquad \text{and} \qquad [\mathcal{S}_{p_0}^n,\mathcal{S}_{p_1}^n]_\theta = \mathcal{S}_p^n$$ (isometrically), where $1/p = (1-\theta)/{p_0} + \theta/{p_1}$ (for the complex interpolation formula for $\ell_p^n(E)$'s see \cite[5.1.2]{BL}, whereas the formula for Schatten classes can be deduced from e.\,g. \cite[Satz~8]{pt} and the complex reiteration theorem \cite[4.6.1]{BL}). For $0 \le \theta <1$ a $\theta$-Hilbert space is an interpolation space $[E_0,E_1]_\theta$ where $E_1$ is a Hilbert space (this notion goes back to Pisier); in particular, $\ell_p^n$ and $\mathcal{S}_p^n$ for $1 < p < \infty$ are $\theta$-Hilbert spaces for $\theta=1-|1-{2/p}|$. \par The following complex interpolation theorem for the mixing norm is our main abstract tool. \begin{theo} \label{mixing} Let $2 \le s_0,s_1 \le \infty$, $0 \le \theta \le 1$ and $s_\theta$ given by $1/{s_\theta}=(1-\theta)/{s_0} + \theta / {s_1}$. Then for two finite-dimensional interpolation couples $[E_0,E_1]$, $[F_0,F_1]$ and each $T \in {\cal L}(E_\theta,F_\theta)$ $$ \mu_{s_\theta,2}(T : E_\theta \rightarrow F_\theta) \le d_\theta[E_0,E_1] \cdot \mu_{s_0,2}(T : E_0 \rightarrow F_0)^{1-\theta} \cdot \mu_{s_1,2}(T : E_1 \rightarrow F_1)^{\theta},$$ where \begin{equation} d_\theta[E_0,E_1]:= \sup_m \norm{{\cal L}(\ell_2^m, E_\theta) \hookrightarrow [{\cal L}(\ell_2^m, E_0), {\cal L}(\ell_2^m, E_1)]_\theta}. \end{equation} \end{theo} \proof Consider for $\eta=0,\theta,1$ the bilinear mapping $$ \begin{array}{lccccc} \Phi_\eta^{n,m}: &\ell_{s_\eta}^n(F_\eta') & \times &{\cal L}(\ell_2^m, E_\eta) &\longrightarrow & \ell_2^m(\ell_{s_\eta}^n) \\ &(y_1', \dots , y_n') &\times & S &\longmapsto & ((\langle y_k',TSe_j \rangle )_k)_j \end{array}, $$ where $(e_j)$ denotes the canonical basis in $\Bbb{C}^m$. By the discrete characterization of the mixing norm (see \cite{maurey} or \cite[32.4]{df}) $\mu_{s_\eta}(T: E_\eta \rightarrow F_\eta)$ is the infimum over all $c \ge 0$ such that for all $n,m$, all $y_1', \ldots, y_n' \in F_\eta'$ and all $x_1, \ldots, x_m \in E_\eta$ $$ \left ( \sum_{j=1}^m \left ( \sum_{k=1}^n |\langle y_k',Tx_j \rangle | ^{s_\eta} \right )^{2/{s_\eta}} \right )^{1/2} \le c \cdot \left ( \sum_{k=1}^n \norm{y_k'}_{F_\eta'} ^{s_\eta} \right )^{1/{s_\eta}} \cdot \sup_{x' \in B_{E_\eta'}} \left ( \sum_{j=1}^m |\langle x',x_j \rangle |^2 \right )^{1/2}. $$ Since for each $S = \sum_{j=1}^m e_j \otimes x_j \in {\cal L}(\ell_2^m,E_\eta)$ $$ \norm{S} = \sup_{x' \in B_{E_\eta'}} \left ( \sum_{j=1}^m | \langle x',x_j \rangle |^2 \right )^{1/2}, $$ we obviously get that $$ \mu_{s_\eta,2}(T:E_\eta \rightarrow F_\eta) = \sup_{n,m} \norm{\Phi_\eta^{n,m}}. $$ Now the proof follows by bilinear complex interpolation: For the interpolated bilinear mapping $$ [\Phi_0^{n,m},\Phi_1^{n,m}]_\theta : [\ell_{s_0}^n(F_0'), \ell_{s_1}^n(F_1')]_\theta \times [{\cal L}(\ell_2^m,E_0), {\cal L}(\ell_2^m,E_1)]_\theta \longrightarrow [\ell_2^m(\ell_{s_0}^n),\ell_2^m(\ell_{s_1}^n)]_\theta $$ by \cite[4.4.1]{BL} $$ \norm{[\Phi_0^{n,m},\Phi_1^{n,m}]_\theta} \le \norm{\Phi_0^{n,m}}^{1-\theta} \cdot \norm{\Phi_1^{n,m}}^\theta. $$ Since by the interpolation theorem for $\ell_p(E)$'s together with the duality theorem \cite[4.5.2]{BL} $$ [\ell_{s_0}^n(F_0'),\ell_{s_1}^n(F_1')]_\theta = \ell_{s_\theta}^n(F_\theta') \qquad \text{and} \qquad [\ell_2^m(\ell_{s_0}^n),\ell_2^m(\ell_{s_1}^n)]_\theta = \ell_2^m(\ell_{s_\theta}^n) $$ (isometrically), we have $$ \norm{\Phi_\theta^{n,m}} \le \norm{{\cal L}(\ell_2^m, E_\theta) \hookrightarrow [{\cal L}(\ell_2^m, E_0), {\cal L}(\ell_2^m, E_1)]_\theta} \cdot \norm{[\Phi_0^{n,m},\Phi_1^{n,m}]_\theta}.$$ Consequently \begin{align*} \mu_{s_\theta,2}(T:E_\theta \rightarrow F_\theta) &= \sup_{n,m} \norm{\Phi_\theta^{n,m}} \\ & \le\sup_{n,m} \{ \norm{{\cal L}(\ell_2^m, E_\theta) \hookrightarrow [{\cal L}(\ell_2^m, E_0), {\cal L}(\ell_2^m, E_1)]_\theta} \cdot \norm{\Phi_0^{n,m}}^{1-\theta} \cdot \norm{\Phi_1^{n,m}}^\theta \} \\ & \le d_\theta[E_0,E_1] \cdot \mu_{s_0,2}(T:E_0 \rightarrow F_0)^{1-\theta} \cdot \mu_{s_1,2}(T:E_1 \rightarrow F_1)^\theta, \end{align*} the desired result. \qed \\[10pt] In the same way an analogous result for the $(r,2)$-summing norm can be obtained. \par Applications of theorem~\ref{mixing} come from ``uniform estimates'' for $d_\theta[E_0,E_1]$. Pisier proved in \cite{pisier90} that \begin{equation} \label{koubal} d_\theta[\ell_1,\ell_2]:= \sup_n d_\theta[\ell_1^n,\ell_2^n] < \infty; \end{equation} the proof is based on the Maurey factorization theorem which says that every operator $T:\ell_2 \rightarrow \ell_p$ ($1 \le p \le 2$) factorizes through an appropriate diagonal operator $D_\lambda: \ell_2 \rightarrow \ell_p$, and spaces of diagonal operators clearly behave well under interpolation. The fact \eqref{koubal} can also be obtained as an application (case (c) of the following estimates) of a deep result of Kouba \cite{kouba} on the complex interpolation of injective tensor products of Banach spaces (see also \cite{dm99})---as a consequence of Kouba's work we get for a finite-dimensional interpolation couple $[E_0,E_1]$ and $0 \le \theta \le 1$ the following estimates for $d_\theta[E_0,E_1]$: \begin{enumerate}[(a)] \item $d_\theta[E_0,E_1]=1$, if $E_0=E_1$. \item $ d_\theta[E_0,E_1] \le \mathbf{T_2} (E_0')^{1-\theta} \cdot \mathbf{T_2} (E_1')^{\theta}.$ \item $ d_\theta[E_0,E_1] \le (2/\sqrt{\pi}) \cdot \mathbf{M_{(2)}}(E_0)^{1-\theta} \cdot \mathbf{M_{(2)}}(E_1)^{\theta},$ if the canonical bases in $E_0$ and $E_1$ are $1$-unconditional and induce the lattice structures. \end{enumerate} \vspace{-5pt} Note that the constants given in (b) and (c) are different from those in Kouba's work. They follow from a short analysis of his proofs in the finite-dimensional case: Since in our setting a Hilbert space is involved, Kouba's formulas (3.8) and (3.10) on p.~47--48 can both be changed into \mbox{$\gamma_z (T) \le \| \mspace{-1.5mu} | T \| \mspace{-1.5mu} |_z$}. Moreover, calculating the terms $W^E(z)$ defined in Kouba's lemma~3.2 and lemma~3.3 (use two spaces instead of a family of Banach spaces, see also \cite[p.~218]{ccrsw}), one obtains $$W^E(z)= \mathbf{T_2}(E_0)^{1-\theta (z)} \cdot \mathbf{T_2}(E_1)^{\theta (z)} \text{ and } W^E(z) = (2/\sqrt{\pi}) \cdot \mathbf{M^{(2)}} (E_0)^{1-\theta (z)} \cdot \mathbf{M^{(2)}} (E_1)^{\theta (z)} $$ (with $\theta (z)$ as in \cite[corollary~5.1]{ccrsw}), respectively. This leads to the above estimates (note that $E_0$ and $E_1$ in Kouba's formulas, in our context have to be replaced by $E_0'$ and $E_1'$, and recall the well-known duality relation between $2$-concavity and $2$-convexity, see e.\,g. \cite[1.d.4]{lt}). \\[10pt] Later we also need a non-commutative version of Pisier's result \eqref{koubal}. Using an extension of Kouba's formulas for the Haagerup tensor product of operator spaces due to \cite{pisierop}, Junge in \cite[4.2.6]{junge} proved an analogue of \eqref{koubal} for Schatten classes: \begin{equation} \label{koubas} d_\theta[\mathcal{S}_1,\mathcal{S}_2]:= \sup_n d_\theta[\mathcal{S}_1^n,\mathcal{S}_2^n] < \infty. \end{equation} \par Finally we state a corollary on $\theta$-Hilbert spaces which together with \eqref{koubal} and \eqref{koubas} is crucial for our purposes. \begin{cor} \label{hilbert} Let $0 \le \theta \le 1$, $E=[E_0,\ell_2^n]_\theta$ be a $n$-dimensional $\theta$-Hilbert space and \mbox{$2 \le s_\theta \le\infty$} given by $s_\theta =2/\theta$. Then $$ \mu_{s_\theta,2}(E \hookrightarrow \ell_2^n) \le d_\theta[E_0,\ell_2^n] \cdot \pi_2(E_0 \hookrightarrow \ell_2^n)^{1-\theta}.$$ \end{cor} \section{\sc Bennett--Carl Inequalities for Symmetric Banach Sequence Spaces} As announced the preceding interpolation theorem implies the Bennett--Carl result and its extension of Carl--Defant as an almost immediate consequence: \begin{samepage} \begin{cor} \label{beca} Let $1 \le u \le 2 $ and $1 \le u \le v \le \infty$. Then for $2 \le s \le \infty $ such that $1/s= 1/2 - 1/u + \max(1/v,1/2)$ $$ \sup_{n} \mu_{s,2}(\ell_u^n \hookrightarrow \ell_v^n) < \infty.$$ In particular, for $2 \le r \le \infty$ such that $1/r = 1/u - \max(1/v,1/2)$ $$ \sup_{n} \pi_{r,2}(\ell_u^n \hookrightarrow \ell_v^n) < \infty.$$ \end{cor} \end{samepage} \proof Only the case $1 \le u < v \le 2$ has to be considered; the case $2 \le v \le \infty$ then easily follows by factorization through $\ell_2^n$, and the case $u=v$ is trivial anyway. In what follows we use the complex interpolation formula for $\ell_p^n$'s without further mentioning. \\ i) Take first $v=2$. It is well-known (see e.\,g. \cite[22.4.8]{pietsch} or \eqref{hilfe1}) that $$ \pi_2 (\ell_1^n \hookrightarrow \ell_2^n)= 1. $$ For $1 \le u \le 2$ choose $0\le \theta \le 1$ such that $1/u = (1-\theta)/1 + \theta/2$. Then $s_\theta:=2/\theta=u'$, and by corollary~\ref{hilbert} together with \eqref{koubal} $$ \mu_{u',2}(\ell_u^n \hookrightarrow \ell_2^n) \le d_\theta[\ell_1,\ell_2] < \infty.$$ ii) Let $1 \le u < v < 2$. Combining case~i), $$ \mu_{u',2}(\ell_u^n \hookrightarrow \ell_2^n) \le d_\theta[\ell_1,\ell_2],$$ and $$ \mu_{2,2} (\ell_u^n \hookrightarrow \ell_u^n ) = \norm{\ell_u^n \hookrightarrow \ell_u^n } =1, $$ we have $$ \mu_{s_{\tilde{\theta}},2}(\ell_u^n \hookrightarrow \ell_v^n) \le \sup_n d_{\tilde{\theta}}[\ell_u^n,\ell_u^n] \cdot d_\theta[\ell_1,\ell_2]^{1-\tilde{\theta}} < \infty,$$ with $\tilde{\theta} := (1/v -1/2)/(1/u - 1/2)$ and $1/{s_{\tilde{\theta}}} := (1-\tilde{\theta})/{u'} + \tilde{\theta}/2 =1/2 - 1/u + 1/v=1/s$. \qed \par As in the original proofs of Bennett and Carl the crucial step in the preceding proof is to show that for the symmetric Banach sequence space $E=\ell_u$ \begin{equation} \label{crucial} \sup_{n} \pi_{r,2} (E_n \hookrightarrow \ell_2^n) < \infty, \end{equation} where $1 \le u \le 2$ and $1/r =1/u -1/2$. We now prove a result within the framework of symmetric Banach sequence spaces which shows that \eqref{crucial} is sharp in a very strong sense. Take an arbitrary $2$-concave and $u$-convex Banach sequence space $E$---these geometric assumptions in particular imply that the continuous inclusions $\ell_u \subset E \subset \ell_2$ hold---which satisfies \eqref{crucial}. The following result shows that there is only one such space: \begin{samepage} \begin{theo} \label{converse} Let $1 \le u \le 2$ and $1/r =1/u -1/2$. For each $2$-concave and $u$-convex symmetric Banach sequence space $E$ the following are equivalent: \em \vspace{-5pt} \begin{enumerate}[(1)] \item $\sup_{n} \mu_{u',2}(E_n \hookrightarrow \ell_2^n) <\infty$. \item $\sup_{n} \pi_{r,2}(E_n \hookrightarrow \ell_2^n) < \infty$. \item $E=\ell_u$. \end{enumerate} \end{theo} \end{samepage} Clearly we only have to deal with the implication $(2) \Rightarrow (3)$; its proof is based on two lemmas. For the first one we invent the notion of ``enough symmetries in the orthogonal group''. Let $E=(\Bbb{C}^n,\norm{\cdot})$ be an $n$-dimensional Banach space. We say that $E$ has {\em enough symmetries in $\mathcal{O}(n)$} if there is a compact subgroup $G$ in $\mathcal{O}(n)$ such that \begin{equation} \label{invariant} \forall \, u \in {\cal L}(E) \, \forall \, g,g' \in G: \norm{u}=\norm{gug'} \end{equation} and \begin{equation} \label{enough} \forall \, u \in {\cal L}(E) \text{ with } ug=gu \text{ for all } g \in G \, \exists \, c \in \Bbb{K}: u = c \cdot \text{{\rm id}}_E. \end{equation} Basic examples of spaces with enough symmetries in the orthogonal group are the finite-dimensional spaces $E_n$ and $\mathcal{S}_E^n$ associated to a symmetric Banach sequence space $E$. The following lemma extends the corresponding results in \cite[p.~233, 236]{CD97}. \begin{lemma} \label{2summing} Let $E_n$ and $F_n$ have enough symmetries in $\mathcal{O}(n)$. Then \begin{equation} \label{hilfe1} \pi_2(E_n \hookrightarrow F_n) = n^{1/2} \cdot \frac{\norm{\ell_2^n \hookrightarrow F_n}}{\norm{\ell_2^n \hookrightarrow E_n}}, \end{equation} and for $1 \le k \le n$ \begin{equation} \label{hilfe2} \left ( \frac{n-k+1}{n} \right )^{1/2} \cdot \frac{\norm{\ell_2^n \hookrightarrow F_n}}{\norm{\ell_2^n \hookrightarrow E_n}} \le x_k(E_n \hookrightarrow F_n) \le \left ( \frac{n}{k} \right )^{1/2} \cdot \frac{\norm{\ell_2^n \hookrightarrow F_n}}{\norm{\ell_2^n \hookrightarrow E_n}}. \end{equation} \end{lemma} \proof \eqref{hilfe1}: Trace duality allows to deduce the lower estimate from the upper one: $$ n \le \pi_2(\ell_2^n \hookrightarrow F_n) \cdot \pi_2(F_n \hookrightarrow \ell_2^n) \le \norm{\ell_2^n \hookrightarrow E_n} \cdot \pi_2(E_n \hookrightarrow F_n) \cdot n^{1/2} \cdot \norm{\ell_2^n \hookrightarrow F_n}^{-1}. $$ For the proof of the upper estimate it may be assumed without loss of generality that $F_n=\ell_2^n$ (factorize through $\ell_2^n$). In this case it suffices to show that $$ \norm{\ell_2^n \hookrightarrow E_n}^{-1} \cdot B_{\ell_2^n} $$ is John's ellipsoid $D_{\max}$ of maximal volume in $B_{E_n}$ (see e.\,g. \cite[3.8]{pisier89} or \cite[6.30]{djt}). By definition there is a linear bijection $u: \ell_2^n \rightarrow E_n$ such that $u(B_{\ell_2^n}) = D_{\max}$. In particular, $\norm{u}=1$ and $N(u^{-1})=n$ ($N$ denotes the nuclear norm, see e.\,g. \cite[3.7]{pisier89} or \cite[6.30]{djt}). On the other hand by a standard averaging argument there is a linear bijection $v:\ell_2^n \rightarrow E_n$ with $\norm{v}=1$, $N(v^{-1})=n$ and $vg=gv$ for all $ g \in G$, where $G$ is a compact group in $\mathcal{O}(n)$ satisfying \eqref{invariant} and \eqref{enough} (see \cite[3.5]{pisier89} which also holds in the complex case). By property \eqref{enough} of $G$ and the fact that $\norm{v}=1$ we have $v=\norm{\ell_2^n \hookrightarrow E_n}^{-1} \cdot \text{{\rm id}}$. Then by Lewis' uniqueness theorem $v^{-1}u \in \mathcal{O}(n)$ (\cite[3.7]{pisier89} or \cite[6.25]{djt}). Altogether we finally obtain $$ \norm{\ell_2^n \hookrightarrow E_n}^{-1}\cdot B_{\ell_2^n} =v(B_{\ell_2^n}) = v [v^{-1}u(B_{\ell_2^n})] = u(B_{\ell_2^n})=D_{\max}. $$ \eqref{hilfe2}: Recall from \eqref{weyl} that $k^{1/2} \cdot x_k(T) \le \pi_2(T)$ for every $2$-summing operator $T$ acting between two Banach spaces. Together with \eqref{hilfe1} this gives the second inequality. The first then follows from the basic properties of the Weyl numbers (see e.\,g. \cite{koenig}): \begin{align*} 1 &= x_n(\text{{\rm id}}_{\ell_2^n}) \\ & \le x_k(\ell_2^n \hookrightarrow F_n) \cdot x_{n-k+1}(F_n \hookrightarrow \ell_2^n) \\ & \le \norm{\ell_2^n \hookrightarrow E_n} \cdot x_k(E_n \hookrightarrow F_n) \cdot \left (\frac{n}{n-k+1} \right )^{1/2} \cdot \norm{\ell_2^n \hookrightarrow F_n}^{-1}. \end{align*} \qed \\[10pt] The following obvious examples will be useful later. \begin{cor} For $1 \le u,v \le \infty$ \begin{equation} \label{formel} \pi_2 (\mathcal{S}_u^n \hookrightarrow \mathcal{S}_v^n) = n \cdot \frac{\max(1, n^{1/v -1/2})}{\max(1, n^{1/u -1/2})} \end{equation} and \begin{equation} \label{weylschatten} x_{[n^2/2]} (\mathcal{S}_u^n \hookrightarrow \mathcal{S}_v^n) \asymp \frac{\max(1, n^{1/v -1/2})}{\max(1, n^{1/u -1/2})}. \end{equation} \end{cor} \par The preceding lemma turns out to be of special interest in combination with a result due to Szarek and Tomczak-Jaegermann \cite[proposition~2.2]{stj} which states that for each $2$-concave symmetric Banach sequence space $E$ \begin{equation} \label{szarek} \norm{\ell_2^n \hookrightarrow E_n} \asymp n^{-1/2} \cdot \norm{\sum\nolimits_1^n e_i}_{E_n}. \end{equation} The second lemma which we need for the proof of theorem~\ref{converse}, is based on \eqref{szarek} and an important result about the interpolation of Banach lattices due to Pisier \cite{pisier79}. \begin{lemma} For $1 \le u \le 2$ let $E$ be a $u$-convex and $u'$-concave symmetric Banach sequence space. Then \begin{equation} \label{zuerst} \norm{E_n \hookrightarrow \ell_u^n} \asymp \frac{n^{1/u}}{\norm{\sum_1^n e_i}_{E_n}}. \end{equation} In particular, if $E$ is even $2$-concave, then \begin{equation} \label{lu} \norm{E_n \hookrightarrow \ell_u^n} \asymp \frac{n^{1/u}}{\norm{\sum_1^n e_i}_{E_n}} \asymp \frac{n^{1/u-1/2}}{\norm{\ell_2^n \hookrightarrow E_n}}. \end{equation} \end{lemma} \proof \eqref{lu} follows directly from \eqref{zuerst} and \eqref{szarek}, and clearly $n^{1/u} \le \norm{E_n \hookrightarrow \ell_u^n} \cdot \norm{\sum_1^n e_i}_{E_n}$. For the upper estimate in \eqref{zuerst} we only have to consider $1 < u <2$: The case $u=1$ is stated below in \eqref{border}, and a $2$-convex and $2$-concave symmetric Banach sequence space necessarily equals $\ell_2$ with equivalent norms. Without loss of generality we may assume $\mathbf{M^{(u)}}(E)= \mathbf{M_{(u')}}(E)=1$ (see \cite[1.d.8]{lt}). Then by \cite[theorem~2.2]{pisier79} there exists a symmetric Banach sequence space $E_0$ such that $E=[E_0,\ell_2]_\theta$ with $\theta=2/{u'}$; moreover, we have $E_n=[E_0^n,\ell_2^n]_\theta$ with equal norms. The conclusion now follows by interpolation: It can be shown easily that \begin{equation} \label{border} \norm{E_0^n \hookrightarrow \ell_1^ n} \le \frac{n}{\norm{\sum_1^n e_i}_{E_0^n}} \end{equation} (see e.\,g. \cite[proposition~2.5]{stj}), hence $$ \norm{E_n \hookrightarrow \ell_u^n} \le \norm{E_0^n \hookrightarrow \ell_1^n}^{1-\theta} \cdot \norm{\ell_2^n \hookrightarrow \ell_2^n}^\theta \le \frac{n^{1-\theta}}{\norm{\sum_1^n e_i}_{E_0^n}^{1-\theta}}. $$ Since $E_n=[E_0^n,\ell_2^n]_\theta$ is of $J$-type $\theta$ (i.\,e. $\norm{x}_{E_n} \le \norm{x}_{E_0^n}^{1-\theta} \cdot \norm{x}_{\ell_2^n} ^\theta$ for all $x \in E_n$), we have $$ \norm{{\textstyle \sum_1^n e_i}}_{E_n} \le \norm{{\textstyle \sum_1^n e_i}}_{E_0^n}^{1-\theta} \cdot n^{\theta/2}, $$ and consequently $$ \norm{E_n \hookrightarrow \ell_u^n} \le \frac{n^{1-{\theta/2}}}{\norm{\sum_1^n e_i}_{E_n}} = \frac{n^{1/u}}{\norm{\sum_1^n e_i}_{E_n}}. $$ \qed \\[10pt] {\em Proof} of the implication $(2) \Rightarrow (3)$ in theorem~\ref{converse}: Assume that $\sup_{n} \pi_{r,2}(E_n \hookrightarrow \ell_2^n) < \infty$. By \eqref{weyl}, \eqref{hilfe2} and \eqref{lu} \begin{equation} \label{lower} \pi_{r,2}(E_n \hookrightarrow \ell_2^n) \ge [n/2]^{1/r} \cdot x_{[n/2]}(E_n \hookrightarrow \ell_2^n) \succ \frac{n^{1/r}}{\norm{\ell_2^n \hookrightarrow E_n}} \asymp \norm{E_n \hookrightarrow \ell_u^n}, \end{equation} which by assumption shows that $\sup_n \norm{E_n \hookrightarrow \ell_u^n} < \infty$. This clearly gives the claim. \qed \\[10pt] Note that \eqref{lower} does not depend on the special choice of $r$. \par If $E$ is a $2$-concave and $u$-convex $(1 \le u \le 2)$ symmetric Banach sequence space different from $\ell_u$ (i.\,e. the inclusion $\ell_u \subset E$ is strict), then by theorem~\ref{converse} for $1/r = 1/u-1/2$ $$ \pi_{r,2}(E_n \hookrightarrow \ell_2^n) \nearrow \infty. $$ The following result gives the precise asymptotic order of the sequence \mbox{$(\pi_{r,2}(E_n \hookrightarrow \ell_2^n))_n$}: \begin{cor} \label{precise} For $1 \le u \le 2$ let $E$ be a $2$-concave and $u$-convex symmetric Banach sequence space. Then for $2 \le r,s \le \infty$ such that $1/r =1/u-1/2$ and $1/s=1/2-1/r$ $$ \pi_{r,2}(E_n \hookrightarrow \ell_2^n) \asymp \mu_{s,2}(E_n \hookrightarrow \ell_2^n) \asymp \frac{n^{1/r+1/2}}{\norm{\sum_1^n e_i}_{E_n}}. $$ \end{cor} \proof The lower estimate has already been shown in \eqref{lower}, and the upper estimate simply follows by factorization through $\ell_u^n$, the Bennett--Carl inequalities and \eqref{lu}. \qed \begin{rem} \label{remark} \begin{enumerate}[(a)] \item Since a $u$-convex Banach lattice is $p$-convex for all $1\le p \le u$ (see \cite[1.d.5]{lt}), the formula in the preceding theorem even holds for all $2\le r \le \infty$ such that $1/r \ge 1/u -1/2$. \item For $1 \le u \le 2$ let $E$ be a $2$-concave and $u$-convex symmetric Banach sequence space, $F$ an arbitrary symmetric Banach sequence space, and let $2 \le r \le \infty$ such that $1/r \ge 1/u -1/2$. Then---by factorization through $\ell_2^n$ for the upper estimate and \eqref{lower} for the lower one---the following formula holds: $$ \pi_{r,2}(E_n \hookrightarrow F_n) \asymp n^{1/r} \cdot \frac{\norm{\ell_2^n \hookrightarrow F_n}} {\norm{\ell_2^n \hookrightarrow E_n}}; $$ in particular, if $F$ is a $2$-concave, then $$ \pi_{r,2}(E_n \hookrightarrow F_n) \asymp n^{1/r} \cdot \frac{\norm{\sum_1^n e_i}_{F_n}} {\norm{\sum_1^n e_i}_{E_n}}. $$ Note that these results can be considered as extensions of \eqref{hilfe1}. \item For the special case $F=\ell_v$ ($1 \le u \le v \le 2$) the formulas in (b) even hold for all $2 \le r \le \infty$ such that $1/r \ge 1/u -1/v$; simply repeat the proof of corollary~\ref{precise} for $1/r =1/u -1/v$ and use the argument from remark (a). \end{enumerate} \end{rem} \section{\sc Bennett--Carl Inequalities for Unitary Ideals} We now use Junge's counterpart \eqref{koubas} of \eqref{koubal} and our interpolation theorem~\ref{mixing} in order to show a ``non-commutative'' analogue. Note first that for all $1 \le u,v \le \infty$ and $2 \le r \le \infty$ \begin{equation} \label{low} n^{1/r} \le \pi_{r,2}(\mathcal{S}_u^n \hookrightarrow \mathcal{S}_v^n), \end{equation} and hence also for $2 \le s \le \infty$ $$ n^{1/2 -1/s} \le \mu_{s,2}(\mathcal{S}_u^n \hookrightarrow \mathcal{S}_v^n);$$ this is a consequence of the trivial estimate $\pi_{r,2}(\ell_2^n \hookrightarrow \ell_2^n) \ge n^{1/r}$ (insert $e_k$'s) and the fact that $\ell_2^n$ is $1$-complemented in each $\mathcal{S}_u^n$ (assign to each $x \in \ell_2^n$ the matrix $x \otimes e_1 \in \mathcal{S}_u^n$). For $u,v$ considered in corollary~\ref{beca} this lower bound is optimal: \begin{cor} \label{schatten} Let $1 \le u \le 2 $ and $1 \le u \le v \le \infty$. Then for $2 \le s \le \infty$ such that $1/s= 1/2 - 1/u + \max(1/v,1/2)$ $$ \mu_{s,2}(\mathcal{S}_u^n \hookrightarrow \mathcal{S}_v^n) \asymp n^{1/2-1/s}.$$ In particular, for $2 \le r \le \infty$ and $1/r = 1/u - \max(1/v,1/2)$ $$\pi_{r,2}(\mathcal{S}_u^n \hookrightarrow \mathcal{S}_v^n) \asymp n^{1/r}.$$ \end{cor} \proof The proof of the upper bound is analogous to that of corollary~\ref{beca}: Of course the complex interpolation formula for $\mathcal{S}_p^n$'s is needed instead of that for $\ell_p^n$'s, and in i) use $\pi_2(\mathcal{S}_1^n \hookrightarrow \mathcal{S}_2^n) = n^{1/2}$ (see \eqref{formel}) and Junge's result \eqref{koubas} in order to obtain $$ \mu_{u',2}(\mathcal{S}_u^n \hookrightarrow \mathcal{S}_2^n) \le d_\theta[\mathcal{S}_1^n,\mathcal{S}_2^n] \cdot n^{(1-\theta)/2} \le d_\theta[\mathcal{S}_1,\mathcal{S}_2] \cdot n^{1/u-1/2}, $$ where $\theta=2/{u'}$. Then in ii) one arrives at $$ \mu_{s_{\tilde{\theta}},2}(\mathcal{S}_u^n \hookrightarrow \mathcal{S}_v^n) \prec n^{(1-\tilde{\theta})(1/u - 1/2)} =n^{1/u-1/v}, $$ with $\tilde{\theta} := (1/v -1/2)/(1/u - 1/2)$ and $1/{s_{\tilde{\theta}}} = (1-\tilde{\theta})/{u'} + \theta/2 =1/2 - 1/u + 1/v$=1/s.\qed \par Exploiting the ideas of the preceding section one easily obtains the asymptotic order of the $(r,2)$-summing and the $(s,2)$-mixing norm of identities between finite-dimensional unitary ideals $\mathcal{S}_E^n$ and $\mathcal{S}_2^n$: \begin{cor} \label{unitary} For $1 \le u \le 2$ let $E$ be a $2$-concave and $u$-convex symmetric Banach sequence space. Then for all $2\le r,s \le \infty$ such that $1/r \ge 1/u -1/2$ and $1/s =1/2-1/r$ $$ \pi_{r,2}(\mathcal{S}_E^n \hookrightarrow \mathcal{S}_2^n) \asymp \mu_{s,2}(\mathcal{S}_E^n \hookrightarrow \mathcal{S}_2^n) \asymp \frac{n^{2/r+1/2}}{\norm{\sum_1^n e_i}_{E_n}}. $$ \end{cor} \proof Recall the simple fact that for all symmetric Banach sequence spaces $E$ and $F$ \begin{equation} \label{ssui} \norm{\mathcal{S}_E^n \hookrightarrow \mathcal{S}_F^n} = \norm{E_n \hookrightarrow F_n}, \end{equation} and by the same reasoning as in remark~\ref{remark} (a) it is enough to deal with the case $1/r=1/u-1/2$. Then factorization through $\mathcal{S}_u^n$ and \eqref{lu} give $$ \mu_{u',2}(\mathcal{S}_E^n \hookrightarrow \mathcal{S}_2^n) \prec \norm{\mathcal{S}_E^n \hookrightarrow \mathcal{S}_u^n} \cdot n^{1/u-1/2} \asymp \frac{n^{2/u-1/2}}{\norm{\sum_1^n e_i}_{E_n}} =\frac{n^{2/r+1/2}}{\norm{\sum_1^n e_i}_{E_n}}, $$ and in order to obtain the lower estimate apply again \eqref{hilfe2} together with \eqref{weyl} and the second asymptotic in \eqref{lu}: $$ \pi_{r,2}(\mathcal{S}_E^n \hookrightarrow \mathcal{S}_2^n) \ge [n^2/2]^{1/r} \cdot x_{[n^2/2]}(\mathcal{S}_E^n \hookrightarrow \mathcal{S}_2^n) \succ \frac{n^{2/r}}{\norm{\ell_2^n \hookrightarrow E_n}} \asymp \frac{n^{2/r+1/2}}{\norm{\sum_1^n e_i}_{E_n}}. $$ \qed \section{\sc Applications} \subsection*{Weyl Numbers} The results of the preceding sections can be used to improve the estimates for Weyl numbers of identities on symmetric Banach sequence spaces and unitary ideals in \eqref{hilfe2}: The exponent $1/2$ in each of the two inequalities there can be replaced by $1/u-1/2$ whenever $u$-convexity and $2$-concavity assumptions are made. \begin{cor} \label{weylcor} For $1 \le u,v \le 2$ let $E$ and $F$ be $2$-concave symmetric Banach sequence spaces where $E$ is $u$-convex and $F$ is $v$-convex. Then there exist constants $C_u,C_v>0$ such that for all $1 \le k \le n$ $$ C_v^{-1} \cdot \left (\frac{n-k+1}{n} \right )^{1/v-1/2} \cdot \frac{\norm{\sum_1^n e_i}_{F_n}}{\norm{\sum_1^n e_i}_{E_n}} \le x_k(E_n \hookrightarrow F_n) \le C_u \cdot \left (\frac{n}{k} \right )^{1/u-1/2} \cdot \frac{\norm{\sum_1^n e_i}_{F_n}}{\norm{\sum_1^n e_i}_{E_n}}, $$ and all $1 \le k \le n^2$ $$ C_v^{-1} \cdot \left (\frac{n^2-k+1}{n^2} \right )^{1/v-1/2} \cdot \frac{\norm{\sum_1^n e_i}_{F_n}}{\norm{\sum_1^n e_i}_{E_n}} \le x_k(\mathcal{S}_E^n \hookrightarrow \mathcal{S}_F^n) \le C_u \cdot \left (\frac{n^2}{k} \right )^{1/u-1/2} \cdot \frac{\norm{\sum_1^n e_i}_{F_n}}{\norm{\sum_1^n e_i}_{E_n}}. $$ \end{cor} \proof The upper estimates follow by using the inequality~\eqref{weyl} and the results from the preceding two sections, and the lower estimates then are immediate consequences of the upper ones---simply repeat the proof of \eqref{hilfe2} with a different exponent. \qed \\[10pt] Recall that for the embedding $\ell_u^n \hookrightarrow \ell_2^n$, $1 \le u \le 2$ by \cite[2.3.3]{CD92} even the following equality is known: $x_k(\ell_u^n \hookrightarrow \ell_2^n)=k^{1/2-1/u}$, $1 \le k \le n$. The second estimate in corollary~\ref{weylcor} implies that for $1 <u<2$ $$ x_k(\mathcal{S}_u^n \hookrightarrow \mathcal{S}_2^n) \le C_u \cdot \left ( \frac{n}{k} \right)^{1/u-1/2}, $$ hence by \cite[2.3.2]{CD92} $$ a_k(\mathcal{S}_2^n \hookrightarrow \mathcal{S}_u^n) \ge C_u^{-1} \cdot \left (\frac{n^2-k+1}{n}\right)^{1/u-1/2}. $$ This disproves the conjecture $$ a_k(\mathcal{S}_2^n \hookrightarrow \mathcal{S}_u^n) \asymp \max \left (1,\left (\frac{n^2-k+1}{n^2}\right)^{1/2} \cdot n^{1/u-1/2} \right) $$ from \cite[p.~249]{CD97} (put $k:=[n^2-n^\alpha+1]$, $1 < \alpha <2$). \subsection*{Identities on Orlicz and Lorentz Sequence Spaces} In the following we apply our results, in particular the corollaries~\ref{precise} and \ref{unitary}, to two natural examples of symmetric sequence spaces: Orlicz and Lorentz sequence spaces (for their definition and basic properties we refer to \cite{lt77}). We only treat the case where the range space of the embedding is the finite-dimensional Hilbert space and leave the formulation for other spaces and the corollaries for Weyl numbers to the reader. \par Let us start with Orlicz sequence spaces $\ell_M$. \begin{cor} Let $1 <u < 2$ and $M$ be a strictly increasing Orlicz function which satisfies the $\Delta_2$-condition at zero. Assume that there exists $K>0$ such that for all $s,t \in (0,1]$ \begin{equation} K^{-1} \cdot s^2 \le {M(st)}/{M(t)} \le K \cdot s^u. \label{orlicz} \end{equation} Then for $2 < r,s < \infty$ such that $1/r >1/u-1/2$ and $1/s=1/2 -1/r$ $$\pi_{r,2}(\ell_M^n \hookrightarrow \ell_2^n) \asymp \mu_{s,2}(\ell_M^n \hookrightarrow \ell_2^n) \asymp \frac{n^{1/r+1/2}}{\norm{\sum_1^n e_i}_{\ell_M^n}} \asymp n^{1/r+1/2} \cdot M^{-1}(1/n) $$ and $$\pi_{r,2}(\mathcal{S}_{\ell_M}^n \hookrightarrow \mathcal{S}_2^n) \asymp \mu_{s,2}(\mathcal{S}_{\ell_M}^n \hookrightarrow \mathcal{S}_2^n) \asymp \frac{n^{2/r+1/2}}{\norm{\sum_1^n e_i}_{\ell_M^n}} \asymp n^{2/r+1/2} \cdot M^{-1}(1/n). $$ \end{cor} Note that \eqref{orlicz} together with the $\Delta_2$-condition assures that $\ell_M$ is $2$-concave and $p$-convex for all $1 \le p < u$ (see \cite[2.b.5]{lt}). \par Now we state an analogue for Lorentz sequence spaces $d(w,u)$. \begin{cor} \label{lorentzcor} Let $1 < u <2$ and $w$ be such that $ n \cdot w_n^q \asymp \sum_{i=1}^n w_i^q,$ where $q=2/(2-u)$. Then for $2 < r,s < \infty$ such that $1/r \ge 1/u -1/2$ and $1/s=1/2-1/r$ $$\pi_{r,2}(d_n(w,u) \hookrightarrow \ell_2^n) \asymp \mu_{s,2}(d_n(w,u) \hookrightarrow \ell_2^n) \asymp n^{1/r +1/2 -1/u} \cdot w_n^{-{1/u}}. $$ and $$\pi_{r,2}(\mathcal{S}_{d(w,u)}^n \hookrightarrow \mathcal{S}_2^n) \asymp \mu_{s,2}(\mathcal{S}_{d(w,u)}^n \hookrightarrow \mathcal{S}_2^n) \asymp n^{2/r +1/2 -1/u} \cdot w_n^{-{1/u}}. $$ \end{cor} Recall that the space $d(w,u)$ is $u$-convex, and if $1 \le u<2$, it is $2$-concave if and only if $w$ satisfies the condition in the assumption of the corollary (see \cite[p.~245--247]{reisner}). \subsection*{Limit Orders} Finally, we consider the asymptotic order of the sequences $(\pi_{r,2} (\mathcal{S}_u^n \hookrightarrow \mathcal{S}_v^n))_n$ for arbitrary $2 \le r \le \infty$, $1 \le u,v \le \infty$. Define the limit orders $$\lambda_\ell(\Pi_{r,2},u,v) := \inf \{ \lambda>0 \, | \, \exists \, \rho>0 \, \forall \, n: \pi_{r,2}(\ell_u^n \hookrightarrow \ell_v^n) \le \rho \cdot n^\lambda\}$$ and $$\lambda_{\mathcal{S}}(\Pi_{r,2},u,v) := \inf \{ \lambda>0 \, | \, \exists \, \rho>0 \, \forall \,n: \pi_{r,2}(\mathcal{S}_u^n \hookrightarrow \mathcal{S}_v^n) \le \rho \cdot n^\lambda\}.$$ Here we only handle the limit order of summing operators since---using the fact that $\Pi_{r,2}$ and ${\cal M}_{s,2}$ for $1/s + 1/r = 1/2$ are almost equal---one can easily see that $\lambda_\ell(\Pi_{r,2},u,v)= \lambda_\ell({\cal M}_{r,2},u,v)$ and $\lambda_{\mathcal{S}}(\Pi_{r,2},u,v)= \lambda_{\mathcal{S}}({\cal M}_{r,2},u,v)$ (with the obvious definition for the right sides of these equalities; see \cite[22.3.7]{pietsch}). \par The calculation of the limit order $\lambda_\ell(\Pi_{r,2},u,v)$ was completed in \cite{CMP}: \begin{center} \small \unitlength3pt \begin{picture}(60,50) \put(10,10){\framebox(40,40)} \put(30,10){\line(0,1){40}} \put(30,30){\line(1,0){10}} \put(40,10){\line(0,1){20}} \put(40,30){\line(1,1){10}} \put(10,20){\line(2,1){20}} \put(40,10){\makebox(10,22){$0$}} \put(30,10){\makebox(10,20){\shortstack{$\frac{1}{r}+ \frac{1}{2}$ \\ $-\frac{1}{u}$}}} \put(30,35){\makebox(20,15){$\frac{1}{r} + \frac{1}{v}-\frac{1}{u}$}} \put(10,10){\makebox(20,12){$\frac{1}{r}$}} \put(10,30){\makebox(20,20){$\frac{1}{v}- (1-\frac{2}{r} ) \frac{1}{u}$}} \put(5,20){\makebox(0,0){$\frac{1}{r}$}} \put(30,5){\makebox(0,0){$\frac{1}{2}$}} \put(40,5){\makebox(0,0){$\frac{1}{2} + \frac{1}{r}$}} \put(55,40){\makebox(0,0){$\frac{1}{r'}$}} \put(7,53){\makebox(0,0){$1/v$}} \put(53,7){\makebox(0,0){$1/u$}} \end{picture} \end{center} Moreover, the proof in \cite{CMP} shows that the limit order is attained: $\pi_{r,2}(\ell_u^n \hookrightarrow \ell_v^n) \asymp n^{\lambda_\ell(\Pi_{r,2},u,v)}$. In view of the results of section~4 the following conjecture seems to be natural: \\[10pt] {\bf Conjecture:} $\lambda_{\mathcal{S}}(\Pi_{r,2},u,v)=1/r + \lambda_\ell(\Pi_{r,2},u,v)$. \par For the border cases $r=2$ (the $2$-summing norm) and $r=\infty$ (the operator norm) this conjecture by \eqref{hilfe1} and \eqref{formel} is true. In the following corollary we confirm the upper estimates of this conjecture for all $u,v$ and the lower ones for all $u,v$ except those in the upper left corner of the picture. \begin{cor} Let $1 \le u ,v\le $ and $2 < r < \infty$. \vspace{-10pt} \begin{enumerate}[(a)] \item $\lambda_{\mathcal{S}}(\Pi_{r,2},u,v)=1/r + \lambda_\ell(\Pi_{r,2},u,v)$ \quad for $1 \le u \le 2$. \item $\lambda_{\mathcal{S}}(\Pi_{r,2},u,v) \le 1/r + \lambda_\ell(\Pi_{r,2},u,v)$ \quad for $2 \le u \le \infty$, with equality whenever \mbox{$1/v \le 1/r + (1- 2/r)(1/u)$.} \end{enumerate} \end{cor} \proof Let $1/s:=1/2-1/r$. The upper estimates for the case $1 \le u \le 2$ follow from corollary~\ref{schatten}: Consider for $u_0 :=(1/2-1/r)^{-1}$ the following alternative: (i) $1/u \le 1/{u_0}$ or (ii) $1/u > 1/{u_0}$. Then the conclusion in case~(i) is a consequence of corollary~\ref{schatten} and the following factorization: $$ \mu_{s,2}(\mathcal{S}_u^n \hookrightarrow \mathcal{S}_v^n) \le \norm{\mathcal{S}_u^n \hookrightarrow \mathcal{S}_{u_0}^n} \cdot \mu_{s,2}(\mathcal{S}_{u_0}^n \hookrightarrow \mathcal{S}_2^n) \cdot \norm{\mathcal{S}_2^n \hookrightarrow \mathcal{S}_v^n} \prec n^{2/r +1/2-1/u + \max(0,1/v-1/2)}, $$ and for (ii) look with $v_0:= (1/u-1/r)^{-1} \le 2$ at \begin{align*} \mu_{s,2}(\mathcal{S}_u^n \hookrightarrow \mathcal{S}_v^n) \le \mu_{s,2}(\mathcal{S}_u^n \hookrightarrow \mathcal{S}_{v_0}^n) \cdot \norm{\mathcal{S}_{v_0}^n \hookrightarrow \mathcal{S}_v^n} \prec n^{1/r + \max(0,1/r+1/v-1/u)}. \end{align*} Now let $2 \le u \le \infty$. Although this part is very close to the calculations made in \cite[Lemma~6]{CMP}, we give a short sketch of the proof for the convenience of the reader. By \eqref{formel} and theorem~\ref{mixing} (with no interpolation in the range or the image), \begin{align*} \mu_{s,2}(\mathcal{S}_u^n \hookrightarrow \mathcal{S}_2^n) \le \pi_2(\mathcal{S}_u^n \hookrightarrow \mathcal{S}_2^n)^{2/r} \cdot \norm{\mathcal{S}_u^n \hookrightarrow \mathcal{S}_2^n}^{1-{2/r}}= n^{1/r +1/2 -(1-2/r)(1/u)}, \end{align*} hence, by factorization, for $1 \le v \le 2$ $$ \mu_{s,2}(\mathcal{S}_u^n \hookrightarrow \mathcal{S}_v^n) \le n^{1/r + 1/v -(1-{2/r})(1/u)}. $$ Furthermore, for $1/{v_1} := 1/r + (1-{2/r})(1/u)$ \begin{align*} \mu_{s,2}(\mathcal{S}_u^n \hookrightarrow \mathcal{S}_{v_1}^n ) \le \pi_2(\mathcal{S}_u^n \hookrightarrow \mathcal{S}_2^n)^{2/r} \cdot \norm{\mathcal{S}_u^n \hookrightarrow \mathcal{S}_u^n}^{1-{2/r}} = n^{2/r}, \end{align*} hence $$\mu_{s,2}(\mathcal{S}_u^n \hookrightarrow \mathcal{S}_v^n) \le n^{2/r} $$ for all $v_1 \le v \le \infty$. Finally, for all $2 < v < v_1$ and $0< \theta <1$ such that $1/v= (1-\theta)/{v_1} + \theta/2$ \begin{align*} \mu_{s,2}(\mathcal{S}_u^n \hookrightarrow \mathcal{S}_v^n) &\le \mu_{s,2}(\mathcal{S}_u^n \hookrightarrow \mathcal{S}_{v_1}^n) ^{1-\theta} \cdot \mu_{s,2}(\mathcal{S}_u^n \hookrightarrow \mathcal{S}_2^n)^\theta \\ & \le n^{1/r + (1-\theta)/r + \theta(1/2 -(1-{2/r})(1/u))} = n^{1/r + 1/v -(1-{2/r})(1/u)}. \end{align*} Looking at the picture for $\lambda_\ell(\Pi_{r,2}, u,v)$ one can see that these are the desired results. For the lower estimates recall \eqref{weyl}: $$ [{n^2}/2 ]^{1/r} \cdot x_{[{n^2}/2 ]}(\mathcal{S}_u^n \hookrightarrow \mathcal{S}_v^n) \le \pi_{r,2}(\mathcal{S}_u^n \hookrightarrow \mathcal{S}_v^n),$$ hence \eqref{weylschatten} implies $$\pi_{r,2}(\mathcal{S}_u^n \hookrightarrow \mathcal{S}_v^n) \succ \begin{cases} n^{2/r + 1/v - 1/u} & \text{if $1 \le u,v \le 2$,}\\ n^{2/r + 1/2 -1/u} & \text{if $1 \le u \le 2 \le v \le \infty$,} \\ n^{2/r} & \text{if $2 \le u,v \le \infty$.} \end{cases} $$ Using \eqref{low}, these estimates can be improved for those $u,v$ for which $\lambda_\ell(\Pi_{r,2},u,v)=0$. \qed \par Our results for $\lambda_{\mathcal{S}}(\Pi_{r,2},u,v)$ can be summarized in the following picture: \begin{center} \small \unitlength3pt \begin{picture}(60,60) \put(10,10){\framebox(40,40)} \put(30,10){\line(0,1){40}} \put(30,30){\line(1,0){10}} \put(40,10){\line(0,1){20}} \put(40,30){\line(1,1){10}} \put(10,20){\line(2,1){20}} \put(40,10){\makebox(10,22){$\frac{1}{r}$}} \put(30,10){\makebox(10,20){\shortstack{$\frac{2}{r}+ \frac{1}{2}$ \\ $-\frac{1}{u}$}}} \put(30,35){\makebox(20,15){$\frac{2}{r} + \frac{1}{v}-\frac{1}{u}$}} \put(10,10){\makebox(20,12){$\frac{2}{r}$}} \put(10,30){\makebox(20,20){\shortstack{$ \le \frac{1}{r} + \frac{1}{v}$ \\ $ \text{ } - (1-\frac{2}{r}) \frac{1}{u}$}}} \put(5,20){\makebox(0,0){$\frac{1}{r}$}} \put(30,5){\makebox(0,0){$\frac{1}{2}$}} \put(40,5){\makebox(0,0){$\frac{1}{2} + \frac{1}{r}$}} \put(55,40){\makebox(0,0){$\frac{1}{r'}$}} \put(7,53){\makebox(0,0){$1/v$}} \put(53,7){\makebox(0,0){$1/u$}} \end{picture} \end{center} \par The contents of this article will be part of the second named author's thesis written at the Carl von Ossietzky University of Oldenburg under the supervision of the first named author. \newcommand{\etalchar}[1]{$^{#1}$} \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
\section{Introduction} The strong interaction at low energies, i.e. elastic nucleon-nucleon scattering, is reasonably described by $\pi$, $\sigma$, $\eta$, $\rho$ and $\omega$-meson exchanges between the nucleons and the detailed experimental data on $NN{\to}NN$ reactions have provided information about the meson-nucleon-nucleon vertices, i.e. coupling constants and form factors. Above the pion production threshold the dominant inelasticy of the $NN$ interactions is due to pion production. Already in 1960 Woodruff~\cite{Woodruff} proposed to extend the $NN$ potential model in order to calculate $NN{\to}NN{\pi}$ reactions. Near the reaction threshold the contribution from $\Delta$ intermediate states is expected to be negligible and the $S$-wave pion production is governed by the ${\pi}NN$ vertex. Thus pion production is suited to verify our knowledge about the ${\pi}NN$ coupling constant. A similar motivation also holds for near threshold $\eta$-meson production, when the $S_{11}$ resonance replaces the $\Delta$, and the $\eta$ production cross section should provide some information about the ${\eta}NN$ vertex. Note that the status of the ${\eta}NN$ coupling constant is still an open problem~\cite{Tiator,Feuster,Sibirtsev1} since within our present knowledge $g_{{\eta}NN}$ might vary between 1 and 9 depending on the model adopted as well as the accuracy of the experimental data. Near threshold $\omega$, $\phi$ and $\eta^\prime$-meson production in $NN$ collisions should provide information about the relevant $MNN$ couplings as well as on intermediate baryonic resonances that might be coupled strongly to these mesons; this is discussed as $hidden$ resonance properties. Obviously the strangeness production in $NN$ collisions involves an additional mechanism due to strange meson exchange ($K, K^*$) and sheds light on the kaon-hyperon-nucleon vertex. We will base our analysis in this work on the combined efforts of many experimental groups that have taken data on near threshold meson production: These type of experiments for $NN$ collisions were started at the Indiana University Cyclotron Facility with data on the $pp{\to}pp\pi^0$ reaction at excess energies $\epsilon{=}\sqrt{s}{-}2m_N{-}m_\pi$ from $\simeq$1 to $\simeq$30~MeV~\cite{Meyer1,Meyer2}. The data at $\epsilon{\leq}$1~MeV were complemented by CELSIUS (Uppsala)~\cite{Bondar}; recently also IUCF reported~\cite{Hardie,Flammang} new cross sections on the $pp{\to}pn\pi^+$ reaction at $\epsilon{<}$20~MeV. The near threshold $\eta$-meson production in $pp$ collisions was studied at SATURNE by the collaborations SATURNE-II~\cite{Bergdolt} and PINOT~\cite{Chiavassa} and at CELSIUS~\cite{Calen1}. These measurements cover the range ${\simeq}1.5{\le}\epsilon{\le}100$~MeV. In 1998 CELSIUS reported also data~\cite{Calen2} on the $pn{\to}pn\eta$ reaction at $16{\le}\epsilon{\le}100$~MeV. Additionally, the $pp{\to}pp\eta^\prime$ reaction was studied at SATURNE by SPES-III~\cite{Hibou} and at the COoler SYnchrotron (J\"ulich) by COSY-11~\cite{Moscal} at $\epsilon{<}$10~MeV. Furthermore, the $pp{\to}pp\omega$ reaction was measured at SATURNE by the DISTO Collaboration; they reported~\cite{Balestra} data on $\omega$-meson production from $pp$ collisions at $\epsilon{\simeq}320$~MeV and $\phi$ production at $\epsilon{\simeq}82$~MeV. The data on the $pp{\to}pp\omega$ reaction at $\epsilon{<}$31~MeV were measured by SPESIII and have been reported only very recently~\cite{Hibou1}. The $pp{\to}p{\Lambda}K^+$ reaction was measured by the COSY-11~\cite{Balewski} and the COSY-TOF~\cite{Bilger} Collaborations; COSY-11 also has reported on the $pp{\to}p\Sigma^0K^+$ reaction~\cite{Sewerin}. It should be noted that apart from the $\pi^0$ and $\eta$ data the experimental results on near threshold meson production in $NN$ collisions have became available only during the last years. This has initialized a lot of theoretical activity and inspired the most recent calculations within meson-exchange models. Here we present a systematical analysis of the data and provide the relation between the experimental observables and the production mechanism, respectively. Our work is organized as follows: In Section 2 we will describe the threshold kinematics and discuss various approaches for the final state interactions (FSI). Section 3 is devoted to an analysis of the available data with the aim to extract average production matrix elements for the mesons measured so far. In Section 4 we will discuss the effect of FSI and resonance amplitudes on differential observables while Section 5 concludes this study with a summary. \begin{figure}[t] \label{memo9} \centerline{ \psfig{file=memo9.ps,height=8cm,width=13.5cm}} \phantom{aa}\vspace{-0.8cm} \caption[]{Experimental data on the $pp{\to}pp\pi^0$ \protect\cite{Meyer1,Meyer2,Bondar,LB} and $pp{\to}pn\pi^+$ \protect\cite{Hardie,Flammang,LB} reaction amplitude as a function of the excess energy $\epsilon$.} \phantom{aa}\vspace{-0.7cm} \end{figure} \section{Threshold kinematics and Final State Interactions} The threshold kinematics have several features that request specific conditions for the experimental measurements as well as the theoretical analysis. We note that within nonrelativistic approaches the three-body phase space $\Phi_3$ is proportional to $\epsilon^2$. Thus data on threshold meson production are frequently analyzed in terms of the $reduced$ $cross$ $section$ $\sigma{/}\epsilon^2$. In~\cite{Sibirtsev2} we have proposed to analyze the data in a more transparent way in terms of an average reaction amplitude (for fixed invariant energy $\sqrt{s}$) as \begin{eqnarray} \label{average} |M_R| &=& 2^4 \ \pi^{3/2} \ \lambda^{1/4}(s,m_N^2,m_N^2) \ \sqrt{\sigma s} \nonumber \\ &\times& \left\lbrack \ \intop^{(\sqrt{s}-m_a)^2}_{(m_b+m_c)^2} \hspace{-0.3cm} \lambda^{1/2}(s,s_1,m_a^2) \ \lambda^{1/2}(s_1,m_b^2,m_c^2) \ \frac{ds_1}{s_1} \right\rbrack^{-1/2} \end{eqnarray} with ${\lambda}(x,y,z){=}(x{-}y{-}z)^2{-}4yz$ and $m_a, m_b, m_c$ denoting the masses of the particles in the final state. Furthermore, among the five variables characterizing the three-body final state, there are two of direct physical relevance: the invariant mass $\sqrt{s_1}$ of two final particles $b$ and $c$ and the 4-momentum squared $t$ transfered from the initial nucleon to particle $a$. These variables allow to express the production amplitude in the meson-exchange mechanism. Note that $\sqrt{s_1}$ varies from $m_b{+}m_c$ up to $m_b{+}m_c{+}\epsilon$. Since the width of the known baryonic resonances is larger than 100~MeV, it is not possible - within a narrow $\epsilon$ range - to detect directly an intermediate baryonic resonance coupled to $bc$ (meson + nucleon or hyperon) and to reconstruct experimentally the relevant production mechanism~\cite{Sibirtsev1,Sibirtsev3}. Therefore complete measurements have to be performed at least up to $\epsilon{\simeq}$100~MeV. Close to threshold both $\sqrt{s_1}$ and $t$ vary only slightly and the production amplitude itself is expected to be almost constant. Fig. \ref{memo9} shows the amplitudes for the $pp{\to}pp\pi^0$ and $pp{\to}pn\pi^+$ reactions extracted from the experimental data~\cite{Meyer1,Meyer2,Bondar,Hardie,Flammang,LB} using Eq. (\ref{average}). The amplitudes substantially depend on the excess energy $\epsilon$ but seem to approach a constant value for large excess energies. Such a deviation from a constant value has been predicted by Watson~\cite{Watson} and Migdal~\cite{Migdal} due to the strong $S$-wave interaction between the final nucleons. Indeed the Watson-Migdal theorem can be understood, for instance, in terms of the $pp$ cross section shown in Fig.\ref{memo6-8}a) as a function of the proton momentum $q$ in the center-of-mass system. The cross section is enhanced at low $q$ due to the $^1S_0$ partial wave~\cite{SP99} as shown by the solid line in Fig. \ref{memo6-8}a). Above about 400 MeV/c the elastic cross section approaches again a constant as indicated by the dashed line. It is thus expected that the production of mesons is enhanced when the protons emerge with a low relative momentum in the final state. \begin{figure}[h] \phantom{aa}\vspace{-0.4cm} \centerline{ \begin{minipage}[l]{6cm} \psfig{file=memo8.ps,width=6.2cm,height=8.6cm} \end{minipage}\begin{minipage}[l]{6cm} \psfig{file=memo6.ps,width=6.2cm,height=8.6cm} \end{minipage}} \phantom{aa}\vspace{-0.6cm} \caption[]{a) Total (circles) and elastic (squares) cross sections for the $pp$ interaction as a function of the momentum in the $pp$ cms. The data are from Ref. \protect\cite{LB}. The solid line shows the contribution from the $^1S_0$ partial wave~\protect\cite{SP99}, while the dashed line indicates the large momentum limit. b) Correction factor due to final-state-interactions (FSI). The squared $pp$ scattering amplitudes are shown for $^1S_0$ (circles), $^3P_0$ (squares) and $^3P_1$ (triangles) partial waves~\protect\cite{Nijmegen}. The dotted line shows the result from the effective range approximation, the dashed line shows the inverse squared Jost function without Coulomb correction, while the solid line includes a Coulomb correction. Further notations are explained in the text.} \label{memo6-8} \end{figure} Note that in the $NN{\to}NNM$ reaction the momentum $q$ varies from zero up to ${\simeq}\sqrt{m_N\epsilon}$. Obviously at large excess energies the contribution from FSI due to the strong $S$-wave to the total $pp{\to}pp\pi^0$ cross section seems to be not dominant, since one should integrate over the wide phase space. However, $S$-wave FSI can be detected by differential observables even at large $\epsilon$ as we will illustrate in the following. For $\epsilon{\le}25$~MeV the FSI between the protons is entirely due to the $^1S_0$-wave. At higher energies the $pp$ cross section deviates from the calculations with the $^1S_0$ phase shift as can be seen from Fig.\ref{memo6-8}a) and is indicated by an arrow. Now the deviation of the $NN{\to}NN\pi$ reaction amplitude shown in Fig.\ref{memo9} from a constant can be understood within the Watson-Migdal approximation. Moreover, the difference in the energy dependence of the $pp{\to}pp\pi^0$ and $pp{\to}pn\pi^+$ reaction amplitudes at $\epsilon{\le}1$~MeV is due to the Coulomb interaction between the final nucleons, which is absent in $np$ scattering, but influences the final $pp$ scattering for $\epsilon{\le}1$~MeV as can be seen from Fig.\ref{memo6-8}b. Taking the near threshold production amplitude $M$ as a constant, it was proposed in Refs.\cite{Watson,Migdal} to factorize the reaction amplitude $M_R$ as \begin{equation} \label{wm} M_R = M \times C_{FSI} , \end{equation} where $C_{FSI}$ stands for the amplitude due to the interaction between the final particles. Strictly one should account for the three-body FSI, which itself is a rigorous problem. As was suggested by Gell-Mann and Watson~\cite{GellMann} the near threshold $NN{\to}NN\pi$ reaction might be examined when considering the dominance of low energy $NN$ scattering as compared to the $S$-wave ${\pi}N$ interaction and taking $C_{FSI}$ as the $S$-wave $NN$ on-shell scattering amplitude $T_s$. Obviously, the produced particles are off-shell before rescattering due to FSI, which in principle involves an additional assumption about the off-shell correction to $T_s$. Fig.\ref{memo6-8}b) shows the squared $^1S_0$, $^3P_0$ and $^3P_1$ $pp$ scattering amplitudes calculated with the phase shifts from the Nijmegen partial wave analysis~\cite{Nijmegen}. At low $\epsilon$ the $S$-wave amplitude dominates and for further implementation to $NN{\to}NNM$ calculations can be expressed within the effective range approximation as \begin{equation} \label{scat} T_s(q) = \left( -\frac{1}{a_s}+\frac{r_s q^2}{2}-iq \right)^{-1}, \end{equation} where $a_s{=}{-}7.8$~fm and $r_s{=}2.79$~fm~\cite{Stoks} denote the scattering length and effective range, respectively. The effective range approximation is shown in Fig.\ref{memo6-8}b) by the dotted line and is valid for excess energies $\epsilon$ from 1 up to 40~MeV. Another way~\cite{Taylor} to account for FSI is to express $C_{FSI}$ as an inverse $S$-wave Jost function \begin{equation} \label{Jost} C_{FSI}(q) = \frac{q+i\beta}{q-i\alpha}, \end{equation} where the parameters $\alpha$ and $\beta$ are related to the effective range parameters as \begin{equation} \label{range} a_s=\frac{\alpha + \beta}{\alpha \beta}, \hspace{2cm} r_s= \frac{2}{\alpha + \beta}. \end{equation} The squared inverse Jost function is shown by the dashed line in Fig.\ref{memo6-8}b) and is close to the effective range approximation only for $\epsilon{\le}5$~MeV. Note that Eq.(\ref{Jost}) approaches unity at large momenta $q$ since the $S$-wave FSI does not contribute at large $q$, which is the proper boundary condition in terms of the factorization (\ref{wm}). Furthermore, to account for the Coulomb repulsion at $\epsilon{\le}1$ one can correct $C_{FSI}$ in line with the Gamov factor (solid line in Fig. 2b). Finally, when calculating the FSI within different approaches as the $NN$ scattering amplitude itself or with the Jost function or an effective range approximation including Coulomb corrections we find no severe differences up to excess energies of ${\simeq}5$~MeV. Furthermore, since the $S$-wave dominates the $NN$ scattering up to $\epsilon{\simeq}25$~MeV, the Jost function is an appropriate way to account for FSI corrections because it approaches unity at large $\epsilon$ in line with the factorization ansatz. The disadvantage of the method is due to the implementation of the on-shell $NN{\to}NN$ amplitude. However, off-shell corrections will introduce new parameters to the calculations that later on should be controlled by data. \begin{figure}[t] \psfig{file=memo3.ps,width=13.0cm,height=8cm} \phantom{aa}\vspace{-1.1cm} \caption[]{Experimental data~\protect\cite{Meyer1,Meyer2,Bondar,LB} on the average $pp{\to}pp\pi^0$ production amplitude as a function of the excess energy $\epsilon$ calculated without (upper part) and with Coulomb correction (lower part). The solid lines show the fit with a constant value for the production matrix element.} \label{memo3} \phantom{aa}\vspace{-0.8cm} \end{figure} \section{Evaluation of the production amplitude from the data} Now we adopt the on-shell approach and use the Jost function in order to account for the FSI correction. Moreover, we perform the data analysis with and without Coulomb correction to demonstrate the systematic uncertainties. To calculate the production amplitude $|M|$ we substitute the function $C_{FSI}(q)$ in the integral of Eq.(\ref{average}). Fig.\ref{memo3} shows the average $pp{\to}pp\pi^0$ production amplitude as a function of the excess energy $\epsilon$. In this representation the data are almost energy independent and approach a constant value. For $\epsilon{<}1$~MeV two data points from Ref.\cite{Bondar} substantially deviate from the constant for calculations without the $pp$ Coulomb repulsion, but become closer to a constant value after Coulomb correction. However, to shed light on the Coulomb effect one needs more data at $\epsilon{<}1$~MeV. We also notice that the 1992 IUCF data~\cite{Meyer2} are better described by a constant amplitude $|M|$ as compared to the 1990 IUCF data~\cite{Meyer1}. Our analysis with Coulomb correction gives $|M|\approx 82$~fm for the $pp{\to}pp\pi^0$ reaction while we get $|M| \approx$ 79 fm without this correction which indicates the systematic uncertainty of our analysis. \begin{figure}[h] \phantom{aa}\vspace{-0.8cm} \centerline{ \begin{minipage}[l]{6cm} \psfig{file=memo2.ps,width=6.2cm,height=8cm} \end{minipage}\begin{minipage}[l]{6cm} \psfig{file=memo1.ps,width=6.2cm,height=8cm} \end{minipage}} \phantom{aa}\vspace{-0.5cm} \caption[]{Experimental data on the average $pp{\to}pp\eta$~\protect\cite{Bergdolt,Chiavassa,Calen1} and $pp{\to}pp\eta^\prime$~\protect\cite{Hibou,Moscal} production amplitudes calculated with and without Coulomb repulsion. The solid lines show the fit with a constant value for the production matrix element.} \label{memo2-1} \phantom{aa}\vspace{-0.7cm} \end{figure} In a similar way we evaluate the average production amplitude from the total cross sections for the $pp{\to}pp\eta$~\protect\cite{Bergdolt,Chiavassa,Calen1}, $pp{\to}pp\omega$~\cite{Hibou1} and $pp{\to}pp\eta^\prime$~\protect\cite{Hibou,Moscal} reactions and show the result in Figs.\ref{memo2-1},\ref{memo19}. The results for the $pp{\to}pp\omega$ reaction are shown for a fixed $\omega$-meson pole mass (squares) and for the calculation with a Breit-Wigner $\omega$ spectral function (circles), which is explicitely given as \begin{eqnarray} \label{wigner} |M_R|= 2^{9/2} \ \pi^2 \ \lambda^{1/4}(s,m_N^2,m_N^2) \ \sqrt{\sigma s} \left\lbrack \ \intop^{\sqrt{s}-2m_N}_{m_\pi}\hspace{-0.3cm} \frac{\Gamma \ dx}{(x-m_\omega)^2+\Gamma^2/4} \right. \nonumber \\ \times \left. \ \hspace{-0.2cm} \intop^{(\sqrt{s}-x)^2}_{4m_N^2} \hspace{-0.3cm} \lambda^{1/2}(s,s_1,x^2) \ \lambda^{1/2}(s_1,m_N^2,m_N^2) \ \ \left| C_{eff}(0.5\sqrt{s_1-4m_N^2}) \right|^2 \ \ \frac{ds_1}{s_1} \right\rbrack^{-1/2} \end{eqnarray} with the vacuum $\omega$-meson width $\Gamma{=}8.41$~MeV. Again the deviation of the matrix element $|M|$ from a constant seems to be small for $\eta$, $\omega$ and $\eta^\prime$ production in $pp$ collisions. The data are only available for $\epsilon{>}1$~MeV and thus we can not observe the effect of the Coulomb $pp$ final state repulsion. The calculations with and without Coulomb correction provide almost the same results for the production amplitudes, i.e. $|M|{\approx}61$~fm for the $\eta$, $|M|{\approx}33$~fm for the $\omega$, and $|M|{\approx}19.0$~fm for the $\eta^\prime$-meson. Note that in case of the $\omega$ meson it is essential to account for the finite width of the spectral function close to threshold. \begin{figure}[h] \psfig{file=memo19.ps,width=12.6cm,height=8cm} \phantom{aa}\vspace{-0.8cm} \caption[]{Experimental data on the average $pp{\to}pp\omega$~\protect\cite{Hibou1} production amplitudes calculated with and without Coulomb repulsion, for fixed $\omega$ mass (squares) and for the finite spectral function of the $\omega$-meson with a width of 8.41 MeV (circles). The solid lines show the fit with a constant value for the production matrix element.} \label{memo19} \phantom{aa}\vspace{-0.5cm} \end{figure} Recently IUCF published data on the $pp{\to}pn\pi^+$~ reaction \cite{Hardie,Flammang} and CELSIUS reported $pn{\to}pn\eta$~\cite{Calen2} total cross sections. Both reactions are crucial for the verification of our approach, since the final $np$ system does not suffer Coulomb repulsion as in case of the meson production data at $\epsilon{\le}1$~MeV. Fig.\ref{memo7} shows the $pp{\to}pn\pi^+$ and $pn{\to}pn\eta$ production amplitude extracted by Eq.\ref{average} with inclusion of the $np$ FSI. Indeed, the two experimental points available at $\epsilon{\le}1$~MeV as well as the data for the $pp{\to}pn\pi^+$ cross section at higher exess energies are reproduced by a constant value of $|M|\approx 234$~fm. Fig.\ref{memo7} illustrates that the data for the $pn{\to}pn\eta$ reaction can be reasonably described by $|M|\approx 157$~fm. Finally, the simple approach outlined above allows to evaluate the average production amplitudes from the total cross sections for $NN{\to}NNM$ reactions and enables one to substract the FSI due to $NN$ rescattering. The systematical analysis of the available experimental data on $\pi^0$, $\pi^+$, $\eta$ and $\eta^\prime$-meson production confirms the validity of the method proposed. Furthermore, the results illustrate a sensitivity to the difference between the $pp$ and $pn$ interactions in the final state and can be tested by data at excess energies below 1~MeV. Since the effective range parameters are the essential ingredients for our calculations, the method should be limited to $\epsilon{\leq}40$~MeV (see Fig.\ref{memo6-8}). However, at $\epsilon{\ge}40$~MeV the $NN$ scattering amplitude is almost energy independent and approaches a constant value, which might provide an explanation for the observation that the method seems to work even at higher energies. \begin{figure}[h] \phantom{aa}\vspace{-0.5cm} \psfig{file=memo7.ps,width=12.6cm,height=8cm} \phantom{aa}\vspace{-0.72cm} \caption[]{The data on the $pp{\to}pn\pi^+$~\protect\cite{Hardie,Flammang} and $pn{\to}pn\eta$~\protect\cite{Calen2} production amplitudes.} \label{memo7} \phantom{aa}\vspace{-0.5cm} \end{figure} Indeed, the results for the $pp{\to}pp\eta$ and $pn{\to}pn\eta$ reactions indicate an almost constant value of $|M|$ up to $\epsilon{\simeq}100$~MeV. This finding is in line with the meson exchange model for $\eta$-meson production due to the $S_{11}(1535)$ intermediate baryonic resonance excitation which provides the dominant $S$-wave production amplitude. A different situation holds for the $NN{\to}NN\pi$ reaction because at large $\epsilon$ the meson exchange model involves the ${\Delta}(1232)$ resonance and a strong contribution to the production amplitude due to the $P$-wave. Therefore, our approach can not be valid for $\pi$-meson production at large $\epsilon$. Recently the $pp{\to}p{\Lambda}K^+$ reaction was measured at COSY~\cite{Balewski,Bilger}. The data indicate a strong deviation from the calculations with the one boson exchange model~\cite{Sibirtsev1,Sibirtsev4} at low $\epsilon$ due to the FSI between the proton and $\Lambda$-hyperon~\cite{Sibirtsev2}. We have evaluated the $pp{\to}p{\Lambda}K^+$ production amplitude with the singlet $^1S_0$ and triplet $^3S_1$ effective range parameters for ${\Lambda}p$ scattering from Ref.\cite{Stoks} (model a) and show the result in Fig.\ref{memo4}. Again the data can be reasonably reproduced with $|M| \approx 43$~fm over the available range of the excess energy. We mention that the parameters for the $YN$ interactions cannot be fitted uniquely to the available $YN$ scattering data since experimental results are very scarce and have large statistical and systematical uncertainties. In turn the $pp{\to}NYK$ reaction might serve as an additional source for the examination of the hyperon-nucleon interaction at low relative momenta. \begin{figure}[h] \phantom{aa}\vspace{-0.7cm} \psfig{file=memo4.ps,width=12.6cm,height=6.5cm} \phantom{aa}\vspace{-0.8cm} \caption[]{Experimental data on the average $pp{\to}p{\Lambda}K^+$\protect\cite{Balewski,Bilger} production amplitudes. The solid lines show a fit with $|M|{=}43$~fm.} \label{memo4} \end{figure} Furthermore, to analyze the $pp{\to}p\Sigma^0K^+$ data~\cite{Sewerin} one needs accurate coupled channel calculations that include the $\Sigma^0p{\leftrightarrow}{\Lambda}p$ transition as well as $\Sigma^0p{\to}\Sigma^0p$ effective range parameters, which are not available by now~\cite{Stoks1}. \begin{table}[h] \begin{center} \phantom{aa}\vspace{-0.9cm} \caption{\label{table1} The $pp{\to}pp\pi^0$, $pp{\to}pp\eta$, $pp{\to}pp\omega$ and $pp{\to}pp\eta^\prime$ production amplitudes $|M|$ evaluated from the data with and without the Coulomb corrections.} \vspace{0.25cm} \begin{tabular}{|l||c|c||c|c|} \hline Reference & \multicolumn{2}{c||} {without Coulomb} & \multicolumn{2}{|c|} {with Coulomb} \\ \cline{2-5} & $|M|$ (fm) & $\chi^2$ & $|M|$ (fm) & $\chi^2$ \\ \hline \multicolumn{5}{|c|} {$pp{\to}pp\pi^0$ } \\ \hline IUCF~~\cite{Meyer1} & 84.2 & 1.8 & 88 & 6.5 \\ IUCF~~\cite{Meyer2} & 79.0 & 0.8 & 81.7 & 4.8 \\ CELSIUS~~\cite{Bondar} & 79.9 & 7.8 & 83.1 & 13.4 \\ \hline \hline \multicolumn{5}{|c|} {$pp{\to}pp\eta$ } \\ \hline SATURNE-II~~\cite{Bergdolt} & 55 & 5.1 & 62 & 3.9 \\ PINOT~~\cite{Chiavassa} & 63 & 1.9 & 60 & 3.3 \\ CELSIUS~~\cite{Calen1} & 61 & 1.1 & 61 & 2.5 \\ \hline \hline \multicolumn{5}{|c|} {$pp{\to}pp\omega$ } \\ \hline SPES-III~~\cite{Hibou1} & 35.3 & 0.5 & 32.7 & 0.5 \\ \hline \hline \multicolumn{5}{|c|} {$pp{\to}pp\eta^\prime$ } \\ \hline SPES-III~~\cite{Hibou} & 17.6 & 0.4 & 18.7 & 1.0 \\ COSY-11~~\cite{Moscal} & 16.1 & 1.3 & 19.3 & 0.6 \\ \hline \end{tabular} \end{center} \phantom{aa}\vspace{-0.65cm} \end{table} The Tables \ref{table1} and \ref{table2} show the averaged production amplitudes evaluated from the data for the reactions discussed above. We separately show the results from different experiments, which are in reasonable agreement with each other. For the $pp{\to}pp\pi^0$, $pp{\to}pp\eta$ and $pp{\to}pp\eta^\prime$ reactions the results are shown with and without Coulomb correction to the $pp$ FSI. \begin{table}[h] \begin{center} \phantom{aa}\vspace{-0.9cm} \caption{\label{table2} The $pp{\to}pn\pi^+$, $pn{\to}pn\eta$ and $pp{\to}p{\Lambda}K^+$ production amplitudes $|M|$.} \vspace{0.2cm} \begin{tabular}{|l||c|c|} \hline Reference & $|M|$ (fm) & $\chi^2$ \\ \hline \multicolumn{3}{|c|} {$pp{\to}pn\pi^+$ } \\ \hline IUCF~~\cite{Hardie} & 240 & 1.8 \\ IUCF~~\cite{Flammang} & 228 & 6.8 \\ \hline \hline \multicolumn{3}{|c|} {$pn{\to}pn\eta$ } \\ \hline CELSIUS~~\cite{Calen2}\ & 157 & 0.5 \\ \hline \hline \multicolumn{3}{|c|} {$pp{\to}p{\Lambda}K^+$ } \\ \hline TOF~~\cite{Bilger} & 38 & 0.05 \\ COSY-11~~~\cite{Balewski} & 46.3 & 0.41 \\ \hline \end{tabular} \end{center} \phantom{aa}\vspace{-0.8cm} \end{table} Finally, due to the FSI the total production cross section is strongly enhanced at low excess energies as illustrated by Fig.\ref{memo18} which shows the $pp{\to}pp\eta^\prime$ cross section as a function of $\epsilon$ calculated in the pion exchange model~\cite{Sibirtsev5}. The dotted line indicates the calculations without the FSI and substantially underestimates the experimental results~\cite{Hibou,Moscal}. Now taking into account the $s$-wave interaction between the final protons we reasonably reproduce the available data. Note that the Coulomb corrections influence the results for $\epsilon{\le}10$~MeV. \begin{figure}[b] \phantom{aa}\vspace{-0.7cm} \centerline{ \psfig{file=memo18.ps,width=12.5cm,height=7cm}} \phantom{aa}\vspace{-0.6cm} \caption[]{The total cross section for the $pp{\to}pp\eta^\prime$ reaction. The experimental data are from Ref.\protect\cite{Hibou,Moscal} while the lines show the calculations with the pion exchange model~\protect\cite{Sibirtsev5} without FSI between the protons (dotted), with FSI (solid) and with Coulomb corrections to FSI (dashed).} \label{memo18} \end{figure} Our calculations illustrate that FSI change the energy dependence of the $pp{\to}pp\eta^\prime$ cross section as compared to the pure phase space $\epsilon^2$. Note that the results without FSI (dotted line in Fig.\ref{memo18}) might, in principle, be renormalized in order to fit the data~\cite{Hibou,Moscal} for $\epsilon \leq$ 10 MeV, however, the increase with $\epsilon$ would be much faster. This indicates that in order to determine the FSI experimentally one needs data on the total production cross section from threshold up to about 100 MeV in excess energy. \section{FSI and differential observables} Obviously the FSI effect differential observables in a more pronounced way than the total production cross section. Fig.\ref{memo6-8} shows that the $s$-wave dominantes the low energy proton-proton scattering and accordingly enhances the low energy part of the $pp$ invariant mass distribution. Thus, due to energy conservation, the high energy part of the final meson-baryon invariant mass distribution is also enhanced. Let us illustrate this for the $pp{\to}pp\eta^\prime$ reaction. Since there are no data on baryonic resonances that couple to the $\eta^\prime$-meson, our calculations~\cite{Sibirtsev5} for the $pp \rightarrow pp \eta^{\prime}$ reaction have been carried out within the pion exchange model without explicitly introducing intermediate baryonic resonances. Thus any deviation of the calculated differential observables for the $pp{\to}pp\eta^\prime$ reaction at low $\epsilon$ from phase space only stems from the $s$-wave FSI between the protons. \begin{figure}[b] \phantom{aa}\vspace{-0.6cm} \centerline{ \begin{minipage}[l]{6cm} \psfig{file=memo12.ps,width=6.1cm,height=7.6cm} \end{minipage}\begin{minipage}[l]{6cm} \psfig{file=memo13.ps,width=6.1cm,height=7.6cm} \end{minipage}} \phantom{aa}\vspace{-0.6cm} \caption[]{The differential observables for the $pp{\to}pp\eta^\prime$ reaction at $\epsilon{=}10$~MeV. a). The Dalitz plot calculated within the pion exchange model and with FSI. b-c). The momentum spectra of the $\eta^\prime$-mesons and protons in the center of mass system calculated with (solid) and without FSI (dashed histograms). The calculations without FSI have been renormalised to the same total cross section.} \label{memo12-13} \end{figure} Fig.\ref{memo12-13}a) shows the Dalitz plot for the $pp{\to}pp\eta^\prime$ reaction at $\epsilon{=}10$~MeV. Indeed the distribution is enhanced at low $pp$ and large $p\eta^\prime$ masses. Figs. \ref{memo12-13}b,c), furthermore, show the c.m.s. momentum spectra of the $\eta^\prime$-mesons and protons produced in the $pp{\to}pp\eta^\prime$ reaction at $\epsilon{=}10$~MeV. The solid histograms display our calculations within the pion exchange model~\cite{Sibirtsev5} including the FSI. The dashed histograms are the results without FSI but corrected by a factor $8.12$ due to the difference in the total cross section calculated with and without FSI (see Fig.\ref{memo18}). The impact of the FSI is obvious and can be easily detected in the $\eta^\prime$-spectra. It is important to note that the distortion of the phase space distribution due to the FSI should be properly taken into account when extrapolating experimental data in a limited acceptance to $4\pi$. Moreover, the FSI produce some resonance structure in the meson-baryon invariant mass distribution as shown in Fig.\ref{memo11}a,b) for the $pp{\to}pp\eta^\prime$ reaction at $\epsilon{=}10$~MeV and $\epsilon{=}100$~MeV. Here the solid histograms are our calculations with FSI while the dotted histograms show the results calculated without FSI which are similar to the pure phase-space distributions. The dashed histograms in Figs. \ref{memo11}a,b) are the calculations without FSI but renormalized to the same total production cross section. Recall that we do not include intermediate baryonic resonances in our model~\cite{Sibirtsev5} and that the $pseudo$ resonance structure in the $p\eta^\prime$ mass spectra stems from the FSI. \begin{figure}[h] \centerline{ \psfig{file=memo11.ps,width=12.5cm,height=10cm}} \phantom{aa}\vspace{-0.6cm} \caption[]{The $p\eta^\prime$ invariant mass spectra (a,b) and the compatibility ratio (c,d) calculated for the $pp{\to}pp\eta^\prime$ reaction at $\epsilon{=}10$~MeV and 100~MeV. The histograms in a),b) show the results with FSI (solid), without FSI (dotted) and without FSI but renormalized to the same cross section(dashed).} \label{memo11} \phantom{aa}\vspace{-0.8cm} \end{figure} Experimentally this effect can be detected when analyzing the compatibility ratio, i.e. the ratio of the measured invariant mass spectra to the phase space distribution that is normalized to the experimental total cross section. The calculated compatibility ratio for the $pp{\to}pp\eta^\prime$ reaction at $\epsilon{=}10$~MeV and 100~MeV is shown in Figs. \ref{memo11}c,d) and visibly deviates from unity. Recall that in the absence of FSI as well as other effects, e.g. an excitation of a baryonic resonance in the meson-baryon system or the appearence of higher partial waves in the production amplitude (which might happen at large $\epsilon$), the compatibility ratio should approach unity. On the other hand, to detect the distortion of the compatibility ratio one needs sufficiently large statistical accuracy as can be seen from Figs. \ref{memo11}c,d). In order demonstrate how an intermediate resonance shows up in the invariant mass spectra we analyze the $pp{\to}pp\eta$ reaction calculating the production amplitude due to the excitation of the $S_{11}(1535)$ resonance. Fig.\ref{memo14} shows the resulting $p\eta$ invariant mass spectra for $\epsilon{=}10$, 100, 150 and 200~MeV. The solid histograms are our calculations with the $S_{11}(1535)$ and FSI, while the dotted histograms indicate the results without FSI between the protons. The dotted lines in Fig.\ref{memo14} show the phase-space distribution normalized to the calculated total cross section. \begin{figure}[h] \centerline{ \psfig{file=memo14.ps,width=12.5cm,height=9cm}} \phantom{aa}\vspace{-0.6cm} \caption[]{The $p\eta$ invariant mass spectra for the $pp{\to}pp\eta$ reaction at $\epsilon{=}10$~MeV, 100, 150 and 200~MeV. The solid histograms are calculations with an excitation of $S_{11}(1535)$ and FSI, while the dashed histograms show the results with $S_{11}(1535)$ but without FSI. The dotted lines indicate the normalized phase-space distributions.} \label{memo14} \phantom{aa}\vspace{-0.7cm} \end{figure} As discussed above, the $S_{11}$ structure cannot be detected at $\epsilon{\le}100$~MeV since the width of the baryonic resonance is larger than the range of the $p\eta$ invariant mass. Furthermore, the shape of the spectra calculated with a $S_{11}$ intermediate resonance and without FSI are similar to the spectra in line with phase space at $\epsilon{\le}100$~MeV. The deviation of the $p\eta$ mass spectra at $\epsilon{=}10$ and 100~MeV from phase space (dotted lines) is entirely due to FSI. The $S_{11}$ structure can be detected at $\epsilon{=}150$ and 200~MeV where the $p\eta$ mass spectra calculated even without FSI (dashed histograms) differ already from pure phase space. Note, however, that FSI substantially distort the spectra and consequently we find two structures in the $p\eta$ invariant mass distributions. The enhancement around $M_{p\eta}$ is due to the $S_{11}$ resonance while the structure close to the kinematical limit of the $p\eta$ mass spectra stems from the FSI. Again the compatibility ratio might serve as a promising tool to detect the reaction mechanism. \begin{figure}[h] \centerline{ \psfig{file=memo15.ps,width=12.5cm,height=8cm}} \phantom{aa}\vspace{-0.6cm} \caption[]{The $\eta$-meson energy spectrum in the center-of-mass system measured for the $pp{\to}pp\eta$ reaction at $\epsilon{=}16$~MeV. The full dots show the experimental results from Ref.\protect\cite{Calen4} while the solid histogram is our calculation with FSI, the dashed histogram without FSI.} \label{memo15} \phantom{aa}\vspace{-0.7cm} \end{figure} Recently CELSIUS reported~\cite{Calen4} the $\eta$-meson c.m.s. energy spectrum measured in the $pp{\to}pp\eta$ reaction at $\epsilon{=}16$~MeV which is shown in Fig.\ref{memo15} together with our calculations. The solid histogram in Fig.\ref{memo15} shows the result with FSI that reasonably reproduces the data; the dashed histogram indicates the result without FSI and substantially differs from the experimental spectrum both in the absolute height and in shape. This comparison, furthermore, demonstrates the validity of our approach which is of sufficient simple form to be used in all data analysis for near threshold reactions. \section{Summary} In this work we have proposed a simple method to analyze or calculate cross sections on near threshold meson production in $pp$ collisions by dividing out kinematical factors and accounting for final-state-interactions (FSI) between the nucleons including approximately also Coulomb corrections. Our analysis of the various models for FSI has shown that the inverse Jost-function method has the largest range of applicability, posesses the correct boundary condition for large excess energies and, furthermore, only involves the effective range parameters $a_s$ and $r_s$ that can be taken from a fit to the respective $s$-wave scattering amplitude. Within this model we have analyzed the available data on $\pi$, $\eta$, $\omega$, $\eta^\prime$ and $K^+\Lambda$ production and found that all data are approximately compatible with constant production matrix elements. This information now in turn can be used to calculate reaction channels with different final states of the baryons if their FSI is known. On the other hand, the constant matrix element hypothesis allows to $measure$ the FSI of baryons that are not available for scattering experiments. Note, however, that precise data up to excess energies of $\approx$ 100 MeV will be necessary. Furthermore, we have shown that a differential data analysis in terms of Dalitz-plots allows to distinguish effects from final state interactions and resonance amplitudes if data are available in a sufficiently wide energy range comparable at least to the width of the resonance amplitude.
\section{Introduction} The influence of the gravitomagnetic field of a rotating compact object on electromagnetic fields has been studied for some 25 years (Wald 1974, Ruffini \& Wilson 1975, Blandford \& Znajek 1977). The coupling of the gravitomagnetic potential with a magnetic field results in an electromotive force. Currents driven by this electromotive force may extract rotational energy from a black hole. This energy could power relativistic jets by Poynting flux (but see also, e.g., Punsly 1996). Cast in the language of the 3+1 split of the Kerr metric, Maxwell's equations, together with the `ingoing wave boundary condition' for electromagnetic fields at the horizon, led to the {\it Membrane Paradigm} (Thorne et al. 1986). Since a black hole does not carry its own magnetic field, not to mention kGauss fields required for the Blandford-Znajek process to be efficient, strong magnetic fields must either be accreted into the black hole from the outer accretion disk, or have to be generated and amplified in the plasma surrounding the black hole. The generation of magnetic fields by a battery operating in the plasma close to a rotating black hole was studied by Khanna (1998b). It was shown that the gravitomagnetic force may play a crucial role in the battery. Khanna \& Camenzind (1996a) studied the possibility of an axisymmetric gravitomagnetic dynamo (or $\omega\Omega$ dynamo), in which the coupling between the gravitomagnetic potential and an electric field is a source for the poloidal magnetic field. This theoretical result invalidates Cowling's anti-dynamo theorem (see also N\`u\~nez 1997), but self-excited growing dynamo modes could not yet be numerically varified for simple kinematics (Khanna \& Camenzind 1996b). Egi et al. (1998) have given a criterion for growing modes of the $\omega\Omega$ dynamo, i.e. that the Poynting flux from close to the horizon (extracted rotational energy of the hole) be positive. Recently, Meier (1998) speculated that, in a zero angular momentum flow into a Kerr black hole, alternatively to the $\omega\Omega$ dynamo, an $\alpha\omega$ dynamo might operate. My simulations of such a sceanrio show that, due to the relativistic accretion velocity, an $\alpha\omega$ dynamo is not very likely. Section \ref{MHD} gives an introduction to the derivation of MHD in the 3+1 split of the Kerr metric. The generalized Ohm's law for an electron-ion plasma is presented in Sec. \ref{gOhm}. The gravitomgnatic battery (along with the relativistic equivalent of Biermann's battery) are discussed in Sec. \ref{bat}. In Sec. \ref{Indeq} the MHD induction equation in the 3+1 split of the Kerr metric is presented. Applications are the gravitomagnetic dynamo (Sec. \ref{gmdyn} and \ref{mfstruc}) and, in Sec. \ref{aOd}, the $\alpha\Omega$ dynamo, or the $\alpha\omega$ dynamo, respectively. Throut the paper I set $G=1=c$. \section{The MHD description of an electron-ion plasma}\label{MHD} The formulation of MHD requires the relativistic definition of a plasma as center-of-mass fluid of its components (Khanna 1998a). The plasma is assumed to be a perfect fluid and is defined by the sum of the ion and electron stress-energy tensors, which contain a collisional coupling term: \begin{equation} (\rho_{\rm m}^{\, '} + p^{\, '})W^{\alpha}W^{\beta} + p^{\, '} g^{\alpha\beta} \equiv T^{\alpha\beta} = \sum_{x=i,e} (\rho_{\rm mx}^{\rm x} + p_{\rm x}^{\rm x}) W_{\rm x}^{\alpha}W_{\rm x}^{\beta} + p_{\rm x}^{\rm x} g^{\alpha\beta} + T^{\alpha\beta}_{\rm x\, coll}\; . \label{defplas} \end{equation} Subscripts $i,e$ refer to ion and electron quantities, respectively. Superscripts denote the rest-frame in which the quantity is defined, where $^{\, '}$ refers to the plasma rest-frame. In the 3+1 split (into hypersurfaces of constant Boyer-Lindquist time $t$, filled with stationary zero angular momentum {\it fiducial observers}) \(T^{\alpha\beta} \) splits into the total density of mass-energy $\epsilon$ and momentum density $\vec{S}$ \begin{equation} \epsilon \equiv (\rho_{\rm m}^{\, '} + p^{\, '} v^2)\gamma^2 \approx \rho_{\rm m}^{\, '}\gamma^2 \qquad \vec{S} \equiv (\rho_{\rm m}^{\, '}+ p^{\, '})\gamma^2\vec{v} \approx \rho_{\rm m}^{\, '} \gamma^2\vec{v} \label{epsSdef} \end{equation} and the stress-energy tensor of 3-space with metric $\buildrel\leftrightarrow\over{h}$ \begin{equation} \buildrel\leftrightarrow\over{T} \equiv (\rho_{\rm m}^{\, '}+ p^{\, '})\gamma^2 \vec{v}\otimes\vec{v} + p^{\, '}\buildrel\leftrightarrow\over{h} \approx \rho_{\rm m}^{\, '}\gamma^2 \vec{v}\otimes\vec{v} + p^{\, '}\buildrel\leftrightarrow\over{h}\; . \end{equation} The approximate expressions hold for a `cold' plasma. Charge density and current density are given by \begin{equation} \rho_{\rm c} \equiv \rho_{\rm ci} + \rho_{\rm ce} = Z e n_{\rm i}\gamma_{\rm i} - e n_{\rm e}\gamma_{\rm e} \qquad \vec{j} \equiv \vec{j}_{\rm i} + \vec{j}_{\rm e} = Z e n_{\rm i}\gamma_{\rm i}\vec{v}_{\rm i} - e n_{\rm e}\gamma_{\rm e}\vec{v}_{\rm e} \; . \end{equation} All quantities resulting from the split are measured locally by FIDOs. \subsection{The generalized Ohm's law in the 3+1 split of the Kerr metric} \label{gOhm} In the `cold' plasma limit, the local laws of momentum conservation for each species can be re-written as equations of motion, which can then be combined to yield the generalized Ohm's law for an electron-ion plasma \begin{eqnarray} \frac{\vec{j}}{\sigma\gamma_{\rm e}}&\approx& \vec{E} +\frac{Z n_{\rm i}\gamma_{\rm i}}{n_{\rm e}\gamma_{\rm e}} \vec{v}\times\vec{B} -\frac{\vec{j}\times\vec{B}}{e n_{\rm e}\gamma_{\rm e}} +\frac{\vec{\nabla}(\alpha_{\rm g} p_{\rm e}^{\rm e})} {e n_{\rm e}\gamma_{\rm e}\alpha_{\rm g} } +\frac{4\pi\gamma_{\rm e}}{\omega^2_{\rm pe}}\rho_{\rm c}^{\; '} \vec{g} +\frac{\rho_{\rm c}{\; '}\gamma\vec{v}}{\sigma\gamma_{\rm e}} \nonumber\\ &-&\frac{4\pi e(Z n_{\rm i}\gamma_{\rm e}^2 - n_{\rm e}\gamma_{\rm i}^2)} {\omega_{\rm pe}^2 \gamma_{\rm e}\gamma_{\rm i}^2} \left(\frac{d(\gamma\vec{v})}{d\tau_{\rm p}} - \buildrel\leftrightarrow\over{H}\cdot(\gamma^2\vec{v})\right) \; , \label{allgOhm} \end{eqnarray} with the conductivity \( \sigma = {e^2 n_{\rm e}}/{m_{\rm e}\nu_{\rm c}}\equiv \omega^2_{\rm pe}/4\pi\nu_{\rm c}\) as measured in the plasma rest frame, the electron plasma frequency $\omega_{\rm pe}$, the factor of gravitational redshift $\alpha_{\rm g}$ (with \(\vec{g}=-\vec{\nabla}\ln\alpha_{\rm g}\)) and the gravitomagnetic tensor field \(\buildrel\leftrightarrow\over{H}\equiv \alpha_{\rm g}^{-1} \vec{\nabla}\vec{\beta}\, .\) \(\vec{\beta} = \beta^{\phi}\vec{e}_{\phi}\equiv -\omega\vec{e}_{\phi} \) is the gravitomagnetic potential, which drags space into differential rotation with angular velocity $\omega$. Note that, in the single fluid description, the gravitomagnetic force drives currents, only if the plasma is charged in its rest frame. $\tau_{\rm p}$ is the proper time in the plasma rest frame. The derivation requires the assumption that the species are coupled sufficiently strong that their bulk accelerations \begin{equation} \frac{d(\gamma_{\rm x}\vec{v}_{\rm x})}{d\tau_{\rm x}}\equiv \left[\frac{\gamma_{\rm x}}{\alpha_{\rm g}}\frac{\partial}{\partial t} +\gamma_{\rm x}\left(\vec{v}_{\rm x} -\frac{\vec{\beta}}{\alpha_{\rm g}}\right)\cdot\vec{\nabla}\right] (\gamma_{\rm x}\vec{v}_{\rm x}) \label{bulkacc} \end{equation} are synchronized. The same is required for the gravitomagnetic accelerations, i.e. \(|\buildrel\leftrightarrow\over{H}\cdot(\gamma_{\rm i}^2\vec{v}_{\rm i}) - \buildrel\leftrightarrow\over{H}\cdot(\gamma_{\rm e}^2\vec{v}_{\rm e})| \ll |\buildrel\leftrightarrow\over{H}\cdot(\gamma_{\rm i}^2\vec{v}_{\rm i})| \). If the MHD-assumption of ``synchronized accelerations'' is not made, Ohm's law contains further current acceleration terms, inertial terms and gravitomagnetic terms (Khanna 1998a), which may be important for collisionless reconnection and particle acceleration along magnetic fields. This topic will be discussed elsewhere. In the limit of quasi-neutral plasma \((Z n_{\rm i}\approx n_{\rm e})\) and \(\gamma_{\rm e}\approx\gamma_{\rm i}\approx\gamma\), Eq.~(\ref{allgOhm}) reduces to \begin{equation} \vec{j} \approx \sigma\gamma(\vec{E} +\vec{v}\times\vec{B}) -\frac{\sigma}{e n_{\rm e}}(\vec{j}\times\vec{B}) +\frac{\sigma}{e n_{\rm e}\alpha_{\rm g}}\vec{\nabla}(\alpha_{\rm g} p_{\rm e}^{\rm e}) \; , \label{allgOhmqn} \end{equation} which contains all the terms, familiar from the non-relativistic generalized Ohm's law, but no gravitomagnetic terms. \subsection{The gravitomagnetic battery}\label{bat} The generation of magnetic fields by a plasma battery was originally devised by Biermann (1950) for stars. He showed that, if the centrifugal force acting on a rotating plasma does not possess a potential, the charge separation owing to the electron partial pressure cannot be balanced by an electrostatic field, and thus currents must flow and a magnetic field is generated. In Khanna (1998b) I have re-formulated Biermann's theory in 3+1 split of the Kerr metric. The base of this battery theory is Ohm's law of eq.~(\ref{allgOhmqn}). Assuming that electrons and ions have non-relativistic bulk velocities in the plasma rest frame, superscripts $i,e,^{\, '} $ can be dropped. With \(p = p_{\rm i}+ p_{\rm e}= (n_{\rm i}+ n_{\rm e})kT\), the {\it impressed electric field} (IEF), \( \vec{E}^{(i)} = {\vec{\nabla}(\alpha_{\rm g} p_{\rm e})} /{e n\gamma \alpha_{\rm g}} \), can be re-expressed with the aid of the equation of motion for a `cold' quasi-neutral plasma to yield \begin{equation} \vec{E}^{(i)} = \frac{m_{\rm i}}{(Z+1)e } \left(\gamma\vec{g} + \buildrel\leftrightarrow\over{H}\cdot(\gamma\vec{v}) -{d(\gamma\vec{v}) \over d\tau}\right) +\frac{Z \left(\vec{j}\times\vec{B} + (\vec{v}\cdot\vec{j})\vec{E}\right) } {(Z+1) e n \gamma}\; . \end{equation} $\tau$ is the proper time in a FIDO frame; i.e. \(d / d\tau_{\rm p} = \gamma d / d\tau\). The criterium for magnetic field generation is that $\vec{\nabla} \times{\alpha_{\rm g}\vec{E}^{(i)}}\ne 0\, .$ Here I restrict the discussion to the gravitomagnetic IEF $\vec{E}^{(i)}_{\rm gm}\; .$ The function part of \(\alpha_{\rm g}\vec{E}^{(i)}_{\rm gm}\) is \begin{eqnarray} \lefteqn{ \left(\vec{\beta}\cdot\vec{\nabla} + \vec{\nabla}\vec{\beta}\, \cdot\right) (\gamma\vec{v}) = \left(\beta^i(\gamma v^j)_{|i} + \gamma\beta^{i|j} v_i\right)\vec{e}_j }\nonumber\\ &&= -\gamma v^{\phi}{\tilde \omega}^2\vec{\nabla}\omega -\omega\left((\gamma v^r)_{,\phi}\vec{e}_r + (\gamma v^{\phi})_{,\phi}\vec{e}_{\phi}\right) \; , \label{Egm} \end{eqnarray} where ${\tilde \omega} = (h_{\phi\phi})^{1/2}$ and ${}_|$ denotes the covariant derivative in 3-space. In axisymmetry \(\alpha_{\rm g}\vec{E}^{(i)}_{\rm gm}\) is clearly rotational, unless some freak $\gamma$ should manage to make \(\gamma v^{\phi}{\tilde \omega}^2\) a function of $\omega$ alone. Thus the gravitomagnetic force drives a poloidal current and generates a toroidal magnetic field. Only if $ v^{\phi}$ is non-axisymmetric, the gravitomagnetic IEF drives to\-ro\-idal currents. The total IEF \((\alpha_{\rm g}\vec{E}^{(i)}_{\rm gm} +\alpha_{\rm g}\vec{E}^{(i)}_{\rm class})\) is likely to rotational in general. This will be quantified for specific velocity fields elsewhere. In presence of a weak poloidal magnetic field the Biermann battery is limited due to modifications of the rotation law by the Lorentz force, rather than by ohmic dissipation. Then the contribution of the centrifugal force to the IEF becomes irrotational already at weak to\-ro\-idal fields (Mestel \& Roxburgh 1962). The gravitomagnetic battery term, on the other hand, is only linearly dependent on $\vec{v}$. The equilibrium field strength should therefore be higher than for the Biermann battery. \section{The MHD induction equation in the 3+1 split of the Kerr metric} \label{Indeq} In this section I review the axisymmetric dynamo equations in the 3+1 split of the Kerr metric (Khanna \& Camenzind 1996a). Ohm's law is assumed to be of the standard form for a quasi-neutral plasma; Hall-term and IEF are neglected. Combining Maxwell's equations (Thorne et al. 1986) with Ohm's law yields the MHD induction equation \begin{equation} \frac{\partial\vec{B}}{\partial t} = \vec{\nabla} \times\left( (\alpha_{\rm g}\vec{v}\times\vec{B} ) - \frac{\eta}{\gamma} \left(\vec{\nabla} \times(\alpha_{\rm g}\vec{B}) + (\vec{E}_{\rm p}\cdot\vec{\nabla}\omega) {\tilde \omega}\vec{e}_{\hat\phi}\right) \right) + (\vec{B}_{\rm p} \cdot\vec{\nabla}\omega){\tilde \omega}\vec{e}_{\hat\phi}\; . \label{MHDindeq} \end{equation} The term standing with the magnetic diffusivity $\eta$ is the current density, which, via Amp\`ere's law, contains the coupling of the gravitomagnetic field with the electric field. In axisymmetry this is simply the shear of the poloidal electric field in the differential rotation of space, $\omega$. Another induction term is the shear of the poloidal magnetic field by $\omega$. This generates to\-ro\-idal magnetic field out of poloidal magnetic field even in a zero-angular-momentum flow. \subsection{The gravitomagnetic dynamo}\label{gmdyn} Introducing the flux $\Psi$ of the poloidal magnetic field and the poloidal current $T$ \begin{equation} \Psi = \frac{1}{2\pi}\int\vec{B}_{\rm p}\cdot d\vec{A} = {\tilde \omega} A^{\hat\phi} \qquad T = 2\int\alpha_{\rm g}\vec{j}_{\rm p}\cdot d\vec{A} = \alpha_{\rm g}{\tilde \omega} B^{\hat\phi} \; , \end{equation} where $A^{\hat\phi}$ is the to\-ro\-idal component of the vector potential, eq.~(\ref{MHDindeq}) splits into \begin{eqnarray} \frac{\partial\Psi}{\partial t} &+& \alpha_{\rm g} (\vec{v}_{\rm p}\cdot\vec{\nabla})\Psi -\frac{\eta{\tilde \omega}}{\gamma} v^{\hat\phi}(\vec{\nabla}\omega\cdot\vec{\nabla}\Psi) -\frac{\eta{\tilde \omega}^2}{\gamma}\vec{\nabla}\cdot \left(\frac{\alpha_{\rm g}}{{\tilde \omega}^2}\vec{\nabla}\Psi\right) \nonumber\\ &=&{} \frac{\eta{\tilde \omega}}{\gamma\alpha_{\rm g}} \left[ \left( T \vec{v}_{\rm p} - \frac{\eta}{\gamma}\vec{\nabla} T\right)\times\vec{e}_{\hat\phi} \right] \cdot\vec{\nabla}\omega \label{dtPsi} \end{eqnarray} \begin{eqnarray} \frac{\partial T}{\partial t} &+& \alpha_{\rm g} (\vec{v}_{\rm p}\cdot\vec{\nabla})T +\alpha_{\rm g}{\tilde \omega}^2 T \left(\vec{\nabla}\cdot\frac{\vec{v}_{\rm p}}{{\tilde \omega}^2} \right) - \alpha_{\rm g} {\tilde \omega}^2 \vec{\nabla}\cdot\left(\frac{\eta}{\gamma{\tilde \omega}^2} \vec{\nabla} T\right) \nonumber\\ &=&\alpha_{\rm g} {\tilde \omega}(\vec{\nabla}\Psi\times \vec{e}_{\hat\phi}) \cdot\vec{\nabla}\Omega \; . \label{dtT} \end{eqnarray} These equations are the relativistic equivalent of the classical axisymmetric dynamo equations. It is important to note, however, that {\it no mean-field approach} was made, but $\Psi$ has source terms anyway. They result from $\vec{E}_{\rm p}\cdot\vec{\nabla}\omega$. Obviously, Cowling's anti-dynamo theorem does not hold close to a rotating black hole. Growing modes of this gravitomagnetic dynamo were shown to exist for steep gradients of the plasma angular velocity $\Omega$ (N\'u\~nez 1997). According to Egi et al. (1998) growing modes require that the Poynting flux carrying rotational energy extracted from the hole be positive. For simple accretion scenarios, growing modes could not be found in kinematic numerical simulations (Khanna \& Camenzind 1996b). If, however, magnetic field is replenished by an outer boundary condition, the gravitomagnetic source terms generate closed loops around the black hole. \subsection{The magnetic field structure in the accretion disk close to the hole}\label{mfstruc} In an accretion disk, magnetic fields may be advected into the near-horizon area, where gravitomagnetic effects may become important. This can be simulated by advection/diffusion boundary conditions for $F=\Psi ,\ T$ \begin{equation} \frac{\partial F}{\partial n} + \frac{\gamma |v^{\hat r}|}{\eta}F = \frac{\partial F_{\rm out}}{\partial n} + \frac{\gamma |v^{\hat r}|}{\eta}F_{\rm out}\; , \end{equation} where $\partial /\partial n$ is the derivative along the outer boundary normal. Figure~\ref{gmdy} shows the stationary final state of a time-dependent simulation (with turbulent magnetic diffusivity), in which $|B_{\rm p,out}| / |B_{\rm t,out}| = 1/50$. For such a dominantly to\-ro\-idal magnetic field the gravitomagnetic source terms are strong enough to change the topology of $\Psi$. This may influence the efficiency of the electromagnetic extraction of rotational energy from the hole. \begin{figure}[] \psfig{width=\textwidth,figure=gmdy.eps,angle=-90} \caption[ ]{Left: Contours of magnetic flux $\Psi$ showing a quadrupolar magnetosphere of an accreting, rapidly rotating black hole ($a=0.998M$, $\alpha_{\rm visc} = 0.25$). Right: The gravitomagnetic current as source of $\Psi$. The disk is marked by long-dashed lines. Solid contours correspond to positive values, short-dashed contours indicate negative values. The range of contours is given below the boxes. Time is measured in diffusive timescales (see below). $r_{\rm g}=GM/c^2$.} \label{gmdy} \end{figure} \subsection{The {\boldmath$\alpha\Omega$} dynamo in the Kerr metric} \label{aOd} \begin{figure}[] \psfig{width=\textwidth,figure=aO1.eps,angle=-90} \caption[ ]{A simulation of an $\alpha\Omega$ dynamo with symmetric initial current in a quasi-Keplerian accretion disk. Kerr parameter $a=0.998M$, $\alpha_{\rm visc}=0.065$. Solid contours have positive values, dashed contours have negative values. Kinks at the outer edge are artefacts of transforming data from the spherical grid into a carthesian plot. Simulation continued in Fig.~\ref{aO2}. } \label{aO1} \end{figure} \begin{figure}[] \psfig{width=\textwidth,figure=aO2.eps,angle=-90} \caption[ ]{The simulation of Fig.~\ref{aO1} has reached a slowly growing, oscillating eigenmode with a period of $\sim 3\, t_{\rm diff}$. Deviations from equatorial (anti-)symmetry are probably due to insufficient resolution in $\theta$-direction. Not shown here: The inclusion of non-linear $\alpha$-quenching leads to severe symmetry breaking and chaotic behavior.} \label{aO2} \end{figure} In the innermost region of an accretion disk around a black hole there may also be a turbulent source term of $\alpha$-type. Without any knowledge of the physical source of the term (convection or magnetic shear instability) or its mathematical form in the relativistic context, one can try to assess the physical regime (magnetic diffusivity, accretion velocity, rotation law) in which growing modes of an $\alpha\Omega$ dynamo exist. For a simple mean-field ansatz (Khanna \& Camenzind 1996a) the equations of the kinematic $\alpha\Omega$ dynamo are identical to Eqs.~(\ref{dtPsi}) and (\ref{dtT}) augmented by the $\alpha$-source term, $\alpha T$, for the flux and $\eta$ replaced by $\eta_{\rm turb}$. In analogy to the expression for $\alpha$ in classical disks, I assume \begin{equation} \alpha = ( \alpha_{\rm g} R_o l_0^2 \Omega /H) \; f(z)/f(H/2) = (3 \alpha_{\rm g} \alpha_{\rm visc} H \Omega) \; f(z)/f(H/2) \; , \label{adyn} \end{equation} where $H$ is the disk scale height, $R_o$ is the Rossby number and $\alpha_{\rm visc}$ is the viscosity parameter of standard accretion disk theory. The factor $\alpha_{\rm g}$ is added in order to suppress the source close to the horizon, where the accretion velocity approaches the speed of light (in properly derived mean-field equations there would probably be a ``Rossby number'' correlated to the accretion velocity instead). The vertical dependence of $\alpha$ is modelled with \(f(z) = \tanh(z) \exp[- (z/H)^2] \). The turbulent diffusivity is described as \begin{equation} \eta_{\rm turb}=\alpha_{\rm visc} H^2 \Omega\, \tilde\omega / r\sin(\theta) \; , \end{equation} with a vertical scaling \((\exp[- (z/H)^2] + 0.1)/1.1\). The boundary condition at $r=10\, r_{\rm g}$ is $\partial\Psi /\partial n = 0$ and $T=0$. Figure~\ref{aO1} shows the first part of a simulation with an initial current $T$, which is symmetric with respect to the equatorial plane, and $\Psi = 0$. The parameters are the Kerr parameter $a=0.998\, M$, $\alpha_{\rm visc}=0.065$, and the angular momentum of the accreting plasma is $99.999 \%$ of the Keplerian value at $r>r_{\rm ms}$ and constant within, which yields an accretion velocity of $\sim 0.003\, c$ at $r=3\, r_{\rm g}$, increasing to $c$ at the horizon. The dynamo is in a slowly growing quadrupolar mode, oscillating with a period of about three diffusive timescales \(t_{\rm diff} = r_{\rm g}^2 / \eta_0 \approx 2\ 10^5\; \sec\, M_9\, \left(\frac{\alpha_{\rm visc}}{0.1}\right)^{-1}\, \left({H(r_{\rm h})\over 0.5 r_{\rm g}}\right)^{-2}\). The same setup, but with lower angular momentum in the accretion disk ($99.9 \%$ of the Keplerian value), which corresponds to a radial velocity of $\sim 0.03 c$ at $r=3\, r_{\rm g}$, is in a decaying mode, which demonstrates that accretion impedes dynamo action (Fig.~\ref{aO3}). \begin{figure}[] \psfig{width=\textwidth,figure=aO3.eps,angle=-90} \caption[ ]{Simulation with same setup as above, except that the accretion velocity at $r\ga 2\, r_{\rm g}$ is 10 times higher. } \label{aO3} \end{figure} \subsection{{\boldmath$\alpha\omega$} dynamo action in a zero-angular momentum flow?}\label{aod} It was mentioned above that the shear of space does also induce a to\-ro\-idal magnetic field (cf. eq.~[\ref{MHDindeq}]). In eq.~(\ref{dtT}) this shear term is obscure, but still there, hidden in \((\vec{\nabla}\Psi\times\vec{e}_{\hat\phi})\cdot\vec{\nabla}\Omega\propto \vec{B}_{\rm p}\cdot\vec{\nabla} \Omega = \vec{B}_{\rm p}\cdot\vec{\nabla} (\alpha_{\rm g} v^{\phi}+ \omega)\). In a zero-angular-momentum flow $v^{\phi}=0$ (or, equivalently $\Omega = \omega$) and thus the current $T$ is solely generated by the shear of space. Moreover, $\vec{\nabla}\omega$ is significantly steeper than $\vec{\nabla}\Omega_{\rm K}$, which means that $\vec{B}_{\rm p}\cdot\vec{\nabla}\omega$ is a strong source term. Meier (1998) speculated that, alternatively to the gravitomagnetic dynamo described above, there could be an $\alpha\omega$ dynamo in a zero-angular momentum accretion flow, with $\alpha$ being due to the magnetic shearing instability. Such a flow, however, accretes at relativistic velocities ($\ga 0.1 \, c$ at $r=10\, r_{\rm g}$), which should suppress any dynamo action. This conclusion is supported by the simulation shown in Fig.~\ref{ao1}. The $\Psi$ and $T$ loops in the corona are transient and depend on the description of $\eta$ and $\alpha$ (here described as in Eq.~[\ref{adyn}], but not suppressed by $\alpha_{\rm g}$ in order to have an upper estimate of the source term). Parameters are $\alpha_{\rm visc}=0.003$ and $H(r_{\rm h})=0.8\, r_{\rm g}$. A wider parameter study is in progress. \begin{figure}[] \psfig{width=\textwidth,figure=ao1.eps,angle=-90} \caption[ ]{Simulation of an $\alpha\omega$ dynamo. The dynamo generates transient structures in the corona close to the horizon. Within the disk there are no signs whatsoever of dynamo action. The field and current are completely determined by the relativistic advection. $T$ is shown in logarithmic contours. } \label{ao1} \end{figure} \acknowledgments This work was partly supported by the Deutsche For\-schungs\-ge\-mein\-schaft (SFB 328).
\section{#1}} \setcounter{equation}{0} \newcommand{\begin{equation}}{\begin{equation}} \newcommand{\end{equation}}{\end{equation}} \newcommand{\begin{eqnarray}}{\begin{eqnarray}} \newcommand{\end{eqnarray}}{\end{eqnarray}} \newcommand{\hspace{0.1cm}}{\hspace{0.1cm}} \newcommand{\hspace{0.7cm}}{\hspace{0.7cm}} \newcommand{\theta}{\theta} \newcommand{\sigma}{\sigma} \newcommand{\rightarrow}{\rightarrow} \newcommand{\beta}{\beta} \newcommand{\bar{z}}{\bar{z}} \newcommand{\tilde{\zeta}}{\tilde{\zeta}} \newcommand{\tilde{\zeta}}{\tilde{\zeta}} \newcommand{\kappa}{\kappa} \newcommand{\bar{k}}{\bar{k}} \newcommand{\bar{\imath}}{\bar{\imath}} \newcommand{\bar{\jmath}}{\bar{\jmath}} \newcommand{\alpha}{\alpha} \newcommand{\Delta}{\Delta} \newcommand{\bar{w}}{\bar{w}} \newcommand{\bar{x}}{\bar{x}} \newcommand{\stackrel{o}{\Gamma}}{\stackrel{o}{\Gamma}} \begin{document} \topmargin 0pt \oddsidemargin 5mm \renewcommand{\thefootnote}{\fnsymbol{footnote}} \newpage \setcounter{page}{0} \begin{titlepage} \begin{flushright} {\tt DIAS-STP-99-03\\ hep-th/9904119 } \end{flushright} \bigskip \bigskip \begin{center} {\Large Renormalization group flow and parallel transport with non-metric compatible connections} \bigskip \bigskip {Brian P. Dolan\footnote{Department of Mathematical Physics, National University of Ireland, Maynooth, Republic of Ireland {\it and}\hskip 0.5cm Dublin Institute for Advanced Studies, 10 Burlington Rd., Dublin 4, Republic of Ireland; \hfill\break e-mail: <EMAIL>} and Alex Lewis\footnote{Department of Mathematical Physics, National University of Ireland, Maynooth, Republic of Ireland; e-mail: <EMAIL>, supported by Enterprise Ireland grant no. SC/98/739}.} \end{center} \begin{center} \footnotesize \end{center} \normalsize \bigskip \bigskip \begin{center} {\bf Abstract} \end{center} A family of connections on the space of couplings for a renormalizable field theory is defined. The connections are obtained from a Levi-Civita connection, for a metric which is a generalisation of the Zamolodchikov metric in two dimensions, by adding a family of tensors which are solutions of the renormalization group equation for the operator product expansion co-efficients. The connections are torsion free, but not metric compatible in general. The renormalization group flows of $N=2$ supersymmetric Yang-Mills theory in four dimensions and the $O(N)$-model in three dimensions, in the large $N$ limit, are analysed in terms of parallel transport under these connections. \end{titlepage} \newpage In this letter we investigate geometrical properties of the renormalization group flow in some exactly solved theories. The renormalization group flow can be seen as a vector flow in the space of theories, with the couplings of the theory $g^a$ being coordinates on this space . In this approach, it has been shown in \cite{dolan1} following a suggestion in \cite{jo} (see also \cite{ocs,sonoda} that the renormalization group equations for multi-point correlation functions, written in a coordinate covariant form, depend on a symmetric connection $\Gamma^a_{bc}$ through a tensor $\tau^a_{bc}$, \begin{equation} \tau^a_{bc} = \nabla_b \nabla_c \beta^a - R^a_{cbd}\beta^d, \label{tauabc}\end{equation} defined by the RG equation for a regularized 3-point function $G_{abc}(p,q,r)=\langle\Phi_a(p)\Phi_b(q)\Phi_c(r)\rangle$ \begin{equation} \left(\Lambda\frac{\partial}{\partial\Lambda} +{\cal L}_\beta\right) G_{abc}(p,q,r)=\tau_{ab}^dG_{dc}(p+q,r) + \tau_{bc}^d G_{da}(q+r,p) + \tau^d_{ac}G_{db}(r+p,q) + \cdots \label{rgeqn}\end{equation} where $G_{ab}(p,q)=\langle\Phi_a(p)\Phi_b(q)\rangle$ and the dots denote contact terms that are only important for large momenta. However, there is no general rule for finding a connection. Moreover, since the RG equations only depend on the connection through the tensor $\tau^a_{bc}$, there is in fact a family of connections which give the same equations. The approach we take is therefore to determine the full family of possible connections for some exactly solvable models, and investigate the geometrical properties of the RG flow for the most general connection. For two connections $\Gamma$ and $\tilde\Gamma$ with covariant derivatives $\nabla$ and $\tilde\nabla$ and curvatures $R$ and $\tilde R$ respectively to both be compatible with eq. (\ref{tauabc}) we must have \begin{equation} \nabla_b \nabla_c \beta^a - R^a_{cbd}\beta^d = \tilde\nabla_b \tilde\nabla_c \beta^a - \tilde{R}^a_{cbd}\beta^d. \end{equation} This equation is satisfied if the Lie derivative ${\cal L}_\beta$ of the difference between the connections vanishes. This enables us to determine the full family of possible connections if one connection $\stackrel{o}{\Gamma}^a_{bc}$ is already known: we can write any connection which is compatible with eq. (\ref{tauabc}) as $\Gamma^a_{bc}= \stackrel{o}{\Gamma}^a_{bc} + {\cal G}^a_{bc}$, where \begin{equation} {\cal L}_\beta {\cal G}^a_{bc} = {\cal G}^a_{dc}\partial_b\beta^d + {\cal G}^a_{bd}\partial_c\beta^d - {\cal G}^d_{bc}\partial_d\beta^a + \beta^d\partial_d{\cal G}^a_{bc} =0. \label{Liedv}\end{equation} We still have to find a connection $\stackrel{o}{\Gamma}^a_{bc}$ to construct the other possible connections. One solution is to use the Levi-Civita connection of a metric on the space of couplings. An example of such a metric is the Zamolodchikov metric in $D=2$, which was used in the proof of the $c$-theorem \cite{zamolodchikov}. More recently, building on ideas laid out in \cite{ocs}, the geometrical properties of metrics in $D>2$ have also been investigated for some models, including: free field theory \cite{dolan2}, the O(N) model \cite{dolan3} and Seiberg-Witten theory for $SU(2)$ \cite{dolan4}. In all these models, it was found that some (but not all) of the RG flow lines are geodesics of the metric. In particular, the lines of crossover between fixed points are geodesics, and this may be related to irreversibility of the renormalization group flow. Since we now have a family of connections which are equivalent, at least as far as the renormalization group equations are concerned, it is natural to ask whether the geodesic flow of the lines of crossover generalises to auto-parallel flow for other connections (a line which is auto-parallel for the Levi-Civita connection is a geodesic)\footnote{We distinguish between auto-parallels, which are curves whose tangent vectors remain tangent vectors under parallel transport along the curve, and geodesics, which are curves of shortest length. In general, these coincide for the Levi-Civita connection only.}. The auto-parallel equation for a vector field is $\nabla_{\vec\beta} \vec\beta = \eta\vec\beta$, where $\eta$ is a function which depends on the parameterization along the curve. With a connection $\Gamma=\stackrel{o}{\Gamma} + {\cal G}$, this becomes \begin{equation} \beta^b\frac{\partial\beta^a}{\partial x^b} + \stackrel{o}{\Gamma}^a_{bc}\beta^b\beta^c + {\cal G}^a_{bc}\beta^b\beta^c = \eta\beta^a. \label{fulleq}\end{equation} Our main aim is to see which of the possible connections, if any, will satisfy this equation for a given renormalization group trajectory. In particular, some of the trajectories in the models we will examine in this letter are geodesics of the metric, so eq. (\ref{fulleq}) is satisfied for ${\cal G}=0$. In that case, eq. (\ref{fulleq}) simplifies to the condition that \begin{equation} {\cal G}^a_{bc}\beta^b\beta^c = \eta'\beta^a, \label{auto}\end{equation} where $\eta'$ is another function. A natural candidate for a metric on the space of couplings is the two-point correlation functions of the model \cite{ocs}. If the action $S$ is linear in the couplings, \begin{equation} S = S_0 + \int d^Dx g^a\Phi_a(x) \end{equation} then a metric can be defined by \begin{equation} G_{ab} = \int d^Dx \langle \tilde\Phi(x)\tilde\Phi(0) \rangle \label{metric}\end{equation} where $\tilde\Phi(x)=\Phi(x)-\langle\Phi(x)\rangle$. Although the individual components of this metric may diverge, the geometry can still be non-singular. As our first example, we consider the $O(N)$ model for large $N$ in 3 dimensions. This is a model of a scalar field $\vec\varphi$ in the vector representation of $O(N)$ with the action \begin{equation} S=\int d^3x \left\{ \frac{1}{2} (\nabla\vec\varphi)^2 +\vec{j}\cdot\vec\varphi +\frac{r}{2}\vec\varphi^2 +\frac{u}{4!}(\vec\varphi^2)^2 \right\} \end{equation} following \cite{dolan3} we analyse the geometry in terms of three bare parameters, $\phi ,X,\lambda$, defined by \begin{equation} \phi = 4\sqrt\frac{\pi }{N\Lambda}\langle\varphi\rangle, ~~~~~X = \frac{1}{2\Lambda}\int d^3x \langle\varphi^2\rangle, ~~~~~\lambda = \frac{Nu}{48\pi\Lambda}. \end{equation} Although these are bare parameters, they are finite as we have a finite cut-off $\Lambda$, so we can use them as our coordinates on the space of couplings (since we are in any case only interested in properties of the RG flow which are independent of the coordinate system). The beta functions, which represent a vector flow on this space are \cite{dolan3} \begin{equation} \beta^\phi=-\frac{1}{2}\phi~~~~~\beta^X=-X~~~~~\beta^\lambda =-\lambda. \label{betas}\end{equation} In \cite{dolan3} the metric (\ref{metric}) was computed, and it was found that only one of the renormalization group trajectories described by these beta functions is actually a geodesic of the metric - the line $X=\phi=0$, which is the line of crossover from the Wilson-Fisher fixed point at $\lambda=\infty$ to the Gaussian fixed point at $\lambda=0$. We now want to see if any of the renormalization group trajectories are auto-parallel for a connection $\Gamma=\stackrel{o}{\Gamma} +{\cal G}$, where $\stackrel{o}{\Gamma}$ is the Levi-Civita connection from \cite{dolan3} and ${\cal G}$ is a solution of eq. (\ref{Liedv}). Using bare rather than renormalized parameters as coordinates makes it easy to solve eq. (\ref{Liedv}) -- the general solution contains a number of arbitrary functions which depend only on the ratios $X/\lambda$ and $\phi^2/\lambda$, but these ratios are constant along any of the RG trajectories described by eq. (\ref{betas}), so we can treat them as just being arbitrary constants, which we write as $f^i_{jk}$. The solution can then be written, provided $X$, $\phi$ and $\lambda$ are non-zero, as \begin{equation} {\cal G}^i_{jk} = \frac{\beta^i}{\beta^j\beta^k}f^{i}_{jk} \label{ONG}\end{equation} (no summation on $i$, $j$, $k$). With this solution, it is clear that for eq. (\ref{fulleq}) to be satisfied for a trajectory, the following differences must be constant along the curve: \begin{equation} \frac{2}{\phi}\stackrel{o}{\Gamma}^\phi_{ij}\beta^i\beta^j- \frac{1}{\lambda}\stackrel{o}{\Gamma}^\lambda_{ij}\beta^i\beta^j=\mbox{constant}, ~~~~~~~~\frac{2}{\phi}\stackrel{o}{\Gamma}^\phi_{ij}\beta^i\beta^j- \frac{1}{X}\stackrel{o}{\Gamma}^X_{ij}\beta^i\beta^j=\mbox{constant}. \label{cond}\end{equation} It can be seen from the expressions for the components of the Levi-Civita connection given in Appendix 2 of \cite{dolan2} that this is not true for any of the RG flow lines except the line of crossover. Thus none of the flow lines with $\phi$, $X$, and $\lambda$ all non-zero can be auto-parallel for any connection. When one of $X$, $\phi$ or $\lambda$ is $0$, the solution (\ref{ONG}) has to be changed by absorbing factors of $X/\lambda$ or $\phi^2/\lambda$ into $f^i_{jk}$ to make it finite, for example if $\lambda\neq 0$ we can write the solution as \begin{equation} {\cal G}^i_{jk} = \lambda^nf^i_{jk}~~~~~~~~n=n_i-n_j-n_k \label{secondG}\end{equation} with $n_\phi=\frac12$ and $n_X=n_\lambda=1$. However, if $X=0$, $\phi=0$ or $\lambda=0$, eq. (\ref{fulleq}) can only be satisfied if $\stackrel{o}{\Gamma}^X_{ij}\beta^i\beta^j=0$, $\stackrel{o}{\Gamma}^\phi_{ij}\beta^i\beta^j=0$ or $\stackrel{o}{\Gamma}^\lambda_{ij}\beta^i\beta^j=0$ respectively, and the only line which satisfies these conditions is $X=\phi=0$. Thus none of the other RG flow lines can ever be auto-parallel. Finally, we want to see if the geodesic $X=\phi=0$ is auto-parallel for other connections (apart from the Levi-Civita connection). This means we have to see if eq. (\ref{auto}) is satisfied by the solution (\ref{secondG}). For example, if only $f^\phi_{\phi\phi}$, $f^\lambda_{\lambda\lambda}$ and $f^X_{XX}$ are non-zero, eq. (\ref{auto}) becomes \begin{eqnarray} -\eta' \phi &=& \lambda^{-1/2}f^\phi_{\phi\phi}\phi^2/2 \nonumber \\ -\eta' \lambda &=& \lambda^{-1}f^\lambda_{\lambda\lambda}\lambda^2 \\ -\eta' X &=& \lambda^{-1}f^X_{XX}X^2 \nonumber \end{eqnarray} which clearly is satisfied for $X=\phi=0$. However this is not true for the most general connection. For example, if ${\cal G}^X_{\lambda\lambda}=\lambda^{-1} f^X_{\lambda\lambda}\ne 0$ or ${\cal G}^\phi_{\lambda\lambda}=\lambda^{-3/2} f^\phi_{\lambda\lambda}\ne 0$ then the line of crossover is not an auto-parallel (since the corresponding Levi-Civita connection components vanish in the large $N$ limit, \cite{dolan3}). The line of crossover is therefore not auto-parallel for the entire family of connections which can be used in the RG equations, but only for the class with ${\cal G}^X_{\lambda\lambda}={\cal G}^\phi_{\lambda\lambda}=0$. Thus we find that introducing the family of non-metric compatible connections changes the conclusions, compared to the the Levi-Civita case, as to which renormalization group flow lines are auto-parallel. There is a large class of connections for which the line of crossover between the Gaussian and the Wilson-Fisher fixed points remains auto-parallel, but this is not true of the most general connection. Our second example of an RG flow shows that, in general, the family of connections do change the auto-parallel nature of the geodesic flows (in the special cases where RG flow is geodesic). This is 4-dimensional $N=2$ Yang-Mills theory \cite{sw}. The geometrical properties of this model were investigated in \cite{dolan4} where two possible metrics were considered, and it was found most of the renormalization group flow lines are not geodesics, but some special lines are geodesics of both metrics. We now want to see which, if any, of the RG flow lines are auto-parallel for other connections. The complex coupling $\tau = \frac{\theta_{eff}}{2\pi} +\frac{4\pi i}{g^2_{eff}}$ is given as a function of $\tilde{u}=u/\Lambda^2$, where $u=Tr\langle\varphi^2\rangle$ parameterizes the symmetry breaking, by \begin{equation} \tau = \frac{iK'}{K}+2n \label{tau}\end{equation} where $K(k^2)$ is a standard elliptic integral \cite{ww} with $k^2 = \frac{2}{\tilde{u}+1}$ and $K'=K(1-k^2)$ and $n$ is an integer. Various beta functions for this model have been investigated in \cite{bm,ritz,ckmm}. In \cite{ckmm} Wilsonian and Novikov, Shifman, Vainshtein, Zakharov, \cite{NSVZ}, beta functions were considered. Here we concentrate on the Callan-Symanzik beta functions of \cite{bm,ritz}. The Callan-Symanzik beta functions, $\vec\beta = \beta\frac{\partial}{\partial\tau} +\bar\beta\frac{\partial}{\partial\bar\tau}$ are defined by \begin{equation} \beta(\tau) = \lambda \left.\frac{\partial\tau}{\partial\Lambda}\right|_u = -2\tilde{u}\frac{d\tau}{d\tilde{u}}\; . \label{betatau}\end{equation} This beta function represents a vector flow on a manifold, parameterized by $\tau$ (or $u$), which has the topology of a sphere with three punctures. This manifold has three singular points: $u=\infty$ (the weak coupling limit) and $u=\pm \Lambda^2$ (where there are extra massless degrees of freedom). The Seiberg-Witten metric on this manifold in the $u$-coordinates is \begin{equation} ds^2=\pi^2 \mbox{Im}(\tau) \left| \frac{\vartheta_3^4\vartheta_4^4}{\vartheta_2^2}\right|^2 d\tau d\bar\tau = \frac{1}{\pi^2} \frac{(K'\bar{K} + \bar{K}'K)}{\sqrt{1+u}\sqrt{1+\bar{u}}}dud\bar{u}, \label{swmetric}\end{equation} where $\vartheta_i$, $i=2,3,4$ are Jacobi $\vartheta$-functions, \cite{ww}. Another metric which can be introduced for this geometry is the Poincare metric \begin{equation} ds^2=\frac{1}{(\mbox{Im}(\tau))^2}d\tau d\bar\tau. \end{equation} For both these metrics, the lines of real $u$ and imaginary $u$ are geodesics, but the other RG flow lines are not \cite{dolan4}. In fact we do not need to know the explicit form of $\beta(\tau)$ to see if the geodesics will be auto-parallel for any connection: if we use $u$ and $\bar u$ as coordinates, so that $\vec\beta = \beta (u)\frac{\partial}{\partial u} +\bar\beta (u) \frac{\partial}{\partial \bar{u}}$, we only need to know that $\frac{\partial\beta}{\partial \bar u}=\frac{\partial\bar\beta}{\partial u}= 0$. The solution of eq. (\ref{Liedv}) is then \begin{eqnarray} {\cal G}^u_{uu} = \frac{1}{\beta}g^u_{uu}(u/\bar{u}) && {\cal G}^{\bar u}_{\bar u\bar u} = \frac{1}{\bar\beta}g^{\bar u}_{\bar{u}\bar{u}}(u/\bar{u}) \nonumber\\ {\cal G}^u_{\bar{u}u} = \frac{1}{\bar\beta}g^u_{\bar{u}u}(u/\bar{u}) && {\cal G}^{\bar u}_{u\bar{u}} = \frac{1}{\beta}g^{\bar u}_{u\bar{u}}(u/\bar{u}) \nonumber \\ {\cal G}^u_{\bar u\bar u} = \frac{\beta}{\bar\beta^2}g^u_{\bar{u}\bar{u}}(u/\bar{u}) && {\cal G}^{\bar u}{uu} = \frac{\bar\beta}{\beta^2}g^{\bar u}_{uu}(u/\bar{u}). \end{eqnarray} The solution contains arbitrary functions of $u/\bar{u}$, but these functions are constant along the radial lines in the $u$-plane, and we can see from eq. (\ref{betatau}) that these are just the RG flow lines (in the $u$-coordinates, $\beta^u=-2u$, $\beta^{\bar u}=-2\bar u$). If we assume that ${\cal G} = {\cal G}^i_{jk}dx^j\otimes dx^k\otimes\frac{\partial}{\partial x^i}$ is a real tensor and substitute this solution into eq. (\ref{fulleq}) or (\ref{auto}), we can see that those equations depend only on one function $g(u/\bar u)$: \begin{equation} g(u/\bar{u}) = g^u_{uu}+g^u_{\bar u u}+g^u_{u\bar u}+g^u_{\bar u\bar u} ~~~~~ \bar{g}(u/\bar{u}) = g^{\bar u}_{\bar u\bar u}+g^{\bar u}_{u\bar u}+g^{\bar u}_{\bar u u}+g^{\bar u}_{uu}. \end{equation} Eq. (\ref{fulleq}) then reduces to the condition that \begin{equation} \stackrel{o}{\Gamma}^u_{uu} u +g(u/\bar u) = \stackrel{o}{\Gamma}^{\bar u}_{\bar u\bar u} \bar u +\bar g(u/\bar u) \end{equation} where for the Seiberg-Witten metric $\stackrel{o}{\Gamma}^{\bar u}_{\bar u\bar u} = \overline{{\stackrel{o}{\Gamma}}^u_{uu}}$ and \begin{equation} \stackrel{o}{\Gamma}^u_{uu} = \frac{\bar{K}'\partial_u K +\bar{K}\partial_uK'}{K'\bar{K} +\bar{K}'K} -\frac{1}{2(u+1)}. \end{equation} Since $g$ and $\bar g$ are constant along an RG trajectory, a trajectory can only be auto-parallel for some choice of $g$ and $\bar g$ if the imaginary part of $\stackrel{o}{\Gamma}^u_{uu}u$ is constant along that trajectory. It can be shown numerically that this is only the case for the geodesics, the lines of real and imaginary $u$, where $\stackrel{o}{\Gamma}^u_{uu}u$ is real. All the other RG flow lines are therefore not auto-parallel for any connection, while the real $u$ line is auto-parallel provided $g(1)=\bar{g}(1)$ and the imaginary $u$ line is auto-parallel provided $g(-1)=\bar{g}(-1)$. Thus the geodesic flow lines are not auto-parallel for the most general connection (as was the case in the $O(N)$ model), but they are auto-parallel, for example, if $\bar{g}(u/\bar u)= g(\bar u/u)$. Since we did not use the explicit expressions for $\tau$ and $\beta$ above, this result can be applied to any theory in which the beta function is an analytic function of one complex variable only. In the case of $N=2$ $SU(2)$ Yang-Mills theory with massless quarks, as in the pure Yang-Mills theory, the complex coupling $\tau$ depends only on $u/\Lambda^2$. However, the singularities of the theories with massless quarks are not all on the real or imaginary $u$ axes - for example, when the number of flavours $N_f=1$, there are singularities at $u=-u_0$ and $u=u_0e^{\pm i\pi/3}$. Of course, we cannot tell if these lines are actually auto-parallel unless we know whether they are geodesics of the Seiberg-Witten metric, but we can say that if they are geodesics of the Seiberg-Witten metric in these models, they be auto-parallel for connections with $g(e^{2\pi i/3})=\bar {g}(e^{2\pi i/3})$, but not for all connections for which $\bar{g}(u/\bar u)= g(\bar u/u)$, as in the case of the pure Yang-Mills theory. In conclusion it has been shown that the renormalization group equation for the operator expansion co-efficients gives rise to a family of non-metric compatible connections which are related to the Levi-Civita connection by $\Gamma^a_{bc}= \stackrel{o}{\Gamma}^a_{bc} + {\cal G}^a_{bc}$, where ${\cal L}_\beta {\cal G}^a_{bc} =0$. In general RG flows which are geodesic for the Levi-Civita connection are not auto-parallel for all members of the family, though in the models examined they are for a large sub-class of connections (it can be shown that the same is also true for the free field models considered in \cite{dolan2}). In none of the examples examined here is it the case that a RG flow line which is not geodesic under the Levi-Civita connection is auto-parallel for some member of the family.
\partial{\partial} \def\rangle{\rangle} \def\subset{\subset} \def\simeq{\simeq} \def\times{\times} \def\wedge{\wedge} \def\widehat{\widehat} \def\widetilde{\widetilde} \def\alpha{\alpha} \def\beta{\beta} \def\delta{\delta} \def\gamma{\gamma} \def\kappa{\kappa} \def\lambda{\lambda} \def\omega{\omega} \def\sigma{\sigma} \def\theta{\theta} \def\varepsilon{\varepsilon} \def\varphi{\varphi} \def\varpi{\varpi} \def\zeta{\zeta} \def\Delta{\Delta} \def\Gamma{\Gamma} \def\Lambda{\Lambda} \def\Omega{\Omega} \def\Sigma{\Sigma} \def\Theta{\Theta} \def\mathop{\rm Der}\nolimits{\mathop{\rm Der}\nolimits} \def\mathop{\rm Diff}\nolimits{\mathop{\rm Diff}\nolimits} \def\mathop{\rm Dom}\nolimits{\mathop{\rm Dom}\nolimits} \def\mathop{\rm End}\nolimits{\mathop{\rm End}\nolimits} \def\mathop{\rm Id}\nolimits{\mathop{\rm Id}\nolimits} \def\mathop{\rm Ind}\nolimits{\mathop{\rm Ind}\nolimits} \def\mathop{\rm Index}\nolimits{\mathop{\rm Index}\nolimits} \def\mathop{\rm Iso}\nolimits{\mathop{\rm Iso}\nolimits} \def\mathop{\rm Ker}\nolimits{\mathop{\rm Ker}\nolimits} \def\mathop{\rm loc}\nolimits{\mathop{\rm loc}\nolimits} \def\mathop{\rm mod}\nolimits{\mathop{\rm mod}\nolimits} \def\mathop{\rm res}\nolimits{\mathop{\rm res}\nolimits} \def\mathop{\rm Res}\nolimits{\mathop{\rm Res}\nolimits} \def\mathop{\rm Sign}\nolimits{\mathop{\rm Sign}\nolimits} \def\mathop{\rm Spin}\nolimits{\mathop{\rm Spin}\nolimits} \def\mathop{\rm vol}\nolimits{\mathop{\rm vol}\nolimits} \font\tenbb=msbm10 \font\sevenbb=msbm7 \font\fivebb=msbm5 \newfam\bbfam \textfont\bbfam=\tenbb \scriptfont\bbfam=\sevenbb \scriptscriptfont\bbfam=\fivebb \def\fam\bbfam{\fam\bbfam} \def{\bb C}{{\fam\bbfam C}} \def{\bb F}{{\fam\bbfam F}} \def{\bb Q}{{\fam\bbfam Q}} \def{\bb R}{{\fam\bbfam R}} \def{\bb T}{{\fam\bbfam T}} \def{\bb Z}{{\fam\bbfam Z}} \def{\cal A}{{\cal A}} \def{\cal E}{{\cal E}} \def{\cal F}{{\cal F}} \def{\cal G}{{\cal G}} \def{\cal H}{{\cal H}} \def{\cal L}{{\cal L}} \def{\cal M}{{\cal M}} \def{\cal O}{{\cal O}} \def{\cal P}{{\cal P}} \def{\cal U}{{\cal U}} \def{\cal X}{{\cal X}} \def\build#1_#2^#3{\mathrel{ \mathop{\kern 0pt#1}\limits_{#2}^{#3}}} \def\diagram#1{\def\normalbaselines{\baselineskip=0pt \lineskip=10pt\lineskiplimit=1pt} \matrix{#1}} \def\hfl#1#2{\smash{\mathop{\hbox to 6mm{\rightarrowfill}} \limits^{\scriptstyle#1}_{\scriptstyle#2}}} \def\hfll#1#2{\smash{\mathop{\hbox to 6mm{\leftarrowfill}} \limits^{\scriptstyle#1}_{\scriptstyle#2}}} \def\vfl#1#2{\llap{$\scriptstyle #1$}\left\downarrow \vbox to 3mm{}\right.\rlap{$\scriptstyle #2$}} \def\limind{\mathop{\oalign{lim\cr \hidewidth$\longrightarrow$\hidewidth\cr}}} \def\limproj{\mathop{\oalign{lim\cr \hidewidth$\longleftarrow$\hidewidth\cr}}} \def\boxit#1#2{\setbox1=\hbox{\kern#1{#2}\kern#1}% \dimen1=\ht1 \advance\dimen1 by #1 \dimen2=\dp1 \advance\dimen2 by #1 \setbox1=\hbox{\vrule height\dimen1 depth\dimen2\box1\vrule}% \setbox1=\vbox{\hrule\box1\hrule}% \advance\dimen1 by .4pt \ht1=\dimen1 \advance\dimen2 by .4pt \dp1=\dimen2 \box1\relax} \def\mathop{>\!\!\!\triangleleft}{\mathop{>\!\!\!\triangleleft}} \catcode`\@=11 \def\displaylinesno #1{\displ@y\halign{ \hbox to\displaywidth{$\@lign\hfil\displaystyle##\hfil$}& \llap{$##$}\crcr#1\crcr}} \def\ldisplaylinesno #1{\displ@y\halign{ \hbox to\displaywidth{$\@lign\hfil\displaystyle##\hfil$}& \kern-\displaywidth\rlap{$##$} \tabskip\displaywidth\crcr#1\crcr}} \catcode`\@=12 \baselineskip=14pt \hsize=118mm \hoffset=20mm \vsize=215mm \voffset=15mm \def{\cal A}{\cal A} \def{\cal B}{\cal B} \def{\cal H}{\cal H} \def{\cal L}{\cal L} \def{\cal U}{\cal U} \font\tenbb=msbm10 \font\sevenbb=msbm7 \font\fivebb=msbm5 \newfam\bbfam \textfont\bbfam=\tenbb \scriptfont\bbfam=\sevenbb \scriptscriptfont\bbfam=\fivebb \def\fam\bbfam{\fam\bbfam} \def{\bb C}{{\fam\bbfam C}} \def{\bb N}{{\fam\bbfam N}} \def{\bb R}{{\fam\bbfam R}} \def{\bb Z}{{\fam\bbfam Z}} \def\alpha{\alpha} \def\beta{\beta} \def\delta{\delta} \def\Delta{\Delta} \def\gamma{\gamma} \def\Gamma{\Gamma} \def\lambda{\lambda} \def\Lambda{\Lambda} \def\omega{\omega} \def\Omega{\Omega} \def\sigma{\sigma} \def\theta{\theta} \def\varepsilon{\varepsilon} \def\varphi{\varphi} \def\backslash{\backslash} \def\infty{\infty} \def\nabla{\nabla} \def\partial{\partial} \def\simeq{\simeq} \def\times{\times} \def\wedge{\wedge} \def\langle{\langle} \def\rangle{\rangle} \def\forall{\forall} \def\oplus{\oplus} \def\otimes{\otimes} \def\subset{\subset} \def\widetilde{\widetilde} \def\rightarrow{\rightarrow} \def\hookrightarrow{\hookrightarrow} \def\build#1_#2^#3{\mathrel{ \mathop{\kern 0pt#1}\limits_{#2}^{#3}}} \def\mathop{>\!\!\!\triangleleft}{\mathop{>\!\!\!\triangleleft}} \def\vrule height 0.5em depth 0.2em width 0.5em{\vrule height 0.3em depth 0.2em width 0.3em} \font\tenbb=msbm10 \font\sevenbb=msbm7 \font\fivebb=msbm5 \newfam\bbfam \textfont\bbfam=\tenbb \scriptfont\bbfam=\sevenbb \scriptscriptfont\bbfam=\fivebb \def\fam\bbfam{\fam\bbfam} \def{\bb C}{{\fam\bbfam C}} \def{\bb H}{{\fam\bbfam H}} \def{\bb N}{{\fam\bbfam N}} \def{\bb R}{{\fam\bbfam R}} \def{\bb T}{{\fam\bbfam T}} \def{\bb Z}{{\fam\bbfam Z}} \def{\cal A}{{\cal A}} \def{\cal B}{{\cal B}} \def{\cal E}{{\cal E}} \def{\cal H}{{\cal H}} \def{\cal I}{{\cal I}} \def{\cal K}{{\cal K}} \def{\cal L}{{\cal L}} \def{\cal R}{{\cal R}} \def{\cal S}{{\cal S}} \def{\cal U}{{\cal U}} \def\alpha{\alpha} \def\beta{\beta} \def\delta{\delta} \def\lambda{\lambda} \def\gamma{\gamma} \def\omega{\omega} \def\sigma{\sigma} \def\theta{\theta} \def\varepsilon{\varepsilon} \def\varphi{\varphi} \def\varpi{\varpi} \def\zeta{\zeta} \def\Delta{\Delta} \def\Gamma{\Gamma} \def\Lambda{\Lambda} \def\Omega{\Omega} \def\Sigma{\Sigma} \def\exists{\exists} \def\forall{\forall} \def\infty{\infty} \def\langle{\langle} \def\nabla{\nabla} \def\oplus{\oplus} \def\otimes{\otimes} \def\partial{\partial} \def\rangle{\rangle} \def\subset{\subset} \def\simeq{\simeq} \def\times{\times} \def\wedge{\wedge} \def\rightarrow{\rightarrow} \def\longrightarrow{\longrightarrow} \def\Rightarrow{\Rightarrow} \def\leftrightarrow{\leftrightarrow} \def\hbox{\hbox} \def\mathop{\rm Aut}\nolimits{\mathop{\rm Aut}\nolimits} \def\mathop{\rm Diff}\nolimits{\mathop{\rm Diff}\nolimits} \def\mathop{\rm Dom}\nolimits{\mathop{\rm Dom}\nolimits} \def\mathop{\rm End}\nolimits{\mathop{\rm End}\nolimits} \def\mathop{\rm id}\nolimits{\mathop{\rm id}\nolimits} \def\mathop{\rm Index}\nolimits{\mathop{\rm Index}\nolimits} \def\mathop{\rm Inf}\nolimits{\mathop{\rm Inf}\nolimits} \def\mathop{\rm Int}\nolimits{\mathop{\rm Int}\nolimits} \def\mathop{\rm Ker}\nolimits{\mathop{\rm Ker}\nolimits} \def\mathop{\rm Rang}\nolimits{\mathop{\rm Rang}\nolimits} \def\mathop{\rm Re}\nolimits{\mathop{\rm Re}\nolimits} \def\mathop{\rm Res}\nolimits{\mathop{\rm Res}\nolimits} \def\mathop{\rm Sign}\nolimits{\mathop{\rm Sign}\nolimits} \def\mathop{\rm Signe}\nolimits{\mathop{\rm Signe}\nolimits} \def\mathop{\rm Spectrum}\nolimits{\mathop{\rm Spectrum}\nolimits} \def\mathop{\rm spin}\nolimits{\mathop{\rm spin}\nolimits} \def\mathop{\rm Spin}\nolimits{\mathop{\rm Spin}\nolimits} \def\mathop{\rm Sup}\nolimits{\mathop{\rm Sup}\nolimits} \def\mathop{\rm Trace}\nolimits{\mathop{\rm Trace}\nolimits} \def\build#1_#2^#3{\mathrel{ \mathop{\kern 0pt#1}\limits_{#2}^{#3}}} \vglue 2cm \font\tencsc=cmcsc10 \def\rightarrow{\rightarrow} \def\otimes{\otimes} \def\varepsilon{\varepsilon} \def\delta{\delta} \def\Delta{\Delta} \def\otimes{\otimes} \def\widetilde{\widetilde} \def\forall{\forall} \def\sigma{\sigma} \def\Lambda{\Lambda} \font\tenbb=msbm10 \font\sevenbb=msbm7 \font\fivebb=msbm5 \newfam\bbfam \textfont\bbfam=\tenbb \scriptfont\bbfam=\sevenbb \scriptscriptfont\bbfam=\fivebb \def\fam\bbfam{\fam\bbfam} \def{\bb C}{{\fam\bbfam C}} \def{\cal H}{{\cal H}} \def\boxit#1#2{\setbox1=\hbox{\kern#1{#2}\kern#1}% \dimen1=\ht1 \advance\dimen1 by #1 \dimen2=\dp1 \advance\dimen2 by #1 \setbox1=\hbox{\vrule height\dimen1 depth\dimen2\box1\vrule}% \setbox1=\vbox{\hrule\box1\hrule}% \advance\dimen1 by .4pt \ht1=\dimen1 \advance\dimen2 by .4pt \dp1=\dimen2 \box1\relax} \def\vrule height 0.5em depth 0.2em width 0.5em{\vrule height 0.5em depth 0.2em width 0.5em} \centerline{\tencsc Cyclic cohomology and Hopf algebras} \vglue 1cm \centerline{A. Connes and H. Moscovici \footnote{(*)}{This material is based upon work supported by the National Science Foundation under Award No. DMS-9706886.}} \bigskip \centerline{\it Dedicated to the memory of Mosh\'e Flato} \vglue 1cm \noindent{\bf Abstract} \smallskip We show by a direct computation that, for any Hopf algebra with a modulus-like character, the formulas first introduced in [CM] in the context of characteristic classes for actions of Hopf algebras, do define a cyclic module. This provides a natural generalization of Lie algebra cohomology to the general framework of Noncommutative Geometry, which covers the case of the Hopf algebra associated to $n$-dimensional transverse geometry [CM] as well as the function algebras of the classical quantum groups. \vglue 1cm \noindent{\bf Introduction} \medskip We shall concentrate in this paper on the interplay between two basic concepts of Noncommutative geometry. The first is cyclic cohomology which plays the same role in Noncommutative geometry as De Rham cohomology plays in differential geometry. The second is Hopf algebras whose actions on noncommutative algebras are analoguous to Lie group actions on ordinary manifolds. \smallskip We shall show by a direct and elementary computation that, for any Hopf algebra with a modulus-like character, the natural cosimplicial module associated to the subjacent coalgebra structure can be upgraded to a {\it cyclic module} (or, rather, a module over the {\it cyclic category} $\Lambda$, cf. [C, III.A]), by invoking both the product and the antipode. This cyclic module was first introduced in [CM] in the context of characteristic classes for actions of Hopf algebras, under a certain condition of existence of sufficiently nondegenerate actions, which made the verification of the axioms tautological. The fact that the latter assumption was superfluous has also been remarked by M.~Crainic [Cr], who recasted our construction in the framework of the Cuntz-Quillen formalism [CQ]. \vglue 1cm \noindent{\bf I Cyclic cohomology and the cyclic category} \medskip The role of cyclic cohomology in Noncommutative geometry can be understood at several levels. In its simplest guise it is a construction of invariants of K-theory, extending to the general framework the Chern-Weil characteristic classes of vector bundles and allowing for concrete computations on Noncommutative spaces. For starters one should prove for oneself the following simple algebraic statement which extends to higher dimension the obvious properties of the K-theory invariant provided by a trace $\tau$ on a noncommutative algebra $A$, by means of the equality, $$ < E \, ,\tau > = \tau(E) \leqno (1) $$ for any idempotent $E, \, E^2 =E$, $E \in M_q(A)$, where the trace $\tau$ is extended to the algebra $M_q(A)$ of matrices over $A$ by, $$ \tau ((a_{i,j})) \, = \sum \tau(a_{i,i}) \leqno (2) $$ To pass from this 0-dimensional situation to, say, dimension 2, one considers a trilinear form $\tau$ on the algebra $A$ which possesses the following compatibility with the algebra structure, reminiscent of the properties of a trace, $$ \eqalign{\tau (a^1, a^2, a^0) \, = \tau(a^0, a^1, a^2) \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \cr \tau (a^0a^1, a^2, a^3) \,-\tau (a^0,a^1a^2, a^3) \,+\tau (a^0, a^1, a^2a^3) \, - \tau(a^3a^0, a^1, a^2) = 0 . \cr \forall a^j \in A}. \leqno (3) $$ \smallskip The statement then asserts that for each idempotent $E, \, E^2 =E$, $E \in M_q(A)$, the scalar $\tau (E, E, E)$ remains constant when $E$ is deformed among the idempotents of $M_q(A)$. This homotopy invariance of the resulting pairing between cyclic cocycles of arbitrary dimension (i.e. multilinear forms on $A$ fulfilling the n-dimensional analogue of (3) ) and K-theory is the starting point of cyclic cohomology. \smallskip At the conceptual level, cyclic cohomology is obtained as an Ext functor by linearisation of the non-additive category of algebras and algebra homomorphisms ([C2]) using the additive category of $\Lambda$-modules where $\Lambda$ is the cyclic category. \smallskip The cyclic category is a small category which can be defined by generators and relations. It has the same objects as the small category $\Delta$ of totally ordered finite sets and increasing maps. Let us recall the presentation of $\Delta$. It has one object $[n]$ for each integer $n$, and is generated by faces $\delta_i, [n-1] \rightarrow [n]$ (the injection that misses $i$), and degeneracies $\sigma_j,[n+1] \rightarrow [n] $ (the surjection which identifies $j$ with $j+1$), with the relations, $$ \delta_j \, \delta_i = \delta_i \, \delta_{j-1} \ \hbox{for} \ i < j \, , \ \sigma_j \, \sigma_i = \sigma_i \, \sigma_{j+1} \qquad i \leq j \leqno (4) $$ $$ \sigma_j \, \delta_i = \left\{ \matrix{ \delta_i \, \sigma_{j-1} \hfill &i < j \hfill \cr 1_n \hfill &\hbox{if} \ i=j \ \hbox{or} \ i = j+1 \cr \delta_{i-1} \, \sigma_j \hfill &i > j+1 \, . \hfill \cr } \right. $$ To obtain $\Lambda$ one adds for each $n$ a new morphism $\tau_n, [n] \rightarrow [n]$ such that, $$ \matrix{ \tau_n \, \delta_i = \delta_{i-1} \, \tau_{n-1} &1 \leq i \leq n , &\tau_n \, \delta_0 = \delta_n \hfill \cr \cr \tau_n \, \sigma_i = \sigma_{i-1} \, \tau_{n+1} &1 \leq i \leq n , &\tau_n \, \sigma_0 = \sigma_n \, \tau_{n+1}^2 \cr \cr \tau_n^{n+1} = 1_n \, . \hfill \cr } \leqno (5) $$ The small category $\Lambda$ is in fact best obtained as a quotient of the following category $E \, \Lambda$. The latter has one object $({\bb Z} , n)$ for each $n$ and the morphisms $f : ({\bb Z} , n) \rightarrow ({\bb Z} , m)$ are non decreasing maps, $(n,m \geq 1)$ $$ f : {\bb Z} \rightarrow {\bb Z} \ , \ f(x+n) = f(x)+m \qquad \forall \, x \in {\bb Z} \, . \leqno (6) $$ One has $\Lambda = (E \, \Lambda) / {\bb Z}$ for the obvious action of ${\bb Z}$ by translation. The original definition of $\Lambda$ (cf. [C2]) used homotopy classes of non decreasing maps from $S^1$ to $S^1$ of degree 1, mapping ${\bb Z} / n$ to ${\bb Z} / m$ and is trivially equivalent to the above. \bigskip Given an algebra $A$ one obtains a module over the small category $\Lambda$ by assigning to each integer $n \geq 0$ the vector space $C^n$ of $n+1$-linear forms $\varphi (x^0 , \ldots , x^n)$ on $A$, while the basic operations are given by $$ \matrix{ (\delta_i \, \varphi) (x^0 , \ldots , x^n) &=& \varphi (x^0 , \ldots , x^i \, x^{i+1} , \ldots , x^n) \hfill &i=0,1,\ldots , n-1 \hfill \cr \cr (\delta_n \, \varphi) (x^0 , \ldots , x^n) &=& \varphi (x^n \, x^0 , x^1 , \ldots , x^{n-1}) \hfill \cr \cr (\sigma_j \, \varphi) (x^0 , \ldots , x^n) &=& \varphi (x^0 , \ldots , x^j , 1 , x^{j+1} , \ldots , x^n) \hfill &j=0,1,\ldots , n \hfill \cr \cr (\tau_n \, \varphi) (x^0 , \ldots , x^n) &=& \varphi (x^n , x^0 , \ldots , x^{n-1}) \, . \hfill \cr } \leqno (7) $$ In the first two lines $\delta_i : C^{n-1} \rightarrow C^n$. In the third line $\sigma_i : C^{n+1} \rightarrow C^n$. Note that $(\sigma_n \, \varphi) (x^0 , \ldots , x^n) = \varphi (x^0 , \ldots , x^n ,1)$, $(\sigma_0 \, \varphi) (x^0 , \ldots , x^n) = \varphi (x^0 , 1 , x^1 , \ldots ,$ $x^n)$. \smallskip These operations satisfy the relations (4) and (5). This shows that any algebra $A$ gives rise canonically to a $\Lambda$-module and allows ([C2][L]) to interpret the cyclic cohomology groups $HC^n(A)$ as $Ext^n$ functors. All of the general properties of cyclic cohomology such as the long exact sequence relating it to Hochschild cohomology are shared by Ext of general $\Lambda$- modules and can be attributed to the equality of the classifying space $B\Lambda$ of the small category $\Lambda$ with the classifying space $BS^1$ of the compact one-dimensional Lie group $S^1$. \vglue 1cm \noindent{\bf II Characteristic classes for actions of Hopf algebras} \medskip Hopf algebras arise very naturally from their actions on noncommutative algebras. Given an algebra $A$, an action of the Hopf algebra ${\cal H}$ on $A$ is given by a linear map, $$ {{\cal H}} \otimes A \rightarrow A \ , \ h \otimes a \rightarrow h(a) $$ satisfying $h_1 (h_2 \, a) = (h_1 \, h_2) (a) \quad \forall \, h_i \in {{\cal H}}, \, a \in A$ and $$ h(ab) = \sum \, h_{(1)} \, (a) \, h_{(2)} \, (b) \qquad \forall \, a,b \in A \, , \ h \in {{\cal H}} \, . \leqno (1) $$ where the coproduct of $h$ is, $$ \Delta(h)= \, \sum \, h_{(1)} \, \otimes \, h_{(2)} \, \leqno (2) $$ In concrete examples, the algebra $A$ appears first, together with linear maps $A \rightarrow A$ satisfying a relation of the form (1) which dictates the Hopf algebra structure. We refer to [CM] for an application of this construction to the leaf space of foliations. \medskip\smallskip \noindent The theory of characteristic classes for actions of ${\cal H}$ extends the construction ([C3]) of cyclic cocycles from a Lie algebra of derivations of a $C^*$ algebra $A$, together with an {\it invariant trace} $\tau$ on $A$. \smallskip In order to cover the nonunimodular case which does appear in the simplest examples, we fix a character $\delta$ of ${\cal H}$ which will play the role of the module of locally compact groups. \smallskip We then introduce the twisted antipode, $$ \widetilde S (y) = \sum \, \delta (y_{(1)}) \, S (y_{(2)}) \ , \ y \in {{\cal H}} \, , \ \Delta \, y = \sum \, y_{(1)} \otimes y_{(2)} \, . \leqno (3) $$ One has $\widetilde S (y) = S (\sigma (y))$ where $\sigma$ is the automorphism $\sigma= \, (\delta \otimes 1) \circ \Delta : {{\cal H}} \rightarrow {{\cal H}}$. \medskip \noindent {\bf Definition 1.} {\it We shall say that a trace $\tau$ on $A$ is $\delta$-invariant under the action of ${\cal H}$ iff the following holds,} $$ \tau (h(a)b) = \tau (a \, \widetilde S (h)(b)) \qquad \forall \, a,b \in A \, , \ h \in {{\cal H}} \, . $$ We have shown in ([CM]) that the definition of the cyclic complex $HC^*_{\delta} ({{\cal H}})$ is uniquely dictated in such a way that the following holds, \medskip \noindent {\bf Proposition 2.} ([CM]) {\it Let $\tau$ be a $\delta$-invariant trace on $A$, then the following defines a canonical map from $HC^*_{\delta} ({{\cal H}})$ to $HC^* (A)$, $$ \matrix{ \gamma (h^1 \otimes \ldots \otimes h^n) \in C^n (A) \, , \ \gamma (h^1 \otimes \ldots \otimes h^n) (x^0 , \ldots , x^n) = \cr \cr \tau (x^0 \, h^1 (x^1) \ldots h^n (x^n)) \, . \cr } $$ } In ([CM]) we needed to assume the existence of enough such actions of ${{\cal H}}$ in order to check that the formulas were actually defining a cyclic module. We shall show below by a direct and elementary computation that, for any Hopf algebra with a modulus-like character $\delta$ as above, the construction of [CM] does yield a cyclic module. \vglue 1cm \noindent{\bf III The cyclic module of a Hopf algebra} \medskip In this section we shall associate a cyclic complex (in fact a $ \Lambda $-module, where $ \Lambda $ is the cyclic category), to any Hopf algebra together with a character $\delta$ such that the $\delta$-twisted antipode is an involution. The resulting cyclic cohomology appears to be the natural candidate for the analogue of Lie algebra cohomology in the context of Hopf algebras, where both the Hochschild cohomology (also called Sweedler cohomology) or the transposed (also called Harrison cohomology) give too naive results. \smallskip Let ${\cal H}$ be a Hopf algebra (over ${\bb C}$) with unit map $\eta : {\bb C} \rightarrow {\cal H}$, counit $\varepsilon : {\cal H} \rightarrow {\bb C}$ and antipode $S : {\cal H} \rightarrow {\cal H}$, $$S * I = I * S = \eta \, \varepsilon.$$ We fix a character $\delta : {\cal H} \rightarrow {\bb C}$, which will play the role of the modular function of a locally compact group. With the usual coproduct notation $$ \Delta h = \sum_{(h)} h_{(1)} \otimes h_{(2)} \quad , \quad h \in {\cal H} \, , $$ we introduce the $\delta$-twisted antipode $$ \widetilde S (h) = \sum_{(h)} \delta (h_{(1)}) \ S (h_{(2)}) \quad , \quad h \in {\cal H} \, . \leqno (1) $$ The elementary properties of $S$ imply immediately that $\widetilde S$ is an algebra antihomomorphism $$ \matrix{ &\widetilde S (h^1 \, h^2) = \widetilde S (h^2) \, \widetilde S (h^1) \quad , \quad \forall \, h^1 , h^2 \in {\cal H} \cr \cr &\widetilde S (1) = 1 \, , \hfill \cr } \leqno (2) $$ a coalgebra twisted antimorphism $$ \Delta \, \widetilde S (h) = \sum_{(h)} S (h_{(2)}) \otimes \widetilde S (h_{(1)}) \quad , \quad \forall \, h \in {\cal H} \, ; \leqno (3) $$ and also that it satisfies $$ \varepsilon \circ \widetilde S = \delta \, . \leqno (4) $$ By transposing the standard simplicial operators underlying the \break Hochschild homology complex of an algebra, one associates to ${\cal H}$, viewed only as a coalgebra, the following natural cosimplicial module: $\{ {\cal H}^{\otimes n} \}_{n \geq 1}$, with face operators $\delta_i : {\cal H}^{\otimes n-1} \rightarrow {\cal H}^{\otimes n}$, $$ \matrix{ &\delta_0 (h^1 \otimes \ldots \otimes h^{n-1}) = 1 \otimes h^1 \otimes \ldots \otimes h^{n-1} \, , \hfill \cr \cr &\delta_j (h^1 \otimes \ldots \otimes h^{n-1}) = h^1 \otimes \ldots \otimes \Delta h^j \otimes \ldots \otimes h^n \, , \ \forall \, 1 \leq j \leq n-1 \, , \hfill \cr \cr &\delta_n (h^1 \otimes \ldots \otimes h^{n-1}) = h^1 \otimes \ldots \otimes h^{n-1} \otimes 1 \hfill \cr } \leqno (5) $$ and degeneracy operators $\sigma_i : {\cal H}^{\otimes n+1} \rightarrow {\cal H}^{\otimes n}$, $$ \sigma_i (h^1 \otimes \ldots \otimes h^{n+1}) = h^1 \otimes \ldots \otimes \varepsilon (h^{i+1}) \otimes \ldots \otimes h^{n+1} \ , \ 0 \leq i \leq n \, . \leqno (6) $$ In [CM, \S~7] the remaining two essential features of a Hopf algebra --{\it product} and {\it antipode} -- are brought into play, to define the {\it cyclic operators} $\tau_n : {\cal H}^{\otimes n} \rightarrow {\cal H}^{\otimes n}$, $$ \tau_n (h^1 \otimes \ldots \otimes h^n) = (\Delta^{n-1} \, \widetilde S (h^1)) \cdot h^2 \otimes \ldots \otimes h^n \otimes 1 \, . \leqno (7) $$ \bigskip \noindent {\bf Theorem 3.} {\it Let ${\cal H}$ be a Hopf algebra endowed with a character $\delta \in {\cal H}^*$ such that the corresponding twisted antipode $(1)$ is an involution: $$ \widetilde{S}^2 = I \, . \leqno (8) $$ Then ${\cal H}_{\delta}^{\natural} = \{ {\cal H}^{\otimes n} \}_{n \geq 1}$ equipped with the operators given by $(5)$--$(7)$ defines a module over the cyclic category $\Lambda$.} \bigskip {\it Proof.} One has to check the relations $$ \matrix{ &\tau_n \, \delta_i = \delta_{i-1} \, \tau_{n-1} \ , \ 1 \leq i \leq n \, , \hfill \cr \cr &\tau_n \, \delta_0 = \delta_n \, , \hfill \cr } \leqno (9) $$ $$ \matrix{ &\tau_n \, \sigma_i = \sigma_{i-1} \, \tau_{n+1} \ , \ 1 \leq i \leq n \, , \hfill \cr \cr &\tau_n \, \sigma_0 = \sigma_n \, \tau_{n+1}^2 \, , \hfill \cr } \leqno (10) $$ $$ \tau_n^{n+1} = I_n \, . \leqno (11) $$ It is the latter which poses a technical challenge. To size it up, let us first look at the case $n=2$. \medskip In what follows we shall only use the basic properties of the product, the coproduct, the antipode and of the twisted antipode (cf. (1)--(4)). We shall also employ the usual notational conventions for the Hopf algebra calculus (cf. [S]). To begin with, $$ \eqalign{ \tau_2 (h^1 \otimes h^2) = & \ \Delta \, \widetilde S (h^1) \cdot h^2 \otimes 1 = \cr = & \ \sum \widetilde S (h^1)_{(1)} \, h^2 \otimes \widetilde S (h^1)_{(2)} \cr = & \ \sum S (h_{(2)}^1) \, h^2 \otimes \widetilde S (h_{(1)}^1) \, . \cr } $$ Its square is therefore: $$ \eqalign{ \tau_2^2 (h^1 \otimes h^2) = & \ \sum S (S (h_{(2)}^1)_{(2)} \, h_{(2)}^2) \, \widetilde S (h_{(1)}^1) \otimes \widetilde S (S (h_{(2)}^1)_{(1)} \, h_{(1)}^2) \cr = & \ \sum S (S (h_{(2)(1)}^1) \, h_{(2)}^2) \, \widetilde S (h_{(1)}^1) \otimes \widetilde S (S (h_{(2)(2)}^1) \, h_{(1)}^2) \cr = & \ \sum S (h_{(2)}^2) \, (S \circ S) \, (h_{(2)(1)}^1) \, \widetilde S (h_{(1)}^1) \otimes \widetilde S (h_{(1)}^2) \, (\widetilde S \circ S) \, (h_{(2)(2)}^1) \cr = & \ \sum S (h_{(2)}^2) \ \hbox{\boxit{6pt}{$S(S(h_{(1)(2)}^1)) \, \widetilde S (h_{(1)(1)}^1)$}} \ \otimes \widetilde S (h_{(1)}^2) \, \widetilde S (S (h_{(2)}^1)) \, . \cr } $$ The term in the box is computed as follows. With $k = h_{(1)}^1$, one has $$ \eqalign{ \sum \, S (S(k_{(2)})) \, \widetilde S (k_{(1)}) = & \ \sum S (S(k_{(2)})) \, \delta (k_{(1)(1)}) \, S (k_{(1)(2)}) \cr = & \ \sum S(S(k_{(2)(2)}) \, \delta (k_{(1)}) \, S (k_{(2)(1)}) = \cr = & \ \sum \delta (k_{(1)}) \, S \left( \sum k_{(2)(1)} \, S (k_{(2)(2)}) \right) \cr = & \ \sum \delta (k_{(1)}) \, S (\varepsilon (k_{(2)}) \, 1) = \cr = & \ \sum \delta (k_{(1)}) \, \varepsilon (k_{(2)}) = \delta \left( \sum k_{(1)} \, \varepsilon (k_{(2)}) \right) \cr = & \ \delta (k) \, . \cr } $$ It follows that $$ \eqalign{ \tau_2^2 (h^1 \otimes h^2) = & \ \sum S (h_{(2)}^2) \, \underbrace{\delta (h_{(1)}^1) \otimes \widetilde S (h_{(1)}^2) \, \widetilde S (S (h_{(2)}^1))} \cr = & \ \sum S (h_{(2)}^2) \otimes \widetilde S (h_{(1)}^2) \, \widetilde S (\widetilde S (h^1)) = \cr = & \ \sum S (h_{(2)}^2) \otimes \widetilde S (h_{(1)}^2) \, h^1 \, , \cr } $$ where we have used first (1) then (8). Thus $$ \eqalign{ \tau_2^2 (h^1 \otimes h^2) = & \sum S (h_{(2)}^2) \otimes \widetilde S (h_{(1)}^2) \cdot 1 \otimes h^1 \cr = & \ \Delta \, \sum \widetilde S (h^2) \cdot 1 \otimes h^1 \, . \cr } $$ In a similar fashion, $$ \eqalign{ \tau_2^3 (h^1 \otimes h^2) = & \ \sum S (S(h_{(2)}^2)_{(2)}) \, \widetilde S (h_{(1)}^2) \, h^1 \otimes \widetilde S (S (h_{(2)}^2)_{(1)}) \cr = & \ \sum S (S(h_{(2)(1)}^2)) \, \widetilde S (h_{(1)}^2) \, h^1 \otimes \widetilde S (S (h_{(2)(2)}^2)) \cr = & \ \sum S (S(h_{(1)(2)}^2)) \, \widetilde S (h_{(1)(1)}^2) \, h^1 \otimes \widetilde S (S (h_{(2)}^2)) \cr = & \ \sum \delta (h_{(1)}^2) \, h^1 \otimes \widetilde S (S (h_{(2)}^2)) = \cr = & \ \sum h^1 \otimes \widetilde{S}^2 (h^2) = h^1 \otimes h^2 \, . \cr } $$ \medskip We now pass to the general case. With the standard conventions of notation, $$ \eqalign{ \tau_n (h^1 \otimes h^2 \otimes \ldots \otimes h^n) = \Delta^{(n-1)} \, \widetilde S (h^1) \cdot h^2 \otimes \ldots \otimes h^n \otimes 1 \cr = \sum S (h_{(n)}^1) \, h^2 \otimes S (h_{(n-1)}^1) \, h^3 \otimes \ldots \otimes S (h_{(2)}^1) \, h^n \otimes \widetilde S (h_{(1)}^1) \, . \cr } $$ Upon iterating once $$ \eqalign{ \tau_n^2 (h^1 \otimes \ldots \otimes h^n) = & \ \sum S (S(h_{(n)}^1)_{(n)} \, h_{(n)}^2 )) \, S (h_{(n-1)}^1) \, h^3 \otimes \cr \otimes & \ S (S (h_{(n)}^1)_{(n-1)} \, h_{(n-1)}^2)) \, S (h_{(n-2)}^1) \, h^4 \otimes \ldots \cr \ldots \otimes & \ S (S (h_{(n)}^1)_{(2)} \, h_{(2)}^2)) \, \widetilde S (h_{(1)}^1) \otimes \widetilde S (S (h_{(n)}^1)_{(1)} \, h_{(1)}^2) \cr = & \ \sum S (h_{(n)}^2) \, S (S (h_{(n)(1)}^1)) \, S (h_{(n-1)}^1) \, h^3 \otimes \cr \otimes & \ S (h_{(n-1)}^2) \, S (S (h_{(n)(2)}^1)) \, S (h_{(n-2)}^1) \, h^4 \otimes \ldots \cr \ldots \otimes & \ S (h_{(2)}^2) \, S (S (h_{(n)(n-1)}^1)) \, \widetilde S (h_{(1)}^1) \otimes \cr \otimes & \ \widetilde S (h_{(1)}^2) \, \widetilde S (S(h_{(n)(n)}^1)) = \cr = & \ \sum S (h_{(n)}^2) \, S (h_{(n-1)}^1 \, S (h_{(n)}^1)) \, h^3 \otimes \cr \otimes & \ S (h_{(n-1)}^2) \, S (h_{(n-2)}^1 \, S (h_{(n+1)}^1)) \, h^4 \otimes \ldots \cr \ldots \otimes & \ S (h_{(2)}^2) \, S (S (h_{(2n-2)}^1 )) \cdot \widetilde S (h_{(1)}^1) \otimes \cr \otimes & \ \widetilde S (h_{(1)}^2) \, \widetilde S ( S (h_{(2n-1)}^1 )) \, . \cr } $$ We pause to note that $$ \sum h_{(n-1)}^1 \, S (h_{(n)}^1) = \sum h_{(n-1)(1)}^1 \, S (h_{(n-1)(2)}^1) $$ equals $$ \varepsilon \, (h_{(n-1)}^1) \, 1 \, , $$ after resetting the indexation. Next $$ \sum \varepsilon (h_{(n-1)}^1) \, h_{(n-2)}^1 $$ gives $h_{(n-2)}^1$ after another resetting. In turn $$ \sum h_{(n-2)}^1 \, S (h_{(n-1)}^1) $$ equals $$ \varepsilon (h_{(n-2)}^1) \, 1 \, , $$ and the process continues. \smallskip In the last step, $$ \eqalign{ & \ \sum S (h_{(n)}^2) \, h^3 \otimes S (h_{(n-1)}^2) \, h^4 \otimes \ldots \cr \ldots \otimes & \ S (h_{(2)}^2) \ \hbox{\boxit{6pt}{$S(S(h_{(2)}^1)) \, \delta (h_{(1)(1)}^1) \, S (h_{(1)(2)}^1)$}} \ \otimes \cr \otimes & \ \widetilde S (h_{(1)}^2) \, \widetilde S ( S (h_{(3)}^1)) \cr = & \ \sum S (h_{(n)}^2) \, h^3 \otimes S (h_{(n-1)}^2) \, h^4 \otimes \ldots \otimes S (h_{(2)}^2) \otimes \widetilde S (h_{(1)}^2) \, h^1 \cr = & \ S (h_{(n)}^2) \otimes S (h_{(n-1)}^2) \otimes \ldots \otimes \widetilde S (h_{(1)}^2) \cdot h^3 \otimes h^4 \otimes \ldots \otimes 1 \otimes h^1 \cr = & \ \Delta^{(n-1)} \, \widetilde S (h^2) \cdot h^3 \otimes h^4 \otimes \ldots \otimes 1 \otimes h^1 \, , \cr } $$ with the boxed term simplified as before. By induction, one obtains for any $j = 1, \ldots , n+1$, $$ \tau_n^j (h^1 \otimes \ldots \otimes h^n) = \Delta^{n-1} \, \widetilde S (h^j) \cdot h^{j+1} \otimes \ldots \otimes h^n \otimes 1 \otimes \ldots \otimes h^{j-1} \, , $$ in particular $$ \tau_n^{n+1} (h^1 \otimes \ldots \otimes h^n) = \Delta^{n-1} \, \widetilde S (1) \cdot h^1 \otimes \ldots \otimes h^n = h^1 \otimes \ldots \otimes h^n \, . $$ The verification of the compatibility relations (9), (10) is straightforward. Indeed, starting with the compatibility with the face operators, one has: $$ \eqalign{ \tau_n \, \delta_0 (1 \otimes h^1 \otimes \ldots \otimes h^{n-1}) = & \ \tau_n (1 \otimes h^1 \otimes \ldots \otimes h^{n-1}) = \cr = & \ \Delta^{n-1} \, \widetilde S (1) \cdot h^1 \otimes \ldots \otimes h^{n-1} \otimes 1 \cr = & \ h^1 \otimes \ldots \otimes h^{n-1} \otimes 1 \cr = & \ \delta_n (h^1 \otimes \ldots \otimes h^{n-1}) \, , \cr } $$ then $$ \eqalign{ \tau_n \, \delta_1 (h^1 \otimes \ldots \otimes h^{n-1}) = & \ \tau_n \,( \Delta \, h^1 \otimes h^2 \otimes \ldots \otimes h^{n-1}) \cr = & \ \sum \tau_n (h_{(1)}^1 \otimes h_{(2)}^1 \otimes h^2 \otimes \ldots \otimes h^{n-1}) \cr = & \ \sum \Delta^{n-1} \, \widetilde S (h_{(1)}^1) \cdot h_{(2)}^1 \otimes h^2 \otimes \ldots \otimes h^{n-1} \otimes 1 = \cr = & \ \sum S (h_{(1)(n)}^1) \, h_{(2)}^1 \otimes S (h_{(1)(n-1)}^1) \, h^2 \otimes \ldots \cr & \ \otimes S (h_{(1)(2)}^1) \, h^{n-1} \otimes \widetilde S (h_{(1)(1)}^1) \cr = & \ \sum \varepsilon (h_{(n)}^1) \, 1 \otimes S (h_{(n-1)}^1) \, h^2 \otimes \ldots \cr & \ \otimes S (h_{(1)}^1) \, h^{n-1} \otimes \widetilde S (h_{(1)}^1) \cr = & \ 1 \otimes S (h_{(n-1)}^1) \, h^2 \otimes \ldots \otimes S (h_{(1)}^1) \, h^{n-1} \otimes \widetilde S (h_{(1)}^1) \cr = & \ \delta_0 \, \tau_{n-1} \, (h^1 \otimes \ldots \otimes h^{n-1}) \, , \cr } $$ and so forth. Passing now to degeneracies, $$ \eqalign{ & \ \tau_n \, \sigma_0 (h^1 \otimes \ldots \otimes h^{n+1}) = \varepsilon (h^1) \, \tau_n (h^2 \otimes \ldots \otimes h^{n+1}) = \cr = & \ \varepsilon (h^1) \, S (h_{(n)}^2) \, h^3 \otimes \ldots \otimes S (h_{(2)}^2) \, h^{n+1} \otimes \widetilde S (h_{(1)}^2) \, , \cr } $$ and on the other hand $$ \eqalign{ & \ \sigma_n \, \tau_{n+1}^2 (h^1 \otimes \ldots \otimes h^{n+1}) = \cr = & \ \sigma_n \left( \sum S (h_{(n+1)}^2) \, h^3 \otimes \ldots \otimes S (h_{(2)}^2) \otimes \widetilde S (h_{(1)}^2) \, h^1 \right) \cr = & \ \sum \varepsilon (\widetilde S (h_{(1)}^2) \, h^1) \, S (h_{(n+1)}^2) \, h^3 \otimes \ldots \otimes S (h_{(2)}^2) \cr = & \ \varepsilon (h^1) \sum \delta (h_{(1)}^2) \, S (h_{(n+1)}^2) \, h^3 \otimes \ldots \otimes S (h_{(2)}^2) \cr = & \ \varepsilon (h^1) \, S (h_{(n)}^2) \, h^3 \otimes \ldots \otimes S (h_{(2)}^2) \, h^{n+1} \otimes \widetilde S (h_{(1)}^2) \, . \cr } $$ In the next step $$ \eqalign{ & \ \tau_n \, \sigma_1 (h^1 \otimes \ldots \otimes h^{n+1}) = \varepsilon (h^2) \, \tau_n (h^1 \otimes h^3 \otimes \ldots \otimes h^{n+1}) \cr = & \ \varepsilon (h^2) \cdot \Delta^{n-1} \, \widetilde S (h^1) \cdot h^3 \otimes \ldots \otimes h^{n+1} \otimes 1 \, , \cr } $$ while on the other hand $$ \eqalign{ & \ \sigma_0 \, \tau_{n+1} (h^1 \otimes \ldots \otimes h^{n+1}) = \cr & \ \sigma_0 (S (h_{(n+1)}^1) \, h^2 \otimes \ldots \otimes S (h_{(2)}^1) \, h^{n+1} \otimes \widetilde S (h_{(1)}^1)) \cr = & \ \varepsilon (h^2) \cdot \varepsilon (h_{(n+1)}^1) \cdot S (h_{(n)}^1) \, h^3 \otimes \ldots \otimes S (h_{(2)}^1) \, h^{n+1} \otimes \widetilde S (h_{(1)}^1) \cr = & \ \varepsilon (h^2) \cdot S (h_{(n-1)}^1) \, h^3 \otimes \ldots \otimes S (h_{(2)}^1) \, h^{n+1} \otimes \widetilde S (h_{(1)}^1) \, , \cr } $$ and similarly for $i = 2, \ldots n$.~\vrule height 0.5em depth 0.2em width 0.5em \bigskip The cohomology of the $(b,B)$-bicomplex corresponding to the cyclic module ${\cal H}_{\delta}^{\natural}$ is, by definition, the {\it cyclic cohomology} $H \, C_{\delta}^* ({\cal H})$ of ${\cal H}$ relative to the modular character $\delta$. \bigskip \noindent When ${{\cal H}} = {{\cal U}} ({\bf G})$ is the envelopping algebra of a Lie algebra, there is a natural interpretation of the Lie algebra cohomology, $$ H^* ({\bf G} , {\bb C}) = H^* ({{\cal U}} ({\bf G}) , {\bb C}) $$ where the right hand side is the Hochschild cohomology with coefficients in the ${{\cal U}} ({\bf G})$-bimodule ${\bb C}$ obtained using the augmentation. In general, given a Hopf algebra ${{\cal H}}$ one can dualise (this is the construction of the Harrison complex), the construction of the Hochschild complex $C^n ({{\cal H}}^* , {\bb C})$ where ${\bb C}$ is viewed as a bimodule on ${{\cal H}}^*$ using the augmentation, i.e. the counit of ${{\cal H}}^*$. This gives the above operations: ${{\cal H}}^{\otimes (n-1)} \rightarrow {{\cal H}}^{\otimes n}$, defining a cosimplicial space. \smallskip When applied to the Hopf algebra ${{\cal H}}$ of functions on an affine algebraic group, this dual-Hochschild or Harrison cohomology gives simply the vector space of invariant twisted forms and ignores the group cohomology. The second assertion of the following proposition shows that the cyclic cohomology $H \, C_{\delta}^* ({{\cal H}})$, gives a highly nontrivial answer, thanks precisely to the action of the $B$ operator ([CM]). \medskip \noindent {\bf Proposition 4.} ([CM]) 1) {\it The periodic cyclic cohomology $H \, C_{\delta}^* ({{\cal H}})$, for ${{\cal H}} = {{\cal U}} ({\bf G})$ the envelopping algebra of a Lie algebra ${\bf G}$ is isomorphic to the Lie algebra homology $H_* ({\bf G} , {\bb C})$ where ${\bb C}$ is a ${\bf G}$-module using the module $\delta$ of $G$.} \noindent 2) {\it Let ${{\cal H}} = {{\cal U}} ({\bf G})_*$ be the Hopf algebra of polynomials in the coordinates on an affine simply connected nilpotent group $G$. The periodic cyclic cohomology $H \, C_{\delta}^* ({{\cal H}})$, is isomorphic to the Lie algebra cohomology of ${\bf G}$ with trivial coefficients.} \smallskip \noindent We refer to [CM] for the proof which was also reproduced in [Cr]. \medskip Finally we should point out that the existence of a twisted antipode $\widetilde S$ of square one is still a partial unimodularity condition on a Hopf algebra. It was crucial for the results of [CM] that this condition is actually fulfilled for the Hopf algebra associated to $n$-dimensional transverse geometry of foliations. Moreover, as we shall see now, it is also fulfilled by the most popular quantum groups. First, recall that if ${\cal H}$ is {\it quasi-triangular} (also called {\it braided} ), with universal $R$-matrix $R$, then (see e.g. [K]) $$ S^2 (x) = u \, x \, u^{-1} \, \, , x \in {\cal H} $$ with $$ u = \sum \, S ( R^{(2)}) \, R^{(1)} \ , \quad \varepsilon (u) = 1 $$ and $$ \Delta \, u = (R_{21} \, R)^{-1} \, (u \otimes u) \, . $$ \noindent If in addition $u \, S(u) = S(u) \, u$, which is central, has a central square root $\theta$, such that $$ \Delta (\theta) = (R_{21} \, R)^{-1} \, (\theta \otimes \theta) \ , \quad \varepsilon (\theta) = 1 \ , \quad S(\theta) = \theta \, , $$ then ${\cal H}$ is called a {\it ribbon} algebra. Any braided Hopf algebra ${\cal H}$ can be canonically embedded in a ribbon algebra (cf. [RT]): $$ {\cal H} (\theta) = {\cal H} \, [\theta] / (\theta^2 - u \, S(u)) $$ If ${\cal H}$ is a ribbon algebra, then $$ \delta = \theta \, u^{-1} \, , $$ is a group-like element: $$ \Delta \, \delta = \delta \otimes \delta \ , \quad \varepsilon (\delta) = 1 \ , \quad S(\delta) = \delta^{-1} \, . $$ Defining $$ \widetilde S = \delta \cdot S \, , $$ one obtains a twisted antipode which satisfies the property ${\widetilde S}^2 = 1 \, .$ \noindent Indeed, $$ \matrix{ {\widetilde S}^2 (x) &= &\delta \, S (\delta \, S (x)) = \delta \, S^2 (x) \, \delta^{-1} \hfill \cr &= &\delta \, u \, x \, u^{-1} \, \delta^{-1} = \theta \, x \, \theta^{-1} = x \, , \hfill \cr } $$ because $\theta$ is central. \medskip By dualizing the above definitions one obtains the notion of a {\it cobraided}, resp. {\it coribbon} algebra. Among the most prominent examples of coribbon algebras are the function algebras of the classical quantum groups $GL_q (N)$, $SL_q (N)$, $SO_q (N)$, $O_q (N)$ and $Sp_q (N)$. For a coribbon algebra ${\cal H}$, the analogue of the above {\it ribbon group-like element} $\delta$ is the {\it ribbon character} $\delta \in {\cal H}^*$. The corresponding twisted antipode $${\widetilde S} = \delta * S \, , $$ satisfies again the condition ${\widetilde S}^2 = 1$. \vglue 1cm \noindent {\bf References} \bigskip \item{[C1]} A.~Connes, {\it Noncommutative Geometry}, London-New York, Academic Press, 1994. \medskip \item{[C2]} A. Connes : Cohomologie cyclique et foncteur $Ext^n$. {\it C.R. Acad. Sci. Paris}, Ser.I Math {\bf 296} (1983). \medskip \item{[C3]} A. Connes : $C^*$ alg\`ebres et g\'eom\'etrie differentielle. {\it C.R. Acad. Sci. Paris}, Ser.~A-B {\bf 290} (1980). \medskip \item{[CM]} A.~Connes and H.~Moscovici, Hopf Algebras, Cyclic Cohomology and the Transverse Index Theorem, {\it Commun. Math. Phys.} {\bf 198} (1998), 199-246. \medskip \item{[Cr]} M.~Crainic, Cyclic Cohomology of Hopf Algebras and a Noncommutative Chern-Weil Theory, {\it Preprint} QA/9812113. \medskip \item{[CQ]} J.~Cuntz and D.~Quillen, Cyclic homology and singularity, {\it J. Amer. Math. Soc.} {\bf 8} (1995), 373-442. \medskip \item{[K]} C.~Kassel, {\it Quantum Groups}, Springer-Verlag, 1995. \medskip \item{[L]} J.L.~Loday, {\it Cyclic Homology}, Springer, Berlin, Heidelberg, New York, 1998. \medskip \item{[RT]} N. Yu.~Reshetikhin and V. G.~Turaev, Ribbon graphs and their invariants derived from quantum groups, {\it Commun. Math. Phys.} {\bf 127} (1990), 1-26. \medskip \item{[S]} M.E.~Sweedler, {\it Hopf Algebras}, W.A.~Benjamin, Inc., New York, 1969. \bye
\section{INTRODUCTION} The Lagrange points $L_4$ and $L_5$ are stable in the restricted three body problem (e.g., Danby 1988). However, the long-term survival of Trojans around the Lagrange points of the planets in the presence of perturbations from the remainder of the Solar System is a difficult and still unsolved problem (e.g., \'Erdi 1997). Jovian Trojan asteroids have been known since the early years of this century, while a number of Saturnian moons (e.g., Dione and Helene, Tethys and Calypso, Tethys and Telesto) also form Trojan configurations with their parent planet. However, whether there exist Trojan-like bodies associated with the other planets has been the matter of both observational activity (e.g., Tombaugh 1961, Kowal 1971) and intense theoretical speculation (e.g., Weissman \& Wetherill 1974, Mikkola \& Innanen 1990, 1992). The answer to this problem came in 1990, with the discovery of 5261 Eureka, the first Trojan around Mars (see Mikkola et al. 1994 for details). The last few months of 1998 have seen further remarkable progress with the discovery of one certain Martian Trojan, namely 1998 VF31, as well as two further candidates, namely 1998 SD4 and 1998 QH56 (see the {\it Minor Planet Electronic Circulars 1998-W04, 1998-R02, 1998-S20 } and the {\it Minor Planet Circular 33085}). The suggestion that 1998 QH56 and 1998 VF31 might be Martian Trojans was first made by G.V. Williams. These recent discoveries raise very directly the following questions. Are there any more Martian Trojans? If so, where should the observational effort be concentrated? Of course, the first question can only be answered at the telescope, but the resolution of the second question is provided in this {\it Letter}. By integrating numerically an ensemble of inclined and in-plane orbits in the vicinity of the Martian Lagrange points for 25 and 60 million years respectively, the stable r\'egimes are mapped out. On re-simulating and sampling the ensemble of stable orbits, the probability density of Martian Trojans as a function of longitude and inclination can be readily obtained. If a comparatively puny body such as Mars possesses Trojans, the existence of such objects around the larger terrestrial planets also merits very serious attention. There are Trojan orbits associated with Venus and the Earth that survive for tens of millions of years (e.g., Tabachnik \& Evans 1998). If objects populating such orbits exist, they must be small else they would have been found by now. \section{MARTIAN TROJANS} Saha \& Tremaine (1992, 1994) have taken the symplectic integrators developed by Wisdom \& Holman (1991) and added individual planetary timesteps to provide a fast code that it is tailor-made for long numerical integrations of low eccentricity orbits in a nearly Keplerian force field. In our simulations, the model of the Solar System consists of the eight planets from Mercury to Neptune, together with test particles starting near the Lagrange points. The effect of Pluto on the evolution of Martian Trojans is quite negligible. Of course, the Trojan test particles are perturbed by the Sun and planets but do not themselves exert any gravitational forces. The initial positions and velocities of the planets, as well as their masses, are provided by the JPL Planetary and Lunar Ephemerides DE405 and the starting epoch is JD 2440400.5 (28 June 1969). All our simulations include the most important post-Newtonian corrections, as well as the effects of the Moon. Individual timesteps are invaluable for this work, as orbital periods are much smaller in the Inner Solar System than the Outer. For all the computations described in this Letter, the timestep for Mercury is $14.27$ days. The timesteps of the planets are in the ratio $1:2:2:4:8:8:64:64$ for Mercury moving outwards, so that Neptune has a timestep of $2.5$ years. The Trojan particles all have the same timestep as Mercury. These values were chosen after some experimentation to ensure the relative energy error has a peak amplitude of $\approx 10^{-6}$ over the tens of million year integration timespans. After each timestep, the Trojan test particles are examined to see whether their orbits have become hyperbolic or if they have entered the planet's sphere of influence (defined as $r_{\rm s} = a_{\rm p} M_{\rm p}^{2/5}$ where $a_{\rm p}$ and $M_{\rm p}$ are the semimajor axis and mass of the planet). If so, they are terminated. In methodology, our calculations are very similar to the magisterial work on the Trojan problem for the four giant planets by Holman \& Wisdom (1993). The earlier calculations of Mikkola \& Innanen (1994, 1995) on the Trojans of Mars for timespans of between tens of thousands and 6 million years have also proved influential. Our integrations of Trojan orbits are pursued for durations ranging from 25 to 60 million years, the longest integration periods currently available. Nonetheless, the orbits have been followed for only a tiny fraction of the age of the Solar System ($\sim 4.5$ Gigayears), so it is wise to remain a little cautious about our results. Figure 1 shows the results of our first experiment. Here, the orbits of 1080 Trojan test particles around Mars are integrated for 25 million years. The initial inclinations of the test particles (with respect to the plane of Mars' orbit) are spaced every $2^\circ$ from $0^\circ$ to $90^\circ$ and the initial longitudes (again with respect to Mars) are spaced every $15^\circ$ from $0^\circ$ to $360^\circ$. The starting semimajor axes and the eccentricities of the Trojans are the same as the parent planet. Only the test particles surviving till the end of the 25 million year integration are marked on the Figure. The survivors occupy a band of inclinations between $10^\circ$ and $40^\circ$ and longitudes between $30^\circ$ and $120^\circ$ (the $L_4$ Lagrange point) or $240^\circ$ and $330^\circ$ (the $L_5$ point). On the basis of 4 million year timespan integrations, Mikkola \& Innanen (1994) claim that stable Martian Trojans have inclinations between $15^\circ$ and $30^\circ$ and between $32^\circ$ and $44^\circ$ with respect to Jupiter's orbit. Our longer integrations seem to suggest a more complex picture. Mikkola \& Innanen's instability strip between $30^\circ$ and $32^\circ$ can be detected in Figure~1, but only for objects near $L_4$ with initial longitudes $\mathrel{\spose{\lower 3pt\hbox{$\sim$} 60^\circ$. In particular, this instability strip does not exist around $L_5$ and here Trojans with starting inclinations $30^\circ < i < 32^\circ $ seem to be stable -- as is also evidenced by the recent discovery of 1998 VF31. Marked on the figure are the instantaneous positions of the two certain Martian Trojans, namely 5261 Eureka (marked as a red circle) and 1998 VF31 (a green circle), as well as the two candidates 1998 QH56 (a blue circle) and 1998 SD4 (a yellow circle). It is delightful to see that the two securely established Trojans lie within the stable zone, which was computed by Tabachnik \& Evans (1998) before the discovery of 1998 VF31. In fact, they live deep within the heart of the zone, suggesting that they may even be primordial. The two candidates (1998 QH56 and 1998 SD4) lie closer to the rim. Let us finally note that Trojans starting off in or near the plane of Mars' orbit are unstable. This has been confirmed by an extensive survey of in-plane Martian Trojans. On integrating 792 test particles with vanishing inclination but with a range of longitudes and semimajor axes, we found that all are unstable on timescales of $60$ million years. Martian Trojans with low inclinations are not expected. It is useful to an observer hoping to discover further Trojans to provide plots of the probability density. Accordingly, let us re-simulate the stable zones with much greater resolution. This is accomplished by placing a total of 746 test particles every $1^\circ$ in initial inclination and every $5^\circ$ in initial longitude so as to span completely the stable regions. This ensemble of orbits is then integrated and the orbital elements are sampled every 2.5 years to provide the plots displayed in Figure 2. The upper panel shows the meshed surface of the probability density as a function of both inclination to the invariable plane and longitude with respect to the planet. The asymmetry between the two Lagrange points is evident. The lower panels show the projections of the meshed surface onto the principal planes -- in particular, for the inclination plot, we have shown the contribution at each Lagrange point separately. There are a number of interesting conclusions to be drawn from the plots. First, as shown by the dotted line, the probability density is bimodal at $L_4$. It possesses a flattish maximum at inclinations between $15^\circ$ and $30^\circ$ and then falls sharply, before rising to a second maximum at $36^\circ$. At $L_5$, all inclinations between $15^\circ$ and $40^\circ$ carry a significant probability, though the smaller inclinations in this band are most favored. It is within these inclination windows that the observational effort should be most concentrated. Second, the probability density is peaked at longitudes of $\sim 60^\circ$ ($L_4$) and $\sim 300^\circ$ ($L_5$). The most likely place to observe one of these Trojans is indeed at the classical locations of the Lagrange points. This is not intuitively obvious, as any individual Trojan is most likely to be seen at the turning points of its longitudinal libration. There are two reasons why this effect is not evident in our probability density plots. First, our figures refer to an ensemble of Trojans uniformly populating the stable zone. So, the shape of the stable zone also plays an important role in controlling the position of the maximum of the probability density. Second, the positions of the Lagrange points themselves are oscillating and so the turning points of the longitudinal libration do not occur at the same locations, thus smearing out the enhancement effect. Table 1 lists the orbital elements of the two secure Martian Trojans and the two candidates, as recorded by the Minor Planet Center. From the instantaneous elements, it is straightforward to simulate the trajectories of the objects. Figure 3 shows the orbits plotted in the plane of longitude (with respect to Mars) versus semimajor axis. As the figures illustrate, both 5261 Eureka and 1998 VF31 are stable and maintain their tadpole character (see e.g., Garfinkel 1977) for durations of 50 million years. Based on preliminary orbital elements, Mikkola et al. (1994) integrated the orbit of 5261 Eureka and found that its longitudinal libration was large, quoting $297^\circ \pm 26^\circ$ as the typical range in the longitudinal angle. Our orbit of 5261 Eureka, based on the latest orbital elements, seems to show a smaller libration of $285-314^\circ$. The remaining two objects that have been suggested as Martian Trojans, 1998 QH56 and 1998 SD4, both enter the sphere of influence of Mars -- in the former case after $\sim 500\,000$ years, in the latter case after $\sim 100\,000$ years. Although the orbits are Mars crossing, their eccentricities remain low and their inclinations oscillate tightly about mean values until the Mars' sphere of influence is entered. It is possible that these objects were once Trojans and have been ejected from the stable zones, a possibility that receives some support from their locations in Figure 1 at the fringes of the stable zones. Of course, another possibility is that they are ejected asteroids from the Main Belt. The fact that both confirmed Martian Trojans lie deep within the stable zones in Figure 1 suggests that these objects may be primordial. If so, we can get a crude estimate of possible numbers by extrapolation from the number of Main Belt asteroids (c.f., Holman 1997, Evans \& Tabachnik 1999). The number of Main Belt asteroids $N_{\rm MB}$ is $N_{\rm MB} \mathrel{\spose{\lower 3pt\hbox{$\sim$} \Sigma_{\rm MB} A_{\rm MB} f$, where $A_{\rm MB}$ is the area of the Main Belt, $\Sigma_{\rm MB}$ is the surface density of the proto-planetary disk and $f$ is the fraction of primordial objects that survive ejection (which we assume to be a universal constant). Let us take the Main Belt to be centered on $2.75$ AU with a width of $1.5$ AU. The belt of Martian Trojans is centered on $1.52$ AU and has a width of $ \mathrel{\spose{\lower 3pt\hbox{$\sim$} 0.0025$ AU. If the primordial surface density falls off inversely proportional to distance, then the number of Martian Trojans $N_{\rm MT}$ is \begin{equation} N_{\rm MT} \mathrel{\spose{\lower 3pt\hbox{$\sim$} \Bigl( {2.75\over 1.52} \Bigr) \Bigl( {1.52 \times 0.0025 \over 2.75 \times 1.5} \Bigr) N_{\rm MB} \approx 0.0016 N_{\rm MB} \end{equation} The number of known Main Belt asteroids with diameters $\mathrel{\spose{\lower 3pt\hbox{$\sim$} 1$ km is $\mathrel{\spose{\lower 3pt\hbox{$\sim$} 40000$, which suggests that the number of Martian Trojans is $\mathrel{\spose{\lower 3pt\hbox{$\sim$} 50$. \section{CONCLUSIONS} Motivated by the recent discovery of a new Mars Trojans (1998 VF31) as well as further possible candidates (1998 QH56, 1998 SD4), this paper has provided maps of the stable zones for Martian Trojans and estimates of the numbers of undiscovered objects. For Mars, the observational effort should be concentrated at inclinations satisfying $15^\circ < i < 30^\circ$ and $32^\circ < i < 40^\circ$ for the $L_4$ Lagrange point and between $15^\circ$ and $40^\circ$ for $L_5$. These are the spots where the probability density is significant (see Figure 2), though the lower inclinations in these bands are slightly more favored than the higher. Trojans in or close the orbital plane of Mars are unstable. Crude estimates suggest there may be as many as $\sim 50$ undiscovered Martian Trojans with sizes $\mathrel{\spose{\lower 3pt\hbox{$\sim$} 1$ km . The orbits of 5261 Eureka and 1998 VF31 remain Trojan-like for durations of at least 50 million years. The other candidates, 1998 QH56 and 1998 SD4, are not currently Trojans, though it is conceivable that they may once have been. Both objects will probably enter the sphere of influence of Mars after $\mathrel{\spose{\lower 3pt\hbox{$\sim$} 0.5$ million years. \acknowledgments NWE is supported by the Royal Society, while ST acknowledges financial help from the European Community. We wish to thank John Chambers, Luke Dones, Seppo Mikkola, Prasenjit Saha and Scott Tremaine for many helpful comments and suggestions. We are also grateful for the remarkable service to the academic community provided by the Minor Planet Center. The anonymous referee helpfully provided improved orbital elements for the Trojan candidates for our integrations.
\section{Introduction} We have carried out an absorption study in the 21 cm line of atomic hydrogen in 25 directions in the Galaxy. These directions have been selected from previous optical absorption studies in the lines of singly ionized calcium (CaII) and neutral sodium (NaI). In this paper, we describe the observations and present the data obtained by us. A discussion of the results and the conclusions drawn from the study are in the accompanying paper (Rajagopal, Srinivasan and Dwarakanath 1998; paper II). Our observations were primarily intended to study the velocities of the HI absorption features, their linewidths, and in combination with existing HI emission measurements obtain the spin temperature of the absorbing gas. Our study was motivated by the following questions: \begin{itemize} \item[(i)] How are interstellar clouds seen in optical absorption related to those seen in HI absorption and emission ? \item[(ii)] What is the nature of the relatively fast clouds so commonly seen in the absorption lines of NaI and CaII ? \end{itemize} To clarify these and related issues we briefly summarize the salient historical background. Some of the earliest information about the Interstellar Medium (ISM) came from observations of optical absorption in the H and K ($\lambda\lambda 3933, 3968{\rm\AA}$) lines of CaII, and the D$_1$ and D$_2$ ($\lambda\lambda 5889, 5895{\rm\AA}$) lines of NaI towards bright stars. Adams (1949) made an extensive study of the absorption lines of CaII towards nearly 300 O and B stars. These observations were later extended to the D lines of NaI (Routly and Spitzer 1952; Hobbs 1969), and to high latitude stars (M\"{u}nch and Zirrin 1961).In the simplest model, these lines were attributed to interstellar gas in the form of clouds. The existence of a {\em hot intercloud medium} was conjectured by Spitzer (1956) and the theoretical basis for a two-phase model followed ( Field 1965; Field, Goldsmith and Habing 1969). Independently, a global picture of the ISM emerged from radio observations of the 21 cm HI line (Clark, Radhakrishnan and Wilson 1962; Clark 1965; Radhakrishnan et. al. 1972). These and later studies have established that an important constituent of the ISM are cool diffuse clouds in pressure equilibrium with a warmer intercloud medium. The notion of interstellar ``clouds'' was thus invoked to explain both optical and radio observations. {\em But one was left speculating as to whether the two populations were the same.} Surprisingly the answer to the above question still remains incomplete. Many properties of the two populations seem to differ. In particular the number of clouds per kiloparsec (Blaauw 1952; Radhakrishnan and Goss 1972; Hobbs 1974; Radhakrishnan and Srinivasan 1980), and the velocity distributions deduced from optical and radio observations, respectively, do not agree. The latter discrepancy is discussed in more detail in a subsequent section. In this study, we have attempted a direct comparison by doing HI absorption measurements towards the bright stars themselves. Of course, this can be done only in those lines of sight (towards stars studied previously) where there are strong enough radio sources. We were actually able to do this in about 25 directions. From this data we obtain the velocities of the absorbing gas, and in combination with previous emission measurements in the same directions, the spin temperature of the HI gas in the very entities seen in optical absorption. This is the first attempt at such a direct comparison. Identifying the optical absorption lines with 21 cm absorption features is particularly important because of a long standing puzzle. Optical absorption along many lines of sight reveal two sets of absorption features. One set occur at near zero (low) velocities and another at higher velocities with respect to the local standard of rest. The faster clouds have measured velocities well in excess of the radial component of Galactic rotation one could attribute to them. In a classic study of Adams' data, Blaauw (1952) clearly showed the existence of a high velocity tail extending upto as high as 100 \kms in the distribution of random velocities of the optical absorption features. Thus the existence of a high velocity population of clouds was firmly established. There was also a hint that these ``fast'' clouds may belong to a different population. They exhibited the well known Routly-Spitzer effect (Routly and Spitzer 1952) \ie the NaI to CaII ratios in these clouds were significantly lower than those in the lower velocity clouds. In the decade following the discovery of the 21 cm line attempts were made to detect the gas seen in optical absorption. These involved measuring HI emission in the direction of stars which show optical absorption features in their spectra in order to compare the HI spectra with the CaII and NaI spectra (Takakubo 1967; Habing 1968, 1969; Goldstein and MacDonald 1969; Goniadzki 1972; Heiles 1974 and others). Habing's study in particular targeted selected stars to attempt a direct face off. The results of these early studies threw up another intriguing fact. The low velocity features appeared to be well correlated in HI emission and optical absorption, \ie whenever the optical spectra showed a low velocity absorption feature (v$<$ 10 \kms), there was HI emission at the corresponding velocity. However, the high velocity features were in general {\em absent} in HI emission down to low limits (T$_b <$ 1 K). This seemed to be the case in all directions in the sky. The typical beam sizes in the early experiments to measure HI emission were $\sim$ 0.5$^\circ$. The angular size of the absorbing gas could have been much smaller leading to considerable beam dilution. This could be the reason why the clouds were not detected in HI emission. This exemplifies the difficulties in comparing the features seen in optical absorption with arc-second resolution with those seen in radio emission using a comparatively large beam. This is one of the major factors which led us to attempt an HI {\it absorption study}. The resolution achieved is of the same order as in optical absorption enabling a surer comparison. Despite the failure of the early emission measurements to detect the ``fast'' clouds there is some indication for a population of high velocity clouds from an independent HI absorption study, though this is far from being firmly established. Radhakrishnan and Srinivasan (1980) from a detailed analysis of the 21 cm absorption profile towards the Galactic centre suggested that the peculiar velocities of HI clouds cannot be understood in terms of a single Gaussian distribution with a dispersion of about 5 $-$ 7 \kms, the velocity distribution for ``standard'' HI clouds. There was an indication of a second population of weakly absorbing clouds with a velocity dispersion $\sim 35$ \kms. But this has not been confirmed by other studies. Our study is an attempt to address this important but neglected problem. \section{Scope of the present observations} We chose 25 stars towards which both low and high velocity absorption features have been seen (from CaII or NaI atoms or both). Two of the stars HD 14134 and 14143 happen to be in the same field (in our radio observations) and hence we have measured absorption in only 24 fields. The positions of all the selected stars as projected on the plane of the Galaxy are shown in Fig 1. The Galactic spiral arms shown in this \begin{figure*} \begin{center} \mbox{ \epsfig{file=galplot.ps,width=12.5cm}} \end{center} \caption[Positions of our sample of stars, projected on the Galactic plane]{24 of our sample of 25 stars (HD 119608 lies outside the region shown) are shown as dots on the contour plot of the Galactic spiral arms from Taylor and Cordes. The squares show the local OB associations from Blaauw (1985). The Sun is at 8.5 kpc from the Galactic center. The object with dense contours to the left of the sun is the Gum Nebula. The outer arm shown is the Perseus arm, while the inner one is the Sagittarius arm. The three stars bunched together well inside the Perseus arm are HD 14134, 14143 and 14818. The Perseus OB1 cluster is seen close to them. The two stars seen projected on the Sagittarius arm are HD 166937 and 159176. These fields show high velocity HI absorption.} \end{figure*} figure are from the electron density model by Taylor and Cordes (1993). In 13 of these directions HI emission measurements had been carried out earlier by Habing (1968,1969). Most of the fields studied by Habing were chosen at $|\rm b|$ $>$ 20 $^{\circ}$ to avoid complicated HI emission arising in the plane. Four stars, however, have $|\rm b|$ $<$ 10$^{\circ}$ and were selected since they show CaII velocities well outside the limits of Galactic rotation at the estimated distances. Out of the remaining 12 fields, 3 have been investigated by Goniadzki (1972) for HI emission and one by Takakubo (1967). The rest of the fields have not been investigated in the radio. However, we have been able to get the brightness temperatures and column densities of HI in these directions also from the Leiden-Green Bank survey (Burton 1985). In the observations described in this paper we have obtained HI absorption spectra towards a radio source whose line of sight is close to that of the star in each of the 24 fields. We have been able to identify fairly strong radio sources (in most cases $\gsim$ 50 mJy at 21 cm) within 30' to 40' of the star in every field. In more than half the fields the radio sources are within 10' and several within 5' of the star in question. If the cloud is halfway to the star, then given a 10' separation between the star and the radio source one will sample a cloud of size less than 2 pc even for the farthest stars in the sample which are about 2 kpc away. This is well within the range of accepted sizes for standard clouds (Spitzer 1978). In some fields we had to arrive at a compromise between the minimum cloud size we could sample (\ie the proximity of the projected position of the background source to the star) and the optical depth sensitivity we could achieve. The latter constrains us to fairly strong radio sources for reasonable integration times. This flux density requirement limits the number of sources that one can find close to the star and hence the trade off. Most of the background sources selected are from the NRAO/VLA Sky Survey (Condon et al. 1996). We aimed at an optical depth sensitivity of $\tau$ = 0.1 and have done better than that in several cases. The worst case limit on the optical depth is 0.4. \section{Observations and data analysis} We obtained the absorption spectra using the Very Large Array (VLA\footnote{The VLA is operated by the National Radio Astronomy Observatory (NRAO). The NRAO is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc.}) in the $\sim$ 1 km and $\sim$ 11 km configurations with a synthesized beam size of 44'' and 4'', respectively, at 21 cm. The observations were carried out with a total bandwidth of 1.56 MHz using both polarizations and 128 channels. We used 0134+329 as the primary flux calibrator. For each source we observed a nearby secondary calibrator to do a phase and bandpass calibration. The calibrator was observed with the frequency band shifted by 1.5 MHz, corresponding to a velocity shift $\sim$ 300 km s$^{-1}$. This shift in velocity is sufficient to move the band away from any Galactic feature which might affect the bandpass calibration. Typically, each of these calibrators were observed for $\sim$ 10 minutes. The typical strengths of these calibrator sources being $>$ 1 Jy, 10 minutes of integration time was sufficient to achieve a signal-to-noise ratio greater than that on the source by a factor of 2. After on line Hanning smoothing over 2 channels, the frequency resolution obtained was $\sim$ 12 kHz which corresponds to a velocity resolution of $\sim$ 2.5 \kms. The integration time on each source was chosen to give an optical depth sensitivity of \mbox{$\tau \sim$ 0.1}, and ranged from a few minutes to more than an hour. In all a total of \mbox{$\sim$ 30} hours were spent on the sources, split over several sessions of observing. The analysis was carried out using the Astronomical Image Processing System (AIPS) developed by the National Radio Astronomy Observatory (NRAO). The first step was to make continuum images of each field and examine them for bright sources, including the target source. Our observations coincided with the ongoing NRAO/VLA Sky Survey (in the B and D configurations) and most of our observations were carried out during the daytime. Hence we had to contend with moderate to strong levels of interference over most of the band owing to which approximately 10\% of the data was lost. The next stage involved removing the continuum level from all channels. The task used for this was UVLIN. The continuum level to be removed is determined from a linear fit to the visibility levels in selected channels, which are chosen to be free of interference as well as any spectral features. Finally the image cubes were made. The imaging was typically done over an area of 512 by 512 pixels. In some cases several cubes were synthesised for different areas to cover all the sources of interest in the field. The spectra towards each of the sources were analysed with the Groningen Image Processing System (GIPSY; van der Hulst et al. 1992). \renewcommand{\arraystretch}{.9} \begin{table}[h] \begin{center} \begin{tabular}{|l|l|l|l|c|c|r|}\hline Field & d & l$^{{\tiny II}}$, b$^{{\tiny II}}$ & $\theta$ & V$_{lsr} $ & $\tau$ & V$ _{Gal}$\\ \hline\hline {} & pc & & arcmin & \kms & & {\scriptsize km s$^{-1}$}\\ \hline 14143$^*$& 2000 & 135, $-$4& 10.0& $-$52.8, $-$50.3, $-$46.1,& 0.09& $-$25 \\ & & & & $-$11.2, $-$8.2& & \\ 14818& 2200& 136, $-$4& 10.0& $-$55.8, $-$14.0, $-$3.7& 0.05& $-$25 \\ 21278& 190& 148, $-$6& 37.0& 2.8& 0.08& $>$0 \\ 21291& 1100& 141, 3& 36.0 & $-$31.2, $-$7.0, $-$4.9& 0.07& $-$10 \\ 24912& 46& 160, $-$13& 2.2& 4.3, 6.0, $-$82.0& 0.39& $-$5 \\ 25558& 220& 185, $-$33& 26.0& 8.1& 0.10& $<$0\\ 34816& 540& 215, $-$26& 40.0& 6.0, 8.0& 0.10& $<$0 \\ 37742& 500& 206, $-$16& 15.0& 9.5& 0.03& 5\\ 38666& 300& 237, $-$27& 21.5& none& 0.02& 5 \\ 41335& 300& 213, $-$13& 3.0& 1.0, 10.0& 0.04& $>$0 \\ 42087& 1200& 188, 2& 42.0, 32.0& 4.4, 6.5, 12.4& 0.13& $-$5 \\ 93521& 2000& 183, 62& 27.0& none& 0.11& $<$0 \\ 119608& 3400& 320, 43& 16.0& $-$5.4& 0.09& $-$20 \\ 141637& 170& 346, 21& 11.0& 0.5, $-$0.2& 0.28& $\sim$0\\ 148184& 150& 358, 21& 0.82& 3.4& 0.26& $\sim$0\\ 156110& 720& 71, 36& $<$10.0& 2.2& 0.18& 5 \\ 159176& 1180& 356, 0& 5.3& $-$20.8, $-$74.0& 0.28& $\sim$0 \\ 166937& 1200& 10, $-$2& 5.0, 7.0& 34.6, 47.2, 35.7& 0.39& 5 \\ 175754& 680& 16, $-$10& $<$10.0& 6.8& 0.09& 10 \\ 199478& 2000& 88, 1& 8.0& 5.0 to $-$75.0& 0.21& $-$7 \\ 205637& 250& 32, $-$45& 18.4& none& 0.30& 5 \\ 212978& 520& 95, $-$15& 12.0& 0.3, $-$12.3& 0.04& $-$5 \\ 214680& 780& 97, $-$17& 12.0& $-$4.8, 1.4& 0.28& $-$5 \\ 220172& 750& 68, $-$63& 2.9& none& 0.33& 10 \\ \hline \end{tabular} \caption[Summary of HI absorption.] {Summary of HI absorption: Column 1 gives the HD number of the star. Columns 2 and 3 show the distance to the star (as listed in the Sky Catalogue 2000.0, Vol 1, Hirshfield and Sinnott 1982) and its galactic coordinates, respectively. Column 4 gives the angular separation of the star from the radio source(s) towards which absorption has been detected. Column 5 lists the HI absorption velocities (LSR). Columns 6 shows the detection limit in $\tau$ (towards the strongest source in the field). Column 7 shows the approximate radial component of the Galactic rotation velocity at the distance to the star. $<$ 0 and $>$ 0 are used to indicate velocities with magnitude less than 5 \kms.\\ {\footnotesize The stars 14134 and 14143 are in the {\em same} field referred to above as 14143$^*$.}} \end{center} \end{table} \renewcommand{\arraystretch}{1} \renewcommand{\arraystretch}{.9} \begin{table} \begin{center} \begin{tabular}{|l|l|l|l|c|}\hline \setcounter{table}{1} Field & V$_{lsr}$(optical, all) & V$_{lsr}$(HI, coincident)& V$_{Gal}$&Ref \\ \hline\hline {} & \kms & \kms & \kms & \\ \hline 14143$^*$& $-$62.3, $-$46.8, $-$6.3& $-$46.1, $-$50.3, $-$11.2& $-$25 & M \\ & $-$65.3, $-$50.8, $-$10.3& & & \\ 14818& $-$42.6, $-$33.6, $-$6.6& $-$3.7& $-$25 & M \\ 21278& $-$0.2, 48.6& 2.8& $>$0 & A \\ 21291& $-$34.0, $-$7.5& $-$31.2, $-$7.0, $-$4.9& $-$10 & M \\ 24912& 4.7, 20.7& 4.3& $-$5& A\\ 25558& 10.1, 19.0& 8.1& $<$0 & A \\ 34816& $-$14.0, 4.14& 6.0& $<$0 & A \\ 37742& $-$21.0, 3.6& none & 5 & A \\ 38666& 1.0, 20.2& no absorption& 5 & MZ \\ 41335& $-$20.8, 0.2& 1.0& $>$0 & A \\ 42087& $-$37.7, $-$4.8, 10.2& 12.4& $-$5 & A \\ 93521& $-$55.0, $-$34.0, $-$10.3, 6.8& no absorption& $<$ 0 & MZ \\ 119608& 1.3, 22.4& none & $-$20 & MZ \\ 141637& $-$22.0, 0.0& 0.5, $-$0.2& 0 & B \\ 148184& 14.2, 2.2& 3.4& 0 & A \\ 156110& $-$19.7, 0.4& 2.2& 5 & MZ \\ 159176& 3.5, $-$22.5& $-$20.8& 0 & A \\ 166937& $-$5.5, 5.9, 25.3, 41.1& 5.4, 47.2, 35.7& 5 & A \\ 175754& $-$73.0, 5.9, 29.5& 6.8& 10 & A \\ 199478& $-$2.1, 8.7, 42.3, 61.2& 3.8 (blend) & $-$7 & A \\ 205637& $-$13.9, 1.8& no absorption& 5 & A \\ 212978& $-$73.0, 0.6& 0.3& $-$5 & A \\ 214680& $-$23.7, $-$14.7, 0.1& 1.4& $-$5 & A \\ 220172& $-$21.5, $-$0.8, 13.5& no absorption& 10 & MZ \\\hline \end{tabular} \caption[Summary of coincident velocities.]{Summary of coincident velocities: Column 2 lists the LSR velocities of all the optical absorption features seen in each line of sight (most of the listed velocities are from CaII observations). Velocities of the matching HI absorption features are given in column 3. Column 4 shows the radial component of the Galactic rotation velocity at the distance to the star. Column 5 has the reference for the optical absorption velocities which are from A: Adams (1949), B: Buscombe and Kennedy (1962), M: M\"{u}nch (1957) and MZ: M\"{u}nch and Zirrin (1961). } \end{center} \end{table} \section{Results} Table 1 lists the details of all the fields observed. In most of these fields there were several radio sources within the primary beam in addition to the source initially targeted. We have obtained spectra towards these sources as well if they turned out to be strong enough to detect reasonable optical depths ($\sim$ 0.1). Column 1 gives the HD number of the star. Columns 2 and 3 give the distance to the star (as listed in the SKY 2000 catalogue of bright stars) and its galactic coordinates, respectively. Column 4 gives the angular separation of the star from the radio source(s) towards which absorption has been detected. Column 5 lists the HI absorption velocities (LSR).These velocities have been derived by fitting Gaussian profiles to the absorption features. Columns 6 shows the detection limit in $\tau$ (towards the strongest source in the field). Column 7 shows the radial component of the Galactic rotation velocity at the distance to the star. The optical absorption velocities are likely to suffer from blending of features due to lack of resolution. A discussion of this and associated problems can be found in Welty, Morton and Hobbs (1996). Moreover, the correction for the solar motion adopted by different authors may differ leading to errors of $\sim$ 1 \kms (see for example Blaauw, 1952). In some cases, we have smoothed the absorption spectra to a resolution of 5 \kms to facilitate the convergence of the gaussian fit. Narrow absorption features are known to have widths less than this (Crovisier 1981). We have checked the unsmoothed data to ensure that there are no serious effects of blending in the estimates for velocities for the features discussed below. However, crowding in velocities inevitably causes some of the absorption widths to be suspect. The formal error in the fitting procedure for our HI absorption velocities is in most cases $\sim$ 1 \kms or less. The errors add to \lsim 2 \kms. However, it must be stated that the blending of features in both optical and radio could easily lead to larger errors than this formal value. {\it Hence we consider a feature in the optical spectrum as ``coincident'' with one seen in HI absorption or emission if the magnitude of the difference in velocities is less than 3 \kms}. This is roughly the same criterion adopted by Habing (1969) and Howard et al. (1963) for comparing HI emission profiles with optical absorption. Table 2 lists the ``matching'' features \ie HI absorption features whose velocities agree with the velocity of the optical absorption feature. Column 2 shows the LSR velocities of all the optical absorption features seen in each line of sight (most of the listed velocities are from CaII observations). Velocities of the matching HI absorption features follow in column 3. Column 4 shows the radial component of the Galactic rotation velocity at the distance to the star to facilitate comparison with the optical absorption velocities. Column 5 has the reference for the optical absorption velocities which are from Adams (1949), M\"{u}nch (1957) and M\"{u}nch and Zirrin (1961). We have used later compilations of these by Takakubo (1967), Siluk and Silk (1974), and Habing (1969). We present the HI optical depth towards all the observed fields at the end of the paper. The coordinates of the radio source (epoch 1950) towards which the spectrum is obtained is labelled at the top right of each panel. The HD number of the star towards which the corresponding optical spectrum is obtained is at the top left. The star co-ordinates (epoch 1950) are given immediately below this. The star co-ordinates have not been repeated if there are several radio sources in the same field. In all the spectra the velocities at which optical absorption is seen is indicated by arrows on the velocity axis (top). The conclusions we draw from these observations and a detailed discussion of the implications for the models of the ISM are presented in paper II. \section*{Acknowledgement} We wish to thank Adriaan Blaauw for extensive discussions which led to this investigation. His continued interest and critical comments have been invaluable to us. We are also indebted to Hugo van Woerden for guiding us through the literature pertaining to the early HI observations by the Dutch group.
\section{Introduction} According to the hierarchical clustering scenario, clusters of galaxies are formed through subcluster mergers. The influence of cluster merger on the intracluster medium (ICM) has been investigated in detail through the comparison between hydrodynamic/N-body simulations and X-ray observations. The simulations predict that the collision of large clusters gives rise to distorted X-ray contours and high temperature gas (e.g. Schindler, M\"{u}ller 1993; Burns et al. 1994; Ishizaka, Mineshige 1996; Roettiger et al. 1997; Ricker 1998; Takizawa 1999). Recent X-ray observations have confirmed that many clusters have the complex structure predicted by the simulations (e.g. Fujita et al. 1996; Honda et al. 1996; Knopp et al. 1996; Donnelly et al. 1998; Markevitch et al. 1999). On the other hand, the influence of merger on the galaxies in the clusters is not understood. Since cluster merger drastically changes the environment of the galaxies, especially static and ram-pressure on the galaxies, in a short time ($\ltsim 10^9$ yr), we can expect that it causes observable change in star formation rate (SFR) of the galaxies. However, it is not obvious whether cluster merger increases or decreases SFR of the galaxies as follows. Cluster merger rapidly raises the static and ram-pressure from ICM. As a result, interstellar medium (ISM) of the galaxies is expected to be compressed and star formation activity may be triggered (e.g. Evrard. 1991; Wang et al. 1997). In fact, in several merging clusters, there are galaxies having the abnormal spectrum which reflects recent star formation (e.g. Caldwell et al. 1993; Wang et al. 1997), although Tomita et al. (1996) find that this is not the case for a merging cluster A168. On the contrary, cluster merger may reduce SFR of the galaxies because ram-pressure strips their ISM away. Thus, in order to investigate the effect of pressure on galaxies, SFR of galaxies must be quantitatively estimated. Using the simple model of molecular cloud evolution, Fujita, Nagashima (1999) have quantitatively estimated the SFR of a disk galaxy under the pressure from ICM. However they consider only a radially infalling galaxy; they do not predict the evolutions of all galaxies in the cluster. In this letter, we investigate the evolution of SFR of disk galaxies when two clusters collide and merge. Moreover, we calculate the color distribution of the galaxies in the clusters. We only consider the effect of pressure from ICM; we do not consider the effect of tidal force from cluster potential and other galaxies for simplicity, although it may cause starburst (Bekki 1999). This is because it is difficult to calculate the intermittent influence of tidal force on the internal structures of hundreds of galaxies. \section{Models} We consider the merger of two typical clusters. In order to calculate the evolution of ICM, we use the smoothed-particle hydrodynamics (SPH) method (Monaghan 1992). We treat ions and electrons separately based on the model of Takizawa (1999), although the two-temperature nature does not affect the SFR of galaxies significantly. Collisionless particles corresponding to dark matter and galaxies are also considered. The initial conditions for ICM and collisionless particles are the same as those of Run B in Takizawa (1999) except for the radii of the two clusters, $r_{\rm out}$. Since we assume that $r_{\rm out}$ is ten times the core radius, which is two times larger than that in Takizawa (1999), the total masses of the two clusters are also larger. Their masses are $8\times 10^{14}$ and $2\times 10^{14}\rm\; M_{\odot}$. The gas fraction of the clusters is 0.1, which is supported by recent observations if $H_0\sim 75\rm\: km\; s^{-1} Mpc^{-1}$ (e.g. Ettori, Fabian 1999). The results in the next section are not sensitive to the fraction within the range of recent observational results. At $t=0$, the separation of the two clusters is 3.3 Mpc. We randomly pick out 125 particles from the collisionless particles (100 for the larger cluster and 25 for the smaller cluster) as disk galaxies. We calculate the orbits of these `galaxies' and the pressure from the surrounding ICM. The effects of static and ram-pressure on the SFR of disk galaxies are estimated by the model of Fujita (1998) and Fujita, Nagashima (1999). In this model, the SFR is derived by calculating the evolution of each molecular cloud using the relations for a vilialized cloud. We think that the model is superior to the approach based on the Schmidt law (Schmidt 1959), which has dominated in this field. This is because while the latter gives the same SFR regardless of the pressure of ISM for a fixed density, the former does not. Moreover the latter does not discriminate between HI gas and molecular clouds. The model adopted here treats them separately, although we sometimes call them together interstellar medium (ISM). Note that the SFR derived through the model adopted here is less sensitive to pressure variation compared to that through the model based on Schmidt law in which the density of ISM is assumed to be proportional to the pressure. The condition of ram-pressure stripping is \begin{eqnarray} \label{eq:strip} &\rho_{\rm ICM}v_{\rm rel}^2 \nonumber\\ >& 2\pi G \Sigma_{\star} \Sigma_{\rm HI} \nonumber\\ =& v_{\rm rot}^2 R^{-1} \Sigma_{\rm HI} \label{eq:grav2} \nonumber\\ =& 2.1\times 10^{-11}{\rm dyn\: cm^{-2}} \left(\frac{v_{\rm rot}}{220\rm\; km\: s^{-1}}\right)^2 \nonumber\\ & \times \left(\frac{R}{10\rm\; kpc}\right)^{-1} \left(\frac{\Sigma_{\rm HI}}{8\times 10^{20} m_{\rm H}\;\rm cm^ {-2}}\right) \label{eq:grav3}\:, \end{eqnarray} where $\rho_{\rm ICM}$ is the density of ICM, $v_{\rm rel}$ is the velocity of a galaxy relative to the surrounding ICM, $G$ is the gravitational constant, $\Sigma_{\star}$ is the gravitational surface mass density, $\Sigma_{\rm HI}$ is the surface density of the HI gas, $v_{\rm rot}$ is the rotation velocity, and $R$ is the characteristic radius of the galaxy (Gunn, Gott 1972; Fujita, Nagashima 1999). Abadi et al. (1999) numerically confirm that this analytic relation provides a good approximation. After this condition is satisfied, the formation of molecular cloud is assumed to stop; the gas ejected from stars and supplied from destroyed molecular clouds {\it directly} flows into ICM. Note that ISM in the central region of a galaxy ($\ltsim 2$ kpc) is not stripped because of large gravitational force. However, the mass is generally far smaller than the total mass of ISM (e.g. Struck-Marcell 1991). Thus, we ignore its contribution to the star formation activity of the galaxy. If a stripped galaxy reaches the outer part of the cluster, the galaxy may recover ISM. Since Fujita, Nagashima (1999) do not take account of recovery of ISM, we adopt the condition of recovery, \begin{eqnarray} \label{eq:ret} &\rho_{\rm ICM}v_{\rm rel}^2 \nonumber\\ < & \frac{16}{v_{\rm rel}} \frac{S}{\pi R^2}v_{\rm rot}^2 \nonumber\\ = & 2.0\times 10^{-11} {\rm dyn\; cm^{-2}} \left(\frac{v_{\rm vel}}{500\rm\; km\; s^{-1}}\right)^{-1} \nonumber\\ & \times \left(\frac{S}{6\rm\; M_{\odot}\;yr^{-1}}\right) \left(\frac{R}{10\rm\; kpc}\right)^{-2} \left(\frac{v_{\rm rot}}{220\rm\; km\; s^{-1}}\right)^{2} \:, \end{eqnarray} where $S$ is the gas supply from stars and destroyed molecular clouds (Takeda et al. 1984). Although this analytic expression assumes spherical symmetry and may not be exact in the case of a disk galaxy, the following results do not significantly alter even if we change the coefficient of the right hand of relation (\ref{eq:ret}) by a factor of five. After the galaxy satisfies this condition, gas ejected from stars is trapped in the potential well of the galaxy, and molecular cloud formation is resumed. We start to calculate the SFR of the model galaxies at $t=1$ Gyr. The initial mass of molecular gas and HI column density are $2.5\times 10^9\rm\; M_{\odot}$ and $8\times 10^{20} m_{\rm H}\;\rm cm^ {-2}$, respectively. Moreover, we take $S=6\rm\; M_{\odot}\:yr^{-1}$, $R=10$ kpc, and $v_{\rm rot}=220\rm\;km\; s^{-1}$. Although the parameters are the typical ones for our Galaxy (e.g. Binney, Tremaine 1987), the following results do not change significantly even if we take the ones for typical galaxies whose masses are five times smaller. We have confirmed that the evolutions of the SFR and ISM are not sensitive to the initial time of the calculation $\gtsim 1$ Gyr after the calculation starts. Using the obtained SFR and the population synthesis code made by Kodama, Arimoto (1997), we also investigate the evolution of color of galaxies. \section{Results} \label{sec:results} Figure 1a shows the X-ray contours and positions of galaxies at $t=3.6$ Gyr, where the two clusters have just passed each other. The origin of the figure is the center of gravity of the clusters. We define a `post-starburst galaxy' (PSB) as the galaxy whose SFR reduces to less than $1/3$ of that for $10^8 - 10^9$ yr before the observation time. Figure~1a shows that red ($B-V>0.7$) galaxies and PSBs are concentrated in the central region of the cluster. These features are always seen during merger. Although several blue ($B-V<0.7$) galaxies are also seen in the central region, they do not gather at a specific position. We present the velocity distributions of blue and red galaxies within $0.5$ Mpc from the center of the merged cluster at $t=3.6$ Gyr in figure~2. Since the average velocity of blue galaxies in this region is $\sim 2500\rm\; km\; s^{-1}$, they pass each other simultaneously. Note that in figure 2 the average velocity of red galaxies is $\sim 1800\rm\; km\; s^{-1}$. This reflects that the blue galaxies at the central region of the cluster are the ones that can reach the cluster center before they become red although their ISM is stripped. In figure 1b, we present the state of the cluster after it is nearly relaxed ($t=5$ Gyr). Segregation of blue and red galaxies is noticeable. This is quantitatively shown in figure~3. The number of blue galaxies in the central region of cluster ($\ltsim 0.7$ Mpc) at $t=5$ Gyr is smaller than that at $t=3.6$ Gyr. This is because there are few galaxies rapidly infalling into the center of the merged cluster at $t=5$ Gyr. To see the evolution of galaxies in detail, we show the fraction of blue galaxies and PSBs in figure~4. The evolution of the latter is calculated only for $t>3$ Gyr to avoid the influence of initial conditions. The median static and ram-pressures on galaxies are shown for comparison. They temporarily increase at $t\sim 3.6$ Gyr when two clusters collide. At that time, the HI gas of most galaxies is stripped because of the increase of ram-pressure. After that, new molecular clouds are not produced; the existing ones disappear within $\sim 10^8$ yr because of consumption by star formation and destruction by young stars (see Fujita, Nagashima 1999). Since molecular clouds are used to make stars, the SFR of galaxies and fraction of blue galaxy decrease as molecular clouds disappear. Although the static and ram-pressure compress ISM of galaxies and trigger the star formation activity before the stripping occurs, the duration of activity is short ($\ltsim 0.4$ Gyr); the activity does not affect the color distribution of galaxies in clusters significantly. Thus, cluster merger does not trigger, but weakens star formation activity of the galaxies. At $t\sim 4$ Gyr, the two clusters once come apart. The ICM of the two clusters expands and their relative velocity reduces. Since the average ram-pressure significantly decreases, the ISM of galaxies recovers and star formation restarts. Thus, the fraction of blue galaxies returns to the initial value ($\sim 60$\%). Note that the fraction of blue galaxies is over 40\% even when clusters are colliding (figure~1). This means that cluster merger does not significantly affect galaxies in the outer region of the clusters. Figure~4 clearly shows that the fraction of PSBs increases from 30\% to 60\%, immediately following the merger. These PSBs are the galaxies whose SFR decreases because of ram-pressure stripping. In fact, the fraction of PSBs begins to increase after that of blue galaxies decreases. However, it may not be easy to use PSBs as the probe of cluster merger, because they always exist in clusters. This reflects that some of cluster galaxies have radial orbits and fall into the center of the cluster regardless of cluster merger. Thus, in order to know the relation between PSBs and cluster merger observationally, it is required to compare the PSB fractions of merging clusters with those of non-merging clusters statistically. \section{Summary} \label{sec:summary} We have investigated the effect of pressure on the galaxies when two clusters merge. We find that because of ram-pressure stripping, star formation rate of most of galaxy decreases during merger contrary to the speculation of Evrard (1991) and Wang et al. (1997). Some blue galaxies can reach the central region of the merging clusters before they become red because of their high velocities. By observing velocities, these galaxies would be distinguished from the blue galaxies in which star formation is triggered by tidal force from the gravitational field of the cluster because the tidal interaction is effective when a galaxy moves slowly. Following the decrease of blue galaxy fraction of the clusters, the fraction of post-starburst galaxies increases. After the two clusters pass by, star formation restarts in the galaxies because the ram-pressures decrease. When a quasi-equilibrium state is reached, the segregation of blue and red galaxies becomes prominent. \par \vspace{1pc} \par This work was supported in part by the JSPS Research Fellowship for Young Scientists. \section*{References} \small \re Abadi M.G., Moore B., Bower R.G.\ 1999, MNRAS in press (astro-ph/9903436) \re Bekki K.\ 1999, ApJL 510, L15 \re Binney, J., Tremaine, S. 1987, Galactic Dynamics \re Burns J.O., Roettiger K., Ledlow M., Klypin A.\ 1994, ApJL 427, L87 \re Caldwell H., Rose J.A., Sharples R.M., Ellis R.S., Bower R.G.\ 1993, AJ 106, 473 \re Donnelly R.H., Markevitch M., Forman W., Jones C., David L.P., Churazov E., Gilfanov M.\ 1998, ApJ 500, 138 \re Ettori S., Fabian A.C.\ 1999, MNRAS in press (astro-ph/9901304) \re Evrard A.E.\ 1991, MNRAS 248, 8p \re Fujita Y.\ 1998, ApJ 509, 587 \re Fujita Y., Koyama K., Tsuru T., Matsumoto H.\ 1996, PASJ 48, 191 \re Fujita Y., Nagashima M., 1999, ApJ in press (astro-ph/9812378) \re Gunn J.E., Gott J.R. 1972, ApJ 176, 1 \re Honda H., Hirayama M., Watanabe M., Kunieda H., Tawara Y., Yamashita K., Ohashi T., Hughs, J.P., Henry, J.P.\ 1996, ApJL 473, L71 \re Ishizaka C., Mineshige S.\ 1996, PASJ 48, L37 \re Knopp G.P., Henry J.P., Briel U.G.\ 1996, ApJ 472, 125 \re Kodama T., Arimoto N.\ 1997, A\&A 320, 41 \re Markevitch M., Sarazin C.L., Vikhlinin A.\ 1999, ApJ, in press (astro-ph/9812005) \re Monaghan J.J.\ 1992, ARA\&A 30, 543 \re Ricker P.M.\ 1998, ApJ 496, 670 \re Roettiger K., Loken C., Burns J.O.\ 1997, ApJS 109, 307 \re Schindler S., M\"{u}ller E.\ 1993, A\&A 272, 137 \re Schmidt M.\ 1959, ApJ 344, 685 \re Struck-Marcell 1991, ApJ, 368, 348 \re Takeda H., Nulsen P.E.J., Fabian A.C.\ 1984, MNRAS 208, 261 \re Takizawa M.\ 1999, ApJ in press (astro-ph/9901314) \re Tomita A., Nakamura F.E., Takata T., Nakanishi K., Takeuchi T., Ohta K., Yamada T.\ 1996, AJ 111, 42 \re Wang Q.D., Ulmer M.P., Lavery R.J.\ 1997, MNRAS 288, 702 \label{last} \newpage \begin{figure} \centering \epsfig{figure=f1a.eps, width=8cm} \caption{(a) The X-ray surface brightness and the positions of galaxies at $t=3.6$ Gyr. The crosses and filled circles indicate red ($B-V>0.7$) and blue ($B-V<0.7$) galaxies, respectively. Open circles indicate post-starburst galaxies.} \end{figure} \begin{figure} \centering \epsfig{figure=f1b.eps, width=8cm} \vspace{3mm} \footnotesize{Fig. 1.. (b) Same as in figure 1 but for $t=5$ Gyr.} \end{figure} \begin{figure} \centering \epsfig{figure=f2.eps, width=8cm} \caption{ Histogram showing the distribution of velocity relative to the space coordinate at $t=3.6$ Gyr. The galaxies within 0.5 Mpc from the origin are picked out. } \end{figure} \begin{figure} \centering \epsfig{figure=f3.eps, width=8cm} \caption{ Histogram showing the distribution of distance from the origin.} \end{figure} \begin{figure}[t] \centering \epsfig{figure=f4.eps, width=8cm} \caption{ The fraction of blue galaxies $f_b$ and post-starburst galaxies $f_{\rm PSB}$. The median static pressure $\langle P_{\rm stat} \rangle$ and ram-pressure $\langle P_{\rm ram} \rangle$ are also presented.} \end{figure} \end{document}
\section{Introduction} String theory is a promising candidate of quantum gravity and the theory of spacetime. It is therefore an interesting question to study the properties of spacetime in string theory. A number of uncertainty relations have been proposed in relation to string theory. See, for example, \cite{douglas,liyon1} for a review of the subject and for further references. It was first proposed \cite{early} that a modified version of the canonical Heisenberg uncertainty relation \begin{equation} \delta X\delta P\geq 1+l_s^2\delta P^2, \end{equation} governs the high energy behavior of string theory. This relation implies the existence of a minimal scale $\delta X\sim l_s$ \cite{early}. Another uncertainty relation is the one proposed by Yoneya \cite{Yon} \begin{equation} \label{TX} \delta T\delta X\geq l_s^2. \end{equation} He further suggested that the spacetime uncertainty relation can be one of the fundamental principles underlying the nonperturbative string theory, and can also be one of the guiding principles for constructing a covariant formulation of M theory \cite{Yon}. In this paper, we propose a new kind of uncertainty relation \begin{equation} \label{ours} \delta X^0\delta X^1\cdots\delta X^p\geq g_s l_s^{p+1} \end{equation} for the worldvolume of a D$p$-brane. Here $X^i$ are the D-brane worldvolume coordinates. We also propose a similar kind of uncertainty relations for the M-branes. It is well known that in string theory different probes could see slightly different spacetime geometries. So it is natural to expect that the uncertainty principle can be different for different probes, as it is manifest in our proposed relations (\ref{ours}). Note that (\ref{ours}) is involved only with the longitudinal directions on a D-branes, that is, the D-brane worldvolume. Another uncertainty relation involving both longitudinal and transverse directions for a D-brane was proposed by Yoneya and collaborators \cite{liyon1,Yon,yoneya}. He proposed that if one interprets $T$ and $X$ as the longitudinal and transverse coordinates of a D-brane, (\ref{TX}) can also be understood as an uncertainty relation for a D-brane. In the brane world scenario \cite{bw}, it is argued that we could be living on a D-brane. It is thus of great interest to determine the form of uncertainty relations on a D-brane worldvolume, which will be interpreted as the uncertainty relation for the four dimensional spacetime on which we are trapped. Since it is also known that a D-brane worldvolume becomes noncommutative \cite{CDS,DH} in the presence of a nontrivial background $B_{NS-NS}$ field in the direction of the brane, it is natural to ask what kind of uncertainty relation follows this worldvolume noncommutativity \cite{SJ,CH}. In this paper we first show how to obtain the uncertainty relation for a D1-brane by integrating out quantum fluctuations of the background field. Then we use various dualities in string theory and M theory to extend the uncertainty relation to other D-branes and strings or membranes. Putting all the old and new dualities together, we find a web of spacetime uncertainty relations associated with all the branes in string theory or M theory. The organization of this paper is as follow. In sec.\ref{open}, we first extend the previous results \cite{CDS,DH,CH} about the noncommutativity of D-brane by generalizing the background to the first nontrivial (yet manageable) order, with a curved background metric $g_{\mu \nu}$ and a nontrivial NS-NS 2-form field. Then we argue in sec.\ref{noncomm} that these commutation relations give rise to uncertainty relations on the D-brane worldvolume. While the precise form of the uncertainty relations cannot be easily fixed this way, one can use the dualities of string theory and M theory to constrain the form of the uncertainty relations. Guided by this idea, we propose in sec.\ref{ur} worldvolume uncertainty relations for D-branes, M2-branes, M5-branes, as well as fundamental strings. We check that they are consistent with various dualities of string theory and M theory. Finally we make a few remarks in sec.\ref{remarks}. \section{Noncommutativity from open string quantization}\label{open} It was first proposed by Connes, Douglas and Schwarz \cite{CDS}, and by Douglas and Hull \cite{DH}, that the Matrix model compactified on a torus with the NS-NS $B$ field background should be described by a field theory living on a noncommutative space. This conjecture has been verified in various ways (see, e.g. \cite{NCT}). In particular, it was pointed out in \cite{AAS} that this can be seen by quantizing an open string ending on a D-brane. Based on the intimate relation between compactified Matrix models and D-brane worldvolume theories, Hauffman and Verlinde \cite{HV} also suggested that the low energy effective theory for a D-brane compactified on a torus with $B$ field background should also live on a noncommutative space. Following \cite{AAS}, the quantization of the open string ending on a D-brane with a constant ${\cal F}=B-F$ field background in flat spacetime was carried out in \cite{CH}. An agreement with previous works was obtained but the framework and results of \cite{CH} were more general. In this section we generalize further the previous work to a curved background and non-constant ${\cal F}$ field. The derivation is in parallel with \cite{CH}. For an open string with an end-point sticking to a D$p$-brane with $U(1)$ field strength $F$ in a $B$-field background, the bosonic part of its action can be written as \begin{equation} \label{action} S_B= {1 \over 4\pi\alpha'} \int_{\Sigma} d^2\sigma \left[ \eta^{\alpha\beta}g_{\mu\nu} \partial_\alpha X^{\mu}\partial_\beta X^{\nu}+ \epsilon}\def\vare{\varepsilon^{\alpha\beta} {\cal F}_{\mu\nu}\partial_\alpha X^{\mu}\partial_\beta X^{\nu} \right], \end{equation} where \begin{equation} {\cal F}=B-F \end{equation} is the modified Born-Infeld field strength on the D-brane, $F=dA$ and $A_i,\ i=0,1,\cdots, p$, is the $U(1)$ gauge potential living on the D$p$-brane. We use the convention $\eta^{\alpha\beta}=\mbox{diag}(-1,1)$ and $\epsilon^{01}=1$ as in \cite{CH}. Note that ${\cal F}$ is invariant under the gauge transformations \begin{equation} A\rightarrow A+d\Lambda, \quad B\rightarrow B, \end{equation} and \begin{equation} B\rightarrow B+d\Lambda, \quad A\rightarrow A+\Lambda. \end{equation} The equations of motion are \begin{equation} \label{EOM} g_{\mu \kappa} (\ddot{X}^\kappa + \Gamma^\kappa{}_{\mu \nu} \dot{X^\mu} \dot{X^\nu}) - g_{\mu \kappa} ({X''}^\kappa + \Gamma^\kappa{}_{\mu \nu} {X'}^\mu {X'}^\nu) + H_{\mu\nu\lambda}\def\L{\Lambda}\dot{X}^{\nu}{X'}^{\lambda}\def\L{\Lambda}=0, \end{equation} where \begin{equation} {\Gamma^{\lambda}\def\L{\Lambda}}_{\mu\nu}=\frac{1}{2}g^{\lambda}\def\L{\Lambda\kappa} (\partial_{\mu}g_{\kappa\nu}+\partial_{\nu}g_{\mu\kappa}-\partial_{\kappa}g_{\mu\nu}) \end{equation} is the Riemannian connection for the metric and \begin{equation} H_{\mu\nu\lambda}\def\L{\Lambda}=\partial_{\mu}{\cal F}_{\nu\lambda}\def\L{\Lambda}+\partial_{\nu}{\cal F}_{\lambda}\def\L{\Lambda\mu}+ \partial_{\lambda}\def\L{\Lambda}{\cal F}_{\mu\nu}, \end{equation} that is, $H=d{\cal F}$. In fact, since $dF=0$, $H=dB$. The boundary conditions are \begin{equation} \label{BC} X^{'\nu} g_{\nu i} +\dot{X}^{\nu} {\cal F}_{\nu i} =0 \quad \mbox{for} \;\; i=0,1,\cdots,p, \end{equation} \begin{equation} X^a=x_0^a \quad \mbox{for} \;\; a=p+1,\cdots,9, \end{equation} at $\sigma=0,\pi$, where $x_0^a$ gives the position of the D$p$-brane. For convenience we have chosen the spacetime coordinates in such a way that $x_0^a$ are constant on the D$p$-brane. We will use the indices $(i,j,\cdots)$ for directions parallel to the brane ($i=0,1,\cdots,p$) and the indices $(a,b,\cdots)$ for directions transverse to the brane ($a=p+1,\cdots,9$). We will choose the coordinates such that $g_{ia}=0$ on the D-brane for convenience. The momentum density is \begin{equation} 2 \pi \alpha' P_{\mu}= \dot{X}^{\nu}g_{\nu\mu} +X^{'\nu}{\cal F}_{\nu\mu} . \end{equation} For a constant background (${\cal F}=$constant) in a flat spacetime ($g_{\mu\nu}=\eta_{\mu\nu}$), one can solve the equations of motion and boundary conditions exactly and carry out the canonical quantization \cite{CH}. The final result is \begin{equation} \label{XX1} [X^i,X^j]= \pm 2\pi i\alpha'(M^{-1}{\cal F})^{ij}, \end{equation} where \begin{equation} \label{M} M_{ij}=\eta_{ij}-{\cal F}_i{}^k{\cal F}_{kj}, \end{equation} and $X^i$ is the spacetime coordinates of the open string at the two end-points $\sigma=0,\pi$. The indices are lowered or raised by the spacetime metric $\eta$. While this result does not need a compactification, in case the spacetime is compactified on a torus, the right hand side of (\ref{XX1}) for $F=0$ is proportional to the dual $B$ field on the dual torus. It agrees with the results for the BFSS matrix model \cite{CDS}. For the case of a non-constant ${\cal F}$ on a flat torus and its relation to deformation quantization, see \cite{GC}. For a generic background, we are unable to find the most general solution in parallel of \eq{XX1}. However it is possible to find a special solution in a certain approximation of weak field and slow variation. Consider a deviation from the trivial background ${\cal F}=0$, $g =\eta$ with small ${\cal F}$ and $\partial g$ considered as first order quantities. We will consider $\partial \partial g$ as second order and so from the Einstein equation for the background, $\partial {\cal F} \sim \partial g$ are also of the first order. In summary, we will consider $(x, \dot{x}, g)$ as terms of the 0th order, $({\cal F}, \partial{\cal F}, \partial g)$ as terms of the 1st order, and $( \partial^{n}{\cal F}, \partial^{n} g), n\geq 2$ as terms of higher orders. We will only keep terms of the 0th and 1st order consistently. To solve the equations of motion and boundary conditions, we use the following ansatz for $X$ \begin{equation} \label{ansatz} X^{i}=x^{i}(\tau)+y^{i}(\tau)\sigma, \quad X^{a}=x_0^a. \end{equation} It is easy to check that this ansatz for the $\sigma$ dependence is consistent with our approximation. Substituting \eq{ansatz} into the boundary conditions \eq{BC}, one finds that $y$ is of first order: \begin{equation} \label{BCy} y^{i}=-\dot{x}^j ({\cal F} g^{-1}(x))_j{}^i. \end{equation} For this ansatz, the equations of motion (\ref{EOM}) give the geodesic equation for $x^i$ \begin{equation} \ddot{x}^{i}+{\Gamma (x)^{i}}_{jk}\dot{x}^{j}\dot{x}^{k}=0, \end{equation} and the free motion for $y$ \begin{equation} \label{y} \ddot{y}^i=0. \end{equation} Eqn.(\ref{y}) is satisfied by (\ref{BCy}) up to 2nd order terms. The lowest energy modes on a string are thus approximately given by \begin{eqnarray} X^{i}&=&x^{i}+ (g^{-1} {\cal F} g^{-1}(x))^{ik} p_k \sigma, \nonumber \\ 2\pi \alpha' P_{i}&=& p_i - p_m p_n g^{mj} g^{nk} (\partial_j{\cal F}_{ki}(x)) \sigma, \label{XP} \end{eqnarray} and \begin{equation} X^a=x_0^a, \quad P_a=0, \end{equation} where we have introduced \begin{equation} p_k = \dot{x}^j g_{jk}(x). \end{equation} The Poisson bracket for the open string is determined by the symplectic two-form \footnote{ We don't need to do the time average prescription here as in \cite{CH}; and if we do it here we will obtain the same result. } \begin{equation} \label{Omusu} \Omega= \int_0^{\pi}d\sigma\delta P_{\mu}\delta X^{\mu}. \end{equation} Substituting \eq{XP} into (\ref{Omusu}), we find \begin{equation}\label{Om0} \Omega = \frac{1}{2\alpha'} \delta\bar{p}_{i} \delta\bar{x}^i. \end{equation} up to second order terms. Here $\bar{p}_i$ and $\bar{x}^i$ are defined as \begin{eqnarray} \bar{p}_i &=& p_i - \frac{\pi}{2} p_m p_n g^{mj} g^{nk} (\partial_j{\cal F}_{ki}(x)), \\ \bar{x}^i &=& x^i + \frac{\pi}{2} (g^{-1} {\cal F} g^{-1}(x))^{ik} p_k. \end{eqnarray} We thus obtain the commutation relations \footnote{ Normally, special attention has to be paid to the operator ordering when we derive commutation relations from Poisson brackets, so that the Jacobi identity is satisfied if possible. Fortunately, in the approximation we use here, the issue of operator ordering does not arise. } \begin{eqnarray} &[\bar{p}_i,\bar{p}_j]=0 , \quad [\bar{x}^i,\bar{x}^j]= 0, \label{xxpp}\\ &[\bar{x}^i,\bar{p}_j]=2 i\alpha' \delta^i_j, \label{xbarp} \end{eqnarray} or equivalently, in terms of $x^i$ and $p_j$, it is \begin{eqnarray} &[p_i,p_j]= 0, \label{pp} \\ &[x^i,x^j]= 2 \pi i \alpha' (g^{-1}{\cal F} g^{-1})^{ij}, \label{xx} \\ &[x^{i},p_{j}]=2 i\alpha' \delta^i_j + i \pi \alpha' p_k g^{km} g^{in} (\partial_m {\cal F}_{nj} + \partial_n {\cal F}_{mj} -\partial_j {\cal F}_{mn}). \label{xp} \end{eqnarray} All of these hold only up to 2nd order terms. It is easy to verify that the commutation relations (\ref{pp})-(\ref{xp}) satisfy the Jacobi identity also up to second order terms. It is now easy to check that for $\sigma =0, \pi$, \begin{equation} [X^{i},X^{j}]=\pm 2 \pi i \alpha' (g^{-1}{\cal F} g^{-1})^{ij}, \label{xx1} \end{equation} where the plus (minus) sign corresponds to the end-point $\sigma=0$ $(\pi)$. This agrees with the result for a constant background (\ref{XX1}) in the leading order. Furthermore, since the right hand side of (\ref{xx1}) is a tensor field, this equation is covariant under general coordinate transformations up to second order terms. In the static gauge $X^{i}$ is the worldvolume coordinate for the D-brane, so the D-brane worldvolume appears to be a noncommutative space. We propose that in the weak field and slow variation approximation we considered, \eq{xx1} gives the commutation relations for the D-brane spacetime coordinates in a generic background. Under an S-duality transformation, a D-string is turned into a fundamental string. For a D-string, the noncommutativity is governed by (\ref{xx1}) with the NS-NS background ${\cal F}= B - dA$. The commutation relation for a fundamental string in the dual theory is thus \begin{equation} \label{F1CR} [X^{i},X^{j}]=\pm 2 \pi i g_s' \alpha' (g^{-1}{\cal F}' g^{-1})^{ij}, \end{equation} where $g_s'= 1/g_s$ is the dual string coupling and ${\cal F}'= B'-dA'$ is the R-R counterpart of ${\cal F}$ in the dual theory. Since we assumed that $g_s$ is small to derive (\ref{xx1}), (\ref{F1CR}) is valid only for large $g_s'$. It would be interesting if one can derive this directly from string theory. \section{Noncommutative gauge theory and uncertainty relations} \label{noncomm} The noncommutativity \eq{xx1} signifies the existence of uncertainty relations on the brane. To give a precise formulation, we first need to explain what we mean by an uncertainty $\delta X$. The suitable framework for discussing uncertainty relation is to employ the language of string field theory \cite{ST}. Let $\Psi$ be the normalized wave function for the D-brane and define $(\Delta X^i)^2$ by \begin{equation} (\Delta X^i)^2=\int[DX(\xi)]\Psi^{\dagger}(X(\xi)) (X^i(\xi)-\overline{X^i} \; )^2\Psi(X(\xi)), \end{equation} where \begin{equation} \overline{X^i} =\int[DX(\xi)]\Psi^{\dagger}(X(\xi))X^i(\xi)\Psi(X(\xi)). \end{equation} In these equations, $\xi=(\xi_0,\cdots,\xi_p)$ denote the worldvolume coordinates of the D-brane. The part $[DX(\xi)]$ of the functional measure denotes an integration over all D-brane configurations. Obviously these definitions mimic their counterparts in the familiar case of the quantum mechanics for a point particle. Applying the standard argument for uncertainty principle to \eq{xx1} and notice that as a background, ${\cal F} $ is independent of the D-brane wavefunction $\Psi$, we get \begin{equation} \label{tt} \Delta X^i \Delta X^j \geq 2\pi l_s^2 |{\cal F}^{ij}|, \end{equation} where ${\cal F}^{ij}=(g^{-1}{\cal F} g^{-1})^{ij}$ and $l_s^2=\alpha'$. Note that these first quantized quantities depend on $\Psi$ and the classical backgrounds. The right hand side of \eq{tt} vanishes for a trivial classical background ${\cal F}=0$. However, as explained in \cite{CH}, even for such a classical background, there could be nontrivial uncertainty relations arising from integrating out the quantum fluctuations. On expanding the string field $\Psi$ into the component fields ${\cal B}$, the string field path integral becomes an infinite product of path integrals over these fields and the expectation value $\langle \cdot \rangle$ of an operator in string field theory is defined by \begin{equation} \langle {\cal O} \rangle= {1 \over Z} \int[D{\cal B}] e^{-S} \: {\cal O} \quad \mbox{where}\quad Z =\int[D{\cal B}] e^{-S}. \end{equation} For example, ${\cal B}$ includes the metric $g$ and the $B$-field. It is thus natural to define the desired uncertainty for $X^i$ as \begin{equation} \label{def} \delta X^i= \sqrt{\langle (\Delta X^i)^2 \rangle}. \end{equation} Using Schwarz inequality, we find \begin{equation} (\delta X^i)^2 (\delta X^j)^2 \geq | \langle \Delta X^i \Delta X^j \rangle |^2 . \end{equation} Thus \begin{equation} \delta X^i \delta X^j \geq 2\pi l_s^2 \langle \; |{\cal F}^{ij}| \; \rangle, \end{equation} and it generally reduces to something of the form \begin{equation} \label{uncer} \delta X^i \delta X^j \geq f(g_s) \: l_s^2, \end{equation} where $f(g_s)$ is some function of $g_s$. We will try to determine $f$ in the weak coupling limit. Notice that this form of uncertainty relation follows more or less from dimensional analysis. The point of the analysis performed above is to give a precise definition of the quantities involved and to show that the right hand side of (\ref{uncer}) is really nonvanishing. To see how the $g_s$ dependence comes in, it is instructive to consider the case of a D-string. The commutation relation for the worldsheet directions reads \begin{equation} \label{xx-flat} [X^0, X^1] = 2\pi i l_s^2 {{\cal F}}, \end{equation} where ${\cal F}={\cal F}_{01}$ and we have replaced $g_{ij}$ by the flat metric in our approximation. We thus need to evaluate \begin{equation} \frac{1}{Z}\int[D{\cal B}] e^{-S} |{\cal F}| . \end{equation} For small $g_s$, we can use the tree level SUGRA action where $S$ contains a piece \begin{equation} S= \frac{1}{g_s^2 l_s^8} \int dB*dB+ \cdots . \end{equation} It is convenient to go to a gauge in which ${\cal F}=B$ and by rescaling \begin{equation} B = g_s B', \end{equation} then for the background $B =0$ \begin{eqnarray} \langle \; |{\cal F}| \; \rangle &\simeq&\frac{1}{Z_B} \int[DB] e^{-\frac{1}{g_s^2 l_s^8} \int ( \partial B_{01})^2 } |B_{01}| \nonumber\\ &=& g_s \{\frac{1}{Z'_B} \int[DB']e^{-\frac{1}{l_s^8} \int (\partial B'_{01})^2} |B'_{01}| \}, \label{cF} \end{eqnarray} where \begin{equation} Z_B=\int[DB]e^{-\frac{1}{g_s^2 l_s^8}\int(\partial B_{01})^2}, \quad\quad Z'_B=\int[DB']e^{-\frac{1}{l_s^8}\int(\partial B'_{01})^2}. \end{equation} The above term $\{\cdot\}$ in (\ref{cF}) is independent of $g_s$. One can further scale $B'_{01}$ by $l_s^4$ so that (\ref{cF}) reads $\langle\;|{\cal F}|\;\rangle \simeq g_s l_s^4 I$, where $I$ is a path integral with no apparent dependence on $g_s$ or $l_s$. Because of the absolute sign in \eq{cF}, it is easy to show that $I$ is non-vanishing and is in fact divergent. A momentum cutoff at $\Lambda$ has to be introduced to make sense of $I$ and one finds $I \sim \Lambda^4$ and hence \begin{equation} \label{D1-0} \langle\;{\cal F}\;\rangle\simeq g_s l_s^4\Lambda^4. \end{equation} The natural cutoff here is $\Lambda\sim l_s^{-1}$ because we have ignored all stringy corrections of higher order in $\alpha'$ in the SUGRA action. While the derivation above is not completely rigorous, we consider the possibility of extending this result consistently to all other branes and strings via string dualities in the rest of this paper as a supporting evidence for (\ref{D1-0}). In principle, there could also be other sources contributing to the uncertainty other than the NS-NS 2-form gauge field. Here we considered only the massless mode of SUGRA. In an approximation better than \eq{xx-flat}, quantum fluctuations of the metric also contribute. One should also take into account string loop effects for a generic $g_s$. These can enter in at least two ways: First, eqn.(\ref{xx1}) was derived from a single string in the first order approximation; in general higher order terms and the string loop effects can modify the commutation relations for the D-brane worldvolume coordinates \cite{Schom}. Second, more precisely one should also use the string loop corrected SUGRA action instead of the tree level one in the above derivation. Including all these factors, we expect the uncertainty relation to take the generic form \footnote{ Uncertainty relations for the D-brane worldvolume has also been discussed in the context of Liouville string theory \cite{ellis}. There an uncertainty relation of a form similar to \eq{D1-1} was found, but with a different dependence ($\sqrt{g_s}$ instead of $g_s$) on the string coupling. } \begin{equation} \label{D1-1} \delta X^0 \delta X^1 \geq g_s l_s^2 +\cdots, \end{equation} up to a numerical factor which will be ignored in this paper; and the omitted terms are of higher order in $g_s$ and $\alpha'$ due to the above-mentioned higher order corrections. There could also be a dependence on $\Psi$ in the higher order corrections. Obviously one can perform the same derivation for a D$p$-brane with $p>1$ and find the same uncertainty relation between any two directions on the D-brane. At this point, one may ask a number of questions. For example, is it possible to determine explicitly the higher order corrections in \eq{D1-1}? How does the above generalize to the case of the other D$p$-brane? How does the uncertainty relation look like on a D$p$-brane? What we will do is to find new uncertainty relations by requiring the uncertainty relations to be consistent with the known dualities of string theory. This consistency requirement will be our guiding principle. \section{Worldvolume uncertainty relations}\label{ur} In this section, we will propose a form of the uncertainty relations for D$p$-branes which is consistent with dualities of string theory. To strengthen the starting ground for our argument which leads to the general result, we first consider the cases of D1 and D0-branes in more detail. \subsection{D1-branes} {}From sec.\ref{noncomm}, we find that the uncertainty relation for a D-string in the small $g_s$ limit in flat spacetime takes the form \begin{equation} \label{D1} \delta X^0 \delta X^1 \geq g_s l_s^2. \end{equation} Eqn.(\ref{D1}) gives a minimal area for the D1-brane worldsheet. An independent support for this result can be obtained by S-duality. Under S-duality, a D1-brane is interchanged with a fundamental string, and the string tension interchanged with the D1-brane tension. Thus the uncertainty relation for a fundamental string should be \begin{equation} \label{F1} \delta X^0 \delta X^1 \geq l_s^2. \end{equation} This is in fact what people have suggested based on properties of string scattering amplitudes; worldsheet conformal invariance and other various arguments \cite{Yon1}. This can also be heuristically argued as follows (first two references of \cite{Yon1}). According to the canonical uncertainty relation in quantum mechanics \begin{equation} \label{Et} \delta E\delta T\geq 1, \end{equation} (where $T$ should be identified with $X^0$,) if $\delta T$ is small, $\delta E$ will be large. Since $E\sim 1/\alpha'$ times string length, it is associated with a large uncertainty $\delta X^1$ in the string length \cite{DKPS}. (For this argument to be more rigorous, we need a virial theorem stating that a certain portion of the energy must be attributed to the potential energy due to string tension.) \subsection{D0-branes} \label{D0-brane} The usual uncertainty principle of quantum mechanics (\ref{Et}) implies that \begin{equation} \delta T\geq\frac{1}{\delta E}\geq\frac{1}{E}, \end{equation} where we assumed that $E\geq\delta E$. (This would be the case if $E$ is positive definite.) In the rest frame of the D0-brane, $E=1/(g_s l_s)$ is the mass of a D0-brane, so we find \begin{equation} \label{D0} \delta T\geq g_s l_s, \end{equation} where $T$ is the proper time for the D0-brane worldline. We can also interpret (\ref{D0}) as follows. If (\ref{D0}) is not satisfied, the energy of a D0-brane can be larger than its rest mass and can thus lead to pair productions of D0 and anti-D0-branes. One can then imagine a quantum path in which the created anti-D0-brane annihilates the original D0-brane so that the created D0-brane survives as the final D0-brane. In such cases the proper time is ill-defined during the process of creation and annihilation. In \cite{DKPS} it was explained that the short distance behavior of D-branes is described by the low energy physics of open strings ending on the D-branes. For D0-branes two different characteristic scales were found. The first scale is the Planck scale for the 11 dimensional SUGRA, which is $g_s^{1/3}l_s$ in accordance with the duality between M theory and type \mbox{\II A\hspace{.2em}} theory. Assuming $g_s<1$, the Planck scale is smaller than the string scale $l_s$ which characterizes the uncertainty relation for fundamental strings. The Planck scale was found as the characteristic scale of the $0+1$ dimensional SYM theory, which describes the low energy theory of D0-branes. Although the Planck scale is believed to be the minimal scale in 11D SUGRA, it is not the scale of uncertainty relations for D0-branes since a smaller characteristic scale $g_s l_s$ was found in \cite{DKPS}. It is called the ``fine structure'' scale, which can be seen only after taking into account the correction of the SYM action by the DBI action. Note that no smaller scale was found in the analysis of \cite{DKPS}. Although the fine structure scale was obtained as the characteristic scale in the transverse (spatial) directions, based on our results above we propose in this paper that the fine structure scale also sets the minimal length on a D0-brane worldline as in (\ref{D0}). Another support for the claim that (\ref{D0}) gives the correct uncertainty relation for D0-branes is obtained via T-duality. If the D1-brane has the uncertainty relation $\delta T\delta X\geq g_s l_s^2$, then for a D1-brane wrapped on a compactified circle with radius $R$, \footnote{ Since we assume that the string couplings are smaller than $1$, the compactification radius needs to satisfy $l_s/g'_s>R>g_s l_s$. } $\delta X$ can not be larger than $R$, which implies that $\delta T\geq g_s l_s^2/R$. By T-duality this is interpreted as a dual D0-brane with the uncertainty $\delta T\geq g'_s l_s$, where $g'_s=g_s l_s/R$ is the string coupling constant in the dual theory. This is exactly what we claimed in (\ref{D0}). Incidentally we note that in terms of M theory, $g_s l_s$ is the radius of the compactification through which the M theory is dual to type \mbox{\II A\hspace{.2em}} theory. By compactifying M theory on a circle smaller than the Planck scale, the smaller scale $g_s l_s$ is introduced into the compactified M theory. This would not be possible if there were an uncertainty relation like $\delta X\geq l_p$ in the uncompactified M theory. On the other hand, this is consistent with the worldvolume uncertainty relations we find for membranes and 5-branes as in (\ref{urm2}), (\ref{urm5}). \subsection{D$p$-branes via T-duality} \label{Td} In the above we have seen that T-duality can be used to derive uncertainty relation of D0-branes from that of D1-branes. Here we will generalize the arguments to all D$p$-branes. For simplicity we first consider the case of a flat background. We know that a D-string can be obtained from a D2-brane under T-duality. By wrapping a leg of the D2-brane on the circle, one introduces on the D2-brane worldvolume an uncertainty of the order \begin{equation} \label{X2R} \delta X^2 \sim R, \end{equation} since the center of the D-brane can be anywhere on the circle. It is thus natural to guess that the uncertainty relation on a D2-brane will involve a product of uncertainties of the form \begin{equation} \label{D2} \delta X^0 \delta X^1 \delta X^2 \geq g_s l_s^3, \end{equation} which is the product of the uncertainty relation (\ref{D1}) for a D1-brane and (\ref{X2R}) in terms of the dual $g_s$ and $l_s$. On the other hand, one can also repeat the derivation of D1-brane uncertainty relation in sec.\ref{noncomm} for D2-branes. Since $\delta X^i\delta X^j\geq g_s l_s^2 + \cdots $ for all $i\neq j$, $i,j=0,1,2$, one can derive $\delta X^0\delta X^1\delta X^2\geq g_s^{3/2}l_s^3 +\cdots $. This is a weaker condition than (\ref{D2}). At this moment we do not know how to derive (\ref{D2}) directly from open strings ending on D2-branes as in the case of D1-branes. We leave this interesting question for future study. For a D$p$-brane in general, we propose that \begin{equation} \label{urdp} \delta X^0 \delta X^1 \cdots \delta X^p \geq g_s l_s{}^{p+1} \end{equation} in flat spacetime. For a D$p$-brane in curved spacetime, the natural generalization of the uncertainty relation is just \begin{equation} \label{Dpur} \delta V_{(p+1)}\geq g_s l_s^{p+1}, \end{equation} where $\delta V_{(p+1)}$ is the uncertainty of the D$p$-brane worldvolume, which is the spacetime volume corresponding to $\delta X^i$. One can check that \eq{urdp} respects the T-duality of string theory. To see this, suppose that we start with a D$p$-brane in a string theory compactified on a circle of radius $R$ with string coupling $g_s$. We take one of the dimensions of the D$p$-brane, says $X^p$, to be wrapped on the circle. Since $\delta X^p \sim R$, we get for the D$(p-1)$-brane, \begin{equation} \delta X^0 \delta X^1 \cdots \delta X^{p-1} \geq g_s' l_s{}^p, \end{equation} where $g_s' = g_s l_s/ R$ is the dual string coupling. Thus \eq{urdp} is consistent with T-duality. It is interesting to note that \eq{urdp} can be derived from (\ref{Et}) and \begin{equation} \label{ev} \delta E = {\cal T} \delta V_p, \end{equation} where ${\cal T} =1/ \; g_s l_s^{p+1}$ is the tension of the D$p$-brane. Consider an experiment conducted on the D$p$-brane which is supposed to measure some point-like process. Because of the nonvanishing spacetime uncertainty relation, the region of the process would appear to have a spatial volume uncertainty of order $\delta V_p$. The associated uncertainty in energy would have a typical order of \eq{ev} and would be consistent with the standard $T-E$ type uncertainty relation (\ref{Et}). However, to really derive \eq{urdp} from \eq{Et}, one should in principle also include other possible sources of uncertainties (e.g. contributions from the oscillation modes) in \eq{ev}. The fact that one may derive \eq{urdp} by simply using \eq{ev} suggests that there might be some sort of stringy virial theorem. \subsection{Dyonic branes via S-duality} S-duality is expected to be an exact symmetry of type \mbox{\II B\hspace{.2em}} string theory. Under an S-duality transformation, a D$p$-brane gets transformed into an $(m,n)$-$p$-brane for $p =1,5$. For the purpose of explicit illustrations, we will spell out the $(m,n)$-string case in details. $(m,n)$-5-branes can be treated similarly. Starting with the uncertainty relation for a D-string (\ref{D1}), we can derive the uncertainty relation for a $(m,n)$-string by the $SL(2,{\bf Z})$ transformation of S-duality. It is \begin{equation} \label{urmnstring} \delta X^0 \delta X^1 \geq l_s^2 \frac{1}{ \sqrt{(m-n \chi)^2 +\frac{n^2}{g_s^2}}}, \end{equation} where $\chi$ is the axion vacuum expectation value. Notice that the right hand side is just the inverse of the tension of the $(m,n)$-string. In particular, we get (\ref{F1}) as a special case. \subsection{ M-branes via M/\mbox{\II B\hspace{.2em}} duality} Just as the D-brane is the end-point of an open string, one can also learn about the physics of the M5-brane by considering it as the boundary of an M2-brane (see for example \cite{CS} for other applications in this direction). It was shown in \cite{CH} that the M5-brane worldvolume can also become noncommutative in the presence of ${\cal F}_{ijk}$, which is a combination of the three-form gauge field and a worldvolume field strength. Applying similar considerations as in the string case above, we propose the following uncertainty relation for the M2-brane and M5-brane, \begin{eqnarray} &\delta X^0 \delta X^1 \delta X^2 \geq l_p^3 \quad \mbox{for M2-brane,} \label{urm2} \\ &\delta X^0 \delta X^1 \cdots \delta X^5 \geq l_p^6 \quad \mbox{for M5-brane,} \label{urm5} \end{eqnarray} where $l_p$ is the 11-dimensional Planck length. An uncertainty relation of the same form as (\ref{urm2}) was proposed in \cite{liyon1} as a result of (\ref{F1}) due to the M/\mbox{\II A\hspace{.2em}} duality, but with a different interpretation, which is analogous to the one they had for (\ref{uryoneya}). We now show that \eq{urmnstring} is related to the membrane uncertainty \eq{urm2} by using the M/\mbox{\II B\hspace{.2em}} duality. It is known \cite{sch,aspin} that \mbox{\II B\hspace{.2em}} string theory can be obtained from compactifying M theory on a shrinking 2-torus with radii $R_1, R_{11}$. A \mbox{\II B\hspace{.2em}} $(m,n)$-string is identified with a membrane wrapped over the $(m,n)$-cycle on the torus with length \begin{equation} \label{lmn} L_{(m,n)} = R_{11} \sqrt{(m-n \tau_1)^2 + n^2 \tau_2^2}. \end{equation} Here $\tau = \tau_1 + i \tau_2$ is the modular parameter of the torus and it is identified \cite{sch,aspin} with the \mbox{\II B\hspace{.2em}} string theory parameters as \begin{equation} \tau_1 =\chi, \quad\quad \tau_2 = 1/g_s. \end{equation} Now starting from (\ref{urm2}) and using \begin{equation} \delta X^2 \sim L_{(m,n)} \end{equation} for the uncertainty for the membrane direction which is wrapped on a cycle, we obtain immediately \eq{urmnstring}. Here we used $l_s^2= l_p^3/ R_{11}$. In fact, since our uncertainty relations can be obtained mathematically from (\ref{Et}) and (\ref{ev}), the matching of the brane spectra for dual theories implies that the uncertainty relations must be consistent with all dualities. For instance, it is automatically true that the M/\mbox{\II A\hspace{.2em}} duality also relates (\ref{urm2}) and (\ref{urm5}) to (\ref{F1}) and (\ref{urdp}) for $p=2,4$. Incidentally, the M/\mbox{\II A\hspace{.2em}} duality also gives rise to the uncertainty relation \begin{equation} \delta X^0\cdots\delta X^5\geq g_s^2 l_s^6 \end{equation} for the NS5-brane. \section{Discussions} \label{remarks} In this paper, we discussed the uncertainty relations for the D-brane worldvolume. We introduced the notion of a worldvolume uncertainty and explained how it is defined within the context of string field theory. We proposed worldvolume uncertainty relations that are consistent with the various dualities in string theory. We have also generalized the commutation relation for the noncommutative gauge theory to a nontrivial background of ${\cal F}$ in the lowest order approximation. It would be interesting to generalize this result to the full generality of an arbitrary background. This could be relevant to the interesting proposal in \cite{JR}. In the following we remark on several related subjects. \subsection{Comments on some other uncertainty relations} \underline{Uncertainty relation of Wigner} In the classical study of Wigner \cite{Wigner}, the effects of quantum mechanics on the measurability of the spatial distance was estimated to be given by \begin{equation} \label{w1} \delta D \geq \sqrt{T/M_c}. \end{equation} Here $T$ is the time scale for the process involved and $M_c$ is the mass of the clock. This analysis has recently been strengthened in \cite{Sasa}, which utilizes the existence of a Schwarzschild horizon $R_s$ for any massive object and thus it is necessary that \begin{equation} \label{w2} \delta D \geq R_s \sim G M_c. \end{equation} Combining with \eq{w1}, one obtains \begin{equation} \label{w3} (\delta D)^3 \geq G T. \end{equation} Notice that in this analysis, the precision of the measurement of time is not limited. The uncertainty relation we proposed is consistent with these results. For example, in the brane world scenario, when $\delta T=0$, our uncertainty relation for a D$3$-brane says that $\delta D = \infty$, which is stronger than \eq{w1} or \eq{w3}. The argument leading to (\ref{w3}) utilizes the most popular reason for the belief in the existence of spacetime uncertainty relations. That is, due to quantum mechanics a large energy is needed to probe a small length scale, and when the energy is too large a black hole is formed, which forbids the measurement of distances behind the horizon. However, in our derivation of the uncertainty relations (\ref{urdp}), we did not mention anything about event horizon at all. It remains to be seen how the consideration of black holes can lead to the determination of uncertainty relations in string theory, and whether it will lead to new uncertainty relations. \underline{Uncertainty relation of Yoneya} Notice that ours uncertainty relations \eq{urdp} are not of the same type as those proposed in \cite{liyon1,Yon,yoneya,MS}. These authors proposed uncertainty relations that involve the transverse coordinates while ours are solely for the brane world. For example, in \cite{liyon1,Yon,yoneya}, it was proposed that \begin{equation} \label{uryoneya} \delta T \delta X \geq l_s^2 \end{equation} for a D$p$-brane. Here $\delta T$ is understood as the uncertainty in the longitudinal directions on the brane (to be more precise, $\delta T=|\delta\sigma|$, where $\sigma$ is the worldvolume coordinates in the static gauge) and $\delta X$ represents the uncertainty in directions transverse to the brane. As it was pointed out in \cite{DKPS}, the short distance regime of D-branes are probed by open strings. The exchange of a closed string state between two D-branes, for example, can be viewed as the creation and annihilation of a pair of open strings (an open string loop diagram) due to the modular invariance of string theory. Therefore, the scattering of D-branes is limited by the uncertainty of open strings, and (\ref{uryoneya}) is a direct result of (\ref{F1}). Our uncertainty relation (\ref{urdp}), on the other hand, is concerned with the uncertainty among longitudinal directions on the brane, and it has a form different from (\ref{uryoneya}). In particular ours have a form that depends on the dimensionality of the brane and have additional dependence on $g_s$. \subsection{UV-IR relations, holography and uncertainty relations} In \cite{yoneya,liyon1}, a generalized conformal symmetry was found for the D$p$-brane super Yang-Mills action. It is easy to check that the $(p+1)$-dimensional YM action with coupling $g_{\mbox{\tiny YM}}$, which is schematically \begin{equation} S=\frac{1}{g_{\mbox{\tiny YM}}^2}\int d^{p+1}\sigma((\partial X)^2+X^4), \end{equation} is invariant under the following scaling transformation \begin{eqnarray} &X^a \rightarrow \lambda}\def\L{\Lambda X^a, \quad \sigma_i \rightarrow \lambda}\def\L{\Lambda^{-1} \sigma_i, \label{gct1}\\ &g_{\mbox{\tiny YM}}^2 \rightarrow \lambda}\def\L{\Lambda^{3-p} g_{\mbox{\tiny YM}}^2. \label{gct2} \end{eqnarray} Here $X^a (a= p+1, \cdots, 9)$ are the transverse scalars, $\sigma_i (i=0,1,\cdots p)$ are the worldvolume coordinates in the static gauge. The coupling $g_{\mbox{\tiny YM}}^2$ is related to the string coupling by \begin{equation} \label{gym} g_{\mbox{\tiny YM}}^2 =g_s l_s^{p-3}. \end{equation} The uncertainty relation (\ref{uryoneya}) is invariant under this scaling \eq{gct1} together with $g_s \rightarrow \lambda}\def\L{\Lambda^{3-p} g_s$ and $l_s$ being invariant. Notice that \eq{gct1} is reminiscent of the UV-IR distance relation \cite{SW,UVIR} in the context of AdS/CFT holography \cite{p1,p2,p3}. Maybe this scaling relation is a generic property for general holographic dualities \cite{hol1,hol2}. It was discussed in \cite{liyon1,Min} that the uncertainty relation \eq{uryoneya} is consistent with the UV-IR relation and it was suggested that the uncertainty principle is the underlying principle that implies the UV-IR relation, which in turn guarantees the holographic bound \cite{hol2} to be satisfied \cite{SW}. Since our uncertainty relations are only involved with the worldvolume uncertainties, the scaling (\ref{gct1})-(\ref{gct2}) does not give any nontrivial implications on our relations. On the other hand, since our uncertainty relation implies the existence of a minimal area, it may also be relevant to the holographic principle and to the verification of the holographic bound. It would be interesting if one can see this explicitly. Another interesting issue is that based on the UV-IR relation and holographic principle, it is natural to ask what kind of spacetime property (presumably a spacetime uncertainty relation) will be implied by the worldvolume uncertainty relations. We leave these issues for future studies. \subsection{Characteristic scale} The scaling transformation (\ref{gct1})-(\ref{gct2}) can also be used to find the characteristic scale of YM theory. Let $\lambda}\def\L{\Lambda=g_s^{\frac{1}{p-3}}$, then we scale $g_s\rightarrow 1$ and $X\rightarrow g_s^{\frac{1}{p-3}}X$. This means that the Higgs vacuum expectation value $g_s^{\frac{1}{p-3}}X$ is independent of $g_s$, and thus the YM characteristic scale is $g_s^{\frac{-1}{p-3}}l_s$. For $p=0$ this gives the Planck scale $g_s^{1/3}l_s$. For $p=3$ it is the string scale $l_s$. For $p>3$ this scale is much larger than the string scale, but for these cases YM theory is not renormalizable and it means that we cannot trust it. As it was mentioned in sec.\ref{D0-brane}, a characteristic scale does not have to be the minimal scale. For the case of D0-branes, it is the characteristic scale of the DBI action that turns out to be the minimal scale. It is thus of interest to work out also the characteristic scale for the DBI theory for a D$p$-brane. The DBI action is \begin{equation} S=\frac{1}{g_s l_s^{p+1}}\int d^{p+1}\sigma\sqrt{-{\rm det\,}(g+{\cal F})}, \end{equation} which is invariant under the scaling \begin{equation} X\rightarrow \lambda}\def\L{\Lambda X, \quad g_s\rightarrow \lambda}\def\L{\Lambda^{p+1}g_s, \end{equation} and an arbitrary scaling of $\sigma$. Letting $\lambda}\def\L{\Lambda=g_s^{\frac{-1}{p+1}}$, we find $g_s\rightarrow 1$ and $X\rightarrow g_s^{\frac{-1}{p+1}}X$. This means that $g_s^{\frac{1}{p+1}}l_s$ is the characteristic scale of the DBI action for the transverse directions. It happens that this characteristic scale is also the one determining the minimal volume in our uncertainty relations (\ref{Dpur}). It is possible that this is also the minimal length scale for the transverse directions of a D$p$-brane much like the case of D0-branes discussed in \cite{DKPS}. Even if both the transverse and longitudinal directions of a D0-brane are bounded by this scale $g_s l_s$, (\ref{uryoneya}) is still a stronger condition than just the product of the two minimal lengths. In sec.\ref{Td} we also mentioned that (\ref{urdp}) is a stronger condition than a product of (\ref{D1}) for each pair of longitudinal directions. It seems that in string theory we need a complicated web of uncertainty relations, which cannot be deduced from a single master relation, to fully state the uncertainty property of spacetime. \section*{Acknowledgment} C.S.C. thanks G. Amelino-Camelia and A. Bilal for helpful discussions. He is also grateful to the Department of Physics and the Center for Theoretical Physics at the National Taiwan University for hospitality where part of this work was carried out. P.M.H. and Y.C.K. thank S. Das and particularly M. Li for helpful discussions. P.M.H. thanks M. M. Sheikh-Jabbari for helpful remarks. The work of C.S.C. is supported by the Swiss National Science Foundation. The work of P.M.H. and Y.C.K. is supported in part by the National Science Council (NSC 88-2112-M-002-042, NSC 88-2112-M-002-034) and the Center for Theoretical Physics, National Taiwan University, Taiwan, R.O.C.
\section{Introduction} In many flows the transition to turbulence proceeds via a sequence of bifurcations to flows of ever increasing spatial and temporal complexity. Analytical and experimental efforts in particular on layers of fluid heated from below \cite{Busse,Kosch} and fluids between rotating concentric cylinders \cite{Kosch,Swinney} have lead to the identification and verification of several routes to turbulence, which typically involve a transition from a structureless laminar state to a stationary spatially modulated one and then to more complicated states in secondary and higher bifurcations. Transitions in shear flows do not seem to follow this pattern \cite{Grossmann1996}. Typically, a transition to a turbulent state can be induced for sufficiently large Reynolds number with finite amplitude perturbations, just as in a subcritical bifurcation. However, in the most spectacular cases of plane Couette flow between parallel plates and Hagen-Poiseuille flow in a pipe \cite{DR}, there is no linear instability of the laminar profile for any finite Reynolds number that could give rise to a subcritical bifurcation. The turbulent state seems to be high dimensional immediately, without clear temporal or spatial patterns (unlike the rolls in Rayleigh-B\'enard flow). And the transition seems to depend sensitively on the initial conditions. Based on these characteristic features it has been argued that a novel kind of transition to turbulence different from the well-known three low-dimensional ones is at work \cite{GG1994}. Recent activity has focussed on three features of this transition: the non-normality of the linear eigenvalue problem \cite{GG1994,BB1988,TTRD1993,wal1,wal2,DM}, the occurence of new stationary states without instability of the linear profile \cite{Nagata1990,Nagata1997,CB1997,wal98} and the fractal properties of the lifetime landscape of perturbations as a function of amplitude and Reynolds number \cite{ES1997}. The non-normality of the linear stability problem implies that even in the absence of exponentially growing eigenstates perturbations can first grow in amplitude before decaying since the eigenvectors are not orthogonal. During the decay other perturbations could be amplified, giving rise to a noise sustained turbulence \cite{IF1993}. The amplification could also cause random fluctuations to grow to a size where the nonlinear terms can no longer be neglected \cite{wal2,DM}. Then the dynamics including the nonlinear terms could belong to a new asymptotic state, different from the laminar profile, perhaps a turbulent attractor. Presumably, this attractor would be built around stationary or periodic solution. Here, the observation of tertiary structures \cite{Nagata1990,Nagata1997,CB1997,wal98} comes in since they could form the basis for the turbulent state. Finally, the observation of fractality in the lifetime distribution suggests that the turbulent state is not an attractor but rather a repeller: Infinite lifetimes occur only along the stable manifolds of the repeller, all other initial conditions will eventually decay. Permanent turbulence would thus correspond to noise induced excitations onto a repeller. In plane Couette flow some of the features described above have been identified, but only with extensive numerical effort \cite{Nagata1990,Nagata1997,CB1997,ES1997}. The aim of the present work is to present a simple model that is based on the Navier-Stokes equation and captures the essential elements of the transition. It is motivated in part by the desire to obtain a numerically more accessible model which perhaps will provide as much insight into the transition as the Lorenz model \cite{Lorenz} for the case of fluids heated from below (presumably at the price of similar shortcomings). The two and three degree of freedom models proposed by various groups (and reviewed in \cite{BT1997}) to study the effects of non-normality mock some features of the Navier-Stokes equations considered essential by their inventors but they are not derived in some systematic way from the Navier Stokes equation. The model used here differs from the one proposed by Waleffe \cite{wal2} in the selection of modes. Attempts to built models for shear flows using Fourier modes immediately reveal an intrinsic difficulty: In the case of fluids heated from below the nonlinearity arises from the coupling of the temperature gradient to the flow field so that two wave vectors, ${\bf k}$ and $2{\bf k}$, suffice to obtain nonlinear couplings. In shear flows, the nonlinearity has to come from the coupling of the flow field with itself through the advection term $({\bf u}\cdot\nabla){\bf u}$. This imposes rather strong constraints on the wave vectors. At least three wave vectors satisfying the triangle relation ${\bf k}_1+{\bf k}_2+{\bf k}_3=0$ are required to collect a contribution from the advection term. A minimal model thus has at least six complex variables. Three of these decay monotonically to zero, leaving three for a nontrivial dynamics. In the subspaces investigated (B.E., unpublished), the most complex behaviour found is a perturbed pitchfork bifurcation, which may be seen as a precursor of the observed dynamics: for Reynolds numbers below a critical value, there is only one stable state. Above that value a pair of stable and unstable states is born in a saddle-node bifurcation. The stable state can be excited through perturbations of sufficient amplitude. The basins of attraction of the two stable states are intermingled, but the boundaries are smooth. Thus more wave vectors are needed and they have to couple in a nontrivial manner to sustain permanent dynamics. The specific set of modes used is discussed in section \ref{sec_model}. It is motivated by boundary conditions for the laminar profile and the observation that wave vectors pointing to the vortices of hexagons satisfy the triangle conditions in a most symmetrical manner. Other than that the selected vectors are a matter of trial and error. In the end we arrive at a model with 19 real amplitudes, two force terms and 212 quadratic couplings. Without driving and damping the dynamics is energy conserving, as would be the corresponding Euler equation (suitably truncated). Moreover, the perturbation amplitudes can be put together to give complete flow fields. Thus the model has a somewhat larger number of degrees of freedom, but the dynamics should provide a realistic approximation to shear flows. The outline of the paper is as follows. In section \ref{sec_model} we present the model, in particular the selected wave vectors, the equations of motion and a discussion of symmetries. In section \ref{sec_dynamic} we focus on the dynamical properties of initial perturbations as a function of amplitude and Reynolds number. In section \ref{sec_stationary} we discuss the stationary states, their bifurcations and their stability properties. We conclude in section \ref{sec_conclusion} with a summary and a few final remarks. \section{The model shear flow} \label{sec_model} Ideal parallel shear flows have infinite lateral extension. Both in experiment and theory this cannot be realized. We therefore follow the numerical tradition and chose periodic boundary conditions in the flow and neutral direction. The flow is confined by parallel walls a distance $d$ apart. A convenient way to built a low dimensional model is to use a Galerkin approximation. Solid boundaries would require the vanishing of all velocity components and complicated Galerkin functions where all the couplings can only be calculated numerically. However, under the assumption that here as well as in many other situations the details of the boundary conditions effect the results only quantitatively but not qualitatively, we can adopt free-free boundary conditions on the walls and use simple trigonometric functions as basis for the Galerkin expansion. Similarly, the nature of the driving (pressure, boundary conditions or volume force) should not be essential so that we take a volume force proportional to some basis function (or a linear combination thereof). This still leaves plenty of free parameters to be fixed below. \subsection{Galerkin approximation} We expand the velocity field in Fourier modes, \beq \fettu(\fettx,t) = \sum_{\fettk} \fettu(\fettk,t) e^{i\fettk\cdot\fettx}\,. \eeq Incompressibility demands \beq \fettu(\fettk,t) \cdot \fettk = 0 \,. \label{incomp} \eeq The Navier-Stokes equation for the amplitudes $\fettu(\fettk,t)$ becomes \bea \partial_t\fettu(\fettk,t) = &-& i p_\fettk \fettk - i \sum_{{\bf p}+{\bf q}=\fettk} \left(\fettu({\bf p},t)\cdot {\bf q} \right) \fettu({\bf q},t) \nonumber\\ &-& \nu \fettk^2 \fettu(\fettk,t) + f_\fettk \eea where $p_\fettk$ are the Fourier components of the pressure (divided by the density), $\nu$ is the kinematic viscosity and $f_\fettk$ are the Fourier components of the volume force sustaining the laminar profile. There are three constraints on the components $\fettu(\fettk)$: incompressibility (\ref{incomp}), reality of the velocity field, \beq \fettu(-\fettk) = \fettu(\fettk)^* \eeq and the boundary conditions that the flow is limited by two parallel, impenetrable plates. The ensuing requirement $u_z(x,y,z)=0$ at $z=0$ and $z=d$ (where $d$ is the separation between plates) is most easily implemented through periodicity in $z$ and the mirror symmetry \beq \left(\matrix{u_x \cr u_y \cr u_z}\right) (x,y,-z) = \left(\matrix{u_x \cr u_y \cr -u_z}\right) (x,y,z) \,, \label{mirrorsymm} \eeq which in Fourier space requires \beq \left(\matrix{u_x \cr u_y \cr u_z}\right) (-k_x,-k_y,k_z) = \left(\matrix{u_x^* \cr u_y^* \cr -u_z^*}\right) (k_x,k_y,k_z) \,. \eeq This is not sufficient to fix the coefficients: the dynamics also has to stay in the relevant subspace, and thus the time derivatives have to satisfy similar requirements. \subsection{The wave vectors} The choice of wave vectors is motivated by the geometry of the flow and the aim to include nonlinear couplings. The basic flow shall be a flow in $y$-direction, neutral in the $x$-direction and sheared in the $z$-direction. Thus we take the first three wave vectors in $z$-direction, \beq \fettk_1 = \left(\matrix{0\cr0\cr1}\right)\,, \qquad \fettk_2 = \left(\matrix{0\cr0\cr2}\right)\,, \qquad \fettk_3 = \left(\matrix{0\cr0\cr3}\right) \,. \eeq The negative vectors $-\fettk_i$ also belong to the set but will not be numbered explicitely. In these units, the periodicity in the $z$-direction is $2\pi$, so that the separation between the plates is $d=\pi$ because of the mirror symmetry (\ref{mirrorsymm}). The amplitude $\fettu(\fettk_1)$ will carry the laminar profile and $\fettu(\fettk_3)$ can be excited as a modification to the laminar profile. $\fettk_2$ is needed to provide couplings through the nonlinear term. These three vectors satisfy a triangle identity $\fettk_1+\fettk_2-\fettk_3=0$, but the nonlinear term vanishes since they are parallel. The next set of wave vectors contains modulations in the flow and neutral direction, \beq \fettk_4 = \left(\matrix{1\cr0\cr0}\right)\,, \quad \fettk_5 = \left(\matrix{1/2\cr\sqrt{3}/2\cr0}\right)\,, \quad \fettk_6 = \left(\matrix{1/2\cr-\sqrt{3}/2\cr0}\right) \,. \eeq Together with $-\fettk_i$ they form a regular hexagon, so that they provide nontrivial couplings in the nonlinear term. The periodicity in flow direction is $4\pi/\sqrt{3}$, in the neutral direction it is $4\pi$. Finally, this hexagon is lifted upwards with $\fettk_1$ and $\fettk_2$ to form the remaining 12 vectors, \bea \fettk_7\;\, &=& \fettk_1 + \fettk_4 \quad\; \fettk_8\;\, = \fettk_1 + \fettk_5 \quad\; \fettk_9\; = \fettk_1 + \fettk_6 \cr \fettk_{10} &=& \fettk_1 - \fettk_4 \quad\; \fettk_{11} = \fettk_1 - \fettk_5 \quad\; \fettk_{12} = \fettk_1 - \fettk_6 \cr \fettk_{13} &=& \fettk_2 + \fettk_4 \quad\; \fettk_{14} = \fettk_2 + \fettk_5 \quad\; \fettk_{15} = \fettk_2 + \fettk_6 \cr \fettk_{16} &=& \fettk_2 - \fettk_4 \quad\; \fettk_{17} = \fettk_2 - \fettk_5 \quad\; \fettk_{18} = \fettk_2 - \fettk_6\,. \eea The full set $\fettk_i$, $i=1\ldots18$ is shown in Fig.~\ref{wavevectors}. The Fourier amplitudes $\fettu(\fettk_i)$ have to be orthogonal to $\fettk_i$ because of incompressibility (\ref{incomp}). If they are expanded in basis vectors perpendicular to $\fettk_i$, the pressure drops out of the equations and need not be calculated. We therefore chose normalized basis vectors \bea \fettn(\fettk_i) &=& \left.\left({-k_x k_z\over k_x^2+k_y^2} , {-k_y k_z\over k_x^2+k_y^2}, 1\right)^T \right/ \sqrt{1+k_z^2/(k_x^2+k_y^2)} \cr \fettm(\fettk_i) &=& \left({k_y} , {-k_x}, 0\right)^T\left/ \sqrt{k_x^2+k_y^2} \right.\, \eea so that $\fettn$, $\fettm$ and $\fettk$ form an orthogonal set of basis vectors. For the negative vectors $-\fettk_i$ we chose the basis vectors $\fettn(-\fettk_i)=\fettn(\fettk_i)$ and $\fettm(-\fettk_i)=-\fettm(\fettk_i)$. If the $x$ and $y$ components of $\fettk$ vanish, the above definitions are singular and replaced by \beq \fettn = (1,0,0)^T \qquad \fettm = (0,1,0)^T \,. \eeq The amplitudes of the velocity amplitude are now expanded as \beq \fettu(\fettk_i,t) = \alpha(\fettk_i,t) \fettn(\fettk_i) + \beta(\fettk_i,t) \fettm(\fettk_i) \,. \eeq The impenetrable plates impose further constraints on the $\alpha(\fettk_i)$ and $\beta(\fettk_i)$. For $i=1$, $2$ and $3$ the wave vector has no components in the $x$- and $y$-directions, so that $\alpha$ and $\beta$ have to be real. For $i=4$, $5$ and $6$ the velocity field cannot have any components in the $z$-direction, hence $\alpha=0$. The remaining wave vectors $\fettk_i$ and $-\fettk_i$ with $i=7,\ldots,18$, a total of 24, divide up into six groups of 4 vectors each, \beq \fettk=(k_x,k_y,k_z),\; \fettk' = (-k_x,-k_y,k_z), \; -\fettk \; \mbox{and}\; -\fettk'\,. \eeq The groups are formed by the vectors and their negatives in the pairs with indices (7,10), (8,11), (9,12), (13,16), (14,17) and (15,18). The amplitudes of the vectors in the sets are related by \bea \alpha(\fettk) &=& \alpha(-\fettk)^* = - \alpha(\fettk')^* \cr \beta(\fettk) &=& \beta(-\fettk)^* = - \beta(\fettk')^* \,. \label{bcs} \eea Thus the full model has $6 + 6 + 6\times4 = 36$ real amplitudes. Restricting the flow by a point symmetry around $\fettx_0= (0,0,\pi/2)^T \,$ eliminates the contributions from ${\bf k}_2$ and some other components, resulting in a 19-dimensional subspace with nontrivial dynamics and the following amplitudes: \bea \alpha(\fettk_1) &=& y_1 \qquad \beta(\fettk_1) = y_2\cr \alpha(\fettk_3) &=& y_3 \qquad \beta(\fettk_3) = y_4\cr \beta(\fettk_4) &=& i y_5 \qquad \beta(\fettk_5) = i y_6 \qquad \beta(\fettk_6) = i y_7 \cr \alpha(\fettk_7) &=& y_8 \qquad \beta(\fettk_7) = y_9\cr \alpha(\fettk_8) &=& y_{10} \qquad \beta(\fettk_8) = y_{11}\cr \alpha(\fettk_9) &=& y_{12} \qquad \beta(\fettk_9) = y_{13}\cr \alpha(\fettk_{13}) &=& i y_{14} \qquad \beta(\fettk_{13}) = i y_{15}\cr \alpha(\fettk_{14}) &=& i y_{16} \qquad \beta(\fettk_{14}) = i y_{17}\cr \alpha(\fettk_{15}) &=& i y_{18} \qquad \beta(\fettk_{15}) = i y_{19}\,; \eea components not listed vanish or are related to the given ones by the boundary conditions (\ref{bcs}). A complete listing of the flow fields ${\bf u}_i$ associated with the coefficients $y_i$ such that ${\bf u}=\sum_i y_i {\bf u}_i$ as well as of the equations of motion are available from the authors. \subsection{The equations of motion} In this 19-dimensional subspace $y_1\ldots y_{19}$ the equations of motion are of the form \beq \dot y_i = \sum_{j,k} A_{ijk} y_j y_k - \nu K_i y_i + f_i \,. \eeq Of the driving force all components but $f_2$ and $f_4$ vanish. Moreover, if the $f$'s are taken to be proportional to $\nu$, the resulting laminar profile has an amplitude independent of viscosity (and thus Reynolds number). These components give rise to a laminar profile that is a superposition of a $\cos(z)$ profile (from $f_2$) and a $\cos(3z)$ profile (from $f_4$). This allows us to approximate the first two terms of the Fourier expansion of a linear profile with velocity $u_y=\pm1$ at the walls, \beq \fettu_0= \frac{8}{\pi^2} (\cos z + {1\over 9} \cos 3z) \, {\bf e}_y \,. \label{laminar_profile} \eeq that can be obtained with a driving $f_2 = 4\nu/\pi^2$ and $f_4=4\nu/9\pi^2$ (see Fig.~\ref{laminarflows}). The nonlinear interactions in the Navier-Stokes equation conserve the energy $E=\frac{1}{2}\int dV \fettu^2$. In the 19-dimensional subspace, the corresponding quadratic form is \beq E = V \left( \sum_{i=1}^{7} y_i^2 + 2 \sum_{i=8}^{19} y_i^2\right)\,. \eeq The above equations conserve this form without driving and dissipation. With dissipation but still without driving, the time derivative is negative definite, indicating a monotonic decay of energy to zero. Finally, we define the Reynolds number using the wall velocity of the linear profile, $u_0=1$, the half width of the gap, $D=d/2=\pi/2$ and the viscosity $\nu$, \beq Re = u_0 D/\nu = \pi/2\nu \,. \eeq The other geometric parameters are a period $4\pi/\sqrt{3}$ in flow direction and $4\pi$ perpendicular to it. \subsection{Symmetries} \label{symmetries} We achieved the impenetrability of the plates by requiring the mirror symmetry: \beq \left(\matrix{u_x \cr u_y \cr u_z}\right) (x,y,-z) = \left(\matrix{u_x \cr u_y \cr -u_z}\right) (x,y,z) \, . \eeq The reduction from 36 to 19 modes was achieved by restricting the dynamics to a subspace where the flow has the point symmetry around $\fettx_0= (0,0,\pi/2)^T \,$, a point in the middle of the shear layer, \beq \left(\matrix{u_x \cr u_y \cr u_z}\right) (x,y,z+\pi/2) =\left(\matrix{-u_x \cr -u_y \cr -u_z}\right) (-x,-y,-z+\pi/2) \,. \eeq In addition, there are further symmetries that can be used to reduce the phase space. There is a reflection on the $y$-$z$-plane, \beq T_1: \left(\matrix{u_x \cr u_y \cr u_z}\right) (x,y,z) \rightarrow \left(\matrix{-u_x \cr u_y \cr u_z}\right) (-x,y,z) \,. \\ \eeq and two shifts by half a lattice spacing, \bea T_2:& & \left(\matrix{u_x \cr u_y \cr u_z}\right) (x,y,z) \rightarrow \left(\matrix{u_x \cr u_y \cr u_z}\right) (x+2\pi,y,z) \, \\ T_3:&& \left(\matrix{u_x \cr u_y \cr u_z}\right) (x,y,z) \rightarrow \left(\matrix{u_x \cr u_y \cr u_z}\right) (x+\pi,y+\pi/\sqrt 3,z) \, . \eea When applied to the flow these transformations induce changes in the variables $y_i$ (typically exchanges or sign changes), but the equations of motion are invariant under these transformations. Thus, if a certain flow has this symmetry, it leads to constraints on the variables $y_i$, and if it does not have this symmetry immediately a new flowfield can be obtained by applying this symmetry transformation. We do not attempt to analyze the full symmetry structure here and confine our discussion to two illustrative examples which are relevant for the stationary states discussed below. Demanding invariance of the flow field to the reflection symmetry $T_1$ leads to the following constraints on the variables $y_i$: \bea y_1&=&y_3=y_5=y_8=y_{15}=0\nonumber\\ y_6&=&y_7\quad y_{10}=-y_{12} \nonumber\\ y_{11}&=&y_{13}\quad y_{16}=-y_{18}\quad y_{17}=y_{19}. \eea The non vanishing components, $y_2$, $y_4$, $y_6=y_7$, $y_9$, $y_{10}=-y_{12}$, $y_{11}=y_{13}$, $y_{14}$, $y_{16}=-y_{18}$, $y_{17}=y_{19}$ thus define a 9 dimensional subspace. For the combined symmetry $T_1T_2$ we find the constraints \bea y_1&=&y_3=y_5=y_8=y_{15}=0\nonumber\\ y_6&=&-y_7\quad y_{10}=y_{12} \nonumber\\ y_{11}&=&-y_{13}\quad y_{16}=y_{18}\quad y_{17}=-y_{19} \eea and again a 9 dimensional subspace with non vanishing components $y_2$, $y_4$, $y_6=-y_7$, $y_9$, $y_{10}=y_{12}$, $y_{11}=-y_{13}$, $y_{14}$, $y_{16}=y_{18}$, $y_{17}=-y_{19}$. The dimensions of the invariant spaces vary from a minimum of 6 (for each a $T_1T_3$ and $T_1T_2T_3$ invariance) and a maximum of 10 (for $T_2T_3$-invariance). As mentioned, one can classify flows according to their symmetries. The most asymmetric flows are eightfold degenerate as the application of the eight combinations of the symmetries give eight distinct flows. The laminar flow profile is invariant under all the linear transformations and is the only member of the class with highest symmetry. The other stationary states discussed below fall into equivalence classes with eight members or four members if they are invariant under $T_1$ or $T_1T_2$. \section{Dynamics of perturbations} \label{sec_dynamic} A stability analysis shows that the laminar flow profile is linearly stable for all Reynolds numbers. The matrix of the linearization is non-normal with a block structure along the diagonal. To bring this structure out more clearly, it is best to order the equations in the sequence 1, 2, 3, 4, 5, 7, 15, 8, 9, 14, 13, 19, 12, 18, 6, 11, 17, 10, 16. The matrix of the linearization then is upper diagonal, with a clear block structure: there are 10 eigenvalues isolated on the diagonal, three $2\times2$ blocks and one $3\times 3$ block as well as several couplings between them in the upper right corner. While some eigenvalues can be complex, all of them have negative real part as shown in Fig.~\ref{eigenvallaminar}. For vanishing viscosity, the eigenvalues become zero or purely imaginary. Large amplitude perturbations, however, need not decay. Already in the linear regime the non-orthogonality of the eigenvectors can give rise to intermediate amplifications into a regime where the nonlinear terms become important \cite{GG1994,BB1988,TTRD1993,wal1,wal2}. In a related study on plane Couette flow \cite{ES1997} we used the lifetime of perturbations to get information on the dynamics in a high-dimensional phase space. As in that case, the amplitude of the velocity field in the $z$-direction indicates the survival strength of a perturbation. Linearizing the equations of motion around the base flow ${\bf u}_0$ gives for the perturbation ${\bf u}'$ the equation \beq \partial_t{\bf u}' = - ({\bf u}_0\cdot\nabla){\bf u}' - ({\bf u}'\cdot\nabla){\bf u}_0- \nabla p' + \nu \Delta {\bf u}' \,. \eeq The second term on the right hand side describes the energy source for the perturbation, and depends, because of ${\bf u}_0 = u_0(z)\, {\bf e}_y$ and thus \beq ({\bf u}'\cdot\nabla){\bf u}_0 = u_z' \partial_z u_0(z) {\bf e}_y \eeq in an essential way on the $z$-components of the perturbation. Thus, if the amplitudes $y_8$, $y_{10}$, $y_{12}$, $y_{14}$, $y_{16}$ and $y_{18}$ become too small, the decay of the perturbation cannot be stopped any more. These modes account also for most of the off-diagonal block-couplings. A model for sustainable shear flow turbulence has to include some of these modes. We chose a fixed initial flow field with a random selection of amplitudes $y_1,\ldots,y_{19}$, scaled it by an amplitude parameter $A$ and measured the lifetime as a function of $A$ and Reynolds number Re. Fig.~\ref{timedependence} shows the time evolution of such a perturbation at $Re=400$ with one mode driven and for different amplitudes. For small $A$ there is an essentially exponential decay, whereas for larger amplitudes the perturbation swings up to large amplitude and shows no sign of a decline at all. The results for many amplitudes and Reynolds numbers are collected in Fig.~\ref{landscape} in a landscape plot. For small Reynolds number and/or small amplitude the lifetimes of perturbations are short, indicated by the light areas. For Reynolds numbers around 100 isolated black spots appear, indicating the occurence of lifetimes larger than the integration time (which increases with $Re$ so that $t_{max}/Re = 4\pi$). The spottiness for $Re$ between about 100 and 1000 is due to rapid changes in lifetimes from pixel to pixel. For Re above 1000 the long lifetimes dominate. These results are in good agreement with what has been observed in plane Couette flow. Fig.~\ref{landscape}b shows a similar plot for the case with two modes driven; it is qualitatively similar, but quantitatively shifted to higher Reynolds numbers. In connection with the non-normality of the linearized eigenvalue problem it has been argued that the upper limit on the size of perturbations for which the non-linear terms in the dynamics can be neglected decreases algebraically like $Re^{-\alpha}$. Different exponents have been proposed, ranging from 1 to 3 \cite{GG1994,wal2,BT1997}. It seems that for large $Re$ (where the model is actually less reliable because of the limited spatial resolution) the envelope of the long lived states in the fractal life time plot decays like $Re^{-1}$. The sensitive dependence of lifetimes on initial conditions and parameters is further highlighted in Fig.~\ref{landscape_magnification} and \ref{lifetimes}. The first shows the lifetime in the plane of the amplitudes $y_{16}$ and $y_{17}$ at Reynolds number $Re=400$ with all other components fixed. There is considerable structure on many scales. One notes 'valleys' of short lifetimes between 'plateaus' of longer lifetimes and a granular structure within both. The striations are reminiscent of features seen near fractal basin boundaries \cite{Ott}. Fig.~\ref{lifetimes} shows successive magnifications of lifetime versus amplitude plots at Re=200. Even after a magnification by $10^{7}$ there is no indication of a continuous and smooth variation of lifetime with amplitude. \section{Stationary states} \label{sec_stationary} Motivated by the observation of new stationary structures in plane Couette flow for sufficiently high Reynolds number \cite{Nagata1990,Nagata1997,CB1997} we searched for non-trivial stationary solutions and studied their generation, evolution and symmetries. We computed the stationary states with the help of a Monte Carlo algorithm. The initial conditions for the $y_i$'s were chosen randomly out of the interval $[-1/2,1/2]$ and the Reynolds number was chosen randomly matrixwith an exponential bias for small $Re$ in the interval $[10,10000]$. With these initial conditions we entered a Newton algorithm. If the Newton algorithm converged, we followed the fixed point in Reynolds number as far as possible. We included about 200000 attempts in the Monte Carlo search. The stationary states found for a single driven mode are collected in Fig.~\ref{stationary_single}. No stationary states (besides the laminar profile) were found for Reynolds numbers below about 190. Between 190 and about 500 there are eight stationary states which divide into two groups of four symmetry related states each. With increasing Reynolds number more and more stationary states are found and they reach down to smaller and smaller amplitude. The envelope of all states reflects the $Re^{-1}$ behaviour found for the borderline where nonlinearity becomes important. For two driven modes (Fig.~\ref{stationary_double}) the situation is similar. The appearance of the branches of the stationary states and in particular their coalescence near $Re=190$ suggests that the states are born out of a saddle-node bifurcation. And indeed, the eigenvalues as a function of $Re$ show two eigenvalues moving closer together and collapsing at zero for $Re=190.41$ (Fig.~\ref{saddle_node_bif}). However, these eigenvalues are not the leading ones, so that one set of states has three unstable eigenvalues, the other two unstable ones. It is thus a `saddle-node' bifurcation into unstable states. With increasing $Re$ more and more stationary states appear, partly through secondary bifurcations, partly through additional saddle-node bifurcations. Their number increases rapidly with Reynolds number (Fig.~\ref{proliferation}) and this increase goes in parallel with the increase in density of long lived states, Fig.~\ref{landscape}. The detailed structure of the bifurcation diagram is rather complex and has not yet been fully explored. We note here that the various stationary states may be grouped according to their symmetries introduced in section \ref{symmetries} and that we found only stationary states which belong to equivalent classes with four or eight members. The stationary states of the classes with four members are invariant under the transformation $T_1$ or $T_1T_2$. In addition, there are forward directed bifurcations generating two new branches with the same symmetry properties (eight or four member class) and inverse bifurcations of two branches belonging to eight member equivalent classes. We also found a backward directed bifurcation generating branches of an eight member class, which is born out of a four member class branch. The scenarios described above are marked in the bifurcation diagram Fig. \ref{stationary_single}. \section{Concluding remarks} \label{sec_conclusion} The few degrees of freedom shear model introduced here lies halfway between the simplest models of non-normality and full simulations. Its dynamics has turned out to be surprisingly rich. There are a multitude of bifurcations introducing new stationary states besides the laminar profile, there are secondary bifurcations, and the distribution of life times shows fractal structures on amazingly small scales. It seems that as one goes from the low-dimensional models\cite{TTRD1993,GG1994} via the present one to full simulations one notes not only an increase in numerical complexity but also the appearence of qualitatively new features \cite{EMS1998}. The simplest models with very few degrees of freedom focus on the non-normality of the linearized Navier-Stokes problem and emphasize the amplification of small perturbations. If the non-linearity is included a transition to another kind of dynamics, sometimes as simple as relaxation to a stationary point, is found \cite{BT1997}. Next in complexity are models like the one presented here that share with the few degree of freedom models the amplification and the transition but the additional degrees of freedom allow for chaos. When nonlinearities become important the dynamics does not settle to a fixed point or a limit cycle but continues irregularly for an essentially unpredictable time. The time is unpredictable because of the fractal life time distribution which seems to persist down to amazingly small scale: tiny variations in Reynolds number or amplitudes of the perturbation can cause major variations in life times. This fractal behaviour is the new quality introduced by the additional degrees of freedom. Indications for this behaviour are seen in the experiments by Mullin on pipe flow \cite{Mullin}. It is interesting to ask just how few degrees of freedom are necessary to obtain this behaviour. Reducing our model to the $T_1$ subspace gives one with just 9 degrees of freedom (comparable in number and flow behaviour to the ones of Waleffe \cite{wal2}) that still shows this fractal life time distributions. Further reduction, as in the four mode model of \cite{wal2}, seems to eliminate them. The full, spatially extended shear flows share essential features with the model but add new problems. Spatially resolved simulations of the present model \cite{wal98} as well as plane Couette flow with rigid-rigid boundary conditions \cite{Nagata1990,CB1997} show the occurence of additional stationary states at sufficiently high Reynolds number that are unstable. A novel and as yet unexplained feature in spatially extended plane Couette flow, which we believe to be connected to the high dimensionality of phase space, is the difference between Reynolds numbers where the first stationary states are born (about 125 in units of half the gap width and half the velocity difference) and the ones where experiments begin to see long lived states (about 300--350) \cite{experiments}. The fractal life time distributions have obvious similiarities to chaotic scattering \cite{EA1988,E1988,Ott}. Drawing on this analogy one would like to identify permanent structures in phase space away from the laminar profile that could sustain turbulence. This has partly been achieved by the search for stationary states. Many have been found but irritatingly only for Reynolds numbers above about 190 while long lived states seem to appear much earlier. The solution to this puzzle must be periodic states and indeed we have found a few periodic states in a symmetry reduced model at lower Reynolds numbers, close to the occurence of the first long lived states. This suggests that the dynamical system picture that long lived states have to be connected to persistent structures in phase space is tenable. There are several features of the model that can be studied further. In particular, quantitative characterizations of the fractal life time distribution, visualizations of the flow field, a detailed analysis of the primary and secondary bifurcation, an investigation of the dependence on the aspect ratio of the periodicity cell are required and look promising. We expect the lessons to be learned from this simple model to be useful in understanding the dynamics of full plane Couette and other shear flows. Work along these directions continues.
\section{Introduction} Complex neutrino mixing for 3 family Dirac neutrinos leads to CP and T violation effects in neutrino oscillations \cite{UNO}. In view of the vigorous experimental programme in this field, the study of CP violation becomes an interesting topic. The neutrino states of definite flavour $\alpha$, as generated by well defined weak interaction properties, are related to neutrino states of definite mass $m_k$ by \begin{equation} \nu_\alpha = \sum_k U_{\alpha k} \nu_k \label{eq:uno} \end{equation} \noindent where $U$ is the unitary mixing matrix which, for 3 families, depends on 3 mixing angles and 1 CP phase. If the "$\alpha$" state is prepared at $t = 0$, the probability amplitude that, at time $t$, it is manifested as the "$\beta$" state is \begin{equation} A (\alpha \rightarrow \beta ; t) = \sum_k U_{\alpha k} U_{\beta k}^* exp[- i E_k t] \label{eq:dos} \end{equation} We observe that the time-dependent amplitude contains the interference of different "$k$" terms, with different weak phases in $U_{\alpha k} \, U_{\beta k}^*$ and different oscillation phases governed by $E_k$. These ingredients are necessary and sufficient to generate CP violation in the oscillation probability. In Section 2 we discuss the CPT-invariance condition, together with the CP-odd and T-odd asymmetries. In Section 3 the CP-asymmetry for 3 family neutrino oscillation is considered and the conditions for a non-vanishing value are obtained. These results will lead to the need of long-base-line (LBL) experiments for CP studies. In Section 4 the CP-odd asymmetry is built in this case for hierarchical neutrino masses. These LBL experiments have to include, however, matter effects and in Section 5 we show that these matter effects are large and they constitute an undesired background for CP violation effects. Due to this fake phenomenon, Section 6 studies T-odd asymmetries which are free from this problem. Section 7 answers the question related to the possible Majorana character of neutrinos. Section 8 summarizes some conclusions and the outlook. \section{CPT, CP, T} >From Eq.~(\ref{eq:dos}) the requirement of CPT invariance leads to the amplitude for conjugate flavour states \begin{equation} A (\bar{\alpha} \rightarrow \bar{\beta}; t) = \sum_k U_{\alpha k}^* U_{\beta k} \, exp[-i E_k t] \label{eq:tres} \end{equation} \noindent so that we obtain the condition \begin{equation} CPT \Rightarrow A (\bar{\alpha} \rightarrow \bar{\beta}; t) = A^* (\alpha \rightarrow \beta; - t) \label{eq:cuatro} \end{equation} Eq.~(\ref{eq:cuatro}) will be assummed through this work. CP-invariance is the statement that the probabilities for the original transition and for its conjugate are equal, i.e., \begin{equation} CP \Rightarrow |A (\alpha \rightarrow \beta; t)|^2 = |A (\bar{\alpha} \rightarrow \bar{\beta}; t)|^2 \label{eq:cinco} \end{equation} T-invariance is the statement that the probabilities for the original transition and for its inverse are equal, i.e, \begin{equation} \begin{array}{ll} T \Rightarrow & | A (\alpha \rightarrow \beta ; t)|^2 = | A (\beta \rightarrow \alpha; t)|^2\\ & |A (\bar{\alpha} \rightarrow \bar{\beta}; t)|^2 = |A (\bar{\beta} \rightarrow \bar{\alpha}; t)|^2 \end{array} \label{eq:seis} \end{equation} From these results, we have the corollaries: i) CP, T Violation effects, i.e, the violation of Eq.~(\ref{eq:cinco}) , Eq.~(\ref{eq:seis}), respectively, can take place in Appearance Experiments only. For Disappearance experiments, $\beta = \alpha$, Eq.~(\ref{eq:seis}) is automatic and Eq.~(\ref{eq:dos}) implies \begin{equation} A^* (\alpha \rightarrow \alpha ; t) = A (\alpha \rightarrow \alpha; - t) \label{eq:siete} \end{equation} The combination of Eq.~(\ref{eq:siete}) and Eq.~(\ref{eq:cuatro}) leads to the verification of Eq.~(\ref{eq:cinco}), q.e.d. As a consequence, no CP or T violation effect can be manifested in reactor or solar neutrino experiments. ii) The (numerator of) CP-odd Asymmetry is given by \begin{equation} D_{\alpha \beta} \equiv |A (\alpha \rightarrow \beta; t)|^2 - |A (\bar{\alpha} \rightarrow \bar{\beta}; t)|^2 \label{eq:ocho} \end{equation} CPT-invariance implies $D_{\alpha \beta} = - D_{\beta \alpha}$ and the use of Eq.~(\ref{eq:dos}) and the Unitarity of the Mixing Matrix implies $\sum_{\beta \neq \alpha} D_{\alpha \beta} = 0$. These constraints lead to a unique $D$ for 3 flavours: \begin{equation} D_{e \mu} = D_{\mu \tau} = D_{\tau e} \label{eq:nueve} \end{equation} iii) The (numerators of) T-odd Asymmetries are given by \begin{equation} \begin{array}{l} T_{\alpha \beta} \equiv | A (\alpha \rightarrow \beta; t)|^2 - | A (\beta \rightarrow \alpha ; t)|^2 = | A (\alpha \rightarrow \beta ; t)|^2 - | A (\alpha \rightarrow \beta; -t)|^2 \\ \bar{T}_{\alpha \beta} \equiv | A (\bar{\alpha} \rightarrow \bar{\beta} ; t)|^2 - | A (\bar{\beta} \rightarrow \bar{\alpha} ; t)|^2 = | A (\bar{\alpha} \rightarrow \bar{\beta} ; t)|^2 - | A (\bar{\alpha} \rightarrow \bar{\beta}; -t)|^2 \\[2ex] \end{array} \label{eq:diez} \end{equation} \noindent where use of Eq.~(\ref{eq:dos}) has been made in the right-hand side. Eq.~(\ref{eq:diez}) leads to the important conclusion that \underline{$T_{\alpha \beta}, \bar{T}_{\alpha \beta}$ are odd functions of time}. One should be aware that Eq.~(\ref{eq:diez}) needs an hermitian Hamiltonian for the evolution of the system. In fact, the above conclusion is not valid for the $K^0 \bar{K}^0$ system. \section{CP-Asymmetry} The numerator Eq.~(\ref{eq:ocho}) of the CP-odd asymmetries can be calculated using Eq.~(\ref{eq:dos}) for the amplitudes. In the limit of ultrarelativistic neutrinos, one obtains \begin{equation} D_{\alpha \beta} = \sum_{k >j} I_{\alpha \beta; j k} \sin \frac{\Delta m_{kj}^2 L}{ 2 E} \label{eq:once} \end{equation} \noindent where $L \simeq t$ is the distance between the source and the detector, $E$ is the neutrino energy and $\Delta m_{kj}^2 \equiv m_k^2 - m_j^2$. The I's containing mixing angles and the CP phase are given by \begin{equation} I_{\alpha \beta ; jk} = 4 Im [U_{\alpha j} U_{\beta j}^* U_{\alpha k}^* U_{\beta k} ] \label{eq:doce} \end{equation} \noindent which show the rephasing invariance of the observables explicitly. Suppose that only the highest $m^2$-value is relevant for the neutrino oscillation experiment, assuming a hierarchy in neutrino masses. This statement, which can be considered as the definition of a short-base-line (SBL) experiment, means that the approximations \begin{equation} \left. \begin{array}{c} \Delta m^2 \simeq \Delta m_{31}^2 \simeq \Delta m^2_{32}\\ \frac{\Delta m^2_{21}}{2 E} L << 1 \end{array} \right\} \label{eq:trece} \end{equation} \noindent are fulfilled. In the limit of neglecting terms of order $\frac{ \Delta m_{21}^2}{2 E} L$, the asymmetry Eq.~(\ref{eq:once}) becomes \begin{equation} D_{\alpha \beta}^{(SBL)} \simeq (I_{\alpha \beta ; 13} + I_{\alpha \beta ; 23}) \sin \frac{\Delta m^2L}{2E} \label{eq:catorce} \end{equation} \noindent which vanishes due to the cyclic character of the I's: $I_{\alpha \beta; 23} = I_{\alpha \beta; 31} = - I_{\alpha \beta; 13}$. The lesson learnt from this limit is immediate: the 3 families have to participate ACTIVELY in order to generate a non-vanishing CP-odd observable. It is not enough to know that there are 3 non-degenerate neutrinos in Nature and the presence of mixing among all of them. $\Delta m_{21}^2$ has to participate. Furthermore, in order to generate a non-vanishing value for the I's one needs ALL the mixing angles and the unique CP phase different from zero \cite{DOS}. One thus concludes that a significant CP-odd asymmetry needs the consideration of neutrino oscillations in long-base-line (LBL) experiments \cite{TRES,CUATRO,CINCO,SEIS}. The meaning of this requirement is that both $\Delta m^2$'s, i.e., $\Delta m_{31}^2 \simeq \Delta m_{32}^2$ and $\Delta m_{21}^2$, have to be accessible. Two comments should be considered to soften the above conclusion: i) one can choose to keep terms of order $\frac{\Delta m_{21}^2}{2E} L <<1$ at the expense to search for appearance transitions with very low probability and enhance the CP-odd ratio which defines the asymmetry; ii) even if the value of $D_{\alpha \beta}$, and thus of the CP-odd asymmetry, vanishes under the conditions leading to Eq.~(\ref{eq:catorce}), the existence of a non-vanishing CP-phase can be inferred from CP-conserving observables if enough probabilities are measured. To have information on both $P (\nu_\mu \rightarrow \nu_\tau)$ and $P (\nu_e \rightarrow \nu_\tau)$ under controllable conditions, one probably needs the neutrino facility based on muon-storage-rings \cite{SIETE}. \section{CP effects in LBL experiments} Contrary to the conditions discussed before in Eq.~(\ref{eq:trece}), we assume in this Section that $L/E$ is such that \begin{equation} \left. \begin{array}{c} \frac{\Delta m_{31}^2 L}{2E} \sim \frac{\Delta m_{32}^2 L}{2 E} >> 1\\ \frac{\Delta m_{21}^2 L}{2 E} \sim 1 \end{array} \right\} \label{quince} \end{equation} The calculation of the appearance probabilities $\alpha \rightarrow \beta$ then gives \begin{equation} \begin{array}{l} P_{\nu_\alpha \rightarrow \nu_\beta} = | U^*_{\beta 1} U_{\alpha 1} + U^*_{\beta 2} U_{\alpha 2} exp (-i \frac{\Delta m_{21}^2 L}{2E}) |^2 + |U_{\beta 3}|^2 |U_{\alpha 3}|^2\\ P_{\bar{\nu}_\alpha \rightarrow \bar{\nu}_\beta} = | U_{\beta 1} U^*_{\alpha 1} + U_{\beta 2} U^*_{\alpha 2} exp (-i \frac{\Delta m_{21}^2 L}{2E}) |^2 + |U_{\beta 3}|^2 |U_{\alpha 3}|^2 \end{array} \label{eq:dieciseis} \end{equation} One notices that the heaviest neutrino contributes to these probabilities only through mixing without any oscillation: this term is CP-even. What is relevant for the CP asymmetry is the interference of two amplitudes $k = 1,2$ with diferent weak phases and different oscillation phases: in going to the CP transformed transition, the weak phase changes its sign whereas the oscillation phase remains the same. The difference of the two probabilities Eq.~(\ref{eq:dieciseis}) gives a CP-odd asymmetry \begin{equation} \left. \begin{array}{l} D_{\alpha \beta}^{(LBL)} \simeq I_{\alpha \beta} \sin \frac{\Delta m_{21}^2 L}{2 E}\\ I_{\alpha \beta} \equiv I_{\alpha \beta; 12} = 4 Im [U_{\alpha 1} U_{\beta 1}^* U_{\alpha 2}^* U_{\beta 2}] \end{array} \right\} \label{eq:diecisiete} \end{equation} It is immediate to realize that, for the 3-neutrino case, one has \begin{equation} I_{e \mu} = I_{\mu \tau} = I_{\tau e} \end{equation} Bilenky et al. \cite{SEIS} have used present exclusion plots for $\mu \rightarrow e$ and $\mu \rightarrow \tau $ transitions, together with amplitude and unitarity bounds, to find allowed values for $|I_{e \mu}|$ and $|I_{\mu \tau}|$. Maximum values of $10^{-2}$ for $|I_{e \mu|}$ and around $10^{-1}$ for $|I_{\mu \tau}|$ are accessible. \section{Matter effects in LBL experiments} LBL experiments, with source and detector at the earth surface, imply that neutrinos cross the earth in their travel. It is mandatory to discuss the matter effect \cite{OCHO,NUEVE,DIEZ} The effective Hamiltonians for neutrinos and antineutrinos are given in the flavour basis for 3 families: \begin{equation} \begin{array}{l} H_\nu = \frac{1}{2E} \left\{ U \left( \begin{array}{lll} m_1^2 && \\ & m^2_2 &\\ && m_3^3 \end{array} \right) U^+ + \left( \begin{array}{lll} a &&\\ & 0 & \\ & & 0 \end{array} \right) \right\}\\[2ex] H_{\bar{\nu}} = \frac{1}{2E} \left\{ U^* \left( \begin{array}{lll} m_1^2 && \\ & m^2_2 &\\ && m_3^2 \end{array} \right) U^T - \left( \begin{array}{lll} a &&\\ & 0 & \\ & & 0 \end{array} \right) \right\} \end{array} \label{eq:diecinueve} \end{equation} \noindent where the matter effect for constant density is given by the forward charged current interaction amplitude with electrons \begin{equation} a = G \sqrt{2} N_e 2 E \simeq 2.3 \times 10^{-4} eV^2 (\frac{\rho}{3 gcm^{-3}}) (\frac{E}{GeV}) \label{eq:veinte} \end{equation} \noindent and $N_e$ is the (number) density of electrons. It is interesting to build the dimensionless quantity \begin{equation} \frac{a L}{2 E} \simeq 0.58 \times 10^{-3} (\frac{L}{km}) \label{eq:veintiuno} \end{equation} \noindent which is E-independent! This is in contrast to the E-dependent oscillation quantity which, for $\Delta m^2$ in the range of the atmospheric neutrino oscillation solution, gives \begin{equation} \frac{\Delta m^2 L}{2E} \simeq 0.75 \times 10^{-2} (\frac{L}{km} ) (\frac{GeV}{E}) \label{eq:veintidos} \end{equation} For neutrino beams with energy $E \sim 10 \, GeV$, as envisaged by the FermiLab and CERN LBL experiments, the two quantities Eq.~(\ref{eq:veintiuno}) and Eq.~(\ref{eq:veintidos}) are comparable and one concludes that large matter effects are expected. The diagonalization of $H_\nu$ and $H_{\bar{\nu}}$ by unitary matrices $U'$ and $\bar{U}'$ leads to different eigenvalues $\tilde{m}_\nu^2$ and $\tilde{m}_{\bar{\nu}}^2$, respectively, \begin{equation} H_\nu = U' \frac{\tilde{m}_\nu^2}{2E} U'^+ \quad , \quad H_{\bar{\nu}} = \bar{U}^{'*} \frac{\tilde{m}_{\bar{\nu}}^2}{2E} \bar{U}'^T \label{eq:veintitres} \end{equation} One notes that the matter effect $a \neq 0$ provokes fake CP and CPT violation associated with the interaction with the asymmetric medium: even the simplest diagonal probability equality for $\nu_e$ and $\bar{\nu}_e$ is violated, $P_{\nu_e \rightarrow \nu_e} \neq P_{\bar{\nu}_e \rightarrow \bar{\nu}_e}$! With the neutrino mixing in the medium described by $U'$ and $\bar{U}'$, the CP violating I's of Eq. ~(\ref{eq:doce}) are replaced by \begin{equation} I_{\alpha \beta ; j k} \Rightarrow \left\{ \begin{array}{l} I'_{\alpha \beta ; jk} = 4 Im [U'_{\alpha j } U^{'*}_{\beta j} U^{'*}_{\alpha k} U'_{\beta k} ]\\ \tilde{I}'_{\alpha \beta ; jk} = 4 Im [\bar{U}'_{\alpha j} \bar{U}^{'*}_{\beta j} \bar{U}^{'*}_{\alpha k} \bar{U}'_{\beta k} ] \end{array} \right. \label{eq:veinticuatro} \end{equation} \noindent It is possible to prove \cite{SEIS} that a real diagonal matter term, as dictated by the Standard Model in Eq.~(\ref{eq:diecinueve}), implies the necessary and sufficient condition \begin{equation} I_{\alpha \beta; j k} = 0 \Longleftrightarrow I'_{\alpha \beta ; jk} = \bar{I}'_{\alpha \beta; jk} = 0 \label{eq:veinticinco} \end{equation} This means that the identification of non-vanishing $I', \bar{I}'$ in matter is still a true signal of CP-violation in Nature. The CP-odd asymmetries contain, however, additional terms which are an undesired background \begin{equation} \begin{array}{c} P_{\nu_\alpha \rightarrow \nu_\beta} = \sum_k |U'_{\beta k}|^2 |U'_{\alpha k}|^2 + 2 \sum_{k>j} Re [U'_{\alpha j} U^{'*}_{\beta j} U^{'*}_{\alpha k} U'_{\beta k}] \cos \frac{\Delta \tilde{m}^2_{\nu k j}}{2E} L\\ + \frac{1}{2} \sum_{k > j} I'_{\alpha \beta ;jk} \sin \frac{\Delta \tilde{m}^2_{\nu k j}}{2E} L\\ P_{\bar{\nu}_\alpha \rightarrow \bar{\nu}_\beta} = \sum_k |\bar{U}'_{\beta k}|^2 |\bar{U}'_{\alpha k}|^2 + 2 \sum_{k>j} Re [\bar{U}'_{\alpha j} \bar{U}^{'*}_{\beta j} \bar{U}^{'*}_{\alpha k} \bar{U}'_{\beta k}] \cos \frac{\Delta \tilde{m}^2_{\bar{\nu} k j}}{2E} L\\ - \frac{1}{2} \sum_{k > j} \bar{I}'_{\alpha \beta;jk} \sin \frac{\Delta \tilde{m}^2_{\bar{\nu} k j}}{2E} L \end{array} \label{eq:veintiseis} \end{equation} >From Eq.~(\ref{eq:veintiseis}) it is clear that, in matter, the transition probabilitites of neutrinos and antineutrinos are different even if CP is conserved, i.e., for $I' = \bar{I}' = 0$ one has $D_{\alpha \beta} \neq 0$. One would need the explicit separation of the odd functions $\sin \frac{\Delta \tilde{m}^2_{kj}} {2E}L$ in the oscillation probabilities to measure true CP violation effects. \section{T-odd asymmetries} Since the matter contribution to the effective neutrino and antineutrino Hamiltonians is real and the matter density is symmetric along the path of the neutrino beam in terrestrial LBL-experiments, matter effects are T-symmetric. A non-vanishing value of $T_{\alpha \beta}$ or $\bar{T}_{\alpha \beta}$, Eq.~(\ref{eq:diez}), in matter can only be due to a fundamental violation of T-invariance. It is straightforward to obtain these T-odd asymmetries \begin{equation} T_{\alpha \beta} = \sum_{k>j} I'_{\alpha \beta; jk} \sin \frac{\Delta \tilde{m}^2_{\nu k j }}{2E} L \quad ; \quad \bar{T}_{\alpha \beta} = \sum_{k>j} \bar{I}'_{\alpha \beta, jk} \sin \frac{\Delta \tilde{m}^2_{\bar{\nu} k j}}{2E} L \label{eq:veintisiete} \end{equation} One needs, however, the joint measurements of $\nu_\mu \rightarrow \nu_e$ and $\nu_e \rightarrow \nu_\mu$ which probably has to wait for neutrino factories in muon-storage-ring facilities. \section{Majorana Neutrinos} If neutrinos are Majorana particles, the main modification for charged current neutrino interaction is that the mixing matrix gets replaced \cite{ONCE} by \begin{equation} U^D \rightarrow U^M = U^D P \quad ; \quad P = \left( \begin{array}{ccc} e^{i \alpha_1} &&\\ & e^{i \alpha_2} &\\ && 1 \end{array} \right) \label{eq:veintiocho} \end{equation} \noindent with $U^D$ the conventional $U$ complex mixing matrix with 3 mixing angles and 1 CP-phase. One sees that 2 additional CP-phases come into the game. Some care is needed because CP-violation through these new phases means $\alpha_k \neq 0, \frac{\pi}{2}$! The new phases are, however, not operative as long as we consider Dirac flavour oscillations for neutrinos propagating through the Green function \\ $<0|T \{\psi (x) \bar{\psi} (0) \} | 0>$. Their manifestation would need the study of "neutrino-antineutrino" propagation mediated by the Green function $< 0|T \{\psi (x) \psi^T (0) \} | 0>$. Langacker et al. \cite{DOCE} have shown that this conclusion is valid not only for neutrino oscillations in vacuum but for flavour oscillations in matter too. The new phases associated with Majorana neutrinos are not seen in Flavour Oscillations in Matter. \section{Outlook} The responses to the questions posed in the Introduction are summarized now: - CP,T violation in neutrino oscillation is possible in Appearance Experiments, for mixing of 3 or more non-degenerate neutrinos. - Non-vanishing CP, T violating observables need the active participation of the 3 (different) masses, mixings and CP-phase. In particular, $\frac{E}{L}$ should be comparable to the smallest $\Delta m^2$. - In LBL experiments, there are large matter effects, inducing fake CP and CPT violation. The identification of true CP violation would need the explicit separation of odd functions of time. -Matter Effects are T-symmetric, so that a non-vanishing T-odd asymmetry in matter is still a signal of fundamental T-violation. - If neutrinos are Majorana particles, there is no change for Flavour Oscillations, either in vacuum or in matter. \section*{Acknowledgments} This work has been supported by the Spanish CICYT, under Grant AEN-96/1718.
\section{Introduction} Following Wigner~\cite{Wigner}, an elementary relativistic quantum system, an elementary particle with mass $m$ and spin $j$ is in the mathematical theory described by the space of an unitary irreducible representation (UIR) of the Poincar\'e group~${\mathcal P}$. From these UIR, the relativistic quantum fields are constructed \cite{Weinberg}. More complicated relativistic systems are described by direct sums of UIR (for ``towers'' of elementary particles) or by direct products of UIR (for combination of two or more elementary particles) \cite{Weinberg}. A direct product of UIR may be decomposed into a continuous direct sum (integral) of irreducible representations \cite{Joos,Macf}. The UIR are characterized by three invariants $(m^2, j,\textrm{sign}(p_0))$, where~$j$ represents the spin and the real number $m$ represents the mass of elementary particle (we restrict ourselves here to $\textrm{sign}(p_0)=+1$). The UIR of the Poincar\'e group ${\mathcal P}$ describe stable elementary particles (stationary systems). There is only a very small number of truly stable particles in nature and most relativistic (and also non-relativistic) quantum systems are decaying states (weakly or electromagnetically), or hadron resonances with an exponential decay law and finite lifetime $\tau_R=\frac{\hbar}{\Gamma}$ (in their rest frame) and a Breit-Wigner energy (at rest) distribution. The UIR of ${\mathcal P}$ therefore describe only a few of the relativistic quantum systems in nature; for the vast majority of elementary particles listed in the Particle Physics Booklet~\cite{PPB}, the UIR provide only a more or less approximate description. We want to present here a special class of (non-unitary) semi-group representations of $\cal P$ which describe quasistable relativistic particles. Phenomenologically, one always takes the point of view that resonances are autonomous quantum physical entities, and decaying particles are no less fundamental than stable particles. Stable particles are not qualitatively different from quasistable particles, but only quantitatively by a zero (or very small) value of $\Gamma$. Therefore both stable and quasistable states should be described on the same footing. This has been accomplished in the non-relativistic case, where a decaying state is described by a generalized eigenvector of the (self adjoint, semi-bounded) Hamiltonian with a complex eigenvalue $z_R=E_R-i\Gamma/2$~\cite{Bohm1} called Gamow vectors. The stable state vectors with real eigenvalues $E_S$ are the special case with $\Gamma=0$, i.e., $z_R=E_R-i\Gamma/2\rightarrow E_S$. In the standard Hilbert space formulation of quantum mechanics, such vectors do not exist and one had to employ the Rigged Hilbert Space (RHS) formulation of quantum mechanics. Dirac's bras and kets are, mathematically, generalized eigenvectors with real eigenvalues, and Gamow vectors are generalizations of Dirac kets. They are described by kets $\psi^G\equiv | z_R^-\rangle\sqrt{2\pi\Gamma}$ with complex eigenvalue $z_R=E_R-i\Gamma/2$, where $E_R$ and $\Gamma$ are respectively interpreted as resonance energy and width. Like Dirac kets, the Gamow kets are functionals of a Rigged Hilbert Space : \begin{equation} \Phi_+\subset{\mathcal H}\subset\Phi^\times_+: \,\,\,\,\,\, \psi^G\in\Phi^\times_+, \label{eq1} \end{equation} and the mathematical meaning of the eigenvalue equation $H^\times|z_R^-\rangle=(E_R-i\Gamma/2)|z_R^-\rangle$ is: \begin{equation} \langle H\psi|z_R^-\rangle \equiv \langle\psi|H^\times|z_R^-\rangle= z_R\langle\psi|z_R^-\rangle \,\,\,\,\, \textup{for all} \,\,\,\,\, \psi\in\Phi_+. \label{eq2} \end{equation} The conjugate operator $H^\times$ of the Hamiltonian $H$ is uniquely defined by the first equality in~(\ref{eq2}), as the extension of the Hilbert space adjoint operator $H^\dagger$ to the space of functionals $\Phi^\times_+$~\footnote{For (essentially) self adjoint $H$, $H^\dagger$ is equal to (the closure of) $H$; but we shall use the definition~(\ref{eq2}) also for unitary operators ${\mathcal U}$ where ${\mathcal U}^\times$ is the extension of ${\mathcal U}^\dagger$, but not of ${\mathcal U}$.}; on the space ${\mathcal H}$, the operators $H^\times$ and $H^\dagger$ are the same. The non-relativistic Gamow vectors have the following properties: \begin{enumerate} \item They have an asymmetric (i.e., $t\geq 0$ only) time evolution and obey then an exponential law: \begin{equation} \psi^G(t)=\textrm{e}_+^{-iH^\times t} |E_R-i\Gamma/2^-\rangle= \textrm{e}^{-iE_R t} \textrm{e}^{-\Gamma t/2} |E_R-i\Gamma/2^-\rangle, \,\,\,\, \textup{only for} \,\,\,\, t\geq0. \label{eq3} \end{equation} There is another Gamow vector $\tilde\psi^G=|E_R+i\Gamma/2^+\rangle\in\Phi^\times_-$, and another semigroup $\textrm{e}_-^{-iH^\times t}$ for $t\leq0$ in another RHS $\Phi_-\subset{\mathcal H}\subset\Phi_-^\times$ (with the same ${\mathcal H}$) with the asymmetric evolution \begin{equation} \tilde\psi^G(t)=\textrm{e}_-^{-iH^\times t} |E_R+i\Gamma/2^+\rangle= \textrm{e}^{-iE_R t} \textrm{e}^{\Gamma t/2} |E_R+i\Gamma/2^+\rangle, \,\,\,\, \textup{only for} \,\,\,\, t\leq0. \label{eq4} \end{equation} \item The $\psi^G$ ($\tilde\psi^G$) is derived as a functional at the resonance pole term located at $z_R=(E_R-i\Gamma/2)$ (at $z^*_R=(E_R+i\Gamma/2)$) in the second sheet of the analytically continued S-matrix. \item The Gamow vectors have a Breit-Wigner energy distribution \begin{equation} \langle^-E|\psi^G\rangle= i\sqrt{\frac{\Gamma}{2\pi}}\frac{1}{E-(E_R-i\Gamma/2)}, \,\,\,\, -\infty_{II}<E<\infty, \label{eq5} \end{equation} where $-\infty_{II}$ means that it extends to $-\infty$ on the second sheet of the S-matrix (whereas the standard Breit-Wigner extends to the threshold $E=0$). \end{enumerate} We want to present here a generalization of these non-relativistic Gamow vectors to the relativistic case. In the non-relativistic case the inclusion of the degeneracy quantum numbers of energy, i.e., the extension of the Dirac-Lippmann-Schwinger kets \begin{eqnarray} &&|E^\pm\rangle=|E\rangle+\frac{1}{E-H\pm i0}V|E\rangle =\Omega^\pm|E\rangle\nonumber\\[6pt] &&H|E^\pm\rangle= E|E^\pm\rangle; \,\,\,\, (H-V)|E\rangle=E|E\rangle \label{eq6} \end{eqnarray} to the basis of the whole Galilei group is trivial. For the two particle scattering states (direct product of two irreducible representations of the Galilei group~\cite{ref5}) one uses eigenvectors of angular momentum $(jj_3)$ for the relative motion and total momentum ${\bbox p}$ for the center of mass motion. Thus \begin{equation} |E^{\textrm{\scriptsize tot}}{\bbox p}jj_3(l,s) \,\,^\pm\rangle= |{\bbox p}\rangle\otimes|Ejj_3\,\,^\pm\rangle \label{eq7} \end{equation} where $E^{\textrm{\scriptsize tot}}=\frac{{\bbox p}^2}{2m}+E$ (the Hamiltonian in~(\ref{eq6}) is $H=H^{\textrm{\scriptsize tot}}-\frac{{\bbox P}^2}{2m}$). The center-of-mass motion is usually separated by transforming to the center-of-mass frame, and there one uses in~(\ref{eq6}) \begin{equation} |{\bbox p}=\bbox{0}\rangle\otimes|Ejj_3\,\,^\pm\rangle =|E,\bbox{p}=0,jj_3\,\,^\pm\rangle= |E\,\,^\pm\rangle. \label{eq8} \end{equation} The generalized eigenvectors \begin{equation} |Ejj_3\,\,^\mp\rangle\in\Phi^\times_\pm\supset{\mathcal H}\supset\Phi_\pm \label{eq9} \end{equation} with \begin{equation} H^\times|Ejj_3\,\,^\pm\rangle= E|Ejj_3\,\,^\pm\rangle,\quad 0\leq E <\infty, \label{eq10} \end{equation} where $E$ runs along the cut on the positive real axis of the 1-st sheet of the $j$-th partial S-matrix, are the scattering states. The proper eigenvectors of~(\ref{eq9}) with $E=-|E_n|$ at the poles on the negative real axis on the 1-st sheet are the bound states $|E_njj_3\rangle\in\Phi$. By the Galilei transformation one can transform these vectors to arbitrary momentum ${\bbox p}$; $E$ and ${\bbox p}$ are not intermingled by Galilei transformations. To obtain the non-relativistic Gamow kets one analytically continues the Dirac-Lippmann-Schwinger ket~(\ref{eq6}) into the second sheet of the $j$-th partial S-matrix to the position of the resonance pole $|z_R=E_R-i\Gamma/2,jj_3\ ^-\rangle$ and obtains the following representation~\cite{Bohm1}: \begin{equation} |z_R=E_R-i\Gamma/2,jj_3\ ^-\rangle= \frac{i}{2\pi}\int_{-\infty_{II}}^{+\infty}dE |Ejj_3\ ^-\rangle\frac{1}{E-z_R}. \end{equation} A Galilei transformation can boost this Gamow ket to any \textit{real} momentum~$\bbox{p}$ \[ |{\bbox p},z_R,jj_3\ ^-\rangle= {\mathcal U}(\bbox{p})|\bbox{0}\rangle\otimes|z_Rjj_3\ ^-\rangle. \] However, complex momenta cannot be obtained in this way since the Galilei transformations commute with the intrinsic energy operator~$H$. In the relativistic case the Lorentz transformation -- in particular Lorentz boosts -- intermingle energy $E^{\textrm{\scriptsize tot}}=p^0$ and momenta $p^m$, $m=1,2,3$. Thus complex energy or complex mass also leads to complex momenta. This has led in the past to consider complicated complex momentum representations of the Poincar\'e group ${\mathcal P}$. To restrict this unwieldy set of Poincar\'e group representations we will consider a special class of ``minimally complex'' irreducible representations of ${\mathcal P}$ to describe relativistic resonances and decaying elementary particles. Our construction will also lead to complex momenta $p^\mu$, but in our case the momenta will be ``minimally complex'' in such a way that the 4-velocities $\hat{p}_\mu\equiv\frac{p_\mu}{m}$ remain real. This construction was motivated by a remark of D.~Zwanziger~\cite{zwanzi} and is based on the fact that the 4-velocity eigenvectors $|\hat{\bbox{p}}j_{3}(mj)\rangle$ furnish as valid a basis for the representation space of ${\mathcal P}$ as the usual Wigner basis of momentum eigenvectors $|\bbox{p}j_{3}(m,j)\rangle$. This means every state of an UIR $(m,j)$, ($\phi\in\Phi\subset{\cal H}(m,j)\subset\Phi^{\times}$, where $\Phi$ denotes the space of well-behaved vectors and $\Phi^{\times}$ the space of kets for the Hilbert space ${\cal H}(m,j)$ of an UIR), can be written according to Dirac's basis vector decomposition as \begin{equation} \label{2.15a} \phi=\sum_{j_{3}}\int \frac{d^{3}\hat{p}}{2\hat{p}^{0}} |\hat{\bbox{p}},j_{3}\rangle\langle j_{3},\hat{\bbox{p}}|\phi\rangle \end{equation} where we have chosen the invariant measure \begin{equation} \label{measurehat} d\mu(\hat{\bbox p}) = \frac{d^{3}\hat{p}}{2\hat{p}^{0}} = {\frac{1}{m^{2}}} \, {\frac{d^{3}p}{2 E({\bbox p})}}, \,\,\,\,\,\, \hat{p}^{0} = \sqrt{1+\hat{\bbox p}^{2}} \, . \end{equation} As a consequence of (\ref{measurehat}), the $\delta$-function normalization of these velocity-basis vectors is \begin{equation} \langle \xi , \hat{\bbox p}\,|\,\hat{\bbox p}', \xi' \rangle = 2 \hat{p}^{0} \delta^{3}(\hat{\bbox p}-\hat{\bbox p}') \, \delta_{\xi \xi'}\\ = 2 p^{0} m^{2} \delta^{3} (\bbox{p}-\bbox{p'})\, \delta_{\xi \xi'} \, . \label{normalizationhat} \end{equation} Here, $|\hat{\bbox{p}},j_{3}\rangle\in \Phi^{\times}$ are the eigenkets of the 4-velocity operator $\hat{P}_{\mu}=P_{\mu}M^{-1}$ and $\phi_{j_{3}}(\hat{\bbox{p}})=\langle j_{3}\hat{\bbox{p}}|\phi\rangle$ represents the 4-velocity distribution of the vector $\phi$. The 4-velocity eigenvectors are often more useful for physical reasoning, because 4-velocities seem to fulfill to rather good approximation ``velocity super-selection rules'' which the momenta do not~\cite{ref7}. The relativistic Gamow vectors will therefore be defined, not as momentum eigenvectors, but as 4-velocity eigenvectors in the direct product space of UIR spaces for the decay products of the resonances $R$. We want to obtain the relativistic Gamow vectors from the pole term of the relativistic S-matrix in complete analogy to the way the non-relativistic Gamow vectors were obtained~\cite{Bohm1}. In the absence of a vector space description of a resonance, we shall also in the relativistic theory define the unstable particle by the pole of the analytically continued partial S-matrix with angular momentum $j=j_R$ at the value ${\mathsf{s}}={\mathsf{s}}_R\equiv(M_R-i\Gamma/2)^2$ of the invariant mass square variable (Mandelstam variable) ${\mathsf{s}}=(p_1+p_2+\cdots)^2=E_R^2-{\bbox p}_R^2$, where $p_1$, $p_2$,\ldots are the momenta of the decay products of $R$~\cite{Eden}. This means that the mass $M_R$ and lifetime $\hbar/\Gamma_R$ or the complex invariant mass $w_{R}=(M_{R}-i\Gamma/2)=\sqrt{{\mathsf{s}}_{R}}$, in addition to spin $j_R$, are the intrinsic properties that define a quasistable relativistic particle~\footnote{Conventionally and equivalently one often writes \[ {\mathsf{s}}_R\equiv M_\rho^2-iM_\rho\Gamma_\rho= M_R^2\left(1-\frac{1}{4} \left(\frac{\Gamma_R}{M_R}\right)^2\right)-iM_R\Gamma_R \] and calls $M_\rho=M_R\sqrt{1-\frac{1}{4}\left(\frac{\Gamma_R}{M_R}\right)^2}$ the resonance mass and $\Gamma_\rho=\Gamma_R\left(1-\frac{1}{4} \left(\frac{\Gamma_R}{M_R}\right)^2\right)^{-1/2}$ its width. For the $\rho$ meson $\left(\frac{\Gamma_R}{M_R}\right)^2\approx0.03$ and for most other resonances it is an order of magnitude or more smaller than this.}. In order to make the analytic continuation of the partial S-matrix with angular momentum $j$, we need the angular momentum basis vectors \begin{eqnarray} &|\hat{\bbox p}j_3(wj)\rangle= \int\frac{d^3\hat{p}_1}{2\hat{E}_1} \frac{d^3\hat{p}_2}{2\hat{E}_2} |\hat{\bbox p}_1\hat{\bbox p}_2[m_1m_2]\rangle \langle\hat{\bbox p}_1\hat{\bbox p}_2[m_1m_2]|\hat{\bbox p}j_3(wj)\rangle \label{eq11}\\ &\mbox{for any $(m_1+m_2)^2\leq w^2<\infty$ \,\,\,\,$j=0,1,\ldots$} \nonumber \end{eqnarray} in the direct product space of the decay products of the resonance $R$ \begin{equation} {\mathcal H}\equiv {\mathcal H}(m_1,0)\otimes {\mathcal H}(m_2,0)= \int_{(m_1+m_2)^2}^{\infty} d{\mathsf{s}} \sum_{j=0}^{\infty}\oplus{\mathcal H}({\mathsf{s}},j), \label{eq12} \end{equation} where ${\mathsf{s}}=w^{2}$, the Mandelstam variable defined above. For simplicity, we have assumed here that there are two decay products, $R\rightarrow\pi_1+\pi_2$ with spin zero, described by the irreducible representation spaces ${\mathcal H}^{\pi_i}(m_i,s_i=0)$~\footnote{Though our discussions apply with obvious modifications to the general case of \[ 1+2+3+\cdots\rightarrow R_i\rightarrow 1^\prime+ 2^\prime+ 3^\prime+\cdots, \] these generalizations lead to enormously more complicated equations.} of the Poincar\'e group ${\mathcal P}$. The kets $|\hat{\bbox p}j_3(wj)\rangle$ are eigenvectors of the 4-velocity operators \begin{equation} \hat{P}_\mu=(P^1_\mu+P^2_\mu)M^{-1},\,\,\,\, M^2=(P^1_\mu+P^2_\mu)({P^1}^\mu+{P^2}^\mu) \label{eq13} \end{equation} with eigenvalues \begin{equation} \hat{p}^\mu=\left( \begin{array}{c} \hat{E}=\frac{p^0}{w}=\sqrt{1+\hat{\bbox p}^2}\\ \hat{\bbox p}=\frac{\bbox p}{w} \end{array} \right) \,\,\,\,\,\,\mbox{and}\,\,\,\, w^2={\mathsf{s}}. \end{equation} In here $\hat{P}^i_\mu$ are the 4-velocity operators in the one particle spaces ${\mathcal H}^{\pi_i}(m_i,s_i)$ with eigenvalues $\hat{p}^i_\mu=\frac{p^i_\mu}{m_i}$. To obtain the Clebsch-Gordan coefficients $\langle \hat{\bbox p}_{1}\hat{\bbox p}_{2}[m_1,m_2]|\hat{\bbox p}j_{3}(wj)\rangle$ in~(\ref{eq11}), one follows the same procedure as given in the classic papers~\cite{Wight,Joos,Macf} for the Clebsch-Gordan coefficients for the Wigner (momentum) basis. This has been done in~\cite{Ref12}. The result is \begin{eqnarray} &\langle\hat{\bbox p}_1\hat{\bbox p}_2[m_1,m_2]|\hat{\bbox p}j_3(wj)\rangle= 2\hat{E}(\hat{\bbox p})\delta^3(\bbox{p}-\bbox{r})\delta(w-\epsilon) Y_{jj_3}({\bbox e})\mu_j(w^2,m_1^2,m_2^2) \label{eq15}\\ &\mbox{with}\,\,\,\epsilon^2=r^2=(p_1+p_2)^2, \,\,\,r=p_1+p_2,\nonumber \end{eqnarray} where $\mu_j(w^2,m_1^2,m_2^2)$ is a function that fixes the $\delta$-function ``normalization'' of $|\hat{\bbox p}j_3(wj)\rangle$. The unit vector ${\bbox e}$ in~(\ref{eq15}) is chosen to be in the c.m.\ frame the direction of $\hat{\bbox p}_1^{\textrm{\scriptsize cm}} =-\frac{m_2}{m_1}\hat{\bbox p}_2^{\textrm{\scriptsize cm}}$. The normalization of the basis vectors~(\ref{eq11}) is chosen to be \begin{eqnarray} &\langle\hat{\bbox p}^\prime j^\prime_3(w^\prime j^\prime) |\hat{\bbox p}j_3(wj)\rangle= 2\hat{E}(\hat{\bbox p})\delta(\hat{\bbox p}^\prime-\hat{\bbox p}) \delta_{j_3^\prime j_3}\delta_{j^\prime j} \delta({\mathsf{s}}-{\mathsf{s}}^\prime) \label{eq16}\\ &\mbox{where}\,\,\,\, \hat{E}(\hat{\bbox p})=\sqrt{1+\hat{\bbox p}^2}=\frac{1}{w} \sqrt{w^2+\bbox{p}^2}\equiv \frac{1}{w}E({\bbox p},w).\nonumber \end{eqnarray} This determines the weight function $\mu_j(w^2,m_1^2,m_2^2)$ to \begin{equation} \left| \mu_j(w^2,m_1^2,m_2^2)\right|^{2}= \frac{2m_1^2m_2^2w^2}{\sqrt{\lambda(1,(\frac{m_1}{w})^2,(\frac{m_2}{w})^2)}} \label{eq17} \end{equation} where $\lambda$ is defined by~\cite{Wight} \begin{equation} \lambda(a,b,c)=a^2+b^2+c^2-2(ab+bc+ac). \label{eq18} \end{equation} The basis vectors~(\ref{eq11}) are the eigenvectors of the free Hamiltonian $H_0=P^1_0+P^2_0$ \begin{equation} H_0^{\times}|\hat{\bbox p}j_3(wj)\rangle= E|\hat{\bbox p}j_3(wj)\rangle,\,\,\,\,\,\,\, E=w\sqrt{1+\hat{\bbox{p}}^2}. \label{eq19} \end{equation} The Dirac-Lippmann-Schwinger scattering states are obtained, in analogy to~(\ref{eq6}) (cf. also~\cite{Weinberg} sect.~3.1) by: \begin{equation} |\hat{\bbox p}j_3(wj)^\pm\rangle= \Omega^\pm|\hat{\bbox p}j_3(wj)\rangle \label{eq20} \end{equation} where $\Omega^\pm$ are the M{\o}ller operators. For the basis vectors at rest, (\ref{eq20}) is given by the solution of the Lippmann-Schwinger equation \begin{equation} |\bbox{0}j_3(wj)^\pm\rangle= \left(1+ \frac{1}{w-H\pm i\epsilon}V \right) |\bbox{0}j_3(wj)\rangle. \label{eq21} \end{equation} They are eigenvectors of the exact Hamiltonian $H=H_0+V$ \begin{equation} H|\bbox{0}j_3(wj)^\pm\rangle= \sqrt{{\mathsf{s}}}|\bbox{0}j_3(wj)^\pm\rangle, \,\,\,\, (m_1+m_2)^2\leq {\mathsf{s}}<\infty. \label{eq22} \end{equation} The vectors $|\hat{\bbox p}j_3(wj)^\pm\rangle$ are obtained from the basis vectors at rest $|\bbox{0}j_3(wj)^\pm\rangle$ by the boost (rotation-free Lorentz transformation) ${\mathcal U}(L(\hat{p}))$ whose parameters are the 4-velocities $\hat{p}^{\mu}$. The generators of the Lorentz transformations are the interaction-incorporating observables \begin{equation} P_0=H, \,\,\,\, P^m, \,\,\,\, J_{\mu\nu}, \label{eq23} \end{equation} i.e., the exact generators of the Poincar\'e group (\cite{Weinberg} sec.\ 3.3). These vectors $|\hat{\bbox p}j_{3}(wj)^{\pm}\rangle$ in (\ref{eq20}), or $|\bbox{0}j_{3}(wj)^{\pm}\rangle$ in (\ref{eq21}) when boosted by ${\mathcal U}(L(\hat{p}))$, which for the fixed value $[jw]$ span an irreducible representation space of the Poincar\'e group with the ``exact generators'', will be used for the definition of the relativistic Gamow vectors. The relativistic Gamow kets are obtained by analytically continuing the Dirac-Lippmann-Schwinger kets~(\ref{eq21}) or~(\ref{eq20}) into the second (or higher) sheet of the $j_R$-th partial wave S-matrix $S_{j_R}({\mathsf{s}})$ to the position of the resonance pole at ${\mathsf{s}}_R=(M_R-i\Gamma/2)^2$. This can be done for any value of $\hat{\bbox p}$ (and $j_3$), e.g., for $\hat{\bbox p}={\bbox 0}$. In complete analogy to the non-relativistic case one obtains the relativistic kets \begin{equation} |\hat{\bbox p}j_3({\mathsf{s}}_Rj_R)^{-}\rangle= \frac{i}{2\pi} \int_{-\infty_{II}}^{+\infty_{II}}d{\mathsf{s}} |\hat{\bbox p}j_3({\mathsf{s}} j_R)^{-} \rangle\frac{1}{{\mathsf{s}}-{\mathsf{s}}_R}. \label{eq24} \end{equation} The Lorenz transformations $\Lambda$ are represented by unitary operators ${\mathcal U}(\Lambda)$: \begin{equation} {\mathcal U}(\Lambda)|\hat{\bbox p}j_3({\mathsf{s}}_R j_R)^{-}\rangle= \sum_{j_3^\prime}D^{j_{R}}_{j_{3}^{\prime}j_{3}} ({\mathcal R}(\Lambda,\hat{p})) |\bbox{\Lambda}\hat{\bbox{p}}j_3^\prime({\mathsf{s}}_R j_R)^{-}\rangle, \label{eq25} \end{equation} where ${\mathcal R}(\Lambda,\hat{p})=L^{-1}(\Lambda \hat{p})\Lambda L(\hat{p})$ is the Wigner rotation. In particular for the rotation free Lorentz boost \begin{equation} {\mathcal U}(L(\hat{p})) |\hat{\bbox p}={\bbox 0},j_3({\mathsf{s}}_R j_R)^{-}\rangle= |\hat{\bbox p}j_3({\mathsf{s}}_R j_R)^{-}\rangle \label{eq26} \end{equation} because the boost is a function of the real $\hat{p}^\mu$ and not of the complex $p^\mu$: \begin{equation} L^\mu_{\hphantom{\mu}\nu}= \left( \begin{array}{cc} \frac{p^0}{m}&-\frac{p_n}{m}\\ \frac{p^k}{m}&\delta^k_n-\frac{\frac{p^k}{m}\frac{p_n}{m}}{1+\frac{p^0}{m}} \end{array} \right), \,\,\,\, L(\hat{p}) \left( \begin{array}{c} 1\\ 0\\ 0\\ 0 \end{array} \right) =\hat{p}. \label{eq27} \end{equation} The relativistic Gamow kets~(\ref{eq24}) are generalized eigenvectors of the invariant mass squared operator $M^2=P_\mu P^\mu$ with eigenvalue ${\mathsf{s}}_R$ \begin{equation} \langle\psi^-|M^2|\hat{\bbox p}j_3({\mathsf{s}}_Rj_R)^{\,\,\,\,-}\rangle= {\mathsf{s}}_R\langle\psi^-|\hat{\bbox p}j_3 ({\mathsf{s}}_Rj_R)^{\,\,\,\,-}\rangle \quad \text{for every } \psi^-\in\Phi_+\subset {\mathcal H}\subset\Phi^\times_+. \label{eq28} \end{equation} To prove~(\ref{eq28}) from~(\ref{eq24}), and also to obtain~(\ref{eq24}) from the pole term of the S-matrix one needs to use the properties of the Hardy class space $\Phi_+$. The continuous linear combinations of the 4-velocity kets (\ref{eq24}) with an arbitrary 4-velocity distribution function $\phi_{j_{3}}(\hat{\bbox{p}})\in {\cal S}$ (Schwartz space), \[ \psi^{\rm G}_{j_{R}{\mathsf{s}}_{R}}= \sum_{j_{3}}\int \frac{d^{3}\hat{p}}{2\hat{p}^{0}} |\hat{\bbox{p}}j_{3}({\mathsf{s}}_{R},j_{R})^{-} \rangle \phi_{j_{R}}(\hat{\bbox{p}}), \] also represents relativistic Gamow states with the complex mass ${\mathsf{s}}_{R}=(M_{R}-i\Gamma/2)^{2}$ and have the representation~(\ref{eq24}). This then implies, much like in the RHS theory of non-relativistic Gamow vectors, that the time translation of the decaying state is represented by a semigroup, e.g., at rest, \begin{equation} \textrm{e}^{-iH\tau} |\hat{\bbox p}=\bbox{0},j_3({\mathsf{s}}_R j_R)^{-}\rangle= \textrm{e}^{-im_R\tau} \textrm{e}^{-\Gamma\tau/2} |\hat{\bbox p}=\bbox{0},j_3({\mathsf{s}}_R j_R)^{-}\rangle \,\,\,\, \text{for }\tau\geq0\,\,\,\textrm{only} \label{eq29} \end{equation} where $\tau$ is time in the rest system. Thus relativistic Gamow states are representations of ${\mathcal P}$ with spin $j_R$ and complex mass ${\mathsf{s}}_R=(M_{R}-i\Gamma/2)^2=m_\rho^2-im_\rho\Gamma_\rho$, for which the Lorentz subgroup is unitarily represented. They are obtained from the resonance pole of the relativistic partial S-matrix $S_{j_R}({\mathsf{s}})$, have an exponential time evolution at rest with lifetime $\hbar/\Gamma$ and have -- due to their association to the S-matrix pole and to the Hardy class spaces -- a Breit-Wigner energy distribution \begin{equation} a_j=\frac{Bm\Gamma}{{\mathsf{s}}-(m-i\frac{\Gamma}{2})^2}, \,\,\,\, -\infty_{II}<{\mathsf{s}}<+\infty_{II}. \label{eq30} \end{equation} These are all features which one may welcome or accept for states that are to describe relativistic resonances. In addition, they have a semigroup time evolution expressing a fundamental quantum mechanical arrow of time. We gratefully acknowledge support of the Welch Foundation and of CoNaCyT (Mexico).
\section{#1}} \renewcommand{\theequation}{\thesection.\arabic{equation}} \sect{Introduction\label{intro}} The introduction of quantum algebras by Jimbo \cite{ji} and Drinfeld \cite{d1} lead to many remarkable developments in diverse areas of mathematical physics. One such was in the field of knot theory whereby a connection between the Yang-Baxter equation and the braid group was quickly established. The quantum algebras, being examples of quasi-triangular Hopf algebras, provide very systematic means to find solutions of the Yang-Baxter equation which in turn gives rise to representations of the braid group. Through a Markov trace formulation defined on each braid group representation, an invariant polynomial can then be computed for the knot or link associated with the closure of the braid \cite{r,t,wda,zgb}. Extensions to accomodate the case of quantum superalgebras can be found in \cite{lgz,z}. Around the same time as the appearance of quantum algebras was Jones's discovery of a new polynomial invariant \cite{jo}, an evaluation of which may be undertaken through the simplest quantum algebra $U_q(sl(2))$ in its minimal (two-dimensional) representation. After this breakthrough researchers proceeded to obtain generalizations with the notable examples being the HOMFLY \cite{h} and Kauffman \cite{k} invariant polynomials. What soon became apparent was that the series of link polynomials associated with the fundamental representations of the (non-exceptional) quantum algebras and superalgebras coincided with the two-variable invariants developed in the wake of the discovery of Jones. More precisely, the invariants associated with the fundamental representations of the $U_q(gl(m|n))$ (which includes $U_q(gl(n))$) series belong to the class of HOMFLY invariants while those of the $U_q(osp(m|2n))$ (including both $U_q(o(m))$ and $U_q(sl(2n))$) series are of the Kauffman invariant type \cite{t,z}. It is important to emphasize, however, that by going to higher representations new results are obtainable. In particular, the type I quantum superalgebras consisting of $U_q(gl(m|n))$ and $U_q(osp(2|2n))$ admit one-parameter families of typical representations which give rise to two-variable link invariants in a natural way \cite{dkl,glz,lg}. The work of Reshetikhin and Turaev \cite{rt} introduced further the notion of a ribbon Hopf algebra as a particular type of quasi-triangular Hopf algebra. All the quantum algebras fall into the class of ribbon Hopf algebras. The algebraic properties of ribbon Hopf algebras is such that an extension to produce invariants of oriented tangles is permissible. A tangle diagram is analogous to a link diagram with the possibilty of having free ends. An associated invariant takes the form of a tensor operator acting on a product of vector spaces. As a generalization of Hopf algebras Drinfeld proposed the concept of quasi-Hopf algebras \cite{Dri90} whereby co-associativity of the co-algebra structure is not assumed. Any quasi-Hopf algebra generally belongs to an equivalence class where each member is related to the others by twisting \cite{Dri90}. The potential for applications of these structures in the study of integrable systems is vast. They underly elliptic quantum algebras \cite{Enr97,Fel95,Fod94,Fro97,Jim97,Zha98} which play an important role in obtaining solutions to the dynamical Yang-Baxter equations \cite{Arn97,Bab96} and also twisting lies at the core of the construction of multiparametric quantum spin chains \cite{flr}. In the context of knot theory, Altschuler and Coste \cite{ac} have identified the corresponding ribbon quasi-Hopf algebras as the necessary underlying algebraic structure to study tangle invariants (including closed link invariants). Here we wish to make two important observations to this field of study. First, we will show that the class of ribbon quasi-Hopf algebras is closed under twisting; i.e a twisted ribbon quasi-Hopf algebra is again a ribbon quasi-Hopf algebra. Secondly, we assert that the link polynomials computed from any finite dimensional representation of a quasi-Hopf algebra are invariant with respect to twisting. Importantly, this implies that link polynomials obtained from twisting the usual quantum algebras give us nothing new. In this respect, one cannot find twist generalizations of the HOMFLY and Kauffman invariants. For a very special class of twists this result has already been noted by Reshetikhin \cite{r1}, in which case the twisted quantum algebra is again a ribbon Hopf algebra. Here we will prove the twisting invariance in full generality. The paper is structured as follows. We begin by presenting the definition of a quasi-Hopf algebra. Next we show how representations of the braid group are obtained from a representation of any quasi-Hopf algebra. The third section deals with defining an appropriate Markov trace on each braid group element which then affords a means to obtain a link invariant. Finally, we demonstrate that the Markov trace is invariant under any twisting. \sect{Quasi-Hopf Algebras} Let us briefly recall the defining relations for quasi-Hopf algebras \cite{Dri90}. \begin{Definition}\label{quasi-bi}: A quasi-Hopf algebra is a unital associative algebra $A$ over a field $K$ which is equipped with algebra homomorphisms $\epsilon: A\rightarrow K$ (co-unit), $\Delta: A\rightarrow A\otimes A$ (co-product), an invertible element $\Phi\in A\otimes A\otimes A$ (co-associator), an algebra anti-homomorphism $S: A\rightarrow A$ (anti-pode) and canonical elements $\alpha,~\beta\in A$, satisfying \begin{eqnarray} && (1\otimes\Delta)\Delta(a)=\Phi^{-1}(\Delta\otimes 1)\Delta(a)\Phi,~~ \forall a\in A,\label{quasi-bi1}\\ &&(\Delta\otimes 1\otimes 1)\Phi \cdot (1\otimes 1\otimes\Delta)\Phi =(\Phi\otimes 1)\cdot(1\otimes\Delta\otimes 1)\Phi\cdot (1\otimes \Phi),\label{quasi-bi2} \label{pent}\\ &&(\epsilon\otimes 1)\Delta=1=(1\otimes\epsilon)\Delta,\label{quasi-bi3}\\ &&(1\otimes\epsilon\otimes 1)\Phi=1,\label{quasi-bi4}\\ && m\cdot (1\otimes\alpha)(S\otimes 1)\Delta(a)=\epsilon(a)\alpha,~~~\forall a\in A,\label{quasi-hopf1}\\ && m\cdot (1\otimes\beta)(1\otimes S)\Delta(a)=\epsilon(a)\beta,~~~\forall a\in A, \label{quasi-hopf2}\\ && m\cdot (m\otimes 1)\cdot (1\otimes\beta\otimes\alpha)(1\otimes S\otimes 1)\Phi^{-1}=1,\label{quasi-hopf3}\\ && m\cdot(m\otimes 1)\cdot (S\otimes 1\otimes 1)(1\otimes\alpha\otimes \beta)(1\otimes 1\otimes S)\Phi=1.\label{quasi-hopf4} \end{eqnarray} \end{Definition} Here $m$ denotes the usual product map on $A$: $m\cdot (a\otimes b)=ab,~ \forall a,b\in A$. Note that since $A$ is associative we have $m\cdot(m\otimes 1)=m\cdot (1\otimes m)$. For all elements $a,b\in A$, the antipode satisfies \begin{equation} S(ab)=S(b)S(a). \end{equation} The equations (\ref{quasi-bi2}), (\ref{quasi-bi3}) and (\ref{quasi-bi4}) imply that $\Phi$ also obeys \begin{equation} (\epsilon\otimes 1\otimes 1)\Phi=1=(1\otimes 1\otimes\epsilon)\Phi.\label{e(phi)=1} \end{equation} Applying $\epsilon$ to definition (\ref{quasi-hopf3}, \ref{quasi-hopf4}) we obtain, in view of (\ref{quasi-bi4}), $\epsilon(\alpha)\epsilon(\beta)=1$. By applying $\epsilon$ to (\ref{quasi-hopf1}), we have $\epsilon(S(a))=\epsilon(a),~\forall a\in A$. A distinguishing feature of quasi-Hopf algebras is that they are in general not co-associative; i.e $$(1\otimes\Delta)\cdot\Delta\neq (\Delta\otimes 1)\cdot\Delta. $$ Thus for a given co-product the action extended to the $n$-fold tensor product space is not uniquely determined. Throughout we will adopt the convention to define a left co-product $\Delta_L$ which acts on the tensor algebra $A^{\ot n}$ according to $$\Delta_L(a\ot b\ot ....\ot c)=\Delta(a)\ot b\ot ....\ot c. $$ We then recursively define the action \begin{equation} \Delta^{(n)}=\Delta_L.\Delta^{(n-1)} \label{act} \end{equation} with $\Delta^{(1)}=\Delta,\,\Delta^{(0)}=I$. The category of quasi-Hopf algebras is invariant under a kind of gauge transformation known as twisting. Let $(A,\Delta,\epsilon,\Phi)$ be a quasi-Hopf algebra, with $\alpha,\beta, S$ satisfying (\ref{quasi-hopf1})-(\ref{quasi-hopf4}), and let $F\in A\otimes A$ be an invertible element satisfying the co-unit properties \begin{equation} (\epsilon\otimes 1)F=1=(1\otimes \epsilon)F.\label{e(f)=1} \end{equation} Throughout we set \begin{eqnarray} &&\Delta_F(a)=F\Delta(a)F^{-1},~~~\forall a\in A,\label{twisted-d}\\ &&\Phi_F=F_{12}(\Delta\otimes 1)F\cdot\Phi\cdot(1\otimes\Delta)F^{-1}F_{23}^{-1}\label{twisted-phi} \end{eqnarray} where the subscripts above refer to the embedding of the elements in the triple tensor product space. Then \begin{Theorem}\label{t-quasi-hopf}: $(A,\Delta_F,\epsilon,\Phi_F)$ defined by (\ref{twisted-d}, \ref{twisted-phi}) together with $\alpha_F,\beta_F, S_F$ given by \begin{equation} S_F=S,~~~\alpha_F=m\cdot(1\otimes\alpha)(S\otimes 1)F^{-1},~~~ \beta_F=m\cdot(1\otimes\beta)(1\otimes S)F,\label{twisted-s-ab} \end{equation} is also a quasi-Hopf algebra. Throughout, the element $F$ is referred to as a twistor. \end{Theorem} \begin{Definition}\label{quasi-quasi}: A quasi-Hopf algebra $(A,\Delta,\epsilon,\Phi)$ is called quasi-triangular if there exists an invertible element ${\cal R}\in A\otimes A$ such that \begin{eqnarray} &&\Delta^T(a){\cal R}={\cal R}\Delta(a),~~~~\forall a\in A,\label{dr=rd}\\ &&(\Delta\otimes 1){\cal R}=\Phi^{-1}_{231}{\cal R}_{13}\Phi_{132}{\cal R}_{23}\Phi^{-1}_{123}, \label{d1r}\\ &&(1\otimes \Delta){\cal R}=\Phi_{312}{\cal R}_{13}\Phi^{-1}_{213}{\cal R}_{12}\Phi_{123}. \label{1dr} \end{eqnarray} We refer to ${\cal R}$ as the universal R-matrix. \end{Definition} Throughout, $\Delta^T=T\cdot\Delta$ with $T$ being the twist map which is defined by \begin{equation} T(a\otimes b)=b\otimes a; \end{equation} and $\Phi_{132}$ {\it etc} are derived from $\Phi\equiv\Phi_{123}$ with the help of $T$ \begin{eqnarray} &&\Phi_{132}=(1\otimes T)\Phi_{123},\nonumber\\ &&\Phi_{312}=(T\otimes 1)\Phi_{132}=(T\otimes 1) (1\otimes T)\Phi_{123},\nonumber\\ &&\Phi^{-1}_{231}=(1\otimes T)\Phi^{-1}_{213}=(1\otimes T) (T\otimes 1)\Phi^{-1}_{123},\nonumber \end{eqnarray} and so on. It is easily shown that the properties (\ref{dr=rd})-(\ref{1dr}) imply the Yang-Baxter type equation, \begin{equation} {\cal R}_{12}\Phi^{-1}_{231}{\cal R}_{13}\Phi_{132}{\cal R}_{23}\Phi^{-1}_{123} =\Phi^{-1}_{321}{\cal R}_{23}\Phi_{312}{\cal R}_{13}\Phi^{-1}_{213}{\cal R}_{12}, \label{quasi-ybe} \end{equation} which is referred to as the quasi-Yang-Baxter equation. \begin{Theorem}\label{t-quasi-quasi}: Denoting by the set $(A,\Delta,\epsilon,\Phi,{\cal R})$ a quasi-triangular quasi-Hopf algebra, then $(A, \Delta_F, \epsilon, \Phi_F, {\cal R}_F)$ is also a quasi-triangular quasi-Hopf algebra, with the choice of $R_F$ given by \begin{equation} {\cal R}_F=F^T {\cal R} F^{-1},\label{twisted-R} \end{equation} where $F^T=T\cdot F\equiv F_{21}$. Here $\Delta_F$ and $\Phi_F$ are given by (\ref{twisted-d}) and (\ref{twisted-phi}), respectively. \end{Theorem} Let us specify some notations, where we adopt a summation convention over all repeated indices. Throughtout the paper, \begin{eqnarray} &&\Phi= X_i\otimes Y_i\otimes Z_i,~~~~ \Phi^{-1}= \bar{X}_i\otimes\bar{Y}_i\otimes\bar{Z}_i,\nonumber\\ &&F= f_i\otimes f^i,~~~~F^{-1}= \bar{f}_i\otimes \bar{f}^i,\nonumber\\ &&{\cal R}=a_\nu\otimes b_\nu,~~~~{\cal R}^{-1}=c_\nu\otimes d_\nu,\nonumber\\ &&(1\otimes\Delta)\Delta(a)=\sum a_{(1)}\otimes\Delta(a_{(2)})=\sum a^R_{(1)}\otimes a_{(2)}^R\otimes a^R_{(3)},\nonumber\\ &&(\Delta\otimes 1)\Delta(a)=\sum \Delta(a_{(1)})\otimes a_{(2)})=\sum a^L_{(1)}\otimes a_{(2)}^L\otimes a^L_{(3)}, \nonumber \\ &&(\Phi^{-1}\ot I).(\Delta\ot I\ot I)\Phi= A_i\ot B_i \ot C_i \ot D_i, \nonumber \\ &&(\Delta\ot I\ot I)\Phi^{-1}.(\Phi\ot I)= K_i\ot L_i\ot M_i\ot N_i .\label{notation} \end{eqnarray} A important type of twistor is that due to Drinfeld \cite{Dri90}. For any quasi-Hopf algebra $A$ observe that the actions $$(S\ot S)\cdot\Delta^T,~~~~~~~~~~~~~~~\Delta^T\cdot S^{-1}$$ both determine algebra anti-homomorphisms. It follows that $$\Delta'\equiv (S\ot S)\cdot\Delta^T\cdot S^{-1}$$ gives rise to an algebra homomorphism and thus a co-product action on $A$. Drinfeld showed that the actions $\Delta$ and $\Delta'$ are related by a twistor; i.e. $$\Delta(a)=\f^{-1}\left((S\ot S)\Delta^T(S^{-1}(a))\right) \f~~~~~~~~\forall a\in A$$ where $$ \f= (S\ot S)\Delta^T(X_i).\gamma.\Delta\left(Y_i\beta S(Z_i)\right) $$ and \begin{equation} \gamma= S(B_i)\alpha C_i\ot S(A_i)\alpha D_i. \label{gamma} \end{equation} It is also useful to define \begin{equation} \delta= K_i\beta S(N_i)\ot L_i \beta S(M_i). \label{delta} \end{equation} Then the following relations can be shown to hold $$\Delta(\alpha)=\f^{-1}\gamma,~~~~~~~~~\Delta(\beta)=\delta\f.$$ A quasi-Hopf algebra is said to be of trace type if there exists an invertible element $u\in A$ such that \begin{equation} S^2(a)=u au^{-1},~~~~\forall a\in A.\label{s2a=u} \end{equation} In the case $A$ is quasi-triangular with R-matrix as in (\ref{notation}) we have \cite{ac} \begin{Theorem}\label{u-operator}: The operator defined by \begin{equation} u= S\left(Y_i\beta S(Z_i)\right)S(\beta_\nu)\alpha a_\nu X_i \label{u} \end{equation} satisfies (\ref{s2a=u}). Moreover the inverse is given by \begin{equation} u^{-1}= S^{-1}(X_i)S^{-1}(\alpha d_\nu)c_\nu Y_i\beta S(Z_i).\label{u-1} \end{equation} \end{Theorem} An important relation satisfied by $u$ is \begin{equation} S(\alpha)u=S(b_\nu)\alpha a_\nu \label{imp} \end{equation} which we will need later. The significance of trace type quasi-Hopf algebras is that they afford a systematic means to construct Casimir invariants. We have the following result from \cite{gzi}. \begin{Theorem}\label{trace-inv}: Let $\pi$ be the representation afforded by the finite-dimensional $A$-module $V$. Suppose $\eta=\mu_i\otimes \nu_i\otimes \rho_i\in A\otimes {\rm End}V\otimes A$ obeys \begin{equation} (1\otimes\pi\otimes 1)(1\otimes\Delta)\Delta(a)\cdot\eta=\eta\cdot (1\otimes\pi\otimes 1)(1\otimes\Delta)\Delta(a),~~~~\forall a\in A, \label{eta}\end{equation} then \begin{equation} {\rm tr}\left(\nu_i\pi\left(\beta S(\rho_i)S(\alpha)u\right)\right)\mu_i \end{equation} is a central element. Similarly if $\bar{\eta}=\bar{\mu}_i\otimes \bar{\nu}_i\otimes\bar{\rho}_i\in A\otimes {\rm End}V\otimes A$ satisfies \begin{equation} \bar{\eta}\cdot(1\otimes\pi\otimes 1)(\Delta\otimes 1)\Delta(a)= (1\otimes\pi\otimes 1)(\Delta\otimes 1)\Delta(a)\cdot\bar{\eta},~~~~\forall a\in A \end{equation} then \begin{equation} \sum {\rm tr}\left(\bar{\nu}_i\pi\left(u^{-1}S(\beta)S(\bar{\mu}_i)\alpha\bar{\nu}_i\right) \right)\bar{\rho}_i \end{equation} is a central element. \end{Theorem} As a consequence of the above we have \begin{Corollary}: Suppose $\omega=\sum \omega_i\otimes \Omega^i\in A\otimes{\rm End}V$ satisfies \begin{equation} (1\otimes\pi)\Delta(a)\cdot\omega=\omega\cdot(1\otimes\pi)\Delta(a),~~~~\forall a\in A. \end{equation} Then (\ref{eta}) implies that \begin{eqnarray} \tau(\omega)&=&{\rm tr}\left(\Omega^i\pi\left(Y_k\beta S(\bar{Z}_j Z_k) S(\alpha)u\bar{Y}_j \right)\right)\bar{X}_j\omega_i X_k\nonumber\\ \label{trace-o-inv} \end{eqnarray} is a central element. \end{Corollary} For an $(n+1)$-fold tensor product space and $\omega=\sum \omega_i\otimes \Omega^i\in A^{\ot n}\otimes{\rm End}V$ we define \begin{equation} \tau_n(\omega)={\rm tr}\left(\Omega^i\pi\left(Y_k\beta S(\bar{Z}_j Z_k) S(\alpha)u\bar{Y}_j \right)\right)\Delta^{(n-1)}(\bar{X}_j)\omega_i \Delta^{(n-1)}(X_k). \label{tn} \end{equation} \sect{Representations of the braid group} Given any representation $\pi$ of a quasi-Hopf algebra $A$ we set \begin{equation} \check R=P.(\pi \ot \pi){\cal R} \label{rh} \end{equation} and $$\Phi_i=(\Delta^{(i-2)}\ot I\ot I)\Phi.$$ In terms of $\check R$ the quasi-Yang-Baxter equation may be written \begin{equation} \Phi\check R_{23}\Phi^{-1}\check R_{12}\Phi\check R_{23}\Phi^{-1} =\check R_{12}\Phi\check R_{23}\Phi^{-1}\check R_{12} \label{qyb} \end{equation} where throughout we use the same symbols $\Phi$ and $\Phi_i$ for both the algebraic objects and their matrix representatives. \begin{Theorem} Define $n$ operators on the $(n+1)$-fold tensor product space by \begin{equation} \sigma_i=\Phi_i\check R_{i(i+1)}\Phi_i^{-1},~~~~i=1,2,...,n\label{bg} \end{equation} These give rise to a representation of the braid group $B_n$ by satisfying the defining relations \begin{eqnarray} \sigma_i\sigma_j&=&\sigma_j\sigma_i ~~~~~j\neq i\pm 1 \label{com} \\ \sigma_i\sigma_{i+1}\sigma_i&=&\sigma_{i+1}\sigma_i\sigma_{i+1}.\label{br} \end{eqnarray} \end{Theorem} The above result was given in \cite{ac}. Here we want to present a detailed proof. First we establish that the braid generators (\ref{bg}) are invariant with respect to the co-product action $\Delta^{(n)}$ of $A$; i.e \begin{equation} [\sigma_i,\,\Delta^{(n)}(a)]=0~~~~~~~\forall a\in A.\label{inv} \end{equation} It is clear from the definition (\ref{rh}) that $$ [\check R,\,\Delta(a)]=0~~~~~~~\forall a\in A$$ which immediately implies that $$[\sigma_1,\,\Delta^{(n)}(a)]=0~~~~~~~\forall a\in A.$$ Next consider \begin{eqnarray} \sigma_2\Delta^{(j)}(a)&=&\Phi\check R_{23}\Phi^{-1}(\Delta\ot I^{\ot(j-1)})\Delta^{ (j-1)}(a) \nonumber \\ &=&\Phi\check R_{23}(I\ot \Delta\ot I^{(j-2)})\Delta^{(j-1)}(a)\Phi^{-1} \nonumber \\ &=&\Phi(I\ot \Delta\ot I^{(j-2)})\Delta^{(j-1)}(a)\check R_{23}\Phi^{-1} \nonumber \\ &=&(\Delta\ot I^{\ot(j-1)})\Delta^{(j-1)}(a)\Phi\check R_{23}\Phi^{-1}\nonumber \\ &=&\Delta^{(j)}(a) \sigma_2 \label{si2} \end{eqnarray} Observing that the action (\ref{act}) enjoys the property $$\Delta^{(i)}\cdot \Delta^{(j)}=\Delta^{(i+j)}.$$ and applying $\Delta^{(k)}\ot I^{\ot j}$ to (\ref{si2}) now yields (\ref{inv}) by choosing $k=i-2,\,j=n-i-2$. Since $\check R$ commutes with the co-product action we immediately deduce for $i>1$ \begin{eqnarray} \sigma_1\sigma_i&=&\check R_{12} \Phi_i \check R_{i(i+1)} \Phi_i^{-1} \nonumber \\ &=& \Phi_i\check R_{12}\check R_{i(i+1)}\Phi_i^{-1} \nonumber \\ &=&\Phi_i\check R_{i(i+1)}\check R_{12}\Phi_i^{-1}\nonumber \\ &=&\Phi_i\check R_{i(i+1)}\Phi_i^{-1}\check R_{12}\nonumber \\ &=&\sigma_i\sigma_1. \nonumber \end{eqnarray} Consider now for $l>3$ \begin{eqnarray} \sigma_2\sigma_l&=&\sigma_2 (\Delta^{(l-2)}\ot I\ot I)\Phi. \check R_{l(l+1)}(\Delta^{(l-2)} \ot I \ot I)\Phi^{-1} \nonumber \\ &=&(\Delta^{(l-2)}\ot I\ot I)\Phi.\sigma_2\check R_{l(l+1)}(\Delta^{(l-2)}\Phi^{-1} \ot I\ot I)\nonumber \\ &=&(\Delta^{(l-2)}\ot I\ot I)\Phi. \Phi_{123}\check R_{23}\Phi_{123}^{-1} \check R_{l(l+1)}(\Delta^{(l-2)}\ot I\ot I)\Phi^{-1} \nonumber \\ &=&(\Delta^{(l-2)}\ot I\ot I)\Phi.\check R_{l(l+1)}\Phi_{123} \check R_{23}\Phi_{123}^{-1} (\Delta^{(l-2)}\ot I\ot I)\Phi^{-1} \nonumber \\ &=&(\Delta^{(l-2)}\ot I\ot I)\Phi.\check R_{l(l+1)}\sigma_2(\Delta^{(l-2)} \ot I\ot I)\Phi^{-1} \nonumber \\ &=& (\Delta^{(l-2)}\ot I\ot I)\Phi. \check R_{l(l+1)}(\Delta^{(l-2)} \ot I\ot I)\Phi^{-1}.\sigma_2 \nonumber \\ &=& \sigma_2\sigma_l. \label{s2com} \end{eqnarray} Applying $\Delta^{(k)}\ot I\ot I$ to (\ref{s2com}) yields (\ref{com}) for $i\geq 2$ by choosing $k=i-2,\,l=j-i+2.$ In order to show that (\ref{br}) is satisfied we see from (\ref{qyb}) that $$\sigma_1\sigma_2\sigma_1=\sigma_2\sigma_1\sigma_2$$ is certainly true. Now through (\ref{qyb}), the invariance of $\check R$ and repeated use of the pentagonal relation (\ref{pent}) we find \begin{eqnarray} &&\sigma_2\sigma_3\sigma_2 \nonumber \\ &&=\Phi_2\check R_{23}\Phi_2^{-1}\Phi_3\check R_{34}\Phi_3^{-1} \Phi_2\check R_{23}\Phi^{-1}_2 \nonumber \\ &&=\Phi_2\check R_{23}\Phi_2^{-1}\Phi_3(I\ot I\ot\Delta)\Phi.\check R_{34} (I\ot I\ot\Delta)\Phi^{-1}.\Phi_3^{-1} \Phi_2\check R_{23}\Phi^{-1}_2 \nonumber \\ &&=\Phi_2\check R_{23}(I\ot \Delta\ot I)\Phi.(I\ot\Phi) \check R_{34}(I\ot\Phi^{-1})(I\ot \Delta\ot I)\Phi^{-1} .\check R_{23}\Phi^{-1}_2 \nonumber \\ &&=\Phi_2(I\ot \Delta\ot I)\Phi.\left[\check R_{23}(I\ot\Phi) \check R_{34}(I\ot\Phi^{-1}) \check R_{23}\right] (I\ot \Delta\ot I)\Phi^{-1}.\Phi^{-1}_2 \nonumber \\ &&=\Phi_2(I\ot \Delta\ot I)\Phi.\left[(I\ot\Phi)\check R_{34}(I\ot\Phi^{-1}) \check R_{23}(I\ot\Phi)\check R_{34}(I\ot\Phi^{-1})\right] (I\ot \Delta\ot I)\Phi^{-1}.\Phi^{-1}_2 \nonumber \\ &&=\Phi_3(I\ot I\ot \Delta)\Phi.\check R_{34}(I\ot\Phi^{-1}) \check R_{23}(I\ot\Phi)\check R_{34}(I\ot I\ot \Delta)\Phi^{-1}. \Phi_3^{-1} \nonumber \\ &&=\Phi_3\check R_{34}(I\ot I\ot \Delta)\Phi.(I\ot\Phi^{-1}) \check R_{23}(I\ot\Phi)(I\ot I\ot \Delta)\Phi^{-1}.\check R_{34} \Phi_3^{-1}\nonumber \\ &&=\Phi_3\check R_{34}\Phi_3^{-1}\Phi_2 (I\ot \Delta\ot I)\Phi.\check R_{23}(I\ot \Delta\ot I)\Phi^{-1}.\Phi^{-1}_2 \Phi_3\check R_{23}\Phi_3^{-1}\nonumber \\ &&=\Phi_3\check R_{34}\Phi_3^{-1}\Phi_2\check R_{23}\Phi_2^{-1}\Phi_3\check R_{34}\Phi_3^{-1} \nonumber \\ &&=\sigma_3\sigma_2\sigma_2 \label{int1} \end{eqnarray} Finally, acting $\Delta^{(i-2)}\ot I^{\ot 3}$ on (\ref{int1}) above yields $$\sigma_i\sigma_{i+1}\sigma_i=\sigma_{i+1}\sigma_i\sigma_{i+1}. $$ \sect{Link Polynomials from Ribbon Quasi-Hopf Algebras} In \cite{ac} the following definition was proposed for the ribbon quasi-Hopf algebras. \begin{Definition} Let $A$ be a quasi-triangular quasi-Hopf algebra. We say that $A$ is a ribbon quasi-Hopf algebra if there exists a central element $v\in A$ such that \begin{itemize} \item[1.] $v^2=uS(u)$ \item[2.] $S(v)=v$ \item[3.] $\epsilon (v)=1$ \item[4.] $\Delta(uv^{-1})=\f^{-1}(S\ot S)\f_{21}.uv^{-1}\ot uv^{-1}$ \end{itemize} where $\f$ is the Drinfeld twist discussed earlier. \end{Definition} Given a ribbon quasi-Hopf algebra, a prescription was also provided in \cite{ac} to define a Markov trace on the braid group representation which in turn may be used to compute link polynomials in the usual way. From here on we will omit the symbol $\pi$ denoting the representation for ease of notation. \begin{Theorem} Let $\Psi$ be a word in the braid generators (\ref{bg}) for a fixed finite dimensional irreducible representation of a ribbon Hopf-algebra $A$. Then a Markov trace $\theta_n $ on the $(n+1)$-fold tensor product space may be defined by $$\theta_n(\Psi)={\rm tr}\left(\Psi\Delta^{(n-1)}(\beta S(\alpha)uv^{-1}) \right)$$ which satisfies the Markov properties \begin{itemize} \item[1.] $\theta_n(\Psi_1\Psi_2)=\theta_n(\Psi_2\Psi_1) ~~~~~\forall\,\Psi_1,\,\Psi_2\in B_n$ \item[2.] $\theta_n(\Psi\sigma^{\pm 1})=z^{\pm}\theta_{n-1}(\Psi) ~~~~~\forall \Psi\in B_{n-1}\subset B_n$ \end{itemize} where $z^{\pm}$ are the eigenvalues of the central operators $v^{\mp 1}$ in the representation $\pi$. \end{Theorem} The importance of the Markov trace is that from it one can define a link polynomial $L(\hat{\Psi})$ through \begin{equation} L(\hat{\Psi})=(z^+z^-)^{n/2}\;\left(\frac{z^-}{z^+}\right)^{e(\theta)/2}\theta(\Psi),~~~~ \theta\in B_n \end{equation} where $e(\Psi)$ is the sum of the exponents of the $\sigma_i$'s appearing in $\Psi$. The functional $L(\hat{\Psi})$ enjoys the following properties: \begin{itemize} \item[1.] $~~L(\widehat{\Psi\eta})=L(\widehat{\eta\Psi}),~\forall \Psi, \eta\in B_M;$ ~ \item[2.] $~~L(\widehat{\Psi\sigma_{n-1}^{\pm 1}})=L(\hat{\Psi}),~ \forall \Psi\in B_{n-1}\subset B_n$ \end{itemize} and is an invariant of ambient isotopy. The first Markov property follows easily from the invariance of the braid generators $\sigma^{\pm 1}$ and the cyclic rule of traces. To establish the second Markov property requires some work and was stated in \cite{ac} without proof. Here we provide the details. Before proceeding, we need to determine the co-product action of the element $S(\alpha)uv^{-1}$. Using the Drinfeld twistor we find \begin{eqnarray} \Delta\left(S(\alpha)\right)&=&\f^{-1}\left((S\ot S)\Delta^T(\alpha)\right)\f \nonumber \\ &=&\f^{-1}\left((S\ot S)(\f^{-1}_{21}\gamma_{21})\right)\f \nonumber \\ &=&\f^{-1}(S\ot S)\gamma_{21}.(S\ot S)\f_{21}^{-1}.\f \nonumber \end{eqnarray} Now through using (\ref{gamma}) and definition 3 we find that \begin{equation} \Delta\left(S(\alpha)uv^{-1}\right)= \f^{-1}\left(S(D_i)S(\alpha)uv^{-1}A_i \ot S(C_i)S(\alpha)uv^{-1}B_i\right). \label{dels} \end{equation} We will also need the following result \begin{Lemma} Let ${\cal C}\in {\rm End} (V\ot V)$ be any invariant operator; i.e $$[{\cal C}, \Delta(a)]=0 ~~~~~~\forall a\in A.$$ Then $$\left(A_i\ot B_i\right){\cal C}\left(K_j\beta S(D_iN_j)\ot L_j\beta S(C_iM_j)\right) =\left(\bar{X}_j\ot\bar{Y}_j\right){\cal C}\left(X_i\beta\ot Y_i\beta S(\Z_j Z_i)\right).$$ \end{Lemma} The above result follows directly from the definitions (\ref{notation}). We may now see that \begin{eqnarray} \theta_n(\Psi)&=&{\rm tr}\left(\Psi\Delta^{(n)}(\beta S(\alpha)uv^{-1}) \right) \nonumber \\ &=&{\rm tr}\left(\Psi.(\Delta^{(n-1)}\ot I)\delta.\Delta^{(n-1)}(S(D_i)S(\alpha)uv^{-1}A_i)\ot S(C_i)S(\alpha) uv^{-1}B_i \right) \nonumber \\ &=&{\rm tr}\left(\Psi.\Delta^{(n-1)}\left(K_j\beta S(N_j)S(D_i)S(\alpha) uv^{-1}A_i\right) \ot L_j\beta S(M_j)S(C_i)S(\alpha) uv^{-1}B_i \right) \nonumber \\ &=&{\rm tr}\left(\Psi. \Delta^{(n-1)}(X_i\beta S(\alpha)uv^{-1}\bar{X}_j)\ot Y_i\beta S(Z_i)S(\Z_j)S(\alpha)uv^{-1}\bar{Y}_j \right) \nonumber \\ &=&{\rm tr}\left(\tau_n(\Psi)\Delta^{(n-1)}(\beta S(\alpha)uv^{-1})\right) \nonumber \\ &=&\theta_{n-1}\left(\tau_n(\Psi)\right) \end{eqnarray} where the element $u$ in the definition (\ref{tn}) has now been replaced by $uv^{-1}$. It is apparent also from (\ref{tn}) that for $\Psi\in B_{n-1}$ then $$\tau_n(\Psi\sigma_n^{\pm 1})=\Psi\tau_n(\sigma_n^{\pm 1})$$ To evaluate $\tau_n(\sigma_n^{\pm 1})$ we can appeal to the pentagonal relation (\ref{pent}) to find that \begin{eqnarray} &&\Phi^{-1}_{n+1}\sigma_n^{\pm 1}\Phi_{n+1}\nonumber \\ &&= \Phi^{-1}_{n+1}\Phi_n\check R_{n(n+1)}^{\pm 1}\Phi^{-1}_n\Phi_{n+1} \nonumber \\ &&=I^{(n-2)}\ot \left(( I \ot I \ot \Delta)\Phi.(I\ot \Phi^{-1})(I\ot \Delta\ot I)\Phi^{-1}.(I\ot \check R^{\pm 1} \ot I)\right. \nonumber \\ &&~~~~~~~~~~~.\left.(I\ot \Delta\ot I)\Phi.(I\ot \Phi) (I\ot I\ot \Delta)\Phi^{-1}\right) \nonumber \\ &&=I^{(n-2)}\ot \left((I\ot I \ot \Delta)\Phi.(I\ot \Phi^{-1}) (I\ot \check R^{\pm 1} \ot I)(I\ot \Phi) (I\ot I\ot \Delta)\Phi^{-1}\right) \nonumber \end{eqnarray} which, upon using (\ref{quasi-hopf1}, \ref{quasi-hopf2}, \ref{e(phi)=1}), leads us to conclude that $$\tau_n(\sigma_n^{\pm 1})=I^{\ot (n-1)}\ot \tau(\check R^{\pm 1} ). $$ An algebraic exercise shows that \begin{eqnarray} \tau(\check R)&=&\bar{X}_jb_{\nu}Y_l\beta S(Z_l)S(\Z_j)S(\alpha)uv^{-1}\bar{Y}_j a_{\nu}X_l, \nonumber \\ \tau(\check R^{-1})&=&X_jc_{\nu}Y_k\beta S(Z_k)S(\Z_j)S(\alpha)uv^{-1}\bar{Y}_jd_{\nu}X_k \nonumber \end{eqnarray} and hence we can conclude that $z^{\pm}$ are given by the eigenvalues of the central operators $v^{\mp 1}$ if we can show that the following relations hold. \begin{Lemma} \begin{eqnarray} \bar{X}_jb_{\nu}Y_l\beta S(Z_l)S(\Z_j)S(\alpha)u\bar{Y}_j a_{\nu}X_l&=&I, \nonumber \\ \bar{X}_jc_{\nu}Y_k\beta S(Z_k)S(\Z_j)S(\alpha)u\bar{Y}_jd_{\nu}X_k &=&v^2.\nonumber \end{eqnarray} \end{Lemma} Through use of (\ref{quasi-hopf3}, \ref{imp}) we obtain \begin{eqnarray} I&=&\bar{X}_i\beta S(\bar{Y}_i)\alpha \Z_i \nonumber \\ &=&\bar{X}_j\beta S(\bar{Y}_i)S(d_{\nu})S(b_{\mu})\alpha a_{\mu}c_{\nu}\Z_i \nonumber \\ &=&\bar{X}_j\beta S(\bar{Y}_i)S(d_{\nu})S(\alpha)uc_{\nu}\Z_i \nonumber \\ &=&\bar{X}_i\beta S(\bar{Y}_i)S(d_{\nu})S(\alpha)S^2(c_\nu)S^2(\Z_i)u. \end{eqnarray} From \reff{1dr} we see that $${\cal R}^{-1}_{13}\Phi^{-1}_{312}\left(I\ot \Delta\right){\cal R}=\Phi^{-1}_{213}{\cal R}_{12} \Phi_{123} $$ which expressed in terms of the tensor components reads $$c_{\nu}\Z_ja_l\ot \bar{X}_jb_l^{(1)}\otimes d_{\nu}\bar{Y}_j b_l^{(2)} =\bar{Y}_j a_\nu X_l\ot \bar{X}_j b_\nu Y_l\ot \Z_jZ_l. $$ We can now write \begin{eqnarray} &&\bar{X}_jb_{\nu}Y_l\beta S(Z_l)S(\Z_j)S(\alpha)u\bar{Y}_j a_{\nu}X_l \nonumber \\ &&~~~~=\bar{X}_jb_l^{(1)}\beta S(b_l^{(2)})S(\bar{Y}_j)S(d_\nu)S(\alpha)u c_\nu\Z_ja_l \nonumber \\ &&~~~~=\epsilon(b_l)\bar{X}_j\beta S(\bar{Y}_j)S(d_\nu)S(\alpha)u c_\nu\Z_ja_l \nonumber \\ &&~~~~=\bar{X}_i\beta S(\bar{Y}_i)S(d_{\nu})S(\alpha)S^2(c_\nu)S^2(\Z_i)u \nonumber \\ &&~~~~=I. \end{eqnarray} Next we see that \begin{eqnarray} u&=& S\left(\bar{X}_i\beta S(\bar{Y}_i)\alpha\Z_i\right)u \nonumber \\ &=&S(\Z_i)S(\alpha)S^2(\bar{Y}_i)S(\beta)S(\bar{X}_i)u \nonumber \\ &=&S(\Z_i)S(\alpha)u\bar{Y}_iS^{-1}(\beta)S^{-1}(\bar{X}_i) \nonumber \\ &=&S(\Z_i)S(b_\nu)\alpha a_\nu\bar{Y}_iS^{-1}(\beta)S^{-1}(\bar{X}_i) \end{eqnarray} where in the last step we have used \reff{imp}. Consequently $$S(u)=\bar{X}_i\beta S(\bar{Y}_i)S(a_\nu)S(\alpha)S^2(b_\nu)S^2(\Z_i). $$ \ From \reff{d1r} we have $${\cal R}_{23}\Phi^{-1}_{123}\left(\Delta\ot I\right){\cal R}^{-1}=\Phi_{132}^{-1} {\cal R}^{-1}_{13}\Phi_{231} $$ which we may express as $$\bar{X}_jc^{(1)}_\nu\ot a_\mu\bar{Y}_jc_\nu^{(2)}\ot b_\mu \Z_jd_\nu =\bar{X}_jc_\nu Y_k\ot \Z_jZ_k\ot \bar{Y}_jd_\nu X_k. $$ This relation leads us to deduce that \begin{eqnarray} &&\bar{X}_jc_{\nu}Y_k\beta S(Z_k)S(\Z_j)S(\alpha)u\bar{Y}_jd_{\nu}X_k \nonumber \\ &&~~~~=\bar{X}_jc_\nu^{(1)}\beta S(c_\nu^{(2)})S(\bar{Y}_j)S(a_\mu) S(\alpha)u b_\mu\Z_jd_\nu \nonumber \\ &&~~~~=\bar{X}_j\beta S(\bar{Y}_j)S(\alpha_\mu)S(\alpha)S^2(b_\mu)S^2(\Z_j)u \nonumber \\ &&~~~~=S(u)u \nonumber \\ &&~~~~=v^2 \end{eqnarray} which proves lemma 2 and completes the proof of Theorem 6. \sect{Twisting Invariance of the Markov trace} Now we are in a position to show twisting invariance of the link polynomials. Let us begin with the following result. \begin{Proposition} Every twisted ribbon quasi-Hopf algebra is again a ribbon quasi-Hopf algebra. \end{Proposition} Recall from definition 3 that the first three conditios of a ribbon quasi-Hopf algebra are properties of the algebra structure rather than the co-algebra. Thus, to this end we need only show that if $$ \Delta(uv^{-1})=\f^{-1}(S\ot S)\f_{21}.\,uv^{-1}\ot uv^{-1} $$ then $$\Delta_F(uv^{-1})=\f_F^{-1}(S\ot S)(\f_F)_{21}.\,uv^{-1}\ot uv^{-1}$$ where $\f_F$ denotes the Drinfeld twistor for the twisted quasi-Hopf algebra. Recalling that the Drinfeld twist $\f$ is determined by $$\f\Delta(a)\f^{-1}=(S\ot S)\left(\Delta^T\left(S^{-1}(a)\right)\right) ~~~~~\forall a\in A $$ shows that \begin{eqnarray} F\Delta(a)F^{-1} &=& \Delta_F(a)\nonumber \\ &=&\f^{-1}_F\left((S\ot S)\Delta^T_F\left(S^{-1}(a)\right)\right)\f_F \nonumber \\ &=&\f_F^{-1}\left((S\ot S)\left(F_{21}\Delta^T\left(S^{-1}(a)\right) F_{21}^{-1}\right)\right)\f_F \nonumber \\ &=&\f_F^{-1}\left((S\ot S)F_{21}^{-1}.(S\ot S)\Delta^T\left(S^{-1}(a)\right).(S\ot S)F_{21}\right) \f_F \nonumber \\ &=&\f^{-1}_F(S\ot S)F_{21}^{-1}.\f\Delta(a)\f^{-1}.(S\ot S)F_{21}\f_F \nonumber \end{eqnarray} which leads us to $$\f_F=(S\ot S) F_{21}^{-1}.\f F^{-1}. $$ Now we observe that \begin{eqnarray} &&\f_F^{-1}(S \ot S)(\f_F)_{21}.uv^{-1}\ot uv^{-1} \nonumber \\ &&~~~~~=F\f^{-1}(S\ot S)F_{21}.(S\ot S)\left((S\ot S)F^{-1}.(\f F^{-1})_{21} \right)uv^{-1}\ot uv^{-1} \nonumber \\ &&~~~~~=F\f^{-1}(S\ot S)\f_{21}.(S^2\ot S^2)F^{-1}.uv^{-1}\ot uv^{-1} \nonumber \\ &&~~~~~=F\f^{-1}(S\ot S)\f_{21}.uv^{-1}\ot uv^{-1}. F^{-1}\nonumber \\ &&~~~~~=F\Delta(uv^{-1})F^{-1} \nonumber \\ &&~~~~~=\Delta_F(uv^{-1}) \nonumber \end{eqnarray} thus establishing that the twisted ribbon quasi-Hopf algebra is also of ribbon type. By induction the co-product action on the $(n+1)$-fold space assumes the form $$\Delta^{(n)}_F(a)=\chi_{n}\Delta^{(n)}\chi_{n}^{-1}$$ where $$\chi_{n}=F_{12}.(\Delta\ot I)F_{12}.(\Delta^2\ot I)F_{12}....(\Delta^{(n-1)}\ot I )F. $$ Consider next \begin{eqnarray} \chi_{n}\sigma_i\chi^{-1}_{n}&=&F_{12}(\Delta\ot I)F_{12}.(\Delta^2\ot I)F_{12}....(\Delta^{(n-1)}\ot I )F \sigma_i \nonumber \\ && ~~~~~~~~~~ (\Delta^{(n-1)}\ot I )F^{-1}....(\Delta\ot I)F_{12}^{-1}.F_{12}^{_1} \nonumber \\ &=&F_{12}(\Delta\ot I)F_{12}.(\Delta^2\ot I)F_{12}....(\Delta^{(i-1)}\ot I )F \sigma_i \nonumber \\ && ~~~~~~~~~ (\Delta^{(i-1)}\ot I )F^{-1}....(\Delta\ot I)F_{12}^{-1}.F_{12}^{_1} \nonumber \\ &=&\chi_i\sigma_i\chi_i^{-1}. \nonumber \end{eqnarray} We now determine the representations of the braid generators under twisting; i.e. \begin{eqnarray} \sigma_i^F&=& (\Delta_F^{(i-2)}\ot I\ot I)\Phi_F.\check R^F_{i(i+1)}. (\Delta_F^{(i-2)}\ot I\ot I) \Phi_F^{-1} \nonumber \\ &=&\chi_{i-2}(\Delta^{(i-2)}\ot I \ot I)\left(F_{12}(\Delta\ot I)F.\Phi. (I\ot \Delta)F^{-1}.F_{23}^{-1}\right)\chi_{i-2}^{-1}F_{i(i+1)}\check R_{i(i+1)}F^{-1}_{i.i+1} \nonumber \\ && ~~~~~~~~~\chi_{i-2}\Delta^{(i-2)}\left(F_{23}(I\ot \Delta)F.\Phi^{-1}(\Delta\ot I)F^{-1}.F_{12}^{-1}\right) \chi^{-1}_{i-2} \nonumber \\ &=&\chi_i\Delta\Phi.(\Delta^{(i-2)}\otimes\Delta)F^{-1}.F_{i(i+1)}^{-1} \chi_{i-2}^{-1}F_{i(i+1)}\check R_{i(i+1)}F^{-1}_{i(i+1)} \nonumber \\ && ~~~~~~~~~\chi_{i-2} F_{i(i+1)}(\Delta^{(i-2)}\ot \Delta)F.\Delta^{(i-2)}\Phi^{-1}. \chi_i^{-1} \nonumber \\ &=&\chi_i\Delta^{(i-2)}\Phi. \check R_{i(i+1)} \Delta^{(i-2)} \Phi^{-1}. \chi_i^{-1} \nonumber \\ &=&\chi_i\sigma_i\chi_i^{-1} \nonumber \\ &=&\chi_{n}\sigma_i\chi_{n}^{-1} \nonumber \end{eqnarray} which shows that the representation of the braid generators under twisting are related to those of the untwisted case by a basis transformation. Thus for any word in the generators of the braid group we can write $$\Psi^F=\chi_{n}\Psi\chi_{n}^{-1}$$ in an obvious notation. Using the relations (\ref{twisted-s-ab}) we may write $$\alpha_F=S(\bar{f}_i)\alpha \bar{f}^i,~~~~~~~~\beta_F=f_i\beta S(f^i)$$ and proceed to calculate \begin{eqnarray} \theta^F_n(\Psi)&=&{\rm tr}\left(\Psi^F \Delta_F^{(n)}\left(\beta_F S(\alpha_F) uv^{-1}\right)\right) \nonumber \\ &=&{\rm tr}\left(\chi_{n}\Psi \Delta^{(n)} \left(\beta_F S(\alpha_F)uv^{-1}\right)\chi^{-1}_{n}\right) \nonumber \\ &=&{\rm tr}\left(\Psi\Delta^{(n)}\left(f_i\beta S(f^i)S(\bar{f}^j)S( \alpha)S^2(\bar{f}_j)uv^{-1}\right)\right) \nonumber \\ &=&{\rm tr}\left(\Psi\Delta^{(n)}\left(f_i\beta S(\bar{f}^jf^i) S(\alpha)uv^{-1}\bar{f}_j\right)\right) \nonumber \\ &=&{\rm tr}\left(\Delta^{(n)}(\bar{f}_j)\Psi\Delta^{(n)}\left(f_i\beta S(\bar{f}^jf^i)S(\alpha)uv^{-1}\right)\right) \nonumber \\ &=&{\rm tr}\left(\Psi\Delta^{(n)}\left(\bar{f}_jf_i\beta S(\bar{f}^jf^i)S(\alpha)uv^{-1}\right)\right) \nonumber \\ &=&{\rm tr}\left(\Psi \Delta^{(n)}\left(\beta S(\alpha) uv^{-1} \right) \right) \nonumber \\ &=&\theta_n(\Psi) \nonumber \end{eqnarray} which proves twisting invariance of the Markov trace and consequently the associated link polynomials. \vskip.3in \noindent {\bf Acknowledgements.} This work has been financed by the Australian Research Council. \newpage
\section{Introduction} There is recent experimental evidence that the effective interaction between like-charged colloidal particles (``macroions") is sensitively affected by a confinement between two parallel charged glass plates \cite{kepler,crocker,nature}. For aqueous polystyrene suspensions studied in experiment, the effective force between two colloidal macroions is found to be repulsive far away from the plates but becomes attractive when the like-charge macroions are located close to an equally charged plate. At first glance, these findings are surprising as one would expect a purely repulsive interaction from the electrostatic part of the traditional Derjaguin-Landau-Verwey-Overbeek (DLVO) theory \cite{DLVO}. In fact a full theoretical explanation is still missing but several steps were performed in different directions: the essential difference in a confining geometry with respect to the bulk is that the counterion density field is inhomogeneous for small coupling between the macroions and counterions. In a straightforward generalization of the DLVO theory to such an inhomogeneous situation \cite{anne,Jay}, the effective force between the macroions remains repulsive close to the charged plates but becomes weaker since the local concentration of counterions is higher which results in a stronger screening of the Coulomb repulsion. It was further realized that a charged wall induces significant effective triplet interactions \cite{HLEA} which are ignored in the usual DLVO approach resulting in a net attraction \cite{dave23} or in a repulsion \cite{Tehver} depending on the system parameters. An explicit calculation was done within density functional perturbation theory which is justified, however, only for weak inhomogeneities. A complementary approach is to solve the nonlinear Poisson-Boltzmann equation with appropriate boundary conditions in a finite geometry. This was done recently for two charged spheres in a charged cylindrical pore \cite{bowen} as well as for two charged cylinders confined by two parallel charged plates \cite{ospeck}. However, the first situation corresponds to a finite system where the Poisson-Boltzmann approach does not lead to attraction \cite{Neu} and the second situation is a quasi-two-dimensional set-up which is known to behave qualitatively different from a three-dimensional situation \cite{Schmitz}. A further complication arises from image charges induced by the different dielectric constants of the glass and the solvent \cite{Chang,Tandon,dave1}. A general problem of any theoretical description (as DLVO, Poisson-Boltzmann) is that close to the walls the counterion concentration is high and any weak-coupling theory fails {\it a priori\/} when applied to a situation of confined macroions. For strong coupling, even in the bulk, it is unclear whether an effective attraction of like-charged spherical macroions is possible although there are hints from experiments \cite{Ise,Horn,Weiss}, theory \cite{Levin,Netz,shklov,Preparata,Tokuyama}, and computer simulations \cite{allah,Pincus,Tang}. At this stage it is important to remark that a phase separation seen in experiment does not necessarily imply an effective attraction. The additional contribution from the counterions to the total free energy may induce such a phase separation although the effective interaction between the macroions is purely repulsive \cite{roij,Graf}. Bearing the difficulties in experimental interpretations and theory in mind, computer simulations represent a helpful alternative tool to extract ``exact" results for certain model systems. The general accepted theoretical model for the description of charged colloidal suspensions is the ``primitive approach" where the discrete structure of the solvent is disregarded and the interaction between the macroions and counterions is modelled by excluded volume and Coulomb forces. The problem with a full computer simulation of the primitive model is the high charge asymmetry between macro- and counterions which restricts the full treatment to micelles rather than charged colloidal suspensions \cite{Linse}. In this paper, we use computer simulations to obtain ``exact" results for the effective interaction between confined charged colloids based upon the primitive model. Instead of solving the full many-body problem with many macroions, we only simulate one or two macroions confined between two parallel charged plates. This enables us to access high charge numbers of the macro-particles. As a result we find that the wall-particle and the interparticle interaction is repulsive for weak Coulomb coupling. For stronger coupling, the behaviour of the force changes from repulsive to attractive and back to repulsive as the interparticle distance is varied. In particular, the plate-particle interaction exhibits a short-range attraction for a small distances. This may explain the occurrence of crystalline colloidal layers on top of the glass plates found in recent experiments \cite{Larsen_crystallites,larsen,jessica}. These crystallites are metastable but very long-lived and cannot be understood in terms of DLVO-theory. The paper is organized as follows: the model and our target quantities are defined in section II\@. Section III contains details of our computer simulation procedure. Results for the counterion density profiles are shown in section IV\@. The case of a single macroion is discussed in section V, and a macroion pair is investigated in section VI\@. Finally, we conclude in section VII\@. \section{The model and target quantities} We consider $N_m$ macroions with bare charge $q_{m}=Ze>0$ ($e>0$ denoting the elementary charge) and mesoscopic diameter $d_m$ confined between two parallel plates that carry surface charge densities $\sigma_{1}$ and $\sigma_{2}$. We assume that the plates and the macroions are likely charged. The separation distance between plates is $2L$. For convenience, we choose the $z$ axis to be perpendicular to the plate surface. The origin of the coordinate system is located on the surface of one plate. Image charges are neglected, i.e.\ we assumed for simplicity that the dielectric constants of the solvent, the plate and the colloidal material are the same. Typically we use a periodically repeated square cell in $x$ and $y$ direction which possesses an area $S_{p}$. Hence the macroion number density is $\rho_{m}=N_m/2LS_p$. We restrict our studies to a small number of macroions in the cell. In particular we are considering the cases $N_m=0,1,2$ subsequently. Both the macroions and the charged plates provide neutralizing counterions which are dissolved in a solvent of dielectric constant $\epsilon$. The counterions have a microscopic diameter $d_c$ and carry an opposite charge $q_c = -qe$ where $q>0$ denotes the valency. Typically, $q=1,2$. For simplicity, we assume that the counterions from the walls and from the macroions are not distinguishable. The total counterion number $N_c$ in the cell (as well as the averaged counterion number density $\rho_{c}=N_c/2LS_p$) is fixed by the condition of global charge neutrality \begin{equation} \label{neutral1} \rho_{m}q_{m}+\rho_{c}q_{c}+\frac{\sigma_{1}+\sigma_{2}}{2L} = 0. \end{equation} The interactions between the particles are described within the framework of the primitive model. We assume the following pair interaction potentials $V_{mm}(r)$, $V_{mc}(r)$, $V_{cc}(r)$ between macroions and counterions, $r$ denoting the corresponding interparticle distance: \begin{equation} V_{mm} (r) = \cases {\infty &for $ r \leq d_m$\cr {{Z^2e^2} \over {\epsilon r}} &for $ r > d_m$\cr} \label{1aLMH} \end{equation} \begin{equation} V_{mc} (r) = \cases {\infty &for $ r \leq (d_m + d_c) /2$\cr - {{Zqe^2} \over {\epsilon r}} &for $ r > (d_m + d_c) /2$\cr} \label{1bLMH} \end{equation} \begin{equation} V_{cc} (r) =\cases {\infty &for $ r \leq d_c$\cr {{q^2e^2} \over {\epsilon r}} &for $ r > d_c$\cr} \label{1cLMH} \end{equation} The interaction between the particles and the wall is described by the potential energy \begin{equation} V_{pi} (z) =\cases {\infty &for $z<d_i/2$ and \cr & $z>2L-d_i/2$\cr {{2\pi(\sigma_2-\sigma_1)q_i z} \over {\epsilon }}&else \cr} \label{2} \end{equation} where $z$ is the altitude of the particle center and $i=m,c$. Note that the interaction between the wall and the particles is zero for equally charged plates. Our target quantities are the equilibrium counterion profiles and the effective forces exerted on the macroions. The counterionic density profile $\rho_c^{(0)}(\vec r)$ is defined as statistical average via \begin{equation} \rho_c^{(0)}({\vec r})= \sum_{j=1}^{N_c}<\delta ( {\vec r} - {\vec r}_j )>_c \label{3} \end{equation} where $\{ {\vec r}_j=(x_j, y_j, z_j); j=1,...N_c \}$ denote the counterion positions. The canonical average $<...>_c$ over an $\{ {\vec r}_j \}$-dependent quantity $\cal A$ is defined via the classical trace \begin{eqnarray} <{\cal A}(\{ {\vec r}_k \}) >_c =&& {1\over {\cal Z}} {1\over {N_c!}} \int_V d^3r_1...\int_V d^3r_{N_c}\nonumber \\ &&\times {\cal A} (\{ {\vec r}_k \}) \exp \left( - {{V_c}\over{k_BT}} \right) \label{33} \end{eqnarray} where $k_BT$ is the thermal energy ($k_B$ denoting Boltzmann's constant) and \begin{eqnarray} V_c =&&\sum_{n=1}^{N_m} \sum_{j=1}^{N_c} V_{mc}( \mid {\vec R}_n - {\vec r}_j \mid ) \nonumber \\ & & + {{1} \over {2}} \sum_{i,j=1; i\not= j}^{N_c} V_{cc}( \mid {\vec r}_i - {\vec r}_j \mid ) + \sum_{j=1}^{N_c} V_{pc}(z_j) \label{9999} \end{eqnarray} is the total counterionic part of the potential energy provided the macroions are at positions\\ \hbox {$\{ {\vec R}_j=(X_j, Y_j, Z_j); j=1,...N_m \}$}. Furthermore, the classical partition function \begin{equation} {\cal Z} = {1\over {N_c!}} \int_V d^3r_1...\int_V d^3r_{N_c} \exp \left( - {{V_c}\over{k_BT}}\right) \label{neu3} \end{equation} guarantees the correct normalization $<1>_c=1$. Note that the counterionic density profile $\rho_c^{(0)}({\vec r})$ depends parametrically on the macroion positions $\{ {\vec R}_j \}$. The total effective force ${\vec F}_j$ acting onto the $j$th macroion contains three different parts \cite{LMH,allah2,allah} \begin{equation} {\vec F}_j = {\vec F}_j^{(1)} + {\vec F}_j^{(2)} + {\vec F}_j^{(3)} \label{neu4} \end{equation} The first term, ${\vec F}_j^{(1)}$, is the direct Coulomb repulsion stemming from neighboring macroions and the plates \begin{eqnarray} {\vec F}_j^{(1)} = -{\vec \nabla}_{\vec R_j} \left( \sum_{i=1; j\not=i}^{N_m} V_{mm} ( \mid {\vec R}_i - {\vec R}_j \mid ) + V_{pm} (Z_j)\right) \label{neu2} \end{eqnarray} The second part ${\vec F}_j^{(2)}$ involves the electric part of the counterion-macroion interaction and has the statistical definition \begin{equation} {\vec F}_j^{(2)}=< \sum_{i=1}^{N_c} {\vec \nabla}_{\vec R_j} {{Zqe^2} \over {\epsilon \mid {\vec R}_j - {\vec r}_i \mid }} >_c \label{8} \end{equation} Finally, the third term ${\vec F}_j^{(3)}$ describes a depletion (or contact) force arising from the hard-sphere part in $V_{mc}(r)$, which can be expressed as an integral over the surface ${\cal S}_j$ of the $j$th macroion \begin{equation} {\vec F}_j^{(3)}=k_BT \int_{{\cal S}_j} d{\vec f} \ \ \rho_c^{(0)}({\vec r}) \label{9} \end{equation} where ${\vec f}$ is a surface vector pointing towards the macroion center. This depletion term is usually neglected in any DLVO or Poisson-Boltzmann treatment but becomes actually important for strong macroion-counterion coupling. We define the strength of Coulomb coupling via the dimensionless coupling parameter \cite{allah} \begin{equation} \Gamma_{mc} ={Z \over q}{2 \lambda_B \over {d_m+d_c}} , \label{gamma_mc} \end{equation} where the Bjerrum length is $\lambda_B = q^2e^2/\epsilon k_B T$. A further interesting quantity is the counterion-averaged total potential energy defined as \begin{equation} U( \{ {\vec R}_j \} ) = \sum_{i,j;i<j}^{N_m} V_{mm}(\mid {\vec R}_i - {\vec R}_j \mid ) + < V_c >_c \label{neu5} \end{equation} In general the effective force (\ref{neu4}) is different from the gradient of $U( \{ {\vec R}_j \} )$ \cite{Hartmut} i.e. \begin{equation} {\vec F}_j \not= {\vec {\bar F}}_j \equiv -{\vec \nabla}_{\vec R_j}U( \{ {\vec R}_i \} ) \label{neu566} \end{equation} In fact, as we shall show below these two quantities behave qualitatively different for strong coupling. We emphasize that it is the effective force (\ref{neu4}) that is probed in experiments. \section{details of the computer simulation} The Coulomb interactions involved in the primitive model are long-ranged but the periodically repeated system is finite which poses a computational problem. This can be solved in different ways. The simplest way to solve the problem is to cut off the range of the Coulomb interaction by half of the system size which is the minimum image convention (MIC). The MIC is easy to implement but serious cut-off errors can be introduced. A better way is to include $\cal N$ periodically repeated images (PRI) of neighbour cells in $x$ and $y$ direction. Also the limit ${\cal N}\to\infty$ can be treated by a suitable generalization of the traditional Ewald summation technique \cite{Ewald,han_pol,val_koh} to a two-dimensional system. A straightforward generalization, however, leads to quite massive computational effort \cite{halley}. A much more effective alternative is the so-called Lekner summation method \cite{Lekner1,Lekner2} which has recently been applied successfully to the problem of effective interactions between rodlike polyelectrolytes and like-charged planar surfaces \cite{mashl12}. A completely different way out of the problem is to study the system on a surface of a four-dimensional (4D) hypersphere which itself is a compact closed geometry with spherical boundary conditions \cite{izenberg}. Then one has to express the Coulomb forces in terms of the appropriate 4D spherical coordinates which can be done analytically, see Appendix A. Such spherical boundary conditions were effectively utilized in computer simulations of two-dimensional (2D) classical electrons \cite{han_lev,cal_lev} and other 2D fluids \cite{caillol,kratky,cail_ela}. Simulations of the 3D system located on the surface of a 4D hypersphere were carried out for Lennard-Jones \cite{schrei}, hard sphere \cite{toboch} and charged \cite{cail1} systems. The hypersphere geometry (HSG) was also tested against Ewald summations to investigate the stability of charged interfaces \cite{cail2} and good agreement was found, even for strongly coupled interfaces. Simulations in HSG are much faster than that for Lekner sums or PRI as there is no sum over images. In most of our investigations we have used HSG simulations but tested them against MIC, PRI and Lekner summations. Good agreement was found except for the MIC which suffers from the early truncation of the Coulomb tail. We have performed Molecular Dynamics (MD) simulations at room temperature $T=293^o K$. A more detailed description of the MD procedure in HSG is given in Appendix B. The width of planar slit is fixed to be $2L=5 d_m$. Different sets of system parameters are summarized in Table I. \twocolumn[\hsize\textwidth\columnwidth\hsize\csname @twocolumnfalse\endcsname \begin{table} \caption{Set of parameters used in our calculations} \begin{tabular}{lccccccccc} Run& $N_{m}$ & $Z$ & $q$ & $\sigma_{p}(e/cm^{2})$& $\epsilon$&$\rho_m (1/cm^{3})$&$ d_m (cm)$&$ d_c (cm)$&$\Gamma_{mc}$ \\ \tableline A & 0 & - & 2 & $ 1.24\times10^{11}$& 78.3&$1.17\times10^{13}$& - & $5.32\times10^{-8}$&- \\ B & 1 & 200 & 2 & $ 0.62\times10^{11}$& 78.3&$1.17\times10^{13}$&$5.32\times10^{-6}$ & $5.32\times10^{-8}$&11 \\ C & 1 & 200 & 2 & $ 1.24\times10^{11}$& 78.3&$varied$&$5.32\times10^{-6}$ & $5.32\times10^{-8}$&11 \\ D & 1 & 100 & 2 & $ 1.49\times10^{11}$& $varied $&$1.17\times10^{13}$&$5.32\times10^{-6}$ & $5.32\times10^{-8}$&$varied$ \\ E & 1 & 100 & 2 & $ 2.98\times10^{11}$& 3.9&$1.17\times10^{13}$&$5.32\times10^{-6}$ & $5.32\times10^{-8}$&110 \\ G & 1 & 100 & 2 & $varied$& 78.3&$9.36\times10^{16}$&$2.66\times10^{-7}$ & $2.66\times10^{-8}$&100 \\ K & 1 & 32 & 2 & $ 1.56\times10^{14}$& 77.3&$1.9\times10^{18}$&$2.56\times10^{-7}$ & $2.56\times10^{-9}$&37 \\ L & 2 & 200 & 2 & $ 1.24\times10^{11}$& 78.3&$2.34\times10^{13}$&$5.32\times10^{-6}$ & $5.32\times10^{-8}$&11 \\ M & 2 & 100 & 2 & $ varied$& 3.9&$2.34\times10^{13}$&$5.32\times10^{-6}$ & $5.32\times10^{-8}$&110 \\ N & 2 & 100 & 2 & $varied$& 78.3&$1.87\times10^{17}$&$2.66\times10^{-7}$ & $2.66\times10^{-8}$&100 \\ \end{tabular} \end{table} ] We take divalent counterions throughout our investigations. The dielectric constant is that for water at room temperature $(\epsilon=78.3$) but we have also investigated cases where $\epsilon$ is smaller in order to enhance the Coulomb coupling formally. The charge asymmetry $Z/q$ ranges from 16 to 100. The time step $\triangle{t}$ of the simulation was typically chosen to be $10^{-3}\,\sqrt{m\,{d^{3}_{m}}/e^{2}}$, (with $m$ denoting the mass of the counterions) such that the reflection of counterions following the collision with the surface of macroions and walls is calculated with high precision. For every run the equilibrium state of the system was checked during the simulation time. This was done by monitoring the temperature, average velocity and the distribution function of velocities and total potential energy of the system. On average it took about $10^4$ MD steps to get into equilibrium. Then during $5\cdot 10^4-5\cdot 10^5$ time steps, we gathered statistics to perform the canonical averages for calculated quantities. \section{counterion density profiles between charged plates} First, as a reference case, let us discuss the situation without any macroion. This set-up is well-studied in the literature \cite{Robbins,Valleau}. We consider equally charged surfaces $\sigma_1 = \sigma_2 =\sigma_p$. The imbalance in the interaction with neighbours will push the counterions toward the plates. Consequently, a great majority of neutralizing counterions reside within a thin surface layer. For strong coupling, the width of this layer can be approximately expressed as \cite{rouzina} \begin{equation} \label{lambda} \lambda_z = \frac {\lambda_{D}^{2}} {2L}, \end{equation} where $\lambda_D$ is the bulk Debye screening length \begin{equation} \label{lambda2} \lambda_{D}^{2} = \frac {\epsilon k_B T} {4 \pi \rho_0 q^2 e^2} \end{equation} where $\rho_0\equiv \rho_c$. Due to symmetry, the equilibrium counterion density profile only depends on $z$. The analytical solution of the Poisson-Boltzmann (PB) equation for this profile is \cite{eng} \begin{equation} \label{hb} \rho^{(PB)}_{c}(z)= \rho_c \frac {2\gamma_{0}^{2} \lambda_{D}^{2} }{L^{2} \cos ^2 (\gamma_{0} (1-{z \over L}))} \end{equation} where $\gamma_0$ is defined via the solution of the implicit equation \begin{equation} \label{defgamma} \frac {(L/\lambda_D)^{2}} {2\gamma_0} - \tan{\gamma_0} = 0 \end{equation} For parameters of moderate Coulomb coupling (run A), the PB result is shown as a solid line in Fig.\ref{fig1allah}. \begin{figure} \epsfxsize=6cm \epsfysize=6cm ~\hfill\epsfbox{allah1.ps}\hfill~ \caption{Reduced counterion density profiles $\rho^{(0)}_c(z)/\rho_0$ versus reduced distance $z/L$. solid line- PB prediction and simulation result with incorporating Lekner summation method. Both data do coincide on the same curve, long-dashed line- result of simulation in HSG, dashed line- result of PRI simulation with ${\cal N}=2$, dotted line- result of MIC simulation.} \label{fig1allah} \end{figure} The corresponding MD simulation data were obtained with 600 counterions in the cell using various boundary conditions. As expected, the PB theory coincides with the Lekner summation method which treats best the long-range nature of the Coulomb interactions. In HSG the counterionic profiles are also very similar to the Lekner summation while the MIC deviates significantly. The MIC can already be improved significantly if ${\cal N}=2$ periodic repeated images are included. In conclusion, the agreement between Lekner summation and HSG justifies the HSG a posteriori and gives evidence that the HSG produces reliable results also for stronger couplings. \section{Single macroion between charged plates} Let us now consider a single macroion in the inter-lamellar area. We put the macroion on the $z$-axis, such that its position is at ${\vec R}_{1}=(0,0,Z_1)$. A corresponding schematic picture is given in Fig.\ref{fig2allah}. \begin{figure} \epsfxsize=6.5cm \epsfysize=5cm ~\hfill\epsfbox{allah2.ps}\hfill~ \caption{Schematic picture for a single macroion between likely charged planes of charge density $\sigma_p$, separated by distance $2L$.} \label{fig2allah} \end{figure} The total force acting on the macroion only depends on $Z_1$ and points along the $z$-axis. Obviously, for the case $\sigma_1=\sigma_2=\sigma_p$ of symmetric plates considered here, the direct part of the total force, ${\vec F}_j^{(1)}$, vanishes. For the second (electrostatic) part, simple PB-theory applied to the case of small macroion charge and small macroion diameter yields the following analytical expression for the effective macroion force \cite{eng,anne} \begin{equation} \label{pbforce} {\vec F}_{1}^{PB}=\frac {2 Z k_{B} T \gamma_{0}} {q L} \tan \left(\gamma_{0} \left(1-z/L \right) \right) {\vec e}_z \end{equation} where ${\vec e}_z$ is the unit vector in $z$-direction and $\gamma_0$ is given by (\ref{defgamma}). Note that only the counterion density stemming from the charged plates has to be inserted in (\ref{defgamma}) i.e. $\rho_o=\frac{\sigma_p}{L\mid qe \mid}$. This force pushes the macroion towards the mid-plane, i.e.\ the wall-particle interaction is repulsive. The expression (\ref{pbforce}) will break down, however, for a large macroion diameter $d_m$ and for strong macroion-counterion coupling parameter $\Gamma_{mc}$. We have tested the PB-theory against ``exact" computer simulation data. For moderate couplings (run B and run C) the results for the total force $F_1={\vec F}_1 \cdot {\vec e}_z$ are shown in Figs.\ref{fig3allah}-\ref{fig4allah}. In Fig.\ref{fig3allah}, a surprising agreement between theory and simulation is obtained despite the fact that the macroion charge is large. \begin{figure} \epsfxsize=6cm \epsfysize=7cm ~\hfill\epsfbox{allah3.ps}\hfill~ \caption{Force $F_1={\vec F}_1 \cdot {\vec e}_z$ acting on a single macro-ion versus reduced macro-ion distance $Z_1/L$. The force is scaled by the (arbitrary) unit $F_{0}= \frac{Zqe^{2}} {\epsilon {d_{m}^{2}}}$. The system parameters are from run B\@. The solid line is the prediction from PB theory. The open circles are simulation results in HSG\@. The statistical error corresponds to the symbol size.} \label{fig3allah} \end{figure} This justifies the theoretical conclusions drawn in Refs.\ \cite{anne,Jay} based on PB theory. The deviation between theory and simulation are larger in Fig.\ref{fig4allah} where the surface charge density was doubled. Here, also the system size dependence (resp.\ the dependence on the macroion density) was studied in the simulation. \begin{figure} \epsfxsize=6cm \epsfysize=7cm ~\hfill\epsfbox{allah4.ps}\hfill~ \caption{Same as Fig.\ref{fig3allah} but now for run C. Symbols are simulation results in HSG for various macroion number densities: circles: $\rho_m = 1.17\times 10^{13} cm^{-3}$, squares: $\rho_m = 2.0\times 10^{12} cm^{-3}$, triangles: $\rho_m = 1.0\times 10^{12} cm^{-3}$.} \label{fig4allah} \end{figure} As expected the agreement becomes better for a larger system size (resp.\ for a smaller macroion density) since the theory is constructed formally for vanishing macroion density. In addition, we repeated the calculations for $\rho_m = 1.17\times 10^{13} cm^{-3}$ (corresp.\ to the circles in Fig.\ref{fig4allah}) using the PRI method with $N=4$ and got the same results as in HSG\@. We now enhance the Coulomb coupling by formally reducing the dielectric constant $\epsilon$. For a fixed macroion position at $Z_1=d_m$, the force $F_1$ is shown in Fig.\ref{fig5allah} for the parameters of run D. \begin{figure} \epsfxsize=6cm \epsfysize=7cm ~\hfill\epsfbox{allah5.ps}\hfill~ \caption{Force $F_1={\vec F}_1 \cdot {\vec e}_z$ acting on a single macro-ion versus dielectric constant $\epsilon$ for a fixed position at $Z_1 = d_m $. The force is scaled by the (arbitrary) unit $F_{0}=0.01 \frac{Zqe^{2}} {\epsilon {d_{m}^{2}}}$. The parameters of the system are from run D. The solid line is the prediction of PB theory. The crosses are simulation results in HSG.} \label{fig5allah} \end{figure} While the PB theory always predicts a repulsive force, the simulation data are in accordance with theory only for large $\epsilon$ but the force changes its sign for $\epsilon <10$. Hence as expected the theory breaks down for large Coulomb coupling where correlation between the counterions become significant. For the same run D, the distance-resolved macroion force $F_1$ is shown in Fig.\ref{fig6allah} for a strongly reduced dielectric constant $\epsilon=3.9$. The simulation data were obtained in HSG but confirmed by PRI calculations with ${\cal N}=4$. The electrostatic part $F_1^{(2)}={\vec F}_1^{(2)} \cdot {\vec e}_z$ and the depletion part $F_1^{(3)}={\vec F}_1^{(3)} \cdot {\vec e}_z$ are shown separately. $F_z^{(3)}$ is always repulsive and increases with decreasing $Z_1$, at least if the macroion is not too close to the surface when the counterion depletion between the macroion and the wall induced by the finite counterion core is negligible. This is an expected behavior, since in general there are more counterions close to the walls. The pure electrostatic contribution, $F_1^{(2)}$, on the other hand, exhibits a more subtle behavior. If the macroion is close to the midplane, it is repulsive, then it becomes attractive as the macroion is getting closer to the plates. \begin{figure} \epsfxsize=6cm \epsfysize=7cm ~\hfill\epsfbox{allah6.ps}\hfill~ \caption{Same as Fig.\ref{fig3allah} but now for run D, $\epsilon=3.9$, and for a force unit of $F_{0}=0.1 \frac{Zqe^{2}} {\epsilon {d_{m}^{2}}}$. The crosses are simulation data in HSG for the total force $F_1$, the squares (resp.\ circles) are simulation data the electrostatic part $F_1^{(2)}$ (resp.\ the depletion part $F_1^{(3)}$).The line is a guide to the eye for the total force. The inset shows the effective potential in units of $k_BT$ versus reduced macro-ion distance $Z_1/L$ together with the energy barrier $\Delta V_{eff}$. The solid line is for run D with $\epsilon =3.9$, dashed line is for run E.} \label{fig6allah} \end{figure} As a function of macroion distance, the total force $F_1$ is repulsive, attractive and becomes repulsive again. For small separations (which are still larger than the microscopic counterionic core) the force is dominated by the repulsive depletion force. Hence the macroion has got three equilibrium positions, two of them are stable, namely the midplane and a position in the vicinity of the plate. In order to extract more information, we have calculated the effective wall-particle potential defined by \begin{equation} V_{eff}(Z_1) = - \int_{0}^{Z_1} F_1(h) dh \label{eff} \end{equation} by integrating our data with respect to the macroion altitude $h$. This quantity is shown as an inset in Fig.\ref{fig6allah}. One first sees that the global minimum is in the vicinity of the walls. Furthermore the barrier height $\Delta V_{eff}$ to escape from there is about $8 k_BT$. This implies that the time for a colloidal particle to escape from the position close to the surface is roughly $\tau_0 \exp \left(\Delta V_{eff}/k_BT \right) = e^8 \tau_0\approx 3000 \tau_0$ \cite{Russel,Sauer} where $\tau_0$ is a Brownian time scale governing the decay of dynamical correlations of the macroion. It can also be seen that, for a doubled surface charge (run E), the height of barrier increases. A similar behaviour occurs for another parameter combinations (run G), see Fig.\ref{fig7allah}, corresponding to aqueous suspensions of micelle-sized macroions. It hence seems to be a generic feature of the primitive model for strong Coulomb coupling. \begin{figure} \epsfxsize=6cm \epsfysize=7cm ~\hfill\epsfbox{allah7.ps}\hfill~ \caption{Same as Fig.\ref{fig6allah} but now for run G and for a force unit of $F_{0}=0.1 \frac{Zqe^{2}} {\epsilon {d_{m}^{2}}}$. The squares are simulation results for the total force in HSG for $\sigma_p =1.19\times10^{14} { e \over cm^{2}} $, while the circles are for $\sigma_p =2.38\times10^{14} {e \over cm^{2}}$. The lines are a guide to the eye. The inset shows the effective potential in units of $k_BT$ versus reduced macro-ion distance $Z_1/L$. The dashed line is for $\sigma_p =1.19\times10^{14} { e \over cm^{2}} $, the solid line is for $\sigma_p =2.38\times10^{14} {e \over cm^{2}}$.} \label{fig7allah} \end{figure} \begin{figure} \epsfxsize=6cm \epsfysize=7cm ~\hfill\epsfbox{allah8.ps}\hfill~ \caption{Effective force $F_1={\vec F}_1 \cdot {\vec e}_z$ (circles) and gradient of the potential energy $\bar F_1={\vec {\bar F}}_1 \cdot {\vec e}_z$ (squares) versus reduced macro-ion distance $Z_1/L$ for run K. The unit of the force is $F_{0}=0.1 {{Zqe^{2}} \over {\epsilon d_{m}^{2}} }$. the lines are a guide to the eye. The dashed line are data from Ref.[60]} \label{fig8allah} \end{figure} We note that the barrier height $\Delta V_{eff}$ is about $70 k_B T$ which implies a very large escape time. Finally we show for a certain parameter combination (run K) which was also used in Ref.\ \cite{bo} that the averaged force ${\vec F}_1$ differs from the gradient of the averaged potential energy ${\vec {\bar F}}_1$. As in Ref.\ \cite{bo}, the system consists of a single macroion in a planar slit of width $5d_m$, with one charged and one neutral wall. Results are given in Fig.\ref{fig8allah}. We conclude that the forces behave even qualitatively different. The average force ${\vec F}_1$ is a short-range attractive force which becomes repulsive only for touching macroion configurations. Contrary to that, the force ${\vec {\bar F}}_1$ is repulsive up to distance about $d_m/2$ from the wall surface. Note that our data actually differ from those of Ref.\ \cite{bo} due to the early truncation of the Coulomb interaction performed there. \section{Two macro-ions between plates} We finally consider two equally charged macroions at the positions ${\vec R}_{1}=(X_1 ,Y_1 ,Z_1)$ and ${\vec R}_{2}=(X_2 ,Y_2 ,Z_2)$ confined between plates. A schematic picture is given in Fig.\ref{fig9allah}. \begin{figure} \epsfxsize=7.5cm \epsfysize=6cm ~\hfill\epsfbox{allah9.ps}\hfill~ \caption{Schematic picture for the macroion pair near a charged wall of surface charge density $\sigma_p$. For the sake of clarity, the position of second wall is omitted. The different forces are shown for the case of a mutual attraction.} \label{fig9allah} \end{figure} In order to reduce the parameter space, we assume for simplicity that both macroions have the same altitude $Z_1=Z_2$. The distance between the macroion centers is $R_{12} =\mid {\vec R}_1 - {\vec R}_2 \mid $ where the difference vector ${\vec R}_{12}={\vec R}_1 - {\vec R}_2$ is in the $xy$-plane. We assume that only one of the walls is charged and that the second wall is neutral. This gives us the possibility to simulate higher surface charge densities. Also for strong coupling, the counterions of the two different walls are practically decoupled such that the set-up with a single charged wall is not expected to differ much from the symmetrical set-up. The total force acting on the two macroions can be split into a part pointing in $z$-direction and another contribution pointing along ${\vec R}_{12}$. Hence we write ${\vec F}_j = {\vec F}^\|_j + {\vec F}^\bot_j$ defining ${\vec F}^\|_j= \left({\vec F}_j \cdot {\vec R}_{12} \right) \cdot {\vec R}_{12}/R_{12}^2$ and ${\vec F}^\bot_j= \left( {\vec F}_j \cdot {\vec e}_z \right) \cdot {\vec e}_z$ for $j=1,2$. Clearly, ${\vec F}^\bot_1={\vec F}^\bot_2$, and ${\vec F}^\|_1=-{\vec F}^\|_2$. It is instructive to compare these force to the DLVO bulk theory which yields \begin{eqnarray} {\vec F}^{DLVO}_1 =&& \frac {{Z}^2 e^2 \exp{((d_m - R_{12})/ \lambda_{D}) } } {\epsilon R_{12}(1+ {d_m/2 \lambda_D})^2} \nonumber\\ &&\times \left({1 \over R_{12}}+{1 \over \lambda_D} \right) {{\vec R}_{12} \over R_{12}} \label{dlvo} \end{eqnarray} Here the Debye screening length $\lambda_D$ is given by eq.(\ref{lambda2}), where $\rho_0$ corresponds to the counterion number density coming only from the macroions, $\rho_{0} = {Z \over {q}}\rho_{m}$. One can also modify the DLVO theory by admitting screening also from the counterions stemming from the wall assuming they follow a Poisson-Boltzmann density profile. This yields the PB force \cite{anne,Jay} \begin{equation} {\vec F}^{PB}_1 = ({\vec F}^{PB}_1)^\| + ({\vec F}^{PB}_1)^\bot \label{PBA} \end{equation} where we get for the parallel part of the force \begin{eqnarray} \label{PBpar} ({\vec F}_1^{PB})^\| =&& \frac {{Z}^2 e^2 \exp{(- R_{12}/ \lambda_{D}(Z_1))} } {\epsilon R_{12}} \nonumber \\ &&\times \left({1 \over R_{12}}+{1 \over \lambda_D(Z_1)}\right) {{\vec R}_{12} \over R_{12}}. \end{eqnarray} The perpendicular part of the force is \begin{eqnarray} \label{PBper} ({\vec F}^{PB}_1)^\bot =&& {\vec F}_1^{PB} - \frac {{Z}^2 e^2 } {\epsilon} \frac{\lambda_D(Z_1) \gamma_0^3}{L^3} \nonumber \\ && \times \frac {\tan \left(\gamma_0 (1-Z_1/L)\right)} {\cos^2 \left(\gamma_0 (1-Z_1/L)\right)} {\vec e}_z \end{eqnarray} \begin{figure} \epsfxsize=6cm \epsfysize=7cm ~\hfill\epsfbox{allah10.ps}\hfill~ \caption{Parallel part of the effective force acting onto a macroion pair, $F^\|_1={\vec F}^\|_1\cdot {{\vec R}_{12} / R_{12}}$, versus reduced interparticle distance $R_{12}/d_m$. The unit of the force is $F_{0}=\left( \frac{Z^{2} e^{2}} {\epsilon {d_{m}^{2}}}\right )\,\times 10^{-2}$. The parameters of system are from run L and for an altitude of macroions of $Z_1=0.6 d_m$. The solid line is the bulk DLVO theory, the dashed line is the Poisson-Boltzmann theory (\ref{PBpar}) and the points are simulation results in HSG\@. The statistical error corresponds to the symbol size.} \label{fig10allah} \end{figure} The space dependent Debye screening length is \begin{equation} \label{ldz} \lambda_D(Z_1) = \left( 4 \pi \lambda_B \left( {Z \over q}\rho_m + \rho_c^{PB}(Z_1)\right) \right)^{-\frac{1}{2}} \end{equation} Here $\rho_c^{PB}(z)$ and ${\vec F}_1^{PB}$ are given by Eqns.\ (\ref{hb}) and (\ref{pbforce}). Contrary to the bulk DLVO force, the PB force has an additional perpendicular part for a pair of particles (second term in (\ref{PBper})). This additional force is attractive. Still the total perpendicular force (\ref{PBper}) is always repulsive. For the parameters of run L corresponding to weak coupling, simulation results for $F^\|_1={\vec F}^\|_1 \cdot {{\vec R}_{12} / R_{12}}$ are shown in Fig.\ref{fig10allah}. The solid line corresponds to the bulk DLVO force, and the dashed line is the Poisson-Boltzmann result. The force is repulsive both in theory and simulation, but the theories overestimate the force significantly. As expected the Poisson-Boltzmann approach yields better agreement than DLVO bulk theory. Results for $F^\|_1$ for stronger coupling (runs M and N) are displayed in Figs.\ref{fig11allah}-\ref{fig12allah}. \begin{figure} \epsfxsize=6cm \epsfysize=7cm ~\hfill\epsfbox{allah11.ps}\hfill~ \caption{Same as Fig.\ref{fig10allah} but now for run M and $F_{0}=\left( \frac{Z^{2} e^{2}} {\epsilon {d_{m}^{2}}}\right )\,\times 10^{-3}$. Simulation results are shown for three different surface charges: squares: $\sigma_p = 0 {e \over cm^{2}}$, triangles: $\sigma_p = 2.98\times10^{11} {e \over cm^{2}}$, circles: $\sigma_p = 5.95\times10^{11} {e \over cm^{2}}$. The lines are a guide to the eye.} \label{fig11allah} \end{figure} For a neutral wall, the interaction force between macroions is already attractive. With increasing surface charge the attraction between macroions becomes stronger. Clearly this attraction is neither contained in DLVO theory nor in the Poisson-Boltzmann approach (\ref{PBpar}). \begin{figure} \epsfxsize=6cm \epsfysize=7cm ~\hfill\epsfbox{allah12.ps}\hfill~ \caption{Same as Fig.\ref{fig11allah} but now for run N and for $Z_1 =0.7\, d_m$. Results are shown for three different surface charges: squares: $\sigma_p = 0 {e \over cm^{2}}$, triangles: $\sigma_p = 1.19\times10^{14} {e \over cm^{2}}$, circles: $\sigma_p = 2.38\times10^{14} {e \over cm^{2}}$.} \label{fig12allah} \end{figure} In Fig.\ref{fig13allah} we fixed the macroion distance and calculated $F^\|_1$ and the force perpendicular to the plates, $F^\bot_1={\vec F}^\bot_1 \cdot {\vec e}_z$ versus altitude $Z_1$ for run N. \begin{figure} \epsfxsize=6cm \epsfysize=7cm ~\hfill\epsfbox{allah13.ps}\hfill~ \caption{Parallel $F^\|_1={\vec F}^\|_1\cdot {{\vec R}_{12} / R_{12}}$ ( squares) and perpendicular $F^\bot_1={\vec F}^\bot_1\cdot {\vec e}_z$ (circles) parts of effective force versus reduced altitude $Z_1/d_m$ for fixed interparticle spacing $R_{12}=1.2\,d_m$. The unit of the force $F^\|_1$ is $F_{0}=\left( \frac{Z^{2} e^{2}} {\epsilon {d_{m}^{2}}}\right )\,\times 10^{-3}$, and for the force $F^\bot_1$ is $F_{0}^*=\left( \frac{Z^{2} e^{2}} {\epsilon {d_{m}^{2}}}\right )\,\times 10^{-2}$. The surface charge density is $\sigma_p = 2.38\times10^{14} {e \over cm^{2}}$. The inset shows the effective potential in units of $k_BT$ versus reduced macro-ion distance $Z_1/d_m$. } \label{fig13allah} \end{figure} There is attraction. Both the interparticle attraction and the wall-particle attraction become stronger in the vicinity of the plate. The effective wall-particle interaction potential for the perpendicular part is shown as an inset in Fig.\ref{fig13allah}. Note that the minimum of $V_{eff}$ is much more than twice as large as in the single macroion case (compare to inset in Fig.\ref{fig7allah}, solid line). Thus, a pair of macroions near a planar surface is more stable than a single macroion. This is also evident from the results for run N shown in Fig.\ref{fig14allah}. \begin{figure} \epsfxsize=6cm \epsfysize=7cm ~\hfill\epsfbox{allah14.ps}\hfill~ \caption{Perpendicular part of the effective force acting onto a macroion pair, $F^\bot_1={\vec F}^\bot_1\cdot {\vec e}_z$ in units of $F_{0}=\left( \frac{Z^{2} e^{2}} {\epsilon {d_{m}^{2}}}\right )\,\times 10^{-3}$ versus dimensionless interparticle distance $R_{12}/d_m$. The parameters of system are from run N and the altitude of macroions is fixed to $Z_1=0.7\, d_m$. Simulation results are shown for two different surface charges: squares: $\sigma_p = 1.19\times10^{14} {e \over cm^{2}}$, circles: $\sigma_p = 2.38\times10^{14} {e \over cm^{2}}$.} \label{fig14allah} \end{figure} Again there is attraction towards the plate for varied $R_{12}$ and fixed $Z_1$. The attraction becomes stronger if the interparticle distance is decreasing. This shows that the attraction between the wall and a single macroion discussed in chapter V is stable and even enhanced if more macroions are close to the wall. This leads us to the final conclusion that the macroions will assemble on top of the surface forming two-dimensional colloidal layers. \section{Conclusions} We have simulated the effective force between macroions confined in a slit-geometry. An effective attraction was found for strong Coulomb coupling. In particular, the effective potential of a single macroion confined between two parallel charged plates was found to have two stable minima where the total force vanishes: the first is in the mid-plane, the second close to the walls. This result was confirmed for two macroions. In this case the attraction towards the walls was even stronger than for a single macroion. Our most important conclusion is that the attractive force will result in two-dimensional colloidal layers on top of the plates. As the depth of the attractive potential is larger than $k_B T$, these layers possess a large life-time with respect to thermal fluctuations. The layers should be crystalline as the interparticle interaction is also attractive. This can explain at least qualitatively the long-lived metastable crystalline layers found in recent experiments on confined samples of charge colloidal suspensions \cite{Larsen_crystallites,larsen}. We want to add some remarks: First, our parameters are actually different form those describing the experiments. The main difference is the high surface charge of the glass plates within an area spanned by a typical macroion separation distance. Such a system cannot be simulated since it requires a huge number of counterions in the simulation box. We have mimicked the high surface charge by dealing with a small dielectric constant, but strictly speaking this corresponds to a different system. Second, the mechanism of our attraction is similar to that proposed recently by us in the bulk case \cite{allah}. It only occurs for strong coupling with divalent counterions and is short-ranged. In this respect, it behaves different than in experiment where the attraction was long-ranged. We emphazise again that the depletion force is crucial in the strong coupling parameter regime. Third, our computer simulation data were tested against simple DLVO- or Poisson-Boltzmann-type theories. It would be interesting to use them as benchmark data for more sophisticated theoretical approaches which predict attraction as e.g.\ the density functional perturbation theory recently proposed by Goulding and Hansen \cite{dave23}. Finally, in our simulations, we neglected any impurity or added salt ions. Their inclusion increases substantially the number of microscopic particles and would lead to more extensive simulation. A further challenge would be to incorporate image charges properly into the model which requires a non-trivial extension of our approach. \acknowledgments We thank D. Goulding, J. P. Hansen, C. N. Likos for helpful comments. Financial support from the Deutsche Forschungsgemeinschaft within (SFB 237 and Schwerpunkt Benetzung und Strukturbildung an Grenzfl{\"a}chen) is gratefully acknowledged. We also thank the IFF at the Forschungszentrum J{\"u}lich for providing CPU time.
\section{Introduction} One of the least expected achievements of the Infra-Red Astronomical Satellite (IRAS; Neugebauer et al.\ 1984) was the detection of a large number of mid-IR point sources in the Large Magellanic Cloud (LMC) just above its limits of sensitivity (IRAS Point Source Catalogue; Schwering \& Israel 1990). Many of these are candidates for intermediate-mass stars at the tip of the Asymptotic Giant Branch (AGB). Their lives drawing to a close, these stars are shedding their stellar mantles at rates of up to $10^{-4}$ M$_\odot$ yr$^{-1}$. Their dusty circumstellar envelopes (CSEs) obscure the optical light from the star and become very bright IR objects. The details of the evolution and mass loss of AGB stars are poorly understood. The study of galactic samples of AGB stars is severely hampered by the difficulty to determine accurate distances to stars in the Milky Way. The distance to the LMC, however, is well known and hence luminosities and mass-loss rates of AGB stars in the LMC may be determined with a high degree of accuracy. Early explorations of the IRAS data in combination with ground-based near-IR observations resulted in the first identifications of mid-IR sources in the LMC with obscured AGB stars (Reid et al.\ 1990; Wood et al.\ 1992). We have successfully increased the sample of known AGB counterparts of IRAS sources in the LMC from a dozen to more than 50 stars (Loup et al.\ 1997; Zijlstra et al.\ 1996; van Loon et al.\ 1997, 1998a: Papers I to IV). We attempted to classify their photospheres and CSEs as oxygen- or carbon-dominated, but for the majority of the stars this could not be done conclusively. There remained therefore considerable uncertainty about the luminosity distribution of the obscured carbon stars. This information is important for testing current understanding of the evolution of AGB stars, including dredge-up of carbon and nuclear burning at the bottom of the convective mantle (Hot Bottom Burning, HBB). \begin{table*} \caption[]{IRAS detected stars observed with ISO: names (LI stands for LI-LMC (Schwering \& Israel 1990), TRM is from Reid et al.\ (1990), HV is from Payne-Gaposchkin (1971), SP is from Sanduleak \& Philip (1977) and WOH is from Westerlund et al.\ (1981); sources will be referenced hereafter by their bold-faced names), ISO pointing coordinates (J2000), and references: 1: Hodge \& Wright (1969); 2: Eggen (1971); 3: Wright \& Hodge (1971); 4: Dachs (1972); 5: Sandage \& Tammann (1974); 6: Glass (1979); 7: Humphreys (1979); 8: Blanco et al.\ (1980); 9: Feast et al.\ (1980); 10: Bessell \& Wood (1983); 11: Wood et al.\ (1983); 12: Rebeirot et al.\ (1983); 13: Prevot et al.\ (1985); 14: Elias et al.\ (1985); 15: Wood et al.\ (1985); 16: Elias et al.\ (1986); 17: Wood et al.\ (1986); 18: Reid et al.\ (1988); 19: Reid (1989); 20: Hughes (1989); 21: Hughes \& Wood (1990); 22: Reid et al.\ (1990); 23: Hughes et al.\ (1991); 24: Wood et al.\ (1992); 25: Roche et al.\ (1993); 26: Groenewegen et al.\ (1995); 27: Zijlstra et al.\ (1996); 28: Ritossa et al.\ (1996); 29: van Loon et al.\ (1996); 30: van Loon et al.\ (1997); 31: Loup et al.\ (1997); 32: Oestreicher (1997); 33: van Loon et al.\ (1998a); 34: Groenewegen \& Blommaert (1998); 35: van Loon et al.\ (1998b); 36: van Loon et al.\ (1999)} \begin{flushleft} \begin{tabular}{llllllll} \hline\hline LI & IRAS & TRM & HV & RA (2000) & Decl (2000) & Other names & References \\ \hline \multicolumn{8}{c}{{\em IRAS detected stars}} \\ \hline 1825 & {\bf 04286$-$6937} & -- & -- & 04 28 30.3 & $-$69 30 49 & -- & 27,31,33 \\ 1844 & {\bf 04374$-$6831} & -- & -- & 04 37 22.8 & $-$68 25 03 & -- & 27,31,33 \\ 4 & {\bf 04407$-$7000} & -- & -- & 04 40 28.4 & $-$69 55 13 & -- & 27,31,33 \\ 57 & {\bf 04496$-$6958} & -- & -- & 04 49 18.6 & $-$69 53 14 & -- & 27,31,33,34,36 \\ 60 & {\bf 04498$-$6842} & -- & -- & 04 49 41.4 & $-$68 37 50 & -- & 27,31,33,36 \\ 77 & {\bf 04509$-$6922} & -- & -- & 04 50 40.2 & $-$69 17 33 & -- & 24,27,28,33,36 \\ 92 & {\bf 04516$-$6902} & -- & -- & 04 51 28.4 & $-$68 57 53 & -- & 24,27,33 \\ 121 & {\bf 04530$-$6916} & -- & -- & 04 52 45.3 & $-$69 11 53 & -- & 24,27,28 \\ 141 & {\bf 04539$-$6821} & -- & -- & 04 53 46.3 & $-$68 16 12 & -- & 27,31,33 \\ 153 & 04544$-$6849 & -- & -- & 04 54 14.4 & $-$68 44 13 & {\bf SP77 30-6}, WOH SG66 & 12,13,20,21,27,31 \\ 159 & {\bf 04545$-$7000} & -- & -- & 04 54 09.8 & $-$69 56 00 & -- & 24,27 \\ 181 & 04553$-$6825 & -- & -- & 04 55 10.1 & $-$68 20 35 & {\bf WOH G64} & 16,17,19,24,25,27,29,31,33,35,36 \\ 198 & {\bf 04557$-$6753} & -- & -- & 04 55 38.9 & $-$67 49 10 & -- & 27,31,33 \\ 203 & 04559$-$6931 & -- & {\bf 12501} & 04 55 41.6 & $-$69 26 25 & SP77 31-20, WOH SG097 & 11,12,13,20,22,27,32,33 \\ 297 & {\bf 05003$-$6712} & -- & -- & 05 00 18.9 & $-$67 08 02 & -- & 27,30,31,33,36 \\ 310 & {\bf 05009$-$6616} & -- & -- & 05 01 03.8 & $-$66 12 40 & -- & 27,31,33,36 \\ 383 & 05042$-$6720 & 48 & {\bf 888} & 05 04 14.3 & $-$67 16 17 & SP77 29-33, WOH SG140 & 5,6,7,11,14,18,22,31,32 \\ 570 & {\bf 05112$-$6755} & 4 & -- & 05 11 10.1 & $-$67 52 17 & -- & 22,27,31,33,36 \\ 571 & {\bf 05113$-$6739} & 24 & -- & 05 11 13.7 & $-$67 36 35 & -- & 22,27,31,33 \\ 578 & -- & {\bf 72} & -- & 05 11 41.2 & $-$66 51 12 & -- & 22,27,31,33 \\ 612 & 05128$-$6728 & 43 & {\bf 2360} & 05 12 46.4 & $-$67 19 37 & SP77 37-24, WOH SG193 & 3,5,6,7,11,14,18,22,31 \\ 1880 & {\bf 05128$-$6455} & -- & -- & 05 13 04.6 & $-$64 51 40 & -- & 27,31,33,36 \\ 663 & 05148$-$6730 & 36 & {\bf 916} & 05 14 49.9 & $-$67 27 19 & SP77 37-35, WOH SG204 & 1,2,4,6,7,11,18,22,31,32 \\ 793 & {\bf 05190$-$6748} & 20 & -- & 05 18 56.7 & $-$67 45 06 & -- & 22,27,31,33 \\ -- & -- & {\bf 88} & -- & 05 20 20.9 & $-$66 36 00 & -- & 22,27,31,33,34,36 \\ -- & -- & {\bf 45} & -- & 05 28 16.3 & $-$67 20 55 & -- & 22,27,31,33 \\ 1157 & {\bf 05295$-$7121} & -- & -- & 05 28 40.8 & $-$71 19 13 & -- & 27,31 \\ 1130 & {\bf 05289$-$6617} & 99 & -- & 05 29 02.6 & $-$66 15 31 & -- & 22,27,31 \\ 1145 & -- & -- & {\bf 5870} & 05 29 03.5 & $-$69 06 47 & SP77 47-17, WOH SG331 & 9,11,20,31 \\ 1153 & {\bf 05294$-$7104} & -- & -- & 05 28 47.8 & $-$71 02 29 & -- & 24,27,31 \\ 1164 & {\bf 05298$-$6957} & -- & -- & 05 29 24.5 & $-$69 55 14 & -- & 24,27,31,36 \\ 1177 & {\bf 05300$-$6651} & 79 & -- & 05 30 04.2 & $-$66 49 23 & -- & 22,27,31,36 \\ 1238 & 05316$-$6604 & 101 & -- & 05 31 45.9 & $-$66 03 51 & {\bf WOH SG374} & 22,27,31 \\ 1281 & 05327$-$6757 & 5 & {\bf 996} & 05 32 36.0 & $-$67 55 08 & SP77 46-59, WOH SG388 & 7,11,17,18,22,31 \\ 1286 & {\bf 05329$-$6708} & 60 & -- & 05 32 52.5 & $-$67 06 25 & -- & 17,22,24,26,27,31,33,36 \\ 1345 & {\bf 05348$-$7024} & -- & -- & 05 34 16.1 & $-$70 22 53 & -- & 27,31,33 \\ 1382 & {\bf 05360$-$6648} & 77 & -- & 05 36 03.3 & $-$66 46 47 & -- & 22,27,31,33 \\ 1506 & {\bf 05402$-$6956} & -- & -- & 05 39 44.6 & $-$69 55 18 & -- & 24,27 \\ 1756 & {\bf 05506$-$7053} & -- & -- & 05 50 09.1 & $-$70 53 12 & -- & 27,31,33 \\ 1790 & {\bf 05558$-$7000} & -- & -- & 05 55 20.8 & $-$70 00 05 & -- & 27,31 \\ 1795 & {\bf 05568$-$6753} & -- & -- & 05 56 38.7 & $-$67 53 39 & -- & 27,31 \\ \hline \end{tabular} \end{flushleft} \normalsize \end{table*} 57 obscured AGB stars and a few red supergiants (RSGs) in the LMC were selected for Guaranteed Time and follow-up Open Time observations with the Infrared Space Observatory (ISO; Kessler et al.\ 1996). The goals were to obtain photometry at 12, 25 and 60 $\mu$m and to spectroscopically determine the chemical types of the CSEs. The photometry, which covers the entire spectral energy distributions (SEDs), can be modelled and used to derive accurate luminosities and mass-loss rates. In this paper we present the ISO data and classify sources as oxygen- or carbon-rich. \section{Source selection} \begin{table*} \caption[]{The list of program stars without IRAS counterpart. The references are as in Table 1. SHV is from Hughes (1989), BMB is from Blanco et al.\ (1980), WBP is from Wood et al.\ (1985) and GRV is from Reid et al.\ (1988).} \begin{flushleft} \begin{tabular}{llllllll} \hline\hline LI & IRAS & TRM & HV & RA (2000) & Decl (2000) & Other names & References \\ \hline \multicolumn{8}{c}{{\em non-IRAS sources classified as C stars}} \\ \hline -- & -- & -- & -- & 04 53 59.7 & $-$67 45 47 & {\bf SHV0454030$-$675031} & 20,21 \\ -- & -- & -- & -- & 05 02 28.7 & $-$69 20 10 & {\bf SHV0502469$-$692418} & 20,21,23 \\ -- & -- & -- & {\bf 2379} & 05 14 46.3 & $-$67 55 47 & -- & 3,10,11,20 \\ -- & -- & -- & -- & 05 20 46.8 & $-$69 01 25 & {\bf SHV0521050$-$690415}, BCB-R046 & 8,20,21,23 \\ -- & -- & -- & -- & 05 25 30.6 & $-$70 09 13 & {\bf SHV0526001$-$701142} & 20,21 \\ -- & -- & -- & -- & 05 26 17.4 & $-$69 08 07 & {\bf WBP14} & 15 \\ -- & -- & -- & -- & 05 35 11.4 & $-$70 22 46 & {\bf SHV0535442$-$702433} & 20,21 \\ \hline \multicolumn{8}{c}{{\em non-IRAS sources classified as M or S stars}} \\ \hline -- & -- & -- & {\bf 2446} & 05 20 01.5 & $-$67 34 43 & WOH G274, GRV0520$-$6737 & 11,18 \\ -- & -- & -- & -- & 05 21 33.1 & $-$70 09 56 & {\bf SHV0522023$-$701242} & 20,21 \\ -- & -- & -- & -- & 05 21 40.5 & $-$70 22 31 & {\bf SHV0522118$-$702517} & 20,21 \\ -- & -- & -- & -- & 05 24 31.3 & $-$69 43 25 & {\bf SHV0524565$-$694559} & 20,21 \\ -- & -- & -- & -- & 05 30 00.3 & $-$70 20 06 & {\bf SHV0530323$-$702216} & 20,21,23 \\ -- & -- & -- & {\bf 12070} & 05 52 27.8 & $-$69 14 12 & WOH SG515 & 9,11 \\ \hline \multicolumn{8}{c}{{\em non-IRAS sources without spectral classification}} \\ \hline -- & -- & -- & -- & 05 00 11.2 & $-$68 12 48 & {\bf SHV0500193$-$681706} & 20,21 \\ -- & -- & -- & -- & 05 00 13.5 & $-$68 24 56 & {\bf SHV0500233$-$682914} & 20,21 \\ -- & -- & -- & -- & 05 19 41.8 & $-$66 57 50 & {\bf GRV0519$-$6700} & 18 \\ \hline \end{tabular} \end{flushleft} \normalsize \end{table*} The sources observed with ISO were selected from the lists presented in Paper I, where all IRAS candidate AGB stars in the MCs are listed. We selected 30 infrared AGB stars or RSGs without optical counterparts from their Table 2. These objects should represent the brightest, most obscured AGB stars. Four objects from this table were excluded because of their red IRAS colours ($S_{25}/S_{12}\gsim2.5$): LI-LMC528, 861, 1137 and 1341. We also selected 8 sources from the optically known M and C stars with IRAS counterparts in Table 1 of Paper I. These include well known Harvard Variables as well as the optically thick source IRAS04553$-$6825 (LI-LMC181, WOH G64; Elias et al.\ 1986; Wood et al.\ 1986). Two unidentified IRAS sources from Table 4 of Paper I have been included in the present sample. LI-LMC203 is near an M1.5 star (HV12501), but there is also an A3 Iab supergiant (Sk$-69$-39a) close to the IRAS position. For LI-LMC1795 we found a bright R-band counterpart (Paper II). Finally one source from Table 7 of Paper I was included (LI-LMC1130). Although listed in Paper I as a foreground star, it was included here in an attempt to establish whether this is true. For these last three stars the higher spatial resolution ISO observations at 12 $\mu$m allow a better identification of the source with one of the possible counterparts found near the IRAS position. The 41 IRAS sources included in this study are listed in Table 1, with the most common names for these objects, their coordinates (J2000) and some references. The coordinates for the pointings of the ISO observations were taken from the SIMBAD astronomical database in 1994. The selection of IRAS detected AGB stars gives a sample that is severely biased towards very luminous stars (including supergiants). We therefore also included 16 non-IRAS stars. These were mostly taken from Wood et al.\ (1983, 1985), Reid et al.\ (1990) and Hughes (1989). Seven of these objects are classified as C stars from optical spectra or near-IR colours. Six objects are classified as M or S stars and for three objects no classification is available. This group of non-IRAS sources also includes the RCB-like variable HV2379 (Bessell \& Wood 1983). These sources are listed in Table 2. \section{IRAS data} \begin{table} \caption[]{Revised IRAS 12, 25, 60 and 100 $\mu$m photometry (in Jy), accompanied by a colon if questionable.} \begin{tabular}{lllll} \hline\hline Star & $F_{12}$ & $F_{25}$ & $F_{60}$ & $F_{100}$ \\ \hline GRV0519$-$6700 & \llap{$<$}0.06 & \llap{$<$}0.02 & \llap{$<$}0.1 & \\ HV12070 & 0.06 & 0.03 & 0.1\rlap{:} & \\ HV12501 & 0.23 & 0.06 & \llap{$<$}1.0 & \\ HV2360 & 0.38 & 0.35 & 0.4\rlap{:} & \\ HV2379 & 0.05 & 0.02 & \llap{$<$}0.8 & \\ HV2446 & 0.05 & 0.02 & \llap{$<$}0.2 & \\ HV5870 & 0.30 & 0.17 & \llap{$<$}5.0 & \\ HV888 & 0.58 & 0.29 & \llap{$<$}4.0 & \\ HV916 & 0.44 & 0.23 & \llap{$<$}2.0 & \\ HV996 & 0.71 & 0.53 & \llap{$<$}0.5 & \\ IRAS04286$-$6937 & 0.28 & 0.20 & \llap{$<$}0.1 & \\ IRAS04374$-$6831 & 0.24 & 0.12 & 0.1\rlap{:} & \\ IRAS04407$-$7000 & 0.76 & 0.76 & 0.1 & \\ IRAS04496$-$6958 & 0.31 & 0.19 & 0.1 & \\ IRAS04498$-$6842 & 1.33 & 0.89 & \llap{$<$}0.2 & \\ IRAS04509$-$6922 & 0.89 & 0.86 & \llap{$<$}2.0 & \\ IRAS04516$-$6902 & 0.86 & 0.55 & 0.4\rlap{:} & \\ IRAS04530$-$6916 & 2.07 & 5.09 & \llap{2}2.0 & \llap{2}8.0 \\ IRAS04539$-$6821 & 0.22 & 0.12 & 0.1\rlap{:} & \\ IRAS04545$-$7000 & 0.46 & 0.83 & \llap{$<$}0.5 & \\ IRAS04557$-$6753 & 0.24 & 0.14 & \llap{$<$}0.3 & \\ IRAS05003$-$6712 & 0.43 & 0.33 & 0.1\rlap{:} & \\ IRAS05009$-$6616 & 0.28 & 0.14 & \llap{$<$}0.4 & \\ IRAS05112$-$6755 & 0.46 & 0.33 & 0.7\rlap{:} & \\ IRAS05113$-$6739 & 0.25 & 0.14 & 0.1\rlap{:} & \\ IRAS05128$-$6455 & 0.23 & 0.24 & 0.1 & \\ IRAS05190$-$6748 & 0.39 & 0.25 & 0.1\rlap{:} & \\ IRAS05289$-$6617 & 0.16 & 0.39 & 0.3 & \\ IRAS05294$-$7104 & 0.69 & 0.56 & \llap{$<$}3.0 & \\ IRAS05295$-$7121 & 0.23 & 0.08 & \llap{$<$}0.3 & \\ IRAS05298$-$6957 & 0.85 & 1.38 & \llap{$<$}3.0 & \\ IRAS05300$-$6651 & 0.28 & 0.17 & 0.1\rlap{:} & \\ IRAS05329$-$6708 & 0.74 & 1.23 & 0.2\rlap{:} & \\ IRAS05348$-$7024 & 0.58 & 0.16 & \llap{$<$}1.0 & \\ IRAS05360$-$6648 & 0.21 & 0.09 & 0.3\rlap{:} & \\ IRAS05402$-$6956 & 0.71 & 1.02 & \llap{$<$}2.0 & \\ IRAS05506$-$7053 & 0.28 & 0.16 & \llap{$<$}0.2 & \\ IRAS05558$-$7000 & 0.85 & 0.80 & 0.2\rlap{:} & \\ IRAS05568$-$6753 & 0.35 & 0.43 & 0.2 & \\ SHV0454030$-$675031 & \llap{$<$}0.03 & \llap{$<$}0.03 & \llap{$<$}0.2 & \\ SHV0500193$-$681706 & 0.11\rlap{:} & 0.07 & \llap{$<$}0.3 & \\ SHV0500233$-$682914 & 0.10\rlap{:} & 0.03 & \llap{$<$}1.5 & \\ SHV0502469$-$692418 & 0.02\rlap{:} & \llap{$<$}0.03 & \llap{$<$}0.1 & \\ SHV0521050$-$690415 & 0.06\rlap{:} & 0.02\rlap{:} & \llap{$<$}0.7 & \\ SHV0522023$-$701242 & \llap{$<$}0.10 & \llap{$<$}0.04 & 0.4\rlap{:} & \\ SHV0522118$-$702517 & 0.06\rlap{:} & 0.05 & \llap{$<$}2.2 & \\ SHV0524565$-$694559 & \llap{$<$}0.14 & \llap{$<$}0.07 & \llap{$<$}1.0 & \\ SHV0526001$-$701142 & 0.07\rlap{:} & 0.01\rlap{:} & 0.1\rlap{:} & \\ SHV0530323$-$702216 & \llap{$<$}0.04 & \llap{$<$}0.04 & 0.4 & \\ SHV0535442$-$702433 & 0.01\rlap{:} & 0.07\rlap{:} & \llap{$<$}1.0 & \\ SP77 30$-$6 & 0.26 & 0.13 & 0.1\rlap{:} & \\ TRM45 & 0.07 & 0.07 & \llap{$<$}2.0 & \\ TRM72 & 0.22 & 0.06 & \llap{$<$}0.3 & \\ TRM88 & 0.17 & 0.04 & \llap{$<$}0.7 & \\ WBP14 & 0.01\rlap{:} & \llap{$<$}0.03 & \llap{$<$}4.0 & \\ WOH G64 & 8.45 & \llap{1}3.53 & 2.2 & \\ WOH SG374 & 0.37 & 0.38 & 0.2\rlap{:} & \\ \hline \end{tabular} \end{table} Here the IRAS data are discussed, for later comparison with the ISO photometry. Data at 12, 25, 60 and 100 $\mu$m was retrieved from the IRAS data base server in Groningen\footnote{The IRAS data base server of the Space Research Organisation of the Netherlands (SRON) and the Dutch Expertise Centre for Astronomical Data Processing is funded by the Netherlands Organisation for Scientific Research (NWO). The IRAS data base server project was also partly funded through the Air Force Office of Scientific Research, grants AFOSR 86-0140 and AFOSR 89-0320.} (Assendorp et al.\ 1995). The Groningen Gipsy data analysis software was used to measure the flux density from a trace through the position of the star (Gipsy command SCANAID). For the 60 and 100 $\mu$m data, $2\times2$ square degree maps were created with $0.5^\prime$ pixels to find point sources coincident with the positions of the stars. The 12 and 25 $\mu$m flux densities have a 1-$\sigma$ error of a few per cent, with a minimum error of $\sim0.01$ Jy. The 60 and 100 $\mu$m flux densities are much less certain, and it is also more difficult to assess reliable error estimates: 10\% would be a typical error. The faintest 60 $\mu$m sources that IRAS detected were assigned $F_{60}=0.1$ Jy. Only one source was well detected at 100 $\mu$m. The flux densities are listed in Table 3. When it is not certain that the measured flux density is physically related to the star of interest it is marked with a colon. All of our sources that are in the IRAS-PSC, plus HV5870 (=LI-LMC1145) and TRM72 (=LI-LMC578) that are in Schwering \& Israel (1990), were recovered with good flux density determinations at 12 $\mu$m. Reliable 12 $\mu$m flux densities could also be determined for IRAS05128$-$6455 and 05289$-$6617, below their upper limits as listed in the PSC. Neither in the PSC, nor in Schwering \& Israel (1990), are HV12070, HV2379, HV2446, TRM45, and TRM88 secure detections. Detection is not certain for WBP14 and the SHV sources for which flux density estimates are listed. The 12 $\mu$m flux densities of the GRV source and four SHV sources are upper limits. IRAS05506$-$7053 looks extended or multiple. At 25 $\mu$m detections seem a little more reliable than at 12 $\mu$m, at a given flux density. Rather surprisingly, the detection limit at 25 $\mu$m is at least as faint as at 12 $\mu$m; sources with $F_{25}\sim0.02$ Jy could be found (see also Reid et al.\ 1990). This is, however, only possible because the positions of the stars are known. For SP77 30$-$6 and all eight (other) IRAS sources the PSC lists only upper limits of $F_{25}<0.25$ Jy. SHV0502469$-$692418 and WBP14 were the only sources that were (tentatively) detected at 12 $\mu$m but not at 25 $\mu$m. Their flux densities are probably below the limit of detection, $F_{25}<0.01$ Jy, if their colours are rather blue. At 60 $\mu$m, IRAS05298$-$6957 is a bright, small but extended source about $10^\prime$ in diameter, with $F_{60}\sim2$ Jy. No point source could be distinguished on top of this emission, that is probably associated with the small cluster of which IRAS05298$-$6957 is a member (Paper IV). Flux densities are listed for two dozen sources, but it is not sure how many among these are real detections and how many are spurious. The only 60 $\mu$m detections in the PSC are IRAS04516$-$6902 ($0.80\pm0.19$ Jy), 04530$-$6916 ($20.51\pm1.85$ Jy) and 05112$-$6755 ($0.91\pm0.11$ Jy), all consistent with our estimates. More stringent upper limits are put on the 60 $\mu$m flux densities of the other sources. At 100 $\mu$m sources may be detected as faint as a few Jy. The only detection, however, is the brightest far-IR source in our sample, IRAS04530$-$6916, which we measured at $F_{100}=28$ Jy. This is consistent with the PSC upper limit of 46.17 Jy. The new flux density estimates can be compared with the literature values from the PSC or Schwering \& Israel (1990) (Fig.\ 1). On average, the new flux densities are only a few per cent fainter than the values from the literature. Flux densities $F_{25}\lsim0.2$ Jy may have been over-estimated in the past. HV12501 with $F_{\rm rev}$/$F_{\rm lit}=0.56$ and 0.32 at 12 and 25 $\mu$m, respectively, and IRAS05506$-$7053 with $F_{\rm rev}$/$F_{\rm lit}=0.67$ and 0.42 at 12 and 25 $\mu$m, respectively, are the most extreme examples of this. Schwering \& Israel (1990) over-estimated the 25 $\mu$m flux density of TRM72 ($F_{\rm rev}$/$F_{\rm lit}=0.55$), but under-estimated the 12 $\mu$m flux density of HV5870 ($F_{\rm rev}$/$F_{\rm lit}=2.00$). The other two flux densities which are obviously under-estimated are for SP77 30$-$6 at 12 $\mu$m ($F_{\rm rev}$/$F_{\rm lit}=1.53$) and IRAS04286$-$6937 at 25 $\mu$m ($F_{\rm rev}$/$F_{\rm lit}=1.67$). \begin{figure}[tb] \centerline{\psfig{figure=h1374.f1,width=88mm,height=45mm}} \caption[]{A comparison of the estimates of IRAS flux densities given here and values from the IRAS Point Source Catalogue and Schwering \& Israel (1990), at 12 (solid symbols) and 25 $\mu$m (open symbols).} \end{figure} IRAS counterparts not listed in IRAS-based catalogues may still be found in the original IRAS data. This is because manual extraction and measuring of the data is a more sophisticated technique than the automatic techniques that created the existing catalogues. In particular, manual flux-density determination enables the background flux levels to be estimated and subtracted better, yielding more reliable photometry. The only sources from our ISO sample that could not be detected in the IRAS data at either 12, 25 or 60 $\mu$m are GRV0519$-$6700, SHV0454030$-$675031 and SHV0524565$-$694559. \section{ISO observations} The programme stars were observed with ISO at 12, 25 and 60 $\mu$m (chopped measurements) and with PHOT-S as part of a Guaranteed Time programme under proposals NTMCAGB1 and NTMCAGB2, and at 60 $\mu$m with mapping mode and CAM-CVF as part of an Open Time programme under proposal LMCSPECT. The 12 $\mu$m photometry was mostly obtained using the ISOCAM instrument (Cesarsky et al.\ 1996) in staring mode with a $3^{\prime\prime}$ pixel field of view in beam LW-s and using the LW10 filter ($\sim$ IRAS 12 $\mu$m). We did 25 exposures of each 2.1 s duration, after a number of read-outs to stabilise the detector (ranging from 10 to 34 frames depending on the expected source flux density). The gain was 2 in most cases, but 1 in the case of sources that were expected to be relatively bright: HV12501 and 996, IRAS04496$-$6958, 04545$-$7000, 05003$-$6712, 05112$-$6755, 05348$-$7024 and 05568$-$6753, and WOH SG374. For most stars this resulted in a clear detection with S/N ratios of 10 to 100. In total 44 sources were observed with ISOCAM at 12 $\mu$m. For sources that were expected to be stronger than 0.5 Jy and which would therefore saturate the ISOCAM detectors with the LW10 filter, the 12 $\mu$m photometry was obtained with the ISOPHOT instrument (Lemke et al.\ 1996) using the 11.5 filter ($\sim$ IRAS 12 $\mu$m). These observations were done using triangular chopping with a chopper throw of $90^{\prime\prime}$. The aperture used for the observations was $52^{\prime\prime}$ in diameter. Integration times were 32 s on-source (and the same off-source), except for IRAS05294$-$7104 that we integrated 64 s. A total of 13 sources were observed in this mode. For 53 sources we obtained PHOT-P photometry at 25 $\mu$m using the 25 filter ($\sim$ IRAS 25 $\mu$m), triangular chopping with a chopper throw of $90^{\prime\prime}$, and an aperture of $52^{\prime\prime}$. Integration times ranged from 32 to 256 s, depending on the expected flux density. In our Guaranteed Time programme we finally observed 40 objects with ISOPHOT at 60 $\mu$m using the PHOT-C100 camera and filter 60 ($\sim$ IRAS 60 $\mu$m) and triangular chopping with a $150^{\prime\prime}$ chopper throw. Integration times ranged from 32 to 128 s, depending on the expected flux density. Unfortunately due to the reduced in-orbit sensitivity of the instrument and the problems with the calibration of the chopped measurements (especially for PHOT-C), we discovered after most observations had already been carried out that this was not the best observing strategy for the 60 $\mu$m photometry. Therefore, 7 objects were observed again in the Open Time using PHOT-C100 and filter 60 in raster mapping mode, with $3\times3$ rasters and $45^{\prime\prime}$ raster steps in X and Y directions (spacecraft coordinates). The integration time per raster point was 128 seconds. In order to establish the carbon- or oxygen-rich nature of some of the programme stars we also obtained IR spectra for a number of them. In the Guaranteed Time 15 objects were observed using PHOT-S in staring mode, with integration times of 256 or 512 s (1024 s for HV2379) depending on the expected flux densities. The advantage of this instrument is that its spectral coverage is rather large (2 to 12 $\mu$m) at a reasonable resolution ($\sim90$). The sensitivity of the PHOT-S instrument, however, limits the detectability to sources with 12 $\mu$m flux densities above $\sim0.3$ Jy. Furthermore, using staring observations the background cannot easily be determined. In this spectral region the diffuse emission is dominated by the zodiacal emission, which, according to IRAS measurements, amounts to about $\sim0.1$ Jy in the PHOT-S aperture at 10 $\mu$m. Considering this, we decided to obtain CAM-CVF spectra for 12 objects with a pixel field-of-view of $6^{\prime\prime}$ in beam LW-l. We did 25 exposures of 2.1 s each at gain 2, after 50 read-outs to stabilise the detector. The unprecedented sensitivity of the ISOCAM instrument allows the observer to obtain spectra even for sources as faint as 100 mJy at 12 $\mu$m. Because of the long duration of a CVF observation, the spectral coverage chosen was only 7 to 9.2 $\mu$m (with step 4) in LW-CVF1 and 9 to 14.1 $\mu$m (with step 2) in LW-CVF2, at a spectral resolution of $\sim40$. A big advantage of the CAM-CVF is that the spectra are obtained using an imaging technique. Therefore, a background spectrum was obtained simultaneously. These background spectra can be used to correct the PHOT-S spectra. We also obtained observations of 3 objects for which PHOT-S spectra had already been taken, in order to cross-check the results from the different instruments. \subsection{Near-IR photometry} Near-IR photometry was determined for each star at the time of the ISO observation, by interpolating near-IR lightcurves from our monitoring campaign at the South African Astronomical Observatory (SAAO) at Sutherland, South Africa (Whitelock et al., in preparation). Nearly always the lightcurve was sampled close in time to the ISO observation, but occasionally some extrapolation was necessary. The quoted uncertainties include an estimate, for each star, of the error introduced by the inter/extrapolation. For TRM45 and for the H-band magnitude of IRAS05360$-$6648 we have made use of the near-IR lightcurves and photometry presented by Wood (1998), after transformation to the SAAO system using Carter (1990). The near-IR photometry is listed in Table 4, along with the ISO photometry and the Julian Dates of the ISO spectroscopy. No near-IR counterparts could be identified with IRAS05568$-$6753 and 05289$-$6617. Two stars with near-IR colours much like those of unobscured M-type stars were monitored in the near-IR, but they show no variability. \section{ISO results and comparison with IRAS photometry} The data were reduced using the PHOT and CAM Interactive Analysis software packages: PIA (Gabriel et al.\ 1997) version V7.1.2(e) and CIA (Ott et al.\ 1996) version V3.0, respectively. For a general description of the data and reduction methods see the ISOPHOT Data Users Manual (Laureijs et al.\ 1998), and the ISOCAM Observer's Manual (1994) and ISOCAM Data Users Manual (Siebenmorgen et al.\ 1998). Details of the steps undertaken in reducing so-called Edited Raw Data (ERD) products to the finally derived flux densities and spectra can be found in Appendices A (photometry) and B (spectroscopy). The resulting ISO photometry is listed in Table 4, and the ISO spectra are presented in Figs.\ 4 \& 5. The flux densities at 12 and 25 $\mu$m for the stars that were detected both by IRAS and ISO (Tables 3 \& 4) are compared in Fig.\ 2. A bright regime where ISO and IRAS are consistent can be distinguished from a faint regime where ISO flux densities are systematically lower than IRAS flux densities. CAM is consistent with IRAS down to fainter levels ($\sim0.2$ Jy) than PHOT ($\sim0.6$ Jy). PHOT seems to under-estimate flux densities at levels between 0.2 and 0.6 Jy by a factor $\sim$two. Below 0.2 Jy, both CAM and PHOT yield flux densities $\sim0.6\times$IRAS. \begin{figure}[tb] \centerline{\psfig{figure=h1374.f2,width=88mm,height=45mm}} \caption[]{Comparison between ISO and our revised IRAS flux densities.} \end{figure} Flux density under-estimation may be caused by the difficulty of the detectors to respond to low signals. CAM 12 $\mu$ flux densities below $\sim0.2$ Jy may have been under-estimated as the $\kappa$-$\sigma$ method does not adequately correct for non-stabilised signals if stabilisation is not reached well within less than half the integration time. The 12 and 25 $\mu$m PHOT measurements, although employing different detector materials, show exactly the same trend. The 12 (either CAM or PHOT) and 25 $\mu$m observations were performed in the same orbit. Default responsivities for the 12 and 25 $\mu$m PHOT measurements were thought to be (much) lower in 1996 than in the current calibrations. Adopting those early values, the ISOPHOT photometry would be consistent with the IRAS photometry to a high degree. Subsequent revisions of the default responsivities have lead to higher values, approaching the values we obtain from the (chopped) FCS measurements. The ratio of ISO and IRAS flux densities at 25 $\mu$m is $0.60\pm0.06$ for the 24 sources with ISO measurements before orbit 190, and $0.68\pm0.09$ for the 21 sources measured after orbit 190. These ratios are very similar, despite the large differences in median IRAS flux density between these two samples: 0.41 and 0.14 Jy, respectively. However, the discrepancy between the ISO and IRAS data may not be as great as it appears if we take plausible selection effects into account. The stars in our sample were largely selected on the basis of their IRAS flux density, but many of them were only just detected by IRAS. It is therefore likely that they were near the maximum of their pulsation cycles at the time of the IRAS observation. In contrast, they will have been at random phases when the ISO observations were made. This will lead to a systematic difference between the IRAS and ISO flux densities for faint sources. A similar effect may explain the discrepancy between the PHOT and CAM behaviour for sources with flux densities in the range 0.2 to 0.6 Jy, as the brighter sources were selected for measurement with PHOT and the fainter sources with CAM. Ground-based 10 $\mu$m (N-band) magnitudes of a subset of our ISO targets were on average $\sim30$\% fainter than measured by IRAS at 12 $\mu$m (Paper IV). Although we explained this in terms of differences between the N-band and IRAS 12 $\mu$m filters, it may actually reflect the same discrepancy seen between the ISO and IRAS photometry. Variability cannot be the complete explanation, though: for instance, the sources IRAS04407$-$7000, 4516$-$6902 and 05003$-$6712 were all near the maxima in their K- and L-band lightcurves at the time of the ISO photometry, yet their PHOT 25 $\mu$m flux densities of 0.584, 0.380, and 0.210 Jy, respectively, are still fainter than the IRAS flux densities by factors of 0.77, 0.69, and 0.64, respectively. Interestingly, Reid et al.\ (1990, their Figure 5) show that for $F_{12,25}\lsim0.3$ Jy both the PSC and Schwering \& Israel (1990) over-estimate flux densities for point sources in the LMC by typically 20 to 50\%. They attribute this to source confusion resulting from the large beam width and crowdedness in typical LMC fields. \begin{landscape} \begin{table} \caption[]{ISO 12, 25 and 60 $\mu$m photometry (in Jy). The near-IR magnitudes for the ISO-epochs is deduced from light curves obtained at SAAO ($JD-2,450,000={\rm orbit}+38$). Values in parentheses are 1-$\sigma$ errors. The last column indicates when a PHOT-S or CAM-CVF spectrum was taken.} \begin{tabular}{lllllllllllll} \hline\hline Star & $JD$ & $J [mag]$ & $H [mag]$ & $K [mag]$ & $L [mag]$ & $F_{12}$(CAM) & $F_{12}$(PHOT) & $F_{25}$(PHOT) & $F_{60}$(chop) & $F_{60}$(map) & Spectrum \\ \hline GRV0519$-$6700 & 318 & \llap{1}2.63(0.03) & \llap{1}1.36(0.02) & \llap{1}0.67(0.02) & & 0.004(0.001) & & & & & \\ HV12070 & 274 & \llap{1}0.03(0.02) & 8.98(0.02) & 8.68(0.02) & 8.24(0.03) & 0.043(0.004) & & \llap{$-$}0.023(0.003) & & & \\ & 787 & \llap{1}0.30(0.10) & 9.20(0.10) & 8.80(0.10) & 8.30(0.10) & & & & & & CAM-CVF \\ HV12501 & 754 & 8.80(0.05) & 8.00(0.05) & 7.75(0.05) & 7.30(0.05) & 0.185(0.005) & & 0.067(0.018) & 0.260(0.108) & & \\ HV2360 & 754 & 9.00(0.05) & 8.15(0.05) & 7.75(0.05) & 7.20(0.05) & 0.331(0.003) & & 0.138(0.019) & 0.207(0.042) & & \\ HV2379 & 288 & \llap{1}2.68(0.03) & \llap{1}2.05(0.04) & \llap{1}1.40(0.04) & 9.70(0.10) & 0.038(0.002) & & & \llap{$-$}0.027(0.048) & & PHOT-S \\ & 626 & & \llap{1}3.20(0.40) & \llap{1}1.90(0.20) & & 0.035(0.001) & & & & & \\ & 729 & & \llap{1}3.20(0.40) & \llap{1}1.90(0.20) & & 0.032(0.001) & & 0.014(0.005) & & & CAM-CVF \\ HV2446 & 288 & \llap{1}0.25(0.02) & 9.23(0.02) & 8.80(0.02) & 8.35(0.04) & 0.066(0.001) & & 0.043(0.005) & 0.109(0.056) & & \\ & 622 & \llap{1}0.60(0.10) & 9.60(0.10) & 9.10(0.10) & 8.35(0.05) & & & & & 0.361(0.135) & \\ & 754 & \llap{1}0.10(0.02) & 9.04(0.02) & 8.67(0.02) & 8.06(0.02) & & & & & & CAM-CVF \\ HV5870 & 754 & 9.34(0.02) & 8.40(0.02) & 8.10(0.02) & 7.60(0.04) & 0.269(0.002) & & 0.092(0.008) & 0.090(0.355) & & \\ HV888 & 195 & 8.10(0.01) & 7.20(0.01) & 6.89(0.01) & 6.46(0.02) & & 0.713(0.037) & 0.201(0.012) & \llap{$-$}0.237(0.626) & & PHOT-S \\ HV916 & 209 & 8.55(0.03) & 7.60(0.03) & 7.25(0.02) & 6.80(0.02) & 0.380(0.004) & & 0.176(0.015) & 0.065(0.171) & & \\ HV996 & 163 & 8.93(0.01) & 8.01(0.01) & 7.58(0.01) & 6.83(0.01) & 0.601(0.006) & & 0.364(0.026) & 0.326(0.074) & & PHOT-S \\ IRAS04286$-$6937 & 217 & & \llap{1}3.00(0.05) & \llap{1}1.25(0.05) & 9.10(0.05) & 0.136(0.002) & & 0.059(0.013) & \llap{$-$}0.101(0.130) & & \\ IRAS04374$-$6831 & 195 & & \llap{1}3.60(0.05) & \llap{1}1.40(0.04) & 8.80(0.03) & 0.185(0.003) & & 0.086(0.027) & 0.228(0.136) & & PHOT-S \\ & 701 & & \llap{1}3.60(0.10) & \llap{1}1.40(0.10) & 8.90(0.10) & & & & & 0.309(0.069) & \\ IRAS04407$-$7000 & 217 & \llap{1}0.18(0.03) & 8.92(0.03) & 8.18(0.03) & 7.30(0.05) & & 0.962(0.024) & 0.584(0.031) & 0.125(0.088) & & \\ & 605 & \llap{1}1.70(0.05) & \llap{1}0.20(0.05) & 9.20(0.05) & 8.10(0.05) & & & & & 0.473(0.132) & \\ IRAS04496$-$6958 & 195 & \llap{1}3.00(0.05) & \llap{1}0.90(0.05) & 9.50(0.04) & 7.70(0.04) & 0.269(0.002) & & 0.126(0.010) & 0.252(0.154) & & PHOT-S \\ & 605 & \llap{1}2.40(0.05) & \llap{1}0.40(0.05) & 8.95(0.05) & 7.60(0.05) & & & & & & CAM-CVF \\ IRAS04498$-$6842 & 217 & \llap{1}0.95(0.05) & 9.65(0.05) & 8.70(0.02) & 7.70(0.02) & & 0.486(0.013) & 0.221(0.028) & 0.292(0.112) & & \\ IRAS04509$-$6922 & 217 & \llap{1}1.50(0.05) & \llap{1}0.00(0.04) & 9.15(0.04) & 8.10(0.04) & & 0.303(0.011) & 0.202(0.028) & \llap{$-$}0.021(0.231) & & \\ IRAS04516$-$6902 & 202 & \llap{1}0.50(0.03) & 8.96(0.03) & 8.24(0.03) & 7.25(0.03) & & 0.854(0.026) & 0.380(0.051) & 0.349(0.135) & & \\ IRAS04530$-$6916 & 202 & \llap{1}3.65(0.05) & \llap{1}1.60(0.04) & 9.73(0.02) & 7.60(0.03) & & 2.144(0.079) & 4.105(0.054) & \llap{3}7.656(2.639) & & \\ IRAS04539$-$6821 & 229 & & \llap{1}4.30(0.05) & \llap{1}1.80(0.04) & 8.80(0.10) & 0.244(0.002) & & 0.089(0.013) & 0.106(0.053) & & \\ IRAS04545$-$7000 & 195 & & \llap{1}1.75(0.03) & 9.40(0.02) & 7.15(0.03) & 0.836(0.008) & & 0.927(0.027) & 0.641(0.192) & & PHOT-S \\ & 605 & & \llap{1}2.80(0.10) & \llap{1}0.30(0.10) & 8.10(0.10) & & & & & 0.286(0.206) & \\ IRAS04557$-$6753 & 229 & & \llap{1}4.60(0.05) & \llap{1}2.45(0.04) & 9.60(0.05) & 0.164(0.002) & & 0.090(0.027) & \llap{$-$}0.072(0.137) & & \\ IRAS05003$-$6712 & 229 & \llap{1}2.30(0.05) & \llap{1}0.65(0.05) & 9.45(0.05) & 8.15(0.05) & 0.362(0.003) & & 0.210(0.007) & 0.173(0.079) & & PHOT-S \\ IRAS05009$-$6616 & 229 & \llap{1}4.70(0.10) & \llap{1}2.65(0.05) & \llap{1}0.70(0.05) & 8.45(0.05) & 0.284(0.003) & & 0.082(0.017) & \llap{$-$}0.032(0.047) & & \\ IRAS05112$-$6755 & 288 & & \llap{1}4.70(0.10) & \llap{1}2.00(0.05) & 8.80(0.05) & 0.414(0.004) & & & 0.078(0.133) & & PHOT-S \\ & 629 & & \llap{1}4.70(0.10) & \llap{1}1.90(0.05) & 8.65(0.05) & 0.360(0.003) & & 0.108(0.011) & & & PHOT-S \\ IRAS05113$-$6739 & 288 & & \llap{1}4.35(0.05) & \llap{1}2.12(0.03) & 9.06(0.02) & 0.316(0.003) & & & 0.398(0.115) & & \\ & 629 & & \llap{1}3.90(0.10) & \llap{1}1.60(0.10) & 8.75(0.05) & 0.209(0.002) & & & & & \\ & 729 & & \llap{1}3.90(0.05) & \llap{1}1.60(0.05) & 8.75(0.05) & 0.254(0.002) & & 0.067(0.013) & & & \\ IRAS05128$-$6455 & 628 & \llap{1}3.60(0.10) & \llap{1}2.10(0.10) & \llap{1}0.55(0.05) & 8.55(0.05) & 0.226(0.002) & & 0.061(0.008) & & & PHOT-S \\ IRAS05190$-$6748 & 288 & & & \llap{1}3.10(0.05) & 9.70(0.10) & 0.346(0.003) & & 0.163(0.022) & 0.190(0.060) & & PHOT-S \\ IRAS05289$-$6617 & 209 & & & & & 0.157(0.002) & & 0.202(0.009) & 0.382(0.049) & & \\ & 668 & & & & & & & & & 0.281(0.069) & \\ & 729 & & & & & & & & & & CAM-CVF \\ \hline \end{tabular} \end{table} \end{landscape} \begin{landscape} \begin{table} \addtocounter{table}{-1} \caption[]{(continued) The near-IR photometry of Wood (1998) is used for TRM45 and in part for IRAS05360$-$6648 (H-band), after transformation to the SAAO system (Carter 1990).} \begin{tabular}{llllllllllll} \hline\hline Star & $JD$ & $J [mag]$ & $H [mag]$ & $K [mag]$ & $L [mag]$ & $F_{12}$(CAM) & $F_{12}$(PHOT) & $F_{25}$(PHOT) & $F_{60}$(chop) & $F_{60}$(map) & Spectrum \\ \hline IRAS05294$-$7104 & 754 & \llap{1}1.80(0.05) & \llap{1}0.00(0.05) & 8.90(0.05) & 7.60(0.05) & & 0.680(0.045) & 0.394(0.035) & 0.079(0.097) & & \\ IRAS05295$-$7121 & 209 & & \llap{1}3.85(0.05) & \llap{1}1.75(0.05) & 9.35(0.05) & 0.143(0.001) & & 0.007(0.011) & 0.233(0.074) & & \\ IRAS05298$-$6957 & 209 & & & \llap{1}1.60(0.20) & 8.60(0.20) & & 0.303(0.015) & 0.359(0.014) & 0.414(0.279) & & PHOT-S \\ & 729 & \llap{1}4.10(0.20) & \llap{1}2.50(0.10) & \llap{1}1.10(0.10) & 8.50(0.20) & & & & & & CAM-CVF \\ IRAS05300$-$6651 & 209 & & \llap{1}4.70(0.20) & \llap{1}2.20(0.10) & 9.10(0.10) & 0.246(0.003) & & 0.114(0.017) & 0.068(0.049) & & \\ IRAS05329$-$6708 & 163 & \llap{1}7.00(0.20) & \llap{1}2.70(0.10) & \llap{1}0.40(0.05) & 8.25(0.05) & & 0.442(0.018) & 0.531(0.035) & 0.479(0.082) & & PHOT-S \\ IRAS05348$-$7024 & 202 & & \llap{1}3.60(0.30) & \llap{1}1.60(0.20) & 8.70(0.20) & 0.525(0.005) & & 0.208(0.031) & 0.170(0.108) & & \\ & 732 & & \llap{1}3.80(0.30) & \llap{1}2.70(0.20) & 9.30(0.20) & & & & & & CAM-CVF \\ IRAS05360$-$6648 & 202 & & \llap{1}5.00(0.40) & \llap{1}2.30(0.05) & 9.50(0.10) & 0.171(0.002) & & 0.082(0.013) & 0.174(0.067) & & \\ IRAS05402$-$6956 & 217 & & \llap{1}3.50(0.10) & \llap{1}0.60(0.05) & 8.00(0.04) & & 0.455(0.020) & 0.429(0.022) & 0.944(0.288) & & \\ & 662 & \llap{1}4.40(0.10) & \llap{1}1.45(0.05) & 9.40(0.03) & 7.15(0.03) & & & & & \llap{$-$}0.245(0.527) & \\ & 782 & \llap{1}4.50(0.10) & \llap{1}1.50(0.05) & 9.50(0.05) & 7.25(0.05) & & & & & & CAM-CVF \\ IRAS05506$-$7053 & 173 & & \llap{1}6.00(0.30) & \llap{1}3.30(0.10) & \llap{1}0.00(0.10) & & & & 0.109(0.062) & & \\ & 228 & & \llap{1}6.40(0.30) & \llap{1}3.60(0.10) & \llap{1}0.30(0.10) & & \llap{$-$}0.035(0.017) & 0.029(0.015) & 0.167(0.073) & & \\ IRAS05558$-$7000 & 209 & \llap{1}1.90(0.04) & \llap{1}0.10(0.04) & 8.90(0.03) & 7.60(0.04) & & 0.667(0.011) & 0.517(0.016) & 0.291(0.100) & & PHOT-S \\ & 781 & \llap{1}3.50(0.05) & \llap{1}1.60(0.03) & \llap{1}0.03(0.01) & 8.20(0.05) & & & & & & CAM-CVF \\ IRAS05568$-$6753 & 186 & & & & & 0.412(0.004) & & 0.285(0.008) & 0.383(0.073) & & PHOT-S \\ SHV0454030$-$675031 & 318 & \llap{1}4.20(0.10) & \llap{1}2.67(0.04) & \llap{1}1.85(0.03) & & 0.006(0.002) & & & & & \\ SHV0500193$-$681706 & 229 & \llap{1}4.00(0.20) & \llap{1}2.00(0.10) & \llap{1}0.70(0.10) & 9.50(0.10) & 0.030(0.005) & & 0.030(0.016) & 0.098(0.026) & & \\ & 754 & \llap{1}3.10(0.10) & \llap{1}1.45(0.05) & \llap{1}0.25(0.05) & 9.00(0.05) & & & & & & CAM-CVF \\ SHV0500233$-$682914 & 229 & \llap{1}3.05(0.05) & \llap{1}1.30(0.05) & \llap{1}0.00(0.05) & 8.50(0.05) & 0.076(0.001) & & 0.004(0.033) & \llap{$-$}0.004(0.097) & & \\ & 729 & \llap{1}3.30(0.10) & \llap{1}1.40(0.10) & \llap{1}0.20(0.05) & 8.60(0.05) & & & & & & CAM-CVF \\ SHV0502469$-$692418 & 304 & \llap{1}2.93(0.03) & \llap{1}1.61(0.02) & \llap{1}0.81(0.02) & \llap{1}0.30(0.05) & 0.002(0.001) & & & & & \\ SHV0521050$-$690415 & 788 & \llap{1}1.07(0.02) & 9.77(0.02) & 9.23(0.02) & 8.40(0.10) & 0.062(0.004) & & \llap{$-$}0.011(0.014) & & & \\ SHV0522023$-$701242 & 788 & \llap{1}2.60(0.10) & \llap{1}1.60(0.10) & \llap{1}1.40(0.10) & & 0.001(0.001) & & 0.003(0.021) & & & \\ SHV0522118$-$702517 & 217 & \llap{1}3.00(0.20) & \llap{1}2.00(0.20) & \llap{1}1.00(0.20) & 9.70(0.20) & 0.028(0.001) & & 0.007(0.002) & 0.170(0.125) & & \\ SHV0524565$-$694559 & 217 & \llap{1}2.50(0.04) & \llap{1}1.40(0.05) & \llap{1}0.75(0.05) & & 0.003(0.001) & & & & & \\ SHV0526001$-$701142 & 217 & \llap{1}3.50(0.10) & \llap{1}2.00(0.10) & \llap{1}0.45(0.05) & 9.15(0.05) & 0.037(0.001) & & \llap{$-$}0.019(0.030) & 0.051(0.063) & & \\ SHV0530323$-$702216 & 788 & \llap{1}1.80(0.10) & \llap{1}0.80(0.10) & \llap{1}0.35(0.05) & & 0.008(0.002) & & 0.008(0.009) & & & \\ SHV0535442$-$702433 & 782 & \llap{1}2.90(0.20) & \llap{1}1.50(0.10) & \llap{1}0.65(0.05) & \llap{1}0.30(0.20) & 0.009(0.001) & & 0.029(0.022) & & & \\ SP77 30$-$6 & 195 & \llap{1}0.80(0.10) & 9.60(0.10) & 9.20(0.10) & 8.50(0.10) & 0.139(0.001) & & 0.077(0.018) & 0.225(0.073) & & \\ & 794 & \llap{1}0.30(0.10) & 9.30(0.10) & 8.80(0.10) & 8.20(0.10) & & & & & & CAM-CVF \\ TRM45 & 788 & \llap{1}6.20(0.20) & \llap{1}3.60(0.10) & \llap{1}1.55(0.05) & 9.50(0.10) & 0.076(0.001) & & 0.037(0.007) & & & \\ TRM72 & 788 & \llap{1}5.22(0.20) & \llap{1}2.81(0.03) & \llap{1}0.96(0.02) & 8.60(0.05) & 0.161(0.001) & & 0.026(0.025) & & & \\ TRM88 & 788 & \llap{1}3.50(0.10) & \llap{1}1.78(0.02) & \llap{1}0.24(0.02) & 8.55(0.05) & 0.101(0.001) & & 0.051(0.014) & & & \\ WBP14 & 209 & \llap{1}3.35(0.05) & \llap{1}1.60(0.05) & \llap{1}0.55(0.03) & 9.50(0.10) & 0.022(0.001) & & \llap{$-$}0.005(0.011) & 1.372(0.539) & & \\ & 616 & \llap{1}3.40(0.10) & \llap{1}1.60(0.10) & \llap{1}0.55(0.05) & 9.50(0.05) & & & & & 0.211(0.293) & \\ WOH G64 & 229 & 9.58(0.02) & 7.96(0.02) & 6.98(0.02) & 5.32(0.02) & & \llap{1}2.112(0.965) & \llap{1}3.786(0.354) & 4.647(0.298) & & PHOT-S \\ WOH SG374 & 788 & 9.90(0.02) & 9.10(0.02) & 8.64(0.02) & 7.69(0.05) & 0.490(0.005) & & 0.188(0.020) & & & \\ \hline \end{tabular} \end{table} \end{landscape} \begin{figure}[tb] \centerline{\psfig{figure=h1374.f3,width=88mm,height=40mm}} \caption[]{Histogram of the distribution of ISO 60 $\mu$m flux densities (chopped measurements). The dotted vertical line indicates 3-$\sigma$ flux density derived from the distribution of negative flux densities. All flux densities over 1 Jy are piled up in the last bin.} \end{figure} A histogram of the distribution of ISO 60 $\mu$m flux densities (Fig.\ 3), leaving out the mapping observations, illustrates the detection rate. Considering negative flux densities indicating non-detection, and assuming a Gaussian distribution around zero flux density for non-detections, we estimate a 1-$\sigma$ detection to have 0.103 Jy. There are 12 sources with flux densities exceeding 3-$\sigma$, i.e.\ probable detections. This does not take into the account the large errors on some of the individual measurements, and a 3-$\sigma$ detection may still turn out to be spurious (an example is WBP14). On the other hand, the distribution below 3-$\sigma$ is certainly skewed towards positive flux densities. Projecting the negative flux density distribution onto the positive domain, we estimate that there are probably 17 more detections between 0 and 3-$\sigma$, and a total of 14 non-detections. In the IRAS 60 $\mu$m data we found 8 detections and 17 tentative detections (Table 3). The 0.1 Jy assigned to the faintest IRAS 60 $\mu$m flux densities compares well with the ISO 1-$\sigma$ detection threshold of the chopped measurements. Of the 8 IRAS 60 $\mu$m detections 6 have ISO chopped measurements, all of which yield higher flux densities than IRAS --- by a factor 1.8 on average. This is in contrast to the 12 and 25 $\mu$m photometry, where ISO flux densities are generally lower than those measured by IRAS. The 14 ISO chopped measurements of IRAS tentative detections also yield higher flux densities than did IRAS --- by a factor of 1.5 on average, although some individual ISO measurements are fainter than the IRAS ones. None of these ISO measurements is negative, indicating that many of the IRAS 60 $\mu$m tentative detections are indeed real. The 7 mapping observations all agree with the chopped measurements within 2-$\sigma$, although these errors can be large. There is no tendency for one of these two methods to yield higher flux densities than the other. As we do not expect strong variability at 60 $\mu$m, which traces cool dust some distance from the stars, flux densities from mapping and chopped measurements are averaged. The error estimates of the mapping measurements are systematically larger than those of the chopped measurements. This may be due to the fact that, for the mapping data, the flux density of the star was determined from the inner $3\times3$ pixels. The contribution of the background to these 9 pixels is considerable. Also, the reliability of the error estimate for the central pixel in the chopped data as produced by PIA is unknown. There is great difficulty in extracting reliable photometry and associated errors from either mapping or chopped measurements at 60 $\mu$m, for stellar sources in fields like the LMC. This is mainly due to the complex background and limited spatial resolution of PHOT at these wavelengths. IRAS05289$-$6617 has a very smooth background, being situated in the line-of-sight to the supergiant shell LMC4 (Meaburn 1980). Indeed, ISO mapping and chopped measurements are relatively precise for this source, and agree nicely with the 60 $\mu$m flux density measured from the IRAS data. \begin{figure}[tb] \centerline{\psfig{figure=h1374.f4,width=88mm}} \caption[]{The CAM-CVF spectra of obscured AGB stars in the LMC. Open symbols represent spectro-photometric points that are prone to have flux densities that are over-estimated due to stabilisation difficulties. The spectral shape is best represented by the solid symbols (squares for the short-, disks for the long-wavelength region). Emission and/or absorption centred at $\sim9.7$ $\mu$m is indicative of oxygen-rich dust (e.g.\ IRAS05402$-$6956 and SP77 30$-$6), whereas carbon-rich dust may show emission at $\sim11.3$ $\mu$m (e.g.\ IRAS05289$-$6617). A featureless continuum around 10 $\mu$m also strongly suggests carbon-rich dust (e.g.\ SHV0500193$-$681706).} \end{figure} \begin{figure*}[tb] \centerline{\psfig{figure=h1374.f5a,width=180mm}} \caption[]{The PHOT-S spectra of obscured AGB stars (and RSGs) in the LMC. Emission and/or absorption centred at $\sim9.7$ $\mu$m suggests oxygen-rich dust (e.g.\ HV888 and IRAS04545$-$7000). Absorption at 3 $\mu$m is seen in carbon star photospheres (e.g.\ IRAS04496$-$6958), but artifacts in the PHOT-S responsivities also mimic weak depression at 3 $\mu$m in the spectra of unambiguous oxygen-rich stars (e.g.\ IRAS04545$-$7000).} \end{figure*} \begin{figure*}[tb] \addtocounter{figure}{-1} \centerline{\psfig{figure=h1374.f5b,width=180mm}} \caption[]{(continued)} \end{figure*} \section{Discussion} \subsection{Chemical types from ISO spectroscopy} The presence or absence in the ISO spectra (Figs.\ 4 \& 5) of discrete dust emission and molecular absorption bands can be used to distinguish between carbon- and oxygen-rich circumstellar envelopes (e.g.\ Merrill \& Stein 1976a,b,c). The results are summarised in Table 5. Amorphous oxygen-rich dust may give rise to strong and broad silicate emission between $\sim8$ and 13 $\mu$m, peaking at $\sim9.7 \mu$m (the exact location may differ from this by a few tenths of $\mu$m). The late-M stars HV2446, 888, 996, and SP77 30$-$6 have prominent silicate emission. In optically thick cases the silicate feature turns into absorption. All spectra of OH maser sources show the silicate feature in self-absorption: IRAS04545$-$7000, 05298$-$6957, 05329$-$6708, 05402$-$6956, and WOH G64. Oxygen-rich molecules do not provide clear diagnostics of the chemical type of CSEs at our signal-to-noise and spectral resolution. We already mentioned that shallow absorption around 3 $\mu$m in oxygen-rich sources is most likely due to an artifact in the responsivities, rather than H$_2$O ice. \begin{table} \caption[]{Chemical types. Optical spectra (Opt Sp) include objective prism and CCD spectroscopy up to $\sim1 \mu$m. ISO spectroscopy (ISO Sp) comprises PHOT-S and CAM-CVF observations. IR colour-colour diagrams (IR col) can in some cases be reasonably conclusive too: we here use $(K-[12])$ and $([12]-[25])$ versus $(K-L)$ diagrams. At radio wavelengths, OH, SiO and/or H$_2$O maser emission is detected from some oxygen-rich sources.} \begin{tabular}{lllll} \hline\hline Star & Opt Sp & ISO Sp & IR col & Maser \\ \hline GRV0519$-$6700 & C & & carbon? & \\ HV12070 & MS3/9 & oxygen? & ? & \\ HV12501 & M1.5 & & oxygen & \\ HV2360 & M2 Ia & & oxygen & \\ HV2379 & C & SiC? & carbon & \\ HV2446 & M5e & silicate & oxygen & \\ HV5870 & M4.5/5 & & oxygen & \\ HV888 & M4 Ia & silicate & oxygen & \\ HV916 & M3 Iab & & oxygen & \\ HV996 & M4 Iab & silicate & oxygen & \\ IRAS04286$-$6937 & & & carbon & \\ IRAS04374$-$6831 & & SiC? & carbon & \\ IRAS04407$-$7000 & & & oxygen & yes \\ IRAS04496$-$6958 & C & car$+$sil? & carbon & \\ IRAS04498$-$6842 & & & oxygen & \\ IRAS04509$-$6922 & M10 & & oxygen & \\ IRAS04516$-$6902 & M9 & & oxygen & \\ IRAS04530$-$6916 & & & oxygen & \\ IRAS04539$-$6821 & & & carbon & \\ IRAS04545$-$7000 & & silicate & oxygen & yes \\ IRAS04557$-$6753 & & & carbon & \\ IRAS05003$-$6712 & & silicate? & oxygen & \\ IRAS05009$-$6616 & & & carbon & \\ IRAS05112$-$6755 & & carbon? & carbon & \\ IRAS05113$-$6739 & & & carbon & \\ IRAS05128$-$6455 & & carbon? & carbon & \\ IRAS05190$-$6748 & & carbon? & carbon & \\ IRAS05289$-$6617 & & SiC & ? & \\ IRAS05294$-$7104 & & & oxygen & \\ IRAS05295$-$7121 & & & carbon & \\ IRAS05298$-$6957 & & silicate & oxygen & yes \\ IRAS05300$-$6651 & & & carbon & \\ IRAS05329$-$6708 & & silicate & oxygen & yes \\ IRAS05348$-$7024 & & SiC & carbon & \\ IRAS05360$-$6648 & & & carbon & \\ IRAS05402$-$6956 & & silicate & oxygen & yes \\ IRAS05506$-$7053 & & & oxygen & \\ IRAS05558$-$7000 & & silicate & oxygen & \\ IRAS05568$-$6753 & & carbon? & ? & \\ SHV0454030$-$675031 & C & & carbon & \\ SHV0500193$-$681706 & & carbon & carbon & \\ SHV0500233$-$682914 & & SiC? & carbon & \\ SHV0502469$-$692418 & C & & carbon? & \\ SHV0521050$-$690415 & C & & carbon & \\ SHV0522023$-$701242 & M3 & & ? & \\ SHV0522118$-$702517 & S? & & carbon & \\ SHV0524565$-$694559 & MS5 & & ? & \\ SHV0526001$-$701142 & C & & carbon & \\ SHV0530323$-$702216 & M6 & & oxygen & \\ SHV0535442$-$702433 & C & & ? & \\ SP77 30$-$6 & M8 & silicate & oxygen & \\ TRM45 & & & carbon & \\ \hline \end{tabular} \end{table} \begin{table} \addtocounter{table}{-1} \caption[]{(continued).} \begin{tabular}{lllll} \hline\hline Star & Opt Sp & ISO Sp & IR col & Maser \\ \hline TRM72 & C & & carbon & \\ TRM88 & C & & carbon & \\ WBP14 & C & & carbon & \\ WOH G64 & M7.5 & silicate & oxygen & yes \\ WOH SG374 & M6 & & oxygen & \\ \hline \end{tabular} \end{table} Crystalline carbon-rich dust sometimes gives rise to a SiC (graphite) emission feature peaking at $\sim11.3 \mu$m, and narrower than the silicate feature. The CVF spectrum of IRAS05289$-$6617 (Fig.\ 4) shows the best example of this. Carbon-rich molecules have several strong absorption bands in our spectral region, all from HCN and C$_2$H$_2$. The strongest is at 3.1 $\mu$m, but the problem with the responsivities limits the number of unambiguous detections to one (IRAS04496$-$6958). Related, but weaker, absorption is visible at 3.8 $\mu$m. More absorption bands are located around 5, 8 and 14 $\mu$m. Unfortunately, the 5 $\mu$m band falls entirely in the blind spectral region of PHOT-S. The 8 and 14 $\mu$m bands are at the edges of the CVF spectra and hence difficult to identify. Some other spectra show merely a featureless dust continuum around 10 $\mu$m. Best examples are the CVF spectrum of SHV0500193$-$681706 and the PHOT-S spectrum of IRAS05568$-$6753. These spectra suggest pure amorphous carbon dust emission. \subsection{IR colour-colour diagrams} The ISO 12, 25 and 60 $\mu$m filters are similar but not identical to the IRAS filters. As the zero-points of these ISO filters are unknown, we adopt here the IRAS zero-points. This results in the following definitions for the (not colour-corrected) mid-IR magnitudes \begin{equation} [12]=-2.5\log(F_{12}/28.3) \end{equation} \begin{equation} [25]=-2.5\log(F_{25}/6.73) \end{equation} \begin{equation} [60]=-2.5\log(F_{60}/1.19) \end{equation} \subsubsection{Diagram of $(K-[12])$ versus $(H-K)$} \begin{figure}[tb] \centerline{\psfig{figure=h1374.f6,width=88mm}} \caption[]{$(K-[12])$ versus $(H-K)$ diagram. Stars are distinguished by their chemical types inferred from spectroscopic methods: carbon stars (solid disks), M stars (open disks), MS stars (open squares), S stars (open triangles), and stars of which the chemical type is a priori unknown (crosses). Carbon stars and oxygen stars define sequences in this diagram, indicated by a dotted and solid curve, respectively.} \end{figure} The $(K-[12])$ versus $(H-K)$ colour-colour diagram separates carbon- from oxygen-rich stars in samples of obscured stars in the MCs (Papers II \& IV). Indeed, the distributions of carbon- and oxygen-rich stars using ISO and SAAO photometry define clear sequences in this diagram (Fig.\ 6). The sequences are fit by eye, with the carbon sequence the same as in Paper IV: \begin{equation} (H-K)=0.36\times(K-[12]) \end{equation} but the oxygen sequence a simple, yet somewhat steeper function than in Paper IV: \begin{equation} (H-K)=0.3+0.0003\times(K-[12])^5 \end{equation} Although the stars with spectral type M follow the oxygen sequence very well, the carbon stars show a large scatter around the carbon sequence with several carbon stars on or beyond the region populated by M stars, at small $(H-K)$ but large $(K-[12])$ magnitudes. This scatter contrasts with the tight carbon sequence that is observed in the Milky Way (Fig.\ 3 in Paper IV). We suspect that this is in part caused by the severe crowding in some LMC fields, affecting the near-IR aperture photometry. Differences in the strength of absorption in the H-band by carbonaceous molecules may cause additional scatter among carbon stars (Bessell \& Wood 1983; Catchpole \& Whitelock 1985). \subsubsection{Diagram of $(K-[12])$ versus $(K-L)$} The $(K-[12])$ versus $(K-L)$ colour-colour diagram shows much less scatter around well-defined carbon and oxygen sequences (Fig.\ 7). This makes it a much more powerful diagnostic diagram than the $(K-[12])$ versus $(H-K)$ diagram in typifying the chemical composition of the circumstellar dust. Noguchi et al.\ (1991a) introduced a very similar diagnostic using $(L-[12])$ and $(K-L)$ colours. We note, however, that some of the peculiar stars in our $(K-[12])$ versus $(H-K)$ diagram were too blue and hence too faint to be detected in the L-band. Still, the tight sequences prove that both the SAAO and ISO photometry are of good quality when comparing individual stars. We fit (by eye) a linear carbon sequence: \begin{equation} (K-L)=\frac{5}{11}\times(K-[12])-\frac{2}{11} \end{equation} and a superposition of even polynomials for the oxygen sequence: \begin{equation} (K-L)=0.35+0.007\times(K-[12])^2+0.0014\times(K-[12])^4 \end{equation} \begin{figure}[tb] \centerline{\psfig{figure=h1374.f7,width=88mm}} \caption[]{$(K-[12])$ versus $(K-L)$ diagram. Symbols as in Fig.\ 6.} \end{figure} \subsubsection{Diagram of $([12]-[25])$ versus $(K-L)$} \begin{figure}[tb] \centerline{\psfig{figure=h1374.f8,width=88mm}} \caption[]{$([12]-[25])$ versus $(K-L)$ diagram. Symbols are as in Fig.\ 6. Oxygen-rich sources and carbon stars occupy distinct areas in this diagram. The dividing line (dotted) between stars with carbon- and oxygen-rich dust is taken from Epchtein et al.\ 1987.} \end{figure} Another colour-colour diagram that separates carbon- from oxygen-rich sources is the $([12]-[25])$ versus $(K-L)$ diagram (Fig.\ 8). The confirmed oxygen-rich sources show a linear relationship between the $([12]-[25])$ and $(K-L)$ colours, possibly flattening out at $(K-L)>1.5$ mag. The LMC stars generally follow the separation determined for galactic stars (dotted line in Fig.\ 8, taken from Epchtein et al.\ 1987). \subsubsection{Diagram of $([25]-[60])$ versus $([12]-[25])$} \begin{figure}[tb] \centerline{\psfig{figure=h1374.f9,width=88mm}} \caption[]{$([25]-[60])$ versus $([12]-[25])$ diagram. Symbols are as in Fig.\ 6. Carbon stars are not well separated from oxygen-rich sources, although carbon stars seem to be relatively bright at 60 $\mu$m.} \end{figure} The aim of obtaining 60 $\mu$m flux densities for stars in the LMC is mainly to probe the coolest circumstellar dust. The 60 $\mu$m flux density is expected to increase as prolonged mass loss first extends the CSE and again as reduced mass loss results in a detached shell. This evolution might be seen in $([25]-[60])$ versus $([12]-[25])$ diagrams (Fig.\ 9, see also van der Veen \& Habing 1988). Unfortunately, the accuracy of the ISO photometry at 60 $\mu$m is not very high for most of these LMC sources, and the diagram contains a lot of scatter. Perhaps the most obvious thing to learn from this diagram is that carbon stars tend to be relatively bright at 60 $\mu$m, yielding $([25]-[60])\sim1.5$ to 3 mag. Although oxygen-rich sources can have similar colours, there are many oxygen-rich sources with $([25]-[60])<2$ and $([12]-[25])>0.6$ mag, colours not seen for any carbon star in our sample. This is similar to the findings of van der Veen \& Habing (1988), but our LMC sources have bluer $([12]-[25])$ and redder $([25]-[60])$ colours than do their Milky Way sources. However, the LMC $([12]-[25])$ colours do not differ much from those discussed by Le Bertre et al.\ (1994). \subsection{Comments on particular objects} \subsubsection{GRV0519$-$6700} The referee Dr.\ Peter Wood conveys that an optical spectrum of GRV0519$-$6700 shows it to be a carbon star, in good agreement with its IR colours of $(H-K)=0.7$ and $(K-[12])=1.0$ mag. \subsubsection{HV12070} The CVF spectrum of HV12070 shows only a hint of the silicate feature, whilst the IR colours cannot distinguish between oxygen- and carbon-rich dust of the optically thin CSE of this MS-type star. \subsubsection{HV2379} The PHOT-S spectrum of HV2379 suggests SiC emission, but its CVF spectrum does not. This may be a result of changes in the properties of the CSE or the dust. Its IR colours leave no doubt about the carbon-rich nature of the dust. \subsubsection{HV2446, 888, 996, and SP77 30$-$6} These late-M stars all have prominent silicate emission and IR colours that unambiguously indicate oxygen-rich dust. \subsubsection{IRAS04286$-$6937, 04539$-$6821, 04557$-$6753, 05009$-$6616, 05113$-$6739, 05295$-$7121, 05300$-$6651, 05360$-$6648, and TRM45 and 72} The position of these objects in the $(K-[12])$ versus $(H-K)$ or $(K-L)$ colour-colour diagrams does not clarify the chemical composition of their CSEs. The $([12]-[25])$ versus $(K-L)$ diagram, however, unambiguously indicates that the dust around these stars is carbon rich. The IR colours of IRAS05113$-$6739 at the three ISO epochs for this star all lie along the carbon sequences in the $(K-[12])$ versus $(H-K)$ and $(K-L)$ diagrams. Ground-based L-band spectra of IRAS05009$-$6616 and 05300$-$6651 show the 3.1 $\mu$m absorption feature due to HCN and C$_2$H$_2$ molecules, indicating carbon-rich photospheres (van Loon et al.\ 1999). \subsubsection{IRAS04374$-$6831} The position of IRAS04374$-$6831 in the $([12]-[25])$ versus $(K-L)$ diagram indicates carbon-rich dust. Its PHOT-S spectrum, which does not clearly reveal the chemical composition of the dust by itself, is then marginally consistent with SiC emission. \subsubsection{IRAS04496$-$6958} IRAS04496$-$6958 shows strong absorption by carbonaceous molecules at 3.1 $\mu$m, already known from ground-based L-band spectroscopy (van Loon et al.\ 1999). Related, but weaker, absorption is visible at 3.8 $\mu$m, and possibly around 8 $\mu$m. Surprisingly, this carbon star has silicate emission too, indicating the presence of oxygen-rich dust (see Trams et al.\ 1999). Its IR colours indicate carbon-rich dust, hence the oxygen-rich dust is only a minor component. \subsubsection{IRAS04530$-$6916} With $(K-L)=2.13$, $(K-[12])=6.9$ and $([12]-[25])=2.3$ mag, the IR colours of IRAS04530$-$6916 imply that the dust around this very luminous and red object must be oxygen rich. \subsubsection{IRAS04545$-$7000, 05298$-$6957, 05329$-$6708, 05402$-$6956, and WOH G64} These OH maser sources all show the silicate feature in self-absorption, and also their IR colours clearly indicate oxygen-rich dust. \subsubsection{IRAS05003$-$6712} The IR colours of IRAS05003$-$6712 unambiguously classify the dust as oxygen rich. The PHOT-S spectrum shows a hint of the silicate feature. A ground-based L-band spectrum of this star shows a featureless continuum around 3.1 $\mu$m, indicating an oxygen-rich photosphere (van Loon et al.\ 1999). \subsubsection{IRAS05112$-$6755} The dust around IRAS05112$-$6755 is classified as carbon rich on the basis of the position in the $([12]-[25])$ versus $(K-L)$ diagram. There is a hint of 8 $\mu$m absorption in the PHOT-S spectrum of IRAS05112$-$6755. A ground-based L-band spectrum of this object shows the strong absorption at 3.1 $\mu$m found in carbon-rich stellar photospheres (van Loon et al.\ 1999). \subsubsection{IRAS05128$-$6455 and 05190$-$6748} The absence of clear indications for the presence of the silicate feature in the PHOT-S spectra of these stars suggest that their dust may be carbon rich, which is also indicated by their $([12]-[25])$ and $(K-L)$ colours. \subsubsection{IRAS05289$-$6617} The CVF spectrum of IRAS05289$-$6617 shows prominent SiC emission. Hence it is probably a mass-losing carbon-rich AGB star in the LMC rather than a foreground object. We have not yet identified its near-IR counterpart. \subsubsection{IRAS05348$-$7024} The CVF spectrum of IRAS05348$-$7024 shows weak SiC emission. The carbon-rich nature of the dust around this object is also indicated by its position in the $([12]-[25])$ versus $(K-L)$ diagram. \subsubsection{IRAS05506$-$7053} IRAS05506$-$7053 is the only star in our sample that could not be detected at 12 $\mu$m. Assuming a 12 $\mu$m flux density $<0.03$ Jy, the $(K-[12])$ colour would be $<6.2$ mag and probably $([12]-[25])>1.5$ mag. At $(K-L)=3.3$ mag, this suggests an oxygen-rich CSE. \subsubsection{IRAS05558$-$7000} The CVF spectrum of IRAS05558$-$7000 is similar to the CVF spectra of IRAS05298$-$6957 and 05402$-$6956, showing silicate emission that is becoming optically thick at 10 $\mu$m. The IR colours of IRAS05558$-$7000 unambiguously imply that the dust is oxygen-rich. \subsubsection{IRAS05568$-$6753} The PHOT-S spectrum of IRAS05568$-$6753 shows a featureless dust continuum around 10 $\mu$m, suggesting pure amorphous carbon dust emission. The near-IR counterpart of this object has yet to be identified. \subsubsection{SHV0500193$-$681706} The CVF spectrum of SHV0500193$-$681706 shows a featureless dust continuum around 10 $\mu$m, suggesting pure amorphous carbon dust emission. The carbon-rich nature of the dust is confirmed by the position in the $(K-[12])$ versus $(K-L)$ diagram. Inaccuracy of its 25 $\mu$m flux density causes the rather odd position among the oxygen-rich stars in the $([12]-[25])$ versus $(K-L)$ diagram. \subsubsection{SHV0500233$-$682914} The CVF spectrum of SHV0500233$-$682914 shows a hint of SiC emission, and also its IR colours clearly indicate that the dust around this star is carbon rich. \subsubsection{SHV0502469$-$692418, 0522023$-$701242 and 0524565$-$694559} The carbon star SHV0502469$-$692418, the M-type star SHV0522023$-$701242 and the MS-type star SHV0524565$-$694559 are surrounded by an optically thin CSE and hence it is difficult to classify the chemical type of their dust from IR colour-colour diagrams. \subsubsection{SHV0522118$-$702517} SHV0522118$-$702517 was tentatively classified an S-type star by Hughes \& Wood (1990). Its IR colours are clearly similar to those of carbon stars. This suggests that carbon-rich dust dominates the absorption and emission characteristics of the CSE despite the under-abundance of carbon atoms in its photosphere. Noguchi et al.\ (1991b) show that the IR colours of the CSE indicate oxygen-rich dust in case of an MS-type star. Also, CS stars show 3 $\mu$m absorption from HCN and C$_2$H$_2$ molecules, whereas SC stars do not (Catchpole \& Whitelock 1985; Noguchi \& Akiba 1986). This suggests that carbon chemistry is dominant in CS stars, but not in SC stars. Thus, we identify SHV0522118$-$702517 with a CS star. Dust-enshrouded S stars --- including MS and CS stars --- that have $(K-L)>1$ mag are very rare in the Milky Way, and none are known with $(K-L)>2$ mag (Noguchi et al.\ 1991b). Hence, with $(K-L)=1.3$ mag, SHV0522118$-$702517 is among the most obscured S stars known. \subsubsection{SHV0530323$-$702216} The late-M type star SHV0535442$-$702433 has $([12]-[25])=1.56$ mag. Its near-IR colours are rather blue and $(K-L)$ is not expected to be larger than unity. Hence the position of this object in the $([12]-[25])$ versus $(K-L)$ diagram suggests that the dust is oxygen rich. \subsubsection{SHV0535442$-$702433} The carbon star SHV0535442$-$702433 is surrounded by an optically thin CSE, and hence the IR colours are difficult to use for classifying the chemical type of the dust. The location among oxygen-rich stars in the $([12]-[25])$ versus $(K-L)$ diagram is caused entirely by the inaccuracy of its 25 $\mu$m flux density yielding a spuriously red $([12]-[25])\sim3$ mag. \section{Conclusions} ISO spectroscopy is used to determine the chemical type of the dust around obscured cool evolved stars in the LMC. ISO photometry at 12, 25 and 60 $\mu$m is presented, together with quasi-simultaneous near-IR photometry from the ground (SAAO). The accuracy and sensitivity of the ISOPHOT photometry is not much better than can be achieved from properly treated IRAS data. The ISOCAM photometry is much more reliable because it is based on imaging, and an order of magnitude more sensitive than was IRAS. Colour-colour diagrams prove that relative photometry is reliable. A combination of $(K-[12])$ and $([12]-[25])$ versus $(K-L)$ diagrams provide a reliable way of distinguishing between carbon- and oxygen-rich dust, provided the CSE has sufficient optical depth. The combination of ISO spectra and photometry enabled us to securely classify the chemical type of the dust around nearly all stars in our sample. This was previously known for only a minority of the stars. Surprisingly, the $(K-[12])$ versus $(H-K)$ diagnostic diagram contains a lot of scatter especially among carbon stars. Many of the obscured AGB stars in our sample are carbon stars: 46\% amongst the LMC stars that were detected by IRAS (Table 1). M stars were always found to be surrounded by oxygen-rich dust. In particular, all detected OH maser sources show self-absorbed silicate emission. As in the Milky Way, the fact that no M star with carbon-rich dust has ever been found suggests that HBB cannot efficiently turn carbon stars back into oxygen-rich stars. The dust around the dust-enshrouded S star SHV0522118$-$702517 has the characteristics of carbon-rich material, suggesting it is actually a CS star. Surprisingly, the dust around the carbon star IRAS04496$-$6958 has a (minor) oxygen-rich component (Trams et al.\ 1999). \acknowledgements{We would like to thank everyone at VilSpa (Madrid) for the discussions and advices during the various stages of ISO data reduction, in particular Drs.\ Jos\'{e} Acosta-Pulido, Carlos Gabriel, Rene Laureijs, Thomas M\"{u}ller, and Bernhard Schulz, and Dr.\ P\'{e}ter \'{A}brah\'{a}m in Heidelberg. The ISOCAM data presented in this paper was analysed using ``CIA'', a joint development by the ESA Astrophysics Division and the ISOCAM Consortium. The ISOCAM Consortium is led by the ISOCAM PI, C. Cesarsky, Direction des Sciences de la Mati\'{e}re, C.E.A., France. The ISOPHOT data presented in this paper was reduced using PIA, which is a joint development by the ESA Astrophysics Division and the ISOPHOT Consortium. We also thank Dr.\ Romke Bontekoe for help with studying the IRAS data, and the referee Dr.\ Peter Wood for his suggestions that improved the presentation considerably. We made use of the SIMBAD database, operated at CDS, Strasbourg, France. This research was partly supported by NWO under Pionier Grant 600-78-333.\\ (JvL: O trabalho mais importante foi feito por um anjo).}
\section{Introduction} The light quark masses, despite being fundamental parameters of the Standard Model, are surprisingly poorly known at present. Since, in the standard (``large condensate'') scenario for the chiral counting in Chiral Perturbation Theory (ChPT), however, the ratio $m_s/(m_u+m_d)=24.4\pm 1.5$ is known\cite{leutwylerqmasses}, fixing either $m_s$ or $m_u+m_d$ is sufficient to determine the other. In this paper we focus on $m_s$, which has been the subject of several recent investigations, employing both QCD sum rules\cite{jm,narison95,cps,cfnp,jamin,km3388,dps,kmtau,pp98,cdh98} and lattice QCD \cite{rt97,al97,alrev,gough97,sesam,cppacs,ape,qcdsf,k98,gimenez,jlqcd,bsw99}. On the lattice, reliable results are presently restricted to the quenched approximation and vary somewhat depending on the fermion implementation and on whether one chooses $m_K$ or $m_\phi$ when matching to physical meson masses. Unquenched calculations are at a more preliminary stage, but suggest unquenching is likely to lower quenched results\cite{rt97,sesam,cppacs,k98}. Recent quenched results for $m_s$ (where here, as in what follows, all masses, unless otherwise specified, will be quoted at a scale $\mu=2$ GeV in the $\overline{MS}$ scheme) are $110\pm 20 \pm 11$ MeV\cite{rt97}, $122\pm 20$ MeV\cite{al97}, $95\pm 16$ MeV\cite{gough97}, $115\pm 2\ (143\pm 6)$ MeV\cite{cppacs} (for $m_K\ (m_\phi )$ input), $111\pm 12$ MeV\cite{ape}, $98.1\pm 2.4$ MeV\cite{qcdsf} (update as quoted by Kenway\cite{k98}), $130\pm 2\pm 18$ MeV\cite{gimenez}, $129\pm 12$ MeV\cite{jlqcd} and $95\pm 26$ MeV\cite{bsw99}. For dynamical ($N_f=2$) simulations, the CP-PACS results $m_s\sim 70\ (80)$ MeV (with $m_K \ (m_\phi )$ as input) and SESAM result $151\pm 30$ MeV\cite{sesam} do not appear compatible. (Note, however, that there are unresolved issues of chiral extrapolations, renormalization constants, and the lack of extrapolation to the continuum limit in the SESAM result; for a further discussion of these issues see Section 6 of the review by Bhattacharya and Gupta\cite{rt97}, and Allton\cite{alrev}.) Recent attempts to extract $m_s$ using QCD sum rules are of four types. Two employ flavor-breaking differences of vector and/or axial vector current correlators, either the isovector-hypercharge difference\cite{narison95,km3388} or the isovector-strange difference\cite{kmtau,pp98,cdh98}. The others involve analyses of the strange scalar\cite{jm,cps,cfnp,jamin} and strange pseudoscalar\cite{jm,dps} channels. For the vector current isovector-hypercharge difference, isospin-breaking corrections are crucial due to the high degree of cancellation on the hadronic side of the sum rule. These are, for natural reasons\cite{km3388,kmcvc}, larger than one would naively expect, and can only be obtained indirectly\cite{km3388}. The high degree of cancellation also magnifies experimental and theoretical uncertainties, leading to significant errors on $m_s$ ($m_s=107\pm\sim 40$ MeV\cite{mw98}, if one takes the isovector spectral function from hadronic $\tau$ decay data\cite{ALEPH,OPAL}). The isovector-strange difference, which employs $\tau$ decay data for the isovector and strange spectral functions, is subject to complications resulting from (1) very poor convergence of the perturbative series for the longitudinal contributions on the OPE side\cite{kmtau,pp98}, and (2) strong cancellation, which again significantly magnifies the effect of the (otherwise rather small) experimental errors. Combining errors in quadrature and translating to the scale $2$ GeV, one finds $m_s=163\pm 50$ MeV\cite{kmtau}, $143\pm 42$ MeV\cite{pp98} and $157\pm 37$ MeV\cite{cdh98}. The third approach involves the correlator $\langle 0\vert T\left( J_{A,s} J^\dagger_{A,s}\right)\vert 0 \rangle$ (where $J_{A,s}\equiv\partial_\mu \bar{s}\gamma_5\gamma^\mu u$)\cite{jm,dps}. In this case, the $K$ pole contribution to $\rho (s)$ is known, but contributions from the continuum ($K(1460)$, $K(1830)$ resonance region) are not. Both analyses of this type employ a sum-of-Breit-Wigner modulation of the near-threshold tree-level ChPT form to deal with this problem, together with some assumptions about the relative size of the two unknown resonance decay constants. As discussed in Refs.~\cite{cfnp,bgm}, this type of normalization procedure for a continuum spectral ansatz can involve potentially significant systematic errors. For example, in the scalar strange channel, it leads to a rather poor representation of the actual spectral function (compare the model versions of Refs.~\cite{jm,cps} with that of Ref.~\cite{cfnp}). In addition, because both analyses employ Borel transformed (SVZ) sum rules\cite{svz}, they involve the conventional ``continuum'' (or local duality) approximation to $\rho (s)$ above $s=s_0$ (where the ``continuum threshold'', $s_0$, is a parameter of the analysis). Since, for Borel masses, $M$, in the stability windows of both analyses, $M^2<<s_0$ is far from being satisfied, there are potential systematic errors resulting from violations of local duality in the ``continuum'' region. (That contributions from the ``continuum'' region are not numerically negligible can also be inferred, e.g., from the figures of Ref.~\cite{dps}). It is worth noting that violations of local duality are, in general, expected in kinematic regimes where typical resonance widths are comparable to or smaller than typical resonance separations. This has been quantified in the isovector vector channel, where the hadronic spectral function is known experimentally; violations of local duality in that channel, even at scales $\sim m_\tau^2$, are seen to be, in general, rather large\cite{kmfesr}. Taking the results from Refs.~\cite{jm,cps} which employ $\Lambda_{QCD}=380$ MeV (which best corresponds to recent experimental values for $\alpha_s(m_\tau^2)$\cite{ALEPH,OPAL}), one finds, for $m_s$, the values $\simeq 100$ MeV\cite{jm} (see Figure 5) and $115\pm 19$\cite{dps} (where the errors do not reflect the systematic uncertainties discussed above). The last approach involves the correlator, relevant to the strange scalar channel, \begin{eqnarray} \Pi (q^2)\ &=&\ i\int\, dx\, e^{iq\cdot x}\langle 0\vert T\left( J_s(x)\, J_s^\dagger (0)\right)\vert 0\rangle \nonumber\\ &=& i \left( m_s-m_u\right)^2\, \int\, dx\, e^{iq\cdot x}\langle 0\vert T\left( \bar{s}u(x)\, \bar{u}s(0) \right)\vert 0\rangle\ , \label{scalarcorrelator} \end{eqnarray} which has been analyzed, in Refs.~\cite{jm,cps,cfnp,jamin}, using the Borel transformed sum rule method. This channel is a favorable one for sum rule analyses because the hadronic spectral function, $\rho (s)$, turns out to be rather well-determinable (albeit indirectly) from experimental data. The reason is that since, experimentally, $K\pi$ scattering is known to be purely elastic up to $s=(1.58\ {\rm GeV})^2$\cite{LASS,dunwoodie}, the phase of the timelike scalar $K\pi$ form factor, $d(s)$, is known up to this point. $d(s)$, moreover, satisfies an Omnes representation \begin{equation} d(s)=d(0)p(s)\, exp\left[ {\frac{s}{\pi}}\int_{th}^\infty\, dt {\frac{\delta (t)}{t(t-s-i\epsilon )}}\right]\ , \label{omnes} \end{equation} where $p(s)$ is a polynomial normalized to $1$ at $s=0$, $\delta (s)$ is the phase of $d(s)$, and the normalization $d(0)$ is known from ChPT and $K_{e3}$\cite{glmff}. In order to satisfy quark counting rules for the asymptotic behavior of $d(s)$, $\delta$ must $\rightarrow (N+1)\pi$ as $s\rightarrow\infty$ if $p(s)$ is of degree $N$. In previous analyses $p(s)$ has been assumed to be $1$ (or, equivalently, very slowly varying in the range of $s$ important to the analysis). The fact that, at the edge of the experimentally explored region, $s=(1.7\ {\rm GeV})^2$, the $I=1/2$ $K\pi$ phase shift is very nearly $\pi$\cite{LASS,dunwoodie,estabrooks}, the asymptotic value required if $p(s)\equiv 1$, suggests, first, that this assumption is rather plausible and, second, that the treatment of the high energy behavior of the phase, $\delta (s)$, employed in Ref.~\cite{cfnp}, {\it i.e.} $\delta (s)=\pi$ for $s>(1.7\ {\rm GeV})^2$, is sensible. We will also see below that, making these assumptions, one obtains extremely well-satisfied sum rules, providing strong {\it post facto} justification as well. With these assumptions, the $K\pi$ contribution to $\rho$ is completely determined, and multiparticle contributions begin only above $s\sim 2.5\ {\rm GeV}^2$. The resulting $K\pi$ contribution to $\rho (s)$ is, moreover, much smaller in the $K_0^*(1950)$ than in the $K_0^*(1430)$ region (see Fig. 2 of Ref.~\cite{cfnp}). Since the $K_0^*(1950)$ $K\pi$ branching fraction is known to be $\sim 50\%$\cite{pdg98}, the remaining multiparticle contributions to $\rho$, in the $K_0^*(1950)$ region, will also be small. This means that, to the extent that one may expect multiparticle contributions to $\rho$ to be resonance dominated, (1) such contributions will play only a small role in the full spectral function out to $s\sim 4\ {\rm GeV}^2$, and (2) in this region, the magnitude of these contributions can be obtained by combining information from the Omnes-generated $K\pi$ component and the known experimental $K_0^*(1950)$ $K\pi$ branching fraction. We do not quote results from the analyses of Refs.~\cite{jm,cps} since the spectral ansatz employed there has been shown to be incompatible with the Omnes representation\cite{cfnp}. Results from Refs.~\cite{cfnp,jamin} are $m_s=\, 91\rightarrow 116\ {\rm MeV}$\cite{cfnp} and $116\pm 22\ {\rm MeV}$\cite{jamin}. The main potential uncertainty in these analyses is that associated with violations of local duality in the ``continuum'' region, $s>s_0$. That continuum contributions are non-negligible may be seen from the lack of a stability plateau for $m_s$ in the analysis of Ref.~\cite{jamin}. Finite energy sum rules constructed using the extracted values of $m_s$ from Refs.~\cite{cfnp,jamin} and the Omnes-generated spectral ansatz of Ref.~\cite{cfnp} (also used in Ref.~\cite{jamin}) as input have been shown to be moderately well satisfied\cite{kmfesr}. The residual deficiencies could arise from either shortcomings in the spectral ansatz or approximations on the OPE side of the sum rules. We will investigate this question in what follows. \section{The Strange Scalar Channel Analysis} Unitarity and analyticity ensure that correlators like $\Pi (q^2)$, defined above, satisfy either appropriately subtracted conventional dispersion relations or finite energy sum rules (FESR's). Borel transformation of the former leads to the conventional\cite{svz} (SVZ) form of QCD sum rules. In what follows we choose to employ FESR's, which are based on Cauchy's theorem, rather than the Cauchy representation theorem, and work with the ``Pac-man'' contour, which traverses both sides of the physical cut, and is closed by a circle of radius $s_0$ in the complex-$s$ plane. One then has \begin{equation} \int_{s_{th}}^{s_0}\, ds\, \rho (s) w(s)\, =\, {\frac{-1}{2\pi i}}\, \oint_{\vert s\vert =s_0}\, ds\, w(s)\Pi (s)\ , \label{basicfesr} \end{equation} valid for any function $w(s)$ analytic in the region of the contour, where $\rho (s)$ is the spectral function and $s_{th}$ the relevant physical threshold. To make this relation useful, one wishes to use the physical spectral function, whose extraction was described in the Introduction, on the LHS, and the OPE for the correlator on the RHS. Since the OPE is proportional to $\left( m_s-m_u\right)^2$, this provides a method for extracting $m_s-m_u$. To make the results of this analysis reliable, $s_0$ should satisfy two criteria: (1) it should be large enough to make the OPE representation of the RHS accurate, and (2) it should be small enough that one still has experimental information on $\rho (s)$ for $s<s_0$ on the LHS. The latter criterion, in the present case, restricts us to the region up to and including the $K_0^*(1950)$ (i.e., $s_0<4\ {\rm GeV}^2$ or so). Since, even at $s=4\ {\rm GeV}^2$, resonance widths and separations are comparable, one expects local duality to be violated, creating potentially significant uncertainties on the OPE side. It has been argued, however, that even if local duality is not exactly satisfied for given timelike $s_0$, the OPE may, nonetheless, provide a good representation of the correlator on most of the circle $\vert s\vert =s_0$ (i.e., except possibly for a region of hadronic scale near the timelike real axis)\cite{pqw}. The isovector vector channel, for which the spectral function has been extracted from hadronic $\tau$ decay data, provides explicit empirical confirmation of this argument. First, the conventional FESR treatment of these decays\cite{taudecay} is very well satisfied, presumably because it involves a weight function (determined by kinematics) with a double zero at $s=s_0$ which cuts out contributions from the potentially dangerous part of the contour. Second, one may demonstrate that using weights $s^k$ (with no such zero at $s=s_0$) leads to significant violation of local duality, even at scale $s_0=m_\tau^2$, whereas FESR's based on weights having either a single or double zero at $s=s_0$ are all extremely well satisfied, even at scales significantly below $s_0=m_\tau^2$\cite{kmfesr}. Using such FESR's one may, thus, hope to construct sum rules in which the use of the ``continuum'' ansatz for the high energy portion of $\rho (s)$ may be avoided. In what follows, therefore, we study FESR's for the correlator, $\Pi (q^2)$, using weights of the form \begin{equation} w_A(s)=\left( 1-{\frac{s}{s_0}}\right)\left( A-{\frac{s}{s_0}}\right)\ . \label{NORM} \end{equation} In Eq.~(\ref{NORM}), $A$ is a free parameter, to be chosen in such a way as to optimize the reliability of the analysis. Since (1) instanton effects are known exactly only in dilute gas single-instanton-background approximation\cite{bbb93,nason} and (2) the instanton density used to obtain these results is known to be altered by quark and gluon condensate effects\cite{svzinst}, the magnitude of the resulting instanton contributions (as stressed by the authors of Refs.~\cite{bbb93,nason}) can be considered only a rough guide. For this reason, we restrict our attention to weights which, after integration, yield instanton effects (including corrections to the coefficients of both the unit operator and quark condensate in the OPE, as given in Ref.~\cite{nason}) which are small for all $s_0$ employed. While any $A>1$ turns out to satisfy this criterion, $A=2$ creates an optimal suppression and we will display results below for this value of $A$. We now discuss the input on the hadronic side of the sum rules. The $K\pi$ and multiparticle portions of $\rho (s)$ are obtained, using the Omnes representation, experimental $K\pi$ phases, and the experimental $K\pi$ branching fractions of the $K_0^*(1430)$ and $K_0^*(1950)$, as described in the Introduction. We employ two slightly different implementations of this procedure, the difference in the resulting $m_s$ values serving to quantify the effect of uncertainties in our knowledge of the hadronic spectral function. In the first implementation, the fit to the experimental $K\pi$ phases obtained in Ref.~\cite{jm} is used as input to the Omnes representation, and multiparticle contributions are allowed in both the $K_0^*(1430)$ and $K_0^*(1950)$ regions, constrained by the Particle Data Group's $K\pi$ branching fractions\cite{pdg98}. In the second implementation we employ the LASS fit\cite{LASS,dunwoodie} to the $K\pi$ phase shifts and inelasticities. For this fit, the form of the parametrization of phase shift data used is slightly different from that of Ref.~\cite{jm}. In addition, since the fit to the $K\pi$ amplitude is still purely elastic in the $K_0^*(1430)$ region, in employing this implementation we allow multiparticle contributions to the spectral function only from the $K_0^*(1950)$ region. (Note that the background effective range parameters, $a$, $b$, and $K_0^*(1430)$ resonance parameters, $m$, $\Gamma$, were mis-recorded in Ref.~\cite{LASS}, and should actually be $m=1.412\ {\rm GeV}$, $\Gamma =.294\ {\rm GeV}$\cite{dunwoodie,gnrmp}, and $a=2.19\ {\rm GeV}^{-1}$, $b=3.74\ {\rm GeV}^{-1}$\cite{dunwoodie}. It has been checked that these provide a satisfactory representation of the amplitude and phase shift data of Ref.~\cite{LASS}.) On the OPE side, it is convenient to work with $d^2\Pi (Q^2)/\left( dQ^2\right)^2\equiv \Pi^{\prime\prime}(Q^2)$, which satisfies a homogenous RG equation, so all scale dependence can be absorbed into the running mass and coupling. To improve the convergence of the perturbative series, and significantly reduce residual scale dependence, we follow standard practice and evaluate the OPE side using the ``contour improvement'' prescription of Ref.~\cite{contourimproved}, first summing logarithms through the renormalization scale choice $\mu^2=Q^2$, and then performing the integral around the circle $\vert s\vert =s_0$ numerically, using the exact solutions for the running mass and coupling which follow from the four-loop-truncated $\beta$ and $\gamma$ functions\cite{beta4,gamma4}. The expressions for the dimension $D=0,2,4,6$ contributions to $\Pi^{\prime\prime}(Q^2)$ are all known from previous work\cite{jm,cps,cfnp} and, for $\mu^2=Q^2$, are: (1) for $D=0$\cite{jm,cps}, \begin{eqnarray} \left[ \Pi^{\prime\prime}(Q^2)\right]_{D=0}&=& \frac{3(m_s-m_u)^2(Q^2)}{8\pi^2 Q^2}\left\{ 1 + \frac{11a(Q^2)}{3} + {a(Q^2)^2}\left( \frac{5071}{144}-\frac{35}{2}\zeta(3)\right)\right.\nonumber\\ & &\left.\qquad +{a(Q^2)^3}\left(\frac{1995097}{5184}-\frac{\pi^4}{36} -\frac{65869}{216}\zeta (3)+\frac{715}{12}\zeta (5)\right)\right\}\ , \label{d0} \end{eqnarray} where $a(Q^2)=\alpha_s (Q^2)/\pi$ and $\zeta (N)$ is the Riemann zeta function; (2) for $D=2$\cite{jm,cps}, \begin{equation} \left[ \Pi^{\prime\prime}(Q^2)\right]_{D=2}= -\frac{3(m_s-m_u)^2(Q^2)m_s^2(Q^2)}{4\pi^2 Q^4}\left\{ 1 + \frac{28a(Q^2)}{3} +{a(Q^2)^2}\left(\frac{8557}{72}-\frac{77}{3}\zeta(3)\right)\right\}\ ; \label{d2} \end{equation} (3) for $D=4$\cite{jm,cps}, \begin{eqnarray} \left[ \Pi^{\prime\prime}(Q^2)\right]_{D=4} &=& \frac{(m_s-m_u)^2(Q^2)}{Q^6}\left\{ 2\langle m_s\bar uu\rangle \left(1 + \frac{23}{3}{a(Q^2)}\right) -\frac{1}{9} I_G\left( 1 + \frac{121}{18} {a(Q^2)} \right) \right.\nonumber\\ & &\left. \ \ + I_s\left( 1+\frac{64}{9}{a(Q^2)}\right) -\frac{3}{7\pi^2}m_s^4(Q^2)\left(\frac{1}{a(q^2)}+ \frac{155}{24}\right)\right\}\ , \label{d4} \end{eqnarray} where $I_G$ and $I_s$ are the RG-invariant $D=4$ condensate combinations of Ref.~\cite{cps}; and (4) for $D=6$\cite{jm}, \begin{equation} \left[ \Pi^{\prime\prime}(Q^2)\right]_{D=6} = \frac{(m_s-m_u)^2(Q^2)}{Q^8}\left\{ 3 m_s\langle g\bar{u}\sigma F u\rangle -\frac{32}{9}\pi^2 \rho_{VSA} a \left( \langle \bar{s}s\rangle^2 + \langle \bar{u}u\rangle^2 +9 \langle \bar{s}s\rangle \langle \bar{u}u\rangle\right)\right\}\ , \label{d6} \end{equation} where $\rho_{VSA}$ is a parameter describing the deviation of the four-quark condensates from their vacuum saturation values. In the above equations, factors of $m_{u,d}$ have been dropped, except in the overall normalization $(m_s-m_u)^2$. For the input required on the OPE side of the sum rule we employ the following: $\alpha_s(m_\tau^2)=.334\pm .022$\cite{ALEPH} (needed as initial condition for the RG running of $\alpha_s$); for the $D=4$ condensates, $\langle \alpha_s G^2\rangle = 0.07\pm .01\ {\rm GeV}^4$\cite{narison97}, $\left( m_u+m_d\right)\langle \bar{u}u\rangle =-f_\pi^2 m_\pi^2$, and $0.7< r_c \equiv \langle \bar{s}s\rangle /\langle \bar{u} u\rangle <1$\cite{jm,cps}; for the $D=6$ condensates, $\langle g\bar{q}\sigma Fq\rangle =\left( 0.8\pm 0.2\ {\rm GeV}^2\right)\langle \bar{q} q\rangle$\cite{jm} and $\rho_{VSA}=5\pm 5$ (i.e., allowing, to be conservative, up to an order of magnitude deviation from vacuum saturation). We evaluate $m_s-m_u$ by optimizing the match of the OPE and hadronic sides of the sum rules in question in the range $3\ {\rm GeV}^2\leq s_0\leq 4\ {\rm GeV}^2$, for which the dominant $D=0$ contour-improved series is well behaved. To convert the extracted values of $m_s-m_u$ to those for $m_s$, we take the central ChPT value for $m_s/m_u$\cite{leutwylerqmasses}. The errors on the extracted value of $m_s$ associated with those on the input OPE parameters, together with their sources, are as follows (where we list only those cases where the error is greater than $0.1$ MeV): due to $\alpha_s(m_\tau^2)$, $\pm 1.0$ MeV; due to $\rho_{VSA}$, $\pm 0.7$ MeV; due to $\langle \alpha_s G^2\rangle$, $\pm 0.2$ MeV; due to $r_c$, $\pm 0.2$ MeV. More significant, on the OPE side, is the error associated with the truncation of the $D=0$ contributions at four-loop order. If we assume continued geometric growth of the expansion coefficients, including this estimate for the next order term lowers $m_s$ by $2.8$ MeV. In the final results below we have included this estimate of the five-loop order term in arriving at the quoted value, and assigned an error of {\it twice} the difference between the four-loop result and five-loop estimate. If added in quadrature, this source of error swamps all others on the OPE side. The sources and errors on $m_s$ associated with uncertainties on the hadronic side of the sum rules (where greater than $0.1$ MeV) are as follows: due to the error on the $K_0^*(1430)$ $K\pi$ branching fraction\cite{pdg98} (in the case of the first implementation described above, where multiparticle contributions are allowed in the $K_0^*(1430)$ region), $\pm 3.5$ MeV; due to that on $m_{K_0^*(1430)}$, $\pm 0.6$ MeV; due to that on $\Gamma_{K_0^*(1430)}$, $\pm 3.8$ MeV; and due to that on $\Gamma_{K_0^*(1950)}$, $\pm 0.4$ MeV. There is also a $0.4$ MeV difference between the extracted central values for the two different implementations of the hadronic spectral ansatz; this uncertainty is, however, swamped by that associated with the treatment of multiparticle contributions in the $K_0^*(1430)$ region in the case of the first implementation. We treat the latter as a systematic error common to both hadronic implementations, even though it is, strictly speaking, not consistent to include it in the case of the second implementation, based on the LASS fit, which is purely elastic in the $K_0^*(1430)$ region. Taking the average of the central values associated with the two different hadronic implementations, and adding errors in quadrature, we then arrive at our final result, \begin{equation} m_s\left( 4 \ {\rm GeV}^2\right)=\left( 115.1\pm 5.7 \pm 5.2\right) \ {\rm MeV}\ , \label{final} \end{equation} where the first error is associated with the OPE side of the sum rule (and is completely dominated by our estimate of the error due to trunctation of the $D=0$ series) and the second with uncertainties on the hadronic side (where errors associated with the inelasticity in the $K_0^*(1430)$ region and with the uncertainty in the $K_0^*(1430)$ width are roughly equally important and, again, dominate all other sources). This result corresponds to \begin{equation} m_s\left( 1 \ {\rm GeV}^2\right)=\left( 158.6\pm 7.9 \pm 7.2\right) \ {\rm MeV}\ . \label{onegev2} \end{equation} In Figure 1 we display the match between the OPE and hadronic sides of the $A=2$ sum rule in the fit window, for the optimized value of $m_s$. The agreement is obviously excellent, increasing our confidence in both the assumptions about the form of the Omnes representation of $d(s)$, and the value of $m_s$ that results. \section{Summary} We have determined $m_s$ from a FESR analysis in which the hadronic side of the sum rules employed is well determined by experimental data, modulo questions about the presence or absence of a polynomial factor in the Omnes representation of $d(s)$ and the high energy behavior of the phase $\delta (s)$. The excellent agreement between OPE and hadronic sides shown in the Figure argues strongly, albeit {\it post facto}, for the reliability of the input assumptions $p(s)=1$ and $\delta (s)=\pi$ for $s>(1.7\ {\rm GeV})^2$. We have attempted to make conservative estimates of the errors associated with uncertainties on the hadronic and OPE sides of the sum rules and conclude that $m_s\left( 4 \ {\rm GeV}^2\right) =115\pm 8\ {\rm MeV}$, compatible with results from recent quenched lattice calculations. It is worth noting that systematic uncertainties in the Borel transformed sum rule approach associated with the use of the ``continuum ansatz'' for the high energy portion of the hadronic spectral function are potentially significant in size on the scale of the errors quoted above. In Ref.~\cite{cps}, for example, a variation of $\pm 9$ MeV in $m_s\left( 1\ {\rm GeV}^2\right)$ results from variations of $\pm 0.5\ {\rm GeV}^2$ in the continuum threshold parameter, $s_0$. Such uncertainties are, of course, absent in the FESR approach, and help to keep the final errors small. \vskip .5in\noindent \begin{figure} [htb] \centering{\ \psfig{figure=ssmsfig.ps,height=10.cm}} \vskip .3in \caption{The OPE ($I_{OPE}$) and hadronic ($I_{had}$) sides of the $A=2$ FESR defined in the text. The solid line gives the OPE side, the dashed line the hadronic side.} \label{figone} \end{figure} \acknowledgements The author would like to thank Rajan Gupta, Tanmoy Bhattacharya and Michael Pennington for numerous useful discussions; Michael Pennington, in addition, for pointing out the existence of the transcription errors in the $K\pi$ parameters of Ref.~\cite{LASS}; and W. Dunwoodie for confirming the revised values of these parameters. The ongoing support of the Natural Sciences and Engineering Research Council of Canada, and the hospitality and support of the Special Research Centre for the Subatomic Structure of Matter at the University of Adelaide, where much of this work was completed, are also gratefully acknowledged.
\section{Introduction} \label{intro} An intriguing aspect of Bose-Einstein condensation (BEC) in a dilute atomic gas is that the internal atomic state of the condensate can be manipulated to produce quite novel systems. A number of interesting experiments have produced uncoupled multi-component condensates, in which two or more internal states of the condensate exist together in a magnetic or optical trap \cite{Wieman1,Ketterle1,Ketterle2,exp1,exp2}. These experimental studies, along with their theoretical counterparts, have investigated various topics such as the ground state of the system \cite{ho1,esry2,pu2,ho2,ohberg2,ao}, the elementary excitations \cite{meystre1,busch1,walls3,pu1,esry1,ohberg3,savage}, and the nonlinear dynamics of component separation \cite{sinatra}. Many fascinating properties can be studied by applying an external electromagnetic field that coherently couples the internal atomic states of the condensate~\cite{exp3,eschmann,williams1,Villain,ohberg,exp5,blakie,ZhangWalls,ZhangLiu}. In the experiment described in \cite{exp3}, the relative phase between two hyperfine components was measured using a technique based on Ramsey's method of separated oscillating fields \cite{Ramsey} and these results have motivated further theoretical investigation \cite{sinatra,eschmann}. In this system, several key parameters can be varied over a wide range of values, such as the coupling-field intensity and frequency, the confining potentials, the total number of atoms, and the temperature, making this a very rich system to explore. Several theoretical papers \cite{williams1,Villain,ohberg} have investigated the weak coupling limit of this system, where the intensity of the external driving field is very low but is turned on for a time long compared to the period of oscillation in the magnetic trap. A clear analogy exists between this system and the Josephson junction \cite{Barone}. In the Josephson junction, identical particles in spatially separated condensates are coupled via the tunneling mechanism \cite{Walls2,Juha,Walls1,Wright,Legget1,Legget2,Smerzi1,Smerzi2,Clark}. In this two-component system, however, two distinct internal states of the condensate are coupled by an applied field. By adjusting the magnetic fields that confine the atoms, the degree of spatial overlap between the two components can be controlled. Recently, there have been both experimental \cite{exp5} and theoretical \cite{blakie} studies on the dressed states of a driven two-component condensate, drawing an analogy to the dressed states of a driven two-level atom in quantum optics. Due to the interplay between the internal and external degrees of freedom, the condensate dressed states have spatial structure that depends on the trap parameters, the mean-field interaction, and the frequency and intensity of the driving field \cite{blakie}. In the experiment reported in \cite{exp5}, the dressed states were created via an adiabatic passage by sweeping the detuning. In this paper, we focus on the limit of a very strong and sustained coupling between hyperfine states, which is the situation achieved experimentally in \cite{exp5}. In that experiment, a BEC of about $8\times 10^5$ atoms was produced in the $|F=1,M_F=-1\rangle$ hyperfine state of ${}^{87}$Rb, at a temperature close to zero, $T\approx 0$. The atoms were confined in a time-averaged, orbiting potential (TOP) magnetic trap by a harmonic potential with axial symmetry along the vertical axis. An external field was then applied that coupled the $|1,-1\rangle$ state to the $|2,1\rangle$ hyperfine state via a two-photon transition. The Rabi frequency was five to ten times larger than the vertical trap frequency and the detuning could be adjusted arbitrarily. Due to their different magnetic moments and the force of gravity, the two hyperfine states sit in shifted traps offset along the vertical axis \cite{exp4}. The degree of separation could be controlled by adjusting the magnetic trapping fields. The subsequent behavior of the system described in \cite{exp5} was quite unexpected: after the coupling field was turned on, the Rabi oscillations between the states appeared to collapse and revive on a time scale which was long compared to the Rabi period of 3 ms. An example of this behavior taken from \cite{exp5} is shown in Fig.~1. It was observed that the period of this modulation increased with decreasing detuning. The behavior of the system also depended critically on the separation between the traps for each state. As the separation was taken to zero, the effect went away. The evolution of the density of each component was also very interesting. Each component cycled between a density profile with one peak and a profile with two peaks. The two-peaked structure was most clearly visible around the collapse time, which is $t \approx 20$ ms for the case shown in Figure 1. \begin{figure} \centerline{\epsfig{file=FIG_1.eps,height=2.5in}} \vspace{0.05cm} \caption{This plot shows the modulation of the fractional population in the (1,-1) state. The top line is experimental data \protect\cite{exp5} while the bottom line is the result of a numerical calculation of the three-dimensional, two-component Gross-Pitaevskii equation Eq.~(\ref{main1}) (below). The coupling strength and detuning were chosen for the calculation to be $\Omega = 350$ Hz and $\delta = -188$ Hz, respectively.} \end{figure} A numerical calculation of the three-dimensional, coupled Gross-Pitaevskii (GP) equations Eq.~(\ref{main1}) (below) describing the system in the zero temperature limit agrees, at least qualitatively, with the outcome of the experiment, as shown in Figure 1. Of course, the agreement between the numerical integration of Eq.~(\ref{main1}) and the laboratory data does not in itself provide an intuitive explanation of the underlying physical mechanism responsible for the collapse-revival behavior. In this paper, we present a detailed analysis of this problem, and arrive at a rather simple model that explains the major features of the system's behavior. Before presenting the details of our analysis, it is useful to first give an overview of the results. There are two main concepts that play key roles in obtaining an intuitive understanding of this problem. First, there is a clear separation of time scales: the period of Rabi oscillations between internal states is much shorter than the period of the trap. That is, the internal dynamics occur on a much shorter time scale than the motional dynamics of the system. It is therefore useful to go to a frame rotating at the effective Rabi frequency. In this rotating frame, we show that there exists a weak coupling between the low lying motional states which is proportional to the offset between the two traps. This weak coupling has the effect of modulating the amplitude of the fast Rabi oscillations in the lab frame. The second key point is to understand exactly which motional states are excited. They are not the linear response collective excitations (normal modes) that have been studied frequently in the BEC literature \cite{griffin,stringari3,fetter}. Instead, they are many-particle topological modes determined by the self-consistent solutions to the two-component GP equations Eq.~(\ref{selfcons}). The well known vortex mode~\cite{edwards1,stringari1,sinha,rokhsar1} is one example of such an excitation in which phase continuity requires quantized circulation around a vortex core. The related excitation which plays a key role in this paper does not exhibit circulation but has a node in the wavefunction amplitude and exhibits odd-parity behavior characteristic of such a dipole mode. For a single component in the limit of the uniform gas, the exact solution of this mode is known as a dark soliton~\cite{Clark}. In that case the scale of the density perturbation around the node is the healing length and is determined by a balance of kinetic and mean-field interaction energies. In the problem we consider here, however, it is necessary to account for the mean field of the the remaining population in the condensate ground state. Consequently the two modes---the ground state and the dipole mode---are inextricably linked and must be determined self consistently. We first present a detailed theoretical analysis in Section~\ref{theory}. After making several reasonable approximations, we arrive in Section~\ref{twomode} at the two-mode model -- a simplified description that encapsulates the essential properties of the system. In Section~\ref{casestudy} we present results of numerical calculations that illustrate the behavior of the system and we compare our model with the exact numerical solution of the coupled GP equations. We finally summarize our work in Section~\ref{conclusion} and suggest further studies based on our understanding of this phenomenon. \section{Theoretical Description} \label{theory} The following theoretical development was motivated by the experiment described in \cite{exp5}. Therefore, we have not tried to keep our calculations general, but instead have made several assumptions based on that particular situation. However, our approach could easily be extended to treat a broader class of systems. We give a brief discussion in the conclusion of the paper about possible extensions of this work to other interesting systems. We begin this section by writing down the coupled mean-field equations, valid for zero temperature, that describe this driven, two-component BEC. In Section~\ref{extint} we rewrite the mean-field equation in a direct-product representation that clearly separates out the external and internal dynamics. We then go to a frame rotating at the effective Rabi frequency in Section~\ref{rotframe} in order to focus on the slower motional dynamics of the system. After making some approximations in Section~\ref{approx}, we finally arrive at the main result of our study in Section~\ref{twomode}, the two-mode model. \subsection{Coupled Mean-field Equations} \label{coupledmfeq} A mean-field description of this many-body system that includes the atom-field interaction has been developed, which generalizes the standard Gross-Pitaevskii (GP) equation to treat systems with internal state coupling~\cite{williams1,Villain,cirac1}. The resulting time-dependent GP equation describing the driven, two-component condensate is \begin{equation} \begin{array}{ccc} i \left( \! \! \begin{array}{c} \dot{\psi}_1 \\ \dot{\psi}_2 \end{array} \!\! \right) \!\!\! &=& \!\!\! \left( \!\! \begin{array}{cc} H_1^0 + H_1^{\rm{MF}} + \delta/2 & \Omega/2 \\ \Omega/2 & H_2^0 + H_2^{\rm{MF}} - \delta/2 \end{array} \! \right) \!\! \left( \!\! \begin{array}{c} \psi_1 \\ \psi_2 \end{array} \!\! \right) \,\, . \end{array} \label{main1} \end{equation} The Hamiltonians describe the evolution in the trap $H_i^0$ and the mean field interaction $H_i^{\rm{MF}}$ for each component \begin{eqnarray} H_i^0 &=& -{1\over{2}} \nabla^2 + {1\over{2}}[({\rho\over\alpha})^2 + (z + \gamma_i \,z_0)^2] \nonumber \\ H_i^{\rm{MF}} &=& N(\lambda_{ii} |\psi_i|^2 + \lambda_{ij} |\psi_j|^2) \,\,, \label{Hs} \end{eqnarray} where $\gamma_1=-1$ and $\gamma_2=1$, and $z_0$ is the shift of each trap from the origin along the vertical axis. The factor $\alpha=\omega_z/\omega_{\rho}$ is the ratio of axial and radial trap frequencies. In the experiment reported in \cite{exp5}, $\alpha > 1$. The mean-field strength is characterized by $\lambda_{ij} = 4\pi a_{ij}/z_{\rm{sho}}$, which depends on the scattering length $a_{ij}$ of the collision. In general there will be three different values, one for each type of collision in this two-component gas: $a_{11}$, $a_{22}$, $a_{12}$. The detuning between the driving field and the hyperfine transition is given by $\delta$ while $\Omega$ denotes the strength of the coupling. We work in dimensionless units: time is in units of $1 / \omega_z$, energy is in terms of the trap level spacing $\hbar \, \omega_z$, and position is in units of the harmonic oscillator length $z_{\rm{sho}} = \sqrt{\hbar / m_{\rm{Rb}} \, \omega_z}$. The complex functions $\psi_i({\bf{r}},t)$ are the mean-field amplitudes of each component, where $i=\{1,2\}$. They obey the normalization condition $\int (|\psi_i|^2 + |\psi_j|^2) d^3x = 1$. The total population is $N$. The coupled mean-field equations Eq.~(\ref{main1}) can be solved numerically using a finite-difference Crank-Nicholson algorithm \cite{holland1}. We show results of such calculations in Fig.~1 for the three-dimensional solution and in Section~\ref{casestudy} for a one-dimensional version. However, in order to gain a more intuitive understanding of the behavior shown in Fig. 1, we formulate a simplified description of the system in the following section. \subsection{External $\otimes$ Internal Representation} \label{extint} The coupled mean-field equations Eq.~(\ref{main1}) can be rewritten in a more illuminating form by making a clear separation of the external and internal degrees of freedom. The system exists in a direct-product Hilbert space ${\mathcal H = H_{\rm{ex}}\otimes H_{\rm{in}}}$, where $\mathcal H_{\rm{ex}}$ is the infinite-dimensional Hilbert space describing the motional state of the system in the trap and $\mathcal H_{\rm{in}}$ is the two-dimensional Hilbert space describing the spin of the system. A general operator in $\mathcal H$ can be written as a sum over the direct-product of operators from $\mathcal H_{\rm{ex}}$ and $\mathcal H_{\rm{in}}$. We rewrite Eq.~(\ref{main1}) in this representation as \begin{equation} i {\partial\over\partial t} |\psi(t)\rangle = [\hat{H}_{0} \otimes \hat{1} + \hat{1} \otimes ({\Omega\over2} \, \hat{\sigma}_x + {\delta \, \over 2} \hat{\sigma}_z) + \hat{H}_{z} \otimes \hat{\sigma}_z] |\psi(t)\rangle \label{main2} \end{equation} where $\{\hat{1},\hat{\sigma}_x,\hat{\sigma}_y,\hat{\sigma}_z\}$ are the standard Pauli spin matrices. The state of the system $|\psi(t)\rangle$ in general has a nonzero projection on the internal states $|1\rangle$ and $|2\rangle$, represented by $\psi_i({\bf{r}},t)=\langle {\bf{r}} | \langle i |\psi(t)\rangle$, where $i = \{1,2\}$. The position representations of $\hat{H}_{0}$ and $\hat{H}_{z}$ are local, i.e. $\langle {\bf{r}}|\hat{H}_{0}|{\bf{r}}'\rangle = H_{0}({\bf{r}}) \, \delta({\bf{r}}-{\bf{r}}')$ and $\langle {\bf{r}}|\hat{H}_{z}|{\bf{r}}'\rangle = H_{z}({\bf{r}}) \, \delta({\bf{r}}-{\bf{r}}')$, where $H_{0}({\bf{r}})$ and $H_{z}({\bf{r}})$ are given by \begin{eqnarray} H_{0}({\bf{r}}) &=& -{1\over{2}} \nabla^2 + {1\over{2}}[({\rho\over\alpha})^2 +z^2] + \langle \psi(t)|\hat{P}_{\bf{r}} \otimes \hat{\lambda}_{+} |\psi(t)\rangle\,\, , \nonumber \\ H_{z}({\bf{r}}) &=& - z_0 \, z + \langle \psi(t)|\hat{P}_{\bf{r}} \otimes \hat{\lambda}_{-} |\psi(t)\rangle \,\, . \label{HSA} \end{eqnarray} The operator $\hat{P}_{\bf{r}}$ is the projector onto the position eigenstates $\hat{P}_{\bf{r}}=|\bf{r}\rangle \langle \bf{r} |$, and the matrix representations of $\hat{\lambda}_{+}$ and $\hat{\lambda}_{-}$ are given as \begin{eqnarray} \hat{\lambda}_{+} &=& {N \over 2} \left( \begin{array}{cc} \lambda_1 + \lambda_{12} & 0 \\ 0 & \lambda_2 + \lambda_{12} \end{array} \right) \,\, , \nonumber \end{eqnarray} \begin{eqnarray} \hat{\lambda}_{-} &=& {N \over 2} \left( \begin{array}{cc} \lambda_1 - \lambda_{12} & 0 \\ 0 & \lambda_{12} - \lambda_2 \end{array} \right) \,\, . \label{lambdaSA} \end{eqnarray} Note that the harmonic potential in $\hat{H}_0$ is centered at the origin. The mean-field interaction has been rewritten in terms of a part that acts identically on both components $\langle \psi|\hat{P}_{\bf{r}} \otimes \hat{\lambda}_{+} |\psi\rangle \otimes \hat{1}$, and a part that acts with the opposite sign on each state $\langle \psi|\hat{P}_{\bf{r}} \otimes \hat{\lambda}_{-} |\psi\rangle \otimes \hat{\sigma}_z$. The first two terms in Eq.~(\ref{main2}) separately describe the external and internal dynamics of the system, respectively. The third term in Eq.~(\ref{main2}), however, couples the internal state evolution to the condensate dynamics in the trap and can lead to interesting behavior. If the term $\hat{H}_{z}$ were identically zero, then the problem would be completely separable in terms of the external and internal degrees-of-freedom. The term $\hat{H}_{z}$ would be zero if the trap separation $z_0$ were zero and if the scattering lengths were all exactly equal. In fact, for ${}^{87}$Rb the three scattering lengths are nearly degenerate, so the main effect of $\hat{H}_{z}$ comes from the term $-z_0 z$, which is the difference in the shifted traps. It causes there to be a spatially varying detuning across the condensate. \subsection{Rotating Frame} \label{rotframe} As previously stated, we are concentrating on the situation where the coupling strength is large, so that the frequency of the Rabi oscillations $\Omega$ is significantly larger than the trap frequency $\nu_z$. In this case, the internal spin dynamics and the motion of the condensate in the trap occur on two different time scales. Therefore, it is useful to go to a rotating frame that eliminates the second term in Eq.~(\ref{main2}) describing the fast Rabi oscillations between the two internal states. In the rotating frame, we will be able to understand more clearly how the third term in Eq.~(\ref{main2}), which couples the motional and spin dynamics of the condensate, effects the system on a time scale much longer than the period of Rabi oscillation. We go to the rotating frame, or interaction picture, by making a unitary transformation using the operator \begin{equation} U_I(t) = e^{-i \, \hat{1}\otimes \, ({\Omega \over2} \hat{\sigma}_x + {\delta \over 2} \hat{\sigma}_z) \, t} \,\, . \label{U1} \end{equation} This can be rewritten in the equivalent form \begin{equation} U_I(t) = \hat{1}\otimes(\cos(\Omega_{\rm{eff}} / 2 \,t) \hat{1} - {i \over \Omega_{\rm{eff}}} \sin(\Omega_{\rm{eff}}/2 \, t)[\Omega \hat{\sigma}_x + \delta \hat{\sigma}_z]) \,\, , \label{U2} \end{equation} where $\Omega_{\rm{eff}}=\sqrt{\Omega^2 + \delta^2}$. The state vector $|\psi^{(I)}(t) \rangle$ in the rotating frame is related to the state vector in the lab frame $|\psi(t) \rangle$ by \begin{equation} |\psi^{(I)}(t) \rangle = U^\dagger_I \, |\psi(t) \rangle \,\, . \end{equation} In the rotating frame, the system evolves according to \begin{equation} i {\partial \over \partial t}|\psi^{(I)}(t) \rangle = \hat{H}^{(I)}(t) |\psi^{(I)}(t) \rangle \,\, , \label{intpic} \end{equation} where $\hat{H}^{(I)}(t)$ is the interaction Hamiltonian \begin{equation} \hat{H}^{(I)}(t)= \hat{H}_{0} \otimes \hat{1} + \hat{H}_{z} \otimes (\alpha_{\rm{x}}(t) \hat{\sigma}_{\rm{x}} + \alpha_{\rm{y}}(t) \hat{\sigma}_{\rm{y}} + \alpha_{\rm{z}}(t) \hat{\sigma}_{\rm{z}}) \,\, . \label{HofI} \end{equation} Note that $\hat{H}_{0}$ and $\hat{H}_{z}$ are unaffected by the unitary transformation to the rotating frame. The time-varying coefficients $\alpha_{\rm{x}}(t)$, $\alpha_{\rm{y}}(t)$, and $\alpha_{\rm{z}}(t)$ are \begin{eqnarray} \alpha_{\rm{x}}(t) &=& {\Omega \over \Omega_{\rm{eff}}} {\delta \over \Omega_{\rm{eff}}} [1 - \cos(\Omega_{\rm{eff}}t)] \nonumber \\ \alpha_{\rm{y}}(t) &=& {\Omega \over \Omega_{\rm{eff}}} \sin(\Omega_{\rm{eff}}t) \nonumber \\ \alpha_{\rm{z}}(t) &=& {\delta^2 \over \Omega_{\rm{eff}}^2} + {\Omega^2 \over \Omega_{\rm{eff}}^2} \cos(\Omega_{\rm{eff}}t) \,\,. \label{alphas} \end{eqnarray} \subsection{Approximations} \label{approx} We now make three simplifications in order to extract out the dominant behavior of the system. We first note that the coefficients $\alpha_i(t)$ given in Eq.~(\ref{alphas}) oscillate rapidly at the Rabi frequency. We expect the system in the rotating frame $|\psi^{(I)}(t) \rangle$ to evolve on a much slower time scale than the period of Rabi oscillation. We can utilize this fact in order to simplify the interaction Hamiltonian $\hat{H}^{(I)}(t)$ given in Eq.~(\ref{HofI}) by taking the average values of the coefficients $\alpha_i(t)$ -- this is equivalent to coarse graining Eq~(\ref{intpic}). The coefficients in Eq.~(\ref{alphas}) become time independent and reduce to: $\alpha_{\rm{x}}=\delta \, \Omega / \Omega_{\rm{eff}}^2$, $\alpha_{\rm{y}}=0$, and $\alpha_{\rm{z}}=\delta^2 / \Omega_{\rm{eff}}^2$. We make the additional assumption that the system is being driven close to resonance, so that $\delta/ \Omega_{\rm{eff}} << 1$. We therefore set $\alpha_{\rm{z}}=0$ since $\alpha_{\rm{z}}$ depends quadratically on this small parameter. Finally, we take advantage of the fact that the scattering lengths for ${}^{87}$Rb are nearly degenerate, with the ratios between inter- and intra-species scattering lengths given by $\{a_2:a_{12}:a_1\} =\{ 0.97:1:1.03\}$ \cite{exp2}. This allows us to simplify the two mean-field terms appearing in Eq.~(\ref{HSA}). We first make the approximation $\hat{\lambda}_{+} \approx \lambda\, N \cdot \hat{1}$ by assuming equal scattering lengths, or $\lambda_1 \approx \lambda_2 \approx \lambda_{12} \rightarrow \lambda$. We can also simplify the other term $\hat{\lambda}_{-}$ by assuming that its predominant effect is to shift the levels slightly. Instead of neglecting it altogether, we simply replace it by a mean-field shift of the levels $\langle \psi|\hat{P}_{\bf{r}} \otimes \hat{\lambda}_{-} |\psi\rangle \rightarrow \delta_{\rm{MF}}$, where the shift is given by $\delta_{\rm{MF}}= \Delta\lambda\int n^2({\bf{r}},0) d^3x / N$. Here $n({\bf{r}},0)$ is the total density at $t=0$ and $\Delta\lambda= (\lambda_1-\lambda_{12})=(\lambda_{12}-\lambda_2)$. We can absorb it into the detuning $\delta$ by defining an effective detuning $\delta'$ that includes this mean-field shift $\delta \rightarrow \delta'= \delta + \delta_{\rm{MF}}$. After making the above approximations, we can now write the interaction Hamiltonian $\hat{H}^{(I)}$ from Eq.~(\ref{HofI}) in a much simpler form \begin{equation} \hat{H}^{(I)} = \hat{H}_{0}' \otimes \hat{1} + \hat{H}_{z}' \otimes \hat{\sigma}_{\rm{x}} \,\, , \label{HofI2} \end{equation} where the position representations of $\hat{H}_{0}'$ and $\hat{H}_{z}'$ are local, i.e. $\langle {\bf{r}}|\hat{H}_{0}'|{\bf{r}}'\rangle = H_{0}'({\bf{r}}) \, \delta({\bf{r}}-{\bf{r}}')$ and $\langle {\bf{r}}|\hat{H}_{z}'|{\bf{r}}'\rangle = H_{z}'({\bf{r}}) \, \delta({\bf{r}}-{\bf{r}}')$, where $H_{0}'({\bf{r}})$ and $H_{z}'({\bf{r}})$ are given by \begin{eqnarray} \hat{H}_{0}'({\bf{r}}) &=& -{1\over{2}} \nabla^2 + {1\over{2}} [({\rho\over\alpha})^2 +z^2] + \lambda \, n({\bf{r}},t) \,\, , \nonumber \\ \hat{H}_{z}'({\bf{r}}) &=& - \beta \, z \,\, . \label{HSA2} \end{eqnarray} The total density is $n({\bf{r}},t)=N \,\langle \psi(t)|\hat{P}_{\bf{r}} \otimes \hat{1}|\psi(t)\rangle$, and $\beta = z_0 \, \delta \, \Omega / \Omega_{\rm{eff}}^2$. For the typical values of the parameters in the experiment, $\Omega = 400$ Hz, $\delta' = 100$ Hz, $z_0/z_{\rm{sho}} = 0.1$, this factor is rather small $\beta \approx 0.02$ in harmonic oscillator units. Note that $\hat{H}_{0}'$ still varies slowly in time through the nonlinear mean-field term, which depends on the density. We refer to this result Eq.~(\ref{HofI2}) as the coarse-grained, small detuning (CGSD) model to distinguish it from the two-mode model presented below, which makes further assumptions. We have managed to greatly simplify the description of the system by going to the rotating frame. The first term in Eq.~(\ref{HofI2}) contains the kinetic energy, a harmonic potential centered at the origin, and a mean-field interaction term depending on the slowly varying density $n({\bf{r}},t)$. The second term in Eq.~(\ref{HofI2}) represents a very weak coupling between the two internal states $|1\rangle$ and $|2\rangle$, and between motional states $|\phi_n\rangle$ and $|\phi_m\rangle$ via the dipole operator $z$. The states $|\phi_n\rangle$ and $|\phi_m\rangle$ are the instantaneous self-consistent eigenmodes of $\hat{H}_{0}'$. In the next subsection we present a model that assumes only two motional states are coupled, the self-consistent ground state $|\phi_0\rangle$ and the self-consistent first-excited state $|\phi_1\rangle$, which has odd-parity along the z-axis. \subsection{Two-mode model}\label{twomode} It is useful to define a basis of motional states with which to describe the system in the rotating frame. A natural choice is the set of instantaneous eigenstates of $\hat{H}_{0}'$, which satisfy \begin{eqnarray} ( -{1\over{2}} \nabla^2 + {1\over{2}} [({\rho\over\alpha})^2 +z^2] + \lambda \, n({\bf{r}},t)) \phi_i({\bf{r}}) &=& \epsilon_i \, \phi_i({\bf{r}}) \nonumber \\ \int_{-\infty}^{\infty} \phi_i({\bf{r}}) \, \phi_j({\bf{r}})\, d^3r &=& \delta_{i,j} \,\, , \label{selfcons} \end{eqnarray} where the index $i$ refers to all of the relevant quantum numbers that uniquely specify each eigenstate, $i=\{n_z,n_{\rho},n_{\phi}\}$, given the cylindrical symmetry of the system. In general, many modes can be occupied and the state vector is written \begin{equation} |\psi^{(I)}(t)\rangle = \sum_i[c_i(t)|\phi_i\rangle|1\rangle + d_i(t)|\phi_i\rangle |2\rangle] \,\, , \label{state_selfcons} \end{equation} where $\phi_i({\bf{r}})=\langle {\bf{r}} | \phi_i \rangle$. The density appearing in Eq.~(\ref{selfcons}) is then \begin{equation} n({\bf{r}},t) = N(| \sum_i c_i(t)\phi_i({\bf{r}})|^2 + | \sum_i d_i(t)\phi_i({\bf{r}})|^2) \,\, . \end{equation} It is clear that the set of coupled eigenvalue equations given in Eq.~(\ref{selfcons}) is nonlinear and requires a numerical procedure that will converge upon the solution in a self-consistent manner. The eigenstates $\phi_i({\bf{r}})$ and eigenenergies $\epsilon_i$ depend on time implicitly through the coefficients $c_i(t)$ and $d_i(t)$, however we do not show this time dependence in order to simplify the notation. We assume that the eigenbasis evolves slowly in time so that the adiabatic condition is satisfied \cite{Messiah}. Based on the experiment reported in \cite{exp5} the initial motional state of the system is $\psi_1^{(\rm{I})}({\bf{r}},0) = \phi_0({\bf{r}}-z_0 {\hat{z}})$; the system is in the ground state of $\hat{H}_{0}'$, but displaced from the origin along the vertical axis by $z_0$. This displacement is small compared to the width $w_z$ of the condensate $z_0/w_z \approx 0.01$. We therefore approximate the initial state of the system as $|\psi^{(I)}(t)\rangle =|\phi_0\rangle |1\rangle$. The system in the rotating frame evolves according to the Hamiltonian described by Eq.~(\ref{HofI2}) and Eq.~(\ref{HSA2}). The term $\hat{H}_{z}' \otimes \hat{\sigma}_{\rm{x}}$ couples the internal states $|1\rangle$ and $|2\rangle$ via ${\hat{\sigma}}_{\rm{x}}$. It also drives transitions between motional states via the dipole operator ${\hat{z}}$. The dipole matrix element $\langle z \rangle_{ij} = \langle \phi_i | {\hat{z}} | \phi_j \rangle$ is the largest between neighboring states and falls off quickly as $|i-j|$ increases. For a small coupling parameter $\beta$, we expect the coupling to the first excited state $|\phi_1\rangle$ to dominate the other transitions, making the evolution of the system predominantly a two state evolution. We therefore make the approximation that the system occupies only two modes \begin{equation} |\psi^{(I)}(t)\rangle = c_0(t)\,|\phi_0\rangle |1\rangle + d_1(t)\,|\phi_1\rangle |2\rangle \,\, , \label{twostate} \end{equation} where $|\phi_0\rangle$ is the ground state $i=\{0,0,0\}$ and $|\phi_1\rangle$ is the first excited state with odd parity along the z-axis $i=\{1,0,0\}$. If we substitute this ansatz into Eq.~(\ref{intpic}), using the Hamiltonian described by Eq.~(\ref{HofI2}) and Eq.~(\ref{HSA2}), we get the equation of motion for the coefficients $c_0(t)$ and $d_1(t)$ \begin{equation} \begin{array}{ccc} i \left( \! \! \begin{array}{c} \dot{c}_0 \\ \dot{d}_1 \end{array} \!\! \right) \!\!\! &=& \!\!\! \left( \!\! \begin{array}{cc} \epsilon_0 & -\beta \, \langle z \rangle_{01} \\ -\beta \, \langle z \rangle_{01} & \epsilon_1 \end{array} \! \right) \!\! \left( \!\! \begin{array}{c} c_0 \\ d_1 \end{array} \!\! \right) \end{array} \,\, , \label{twostateevolve} \end{equation} where we have neglected the time rate-of-change of the slowly varying adiabatic eigenbasis. This coupled pair of equations must be solved numerically by updating the energies $\epsilon_i$ and the dipole matrix element $\langle z \rangle_{01}$ from solving Eq.~(\ref{selfcons}) at each time step. However, in order to see how the behavior depends on the various physical parameters, one can obtain a simple estimate of the solution by fixing $\epsilon_i$ and $\langle z \rangle_{01}$ to their initial values. In this case the solution of Eq.~(\ref{twostateevolve}) is trivial and is given by $c_0(t) = \cos(\Omega_{01}/2 \, t) - i (\Delta \epsilon_{01} / \Omega_{01}) \sin(\Omega_{01}/2 \, t)$ and $d_1(t) = -i (2\beta \langle z \rangle / \Omega_{01}) \sin(\Omega_{01}/2 \, t)$, where $\Delta \epsilon_{01} = \epsilon_1-\epsilon_0$ and $\Omega_{01} = \sqrt{4\beta^2 \langle z \rangle^2 + \Delta \epsilon_{01}^2}$. In the rotating frame, the system oscillates between the two states at a frequency of $\Omega_{01}$, which is much slower than the effective Rabi frequency $\Omega_{\rm{eff}}$. The oscillation frequency $\Omega_{01}$ increases with increasing detuning $\delta'$ and increasing trap separation $z_0$ through the coupling parameter $\beta$. The amplitude of oscillation depends on the energy spacing between modes $\Delta \epsilon_{01}$. Based on numerical calculations, we have found that this effect is enhanced by the mean-field interaction because $\Delta \epsilon_{01}$ decreases with increasing population $N$. Also, the dipole matrix element $\langle z \rangle$ increases with increasing $N$, since the width of the condensate increases with increasing population. The solution in the lab frame can be obtained by applying $U_{\rm{I}}(t)$ from Eq~(\ref{U2}) to $|\psi^{\rm{(I)}}\rangle$ in Eq.~(\ref{twostate}) to yield \begin{eqnarray} |\psi(t)\rangle &=& (\alpha_1(t) c_0(t)\,|\phi_0\rangle + \alpha_2(t) d_1(t)\,|\phi_1\rangle) |1\rangle \nonumber \\ &+& (\alpha_2(t) c_0(t)\,|\phi_0\rangle + \alpha_1^*(t) d_1(t)\,|\phi_1\rangle) |2\rangle \,\, , \label{twostatelab} \end{eqnarray} where $\alpha_1(t) = \cos(\Omega_{\rm{eff}}/2 \,t) - i(\delta' / \Omega_{\rm{eff}}) \sin(\Omega_{\rm{eff}}/2 \,t)$ and $\alpha_2(t) = -i (\Omega / \Omega_{\rm{eff}}) \sin(\Omega_{\rm{eff}}/2 \,t)$. Eq.~(\ref{twostatelab}) is the main result of our study, with which we can explain the essential properties of the system. During the first few Rabi cycles $t \approx 1/\Omega_{\rm{eff}}$, the coefficient $d_1(t) \approx 0$, so that the solution for short times is $|\psi(t)\rangle = (\alpha_1(t)|1\rangle + \alpha_2(t)|2\rangle) \, |\phi_0\rangle$. That is, for short times, the internal and external degrees of freedom appear to be decoupled and the system simply oscillates rapidly between internal states. However, for longer times, the coefficient $d_1(t)$ grows in magnitude as $c_0(t)$ correspondingly decreases. This results in a modulation of the Rabi oscillations. Furthermore, a two-peaked structure in the density appears, associated with the first-excited state $|\phi_1 \rangle$. \section{Results}\label{casestudy} The main goal of this section is to illustrate the behavior of the system by showing results of numerical calculations. For this purpose, it is useful to treat the system in only one dimension---along the vertical axis \cite{williams1}. We also assume equal scattering lengths throughout this section, so that $\delta_{\rm{MF}}=0$. Values of most of the physical parameters are given in Table 1. Values of the remaining parameters are stated for each case considered in the text. \subsection{Understanding the dual dynamics} In Figure 2 we plot the fractional population of state $|1\rangle$, given by $N_1(t) = \int |\langle z | \langle 1|\psi(t)\rangle|^2 dz$, for the case of $\Omega = 700$ Hz and $\delta = 100$ Hz. This is a numerical solution of Eq.~(\ref{main1}). The population is cycling rapidly at the effective Rabi frequency $\Omega_{\rm{eff}}=707$ Hz, while simultaneously being modulated at a much lower frequency of about 11 Hz. In order to visualize how the spin and motional dynamics become entangled over a time long compared to the Rabi period, we show snapshots of the density of each state in Figure 3. Three different sets of snapshots are shown, corresponding to the three circled numbers in Figure 2. A full Rabi cycle is shown for each set. The first set begins at $t=0$ with all of the atoms in the $|1\rangle$ internal state and in the mean-field ground state of the trap $|\phi_0\rangle$. During this first Rabi cycle, the shape of the density profile for each internal state does not change much---only the height changes. That is, the motional state remains the ground state while population cycles rapidly between internal states, as discussed below Eq.~(\ref{twostatelab}). \begin{figure} \centerline{\epsfig{file=FIG_2.eps,height=2.7in}} \vspace{0.05cm} \caption{This plot shows an example of the modulation of the Rabi oscillations. The fractional population in state $|1\rangle$ is plotted as a function of time, obtained from a numerical solution of the one dimensional version of Eq.~(\ref{main1}). The values of the various parameters are given in the text. In Figure 3, the densities for both states are shown for three different Rabi cycles designated by the circled numbers in this plot.} \end{figure} \begin{figure} \centerline{\epsfig{file=FIG_3.eps,height=3.2in}} \vspace{0.05cm} \caption{This plot shows snapshots of the density of each state for three different Rabi cycles corresponding to the three circled numbers in Figure 2. The solid line is the density of the $|1\rangle$ state, while the dashed line is that of the $|2\rangle$ state. Each snapshot within a set is numbered in sequential order. The first set starts at $t = 0$ ms, and runs for a full Rabi cycle 1.41 ms. The second and third sets begin at $t = 45.2$ ms and $t = 90.3$ ms, respectively. The time increment between snapshots is $\Delta t= 0.28$ ms for all three sets.} \end{figure} \begin{figure} \centerline{\epsfig{file=FIG_4.eps,height=2.7in}} \vspace{0.05cm} \caption{The fractional population of the $|1\rangle$ state in the rotating frame is shown. The solid line is the solution given by the CGSD model, while the dot-dashed line corresponds to the solution of the two-mode model. If the two-mode model is extended to include coupling to the first even-parity excited mode, then we get better agreement to the CGSD model, as shown by the dashed line.} \end{figure} \begin{figure} \centerline{\epsfig{file=FIG_5.eps,height=2.2in}} \vspace{0.05cm} \caption{The top strip of this plot shows snapshots of the density of each state corresponding to the solution of the CGSD model given by the solid line in Figure 4. The bottom strip shows the corresponding two self-consistent eigenmodes given by the solution of Eq.~(\ref{selfcons}). The times of each snapshot are shown in the region between the two strips, in units of milliseconds. The solid line corresponds to the density of the $|1\rangle$ state, while the dashed is that of the $|2\rangle$ state.} \end{figure} The second set of snapshots in Figure 3 is taken at around $t=45$ ms, which is halfway through the modulation. The density profiles for each spin state cycle rapidly between a single-peaked and a double-peaked structure. For example, in the first snapshot, the $|1\rangle$ state is in the single-peaked structure, while the $|2\rangle$ state is in the double-peaked structure, but halfway through the Rabi cycle the situation is reversed, as shown in the third and fourth snapshots. Finally, at about $t=90$ ms when the amplitude of the Rabi oscillations has revived, the third set shows that the motional and spin degrees of freedom appear to be decoupled again, with the density profile of each spin state appearing as it did during the first Rabi cycle. As outlined in Section II, this peculiar behavior is most easily understood by going to the rotating frame. In Figure 4, we plot the fractional population in the $|1\rangle$ state in the rotating frame $N_1^{(I)}(t) = \int |\langle z | \langle 1|\psi^{(I)}(t)\rangle|^2 dz$. The solid line corresponds to the CGSD model presented in Section IID. In the rotating frame, population is slowly transferred out of the $|1\rangle$ state due to the coupling from ${\hat{H}}_z'\otimes {\hat{\sigma}}_x$ in Eq.~(\ref{HofI2}). In the rotating frame, the system is being excited out of the ground state $|\phi_0\rangle$ due to the dipole coupling $H_z'$. This can be seen in the top strip of snapshots in Figure 5, where the density of each spin state in the rotating frame is shown, corresponding to the solid line in Figure 4. Initially, all of the atoms are in the $|1\rangle$ internal state and the mean-field ground state of the trap $|\phi_0\rangle$. Due to the dipole coupling, population is transferred out of the ground state. The strongest coupling is between the ground $|\phi_0\rangle$ and the first excited $|\phi_1\rangle$ modes. These eigenmodes are shown in the bottom strip of Figure 5. They evolve slowly in time as the coefficients $c_0(t)$ and $d_1(t)$ change. For example, initially the ground state is just the Thomas-Fermi-like ground state, since all of the population is in that state. However, at $t=45$ ms, about one-third of the population is in the first excited mode, which pinches the ground state due to the mean-field interaction term in Eq.~(\ref{HSA2}). That is why the self-consistent ground state at $t=45$ ms is narrower than at $t=0$. It is clear from Figure 4 that the low-frequency modulation of the rapid Rabi oscillations in the lab frame is just the frequency of oscillation in the rotating frame between $|\phi_0\rangle|1\rangle$ and $|\phi_1\rangle |2\rangle$. This is reflected in the two-mode solution given by Eq.~(\ref{twostatelab}), which also helps explain the peculiar behavior of the densities shown in Figure 3. In the lab frame the system is cycling rapidly between the two modes shown in Figure 5. The initial values of the energies are $\epsilon_0=13.6 \, \hbar \omega_z$ and $\epsilon_1=13.7 \, \hbar \omega_z$, which makes $\Delta_{01}=0.1 \, \hbar \omega_z$. This small energy splitting is due to the effect of the mean field, since in the limit $N\rightarrow1$ these energies move apart by a factor of ten, which greatly reduces the coupling between the modes and thus greatly reduces the modulation effect. If we make the two-mode ansatz and solve Eq.~(\ref{twostateevolve}), we get the dot-dashed line in Figure 4. The discrepancy from the solid line arises due to a weak coupling between the first $|\phi_1 \rangle$ and second $|\phi_2\rangle$ excited modes. If we extend our two-state model to include this third mode, we get the dashed line in Figure 4, which nearly sits on top of the solid line. In this case, the second excited mode $|\phi_2\rangle$ gains less than $5 \%$ of the total population. \begin{figure} \centerline{\epsfig{file=FIG_6.eps,height=3.1in}} \vspace{0.05cm} \caption{This plot shows the fractional population in the $|1\rangle$ state for four different values of the detuning, obtained from a numerical solution of Eq~(\ref{main1}). Starting from the top, the detuning is $\delta = 0$, $\delta = 50$ Hz, $\delta = 100$ Hz, and $\delta = 200$ Hz. The values of the other parameters are given in the text.} \end{figure} \begin{figure} \centerline{\epsfig{file=FIG_7.eps,height=2.7in}} \vspace{0.05cm} \caption{This plot shows the period of modulation as a function of detuning $\delta$. The dashed line corresponds to the numerical solution of the one-dimensional version of Eq.~(\ref{main1}), while the solid line was obtained from a numerical solution of the two-mode model Eq.~(\ref{twostateevolve}). The Rabi frequency was $\Omega = 700$ Hz. } \end{figure} \subsection{Dependence on detuning} In Figure 6, we show how the behavior of the system depends on the detuning $\delta$. The Rabi frequency $\Omega=700$ Hz was held fixed for each plot while the detuning was varied from zero at the top $\delta = 0$ to $\delta = 200$ Hz in the bottom plot. As predicted by the coupling parameter $\beta =z_0 \, \delta \, \Omega / \Omega_{\rm{eff}}^2$ in the CGSD model, no coupling between motional states occurs if $\delta = 0$, and thus the Rabi oscillations experience no modulation. As $\delta$ is increased the motional-state coupling becomes stronger and we expect the modulation frequency to increase. The amplitude of modulation also increases as the detuning is increased. We show the dependence of the period of modulation on detuning more explicitly in Figure 7. The dashed line is the numerical solution of the full problem given by Eq.~(\ref{main1}), while the solid line is the numerical solution of the two-mode model given by Eq.~(\ref{twostateevolve}). \subsection{Dependence on trap displacement} In Figure 8, we show how the behavior of the system depends on the trap displacement $z_0$. The Rabi frequency $\Omega=700$ Hz and the detuning $\delta = 100$ Hz were held fixed, while the trap displacement was varied from zero $z_0=0$ in the top plot to $z_0 = 1 \, \mu$m in the bottom plot. Again, the coupling parameter $\beta$ predicts no modulation if $z_0=0$. As $z_0$ is increased, the frequency of modulation increases as the system is driven harder. However, for the large separation in the bottom plot, the modulation becomes highly irregular and the two-mode model most certainly breaks down. This behavior may be chaotic and warrants further investigation. \begin{figure} \centerline{\epsfig{file=FIG_8.eps,height=3.1in}} \vspace{0.05cm} \caption{This plot shows the fractional population in the $|1\rangle$ state for four different values of the trap displacement $z_0$, obtained from a numerical solution of Eq~(\ref{main1}). Starting from the top, the displacement is $z_0=0$, $z_0 = 0.1 \, \mu $m, $z_0 = 0.4 \, \mu $m, and $z_0 = 1.0 \, \mu $m. The values of the other parameters are given in the text.} \end{figure} \section{Conclusions}\label{conclusion} The gross features predicted by our model, such as double-peaked structure in the density distribution, and the presence of collapses and revivals in the relative population dynamics, are supported by experimental observation~\cite{exp5}. Experiment-theory agreement on finer points is only fair. The theory tends to underestimate the contrast ratio of the collapses, for instance. Moreover, to match the detuning trends shown in Fig. 6 and Fig. 7 one needs to add by hand an unexplained overall detuning offset. This is most likely due to there actually being a spatial dependence of the bare Rabi frequency due to the influence of gravity on the untrapped intermediate state of the two-photon transition. To model the experimental situation in more detail one would have to include this effect as well as inelastic loss processes and finite-temperature effects neglected here. It may be also that treating the TOP trap potential as purely static may be an oversimplication. In this paper we have demonstrated the possibility for quantum state engineering of topological excitations through the interplay between the internal and spatial degrees of freedom in a Bose condensed gas. Due to the symmetry of the system we have analyzed, the excitation in our case was the odd-parity dipole mode. The intriguing possibility of exciting modes with alternative symmetries, such as a vortex mode~\cite{Marzlin1,Walls,Zoller,Bagnato}, would require a different trap geometry, but is a straight-forward extension of the analysis presented here. Although we have focussed in this work on a particular parameter regime, the system is a rich one for study and exhibits complex and perhaps chaotic dynamics under strong excitation conditions. \section{Acknowledgments} We would like to thank Howard Carmichael for highlighting the analogies between this system and the bichromatically driven two-level atoms~\cite{howard}, and also Allan Griffin and Eugene Zaremba for insightful discussions. Finally, we would like to thank David Hall, Mike Matthews, and Paul Haljan for working in parallel on the experimental side of this project and for sharing the results of their observations in the laboratory~\cite{exp5}. This work was supported by the National Science Foundation. E.C. would also like to thank the Office of Naval Research and the National Institute for Standards and Technology for funding support. \bibliographystyle{prsty}
\section{Introduction} The Minimal Supersymmetric Standard Model (MSSM) currently provides the ``standard'' framework for the study of the physics beyond the Standard Model. The MSSM model takes into account the constraints from the negative searches so far for an experimental signature for Supersymmetry and gives circumstantial evidence for supersymmetric unified theories, such as the unification of the gauge couplings or the weak mixing angle prediction. The aforementioned constraints and predictions may also be fulfilled by other models, of string origin, and this would suggest that the MSSM is only a {\it minimal} candidate model for the physics beyond the Standard Model. Such string inspired models are regarded as ``low-energy'' limits of string theory, and they predict below the compactification scale a much richer spectrum (than that of the MSSM) including additional states at the high scale corresponding to predicted vector-like states\footnote{Vector-like states are not protected by the chiral symmetry of the SM and are therefore heavy \cite{georgi}.} under standard model gauge group. The presence of such additional states in conjunction with the low-energy input can lead to a phenomenology different from that of the MSSM. As a specific example which we address in this letter, consider the measurement of the strong coupling at the electroweak scale. The world average value \cite{particledata} of $\alpha_3(M_z)=0.119\pm 0.002$, situated rather close to the one loop value predicted in the MSSM is {\it below} the two-loop MSSM ``bottom-up'' prediction (of $\approx 0.125$) obtained from the RGE equations for the gauge couplings. While at one-loop level this prediction depends on the symmetry of the model and the multiplet content, at two-loop level a dependence of $\alpha_3(M_z)$ on the high scale thresholds is induced, and this may indicate that the mismatch between the MSSM two-loop prediction and the experiment is due to our lack of understanding of the physics at the high scale. For this reason at least, exploring the phenomenological implications of string theory {\it predicted} models with additional intermediate scales finds enough justification and it may also help in selecting and restricting the number of viable string inspired models. The exact structure of the spectrum that string-inspired models predict to exist (in addition to the MSSM spectrum\footnote{ In the following by MSSM spectrum we understand the three generations of quarks and leptons with their superpartners, an appropriate gauge sector containing the gauge bosons and gauginos and the two Higgs doublets and superpartners, {\it without} any $SU(3)$ Higgs triplets.}) depends in general on the particular class of models one considers to investigate. In reference \cite{witold1} the authors presented a class of extended supersymmetric models, as the low-energy limit of a string model with Calabi-Yau compactification \cite{116} and Wilson line breaking \cite{wilson} mechanism for the $E_6$ symmetry. In a generic example of this class of models, the ``low-energy'' spectrum below the compactification scale contained \cite{witold1} the MSSM spectrum plus (pairs of) complete five and ten-dimensional $SU(5)$ multiplets, ``vector-like'' under the Standard Model gauge group. The phenomenological implications of this case were discussed in \cite{grl,grlyuk} for the case when perturbation theory applies up to the unification scale. It was found that, due to a ``mixing'' between the heavy thresholds and the two-loop contributions of the vector-like states to the running of the gauge couplings, there is only a small (two-loop) increase in the unification scale from the MSSM prediction. The increase factor was $\approx 3$, too small to make an agreement with the weakly coupled heterotic string prediction \cite{scale} which gives an unification scale $\approx 20$ times larger than that of the MSSM. The aforementioned factor of increase was accompanied by a small (two-loop) increase from the MSSM prediction for the strong coupling at electroweak scale\footnote{For the case when the unification of the gauge couplings takes place in the non-perturbative regime, when the unified coupling is assumed to be large, one way to make predictions was presented in \cite{sun}, although the errors which affect them might be large as discussed in \cite{grlyuk} (Section 5).}. Reference \cite{witold1} also predicted another interesting possibility for the ``low-energy'' spectrum predicted by the same class of string-inspired models and this will be further analysed in this paper. This more specific model predicts not only complete five dimensional representations of $SU(5)$ in addition to the MSSM spectrum, but also a ``split multiplet'' structure, triplet under the $SU(3)$ group of the Standard Model. The model has therefore \cite{witold1} the nice feature of avoiding the ``doublet-triplet'' splitting problem which appears in the context of (Supersymmetric) Grand Unified Group Theories (GUT). In these theories, the Higgs multiplet content is a $5+{\overline 5}$ pair and consequently, the bare masses of the Higgs doublet and Higgs colour triplet have to be equal. To avoid proton decay mediated by the latter at a rate forbidden by the experimental constraints, the Higgs (colour) triplet has to be heavy enough to suppress such processes. In the meantime, the $SU(2)$ Higgs doublet must be light enough to explain the mass origin at the electroweak scale. This leads to the so-called ``doublet-triplet'' splitting problem, specific to supersymmetric GUT theories as well as to other theories which have such Higgs spectrum assignment. In addition to avoiding the ``doublet-triplet'' splitting problem, this last specific model provides \cite{witold1} the unification of the gauge couplings, even in the absence of a grand unified group such as $SU(5)$ or larger. The spectrum predicted by this string inspired model contains below the compactification scale\footnote{We refer here to the model which in \cite{witold1} was called the ``unconventional'' case.} a pair $3+{\overline 3}$ and an arbitrary number (say $n+1$) of extra pairs $5+{\overline 5}$ of $SU(5)$ {\it in addition} to the matter fields (three families) and the gauge sector of the Minimal Supersymmetric Standard Model, but without its Higgs content \footnote{Equivalently, we can say that the spectrum just below the compactification scale (and before the decoupling of any heavy state as we lower the scale) contains $n$ pairs $5+{\overline 5}$ and 2 pairs $3+{\overline 3}$ in addition to the MSSM spectrum.}. A coupling of the triplet of the ``incomplete'' $SU(5)$ representation, $3+{\overline 3}$, to the triplet component of a five dimensional representation of the form $\lambda {\tilde\phi} 3 {\bar 5}$ can naturally lead to a {\it large} mass (due to their vector-like character under SM group) term for the triplets, via a symmetry breaking mechanism when the (Standard Model singlet) Higgs field $\tilde\phi$ acquires a v.e.v., through a mechanism detailed in \cite{grl}. The same mechanism applies to the {\it extra} $n$ (vector-like) pairs of $5+{\overline 5}$ which also acquire a {\it large} mass (through couplings $\lambda {\tilde\phi} 5 {\bar 5}$ via the same mechanism \cite{grl} which provides a natural explanation for the origin of such mass terms). The {\it bare} value of this mass, assumed to be the same for all $5$'s is denoted by $\mu_g$ for further reference. The (remaining) doublet components of the initial pair $5+{\overline 5}$ are left uncoupled and thus light and can therefore account for the Higgs content of the MSSM. It is the purpose of this letter to examine in some detail the phenomenological implications of this model, in a two-loop analytical approach. The predictions we make refer to the value of $\alpha_3(M_z)$, the unification scale itself and the scale of the intermediate matter\footnote{We take as an input parameter the unified gauge coupling, to predict the intermediate scale and not vice-versa for the reason that the intermediate scale tends to have a flat behaviour for large range of values for $\alpha_g$, which can induce numerical instabilities of the solution, see Figures 3 and 4 of ref. \cite{grl} for a similar case.}, $\mu_g$. The reason for performing a two-loop analytical investigation of this model is three-fold; as mentioned, the discrepancy between the MSSM prediction for $\alpha_3(M_z)$ and the experiment arises mainly due to the two-loop corrections, and thus threshold dependence at the high scale plays a significant role. Moreover, any better candidate model than the MSSM should eliminate such discrepancy at this level of accuracy. Finally, the spectrum predicted by our model contains in addition to the $SU(3)$ triplet components, complete $SU(5)$ representations which are known to change the low energy prediction for $\alpha_3(M_z)$ at two-loop level \cite{grl} only\footnote{A similar situation exists for the unification scale prediction as well.}. The present approach also provides an analytical method to examine, in two loop order the RGE prediction for the strong coupling and may be applied to other (string-inspired) models with spectrum different from that of the MSSM. We show that $\alpha_3(M_z)$ can be reduced from the MSSM value (of $\approx 0.125$ or larger\footnote{The exact value of the strong coupling prediction in the MSSM depends on the assumptions made for the low energy (TeV scale) supersymmetric spectrum.}) and be brought within the experimental limits \cite{particledata} of $0.119\pm 0.002$, while keeping the unification scale $M_g$ close to that of the MSSM for an intermediate scale $\mu_g$ within one order of magnitude below $M_g$. \section{Predictions from the Renormalisation Group Evolution} The standard tool to exploring the phenomenological consequences of our model is the Renormalisation Group Evolution (RGE) for the gauge couplings, for which we take as (low-energy) boundary conditions the well known values of $\alpha_1(M_z)$ and $\alpha_2(M_z)$, obtained from measured electromagnetic coupling and weak mixing angle at the electroweak scale. To apply this tool we need to know the multiplet content, which was detailed in the Introduction, and the symmetry group, which is {\it just} the Standard Model gauge group, with $SU(5)$ normalisation for the $U(1)_Y$ coupling. Thus, the one-loop beta function before the decoupling of any extra (complete or incomplete) multiplet is given by \begin{equation} b^*_i=b_i+\Delta b_i+n \end{equation} with $b_i$ the MSSM one-loop beta function, $b_i=(33/5,1,-3)$, and with $\Delta b_i=4 \times\{1/5, 0, 1/2\}$ to account for two pairs of triplets or equivalently four triplet states, hence the factor 4 in the definition of $\Delta b_i$. After the decoupling of all additional states, the one-loop beta function is just that of the MSSM, namely $b_i$. With some loss of generality we restrict ourselves to the case when the extra states (triplets and 5-plets) have the same {\it bare} mass which we called $\mu_g$ and this assumption does not reintroduce the ``doublet-triplet'' problem. The general case of considering different masses for $3+\bar 3$ and $5+\bar 5$ pairs can be done following the present approach, although introducing one further mass parameter would make the analysis less tractable. To evaluate the full two-loop ``running'' of the gauge couplings, including the effects of the heavy thresholds, we use the integral form of the ``NSVZ beta function''. This has been computed in \cite{novikov} and \cite{murayama1} (see also \cite{shifman2}) and is given by \begin{equation} \beta (\alpha )^{NSVZ}\equiv \frac{d\alpha }{d(\ln \mu )}=-\frac{\alpha ^{2}% }{2\pi }\left[ 3T(G)-\sum_{\sigma }^{{}}T(R_{\sigma })(1-\gamma _{\sigma }^{NSVZ})\right] \left( {1-T(G)\frac{\alpha }{2\pi }}\right) ^{-1} \label{shifmanbetaalpha} \end{equation} with the definition ($\mu $ is the running scale) \[ \gamma _{\sigma }^{NSVZ}=-\frac{d\ln Z_{\sigma }}{d\ln \mu } \] and where $T(G)$ and $T(R_{\sigma })$ represent the Dynkin index for the adjoint representation and for $R_{\sigma }$ representation respectively (not necessarily the fundamental one). The above sum runs over {\it all} matter fields $\sigma $ in representation $R_{\sigma }$ and this includes the extra heavy states in addition to the low energy spectrum of the MSSM. Following the details given in \cite{grlyuk} to integrate the beta function given above, we find that, to all orders in perturbation theory, the gauge couplings run, in the presence of the extra matter, as follows \begin{eqnarray} \label{intalph} \alpha _{i}^{-1}(M_{z}) &=&-\delta _{i}+\alpha _{g}^{-1}+\frac{b_{i}}{2\pi } \ln \frac{M_{g}}{M_{z}}+\frac{n+\Delta b_i} {2\pi }\ln \frac{M_{g}}{\mu _{g}}-\frac{ \beta _{i,H_{1}}}{2\pi }\ln Z_{H_{1}}(M_{z})-\frac{\beta _{i,H_{2}}}{2\pi } \ln Z_{H_{2}}(M_{z}) \nonumber \\ &&-\frac{\beta _{i,g}}{2\pi }\ln \left[ \frac{\alpha _{g}}{\alpha _{i}(M_{z} }\right] ^{1/3}-\sum_{j=1}^{3}\sum_{\phi _{j}}{}\frac{\beta _{i,\phi _{j}}} 2\pi }\ln Z_{\phi _{j}}(M_{z}) \end{eqnarray} where $b_{1}=33/5$, $\,b_{2}=1$, $\,b_{3}=-3$ and where $\beta _{i,\phi _{j}}\equiv T(R_{\phi _{j}}^{i})$, $i=\{1,2,3\}$, are the contributions to one-loop beta function\footnote{% We also used that the one-loop beta function is $b=-3 T(G)+ \sum T(R_\phi)$, where the sum runs over all chiral supermultiplets in representation $R_\phi$% .} of the matter fields $\phi _{j}$ (j=generation index), while $\beta _{i,g}\equiv T^{i}(G)$ is the one-loop beta function for the pure gauge (+gaugino) sector; the Higgs (+higgsino) sector contribution is included separately via the terms proportional to $\beta _{i,H1,2}$; finally, $\alpha_g$ is the unified coupling while $M_g$ stands for the unification scale of our model. We have \begin{equation} \beta _{i,\phi _{j}}=\left( \begin{array}{ccccc} \frac{3}{10} & \frac{1}{10} & \frac{3}{5} & \frac{4}{5} & \frac{1}{5} \\ & & & & \\ \frac{1}{2} & \frac{3}{2} & 0 & 0 & 0 \\ & & & & \\ \ 0 & 1 & 0 & \frac{1}{2} & \frac{1}{2} \end{array} \right) _{i,\phi _{j}}\;\;\;\;\;\;\;\beta _{i,g}=\left( \begin{array}{c} 0 \\ \\ -6 \\ \\ -9 \end{array} \right) ;\;\;\;\;\;\;\;\beta _{i,H_{1,2}}=\left( \begin{array}{c} \frac{3}{10} \\ \\ \frac{1}{2} \\ \\ 0 \end{array} \right) \end{equation} with $\beta _{i,\phi _{j}}$ independent of the values of $j$. The field $\phi _{j}$ runs over the set $% \phi _{j}=\{l_{L},q_{L},e_{R},u_{R},d_{R}\}_{j}$, in this order, with $j$ as generation index. The coefficients $\delta_i$ in eq.(\ref{intalph}) represent the low energy supersymmetric thresholds and they would be equal to zero if supersymmetry were valid at the electroweak scale. Their exact expressions will not be of concern to us as we will present our phenomenological predictions as a change to the MSSM predictions\footnote{ The MSSM quantities used as an input in our calculation will be the unified coupling, the unification scale and the value of the strong coupling at electroweak scale.} which implicitly contain the dependence on $\delta_i$. ($\delta_i$ also contain conversion scheme factors (${\overline {MS}}\rightarrow {\overline {DR}}$) which in our calculation will cancel against those of the MSSM of equal value). The uncertainty in the low-energy supersymmetric spectrum (i.e. the value of $\delta_i$) can be taken into account in our present approach by allowing in our final results, a {\it range} of values for the MSSM variables which are present in our final expressions. To compute the two-loop running for the gauge couplings, only a one-loop expression for the wavefunction renormalisation coefficients is required. Note that in the two-loop approximation there is no regularisation ambiguity which arises only in three-loop order \cite{jones}. At $M_z$ scale the one-loop expressions for $Z$'s have the following structure \begin{eqnarray} \label{zzss} Z_{F}(M_z)&=&\prod_{k=1}^{3} \left[\frac{\alpha_g}{\alpha_k(\mu_g)} \right]^{-\frac{2 C_k(F)}{b^*_k}} \left[\frac{\alpha_k(\mu_g)}{\alpha_k(M_z)}\right]^{-\frac{2 C_k(F)}{b_k}} \nonumber\\ &=&\prod_{k=1}^{3} \left[\frac{\alpha_g}{\alpha_k(\mu_g)} \right]^{\frac{2 C_k(F)}{b_k}\frac{n+\Delta b_k}{b^*_k}} \left[\frac{\alpha_g}{\alpha_k(M_z)}\right]^{-\frac{2 C_k(F)}{b_k}} \end{eqnarray} where $F$ stands for any Higgs or MSSM chiral field. Strictly speaking in the expressions of $Z$ factors we should have used the mean mass of the extra states ${\tilde \mu}$ instead of $\mu_g$; however this difference is an additional radiative effect and thus is of two-loop order for $Z$'s or of three loop order for the gauge couplings, and can be neglected in our two loop calculation. From equations (\ref{intalph}) and (\ref{zzss}) we find the following RGE equations \begin{eqnarray} \label{HHMSSM} &\alpha_i^{-1}(M_z)=&-\delta_i+\alpha_g^{-1}+ \frac{b_i}{2\pi}\ln \left[\frac{% M_g}{M_z}\right]+\frac{n+\Delta b_i}{2\pi} \ln\left[\frac{M_g}{\mu_g} \right]-\frac{1}{% 2\pi}\sum_{j=1}^{3}{\tilde Y}_{ij}\ln \left[\frac {\alpha_g}{\alpha_j(\mu_g)}\right] \nonumber \\ & &+\frac{1}{4\pi}\sum_{j=1}^{3} \frac{b_{ij}}{b_j} \ln\left[\frac{\alpha_g}{% \alpha_j(M_z)}\right] \end{eqnarray} where \begin{equation} \label{whywhy} {\tilde Y}_{ij}=\frac{n+\Delta b_j}{b^*_j} \left[\frac{1}{2}\frac{b_{ij}}{b_j}-\delta_{ij}\lambda_j\right] \end{equation} with $\lambda_1=0,\,\lambda_2=2, \,\lambda_3=3$ while $\delta_i$, $i=\{1,2,3\}$, stand for the low-energy (TeV scale) supersymmetric thresholds. From eq.(\ref{HHMSSM}) we can again see that the presence of $\mu_g$ in the two-loop term $\ln(\alpha_g/\alpha(\mu_g))$ instead of the mean physical mass ${\tilde\mu}$ of the additional multiplets we consider would account for an additional (three-loop) radiative effect and we neglect it, as it is beyond our two-loop approximation for the running of the gauge couplings. Thus, we can say that in two-loop order the extra states contribute to the gauge couplings running through their common bare mass only. At this point we would like to emphasize that the result of equation (\ref{HHMSSM}) can also be obtained using the ``standard'' RGE equations, integrated in two-loop order, with appropriate taking into account of the heavy thresholds that the additional states we consider bring in. Using a radiative dressing of the masses of the additional states and following the approach of ref. \cite{grl} one will obtain the same result. As it was the case there, there is a cancellation of the heavy thresholds against the two-loop contributions of the additional states we considered and as a consequence in eq.(\ref{HHMSSM}) only the two-loop MSSM beta function appears, and {\it not} that in the presence of additional states\footnote{Also note the strong similarity of eq.(\ref{whywhy}) with that of eq.(7) of ref.\cite{grl}.} $5+{\overline 5}$ or $3+{\overline 3}$. Eq.(\ref{HHMSSM}) is actually more general; if one considers vector-like matter in addition to the MSSM sector, with some arbitrary $\delta b_j$ contribution to one-loop beta function, the two-loop RGE equations have a form similar to that of eqs.(\ref{HHMSSM}) and (\ref{whywhy}) with the replacements $n+\Delta b_i\rightarrow \delta b_i$ and $b_j^*\rightarrow b_j+\delta b_j$. To compute the unification scale $M_g$, the strong coupling $\alpha_3(M_z)$ and the value of the mass scale $\mu_g$ we must make some assumptions about the low energy supersymmetric spectrum which affects the running of the gauge couplings through the terms $\delta_i$, as seen from eq.(\ref{HHMSSM}). Since the effects of the low energy supersymmetric thresholds on the predictions of the MSSM are relatively known \cite{lan}, we prefer to express our predictions as a change to the MSSM predictions which all have this dependence included (and assume that $\delta_i$'s have equal values to those of the MSSM). We therefore consider the two-loop running of the gauge couplings in the MSSM, which is of the form (the MSSM variables are labelled with an ``o'' index to distinguish them from those of our extended model) \begin{equation} \alpha _{i}^{o-1}(M_{z})=-\delta _{i}+\alpha _{g}^{o-1}+\frac{b_{i}}{% 2\pi }\ln \left[ \frac{M_{g}^{o}}{M_{Z}}\right] +\frac{1}{4\pi }% \sum_{j=1}^{3}\frac{b_{ij}}{b_{j}}\ln \left[ \frac{\alpha _{g}^{o}}{\alpha _{j}^{o}(M_{Z})}\right] \label{MSSM} \end{equation} We can then substitute\footnote{In the MSSM we have $\alpha_g^o\approx 0.0433$, $M_g^o\approx 3\times 10^{16}$ GeV and $\alpha_3^o(M_z)=0.125$ or larger. These results are obtained in two loop approximation, using $\alpha_1(M_z)$ and $\alpha_2(M_z)$ as an input from experimental values for electromagnetic coupling and weak mixing angle.} the values of $\delta_i$ from the above equation into eq.(\ref{HHMSSM}) and impose that in both the MSSM and our model, the values of $\alpha_1(M_z)$ and $\alpha_2(M_z)$ are taken equal with the corresponding experimental value, so $\alpha_1^o(M_z)=\alpha_1(M_z)$ and $\alpha_2^o(M_z)=\alpha_2(M_z)$. We then compute the analytical expressions for the factor of increase of the unification scale $M_g/M^o_g$, the strong coupling $\alpha_3(M_z)$ (in terms of $\alpha_3^o(M_z)$), and the bare mass of the extra states, $\mu_g$. This can be done following the approach of \cite{grl} and using that, under two-loop terms we can substitute the arguments of the ``log'' terms by their one loop values, as the difference would be of higher order. This means that $\ln(\alpha_g/\alpha_j(\mu_g))=\ln(1+b_j^*\alpha_g/(2\pi) \ln(M_g/\mu_g))$ and we further replace $\ln(M_g/\mu_g)$ by its one loop analytical expression, which is correct for a two loop running for the gauge couplings. After some tedious algebra we find the following two-loop analytical results \begin{eqnarray} \label{ores1} \frac{M_g}{M_g^o}&=& exp\left\{\frac{2\pi}{7n-1}\left(\alpha_g^{-1}-\alpha_g^{o-1}\right)\right\} \left[\frac{\alpha_g}{\alpha_g^o} \right]^{{\cal E}_1} \left\{1+\frac{18\,\alpha_3^o(M_z)}{1-7n} \left(\alpha_g^{-1}-\alpha_g^{o-1}\right)\right\}^{{\cal E}_2} \nonumber\\ &&\times\prod_{j=1}^{3} \left\{1+\frac{7b^*_j}{1-7n} \left[1-\frac{\alpha_g}{\alpha_g^o}\right]\right\}^{{\cal D}_j} \end{eqnarray} \begin{eqnarray} \label{ores4} \alpha_3^{-1} (M_z)&=&\alpha_3^{o-1}(M_z)+ \frac{18(\alpha_g^{-1}-\alpha_g^{o-1})}{1-7n} +\frac{73+17n}{2(7n-1)\pi}\ln\left\{1+\frac{18\,\alpha_3^o(M_z)}{1-7n} \left(\alpha_g^{-1}-\alpha_g^{o-1}\right)\right\}\nonumber\\ &&+\frac{2(467+235n)}{11\pi(1-7n)}\ln\left[\frac{\alpha_g}{\alpha_g^o}\right] + \sum_{j=1}^{3} \ln\left\{1+\frac{7b^*_j}{1-7n} \left[1-\frac{\alpha_g}{\alpha_g^o}\right]\right\}^{{\cal A}_j} \end{eqnarray} \begin{eqnarray} \label{ores2} \frac{\mu_g}{M_g^o}&=& exp\left\{\frac{16\pi}{7n-1}\left(\alpha_g^{-1}-\alpha_g^{o-1}\right)\right\} \left[\frac{\alpha_g}{\alpha_g^o} \right]^{{\cal E}_3} \left\{1+\frac{18\,\alpha_3^o(M_z)}{1-7n} \left(\alpha_g^{-1}-\alpha_g^{o-1}\right)\right\}^{{\cal E}_4} \nonumber\\ &&\times\prod_{j=1}^{3}\left\{1+\frac{7b^*_j}{1-7n} \left[1-\frac{\alpha_g}{\alpha_g^o}\right]\right\}^{{\cal H}_j} \end{eqnarray} where \begin{equation} {\cal E}_1={\frac{285+341n}{33(7n-1)}},\,\,\,\,\,\,\,\, {\cal E}_2=\frac{4(3+n)}{3(1-7n)},\,\,\,\,\,\,\,\, {\cal E}_3=\frac{2621+341n}{33(7n-1)},\,\,\,\,\,\,\,\, {\cal E}_4=\frac{100+4n}{3(1-7n)} \end{equation} \begin{equation} {\cal D}_j=\left\{\frac{(18-77n)(-4-5n)}{660(7n-1)b^*_1}, \frac{-3n(14+13n)}{4(7n-1)b^*_2}, \frac{4(2+n)(3+n)}{3(7n-1)b^*_3}\right\}_j \end{equation} \begin{equation} {\cal A}_j=\left\{\frac{(-29+56n)(-4-5n)}{220(7n-1)\pi b^*_1}, \frac{81n(5+2n)}{4(7n-1)\pi b^*_2}, \frac{2(2+n)(-19+n)}{(7n-1)\pi b^*_3}\right\}_j \end{equation} \begin{equation} {\cal H}_j=\left\{\frac{-268-27n+385n^2}{660(7n-1)b^*_1}, \frac{-3n(125+13n)}{4(7n-1)b^*_2}, \frac{4(50+27n+n^2)}{3(7n-1)b^*_3}\right\}_j \end{equation} The above analytical solution to eq.(\ref{HHMSSM}) agrees well with the numerical one. To find a numerical solution we just solved numerically the system of three equations obtained from subtracting eq.(\ref{MSSM}) from (\ref{HHMSSM}) to eliminate the $\delta_i$'s and also replaced $\ln(\alpha_g/\alpha_j(\mu_g))$ by $\ln(\alpha_g/\alpha_j(\mu_g))=\ln(1+b_j^*\alpha_g/(2\pi) \ln(M_g/\mu_g))$. The agreement between the two approaches is good, within less than $1\%$ relative error for $\alpha_3(M_z)$, $5\%$ relative error for the factor $M_g/M_g^o$, and $10\%$ relative error for $\mu_g/M_g^o$. The larger error exists when the coupling $\alpha_g$ is larger and is also due to the presence of the logarithmic dependence the terms involving $M_g$ and $\mu_g$ come with in the RGE equations. \section{Numerical Results} In this section we analyse the results and the phenomenological implications of eqs.(\ref{ores1}),(\ref{ores4}) and (\ref{ores2}). \begin{figure}[tbh] \begin{center} \parbox{8cm} {\psfig{figure=mgmgo.ps,height=8 cm,width=9cm}} \end{center} \caption{ The values of $M_g/M_g^o$ plotted in function of the ratio $\alpha_g/\alpha_g^o$ for different values of $n$.} \label{fig6} \end{figure} \begin{figure}[tbhp] \begin{center} \parbox{8cm} {\psfig{figure=alfap.ps,height=8cm,width=9cm}} \end{center} \caption{ The values of $\alpha_3(M_z)$ plotted in function of the ratio $\alpha_g/\alpha_g^o$ for different values of $n$.} \label{fig7} \end{figure} \begin{figure}[tbhp] \begin{center} \parbox{8cm} {\psfig{figure=logmiugmgo.ps,height=8 cm,width=9cm}} \end{center} \caption{ The values of $Log_{10}[\mu_g/M_g^o]$ plotted in function of the ratio $\alpha_g/\alpha_g^o$ for different values of $n$.} \label{fig8} \end{figure} Figure 1 shows the ratio of the unification scales $M_g/M_g^o$ for different values of $n$ in function of the ratio $\alpha_g/\alpha_g^o$. We observe that this ratio is less than unity for most of the parameter space and the effect of extra states we added does not bring the unification scale closer to the weakly coupled heterotic string scale which is a factor of $\approx 20$ above the MSSM value. However, for {\it large} $n$ the ratio $M_g/M_g^o$ is very close to unity, and therefore the change induced in this case from the MSSM scale, is very small. (Note that the perturbative calculation is valid for large $n$, as long as $n\alpha\approx \kappa {\cal O}(4\pi)$, with $\kappa <1$). For $n\geq 20$ we find $M_g$ above $0.8M_g^o$ for most values of $\alpha_g$, and therefore the change induced by the extra matter to the MSSM unification scale is insignificant. This is a result of the presence of two competing effects, the reducing of the scale (at one-loop level) due to the $SU(3)$ triplet states and the opposite effect of increasing the scale due to the complete five dimensional multiplets. Such opposite effects are also manifest in the predicted value of $\alpha_3(M_z)$. In Figure 2 we presented this value for different $n$ in function of the ratio $\alpha_g/\alpha_g^o$. We observe that we can accommodate values of $\alpha_3(M_z)$ smaller than in the MSSM and in better agreement with the experimental value \cite{particledata} $\alpha_3(M_z)_{exp}=0.119\pm 0.002$, provided that the value of the unified coupling is marginally increased from the MSSM value, for the case of small $n$. For $n\geq 20$ $\alpha_3(M_z)$ is within the experimental limits for a larger range of values of $\alpha_g$. The effect of reducing the strong coupling is essentially due to the presence of the colour triplets we considered. The result is somewhat expected as, unlike models which include complete $SU(5)$ representations to the MSSM spectrum and where complete representations introduced the same term $n\ln(M_g/\mu_g)$ in the running of the gauge couplings \cite{grl}, the situation here is different because the similar contribution is now $(n+\Delta b_i)\ln(M_g/\mu_g)$, with $\Delta b_i$ standing for the triplets' contribution (see eqs.(\ref{intalph}), (\ref{HHMSSM})). This means that the {\it relative} behaviour of the gauge couplings running is already changed at one-loop order from the MSSM prediction due to the presence of the $SU(3)$ triplets while the complete five dimensional multiplets bring a two-loop additional increasing effect \cite{grl}. We would like to note that the input MSSM value for $\alpha_3^o(M_z)$ considered here was $0.125$; this represents the lower limit prediction of a ``bottom-up'' approach in the MSSM case, and therefore the predictions we made for $\alpha_3(M_z)$ could increase slightly if the input for $\alpha_3^o(M_z)$ is above this value. Figure 3 shows the ratio $\mu_g/M_g^o$ in terms of $\alpha_g$ and this determines the value of one of these when the other is fixed. For the parameter space with good predictions for $\alpha_3(M_z)$ we find that the (bare) intermediate scale is in the region of $3\times 10^{15}$ GeV, only a factor of $\approx 10$ below the standard MSSM unification scale. The large value for the intermediate scale avoids an enhancement of the proton decay rate by the colour triplet states. \section{Conclusions} In this work we have considered the phenomenological implications of a string-motivated model, which predicts below the compactification scale the existence of $n$ extra pairs $5+{\overline 5}$ of $SU(5)$ states and 2 pairs of $SU(3)$ triplets in addition to the MSSM spectrum. The motivation for studying this model originates in the suggestion that this might also solve the ``doublet-triplet'' splitting problem, commonly faced by Grand Unified Group-based theories. The strong coupling at the electroweak scale can be reduced below the value of the two-loop MSSM prediction and be brought into better agreement with the experiment while the value of the unification scale, in two-loop order, remains close to the MSSM prediction. \section{Acknowledgements} The author thanks Graham Ross and Witold Pokorski for useful discussions. D.G. acknowledges the financial support from Oxford University (Leverhulme Trust research grant).
\section{Introduction} Numerical computations are still very important in computer applications. But until recently there was a discrepancy between numerical methods and software/hardware tools for scientific calculations. In particular, numerical programming was not much influenced by the progress in Mathematics, programming languages and technology. Modern tools for numerical calculations are not unified, standardized and reliable enough. It is difficult to ensure the necessary accuracy and safety of calculations without loss of the efficiency and speed of data processing. It is difficult to get correct and exact estimations of calculation errors. For example, standard methods of interval arithmetic \cite{2} do not allow to take into account the error auto-correction effects \cite{19} and, as a result, to estimate calculation errors accurately. However, new ideas in Mathematics and Computer Science lead to a very promising approach (initially presented in \cite{20}--\cite{22}). An essential aspect of this approach is developing a system of algorithms, utilities and programs based on a new mathematical calculus which is called {\it Idempotent Analysis} or {\it Idempotent Calculus}, or {\it Idempotent Mathematics} etc. For many problems in optimization and mathematical modeling this Idempotent Calculus plays the same unifying role as Functional Analysis in Mathematical Physics, see, e.g., \cite{14}, \cite{17}, \cite{28}--\cite{30} and surveys \cite{15}, \cite{21}. Idempotent Analysis is based on replacing the usual arithmetic operations by a new set of basic operations (such as maximum or minimum). There are a lot of such new arithmetics which are associated with sufficiently rich algebraic structures called idempotent semirings. It is very important that many problems, nonlinear in the usual sense, become linear with respect to an appropriate new arithmetic, i.e. linear over a suitable semiring (the so-called {\it idempotent superposition principle} \cite{26}, \cite{27}, \cite{17}, which is a natural analog of the well-known superposition principle in Quantum Mechanics). This `linearity' considerably simplifies explicit constructions of their solutions. Examples are the Bellman equation and its generalizations, the Hamilton--Jacobi equation etc. The idempotent analysis is a powerful heuristic tool to construct new algorithms and apply unexpected analogies and ideas borrowed, e.g., from mathematical physics and quantum mechanics. The abstract theory is well advanced and includes, in particular, a new integration theory, linear algebra and spectral theory, idempotent functional analysis, idempotent Fourier transforms and so on. Its applications include various optimization problems such as multi-criteria decision making, optimization on graphs, discrete optimization with a large parameter (asymptotic problems), optimal design of computer systems and computer media, optimal organization of parallel data processing, dynamic programming, applications to differential equations, numerical analysis, discrete event systems, computer science, discrete mathematics, mathematical logic, etc. (see, e.g. \cite{1}, \cite{3}, \cite{5}--\cite{15}, \cite{17}, \cite{21}, \cite{26}--\cite{32} and references therein). It is possible to obtain an implementation of the new approach to scientific and numeric calculations in the form of a powerful software system based on unified algorithms. This approach ensures the arbitrary necessary accuracy and safety of numerical calculations and a working time reduction for programmers and users because of a software unification. Our approach uses the techniques of object oriented and functional programming (see, for example, \cite{25}, \cite{16}) which is very convenient for the design of our (suggested) software system. A computer algebra techniques \cite{4} is also used. The modern techniques of systolic processors and VLSI realizations of numerical algorithms including parallel algorithms of linear algebra (see, for example, \cite{18}, \cite{31}) is convenient for effective implementations of the proposed approach to hardware design. There is a regular method based on the theory for constructing back-end processors and technical devices intended for a realization of basic algorithms of idempotent calculus and mathematics of semirings. These hardware facilities can increase the speed of data processing. \section{Mathematical objects and their computer representations} Numerical algorithms are combinations of basic operations. Usually these basic operations deal with `numbers'. In fact these `numbers' are thought of as members of some numerical {\it domains} (real numbers, integers etc.). But every computer calculation deals with finite {\it models} (or finite {\it computer representations}) of these numerical domains. For example, integers can be modeled by integers modulo a power of number 2, real numbers can be represented by rational numbers or floating-point numbers etc. Discrepancies between mathematical objects (e.g. `ideal' numbers) and their computer models (representations) lead to calculation errors. Due to imprecision of sources of input data in real-world problems, the data usually come in the form of confidence intervals or other number sets rather than exact quantities. Interval Analysis (see, e.g.,~\cite{2}) extends operations of traditional calculus from numbers to number intervals to make possible processing such imprecise data and controlling rounding errors in computational mathematics. However, there are no universal models which are good in all cases and we have to use varieties of computer models. For example, real numbers can be represented by the following computer models: standard floating-point numbers, double precision floating-point numbers, arbitrary precision floating-point numbers, rational numbers, finite precision rational numbers, floating-slash and fixed-slash rational numbers, interval numbers, etc. To examine an algorithm it is often useful to have a possibility to change computer representations of input/output data. For this aim the corresponding algorithm (and its software implementation) must be universal enough. \section{Universal algorithms} It is very important that many algorithms do not depend on particular models of a numerical domain and even on this domain itself. Algorithms of linear algebra (matrix multiplication, Gauss elimination etc.) are good examples of algorithms of this type. Of course, one algorithm may be more universal than another algorithm of the same type. For example, numerical integration algorithms based on the Gauss--Jacobi quadrature formulas actually depend on computer models because they use finite precision constants. On the contrary, the rectangular formula and the trapezoid rule do not depend on models and in principle can be used even in the case of idempotent integration (see below). The so-called object oriented software tools and programming languages (like $C^{++}$ and Java, see, e.g., \cite{25}) are very convenient for computer implementation of universal algorithms. In fact there are no reasons to restrict ourselves with numerical domains only. Actually it may be a ring of polynomials, a field of rational functions, or an idempotent semiring. The case of idempotent semirings is extremely important because of numerous applications. \section{Idempotent correspondence principle} There is a nontrivial analogy between Mathematics of semirings and Quantum Mechanics. For example, the field of real numbers can be treated as a `quantum object' with respect to idempotent semirings. So idempotent semirings can be treated as `classical' or `semi-classical' objects with respect to the field of real numbers. Let $\Bbb{R}$ be the field of real numbers and $\Bbb{R}_+$ the subset of all non-negative numbers. Consider the following change of variables: $$ u \mapsto w = h \ln u, $$ where $u \in \Bbb{R}_+ \setminus \{0\}$, $h > 0$; thus $u = e^{w/h}$, $w \in \Bbb{R}$. Denote by $\bf{0}$ the additional element $-\infty$ and by $S$ the extended real line $\Bbb{R} \cup \{\bf{0}\}$. The above change of variables has a natural extension $D_h$ to the whole $S$ by $D_h(0) = \bf{0}$; also, we denote $D_h(1) = 0 = \bf{1}$. Denote by $S_h$ the set $S$ equipped with the two operations $\oplus_h$ (generalized addition) and $\odot_h$ (generalized multiplication) such that $D_h$ is a homomorphism of $\{\Bbb{R}_+, +, \cdot\}$ to $\{S, \oplus_h, \odot_h\}$. This means that $D_h(u_1 + u_2) = D_h(u_1) \oplus_h D_h(u_2)$ and $D_h(u_1 \cdot u_2) = D_h(u_1) \odot_h D_h(u_2)$, i.e., $w_1 \odot_h w_2 = w_1 + w_2$ and $w_1 \oplus_h w_2 = h \ln (e^{w_1/h} + e^{w_2/h})$. It is easy to prove that $w_1 \oplus_h w_2 \to \max\{w_1, w_2\}$ as $h \to 0$. Denote by $\rset_{\max}$ the set $S = \Bbb{R} \cup \{\bf{0}\}$ equipped with operations $\oplus = \max$ and $\odot = +$, where ${\bf{0}} = -\infty$, ${\bf{1}} = 0$ as above. Algebraic structures in $\Bbb{R}_+$ and $S_h$ are isomorphic; therefore $\rset_{\max}$ is a result of a deformation of the structure in $\Bbb{R}_+$. We stress the obvious analogy with the quantization procedure, where $h$ is the analog of the Planck constant. In these terms, $\Bbb{R}_+$ (or $\Bbb{R}$) plays the part of a `quantum object' while $\rset_{\max}$ acts as a `classical' or `semi-classical' object that arises as the result of a {\it dequantization} of this quantum object. Likewise, denote by $\rset_{\min}$ the set $\Bbb{R} \cup \{\bf{0}\}$ equipped with operations $\oplus = \min$ and $\odot = +$, where ${\bf{0}} = +\infty$ and ${\bf{1}} = 0$. Clearly, the corresponding dequantization procedure is generated by the change of variables $u \mapsto w = -h \ln u$. Consider also the set $\Bbb{R} \cup \{\bf{0}, \bf{1}\}$, where ${\bf{0}} = -\infty$, ${\bf{1}} =+\infty$, together with the operations $\oplus = \max$ and $\odot=\min$. Obviously, it can be obtained as a result of a `second dequantization' of $\Bbb{R}$ or $\Bbb{R}_+$. The algebras presented in this section are the most important examples of idempotent semirings, the central algebraic structure of Idempotent Analysis. Consider a set $S$ equipped with two algebraic operations: {\it addition} $\oplus$ and {\it multiplication} $\odot$. The triple $\{S, \oplus, \odot\}$ is an {\it idempotent semiring} if it satisfies the following conditions (here and below, the symbol $\star$ denotes any of the two operations $\oplus$, $\odot$): \begin{itemize} \item the addition $\oplus$ and the multiplication $\odot$ are associative: $ x \star (y \star z) = (x \star y) \star z$ for all $x, y, z\in S$; \item the addition $\oplus$ is commutative: $x \oplus y = y \oplus x$ for all $x,y \in S$; \item the addition $\oplus$ is {\it idempotent}: $x \oplus x = x$ for all $x\in S$; \item the multiplication $\odot$ is {\it distributive} with respect to the addition $\oplus$: $x\odot(y\oplus z) = (x\odot y)\oplus(x\odot z)$ and $(x\oplus y)\odot z = (x\odot z)\oplus(y\odot z)$ for all $x, y, z\in S$. \end{itemize} A {\it unity} of a semiring $S$ is an element ${\bf{1}}\in S$ such that for all $x \in S$ $$ {\bf{1}} \odot x = x \odot {\bf{1}} = x. $$ A {\it zero} of a semiring $S$ is an element ${\bf{0}} \in S$ such that $\bf{0} \neq \bf{1}$ and for all $x \in S$ $$ {\bf{0}}\oplus x = x,\qquad {\bf{0}}\odot x = x\odot {\bf{0}} = {\bf{0}}. $$ A semiring $S$ is said to be {\it commutative} if $x\odot y=y\odot x$ for all $x,y\in S$. A commutative semiring is called a {\it semifield} if every nonzero element of this semiring is invertible. It is clear that $\Bbb R_{\max}$ and $\Bbb R_{\min}$ are semifields. Note that different versions of this axiomatics are used, see, e.g., \cite{1}, \cite{3}, \cite{5}, \cite{6}, \cite{12}, \cite{13}--\cite{15}, \cite{17}, \cite{21}, \cite{23}, \cite{30} and some literature indicated in these books and papers. Many nontrivial examples of idempotent semirings can be found, e.g., in \cite{1}, \cite{5}, \cite{6}, \cite{12}, \cite{13}, \cite{14}, \cite{17}, \cite{21}, \cite{23}, \cite{24}, \cite{30}. For example, every vector lattice or ordered group can be treated as an idempotent semifield. The addition $\oplus$ defines the following {\it standard partial order} on $S$: $x\preceq y$ if and only if $x\oplus y=y$. If $S$ contains a zero $\bf{0}$, then ${\bf{0}}\preceq x$ for all $x\in S$. The operations $\oplus$ and $\odot$ are consistent with this order in the following sense: if $x\preceq y$, then $x\star z\preceq y\star z$ and $z\star x\preceq z\star y$ for all $x,y,z \in S$. The basic object of the traditional calculus is a {\it function} defined on some set $X$ and taking its values in the field $\Bbb{R}$ (or $\Bbb{C}$); its idempotent analog is a map $X \to S$, where $X$ is some set and $S =\rset_{\min}$, $\rset_{\max}$, or another idempotent semiring. Let us show that redefinition of basic constructions of traditional calculus in terms of Idempotent Mathematics can yield interesting and nontrivial results (see, e.g., \cite{17}, \cite{21}, \cite{23}, \cite{24}, for details, additional examples and generalizations). {\sc Example 1. Integration and measures.} To define an idempotent analog of the Riemann integral, consider a Riemann sum for a function $\varphi(x)$, $x \in X = [a,b]$, and substitute semiring operations $\oplus$ and $\odot$ for operations $+$ and $\cdot$ (usual addition and multiplication) in its expression (for the sake of being definite, consider the semiring $\rset_{\max}$): $$ \sum_{i = 1}^N \varphi(x_i) \cdot \Delta_i \quad\mapsto\quad \bigoplus_{i = 1}^N \varphi(x_i) \odot \Delta_i = \max_{i = 1, \ldots, N}\, (\varphi(x_i) + \Delta_i), $$ where $a = x_0 < x_1 < \cdots < x_N = b$, $\Delta_i = x_i - x_{i - 1}$, $i = 1,\ldots,N$. As $\max_i \Delta_i \to 0$, the integral sum tends to $$ \int_X^\oplus \varphi(x)\, dx = \sup_{x \in X} \varphi(x) $$ for any function $\varphi$:~$X \to \rset_{\max}$ that is bounded. In general, for any set $X$ the set function $$ m_\varphi(B) = \sup_{x \in B} \varphi(x), \quad B \subset X, $$ is called an $\rset_{\max}$-{\it measure} on $X$; since $m_\varphi(\bigcup_\alpha B_\alpha) = \sup_\alpha m_\varphi(B_\alpha)$, this measure is completely additive. An idempotent integral with respect to this measure is defined as $$ \int_X^\oplus \psi(x)\, dm_\varphi = \int_X^\oplus \psi(x) \odot \varphi(x)\, dx = \sup_{x \in X}\, (\psi(x) + \varphi(x)). $$ Using the standard partial order it is possible to generalize these definitions for the case of arbitrary idempotent semirings. {\sc Example 2. Fourier--Legendre transform.} Consider the topological group $G = \Bbb{R}^n$. The usual Fourier--Laplace transform is defined as $$ \varphi(x) \mapsto \widetilde\varphi(\xi) = \int_G e^{i\xi \cdot x} \varphi(x)\, dx, $$ where $\exp(i\xi \cdot x)$ is a {\it character} of the group $G$, i.e., a solution of the following functional equation: $$ f(x + y) = f(x)f(y). $$ The idempotent analog of this equation is $$ f(x + y) = f(x) \odot f(y) = f(x) + f(y). $$ Hence `idempotent characters' of the group $G$ are linear functions of the form $x \mapsto \xi \cdot x = \xi_1 x_1 + \cdots + \xi_n x_n$. Thus the Fourier--Laplace transform turns into $$ \varphi(x) \mapsto \widetilde\varphi(\xi) = \int_G^\oplus \xi \cdot x \odot \varphi(x)\, dx = \sup_{x \in G}\, (\xi \cdot x + \varphi(x)). $$ This is the well-known Legendre (or Fenchel) transform. These examples suggest the following formulation of the idempotent correspondence principle \cite{20}, \cite{21}: \begin{quote} {\it There exists a heuristic correspondence between interesting, useful, and important constructions and results over the field of real (or complex) numbers and similar constructions and results over idempotent semirings in the spirit of N. Bohr's correspondence principle in Quantum Mechanics.} \end{quote} So Idempotent Mathematics can be treated as a `classical shadow (or counterpart)' of the traditional Mathematics over fields. In particular, an idempotent version of Interval Analysis can be constructed \cite{24}. The idempotent interval arithmetics appear to be remarkably simpler than their traditional analog. For example, in the traditional interval arithmetic multiplication of intervals is not distributive with respect to interval addition, while idempotent interval arithmetics conserve distributivity. Idempotent interval arithmetics are useful for reliable computing. \section{ Idempotent linearity} Let $S$ be a commutative idempotent semiring. The following example of a noncommutative idempotent semiring is very important. {\sc Example 3.} Let $\rm{Mat}_n(S)$ be a set of all $S$-valued matrices, i.e. coefficients of these matrices are elements of $S$. Define the sum $\oplus$ of matrices $A = \|a_{ij}\|$, $B = \|b_{ij}\| \in \rm{Mat}_n(S)$ as $A \oplus B = \|a_{ij} \oplus b_{ij}\| \in \rm{Mat}_n(S)$. The {\it product} of two matrices $A \in \rm{Mat}_n(S)$ and $B \in \rm{Mat}_n(S)$ is a matrix $AB \in \rm{Mat}_n(S)$ such that $AB = \|\bigoplus_{k = 1}^m a_{ik} \odot b_{kj}\|$. The set $\rm{Mat}_n(S)$ of square matrices is an idempotent semiring with respect to these operations. If $\bf{0}$ is the zero of $S$, then the matrix $O = \|o_{ij}\|$, where $o_{ij} = \bf{0}$, is the zero of $\rm{Mat}_n(S)$; if $\bf{1}$ is the unity of $S$, then the matrix $E = \|\delta_{ij}\|$, where $\delta_{ij} = \bf{1}$ if $i = j$ and $\delta_{ij} = \bf{0}$ otherwise, is the unity of $\rm{Mat}_n(S)$. Now we discuss an idempotent analog of a linear space. A set $V$ is called a {\it semimodule} over $S$ (or an $S$-semimodule) if it is equipped with an idempotent commutative associative addition operation $\oplus_V$ and a multiplication $\odot_V$:~$S \times V \to V$ satisfying the following conditions: for all $\lambda$, $\mu \in S$, $v$, $w \in V$ \begin{itemize} \item $(\lambda \odot \mu) \odot_V v = \lambda \odot_V (\mu \odot_V v)$; \item $\lambda \odot_V (v \oplus_V w) = (\lambda \odot_V v) \oplus_V (\lambda \odot_V w)$; \item $(\lambda \oplus \mu) \odot_V v = (\lambda \odot_V v) \oplus_V (\mu \odot_V v)$. \end{itemize} An $S$-semimodule $V$ is called a {\it semimodule with zero} if ${\bf{0}} \in S$ and there exists a {\it zero} element ${\bf{0}}_V \in V$ such that for all $v \in V$, $\lambda \in S$ \begin{itemize} \item ${\bf{0}}_V \oplus_V v = v$; \item $\lambda \odot_V {\bf{0}}_V = {\bf{0}} \odot_V v = {\bf{0}}_V$. \end{itemize} {\sc Example 4. Finitely generated free semimodule.} The simplest $S$-semimodule is the direct product $S^n = \{\, (a_1, \ldots, a_n) \mid a_j \in S, j = 1, \ldots, n \,\}$. The set of all endomorphisms $S^n \to S^n$ coincides with the semiring $\rm{Mat}_n(S)$ of all $S$-valued matrices (see example~3). The theory of $S$-valued matrices is similar to the well-known Perron--Fro\-be\-ni\-us theory of nonnegative matrices, well advanced and has very many applications, see, e.g., \cite{1}, \cite{3}, \cite{5}--\cite{15}, \cite{17}, \cite{21}, \cite{24}, \cite{29}, \cite{30}--\cite{32}). \medskip {\sc Example 5. Function spaces.} An {\it idempotent function space} ${\cal{F}}(X;S)$ consists of functional defined on a set $X$ and taking their values in an idempotent semiring $S$. It is a subset of the set of all maps $X \to S$ such that if $f(x)$, $g(x) \in {\cal{F}}(X;S)$ and $c \in S$, then $(f \oplus g)(x) = f(x) \oplus g(x) \in {\cal{F}}(X;S)$ and $(c \odot f)(x) = c \odot f(x) \in {\cal{F}}(X;S)$; in other words, an idempotent function space is another example of an $S$-semimodule. If the semiring $S$ contains a zero element $\bf{0}$ and ${\cal{F}}(X;S)$ contains the zero constant function $o(x) = \bf{0}$, then the function space ${\cal{F}}(X;S)$ has the structure of a semimodule with zero $o(x)$ over the semiring $S$. If the set $X$ is finite we get the previous example. Recall that the idempotent addition defines a standard partial order in $S$. An important example of an idempotent functional space is the space ${\cal{B}}(X;S)$ of all functions $X \to S$ bounded from above with respect to the partial order $\preceq$ in $S$. There are many interesting spaces of this type including ${\cal{C}}(X;S)$ (a space of continuous functions defined on a topological space $X$), analogs of the Sobolev spaces, etc (see, e.g., \cite{17}, \cite{21}, \cite{23}, \cite{28}--\cite{30} for details). According to the correspondence principle, many important concepts, ideas and results can be converted from usual Functional Analysis to Idempotent Analysis. For example, an idempotent scalar product in ${\cal{B}}(X;S)$ can be defined by the formula $$ \langle\varphi,\psi\rangle = \int_X^\oplus \varphi(x) \odot \psi(x)\, dx, $$ where the integral is defined as the `$\sup$' operation (see example 1). \medskip {\sc Example 6. Integral operators.} It is natural to construct idempotent analogs of integral operators of the form $$ K:\, \varphi(y) \mapsto (K\varphi)(x) = \int_Y^\oplus K(x,y) \odot \varphi(y)\, dy, $$ where $\varphi(y)$ is an element of a functional space ${\cal{F}}_1(Y;S)$, $(K\varphi)(x)$ belongs to a space ${\cal{F}}_2(X;S)$ and $K(x,y)$ is a function $X \times Y \to S$. Such operators are {\it linear}, i.e. they are homomorphisms of the corresponding functional semimodules. If $S = \rset_{\max}$, then this definition turns into the formula $$ (K\varphi)(x) = \sup_{y \in Y}\, (K(x,y) + \varphi(y)). $$ Formulas of this type are standard for optimization problems. \section{Superposition principle} \quad\ In Quantum Mechanics the superposition principle means that the Schr\"odi\-n\-ger equation (which is basic for the theory) is linear. Similarly in Idempotent Mathematics the idempotent superposition principle means that some important and basic problems and equations (e.g., optimization problems, the Bellman equation and its versions and generalizations, the Hamilton-Jacobi equation) nonlinear in the usual sense can be treated as linear over appropriate idempotent semirings, see \cite{26}--\cite{30}, \cite{17}. The linearity of the Hamilton-Jacobi equation over $\Bbb{R}_{\min}$ (and $\Bbb{R}_{\max}$) can be deduced from the usual linearity (over $\Bbb{C}$) of the corresponding Schr\"odinger equation by means of the dequantization procedure described above (in Section 4). In this case the parameter $h$ of this dequantization coincides with $i\hbar$ , where $\hbar$ is the Planck constant; so in this case $\hbar$ must take imaginary values (because $h>0$; see \cite{23} for details). Of course, this is closely related to variational principles of mechanics. The situation is similar for the differential Bellman equation, see \cite{17}. It is well-known that discrete versions of the Bellman equation can be treated as linear over appropriate idempotent semirings. The so-called {\it generalized stationary} (finite dimensional) {\it Bellman equation} has the form $$ X = AX \oplus B, $$ where $X$, $A$, $B$ are matrices with elements from an idempotent semiring and the corresponding matrix operations are described in example 3 above; the matrices $A$ and $B$ are given (specified) and it is necessary to determine $X$ from the equation. B.A. Carr\'e \cite{5} used the idempotent linear algebra to show that different optimization problems for finite graphs can be formulated in a unified manner and reduced to solving these Bellman equations, i.e., systems of linear algebraic equations over idempotent semirings. For example, Bellman's method of solving shortest path problems corresponds to a version of the Jacobi method for solving systems of linear equations, whereas Ford's algorithm corresponds to a version of the Gauss-Seidel method. \section{Correspondence principle for computations} Of course, the idempotent correspondence principle is valid for algorithms as well as for their software and hardware implementations \cite{20}--\cite{22}. Thus: \begin{quote} {\it If we have an important and interesting numerical algorithm, then there is a good chance that its semiring analogs are important and interesting as well.} \end{quote} In particular, according to the superposition principle, analogs of linear algebra algorithms are especially important. Note that numerical algorithms for standard infinite-dimensional linear problems over idempotent semirings (i.e., for problems related to idempotent integration, integral operators and transformations, the Hamilton-Jacobi and generalized Bellman equations) deal with the corresponding finite-dimensional (or finite) `linear approximations'. Nonlinear algorithms often can be approximated by linear ones. Thus the idempotent linear algebra is a basis for the idempotent numerical analysis. Moreover, it is well-known that linear algebra algorithms are convenient for parallel computations; their idempotent analogs admit parallelization as well. Thus we obtain a systematic way of applying parallel computation to optimization problems. Basic algorithms of linear algebra (such as inner product of two vectors, matrix addition and multiplication, etc.) often do not depend on concrete semirings, as well as on the nature of domains containing the elements of vectors and matrices. Algorithms to construct the closure $A^*={\bf{1}}\oplus A\oplus A^2\oplus\cdots\oplus A^n\oplus\cdots= \bigoplus^{\infty}_{n=1} A^n$ of an idempotent matrix $A$ can be derived from standard methods for calculating $({\bf{1}} -A)^{-1}$. For the Gauss--Jordan elimination method (via LU-decomposition) this trick was used in \cite{31}, and the corresponding algorithm is universal and can be applied both to the Bellman equation and to computing the inverse of a real (or complex) matrix $({\bf{1}} - A)$. Computation of $A^{-1}$ can be derived from this universal algorithm with some obvious cosmetic transformations. Thus it seems reasonable to develop universal algorithms that can deal equally well with initial data of different domains sharing the same basic structure \cite{21}, \cite{22}. \section{Correspondence principle for hardware design} A systematic application of the correspondence principle to computer calculations leads to a unifying approach to software and hardware design. The most important and standard numerical algorithms have many hardware realizations in the form of technical devices or special processors. {\it These devices often can be used as prototypes for new hardware units generated by substitution of the usual arithmetic operations for its semiring analogs and by addition tools for performing neutral elements $\bf{0}$ and} $\bf{1}$ (the latter usually is not difficult). Of course, the case of numerical semirings consisting of real numbers (maybe except neutral elements) is the most simple and natural \cite{20}--\cite{22}. Note that for semifields (including $\Bbb R_{\max}$ and $\Bbb R_{\min}$) the operation of division is also defined. Good and efficient technical ideas and decisions can be transposed from prototypes into new hardware units. Thus the correspondence principle generated a regular heuristic method for hardware design. Note that to get a patent it is necessary to present the so-called `invention formula', that is to indicate a prototype for the suggested device and the difference between these devices. Consider (as a typical example) the most popular and important algorithm of computing the scalar product of two vectors: \begin{equation} (x,y)=x_1y_1+x_2y_2+\cdots + x_ny_n. \end{equation} The universal version of (1) for any semiring $A$ is obvious: \begin{equation} (x,y)=(x_1\odot y_1)\oplus(x_2\odot y_2)\oplus\cdots\oplus (x_n\odot y_n). \end{equation} In the case $A=\Bbb R_{\max}$ this formula turns into the following one: \begin{equation} (x,y)=\max\{ x_1+y_1,x_2+y_2, \cdots, x_n+y_n\}. \end{equation} This calculation is standard for many optimization algorithms, so it is useful to construct a hardware unit for computing (3). There are many different devices (and patents) for computing (1) and every such device can be used as a prototype to construct a new device for computing (3) and even (2). Many processors for matrix multiplication and for other algorithms of linear algebra are based on computing scalar products and on the corresponding `elementary' devices respectively, etc. There are some methods to make these new devices more universal than their prototypes. There is a modest collection of possible operations for standard numerical semirings: max, min, and the usual arithmetic operations. So, it is easy to construct programmable hardware processors with variable basic operations. Using modern technologies it is possible to construct cheap special-purpose multi-processor chips implementing examined algorithms. The so-called systolic processors are especially convenient for this purpose. A systolic array is a `homogeneous' computing medium consisting of elementary processors, where the general scheme and processor connections are simple and regular. Every elementary processor pumps data in and out performing elementary operations in a such way that the corresponding data flow is kept up in the computing medium; there is an analogy with the blood circulation and this is a reason for the term `systolic', see e.g. \cite{18}, \cite{31}. Of course, hardware implementations for important and popular basic algorithms can increase the speed of data processing. \section{ Correspondence principle for software design} Software implementations for universal semiring algorithms are not so efficient as hardware ones (with respect to the computation speed) but are much more flexible. Program modules can deal with abstract (and variable) operations and data types. Concrete values for these operations and data types can be defined by the corresponding input data. In this case concrete operations and data types are generated by means of additional program modules. For programs written in this manner it is convenient to use a special techniques of the so-called object oriented (and functional) design, see, e.g., \cite{25}, \cite{16}. Fortunately, powerful tools supporting the object-oriented software design have recently appeared including compilers for real and convenient programming languages (e.g. $C^{++}$ and Java). There is a project to obtain an implementation of the correspondence principle approach to scientific calculations in the form of a powerful software system based on a collection of universal algorithms. This approach ensures a working time reduction for programmers and users because of the software unification. The arbitrary necessary accuracy and safety of numeric calculations can be ensured as well. The system contains several levels (including programmer and user levels) and many modules. Roughly speaking, it is divided into three parts. The first part contains modules that implement domain modules (finite representations of basic mathematical objects). The second part implements universal (invariant) calculation methods. The third part contains modules implementing model dependent algorithms. These modules may be used in user programs written in $C^{++}$ and Java. \centerline{The following modules and algorithms implementations are in progress:} \medskip \centerline{Domain modules:} \smallskip infinite precision integers; rational numbers; finite precision rational numbers; finite precision complex rational numbers; fixed- and floating-slash rational numbers; complex rational numbers; arbitrary precision floating-point real numbers; arbitrary precision complex numbers; p-adic numbers; interval numbers; ring of polynomials over different rings; idempotent semirings $R(\max, \min)$, $R(\max, +)$, $R(\min, +)$, interval idem- potent semirings and others. \centerline{Algorithms:} linear algebra; numerical integration; roots of polynomials; spline interpolations and approximations; rational and polynomial interpolations and approximations; special functions calculation; differential equations; optimization and optimal control; idempotent functional analysis and others. This software system may be especially useful for designers of algorithms, software engineers, students and mathematicians.
\section{Introduction} \setcounter{equation}{0} The averaged null energy condition (ANEC, for short) has attracted some interest during the past several years as a possible candidate for a stability condition in semiclassical gravity. In its simplest form, this condition requires that in quantum field theory (on any spacetime manifold) the integral of the expectation value, $\langle T_{\mu\nu} \rangle$, of the energy-momentum tensor in any physical state, along any complete, lightlike geodesic $\gamma} \def\G{\Gamma$ is always non-negative: $$ \int_{-\infty}^{\infty} \langle T_{\mu\nu}(\gamma} \def\G{\Gamma(s))\rangle k^{\mu}k^{\nu}\,ds \ge 0\,,$$ where $s$ is an affine parameter and $k^{\mu}$ the (parallelly propagated) tangent of $\gamma} \def\G{\Gamma$. (For a formulation not requiring the existence of the integral, see below.) We shall briefly indicate the origin and development of this condition, however, we are not attempting to properly review this area of research and refer the reader to the articles \cite{FlaWa,Y1,Y2,FR,WalYu} for further discussion and additional references. In the theory of classical gravity, one central object of study is the behaviour of solutions to Einstein's equations, $$ G_{\mu \nu}(x) = 8\pi T_{\mu\nu}(x)\,,$$ for classical matter described by the energy-momentum tensor $T_{\mu\nu}$. There are important results asserting that a certain qualitative behaviour of these solutions must necessarily occur, within a broad range of initial conditions, as soon as certain stability requirements are imposed on $T_{\mu\nu}$. It is significant that such qualitative behaviour typically reflects a stability of causality, i.e.\ an initially causally well-behaved spacetime will not end up to develop, e.g., closed timelike curves. Most prominent among those results are the singularity theorems \cite{HE,WaldI}; the typical stability requirements in this context are the null energy condition, \begin{equation} T_{\mu\nu}(x)k^{\mu}k^{\nu} \ge 0 \end{equation} for all lightlike vectors $k^{\mu}$ at any point $x$ in spacetime, or the weak energy condition, where (1.1) is to hold for all causal vectors $k^{\mu}$ at any point $x$, and related variants, like the strong energy condition or the dominant energy condition, cf.\ \cite{HE,WaldI}. The common feature of these conditions is that they impose a local (even pointlike) positivity constraint like in eq.\ (1.1) on the energy-momentum tensor. For energy-momentum tensors of phenomenological models for classical matter, such local positivity constraints have largely been found to be physically realistic. In contrast, it is known that, under very general hypotheses, similar local positivity constraints cannot hold for the expectation values of the energy-momentum tensor $\langle \psi,T_{\mu\nu}(x)\psi \rangle$ of a quantum field in Minkowski-spacetime for a dense set of state vectors $\psi$ \cite{EGJ}. Now, in semiclassical gravity, one investigates the semiclassical Einstein equation \begin{equation} G_{\mu \nu}(x) = 8\pi\langle T_{\mu\nu}(x)\rangle \end{equation} where $\langle T_{\mu\nu}(x)\rangle$ is the expectation value of the energy-momentum tensor in a physical state of a quantum field propagating in a classical background spacetime whose Einstein tensor is $G_{\mu \nu}$. The question arises if there is a realistic replacement for the local positivity constraints on $\langle T_{\mu\nu}(x)\rangle$ leading to similar implications, i.e.\ the necessity of a certain, causally stable behaviour of solutions to (1.2) to occur. And in fact, candidates for such replacements have been found. In \cite{Tip} it was observed that nonlocal, ``averaged'' versions of the local positivity constraints on the classical energy-momentum tensor still lead to essentially the same singularity theorems which result from imposing local positivity constraints (see also \cite{ChiEh,Borde,Rom} for discussion and further results). The ``averaged'' refers to integrating the energy-momentum tensor along causal geodesics. The condition used in \cite{Tip} is that $$ \int_{-\infty}^{\infty} \left(T_{\mu\nu} - \mbox{$ \frac{1}{2}$} g_{\mu \nu}T^{\sigma}{}_{\sigma}\right)\!\!(\gamma} \def\G{\Gamma (s)) \, k^{\mu}k^{\nu}\,ds \ge 0 $$ for any complete causal geodesic with affine parameter $s$ and tangent $k^{\mu}$; $g_{\mu \nu}$ is the spacetime metric. This is referred to as averaged strong energy condition. In \cite{Rom} it was shown that an averaged null energy condition for certain half-complete geodesics, i.e.\ essentially \begin{equation} \liminf_{r \to \infty}\, \int_0^{r}T_{\mu\nu}(\gamma} \def\G{\Gamma (s))\,k^{\mu}k^{\nu}\,ds \ge 0 \end{equation} for all lightlike geodesics $\gamma} \def\G{\Gamma$ with affine parameter $s$ and tangent $k^{\mu}$ emanating at $s =0$ from a closed trapped surface, implies singularity theorems. Moreover, it is proved in \cite{Borde} that singularity theorems are implied by ANEC, roughly, \begin{equation} \liminf_{r_{\pm} \to \infty}\,\int_{-r_-}^{r_+} T_{\mu\nu}(\gamma} \def\G{\Gamma (s))\,k^{\mu}k^{\nu}\,ds \ge 0 \end{equation} for all complete lightlike geodesics with affine parameter $s$ and tangent $k^{\mu}$. (The precise formulations in \cite{Rom} and \cite{Borde} are slightly different from ours in (1.3) and (1.4). The reader is referred to these references for the technical details. The significant point is that the averaged energy conditions don't assume that the integrals converge, nor that they are bounded above.) It was also shown in \cite{MTY} and \cite{FSW} that the averaged null energy conditions (1.3) and (1.4), respectively, prevent the occurence of traversable wormholes in solutions to Einstein's equations. In the light of these findings, an interesting issue is whether such averaged energy (positivity) conditions are fulfilled for the expectation values of the energy-momentum tensor for (suitable) states in quantum field theories. There have been several works dealing with this question and we continue by summarizing, however briefly, the results found so far. To fix our terminology, we say that a state $\omega} \def\O{\Omega$ of a quantum field theory on some background spacetime fulfills the ANEC (resp., AWEC = averaged weak energy condition) if the expectation value $\langle T_{\mu\nu}(x)\rangle_{\omega} \def\O{\Omega}$ of the energy-momentum tensor exists in this state as a (smooth) function of all $x$ in spacetime, and if \begin{equation} \liminf_{r_{\pm} \to \infty}\, \int^{r_+}_{-r_-}\langle T_{\mu\nu}(\gamma} \def\G{\Gamma (s)) \rangle_{\omega} \def\O{\Omega}\, k^{\mu}k^{\nu}\, ds \ge 0 \end{equation} holds for all complete lightlike (resp., timelike) geodesics $\gamma} \def\G{\Gamma$ with affine parameter $s$ and tangent $k^{\mu}$. (However, it should be noted that in some references slightly different formulations are used.) In \cite{Klin} it is shown that ANEC and AWEC are fulfilled for the free scalar field in $n$-dimensional Minkowski spacetime for states which are bounded in particle number and energy. It was also found in this work that AWEC is violated in some states of the free scalar field on a static, spatially closed two-dimensional spacetime. The work \cite{Fol} establishes ANEC for states bounded in particle number and energy of the free electromagnetic field in four-dimensional Minkowski spacetime. In the article \cite{WalYu} it is shown that ANEC holds for all Hadamard states of the massless free scalar field in any two-dimensional globally hyperbolic spacetime, and for all Hadamard states of the massive free scalar field on two-dimensional Minkowski spacetime. Moreover it is proved that ANEC holds for the massive and massless free scalar fields fulfilling some additional condition (implying that the limit $r_{\pm} \to \infty$ in (1.5) exists) in four-dimensional flat spacetime. In that work there appears also an argument indicating that ANEC cannot be expected to hold in general for the massless free scalar field on all four-dimensional curved spacetimes. Conditions implying that ANEC and AWEC will fail to hold generally for a large class of curved spacetimes are given in \cite{Vis}. In \cite{Y1,Y2} it has therefore been suggested that the original formulation of ANEC should be altered via replacing the integrand of (1.5) by $$ \langle T_{\mu\nu}(\gamma} \def\G{\Gamma (s))\rangle_{\omega} \def\O{\Omega} - D_{\mu \nu}(\gamma} \def\G{\Gamma (s)) $$ where $D_{\mu \nu}(x)$ is some state-independent tensor, e.g.\ the expectation value of the energy-momentum tensor in some reference state (like the vacuum in flat spacetime) or some quantity locally constructed from curvature terms. Such formulation of ANEC has been termed ``difference inequality''. Results in \cite{Y1,Y2} and \cite{FR} (cf.\ also \cite{FlaWa}) indicate that such difference inequalities may have a better chance to hold generally in curved spacetime. We refer to the references for further discussion. At any rate, investigations about the validity of ANEC (or difference inequalities) so far have been limited to the consideration of free fields only. The proofs of ANEC presented up to now strongly rely either on the fact that the quantum field obeys a linear hyperbolic equation of motion, or on the explicit form of the Wick-ordered energy-momentum tensor operator as bilinear expression in annihilation and creation operators in Fockspace. This is clearly unsatisfactory if one wishes to assess the general validity of ANEC in quantum field theory (say, in flat spacetime). Moreover, one would like to understand the connection of ANEC to the standard stability requirement in general quantum field theory, i.e.\ the spectrum condition and existence of a vacuum state. In the present work, we make a first attempt towards clarifying the status of ANEC in general quantum field theory. We shall consider a general quantum field theory on two-dimensional Minkowski spacetime obeying the usual assumptions like locality, translation covariance, spectrum condition with mass gap and existence of a unique vacuum. Furthermore we assume that such a theory possesses an energy-momentum tensor, which is essentially supposed to be a Wightman field (operator valued distribution) characterized by being local relative to the observables, divergence-free, generating locally the translations, and fulfilling an energy-bound. The precise assumptions are given in Section 2.1. Comments on these assumptions and some well-known consequences (needed later) appear in Section 2.2. In Section 2.3 we prove that ANEC is fulfilled for a dense, translationally invariant set of vector states of any quantum field theory in two-dimensional Minkowski-spacetime fulfilling the general assumptions of Section 2.1. In Section 3 we show that ANEC will in general fail to hold if the integral averaging is carried out only along a lightlike geodesic half-line as in (1.3). This is of course expected in view of locality and the Reeh-Schlieder property. Some concluding remarks appear in Section 4. We have opted to stage our discussion in the framework of the operator-algebraic approach to local quantum field theory (cf.\ \cite{Haag,BW}) since this makes the structures involved in the argument particularly transparent. One could also obtain similar results working entirely in the setting of Wightman fields \cite{SW}. \section{ANEC in quantum field theory on two-dimensional Minkowski spacetime} \setcounter{equation}{0} \subsection{Assumptions} Our discussion of the ANEC in general quantum field theory on two-dimensional Minkowski spacetime begins by formulating the relevant assumptions. \\[6pt] {\it Notation. } Two-dimensional Minkowski-spacetime will be identified, as usual, with ${\mathbb R}^2$ with metric $(\eta_{\mu\nu}) = {\rm diag}(1,-1)$. The open forward lightcone is the set $V_+ :=\{x \in {\mathbb R}^2: (x^0)^2 - (x^1)^2 > 0,\ x^0 > 0\}$, the open backward lightcone is $V_- := - V_+$. The causal complement, ${\cal O}^{\perp}$, of a set ${\cal O} \subset {\mathbb R}^2$ is the largest open complement of the union of all sets $(V_+ \cup V_- ) + x$, $x \in {\cal O}$. A {\it double cone} is a set of the form ${\cal O}_{I} := (S\backslash I)^{\perp}$ where $S$ is any spacelike line in ${\mathbb R}^2$ (a spacelike hypersurface) and $I$ any finite open subinterval of $S$. Any double cone is of the form ${\cal O} = (V_+ + y) \cap (V_- + x)$ for pairs of points $x,y \in {\mathbb R}^2$ with $x \in V_+ +y$. A {\it wedge region} is of the form $W = L(W_R)$ for any Poincar\'e transformation $L$ where $W_R$ is the right wedge, $W_R :=\{(x^0,x^2) \in {\mathbb R}^2 : 0<x^1,\ |x^0| < x^1 \}$. There will often appear the following special elements in ${\mathbb R}^2$: $$ e_0 := \left(^1_0\right)\,,\ \ e_1 := \left(^0_1\right)\,, \ \ e_+ := \mbox{$\frac{1}{\sqrt{2}}$} (e_0 + e_1)\,,\ \ e_- := \mbox{$\frac{1}{\sqrt{2}}$}(e_0 - e_1)\,. $$ The summation convention is used throughout. \\[6pt] We shall now define what we mean by a quantum field theory with an energy-momentum tensor on two-dimensional Minkowski-spacetime: This is described in terms of a collection of objects $\{{\cal H},{\cal A},U,\O,T_{\mu\nu}\}$ whose properties are assumed to be as follows: \begin{itemize} \item[(i)] ${\cal H}$ is a Hilbertspace, and there is a map ${\cal O} \mapsto {\cal A}({\cal O})$ assigning to each double cone ${\cal O}$ in ${\mathbb R}^2$ a von Neumann algebra in ${\cal B}({\cal H})$, with the properties: \\[2pt] ${}$ \quad $\tilde{{\cal O}} \subset {\cal O} \quad \ \Rightarrow \quad {\cal A}(\tilde{{\cal O}}) \subset {\cal A}({\cal O})$ \quad (isotony), \\[2pt] ${}$ \quad $\tilde{{\cal O}} \subset {\cal O}^{\perp} \quad \Rightarrow \quad {\cal A}(\tilde{{\cal O}}) \subset{\cal A}({\cal O})'$ \quad (locality). \footnote{Recall that ${\cal A}({\cal O})'$ is the commutant of ${\cal A}({\cal O})$, i.e.\ the algebra formed by all operators in ${\cal B}({\cal H})$ that commute with every element in ${\cal A}({\cal O})$.} \item[(ii)] There is a weakly continuous representation ${\mathbb R}^2 \owns a \mapsto U(a)$ of the two-dimensional translation group by unitary operators on ${\cal H}$, fulfilling for all double cones ${\cal O}$, $$ U(a){\cal A}({\cal O})U(a)^* = {\cal A}({\cal O} + a)\,, \quad a \in {\mathbb R}^2 \quad {\rm (covariance)}.$$ \item[(iii)] There is an up to a phase unique unit vector $\O \in {\cal H}$ which is left invariant by the unitary group $U(a)$, $a \in {\mathbb R}^2$ \quad (existence of a unique vacuum). \item[(iv)] Denote by $P = (P_0,P_1)$ the generator of $U(a)$, $a \in {\mathbb R}^2$, i.e.\ $U(a) = {\rm e}^{iP_{\mu}a^{\mu}}$. Its spectrum fulfills \\[2pt] $${\rm sp}(P) \subset \{0\} \cup \{(p_0,p_1) \in {\mathbb R}^2: (p_0)^2 -(p_1)^2 \ge m > 0, \ p_0 > 0\} $$ for some fixed $m > 0$ \ \,(spectrum condition with mass gap). \item[(v)] The vacuum vector $\O$ is cyclic for union of the local von Neumann algebras $\bigcup_{{\cal O}} {\cal A}({\cal O})$, i.e.\ the set $\bigcup_{{\cal O}}{\cal A}({\cal O})\O$ is dense in ${\cal H}$ \quad (cyclicity of the vacuum). \item[(vi)] We denote by ${\cal A}_{\infty}$ the $*$-subalgebra of ${\cal B}({\cal H})$ generated by all operators $A$ of the form $$A = \int h(a)\,U(a)B U(a)^*\, d^2a$$ for $h \in C^{\infty}_{0}({\mathbb R}^2)$ and $B \in \bigcup_{{\cal O}} {\cal A}({\cal O})$, and define: $${\cal A}_{\infty}({\cal O}) := {\cal A}_{\infty} \cap {\cal A}({\cal O})\,.$$ \par The energy-momentum tensor, $T_{\mu\nu}$, $\nu , \mu = 1,2$, is a set of operator valued distributions; more precisely, there is a dense domain $D \subset {\cal H}$, with $U(a)D \subset D$, $a \in {\mathbb R}^2$, and ${\cal A}_{\infty}\O \subset D$, so that for each $f \in C^{\infty}_{0}({\mathbb R}^2)$, $T_{\mu\nu}(f)$ is a closable operator on $D$ with $T_{\mu\nu}(\overline{f}) \subset T_{\mu\nu}(f)^*$. For each $\psi,\psi' \in D$, the map $$C^{\infty}_{0}({\mathbb R}^2) \owns f \mapsto \langle \psi,T_{\mu\nu}(f) \psi' \rangle$$ is a distribution in ${\cal D}'({\mathbb R}^2)$. \item[(vii)] Translation-covariance holds: $$U(a)T_{\mu\nu}(f)U(a)^* = T_{\mu\nu}(f_a)\,, \quad f \in C^{\infty}_{0}({\mathbb R}^2),\ a \in {\mathbb R}^2\,,$$ with $f_a(x) := f(x -a)$. Moreover, $T_{\mu\nu}$ has vanishing vacuum-expectation value: $$\langle \O,T_{\mu\nu}(f)\O \rangle = 0\,, \quad f \in C^{\infty}_{0}({\mathbb R}^2)\,. $$ \item[(viii)] $T_{\mu\nu}$ is local on the vacuum: \quad $\langle A\O,\mbox{\boldmath $[$} T_{\mu\nu}(f),B \mbox{\boldmath $]$} \O \rangle = 0$ \\[2pt] for all $A \in {\cal A}_{\infty}$, $B \in {\cal A}_{\infty}({\cal O})$ and $f \in C^{\infty}_{0}({\cal O}^{\perp})$. \item[(ix)] $T_{\mu\nu}$ is divergence-free on the vacuum: $$ \langle A \O,\mbox{\boldmath $[$} T_{\mu\nu}(\partial^{\mu}f),B \mbox{\boldmath $]$} \O \rangle = 0\,, \quad A,B \in {\cal A}_{\infty} \,.$$ \item[(x)] $T_{\mu\nu}$ generates (locally) the translations on the vacuum: Let $S$ be the $x^0 = 0$ hyperplane (= spacelike line) with unit normal vector $e_0$. Whenever $a,b \in C^{\infty}_{0}({\mathbb R})$ are any two non-negative functions with the properties \\[2pt] ${}$ \quad $a(x^0) =0$ outside of some $x^0$-interval $(-\varepsilon_a,\varepsilon_a)$ and $\int a(x^0)\,dx^0 = 1$, \\[2pt] ${}$ \quad $b(x^1) = 1$ on an open $x^1$-interval $(\xi_b -\varepsilon_a -\delta} \def\D{\Delta_b,\xi_b + \varepsilon_a + \delta} \def\D{\Delta_b)$, \\[2pt] ${}$ \quad where $\varepsilon_a,\delta} \def\D{\Delta_b > 0$, $\xi_b \in {\mathbb R}$, \\[2pt] we require that, upon setting $\chi(x^0,x^1) := a(x^0)b(x^1)$, there holds \begin{eqnarray*} \langle A \O,\mbox{\boldmath $[$} T_{\mu\nu}(\chi),B\mbox{\boldmath $]$} \O \rangle e_0^{\mu} & = & \langle A \O,\mbox{\boldmath $[$} P_{\nu},B \mbox{\boldmath $]$} \O \rangle \\ & = & \langle A \O,P_{\nu}B \O \rangle \end{eqnarray*} for all $A \in {\cal A}_{\infty}$ and all $B \in {\cal A}_{\infty}({\cal O}_I)$ with $I = (\xi_b -\delta} \def\D{\Delta_b,\xi_b + \delta} \def\D{\Delta_b)$. \item[(xi)] Energy bounds for $T_{\mu\nu}$: \quad There is a pair of numbers $c,\ell >0$ such that $(1 + P_0)^{-\ell}T_{\mu\nu}(f)(1+P_0)^{-\ell}$ is for each $f \in C^{\infty}_{0}({\mathbb R}^2)$ a bounded operator whose operator norm satisfies the estimate $$ ||\,(1 + P_0)^{-\ell}T_{\mu\nu}(f)(1+P_0)^{-\ell}\,|| \le c\,||\,f\,||_{L^1}\,, \quad f \in C^{\infty}_{0}({\mathbb R}^2)\,.$$ \end{itemize} \subsection{Comments and some implications} The conditions (i)--(v) imply that we are considering a translation-covariant quantum field theory in a vacuum representation with mass gap, in operator algebraic formulation. These conditions are quite standard; the selfadjoint elements in ${\cal A}({\cal O})$ are viewed as observables of the theory localized in the spacetime region ${\cal O}$, cf.\ \cite{Haag} for further discussion. Note that (v) and uniqueness of the vacuum vector imply irreducibility of the observable algebra, i.e.\ $(\bigcup_{{\cal O}}{\cal A}({\cal O}))' = {\mathbb C}\,1$. Note also that locality, spectrum conditon and (v) imply the Reeh-Schlieder property of the algebras associated with wedge-regions $W$, defined as ${\cal A}(W) := (\bigcup_{{\cal O} \subset W}{\cal A}({\cal O}) )''$, i.e.\ the sets ${\cal A}(W)\O$ are dense in ${\cal H}$ for any wedge-region. It follows easily that then also the sets ${\cal A}_{\infty}(W)\O$ are dense in ${\cal H}$ for all wedge regions $W$. A slightly stronger assumption would be the Reeh-Schlieder property for the local algebras, i.e.\ that ${\cal A}({\cal O})\O$ is dense in ${\cal H}$ for each double cone ${\cal O}$; this is the case when the local von Neumann algebras are weakly additive, as e.g.\ when there is a Wightman field generating the local algebras \cite{ReS,SW}. Then it follows that ${\cal A}_{\infty}({\cal O})\O$ is dense in ${\cal H}$. We will make such an assumption in Section 3. The conditions (vi)--(xi) serve to characterize an energy-momentum tensor in the present abstract setting. Conditions (vi)--(viii) basically say that the energy-momentum tensor is a Wightman field which is local relative to the observables. Particularly important for the interpretation of $T_{\mu\nu}$ as an energy-momentum tensor are clearly (ix) and (x) expressing that, in a weak sense, $T_{\mu\nu}$ is divergence-free and generates locally the translations. Notice that on account of the assumed translation-covariance the condition formulated in (x) implies its validity for any translated copy $S +a$, $a \in {\mathbb R}^2$, of $S$ in place of $S$. It is worth pointing out that we could have also taken for $S$ any other spacelike hyperplane (= spacelike line) instead of the $x^0 = 0$ hyperplane, the proof of Theorem 2.5 below would then only involve changes in notation. Specializing to the $x^0 = 0$ hyperplane is thus just a matter of notational convenience. Notice that for each $A \in {\cal A}_{\infty}$ one has $A\O \in \bigcap_{j \in {\mathbb N}}{\rm dom}\,(1+ P_0)^j$. In view of the assumed energy bound, it actually follows that ${\cal A}_{\infty}\O$ is contained in the domain of $T_{\mu\nu}(f)$. (An assumption of this kind is clearly needed, otherwise it would be difficult to formulate that $T_{\mu\nu}(f)$ is local relative to the observables.) The energy bound (xi) has the simple interpretation that the local energy-momentum density integrated over a finite spacetime volume should be dominated by the total energy (or at least a sufficiently high moment of it). We mention as an aside that, if one assumes the domain $D$ of $T_{\mu\nu}$ to coincide with the set $\bigcap_{j \in {\mathbb N}}{\rm dom}\,(1+P_0)^j$ and takes as testfunction-space the Schwartz-functions ${\cal S}({\mathbb R}^2)$ instead of $C^{\infty}_{0}({\mathbb R}^2)$, then this implies already an energy bound of the form as in (xi) \cite[Prop. 12.4.10]{BW}. Finally, there arises the question if our assumptions regarding $T_{\mu\nu}$ are realistic. For free fields, the canonically constructed energy-momentum tensor fulfills the assumptions. (It fulfills, in particular, a linear energy bound, i.e.\ (xi) holds with $\ell = 1$.) While we have made no attempt to check this, it is to be expected that the quantum field models which have been constructed in two dimensions, like $P(\phi)_2$ or Yukawa$_2$, also comply with all of our assumptions. \\[6pt] The assumptions (i)--(xi) for a theory with energy-momentum tensor, $\{{\cal H},{\cal A},U,\O,T_{\mu\nu}\}$, are known to imply certain properties which will be used in deriving ANEC in the next section. Hence we subsequently collect these properties, mainly referring to the literature for proofs. \begin{Prop} {\rm \cite{Bor,Dri}} One has weak asymptotic lightlike clustering: For any lightlike $k \in {\mathbb R}^2\backslash\{0\}$ and any pair of vectors $\psi,\psi' \in {\cal H}$, it holds that \begin{equation} \lim_{s \to \infty}\, \langle \psi,U(s\cdot k)\psi'\rangle = \langle \psi,\O\rangle\langle \O,\psi'\rangle\,. \end{equation} \end{Prop} \noindent {\it Sketch of Proof: } Let $W_R$ be the right wedge region and ${\cal A}(W_R)$ the associated von Neumann algebra. Let $\D^{it}$, $t \in {\mathbb R}$, be the modular group corresponding to ${\cal A}(W_R),\O$. Then a theorem by Borchers \cite[Thm. II.9]{Bor} establishes the relation $$ \D^{it}U(s\cdot e_+)\D^{-it} = U({\rm e}^{-2\pi t}s \cdot e_+) $$ for all $t,s \in {\mathbb R}$. Consequently, one can apply the argument of Prop.\ I.1.3 in \cite{Dri} to gain relation (2.1). We point out that the mass gap assumption enters in that argument. \begin{Lemma} Let $E := 1 - |\O \rangle \langle \O |$ be the projection orthogonal to the vacuum vector, and let $P_{\pm} := P_0 \pm P_1$. Let $\psi,\psi' \in {\rm dom}\,(P_0)$. Then there exist vectors $\psi_{\pm} \in {\cal H}$ such that $$ \langle \psi, E \psi' \rangle = \langle \psi_{\pm}, E P_{\pm} \psi' \rangle\,. $$ \end{Lemma} \begin{proof} From the mass-gap assumption we obtain \begin{equation} \frac{1}{|p_{\pm}|^2} \le \frac{|p_0|^2}{m^2} \end{equation} for all $p = (p_0,p_1) \in {\rm sp}(P) \backslash \{0\}$, where $p_{\pm} := p_0 \pm p_1$. We claim that the vectors $\psi_{\pm} := (P_{\pm})^{-1}E\psi$ exist (in the sense of the functional calculus). Indeed, denoting the spectral measure of $P$ by $F$, eqn.\ (2.2) implies \begin{eqnarray*} \lefteqn{||\,(P_{\pm})^{-1}E\psi\,||^2 = \int_{{\rm sp}(P) \backslash \{0\}}\frac{1}{|p_{\pm}|^2} \langle \psi,dF(p)\psi\rangle} \\ & \le & \int_{{\rm sp}(P) \backslash \{0\}}\frac{|p_0|^2}{m^2}\langle \psi,dF(p)\psi \rangle \ \le \ \frac{1}{m^2} ||\,P_0\psi\,||^2\,. \end{eqnarray*} Thus, by the functional calculus, $$ \langle\psi,E\psi'\rangle = \langle E\psi,E\psi'\rangle = \langle P_{\pm} (P_{\pm})^{-1} E\psi,E\psi'\rangle = \langle \psi_{\pm},E P_{\pm}\psi' \rangle\,, $$ where we used that $E$ commutes with $P_{\pm}$. \end{proof} \begin{Prop} Let $f \in C^{\infty}_{0}({\mathbb R}^2)$ with $f \ge 0$, $\int f(x)\,d^2x = 1$, and define $f_{x,\lambda} \def\L{\Lambda}(y):= \lambda} \def\L{\Lambda^{-2}f(\lambda} \def\L{\Lambda^{-1}(y-x))$ so that $f_{x,\lambda} \def\L{\Lambda}$ approaches for $\lambda} \def\L{\Lambda \to 0$ the delta-distribution concentrated at $x$. Then for each pair $A,B \in {\cal A}_{\infty}$, the limit $$ \langle A\O,T_{\mu\nu}[x]B\O\rangle := \lim_{\lambda} \def\L{\Lambda \to 0}\, \langle A\O,T_{\mu\nu}(f_{x,\lambda} \def\L{\Lambda}) B\O \rangle $$ exists and defines a quadratic form on ${\cal A}_{\infty}\O \times {\cal A}_{\infty}\O$. Moreover, \begin{itemize} \item[{\rm (a)}] \quad ${\mathbb R}^2 \owns x \mapsto \langle A\O, T_{\mu\nu}[x] B\O\rangle$ is $C^{\infty}$, \item[{\rm (b)}] \quad $\langle U(a)A\O,T_{\mu\nu}[x]U(a)B\O\rangle = \langle A\O,T_{\mu\nu}[x - a]\,B\O \rangle$\,, \quad $x,a \in {\mathbb R}^2$, \item[{\rm (c)}] \quad $(1 + P_0)^{-\ell}T_{\mu\nu}[x](1 + P_0)^{-\ell} := \lim_{\lambda} \def\L{\Lambda \to 0}\, (1 + P_0)^{-\ell}T_{\mu\nu}(f_{x,\lambda} \def\L{\Lambda})(1+P_0)^{-\ell}$ \\[2pt] is a bounded operator on ${\cal H}$, \item[{\rm (d)}] \quad $\langle A\O,\mbox{\boldmath $[$} T_{\mu\nu}[x],B\mbox{\boldmath $]$}\O\rangle = 0$ \quad for all $A \in {\cal A}_{\infty}$, $B \in {\cal A}_{\infty}({\cal O})$ and $x \in {\cal O}^{\perp}$, \item[{\rm (e)}] \quad $\partial^{\mu}\langle A\O,\mbox{\boldmath $[$} T_{\mu\nu}[x],B\mbox{\boldmath $]$} \O \rangle = 0$, \quad $A,B \in {\cal A}_{\infty}$, \item[{\rm (f)}] \quad $\int \langle A\O,\mbox{\boldmath $[$} T_{\mu\nu}[x^1e_1],B\mbox{\boldmath $]$}\O\rangle e_0^{\mu}\,dx^1 = \langle A\O,P_{\nu}B\O \rangle$, \quad $A,B \in {\cal A}_{\infty}$. \end{itemize} \end{Prop} This proposition is a fairly direct consequence of assumption (xi), see \cite[Thm.\ 12.4.8]{BW} (cf.\ also references cited there). The commutator is defined as difference of quadratic forms: $$ \langle A\O,\mbox{\boldmath $[$} T_{\mu\nu}[x],B\mbox{\boldmath $]$} \O \rangle := \langle A\O,T_{\mu\nu}[x]B\O\rangle - \langle B^*A\O,T_{\mu\nu}[x]\O \rangle\,. $$ Observe that the integrand in (f) is supported on a finite interval because of (d). It should also be noted that $T_{\mu\nu}[x]$ will in general not exist as an operator. \begin{Lemma} Let $W$ be a wedge region, $B \in {\cal A}_{\infty}$, and $j \in {\mathbb N}$. Then for each $\varepsilon > 0$ there is some $B_{\varepsilon} \in {\cal A}_{\infty}(W)$ such that $$ ||\,(1 + P_0)^j(B - B_{\varepsilon})\O\,|| < \varepsilon\,. $$ \end{Lemma} The proof can be given along similar lines as the proof of \cite[Prop.\ 14.3.2]{BW}; we may therefore skip the details. The cyclicity of $\O$ for the algebras ${\cal A}_{\infty}(W)$ enters here. In combination with (b) and (c) of Prop.\ 2.3 one obtains as a simple corollary: For each wedge region $W$, any $A,B \in {\cal A}_{\infty}$ and given $\varepsilon > 0$ there is some $B_{\varepsilon} \in {\cal A}_{\infty}(W)$ so that \begin{equation} |\langle A\O,T_{\mu\nu}[x](B-B_{\varepsilon})\O\rangle | < \varepsilon \end{equation} holds uniformly in $x \in {\mathbb R}^2$. \subsection{Main result} In the present section we state and prove our main result about ANEC in quantum field theory on two-dimensional Minkowski spacetime. \begin{Thm} Let $\{{\cal H},{\cal A},U,\O,T_{\mu\nu}\}$ be a quantum field theory with energy-momentum tensor on two-dimensional Minkowski spacetime fulfilling the assumptions (i)--(xi) of Section 2.1. Let $k$ be any non-zero lightlike vector in ${\mathbb R}^2$ and let $A,B \in {\cal A}_{\infty}$, $a \in {\mathbb R}^2$. Then it holds that $$ \lim_{r_{\pm} \to \infty} \, \int_{-r_-}^{r_+} \langle A\O,T_{\mu\nu}[s\cdot k + a]\,B\O \rangle k^{\mu}\, ds = \langle A\O,P_{\nu} B\O \rangle\,. $$ \end{Thm} \begin{Cor} This implies the ANEC for all vector states induced by the dense, translation-invariant set of vectors $\{\psi = A\O$: $A \in {\cal A}_{\infty}\}$, corresponding to energetically strongly damped, local excitations of the vacuum: $$ \lim_{r_{\pm}\to \infty} \, \int_{-r_-}^{r_+} \langle \psi,T_{\mu\nu}[s \cdot k + a]\,\psi\rangle k^{\mu}k^{\nu}\, ds = \langle \psi,k^{\nu}P_{\nu}\psi \rangle \ge 0 $$ since $k$ is lightlike and since the relativistic spectrum condition holds. \end{Cor} \begin{proof} The proof proceeds in three simple steps. For simplicity of notation, we will give the proof only for the case $k = e_+$, the proof for $k = e_-$ is obtained by analogous arguments. In view of translation covariance, it suffices also to consider only the case $a=0$. \\[6pt] {\it 1) } We will first show that for all $C \in {\cal A}_{\infty}$ \begin{eqnarray} \lim_{s \to \pm\infty}\,\langle C\O,T_{\mu\nu}[s\cdot e_+]\O \rangle& =& 0\,, \\ \lim_{r_{\pm}\to\infty}\, \int_{-r_-}^{r_+} \langle C\O,T_{\mu\nu}[s \cdot e_+]\O \rangle \,ds &=& 0\,. \end{eqnarray} To this end, let \begin{eqnarray*} \psi &:= & (1+P_0)^{\ell + 1}C\O\,,\\ \psi_{\mu \nu}' & : = & (1+ P_0)^{-(\ell + 1)}T_{\mu\nu}[0](1 + P_0)^{-\ell}\O\,. \end{eqnarray*} One can see from Prop.\ 2.3(c) that $\psi,\psi_{\mu\nu}' \in {\rm dom}\,(1 + P_0)$. Moreover, denoting by $E := 1 -|\O\rangle \langle \O |$ the projection orthogonal to the vacuum, we deduce, upon using assumption (vii) (implying $\langle \O,T_{\mu\nu}[x]\O\rangle = 0$) and Prop.\ 2.3(b) together with the fact that $E$ commutes with $U(a)$, $a \in {\mathbb R}^2$, that $$ \langle C\O,T_{\mu\nu}[s\cdot e_+]\O \rangle = \langle \psi,E\,U(s\cdot e_+)\psi_{\mu\nu}'\rangle \,, \quad s \in {\mathbb R}\,.$$ Then relation (2.4) follows from weak asymptotic lightlike clustering, Prop.\ 2.1. Furthermore, by Lemma 2.2 it follows that there is a vector $\psi_+ \in {\cal H}$ so that $$ \langle \psi,E\,U(s\cdot e_+)\psi_{\mu\nu}'\rangle = \langle \psi_+,E P_+U(s\cdot e_+)\psi_{\mu\nu}' \rangle = \frac{1}{i}\frac{d}{ds}\langle \psi_+,E\,U(s\cdot e_+)\psi_{\mu\nu}'\rangle \,.$$ \newpage\noindent ${}$ \par \vspace*{-1.5cm} \noindent Thus one obtains \begin{eqnarray*} \lefteqn{\int_{-r_-}^{r_+}\langle C\O,T_{\mu\nu}[s\cdot e_+]\O\rangle\,ds = \int_{-r_-}^{r_+} \frac{1}{i}\frac{d}{ds}\langle\psi_+,E\, U(s\cdot e_+)\psi_{\mu\nu}'\rangle\,ds}\\ & = & \frac{1}{i}\left(\langle\psi_+,E\,U(r_+\cdot e_+)\psi_{\mu\nu}'\rangle - \langle\psi_+,E\,U(r_-\cdot e_+)\psi_{\mu\nu}'\rangle \right) \end{eqnarray*} and the last expression tends to 0 in the limit $r_{\pm} \to \infty$ in view of weak asymptotic lightlike clustering, Prop.\ 2.1. This establishes relation (2.5). \\[6pt] {\it 2) } Relation (2.5) shows, for any $A,B \in {\cal A}_{\infty}$, $$ \lim_{r_{\pm} \to \infty}\, \left( \int_{-r_-}^{r_+}\langle A\O,T_{\mu\nu}[s\cdot e_+]B\O\rangle\,ds\ - \int_{-r_-}^{r_+} \langle A\O,\lbT_{\mu\nu}[s\cdot e_+],B\mbox{\boldmath $]$}\O\rangle\,ds \right) = 0 $$ and hence, to prove the theorem, it suffices to demonstrate \begin{equation} \lim_{r_{\pm} \to \infty}\,\int_{-r_-}^{r_+} \langle A \O,\mbox{\boldmath $[$} T_{\mu\nu}[s\cdot e_+],B \mbox{\boldmath $]$} \O \rangle e_+^{\mu}\,ds = \langle A\O,P_{\nu} B\O \rangle\,. \end{equation} To show this, we fix any $A,B \in {\cal A}_{\infty}$ and use the abbreviation $$ \tau_{\mu\nu}(x) := \langle A\O,\mbox{\boldmath $[$} T_{\mu\nu}[x],B \mbox{\boldmath $]$} \O \rangle\,. $$ Now we define two maps with values in ${\mathbb R}^2$, $$ h_+(s,\rho) := s\cdot e_+ - \rho\cdot e_-\,, \quad \ h_-(s,\rho) := -s\cdot e_+ + \rho \cdot e_-\,, \quad s,\rho \ge 0\,,$$ and the two triangle-shaped regions \begin{eqnarray*} R_{+,r_+} &:=& \{h_+(s,\rho): 0 \le s \le r_+,\ 0 \le \rho \le s\}\,,\\ R_{-,r_-} &:= &\{h_-(s,\rho): 0 \le s \le r_-,\ 0 \le \rho \le s\}\,. \end{eqnarray*} The region $R_{+,r_+}$ is bounded by the two lightlike line segments $L_+(r_+) :=\{s\cdot e_+ : 0 \le s \le r_+\}$ and $H_+(r_+):= \{h_+(r_+,\rho): 0 \le \rho \le r_+\}$, and by the spacelike line segment $S_+(r_+) := \{x^1e_1 : 0 \le x^1 \le \sqrt{2}\,r_+\}$. Similarly, $R_{-,r_-}$ is bounded by the line segments $L_-(r_-) := \{- s \cdot e_+: 0 \le s \le r_-\}$, $H_-(r_-):= \{h_-(r_-,\rho): 0 \le \rho \le r_-\}$, and $S_-(r_-):= \{-x^1e_1 : 0 \le x^1 \le \sqrt{2}\,r_-\}$. (Cf.\ Figure 1.) \begin{center} \epsfig{file=fig1.eps, width=11.0cm}\\ {\small {\bf Figure 1. } \quad Sketch of the regions and bounding line segments described in the text.} \end{center} Now we use $\partial^{\mu}\tau_{\mu\nu}(x) = 0$ and thus, applying Gau{\ss}' law to the region $R_{+,r_+}$, we convert the integral of $ v^{\mu}=\tau^{\mu}{}_{\nu}$ paired with the outer normal along $L_+(r_+)$ into a sum of two integrals of $v^{\mu}$ paired with the inner normals along $H_+(r_+)$ and $S_+(r_+)$. Doing the same with respect to the region $R_{-,r_-}$ (with the roles of inner and outer normals interchanged) yields, with the above parametrizations of the various line segments inserted, \begin{eqnarray} \lefteqn{ \int_{-r_-}^{r_+} \tau_{\mu\nu}(s\cdot e_+)e^{\mu}_+\,ds = \int_{-\sqrt{2}\,r_-}^{\sqrt{2}\,r_+}\tau_{\mu\nu}(x^1e_1)e^{\mu}_0\,dx^1} \\ & & - \int_0^{r_+}\tau_{\mu\nu}(h_+(r_+,\rho))e_-^{\mu}\,d\rho\ -\int_{0}^{r_-}\tau_{\mu\nu}(h_-(r_-,\rho))e^{\mu}_+ \,d\rho \,. \nonumber \end{eqnarray} In view of Prop.\ 2.3(d,f), we deduce that the first integral on the right hand side of (2.7) equals $\langle A\O,P_{\nu}B\O\rangle$ as soon as $r_+$ and $r_-$ are large enough. This implies that (2.6), and hence the statement of the theorem, is proved once it is shown that the two remaining integrals on the right hand side of (2.7) vanish in the limit $r_{\pm} \to \infty$. \\[6pt] {\it 3) } The remaining step in the proof is therefore to show \begin{equation} \lim_{r_{\pm} \to \infty}\, \int_0^{r_{\pm}}\tau_{\mu\nu}(h_{\pm}(r_{\pm},\rho)) \,d\rho = 0\,. \end{equation} We will demonstrate this only for the ``$+$'' case, the reasoning for the ``$-$'' case is similar. It holds that $B \in {\cal A}_{\infty}({\cal O}_I)$ for $I =\{x^1e_1: |x^1| < \sqrt{2}\,\xi\}$ with some sufficiently large $\xi > 0$. By Prop.\ 2.3(d), $\tau_{\mu\nu}(x) = 0$ for $x \in ({\cal O}_I)^{\perp}$, implying that \begin{equation} \int_0^{r_+}\tau_{\mu\nu}(h_+(r_+,\rho))\,d\rho = \int_0^{\xi}\tau_{\mu\nu}(h_+(r_+,\rho))\,d\rho \,, \end{equation} i.e.\ the integral extends for all $r_+ > 0$ only over a fixed interval of finite length. Now choose some wedge region $W$ in the causal complement of $\bigcup_{r_+ \ge 0}H_+(r_+) \subset W_R$, and let $\delta} \def\D{\Delta > 0$ be arbitrary. According to (2.3), one can find some $B_{\delta} \def\D{\Delta} \in {\cal A}_{\infty}(W)$ so that $$ |\langle A\O,T_{\mu\nu}[x](B - B_{\delta} \def\D{\Delta})\O \rangle | < \frac{\delta} \def\D{\Delta}{2\xi} $$ uniformly in $x \in {\mathbb R}^2$. Then $\langle A\O,\lbT_{\mu\nu}[x],B_{\delta} \def\D{\Delta}\mbox{\boldmath $]$}\O\rangle = 0$ for all $x \in H_+(r_+)$, and \begin{eqnarray*} \lefteqn{\int_0^{\xi}\tau_{\mu\nu}(h_+(r_+,\rho))\,d\rho = \int_0^{\xi} \langle A\O,T_{\mu\nu}[h_+(r_+,\rho)](B - B_{\delta} \def\D{\Delta})\O\rangle\,d\rho}\\ &+ & \int_0^{\xi}\langle(B^*_{\delta} \def\D{\Delta} - B^*)A\O,T_{\mu\nu}[h_+(r_+,\rho)]\O\rangle\,d\rho\,. \end{eqnarray*} The absolute value of the first integral on the right hand side of the last equation can be estimated by $\xi\cdot\delta} \def\D{\Delta/2\xi = \delta} \def\D{\Delta /2$. Owing to (2.4), the other integral on the right hand side of the last equation converges to 0 for $r_+ \to \infty$ (note that the integrands are bounded uniformly in $r_+$). Therefore we can find for the given $\delta} \def\D{\Delta > 0$ some $r > 0$ so that $|\int_0^{\xi}\tau_{\mu\nu}(h_+(r_+,\rho))\,d\rho| < \delta} \def\D{\Delta$ for all $r_+ > r$. By (2.9), this establishes the required relation (2.8), and thus the proof is complete. \end{proof} \section{A result for lightlike half-lines} \setcounter{equation}{0} In this section we present a result indicating that ANEC fails to hold in general for dense subsets of the vectors considered in Theorem 2.5 when the expectation value of the energy-momentum tensor is integrated only over a lightlike half-line. This is of course no surprise in view of the fact that a lightlike half-line has a large causal complement together with the assumed properties of the energy-momentum tensor. The precise formulation of the result is as follows. \begin{Prop} Let $\{{\cal H},{\cal A},U,\O,T_{\mu\nu}\}$ be a quantum field theory with energy-momen\-tum tensor on two-dimensional Minkowski-spacetime with the properties assumed in Section 2.1. Let $k$ be a non-zero lightlike vector in ${\mathbb R}^2$, $a \in {\mathbb R}^2$, and let ${\cal O}$ be a double cone lying in the causal complement of the lightlike half-line $L := \{s\cdot k + a : s \ge 0\}$. Suppose that ${\cal A}_{\infty}({\cal O})\O$ is dense in ${\cal H}$ (Reeh-Schlieder property) and that for all $A \in {\cal A}_{\infty}({\cal O})$ there holds $$ \liminf_{r \to \infty}\,\int_0^r \langle A\O,T_{\mu\nu}[s\cdot k + a]A\O\rangle k^{\mu}k^{\nu}\, ds \ge 0\,.$$ Then the Hilbertspace ${\cal H}$ is one-dimensional and spanned by the vacuum vector $\O$, and $T_{\mu\nu}(f) = 0$ for all $f \in C^{\infty}_{0}({\mathbb R}^2)$. \end{Prop} \begin{proof} We consider only the case $k = e_+$ and $a = 0$, the general case is proved analogously. Then we observe that \begin{equation} \langle A\O,T[L]B\O \rangle := \lim_{r \to \infty}\,\int_0^r\langle A\O,T_{\mu\nu}[s\cdot e_+]A\O\rangle e_+^{\mu}e_+^{\nu}\,ds = i\langle A\O,T_{\mu\nu}[0]B\O\rangle e_+^{\mu}e_+^{\nu} \end{equation} holds for all $A,B \in {\cal A}_{\infty}({\cal O})$ as can be seen from (2.5) together with the fact that $\langle A\O,\lbT_{\mu\nu}[s\cdot e_+],B\mbox{\boldmath $]$}\O\rangle = 0$, $s \ge 0$. Equation (3.1) defines a quadratic form $\langle \,.\,,T[L]\,.\,\rangle$ on ${\cal A}_{\infty}({\cal O})\O \times {\cal A}_{\infty}({\cal O})\O$ which is by assumption positive, i.e.\ $\langle A\O,T[L]A\O\rangle \ge 0$, $A \in {\cal A}_{\infty}({\cal O})$. It follows that there is an essentially selfadjoint, positive operator $T_L^{1/2}$ with domain ${\cal A}_{\infty}({\cal O})\O$ so that $$ \langle T_L^{1/2}A\O,T^{1/2}_L B\O\rangle = \langle A\O,T[L]B\O\rangle = \langle B^*A\O,T[L]\O\rangle\,, \quad A,B \in {\cal A}_{\infty}({\cal O})\,.$$ Using $\langle \O,T[L]\O\rangle = 0$ this implies $\langle A\O,T[L]B\O\rangle = 0$ and hence, by (3.1), \begin{eqnarray*} \lefteqn{\langle(1+P_0)^{\ell}A\O,(1+P_0)^{-\ell}T_{\mu\nu}[0] (1+P_0)^{-\ell}(1+P_0)^{\ell}B\O\rangle e_+^{\mu}e_+^{\nu} }\\ &=& \langle A\O,T_{\mu\nu}[0]B\O\rangle e_+^{\mu}e_+^{\nu} \ = \ -i\langle A\O,T[L]B\O\rangle \ = \ 0 \end{eqnarray*} for all $A,B \in {\cal A}_{\infty}({\cal O})$. The set of vectors $(1+P_0)^{\ell}A\O$, $A \in {\cal A}_{\infty}({\cal O})$ is dense in ${\cal H}$, therefore, using also covariance (Prop. 2.3(b)), one arrives at $$ (1+P_0)^{-\ell}T_{\mu\nu}[x](1 + P_0)^{-\ell}e_+^{\mu}e_+^{\nu} = 0\,, \quad x \in {\mathbb R}^2\,.$$ Thus $\langle A\O,T_{\mu\nu}[x]B\O\rangle e_+^{\mu}e_+^{\nu} = 0$ for all $A,B \in {\cal A}_{\infty}$ and $x \in {\mathbb R}^2$, and in view of Theorem 2.5, this entails $$ (P_0 + P_1)B\O = 0\,, \quad B \in {\cal A}_{\infty}\,.$$ Since we have imposed the mass gap assumption (iv), we may apply Proposition I.1.2 of \cite{Dri} to conclude that this is only possible if $B\O$ is parallel to the vacuum vector $\O$. As the set of vectors ${\cal A}_{\infty}\O$ is dense in ${\cal H}$, this implies ${\cal H} = {\mathbb C}\cdot \O$, and by the vanishing of the vacuum-expectation value of $T_{\mu\nu}$, finally $T_{\mu\nu}(f) = 0$ for all $f \in C^{\infty}_{0}({\mathbb R}^2)$. \end{proof} \section{Concluding remarks} It has been shown that in two-dimensional Minkowski spacetime the ANEC can be derived under very general hypotheses for quantum field theories endowed with an energy-momentum tensor. The two-dimensionality was quite essential in exploiting the vanishing of the divergence of $T_{\mu\nu}$ in the proof of Thm.\ 2.5 and it is not at all clear if our simple argument can be generalized to higher dimensions. So the general validity of ANEC in higher dimensions remains unsettled. Concerning quantum field theory in curved spacetime, a familiar problem is that there are no candidates for a vacuum state of a quantum field theory owing to the circumstance that in general there are no spacetime symmetries. Then already the characterization of physical states and the definition of expectation values of an energy-momentum tensor poses considerable problems (see \cite{WaldII} for discussion how this problem is treated in the case of free fields). Clearly the the question if, and in which sense, ANEC may hold in quantum field theory in curved spacetime is connected to this circle of problems, particularly to the issue of how to characterize states which may be viewed as playing the role of preferred, vacuum-like states. Here ANEC is in some ways attractive as imposing a global constraint on candidates for such states, complementary to other prominent conditions, like the Hadamard condition (cf.\ \cite{WaldII}) or the microlocal spectrum condition (see \cite{BFK}) which constrain the short-distance properties of physical states. A drawback is that ANEC is a condition which cannot be tested locally. It is to be hoped that more progress in understanding the relation between the said local conditions and ANEC will be made in the future. \\[18pt] {\bf Acknowledgement.} It would like to thank D.\ Buchholz for useful comments on the topic.
\section*{Introduction} If one places a semiconducting heterostructure over a piezoelectric which SAW is being propagated, the SAW undergoes attenuation associated with the interaction of the electrons of heterostructure with the electric field of SAW. This is the basis of the acoustic method pioneered by Wixforth \cite{Wix1} for the investigation of GaAs/AlGaAs heterostructures. In the paper \cite{ild971} it has been found that in a GaAs/AlGaAs heterostructure in the IQHE regime the acoustically measured conductivity $\sigma^{hf}$ does not coincide with the $\sigma^{dc}$ obtained from the direct-current measurements: $\sigma^{dc}$=0, whereas $\sigma^{hf}$ has a finite value. This difference was explained by means of a conventional model of electrons being localized in the IQHE regime, therefore the conductivity mechanism for a direct current differs from that for an alternative current. For the localized electrons in the hopping conduction regime hf-conductivity can be expressed as a complex value: $\sigma^{hf}=\sigma_1^{hf}-i\sigma_2^{hf}$ \cite{aleiner}. Absorption coefficient of SAW, $\Gamma$, and the change of SAW velocity- $\Delta V/V$-can be presented in the way: \begin{eqnarray} \Gamma=8.68 \frac{K^2}{2} kA\frac{(\frac{4\pi \sigma_{1}}{\varepsilon_s V})t(k)} {[1+(\frac{4\pi \sigma_{2}}{\varepsilon_s V})t(k)]^2+ [(\frac{4\pi \sigma_{1}}{\varepsilon_s V})t(k)]^2}, \label{gam&vel} \end{eqnarray} $$ A=8b(k)(\varepsilon_1+\varepsilon_0)\varepsilon_0^2\varepsilon_sexp(-2k(a+d)), $$ $$ \frac{\Delta V}{V}= \frac{K^2}{2} A\frac{(\frac{4\pi \sigma_{2}}{\varepsilon_s V})t(k)+1} {[1+(\frac{4\pi \sigma_{2}}{\varepsilon_s V})t(k)]^2+ [(\frac{4\pi \sigma_{1}}{\varepsilon_s V})t(k)]^2}, $$ where $K^2$ is the electromechanical coupling coefficient of piezoelectric substrate, $k$ and $V$ are the SAW wavevector and the velocity respectively, $a$ is the vacuum gap width of 2DEG, $d$ is the depth of the 2D-layer, $\varepsilon_1$, $\varepsilon_0$ and $\varepsilon_s$ are the dielectric constants of lithium niobate, vacuum and semiconductor respectively, $b$ and $t$ are complex function of $a$, $d$, $\varepsilon_1$, $\varepsilon_0$ and $\varepsilon_s$. The aim of the work is to determine Re$\sigma^{hf}$(H,T) and Im$\sigma^{hf}$(H,T) in IQHE regime from the $\Gamma$ and $\Delta V/V$ of SAW measurements ($f$=30MHz, T=1.5-4.2K, H up to 7T) and to analyze the 2D-electrons localization mechanism. \section*{The Experimental results and discussion} Fig.1 illustrates the experimental dependencies of $\Gamma$ and $\Delta V/V$ on H at T=1.5K for the sample 1 ($n=2.7 \cdot 10^{11} cm^{-2}$). As long as $\Gamma$ and $\Delta V/V$ are determined by the 2DEG conductivity (Eq.(1)), quantizing of the electro nspectrum in the magnetic field, leading to the SdH oscillations, results in similar peculiarities in $\Gamma$ and $\Delta V/V$ of fig.1. Fig.2a presents the $\sigma_1(T)$ dependencies determined from $\Gamma$ and $\Delta V/V$ (Fig.1) using Eq.(1) for H=5.5; 2.7 and 1.8T ($\nu$=2, 4, 6 respectively), $\nu$=nch/eH is the filling factor. Fig.2b shows the $\sigma_2(T)$ dependencies derived from $\Gamma$ and $\Delta V/V$ for H=5.5, 2.7 and 1.8 T. Fig.3 illustrates the dependencies $\sigma_1$ and $\sigma_2$ on magnetic field near H=5.5T ($\nu$=2) at different temperatures. In a number of papers (see f.e.\cite{furlan}) devoted to the study of magnetoresistance in IQHE regime it was established that in IQHE plateau regions at low T the dominant conductivity mechanism is variable range hopping. HF-conductivity in this case is determined by the two-site model and is associated with the electronic transition between localized states of "tight" pairs of impurities $\sigma_1$. In this case the relation is valid: \begin{eqnarray} \frac{Re\sigma}{Im\sigma}=\frac{\sigma_1}{\sigma_2}= \frac{\pi}{2} ln\frac{\omega}{\omega_{ph}}, \label{reim} \end{eqnarray} where $\omega=2\pi f$ is the frequency of SAW, $\omega_{ph}$ is the characteristic phonon frequency of the order of $10^{12}-10^{13} sec^{-1}$. The calculation using Eq.(\ref{reim}) gives $\sigma_1/\sigma_2=0.15$ (f=30 MHz). For all samples it was experimentally found $\sigma_1/\sigma_2=0.14 \pm 0.03$ at T=1.5K and H in the Hall plateaux middle. This fact allows one to suppose that the mechanism that determines the $\Gamma$ and $\Delta V/V$ of SAW interacting with localized electrons is that of the hf hopping conductivity \cite{efros85}. For the analysis of the dependencies $\sigma_1(H,T)$ and $\sigma_2(H,T) $ we supposed it to be determined by two mechanisms: hf-hopping on the localized states of impurities and thermal activation to the upper Landau band. The value of $\sigma_2 $ is proportional to the number of "tight" pairs, which is changed with a thermal activation processes ($\sigma_2=0$ for delocalized electrons \cite{ild971}). So the dependence $ \sigma_2 (H, T)$ reflects a change of number of these pairs. The dependence $\sigma_1(H)$ is similar to behaviour of $\sigma^{dc}$ when it makes SdH-oscillations, but $\sigma_1 > \sigma^{dc}=0$. $\sigma_1(T)$ can be presented as a sum of two terms: $\sigma_1=\sigma_1^h+\sigma_1^a$, where $\sigma_1^h=0.11 \cdot \sigma_2^h$ is hf-conductivity, $ \sigma_1^a=\sigma_0 \cdot exp(-\Delta E/kT)$ is the conductivity on upper Landau band due to the electrons activated from the states on Fermi level where the activation energy $\Delta E=\hbar \omega_c/2-C/2$, ($\hbar \omega_c$ is the cyclotron frequency, $C$ is the Landau band width). From the plotted dependence $ln(\sigma_1-\sigma_1^h)$ on 1/T the $\Delta E$ was found for H=5.5, 2.7, 1.8T. One can see (inset of Fig.2a) that $\Delta E$ is linear function of H with the slope $0.5\hbar \omega_c$ what allow to derive the width of Landau band, broadened, probably, by the impurity random potential; the width is appeared to be $C$=2meV. \section*{Acknowledgements} The work was supported by RFFI (N 98-02-18280) and Minnauki (N 97-1043).
\section{Preliminaries and notation} In this section we introduce the notation and recall some well-known facts on Prym varieties. Throughout the paper we will suppose the genus of $C$ to be at least $5$. Let $\omega$ be the dualizing sheaf of $C$ and consider the two additional Prym varieties $$ Nm^{-1}(\omega) = P_{even} \cup P_{odd} $$ which are characterized by the fact that $\mathrm{dim} \: H^0(\tilde{C},\lambda)$ is even (resp. odd) for $\lambda \in P_{even}$ (resp. $P_{odd}$). The variety $P_{even}$ carries the naturally defined reduced Riemann theta divisor \[ \Xi :=\{\lambda\in P_{even} \ | \ h^0 (\lambda ) > 0\}\; \] a translate of which is $\Xi_0$. Let ${\mathcal SU}_C(2,\alpha)$ and ${\mathcal SU}_C(2,\omega \alpha)$ be the moduli spaces of semi-stable vector bundles of rank $2$ with determinant $\alpha$ and $\omega\alpha$ respectively. Taking direct image gives morphisms $$ \varphi: P \cup P' \longrightarrow {\mathcal SU}_C(2,\alpha) \qquad \varphi: P_{even} \cup P_{odd} \longrightarrow {\mathcal SU}_C(2,\omega \alpha). $$ Let $\mathcal{L}_\alpha$ (resp. $\mathcal{L}_{\omega \alpha}$) be the generator of the Picard group of ${\mathcal SU}_C(2,\alpha)$ (resp. ${\mathcal SU}_C(2,\omega \alpha)$). It is known that $$ (\varphi_{|P})^* \mathcal{L}_\alpha = {\mathcal O}(2\Xi_0) \qquad (\varphi_{|P_{even}})^* \mathcal{L}_{\omega \alpha} = {\mathcal O}(2\Xi). $$ We denote by ${\mathcal O}(2\Xi'_0)$ (resp. ${\mathcal O}(2\Xi')$) the pull-back of the line bundle $\mathcal{L}_\alpha$ (resp. $\mathcal{L}_{\omega \alpha}$) to the Prym $P'$ (resp. $P_{odd}$), i.e., $(\varphi_{|P'})^* \mathcal{L}_\alpha = {\mathcal O}(2\Xi'_0)$ and $(\varphi_{|P_{odd}})^* \mathcal{L}_{\omega \alpha} = {\mathcal O}(2\Xi')$. We consider the following morphisms $$ \psi : JC \longrightarrow {\mathcal SU}_C(2,\alpha) \qquad \xi \longmapsto \xi \oplus \alpha \xi^{-1} $$ $$ \psi : \mathrm{Pic}^{g-1}(C) \longrightarrow {\mathcal SU}_C(2,\omega \alpha) \qquad \xi \longmapsto \xi \oplus \omega \alpha \xi^{-1}. $$ One computes the pull-backs $$ \psi^* \mathcal{L}_\alpha = \Theta_0 + T^*_\alpha \Theta_0 \qquad \psi^* \mathcal{L}_{\omega \alpha} = \Theta + T^*_\alpha \Theta $$ where \[ \Theta :=\{ L\in Pic^{ g-1} (C) \ | \ h^0 (L )> 0\}\; \] and $\Theta_0$ is a symmetric theta divisor in the Jacobian $JC$, i.e., a translate of $\Theta$ by a theta-characteristic. By abuse of notation, we will also write $\mathcal{L}_{\alpha}$ and $\mathcal{L}_{\omega \alpha}$ for $\psi^*\mathcal{L}_{\alpha}$ and $\psi^*\mathcal{L}_{\omega\alpha}$ respectively. Note that $\psi$ induces linear isomorphisms at the level of global sections: \begin{equation} \label{iso1} \psi^* : H^0({\mathcal SU}_C(2,\alpha),\mathcal{L}_\alpha) \cong H^0(JC,\mathcal{L}_\alpha), \ \ \psi^* : H^0({\mathcal SU}_C(2,\omega \alpha),\mathcal{L}_{\omega \alpha}) \cong H^0(\mathrm{Pic}^{g-1}(C),\mathcal{L}_{\omega \alpha}). \end{equation} There is a well-defined morphism $$ D : {\mathcal SU}_C(2,\alpha) \longrightarrow |\mathcal{L}_{\omega \alpha}| \qquad (\text{resp.} \ D : {\mathcal SU}_C(2,\omega \alpha) \longrightarrow |\mathcal{L}_\alpha|) $$ where the support of $D(E)$ (reduced for $E$ general) is $$ D(E) = \{ \xi \in JC \ (\text{resp.} \ \ \mathrm{Pic}^{g-1}(C)) \: | \: h^0(C, E\otimes \xi) >0 \}. $$ \noindent The two involutions of the Jacobian $JC$ given by $$ T_{\alpha}: \xi \longmapsto \xi \otimes \alpha \qquad (-1): \xi \longmapsto \xi^{-1} $$ induce (up to $\pm 1$) linear involutions $T_{\alpha}^*$ and $(-1)^*$ on the spaces of global sections $H^0(JC,\mathcal{L}_\alpha)$ and $H^0(\mathrm{Pic}^{g-1}(C),\mathcal{L}_{\omega \alpha})$. \begin{lem} The projective linear involutions $T_{\alpha}^*$ and $(-1)^*$ acting on $\mathbb{P} H^0(JC,\mathcal{L}_\alpha)$ are equal. \end{lem} \begin{proof} We observe that the composite map $T_{\alpha} \circ (-1) : \xi \mapsto \alpha \xi^{-1}$ verifies $\psi \circ (T_{\alpha} \circ (-1)) = \psi$. Since $\psi^*$ is a linear isomorphism \eqref{iso1}, we have $(T_{\alpha} \circ (-1))^* = \pm id_{H^0}$. Therefore $T_{\alpha}^* = \pm (-1)^*$. \end{proof} \noindent Thus the two spaces decompose into $\pm$eigenspaces. Note that in order to distinguish the two eigenspaces, we need a lift of the $2$-torsion point $\alpha$ into the Mumford group. We will take the following convention: the $+$eigenspace (resp. $-$eigenspace) contains the Prym varieties $P$ and $P_{even}$ (resp. $P'$ and $P_{odd}$), i.e., we have canonical (up to multiplication by a nonzero scalar) isomorphisms. \begin{equation} \label{iso2} H^0(JC,\mathcal{L}_\alpha)_+ = H^0(P, 2\Xi_0), \qquad H^0(JC,\mathcal{L}_\alpha)_- = H^0(P',2\Xi'_0), \end{equation} \begin{equation} \label{iso3} H^0(\mathrm{Pic}^{g-1}(C),\mathcal{L}_{\omega \alpha})_+ = H^0(P_{even}, 2\Xi), \qquad H^0(\mathrm{Pic}^{g-1}(C), \mathcal{L}_{\omega \alpha})_- = H^0(P_{odd},2\Xi'). \end{equation} \noindent Since the surface $C-C$ is invariant under the involution $(-1) : \xi \mapsto \xi^{-1}$, the subspace $\Gamma^{\alpha}_{C-C}$ is invariant under $(-1)^*$ and decomposes into a direct sum of $\pm$eigenspaces for $(-1)^* = T_{\alpha}^*$: $$ \Gamma^{\alpha}_{C-C} = \Gamma^{\alpha +}_{C-C} \oplus \Gamma^{\alpha -}_{C-C}. $$ \bigskip \noindent {\em Prym-Wirtinger duality} \bigskip For the details see \cite{beau2} lemma 2.3. There exists an integral Cartier divisor on the product ${\mathcal SU}_C(2,\alpha) \times {\mathcal SU}_C(2,\omega \alpha)$ whose support is given by $$ \{ (E,F) \in {\mathcal SU}_C(2,\alpha) \times {\mathcal SU}_C(2,\omega \alpha) \: | \: h^0(C, E \otimes F) >0 \}. $$ Its associated section can be viewed as an element of the tensor product $$ H^0({\mathcal SU}_C(2,\alpha), \mathcal{L}_\alpha) \otimes H^0({\mathcal SU}_C(2,\omega \alpha), \mathcal{L}_{\omega \alpha}) $$ and it can be shown that the corresponding linear map \begin{equation} \label{wirtdual} H^0({\mathcal SU}_C(2,\alpha), \mathcal{L}_\alpha)^* \longrightarrow H^0({\mathcal SU}_C(2,\omega \alpha), \mathcal{L}_{\omega \alpha}) \end{equation} is an isomorphism and is equivariant for the linear involutions induced by the map $E \mapsto E \otimes \alpha$. Hence using the identifications \eqref{iso2} and \eqref{iso3} we obtain canonical isomorphisms, \begin{equation} \label{wd1} H^0(P, 2\Xi_0)^* \map{\sim} H^0(P_{even}, 2\Xi) \qquad H^0(P', 2\Xi'_0)^* \map{\sim} H^0(P_{odd}, 2\Xi'). \end{equation} \section{The base locus of $\mathbb{P} \Gamma_{\tilde{C}}$} In this section we compute the set-theoretical base locus of the subseries $\mathbb{P} \Gamma_{\tilde{C}}$ on the Prym variety $P'$. Suppose $C$ non-hyperelliptic. We need some additional notation: denote by $\tilde{C}_m$ the $m$-th symmetric power of $\tilde{C}$ and let $S$ be the subvariety of $\tilde{C}_{2g-2}$ defined as $$ S = \{ D \in \tilde{C}_{2g-2} \ | \ Nm(D) \in |\omega| \ \text{and} \ h^0(D) \equiv 1 \ \text{mod} \ 2 \}. $$ Then, by \cite{beau} Corollaire page 365, the variety $S$ is normal and irreducible of dimension $g-1$. The variety $S$ comes equipped with two natural surjective morphisms $$ Nm : S \longrightarrow |\omega| \qquad u : S \longrightarrow P_{odd} $$ where $u$ associates to an effective divisor $D$ its line bundle ${\mathcal O}_{\tilde{C}}(D)$. Note that $u$ is birational and $Nm$ is finite of degree $2^{2g-3}$. Also denote by $u$ the extended morphism $u : \tilde{C}_{2g-2} \rightarrow \mathrm{Pic}^{2g-2}(\tilde{C})$ and consider the commutative diagram \begin{equation} \label{comdiags} \begin{array}{ccc} S & \hookrightarrow & \tilde{C}_{2g-2} \\ & & \\ \downarrow^u & & \downarrow^u \\ & & \\ P_{odd} & \hookrightarrow & \mathrm{Pic}^{2g-2}(\tilde{C}). \end{array} \end{equation} Consider the Brill-Noether locus in $P_{odd}$ which is defined set-theoretically by $$ \Xi_3 := \{ \lambda \in P_{odd} \ | \ h^0(\lambda) \geq 3 \}. $$ The scheme structure on $\Xi_3$ is defined by taking the scheme-theoretical intersection \cite{welt2} $$ \Xi_3 := W^2_{2g-2}(\tilde{C}) \cap P_{odd} $$ where $W^2_{2g-2}(\tilde{C}) \subset Pic^{2g-2}\tilde{C}$ is the Brill-Noether locus of line bundles having at least $3$ sections (see \cite{acgh}). \begin{lem} The subscheme $\Xi_3 \subset P_{odd}$ is not empty and is of pure codimension $3$. \end{lem} \begin{proof} Theorem 9 \cite{dcp} asserts that $\Xi_3$ is not empty and every irreducible component has dimension at least $g-4$. Suppose that there is an irreducible component $I$ of dimension $\geq g-3$. Then its inverse image $u^{-1}(I)$ has dimension $\geq g-1$, hence, since $S$ is irreducible, $u^{-1}(I) = S$ and $\Xi_3 = P_{odd}$. The last equality can not happen, since otherwise, using translation by an element of the form ${\mathcal O}_{\tilde{C}}(\tilde{p} - \sigma \tilde{p})$, we would have $\Xi = P_{even}$. \end{proof} Observe that $u$ is equivariant for the action of $\sigma$ on $S$ and $P_{odd}$. Denote by $Z = u^{-1}(\Xi_3)$ the inverse image of the subscheme $\Xi_3$. By the previous lemma $Z$ is of pure codimension $1$ in $S$. We will see in a moment that there is a Cartier divisor $\mathcal{D}$ on $S$ whose support is the support of $Z$. Let $\omega_{\tilde{C}}$ be the dualizing sheaf of $\tilde{C}$. Consider the following divisors in $\tilde{C}_{2g-2}$ \[ U_{\tilde{p}} := \{ D \in \tilde{C}_{2g-2} \ | \ \exists D'\in \tilde{C}_{ 2g-3 } \text{ with } D = D'+\tilde{p}\} = \tilde{p} + \tilde{C}_{2g-3} \] \[ V_{\tilde{p}} := \{ D \in \tilde{C}_{2g-2} \ | \ h^0(\omega_{\tilde{C}}(-D-\tilde{p})) \geq 1 \} \] and let $\bar{U}_{\tilde{p}}$ and $\bar{V}_{\tilde{p}}$ be their intersections with $S$. A straightforward calculation involving Zariski tangent spaces then shows that $\bar{U}_{\tilde{p}}$ is a reduced divisor. We denote by ${\mathcal O}_S(1)$ the pull-back by the norm map of the hyperplane line bundle on $|\omega|$. Then it is easily seen that, for any $\tilde{p} \in \tilde{C}$, \begin{equation} \label{pbnm} Nm^*(|\omega(-p)|) = \bar{U}_{\tilde{p}} + \bar{U}_{\sigma \tilde{p}} \in |{\mathcal O}_S(1)|. \end{equation} Let $\tilde{\Theta}_{\lambda}$ denote the translate of $\tilde{\Theta}$ by $\lambda$. Then, for any points $\tilde{p}, \tilde{q} \in \tilde{C}$, we have an equality among divisors on $\tilde{C}_{2g-2}$ (see \cite{welt1} page 6) \begin{equation} \label{pbthetapq} u^*(\tilde{\Theta}_{\tilde{p} - \tilde{q}}) = U_{\tilde{p}} + V_{\tilde{q}}. \end{equation} The analogue on the even Prym variety of the following lemma was previously proved by R. Smith and R. Varley. In the case of genus $3$ it is in their paper \cite{smithvarley99} (Prop. 1 page 358) and for higher genus it will be published in their upcoming paper \cite{smithvarley00}. \begin{lem} There exists an effective Cartier divisor $\mathcal{D}$ on $S$ whose support is equal to $$ \text{supp} \ Z = \{ D \in S \ | \ h^0({\mathcal O}_{\tilde{C}}(D)) \geq 3 \}. $$ Moreover, we have the following equality among effective Cartier divisors \begin{equation} \label{eqxiD} u^* (\Xi_{\tilde{p} -\sigma\tilde{p}} + \Xi_{\sigma\tilde{p} -\tilde{p}} ) = \bar{U}_{\tilde{p}} + \bar{U}_{\sigma \tilde{p}} + \mathcal{D} \ \ \ \forall \tilde{p} \in \tilde{C}. \end{equation} In particular, $u^* {\mathcal O}_{P_{odd}}(2\Xi') = {\mathcal O}_S(1) \otimes {\mathcal O}_S(\mathcal{D})$. \end{lem} \begin{proof} We are going to define $\mathcal{D}$ as the residual divisor of the restricted divisor $\bar{V}_{\tilde{q}}$, for a given point $\tilde{q} \in \tilde{C}$ and then show that it does not depend on the choice of $\tilde{q}$. We first observe that we have an equality of sets $$\bar{V}_{\tilde{q}} = \bar{U}_{\sigma \tilde{q}} \cup Z $$ which can be seen as follows: for $D\in\tilde{C}_{ 2g-2 }$ such that $h^0 (D) = h^0(\omega_{\tilde{C}}(-D)) = 1$ the assumption $D \in \bar{V}_{\tilde{q}}$ and the formula $D + \sigma D = \pi^* (Nm (D))$ imply that $ \tilde{q} \in \text{supp} \ \sigma D \ \iff \ D \in \bar{U}_{\sigma\tilde{q}}$ . If $h^0(D) = h^0(\omega_{\tilde{C}}(-D)) \geq 2$, then $D \in \text{supp}\ Z$. Again a calculation involving Zariski tangent spaces shows that $\bar{V}_{\tilde{q}}$ is reduced generically on $\bar{U}_{\sigma \tilde{p}}$. Hence we can define $\mathcal{D}$ by $\bar{V}_{\tilde{q}} = \bar{U}_{\sigma \tilde{q}} + \mathcal{D}$. Now we substitute this expression into \eqref{pbthetapq}, which we restrict to $S$ $$ u^*(\tilde{\Theta}_{\tilde{p} - \tilde{q}})_{|S} = \bar{U}_{\tilde{p}} + \bar{U}_{\sigma \tilde{q}} + \mathcal{D}.$$ Now we fix $\tilde{q}$ and we take the limit when $\tilde{p} \rightarrow \tilde{q}$. Since ${\mathcal O}_{P_{odd}}(\tilde{\Theta}) = {\mathcal O}_{P_{odd}}(2\Xi')$, we see that $\bar{U}_{\tilde{p}} + \bar{U}_{\sigma \tilde{p}} + \mathcal{D} \in |u^*{\mathcal O}_{P_{odd}} (2\Xi')|$. So by \eqref{pbnm} we get the line bundle equality claimed in the lemma and we see that the scheme-structure on $\mathcal{D}$ does not depend on the point $\tilde{q}$. To prove \eqref{eqxiD}, we compute using \eqref{pbthetapq} $$u^* (\tilde{\Theta}_{\tilde{p} -\sigma\tilde{p}} + \tilde{\Theta}_{\sigma\tilde{p} -\tilde{p}}) = \bar{U}_{\tilde{p}} + \bar{V}_{\sigma \tilde{p}} + \bar{U}_{\sigma \tilde{p}} + \bar{V}_{\tilde{p}}.$$ Now we restrict to $S$ and use the commutativity of diagram \eqref{comdiags} and the divisorial equality $\tilde{\Theta}_{\tilde{p} - \sigma \tilde{p}} \cap P_{odd} = 2 \Xi_{\tilde{p} - \sigma \tilde{p}}$ to obtain $$u^* (2\Xi_{\tilde{p} -\sigma\tilde{p}} + 2\Xi_{\sigma\tilde{p} -\tilde{p}}) = 2\bar{U}_{\tilde{p}} + 2\bar{U}_{\sigma \tilde{p}} + 2\mathcal{D}.$$ Since $\bar{U}_{\tilde{p}} + \bar{U}_{\sigma \tilde{p}} + \mathcal{D} \in |u^*{\mathcal O}_{P_{odd}} (2\Xi')|$ we can divide this equality by $2$ and we are done. \end{proof} Let $\mu$ be a point of $Bs(\mathbb{P} \Gamma_{\tilde{C}})$. By lemma \ref{lemtaui} the linear map $i^* : |\omega|^* \longrightarrow |2 \Xi'_0|^*$ is injective and, since $|\omega|^*$ is the span of the image of $\tilde{C}$ in $|2\Xi'_0|^*$, the space $\mathbb{P} \Gamma_{\tilde{C}}$ is the annihilator of $|\omega|^* \subset |2\Xi'_0|^*$. So $Bs(\mathbb{P} \Gamma_{\tilde{C}}) = |\omega |^*\cap Kum(P')$ and $\mu$ corresponds to a hyperplane $H_\mu \in |\omega|^*$. Since $\mu\in Kum(P')$, the image of $\mu$ by Wirtinger duality is the divisor $\Xi_{\mu} +\Xi_{\mu^{ -1}}\in |2\Xi' |$. \begin{lem} With the previous notation, we have an equality \begin{equation} \label{pbhmu} \forall \mu \in Bs(\mathbb{P} \Gamma_{\tilde{C}}) \qquad Nm^*(H_\mu) + \mathcal{D} = u^*(\Xi_\mu + \Xi_{\mu^{-1}}). \end{equation} \end{lem} \begin{proof} The equality follows from the commutativity of the right-hand square of the diagram $$ \begin{array}{ccccccccc} \tilde{C} & \map{\pi} & C & \map{\varphi_{can}} & |\omega|^* & \map{Nm^*} & |{\mathcal O}_S(1)| & \map{+\mathcal{D}} & |{\mathcal O}_S(1) \otimes {\mathcal O}_S(\mathcal{D})| \\ & & & & & & & & \\ \downarrow^i & & & & \downarrow^{i^*} & & & \nearrow^{u^*} & \\ & & & & & & & & \\ P' & & \longrightarrow & & |2\Xi'_0|^* & \cong & |2\Xi'|. & & \end{array} $$ The commutativity of the right-hand square follows from that of the outside square because $\varphi_{can}(C)$ generates $|\omega|^*$. In other words we need to check the assertion of the lemma only for hyperplanes of the form $|\omega(-p)|$ for $p \in C$. This follows immediately from \eqref{pbnm} and \eqref{eqxiD}. \end{proof} \begin{cor} For every $\mu\in Bs(\mathbb{P} \Gamma_{\tilde{C}})$, the hyperplane $Nm^* (H_{\mu })$ is reducible. \end{cor} \begin{proof} By the above Lemma we have \[ u^*(\Xi_\mu + \Xi_{\mu^{-1}}) -\mathcal{D} = Nm^*(H_\mu). \] If $Nm^*(H_\mu)$ is irreducible, then the support of one of the divisors $u^*(\Xi_\mu )$ or $u^*(\Xi_{\mu^{-1}})$, say $u^*(\Xi_\mu )$, is contained in the support of $\mathcal{D}$. This is impossible because $u^*(\Xi_\mu )$ is the inverse image of a divisor in $P_{ odd}$ and $supp \mathcal{D}$ is the inverse image of the codimension $3$ support of $\Xi_3$. \end{proof} The set-theoretical assertion of theorem \ref{mainthm1} now follows from the following lemma. \begin{lem}\label{lemHred} If $C$ is not bi-elliptic, we have a set-theoretical equality $$ \{ H \in |\omega|^* \ : \ Nm^*(H) \ \text{is reducible} \} = \varphi_{can}(C). $$ If $C$ is bi-elliptic, the LHS is contained in the union of $\varphi_{can}(C)$ and the finite set of points $t\in |\omega|^*$ such that the projection from $t$ induces a morphism of degree $2$ from $C$ onto an elliptic curve. \end{lem} \begin{proof} Suppose that $Nm^*(H)$ is reducible. Then a local computation shows that the hyperplane $H$ is everywhere tangent to the branch locus of $Nm$. It is immediately seen that the branch locus $B$ of $Nm$ is the dual hypersurface of the canonical curve. The components of the singular locus $Sing(B)$ of $B$ are of two different types which can be described as follows \begin{description} \item[type 1] whose points are hyperplanes tangent to $\varphi_{can}(C)$ in more than one point. \item[type 2] whose points are hyperplanes osculating to $\varphi_{can}(C)$. \end{description} To prove that $\mu\in\varphi_{can}(C)$, we need to prove that there is a point on $H \cap B$ which is smooth on $B$ because the dual variety of $B$ is the closure of the set of hyperplanes tangent to $B$ at a smooth point and this is equal to $\varphi_{can}(C)$. In other words we need to show that $H \cap B$ is not contained in $Sing (B)$. Since $H\cap B$ has pure codimension $2$, it suffices to show that no codimension $2$ component of $Sing (B)$ is contained in a hyperplane. Suppose a codimension $2$ component $B_i$ of type $i$ ($i=1$ or $2$) is contained in a hyperplane $H$ in $|\omega |$ and let $t\in |\omega |^*$ be the corresponding point. Then the set of hyperplanes in $|\omega |^*$ through $t$ and doubly tangent (resp. osculating) to $\varphi_{ can }(C)$ has dimension $g-3$. We have \begin{lem} For any $t\in\varphi_{ can }(C)$ the restriction $\rho$ of the projection from $t$ to $\varphi_{ can }(C)$ is birational onto its image. If $t\in |\omega |^*\setminus\varphi_{ can }(C)$, then $\rho$ is either birational onto its image or of degree two onto an elliptic curve. \end{lem} \begin{proof} First note that the degree of the image $C_t$ of $C$ by the projection is at least $g-2$ because $C_t$ is a non-degenerate curve in a projective space of dimension $g-2$. If $t\in\varphi_{ can }(C)$, then the degree of $\rho$ is equal to $2g-3$. The degree $r$ of the restriction of $\rho$ to $C_t$ verifies $r\cdot\hbox{deg}(C_t) = 2g-3$. Therefore $\frac{2g-3 }{r}\geq g-2$. Or $r\leq 2 +\frac{1 }{g-2 }$ which implies $r\leq 2$. However, $r$ cannot be equal to $2$ because $2g-3$ is odd. If $t\not\in\varphi_{ can }(C)$, then the same argument gives again $r\leq 2$ because $g\geq 5$. Hence, if $\rho$ is not generically injective, then $r =2$ and $\hbox{deg} (C_t ) = g-1$. Therefore $C_t$ is either smooth rational or an elliptic curve. Since $C$ is not hyperelliptic, we have that $C_t$ is an elliptic curve. \end{proof} First suppose that $C\rightarrow C_t$ is birational. If $i=1$, projecting from $t$, we see that the set of hyperplanes in $|\omega |^*/t$ doubly tangent to $C_t$ has dimension $(g-3)$=dimension of the dual variety of $C_t$ which is impossible. If $i=2$, then the set of hyperplanes in $|\omega |^*/t$ osculating to $C_t$ has dimension $g-3$ which is also impossible. If $C\rightarrow C_t$ is of degree $2$, then indeed every hyperplane tangent to $C_t$ pulls back to a hyperplane twice tangent (or osculating if the point of tangency is a branch point of $C\rightarrow C_t$) to $\varphi_{ can } (C)$ and we have a codimension $2$ family of type $B_1$ contained in the hyperplane corresponding $H$ to $t$. Then $Nm^* (H)$ could be reducible. \end{proof} The previous lemma proves theorem 1 set-theoretically for a non bi-elliptic curve. In the bi-elliptic case, we have to work a little more. By Lemma \ref{lemHred} a hyperplane $H \not\in \varphi_{can}(C)$, such that $Nm^*(H)$ might be reducible, corresponds to a point $e \in |\omega|^*$ such that the projection from $e$ induces a morphism $\gamma$ of degree $2$ from $C$ to an elliptic curve $E$. In other words, $e$ is the common point of all chords $\langle \gamma^* q \rangle ; \ (q \in E)$. In that case there exists a 1-dimensional family (parametrized by $E$) of trisecants, namely the chords $\langle \gamma^*q \rangle$, to the Kummer variety $Kum(P')$. By \cite{debarre} the Prym variety is a Jacobian and by \cite{sho} (see also \cite{beau3} page 610) the double cover $\pi : \tilde{C} \rightarrow C$ is of the following two types \begin{enumerate} \item $C$ is trigonal \item $C$ is a smooth plane quintic and $h^0({\mathcal O}_C(1)\otimes \alpha) = 0$ \end{enumerate} \begin{lem} No double cover of a bi-elliptic curve $C$ of genus $g \geq 6$ is of the above two types. \end{lem} \begin{proof} For a bi-elliptic curve $C$, the Brill-Noether locus $W^1_{g-1}(C)$ has two irreducible components, which are fixed by the reflection in $\omega$ (\cite{welt1} Corollary 3.10). For a smooth plane quintic this Brill-Noether locus is irreducible, ruling out 1. For a trigonal curve this Brill-Noether locus has two irreducible components, which are interchanged by reflection in $\omega$, ruling out 2. \end{proof} \begin{rem} If $g = 5$ and $C$ is bi-elliptic, we do not know whether the common point of all the chords for a given bi-elliptic structure lies on $Kum(P')$ (see also \cite{beau3} Remark (1) page 611). We expect it not to be on $Kum(P')$. \end{rem} \section{Rank $2$ bundles and $2\Xi$-divisors} Consider the induced action of the involution $\sigma$ on the moduli space $\mathcal{SU}_{\tilde{C}}(2,{\mathcal O})$ given by $\tilde{E}\mapsto \sigma^* \tilde{E}$. Since the covering $\pi$ is unramified, the fixed point set for the $\sigma$-action $$Fix_{\sigma} \mathcal{SU}_{\tilde{C}}(2,{\mathcal O}) = \{ [\tilde{E}] \in \mathcal{SU}_{\tilde{C}}(2,{\mathcal O}) \ | \ \exists \theta : \ \sigma^* \tilde{E} \map{\sim} \tilde{E} \} $$ has two connected components which are the isomorphic images of ${\mathcal SU}_C(2,{\mathcal O})$ and ${\mathcal SU}_C(2,\alpha)$ by $\pi^*$. Similarly, since $\sigma^* \omega_{\tilde{C}} \map{\sim} \omega_{\tilde{C}}$, the involution $\sigma$ acts on $\mathcal{SU}_{\tilde{C}}(2,\omega_{\tilde{C}})$ and $$Fix_{\sigma} \mathcal{SU}_{\tilde{C}}(2,\omega_{\tilde{C}}) = \pi^*{\mathcal SU}_C(2,\omega) \cup \pi^*{\mathcal SU}_C(2,\omega \alpha) $$ \begin{prop}\label{propDelta} Consider a bundle $E \in {\mathcal SU}_C(2,\omega \alpha)$ such that $E \not\in \varphi(P_{odd})$ and put $\tilde{E} = \pi^* E$. Then there is a divisor $\Delta(E) \in |2\Xi_0|$ with the following properties. \begin{enumerate} \item If $D(\tilde{E})$ does not contain $P$, then \[ D(\tilde{E}) = 2\Delta(E). \] For $E$ general, $P$ is not contained in $D(\tilde{E})$ and $\Delta(E)$ is reduced. \item Let $pr_+$ be the projection $|\mathcal{L}_\alpha|\rightarrow |2\Xi_0|$ with center $| 2\Xi_0 '|$ (see (\ref{iso2})). Then we have a commutative diagram $$ \begin{array}{ccc} {\mathcal SU}_C(2,\omega \alpha) & \map{D} & |\mathcal{L}_\alpha| \\ & & \\ & \searrow^{\Delta} & \downarrow^{pr_+} \\ & & \\ & & |2\Xi_0| = |\mathcal{L}_\alpha|_+ \end{array} $$ \end{enumerate} \end{prop} \begin{rem} Similarly, when $E\in{\mathcal SU}_C(2,\omega \alpha)$ such that $E\not\in\varphi(P_{even})$, we get divisors $\Delta(E) \in |2\Xi'_0|$ as described in prop. 3.2 by projecting on the $-$eigenspace $pr_- : |\mathcal{L}_\alpha| \longrightarrow |\mathcal{L}_\alpha|_- = |2\Xi'_0|$. \end{rem} \begin{proof} 1. Given a bundle $F \in Fix_\sigma \mathcal{SU}_{\tilde{C}}(2,\omega_{\tilde{C}})$ and a line bundle $\xi \in J\tilde{C}$ which is anti-invariant under $\sigma$, i.e., $\sigma^* \xi \map{\sim} \xi^{-1}$, we have a natural non-degenerate quadratic form with values in the canonical bundle $\omega_{\tilde{C}}$ \begin{eqnarray*} q: F \otimes \xi & \longrightarrow & \omega_{\tilde{C}} \\ s & \longmapsto & s \wedge \sigma^* s \end{eqnarray*} where $s$ is a local section of $F \otimes \xi$. Note that we have canonical isomorphisms $$ \sigma^* (F\otimes \xi) = F \otimes \xi^{-1} = \mathrm{Hom}(F\otimes \xi,\omega_{\tilde{C}}) $$ Therefore we are in a position to apply the Atiyah-Mumford lemma \cite{mum} to the family of bundles (here $F$ is fixed, with $\sigma^* F \map{\sim} F$) $$ \{F \otimes \xi \}_{\xi \in P }$$ which states that the parity of $h^0(\tilde{C}, F\otimes \xi)$ is constant when $\xi$ varies in $P$. \bigskip \noindent >From now on, we suppose $F=\tilde{E} = \pi^*E$, with $E \in {\mathcal SU}_C(2,\omega \alpha)$, then $$ h^0(\tilde{C}, \tilde{E}) = 2h^0(C,E) \equiv 0 \ \ \text{mod} \ \ 2. $$ For the first equality we use the fact that $H^0(\tilde{C},\tilde{E}) = H^0(C,E) \oplus H^0(C, E\alpha)$ and, by Riemann-Roch and Serre duality, $h^0(C,E) = h^1 (C, E) = h^0 (C ,\omega\otimes E^* ) = h^0 (C, E\alpha )$. First suppose that $E\in{\mathcal SU}_C(2,\omega \alpha)$ is general. Then the divisor $D(\tilde{E})$ does not contain the Prym variety $P$ (e.g. because, for general $E$, $h^0(E)=0 \ \iff \ h^0 (\tilde{E} )=0 \ \iff \ {\mathcal O} \not\in D(\tilde{E})$), so the restriction of the divisor $D(\tilde{E}) \in |2 \Theta_{\tilde{C}}|$ to $P$ is a divisor in the linear system $|4\Xi_0|$. Moreover, for $\xi \in D(\tilde{E})\cap P$ $$\mathrm{mult}_\xi D(\tilde{E}) \geq h^0(\tilde{C}, \tilde{E} \otimes \xi) \geq 2 $$ because $h^0(\tilde{C}, \tilde{E} \otimes \xi) \equiv h^0(\tilde{C}, \tilde{E} ) \equiv 0 \ \ \text{mod} \ \ 2$. Hence any point $\xi \in D(\tilde{E})\cap P$ is a singular point of $D(\tilde{E})$, which implies that $D(\tilde{E}) \cap P$ is an everywhere non-reduced divisor. We have \begin{lem} Suppose that $D(\tilde{E}) \cap P$ is a divisor in $P$. Then there is a divisor $\Delta(E) \in |2\Xi_0|$ such that $D(\tilde{E}) \cap P = 2 \Delta(E)$. \end{lem} \begin{proof} A local equation of $\Delta(E)$ is given by the pfaffian of a skew-symmetric perfect complex of length one $L \longrightarrow L^*$ representing the perfect complex $Rpr_{1*}(\mathcal{P} \otimes pr^*_2 \tilde{E})$ where $\mathcal{P}$ is the Poincar\'e line bundle over the product $P \times \tilde{C}$ and $pr_1, pr_2$ are the projections on the two factors. The construction of the complex $L \longrightarrow L^*$ is given in the proof of Proposition 7.9 \cite{ls}. \end{proof} If $E$ is of the form $E = \pi_* L$ for some $L \in P_{even}$, we have $\Delta(E) = T^*_L \Xi + T^*_{\omega L^{-1}} \Xi$. It follows from this equality that $\Delta(E)$ is reduced for general $E$. So far we have defined a rational map $\Delta: {\mathcal SU}_C(2,\omega \alpha) \longrightarrow |2\Xi_0|$. It will follow from part 2 of the proposition that $\Delta$ can be defined away form $\varphi(P_{odd})$. \bigskip \noindent 2. First we consider the composite (rational) map $$ \mathrm{Pic}^{g-1}(C) \map{\psi} {\mathcal SU}_C(2,\omega \alpha) \map{\Delta} |2\Xi_0|. $$ A straight-forward computation shows that for all $\xi \in \mathrm{Pic}^{g-1}(C)$ such that $\pi^*\xi\not\in P_{odd}$ the divisor $\Delta(\psi(\xi)) = \Delta(\xi \oplus \omega \alpha \xi^{-1})$ equals the translated divisor $T_{\pi^* \xi} \tilde{\Theta}$ restricted to $P$. Hence, by \cite{mum2}, the map $\Delta \circ \psi$ is given by the full linear system $|\mathcal{L}_{\omega \alpha}|_+$ of invariant elements of $|\mathcal{L}_{\omega \alpha}|$. By Prym-Wirtinger duality \eqref{wirtdual} and \eqref{wd1} $|\mathcal{L}_{\omega \alpha}|^*_+ \cong |\mathcal{L}_\alpha|_+ \cong |2\Xi_0|$ and we obtain the commutative diagram in the proposition. Geometrically, $\Delta$ is obtained by restricting the projection with center the $-$eigenspace $|\mathcal{L}_\alpha|_-$ to the embedded moduli space ${\mathcal SU}_C(2,\omega \alpha) \subset |\mathcal{L}_\alpha|$. Since by \cite{nr} $|\mathcal{L}_\alpha|_- \cap {\mathcal SU}_C(2,\omega \alpha) = \varphi(P_{odd})$ we see that $\Delta$ is well-defined for $E\not\in\varphi(P_{odd})$ even if $D(\tilde{E})\supset P$. \end{proof} \begin{rem} We observe that we obtain by the same construction a rational map $$ \Delta : {\mathcal SU}_C(2,\omega) \longrightarrow |2 \Xi_0|\; . $$ The images under $\Delta$ of the two moduli spaces ${\mathcal SU}_C(2,\omega)$ and ${\mathcal SU}_C(2,\omega \alpha)$ coincide, which is easily deduced from the following formula. Let $\beta$ be a $4$-torsion point such that $\beta^{\otimes 2} = \alpha$ and $\pi^* \beta \in P[2]$. Then, for any $E \in {\mathcal SU}_C(2,\omega)$, we have $E \otimes \beta \in {\mathcal SU}_C(2,\omega \alpha)$ and $$ T_{\pi^* \beta}^* \Delta(E) = \Delta(E\otimes \beta)\; . $$ \end{rem} \noindent Similar statements hold for ${\mathcal SU}_C(2,\alpha)$. \section{Proof of theorem 2} \subsection{Proof of $\Gamma^{\alpha +}_{C-C} = \Gamma_{00}$} The strategy is to show that the two linear maps $$\phi_1 : H^0(P,2\Xi_0)_0 \longrightarrow \mathrm{Sym}^2 T_0^*P = \mathrm{Sym}^2 H^0(\omega \alpha)$$ and $$\phi_2 : H^0(JC,\mathcal{L}_\alpha)_{+0} \longrightarrow H^0(C\times C, \delta^* \mathcal{L}_\alpha - 2\Delta)_+ = \mathrm{Sym}^2 H^0(\omega \alpha)$$ differ by multiplication by a scalar under the isomorphism \eqref{iso2} $H^0(JC,\mathcal{L}_\alpha)_{+0 }\cong H^0(P,2\Xi_0)_0$. Here the subscript $0$ denotes the subspace (on $P$ or $JC$) consisting of global sections vanishing at the origin. The map $\phi_1$ sends $s\in H^0(P,2\Xi_0)_0$ to the quadratic term of its Taylor expansion at the origin ${\mathcal O} \in P$ and $\phi_2$ is the pull-back of invariant sections of $\mathcal{L}_\alpha$ under the difference map \begin{eqnarray*} \delta : C \times C & \longrightarrow & JC \\ (p,q) & \longmapsto & {\mathcal O}_{C}(p-q). \end{eqnarray*} By restricting to the fibers of the two projections $p_i :C\times C\rightarrow C$ and using the See-saw Theorem, we compute that $\delta^* \mathcal{L}_\alpha = p_1^* (\omega \alpha) \otimes p_2^* (\omega \alpha )(2 \Delta_C)$ where $\Delta_C\subset C\times C$ is the diagonal. Since $\phi_2^{ -1} (0) =\Delta_C$ and the sections of ${\mathcal L}_{\alpha }$ are symmetric, we see that $\mathrm{im} \: \phi_2 \subset \mathrm{Sym}^2 H^0(\omega \alpha )\subset H^0 (\omega\alpha )^{\otimes 2} = H^0 (p_1^* (\omega \alpha) \otimes p_2^* (\omega \alpha ))\subset H^0 (p_1^* (\omega \alpha) \otimes p_2^* (\omega \alpha ) (2\Delta_C ))$. So if $\phi_1$ and $\phi_2$ are proportional, we will have $$ \Gamma_{00} = \mathrm{ker} \: \phi_1 = \mathrm{ker} \: \phi_2 = \Gamma^{\alpha +}_{C-C}. $$ \noindent To show that $\phi_1 = \lambda\phi_2$ for some $\lambda\in\mathbb{C}^*$, we compute $\phi_1(s_E)$ and $\phi_2(s_E)$ for special sections, namely those with divisor of zeros $Z(s_E) = \Delta(E)$ for some vector bundle $E \in {\mathcal SU}_C(2,\omega \alpha)$ with $h^0(E) = h^0(E \otimes \alpha) = 2$. Recall that by Riemann-Roch and Serre duality we have for $h^0(E) = h^0(E \otimes \alpha)$ for $E \in {\mathcal SU}_C(2,\omega \alpha)$. Now to compute $\phi_1(s_E)$, we need to determine the tangent cone to $\Delta(E)$ at ${\mathcal O} \in P$. As before we put $\tilde{E} = \pi^*E$. By \cite{laszlo} prop. V.2, this tangent cone is the intersection of the anti-invariant part $H^0(\omega \alpha) = H^0(\omega_{\tilde{C}})_-$ of $H^0(\omega_{\tilde{C}}) = T^*_0 J\tilde{C}$ with the affine cone over the projective cone over the Grassmannian $Gr(2, H^0(\tilde{E})^*) \subset \mathbb{P} \Lambda^2 H^0(\tilde{E})^*$ under the linear map \begin{equation} \label{mapmu} \mu^* : H^0(\omega_{\tilde{C}})^* \longrightarrow \Lambda^2 H^0(\tilde{E})^* \end{equation} which is the dual of the map $\mu :\Lambda^2 H^0(\tilde{E})\rightarrow H^0(\omega_{\tilde{C}})$ obtained from exterior product by the isomorphism $\wedge^2\tilde{E}\cong\omega_{\tilde{C} }$. Note that the $\sigma$-invariant part $[\Lambda^2 H^0(\tilde{E})^*]_+$ is canonically isomorphic to the $2$-dimensional subspace $\Lambda^2 H^0(E)^* \oplus \Lambda^2 H^0(E\alpha)^* \subset \Lambda^2 H^0(\tilde{E})^*$ because $H^0(\tilde{E})_+ = H^0(E)$ and $H^0(\tilde{E})_- = H^0(E \alpha)$. Since $\wedge^2 E\cong\wedge^2 (E\otimes\alpha )\cong\omega\alpha$, the restriction of $\mu$ to $\wedge^2 H^0 (E)$ (resp. $\wedge^2 H^0 (E\otimes\alpha )$) which is obtained from exterior product by the isomorphism $\wedge^2 E\cong\omega\alpha$ (resp. $\wedge^2 (E\otimes\alpha )\cong\omega\alpha$) maps into $H^0 (\omega\alpha )$. Therefore the linear map $\mu^*$ \eqref{mapmu} maps $\sigma$-anti-invariant sections into $\sigma$-invariant sections, i.e., \begin{equation} \label{mapmuplus} \mu^*_+ : H^0(\omega \alpha)^* \longrightarrow \Lambda^2 H^0(E)^* \oplus \Lambda^2 H^0(E\alpha)^*. \end{equation} Since the intersection $\mathbb{P} (\Lambda^2 H^0(E)^* \oplus \Lambda^2 H^0(E\alpha)^*) \cap Gr(2,H^0(\tilde{E})^*)$ consists of the two points $\mathbb{P} \Lambda^2 H^0(E)^*$ and $\mathbb{P} \Lambda^2 H^0(E\alpha)^*$, it follows that the intersection of $H^0 (\omega\alpha )\subset H^0 (\omega_{\tilde{C} })$ with the cone over $Gr(2, H^0(\tilde{E})^*)$ is the union of the two lines $\wedge^2 H^0 (E)$ and $\wedge^2 H^0 (E\otimes\alpha )$. Therefore the tangent cone of $\Delta(E)$ at the origin is the union of the two hyperplanes in $|\omega \alpha|^*$ which are the zeros of $a,b \in H^0(\omega \alpha)$ such that \begin{equation} \label{defab} a \mathbb{C} = \mathrm{im} \: (\Lambda^2 H^0(E) \longrightarrow H^0(\omega \alpha)) \qquad b \mathbb{C} = \mathrm{im} \: (\Lambda^2 H^0(E\alpha) \longrightarrow H^0(\omega \alpha)). \end{equation} In other words, up to multiplication by a nonzero scalar, $$ \phi_1(s_E) = a \otimes b + b \otimes a \in \mathrm{Sym}^2 H^0(\omega \alpha). $$ \noindent We now compute $\phi_2(s_E)$. First we note that the pull-back map induced by $\delta$ is equivariant for the involution $(-1) : \xi \mapsto \xi^{-1}$ acting on $JC$ and the involution $(p,q) \mapsto (q,p)$ acting on $C \times C$. Since $\Delta(E) = pr_+ (D(E))$ by Proposition \ref{propDelta}, this implies that \begin{equation} \label{comprdelta} \phi_2 (s_E) =\phi_2(pr_+(s_E)) = pr_+ (\delta^* (s_E)) \end{equation} On the RHS $pr_+$ denotes the projection $H^0(\omega \alpha) \otimes H^0(\omega \alpha) \longrightarrow \mathrm{Sym}^2 H^0(\omega \alpha)$. Therefore we compute $$\delta^* (D(E)) = \{ (p,q) \in C \times C \: | \: h^0(E(p-q))>0 \} $$ and take its symmetric part. It follows from \cite{geem-iz} lemma 3.2 that \begin{equation} \label{pbdelta} \delta^* (D(E)) = C \times Z_a + Z_b \times C + 2 \Delta_C \end{equation} where $Z_a$ (resp. $Z_b$) is the divisor of zeros of $a$ (resp. $b$). Hence it follows from \eqref{comprdelta} and \eqref{pbdelta} that $\phi_2(pr_+(s_E)) = a \otimes b + b \otimes a$ up to multiplication by a nonzero scalar. We can now conclude that $\phi_1 = \lambda\phi_2$ for some $\lambda\in\mathbb{C}^*$ because, by the following lemma (prop. 3.7 \cite{geem-iz}), we have enough bundles $E \in {\mathcal SU}_C(2,\omega \alpha)$ with $h^0(E) = 2$ to generate linearly the image $\mathrm{Sym}^2 H^0(\omega \alpha)$ of $\phi_1$ and $\phi_2$. \begin{lem}(prop. 3.7 \cite{geem-iz}) For general sections $a,b \in H^0(\omega \alpha)$, we can find a semi-stable bundle $E \in {\mathcal SU}_C(2,\omega \alpha)$ with $h^0(E)=2$ such that \eqref{pbdelta} holds. \end{lem} \subsection{Proof of $\Gamma^{\alpha -}_{C-C} = \Gamma^{(2)}_{\tilde{C}}$} First note that any anti-invariant section of $\mathcal{L}_\alpha$ vanishes at ${\mathcal O} \in JC$. Denote by $$ \tau : H^0(JC,\mathcal{L}_\alpha)_- \longrightarrow T_0^*JC = H^0(\omega) $$ the map which sends an element $s$ of $H^0(JC,\mathcal{L}_\alpha)_-$ to the linear term of its Taylor expansion at the origin (Gauss map). Recall the natural embedding of the curve $\tilde{C}$ into the Prym variety $P'$ \begin{equation} \label{embtc} i: \tilde{C} \longrightarrow P' \qquad \tilde{p} \longmapsto {\mathcal O}_{\tilde{C}}(\tilde{p} - \sigma \tilde{p}). \end{equation} Then $i^* {\mathcal O}(2\Xi'_0)\cong \omega_{\tilde{C}}$ and since all $2\Xi'_0$-divisors are symmetric and $i$ is equivariant for the involution, $i$ induces a linear map \begin{equation} \label{pbi} i^* : H^0(P',2 \Xi'_0) \longrightarrow H^0(C, \omega) = H^0(\tilde{C}, \omega_{\tilde{C}})_+ \end{equation} \begin{lem}\label{lemtaui} The linear maps $\tau$ and $i^*$ are proportional via the isomorphism \eqref{iso2} and are surjective. \end{lem} \begin{proof} It will be enough to show that the canonical divisors $i^*(\Delta(\pi_* \lambda))$ and $\tau(D(\pi_* \lambda))$ are equal for a general element $\lambda \in P_{odd}$. In both cases the divisor coincide with the divisor $Nm(\delta)$, where $\delta$ is the unique effective divisor in the linear system $|\lambda|$. The computations are straight-forward and left to the reader. \end{proof} \bigskip \noindent Therefore we can conclude that $$H^0(JC, \mathcal{L}_\alpha)_{0-}^{(3)} = \mathrm{ker} \: \tau = \mathrm{ker} \: i^* = \Gamma_{\tilde{C}}.$$ where $H^0(JC, \mathcal{L}_\alpha)_{0-}^{(3)}$ denotes the subspace of $H^0 ( JC,\mathcal{L}_\alpha )_-$ of elements with multiplicity $\geq 2$ (hence $\geq 3$ by anti-symmetry) at the origin. We now proceed as in the proof of part 1 of Theorem 2. We consider the two linear maps $$\phi_1 : \Gamma_{\tilde{C}} \longrightarrow \Lambda^2 H^0(\omega \alpha)$$ $$\phi_2 : H^0(JC, \mathcal{L}_\alpha)_{0-}^{(2)} \longrightarrow H^0(C \times C, \delta^* \mathcal{L}_\alpha (-2\Delta))_- = \Lambda^2 H^0(\omega \alpha)$$ which are defined as follows. As in part 1, $\phi_2$ is the map given by pull-back under the difference map $\delta$. To define $\phi_1$, let $N_{\tilde{C}/P'}$ denote the normal bundle of $i (\tilde{C})$ in $P'$. Then $\phi_1$ is obtained by restricting a section $s \in \Gamma_{\tilde{C}}$ to the first infinitesimal neighborhood of $\tilde{C}$. In other words $$\Gamma_{\tilde{C}}^{(2)} = \mathrm{ker} \: \{ \phi_1 : \Gamma_{\tilde{C}} \longrightarrow H^0(\tilde{C}, N_{\tilde{C}/P'}^* \otimes i^* {\mathcal O}(2\Xi'_0))_- = H^0(\tilde{C}, N_{\tilde{C}/P'}^* \otimes \omega_{\tilde{C}})_- \} $$ The vector bundle $N_{\tilde{C}/P'}^*$ fits into the exact sequence \begin{equation}\label{exseqN} 0\longrightarrow N_{\tilde{C}/P'}^*\longrightarrow H^0 (\omega\alpha )\otimes{\mathcal O}_{\tilde{C} }\longrightarrow\omega_{\tilde{C} }\longrightarrow 0 \end{equation} where the right-hand map is the embedding $H^0 (\omega\alpha )\otimes{\mathcal O}_{\tilde{C} }\hookrightarrow H^0 (\omega_{\tilde{C} } )\otimes{\mathcal O}_{\tilde{C} }$ followed by evaluation $H^0 (\omega_{\tilde{C} } )\otimes{\mathcal O}_{\tilde{C} }\rightarrow\omega_{\tilde{C} }$. Therefore this map is the pull-back of evaluation $H^0(\omega \alpha) \otimes {\mathcal O} \stackrel{ev}{\rightarrow} \omega \alpha$. Let $M$ be the kernel of the latter, i.e., we have the exact sequence \begin{equation}\label{exseqM} 0 \longrightarrow M \longrightarrow H^0(\omega \alpha) \otimes {\mathcal O} \map{ev} \omega \alpha \longrightarrow 0, \end{equation} whose pull-back by $\pi$ is \eqref{exseqN}. We twist \eqref{exseqM} by $\omega \alpha$ and take cohomology $$ 0 \longrightarrow H^0(C, M \otimes \omega \alpha) \longrightarrow H^0(\omega \alpha) \otimes H^0(\omega \alpha) \map{m} H^0(\omega^2) \longrightarrow \ldots $$ where $m$ is the multiplication map. We deduce that $$H^0(\tilde{C}, N_{\tilde{C}/P'}^* \otimes \omega_{\tilde{C}})_- = H^0(C,M \otimes \omega \alpha) = \mathrm{ker} \: m = \Lambda^2 H^0(\omega \alpha) \oplus I_C^{Pr}(2)$$ where $I_C^{Pr}(2)$ is the space of quadrics through the Prym-canonical curve. It remains to show that $\mathrm{im} \: \phi_1 = \Lambda^2 H^0(\omega \alpha)$. This will follow from the next two lemmas. First we will compute, as in part 1, the image under $\phi_1$ of some special sections $s_E \in \Gamma_{\tilde{C}}$, namely $s_E$ such that $Z(s_E) = \Delta(E)$ with $E$ a general bundle in ${\mathcal SU}_C(2,\omega \alpha)$ with $h^0(E) = 2$, i.e., we determine the tangent spaces to $\Delta(E)$ along the curve $i(\tilde{C})$. This is done in the following lemma. \begin{lem}\label{lemab} Let $a,b$ be the sections defined by \eqref{defab}. Then we have $$\phi_1(s_E) = a \wedge b \in \Lambda^2 H^0(\omega \alpha)$$ up to multiplication by a nonzero scalar. \end{lem} \begin{proof} First we need to show that for a general semi-stable bundle $E$ with $h^0(E) =2$ the divisor $\Delta(E)$ is smooth at a general point $i(\tilde{p}) \in \Delta(E)$. For this decompose a general Prym-canonical divisor into two effective divisors of degree $g-1$, i.e., $D + D' \in |\omega \alpha|$. Put $L = {\mathcal O}(D)$. Then $h^0(D) = 1 = h^0(\omega(-D)) = h^0(\omega \alpha(-D)) = h^0(\alpha(D))$. If $E = L \oplus \omega \alpha L^{-1}$, then $\tilde{E} = \pi^*E = \pi^*L \oplus \omega_{\tilde{C}} \pi^*L^{-1}$, $D(\tilde{E}) = \tilde{\Theta}_{\pi^*L} + \tilde{\Theta}_{\omega_{\tilde{C}}\pi^*L^{-1}}$ and $\Delta(E) = \tilde{\Theta}_{\pi^*L|P'} + \tilde{\Theta}_{\omega_{\tilde{C}}\pi^*L^{-1}|P'}$. At a general point $i(\tilde{p}) \in \tilde{\Theta}_{\pi^*L}$, we see immediately that the tangent space to $\tilde{\Theta}_{\pi^*L}$ does not contain the tangent space to $P'$, i.e., $\Delta(E)$ is smooth at $i(\tilde{p})$. Next we compute the tangent space to the divisor $\Delta(E)$ at a smooth point $i(\tilde{p}) \in \Delta(E)$. The smoothness of $\Delta(E)$ at $i(\tilde{p})$ implies that $h^0(\tilde{C}, \tilde{E}(\tilde{p} - \sigma \tilde{p})) = 2$. We choose a basis $\{ u,v \}$ of the $2$-dimensional vector space $H^0(\tilde{C},\tilde{E}(\tilde{p} - \sigma \tilde{p}))$. Then by \cite{laszlo} prop. V.2 and the same reasoning as in the proof of part 1 of Theorem 2, we see that the projectivized tangent space $\mathbb{T}_{i(\tilde{p})}\Delta(E)$ to $\Delta(E)$ at $i(\tilde{p})$, which is a hyperplane in $\mathbb{P} T_{i(\tilde{p})}P' \cong |\omega \alpha|^*$ is the zero locus of the section in $\gamma(\tilde{p}) \in H^0(\omega \alpha)$, which is the image of $u \wedge \sigma^* v := u \otimes \sigma^*v - v \otimes \sigma^*u$ under the exterior product map $$ H^0(\tilde{E}(\tilde{p} - \sigma \tilde{p})) \otimes \sigma^* H^0(\tilde{E}(\tilde{p} - \sigma \tilde{p})) = H^0(\tilde{E}(\tilde{p} - \sigma \tilde{p})) \otimes H^0(\tilde{E}(\sigma \tilde{p} - \tilde{p})) \map{\mu} H^0(\omega_{\tilde{C}}) $$ Since $\mathrm{det} \: E = \omega \alpha$, we see that $\gamma(\tilde{p}) = \mu(u \wedge \sigma^* v) \in H^0(\omega \alpha) \subset H^0(\omega_{\tilde{C}})$. We will now describe the map $\gamma: \tilde{C} \rightarrow |\omega \alpha| \ : \ \tilde{p} \mapsto \gamma(\tilde{p})$. Note that, since $h^0 (\tilde{E}) =4$, we have $h^0 (\tilde{E} ( -\sigma\tilde{p} )) = 2$ for $\tilde{p}$ general. Hence $\{ u,v\}$ is also a basis for $H^0 (\tilde{E} (- \sigma\tilde{p} ))$. Consider the inclusion $$H^0 (\tilde{E} (- \sigma\tilde{p} )) \subset H^0(\tilde{E}) = H^0(E) \oplus H^0(E\alpha)$$ and decompose $u = u_+ + u_-$, $v = v_+ + v_-$ with $u_+,v_+ \in H^0(E) = H^0(\tilde{E})_+$ and $u_-,v_- \in H^0(E\alpha) = H^0(\tilde{E})_-$. Then the element $\gamma(\tilde{p})$ is the image of $(u_+ \wedge v_+, -u_- \wedge v_-) \in \Lambda^2 H^0(E) \oplus \Lambda^2 H^0(E\alpha)$ under the exterior product map $\Lambda^2 H^0(E) \oplus \Lambda^2 H^0(E\alpha) \rightarrow H^0(\omega \alpha)$, i.e., $\gamma(\tilde{p}) \in \mathbb{P}(\mathbb{C} a \oplus \mathbb{C} b) \subset |\omega \alpha|$. Since $\tilde{C} \subset \Delta(E)$, we have $\varphi_{\alpha can}(p) \in \mathbb{T}_{i(\tilde{p})}(\Delta(E))$. So for general $\tilde{p}$, $\gamma(\tilde{p})$ is the unique divisor of the pencil $\mathbb{P}(\mathbb{C} a \oplus \mathbb{C} b)$ containing $\tilde{p}$. Hence we can conclude that the section $\phi_1(s_E) \in H^0(M \otimes \omega \alpha)$ considered as a tensor in $H^0(\omega \alpha) \otimes H^0(\omega \alpha)$ is $a \wedge b$. \end{proof} Since, a priori, we do not know that $\mathbb{P} \Gamma_{\tilde{C}}$ is spanned by divisors of the form $\Delta(E)$, we need to establish a symmetry property for any divisor $D \in \mathbb{P} \Gamma_{\tilde{C}}$. This is done as follows. Let $\tilde{s},\tilde{t} \in \tilde{C}$ be two points of $\tilde{C}$ with respective images $s,t \in C$ and let $D$ be an element of $\mathbb{P} \Gamma_{\tilde{C}}$. Assume that $i(\tilde{s}),i(\tilde{t}) \in D$ are smooth points of $D$ and let $\mathbb{T}_s D$ and $\mathbb{T}_t D$ denote the projectivized tangent spaces to the divisor $D$ at the points $i(\tilde{s})$ and $i(\tilde{t})$. Since we can identify the projectivized tangent space to the Prym variety $P'$ at any point with the Prym-canonical space $|\omega \alpha|^*$, we may view $\mathbb{T}_s D$ and $\mathbb{T}_t D$ as hyperplanes in $|\omega \alpha|^*$. Note that $\mathbb{T}_s D$ only depends on $s \in C$ and not on the lift $\tilde{s} \in \tilde{C}$. Then we have \begin{lem}\label{lemst} With the preceding notation, we have an equivalence $$ \varphi_{\alpha can}(s) \in \mathbb{T}_t D \ \iff \ \varphi_{\alpha can}(t) \in \mathbb{T}_s D $$ \end{lem} \begin{proof} Consider the invertible sheaf $x = {\mathcal O}_{\tilde{C}}(\tilde{s} - \sigma \tilde{s} + \tilde{t} - \sigma \tilde{t}) \in P$ and the corresponding embedding $$ i_x : \tilde{C} \longrightarrow P' \qquad \tilde{p} \longmapsto {\mathcal O}_{\tilde{C}}(\tilde{p} - \sigma \tilde{p}) \otimes x. $$ The curve $i_x(\tilde{C})$ is the curve $i(\tilde{C})$ translated by $x$. A straight-forward computation shows that $i_x^{ -1 }({\mathcal O}_{P'} (2\Xi'_0)) = \omega_{\tilde{C}}x^{-2}$ and by a result of Beauville (see \cite{izadivanstraten} page 569) the induced linear map on global sections $H^0(P',2\Xi'_0) \rightarrow H^0(\omega_{\tilde{C}}x^{-2})$ is surjective. We observe that \[ i_x(\sigma \tilde{t}) = i(\tilde{s}), \qquad i_x(\sigma \tilde{s}) = i(\tilde{t}), \] and that the projectivized tangent line to the curve $i_x(\tilde{C})$ at the point $i_x(\sigma \tilde{t})$ (resp. $i_x(\sigma \tilde{s})$) is the point $\varphi_{\alpha can}(t)$ (resp. $\varphi_{\alpha can}(s)$) in $|\omega \alpha|^* \cong \mathbb{P} T_{i(\tilde{s})}P'$ (resp. $\cong \mathbb{P} T_{i(\tilde{t})}P'$). Let $\mathbb{T}_{\tilde{s}}$ (resp. $\mathbb{T}_{\tilde{t}}$) denote the embedded tangent line in $| 2\Xi_0'|^*$ to the curve $i_x(\tilde{C})$ at the point $i_x(\sigma \tilde{t})$ (resp. $i_x(\sigma \tilde{s})$), so that $\mathbb{T}_{\tilde{s}}$ (resp. $\mathbb{T}_{\tilde{t}}$) passes through the point $i(\tilde{s})$ (resp. $i(\tilde{t})$) with tangent direction $\varphi_{\alpha can}(t)$ (resp. $\varphi_{\alpha can}(s)$). Then the lemma will follow if we show that these two tangent lines intersect in a point $I(\tilde{s},\tilde{t})$, i.e. \begin{equation} \label{intpoint} \mathbb{T}_{\tilde{s}} \cap \mathbb{T}_{\tilde{t}} = I(\tilde{s},\tilde{t}) \in |2\Xi'_0|^*. \end{equation} This property follows from a dimension count: since $C$ is non-hyperelliptic, we have $x^{-2} \not= {\mathcal O}_{\tilde{C}}$, so $h^0(\omega_{\tilde{C}}x^{-2}) = 2g-2$. Since $h^0(\omega_{\tilde{C}}x^{-2} (-2\sigma \tilde{s} - 2\sigma \tilde{t})) = h^0(\omega_{\tilde{C}}(-2\tilde{s} - 2\tilde{t})) \geq 2g-5$, the tangent lines $\mathbb{T}_{\tilde{t}}$ and $\mathbb{T}_{\tilde{s}}$ are contained in a projective $2$-plane, hence intersect. To get the equivalence stated in the lemma, let $H_D$ denote the hyperplane in $|2\Xi'_0|^*$ corresponding to the divisor $D \in \mathbb{P} \Gamma_{\tilde{C}}$. Assume e.g. that $\varphi_{\alpha can}(s) \in \mathbb{T}_t D$. This means that $H_D$ contains $\mathbb{T}_{\tilde{t}}$. Since $i(\tilde{s}) \in H_D$, it follows from \eqref{intpoint} that $H_D$ also contains $\mathbb{T}_{\tilde{s}}$, so $\varphi_{\alpha can}(t) \in \mathbb{T}_s D$. \end{proof} At this stage we can conclude: by lemma \ref{lemst} we know that for all $s \in \Gamma_{\tilde{C}}$, $\phi_1(s) \in H^0(\omega \alpha) \otimes H^0(\omega \alpha)$ lies either in the symmetric or skew-symmetric eigenspace, i.e. $\mathrm{im} \: \phi_1 \subset I_C^{Pr}(2) \subset \mathrm{Sym}^2 H^0(\omega \alpha)$ or $\mathrm{im} \: \phi_1 \subset \Lambda^2 H^0(\omega \alpha)$. Lemma \ref{lemab} asserts that $\mathrm{im} \: \phi_1 \subset \Lambda^2 H^0(\omega \alpha)$. As in \eqref{comprdelta}, we have that $\phi_2(pr_-(s_E)) = pr_-(\delta^*(s_E))$, where $pr_-$ denotes the projection $ H^0(\omega \alpha) \otimes H^0(\omega \alpha) \longrightarrow \Lambda^2 H^0(\omega \alpha)$ and $s_E$ is as above. Hence we see that $\phi_2(pr_-(s_E)) = a \wedge b$. By lemma \ref{lemab} the projectivizations of $\phi_1$ and $\phi_2$ coincide on all divisors of the form $\Delta(E)$ whose images generate $\mathbb{P} \wedge^2 H^0 (\omega\alpha)$. Hence $\phi_1 = \phi_2$ up to a nonzero scalar and $\phi_1$ and $\phi_2$ are surjective. \begin{rem} An alternative way of proving that $\mathrm{im} \: \phi_1 \subset \Lambda ^2 H^0(\omega \alpha)$ would be to twice take the derivative of the quadrisecant identity for Prym varieties \cite{fay} prop. 6 (fix two points and consider the other two as canonical coordinates on the universal cover of $\tilde{C}$.) \end{rem} \section{The scheme-theoretical base locus of $\mathbb{P} \Gamma_{\tilde{C}}$} >From section 2 we know that the sets $Bs(\mathbb{P} \Gamma_{\tilde{C}})$ and $i(\tilde{C})$ are equal. To prove the scheme-theoretical equality, it will be enough to show that, $\forall \tilde{p} \in \tilde{C}$, the projectivized tangent spaces at $i(\tilde{p})$ to divisors $D \in \mathbb{P} \Gamma_{\tilde{C}}$ cut out the projectivized tangent space at $i(\tilde{p})$ to $i(\tilde{C})$, which is $\varphi_{\alpha can}(p) \in |\omega \alpha|^* = \mathbb{P} T_{i(\tilde{p})}P'$, i.e., \begin{equation} \label{inttan} \bigcap_{D \in \mathbb{P} \Gamma_{\tilde{C}}} \mathbb{T}_{i(\tilde{p})} D = \varphi_{\alpha can}(p). \end{equation} If we take $D = \Delta(E)$ for some semi-stable vector bundle $E$ with $h^0(E) = 2$ (see section 4.2) then the hyperplane $\mathbb{T}_{i(\tilde{p})}(\Delta(E)) \subset |\omega \alpha|^*$ corresponds to the unique section of the pencil $\mathbb{P}(\mathbb{C} a \oplus \mathbb{C} b)$ vainshing at $p$ (proof of lemma 4.3). Since for general $a,b \in |\omega \alpha|$ we can find a vector bundle $E$ (lemma 4.1) such that equality in lemma 4.3 holds, we can conclude \eqref{inttan}.
\section{A Radio Perspective of Galaxies} Determining how galaxies form and subsequently evolve remains a subject of intense study despite decades of research. Traditionally, astronomers primarily have used optical methods to study the characteristics of local and distant galaxies. With the discovery of the infra-red ultraluminous galaxies in the early 1980s (e.g., Soifer et al. 1984), it was soon recognized that optical studies can severely bias the understanding of galaxy properties due to dust obscuration, especially in those systems undergoing enhanced episodes of star-formation activity (i.e., a starburst galaxy). The far infrared (FIR) radiation (30 $\mu$m $< \lambda <$ 300 $\mu$m) of a starburst is composed of reprocessed ultraviolet(UV) and optical light from young, recently formed stellar populations. This radiation is absorbed by dust in the interstellar medium and thermally reradiated at FIR wavelengths. Closely related to the FIR emission in starburst galaxies is the radio continuum. Although the radio emission is linked to active star-formation by different physical mechanisms than that of the FIR, there is a tight correlation between the FIR and radio luminosity of a starburst (Helou et al. 1985). In normal galaxies (i.e., without a powerful AGN), the centimeter radio luminosity is dominated by diffuse synchrotron emission believed to be produced by relativistic electrons accelerated in supernovae remnants. Detailed radio studies of nearby starburst galaxies such as M82 (Kronberg et al. 1985, Muxlow et al. 1994) and Arp 220 (Smith et al. 1998) have revealed large numbers of young radio supernovae, embedded in extended synchrotron haloes formed by a combination of old, coalesced SNRs and cosmic ray injection into the surrounding disks of these galaxies. As the synchrotron radiation of a starburst dissipates on a physical time scale of $10^7-10^8$ years, the radio luminosity is a true measure of the {\em instantaneous} SFR in a galaxy, uncontaminated by older stellar populations. Furthermore, since supernovae progenitors are dominated by $\sim$8 {\ $M_{\odot}$} stars, synchrotron radiation has the additional advantage of being less sensitive to uncertainties in the initial mass function as opposed to UV and optical recombination line studies. Because galaxies and the inter-galactic medium are transparent at centimeter wavelengths, radio emission is a sensitive measure of star-formation in distant galaxies. \begin{figure} \centerline{\epsfig{file=local_sfr.ps,width=25pc}} \caption{Shown is the contribution to the local star-formation luminosity density (u) per luminosity interval of star-forming galaxy. The radio, IRAS, H$\alpha$, and far-ultraviolet (FUV) luminosity functions have been converted to SFRs assuming a Salpeter IMF over 0.1-100 {\ $M_{\odot}$} . The shaded region represents what may be a dust curtain beyond which optical surveys are blind to star formation. If the SFR luminosity function evolves as L $\propto (1+z)^{3.5}$, then by $z$ = 1, it will appear as the solid line. This analysis suggests that the bulk of global star-formation at high$-z$ is hidden from optical surveys.} \end{figure} Comparison of the local radio luminosity function (LF) of star-forming galaxies (Condon 1989) with those derived independently at FIR (Soifer et al. 1987), H$\alpha$ (Gallego et al. 1995), and UV wavelengths (Treyer et al. 1998) shows good agreement (Cram 1998). Figure 1 shows the four LFs in units of SFRs. This analysis suggests the bulk of local star formation is occurring in modest starbursts with SFR $\sim$ 10 {\ $M_{\odot}$} yr$^{-1}$. However, past the peak in the LF, the H$\alpha$ and UV estimates begin to fall below the radio/FIR rates, and at about 50{\ $M_{\odot}$} yr$^{-1}$ has entirely vanished. The radio source counts and redshift statistics are both consistent with pure luminosity evolution of the local population to $z \sim$ 1 with L $\propto (1+z)^{3.5}$ (e.g., Rowan-Robinson et al. 1993). Thus the peak in the star-forming RLF at $z \sim$ 1 likely moves past a few hundred {\ $M_{\odot}$} yr$^{-1}$ where optically selected surveys become severely biased. Deep radio surveys, sensitive to star-forming galaxies to $z \sim$2, provide unique information on distant, rapidly evolving galaxy populations. \section{Radio Observations of the Hubble Deep Field} The HDF has been observed previously with the VLA at 8.5 GHz to a one sigma sensitivity of 1.8 $\mu$Jy (Richards et al. 1998). In June 1997 we observed the HDF region for an additional 40 hours at 8.5 GHz. We mosaiced an area defined by four separate pointings offset from the center of the HDF by the half-power point of the primary beam response (2.7\arcm ) in the cardinal directions for about 10 hours duration each. The final combined 8.5 GHz images have an effective resolution of 3.5\arcs ~and a completeness limit of 8 $\mu$Jy which rises to 40 $\mu$Jy at 6.6\arcm from the HDF center. Within this area we detected 40 sources in a complete (5 $\sigma$) with an additional 19 sources in a supplementary sample (3.5-5 $\sigma$). In November 1996, we observed the HDF at 1.4 GHz with the VLA. The observational details and data processing are discussed by Richards (1999). The 1.4 GHz VLA image covers 40\arcm ~diameter with an effective resolution of 1.8\arcs ~ and an rms noise of 7.5$\mu$Jy. We defined a completeness limit of 40$\mu$Jy to compose a catalog of 371 radio sources of which 30 were detected at 8.5 GHz. In total 16 radio sources lie in the HDF. In February 1996 and April 1997, we observed the HDF with the MERLIN interferometer at 1.4 GHz for a total of 17 $\times$ 24 hours. These data were combined with the VLA data to produce sky images around all 89 previously known radio sources in the inner 10\arcm $\times$ 10\arcm of the field. These high resolution 1.4 GHz images have a rms noise of 3.3 $\mu$Jy at 0.2\arcs ~resolution (Muxlow et al. 1999). \section{Radio Angular Sizes and Spectra} All previous high resolution studies of submillijanksy radio sources have been limited to approximately 2\arcs resolution. Because the median angular size is known to change rather sharply below a few millijansky at 1.4 GHz (presumably due to the emergence of a new population) from $\sim$10\arcs ~to a few arcsec, our present observations are uniquely suited to study the radio morphologies of the faintest radio sources for the first time. \begin{figure} \centerline{\epsfig{file=ang_sizes_new.ps,width=25pc}} \caption{Angular sizes of sources detected in the VLA+MERLIN images.} \end{figure} Figure 2 shows the angular sizes of the 89 sources detected by both MERLIN and the VLA. Virtually all radio sources are resolved at 0.2\arcs resolution. There are very few radio sources with sizes greater than a few arcsec which are generally associated with classical FR I and II radio galaxies. Rather, the median angular size for our sample is between 1-2\arcs , indicative of radio emission on galactic or sub-galactic scales. The spectral index (S $\propto$ $\nu ^{-\alpha}$) of a source can be used to diagnose the origin of the radio emission. Inverted spectrum sources are invariably associated with self-absorbed synchrotron emission associated with an AGN. Flat spectrum sources (0 $< \alpha$ 0.5) can be produced by AGN or optically thin Bremsstrahlung radiation from star-formation at higher ($\nu >$ 5 GHz) radio frequencies. Steep spectrum sources ($ \alpha >$ 0.5) consist of diffuse synchrotron emission, often associated either with radio jets or star-formation in galaxies. For the 8.5 GHz selected sample in the HDF, the median spectral index is ${\alpha } _{8.5}$ = 0.35 . Less than 15\% of these sources are inverted. Although several of these radio sources are dominated by an AGN, many show diffuse radio emission which is likely a combination of diffuse synchrotron and free-free radiation associated with wide-scale star formation. The 1.4 GHz selected sample has a median spectral index ${\alpha } _{1.4}$ = 0.85 . Thus the microjanksy radio population at 1.4 GHz is dominated by sources with diffuse synchrotron emission. \section{Optical Identifications} The absolute astrometric accuracy of our interferometric images is set by our phase calibrator which has a position error of 0.02\arcs . The independent VLA and MERLIN radio positions for sources detected in the HDF agree to 0.04\arcs . The HDF WFPC2 images were previously aligned to the radio reference frame to about 0.1\arcs ~accuracy (Williams et al. 1996). We bootstrapped each of the eight individual WFPC2 flanking field images to the radio grid by first aligning a widefield I-band image provided to us by Barger et al. (1999) with our radio sources. Optically bright galaxies detected in both the HST and the ground-based images were used to register the individual WFPC2 frames to an accuracy of 0.1-0.2\arcs . \begin{figure} \centerline{\epsfig{file=mag_hist_new.ps,width=26pc}} \caption{I magnitude histogram of radio sources in the HDF region.} \end{figure} Of the 91 radio sources contained in published optical images (Barger et al. 1999), 72 have clear identifications with reliabilities ranging from 95-99\% (Richards et al. 1998). Figure 3 presents the magnitude histogram of these galaxies. The mean of the identifications is I = 22 mag, with a clear decline in the distribution past I = 23 mag. Sixteen radio sources cannot be identified in the HFFs or ground-based images, and three fields are blank in the HDF itself. Of the radio sources identified, the majority reside in disk systems composed of mergers, irregulars, and/or isolated spiral galaxies (Richards et al. 1998). A few red ellipticals are apparent which are almost certainly AGN. Of the radio sources with redshifts, most are at $z$ = 0.4-1 but we caution that there still exists 70\% incompleteness in the sample. Many of the disk systems identified as radio sources have clear indications of active star-formation, including prominent but narrow emission lines ([OII] and H$\alpha$), mid-infrared excesses (Aussel et al. 1999) or peculiar optical morphology. These clues coupled with the diffuse, steep spectrum radio emission give strong evidence that the {\em majority of radio sources in the HDF region are starburst galaxies.} The implied starformation rates for those galaxies with redshifts range from 10{\ $M_{\odot}$} /yr to 1000s {\ $M_{\odot}$} /yr. Two of the radio starbursts we typically detct are shown in Figure 4. \begin{figure} \centerline{ \epsfig{file=HFF36341212.ps,width=20pc}\qquad \epsfig{file=HFF36341240.ps,width=20pc}} \caption{Left) Radio contours for VLA J123634+621212 are overlaid on the WFPC2 image of this I = 19 mag. merging system at $z $ = 0.46 (Cohen et al. 1996). This radio source also has a firm ISO detection (Aussel et al. 1999). The radio SFR = 140 {\ $M_{\odot}$} /yr. The radio emission peak is coincident with a prominent dust lane. Right) VLA J123634+621240 is another candidate merger at $z $ = 1.22 (Cohen 96) and with a bright ISO detection (Aussel et al. 1999). The steep spectrum radio emission implies a SFR = 2100 {\ $M_{\odot}$} /yr . } \end{figure} \subsection{Candidate High-$z$ Radio Sources} Although the majority of radio sources in the HDF region can be identified with rather bright optical galaxies (I $\sim$ 22 mag.), 20\% of the sources remain in blank fields. These sources range in significance from 6 - 100 $\sigma$. Deep infrared imaging exists for several of these radio sources (Barger et al. 1999, Waddington et al. 1999) which shows some fraction to have very red colors consistent with them being high redshift galaxies ($z >$ 2). Figure 5 shows two such optically unidentified radio objects. \begin{figure} \centerline{ \epsfig{file=HDF36511221.ps,width=20pc}\qquad \epsfig{file=hdf36571207.ps,width=15pc}} \caption{Left) Radio contours for VLA J123651+621221 are overlaid on the HDF-I image. The sub-mm source HDF850.1 lies 6\arcs to the northeast. Right) VLA J123657+621206 is located at the very edge of the HDF but remains unidentified to I =28. The source is heavily resolved by the MERLIN beam, but is firmly detected in the low resolution VLA data alone (Richards 1999b). SCUBA source HDF850.1 is located only 3\arcs away, suggesting the two are the same object. } \end{figure} Recently, the HDF has been imaged to the confusion limit by the JCMT/SCUBA at 850 $\mu$m (Hughes et al. 1998). Intriguingly, two of the radio sources unidentified in the HDF lie within a few arcsec of the brightest two sub-mm sources. This is especially significant given that the SCUBA beam is 15\arcs in diameter. VLA J123651+621221 is a steep spectrum radio source contained in both the complete 1.4 and 8.5 GHz samples and is partially resolved by the 0.2\arcs MERLIN beam. Dickinson (private communication) reports a very red object at the position of the radio source giving further evidence to its unusual nature. In Cycle 7, we obtained NICMOS imaging of our brightest unidentified radio object (VLA J123642+621331) in J and H filters (Waddington et al. 1999). We obtained a firm detection in H, yielding a color of I - H $>$ 3.3. Surprisingly, the light profile of the underlying galaxy is exponential with a half-light radius of 0.13\arcs . This is strongly suggestive of a nuclear starburst galaxy at substantial redshift. What are these unidentified radio objects? We consider four possibilities: 1. moderate {\em z} ellipticals - Red ellipticals (I - K $>$ 4) at redshifts 1-2 are not uncommon. However, the radio sources under discussion would have to be associated with particularly underluminous or dusty parent ellipticals. VLA J123642+621331 clearly does not fall into this category. 2. one sided radio jets: We cannot discount the possibility that some of our radio sources are the brightest jet of a nearby but displaced optical galaxy. In this case we would expect the parent galaxy to be a luminous elliptical with an AGN. We find no obvious candidates for this scenario. 3. very high-$z$ AGN ($z >6$) - Another possibility, is that some of the radio sources with very faint optical fluxes are at extreme redshifts, where the Lyman break blanks out the I continuum, placing them at $z >$ 6. With our present radio sensitivity, we could have detected a nominal FR I galaxy to approximately $z$ = 10. However, in several cases the radio emission is so resolved, it is unlikey to emanate from a compact AGN. On the other hand, at least two radio objects in the HDF remain unidentified to H = 26 (Dickinson, private communication) and H = 28 (Thompson et al. 1999). 4. high $z$-starbursts (1 $< z <$ 3): Given that the majority of submillijansky radio sources are associated with star-forming galaxies, it is plausible that a tail of the parent galaxy population is so obscured by dust that only the radio emission is visible. In this case we would expect the radio emission to be steep spectrum (which it is in 18/19 of our objects) and the underlying galaxy to be a very red disk galaxy. Sevral of our objects best fit this description and we consider it the most plausible physical explanation. We note that the surface density of these objects is about 0.1 square arcmin. There is likely some overlap with the faint sub-millimeter population, although at this point the numbers are too sparse to make any definitive statements. Further observations at near infrared and sub-millimeter wavelengths are necessary to discern the nature of the optically unidentified radio population. \section{Conclusions and Future Directions} We have shown that the microjansky radio population is associated primarily with star-forming galaxies. The clues that point to the radio emission being related to star-formation are 1) the steep radio spectra, 2) the small, but extended angular sizes of the sources, and 3) the identification with luminous disk galaxies. We have also isolated a population of optically faint radio sources (I $>26 - 28$) which are possibly distant protogalaxies. \begin{figure} \centerline{ \epsfig{file=DEEP.PS,width=25pc} \caption{Shown is a simulation of a confusion limited HDF sized region at 1.4 GHz as seen by the proposed expanded VLA. The true field of view is 200 times larger, with a total of 40,000 detections above 0.5 $\mu$Jy (a factor of 80 deeper than the current VLA/HDF survey). Such observations could detect the Milky Way at $z = 1$ and Arp 220 up to $z = 10$ free from dust obscuration.}} \end{figure} Radio observations provide a powerful tool in the study of star-formation to the earliest epochs. Together with deep near infrared and sub-mm observations, they have the potential to uncover all star-forming galaxies out to $z \sim$ 2, free from dust extinction. The radio luminosity function of star-forming galaxies at moderate to high$-z$ may ultimately reveal the global star-formation history, free from optical selection biases. The next generation of centimeter wavelength telescopes (the expanded VLA, Square Kilometer Array) will extend our knowledge of the radio properties of distant galaxies to redshifts of about 10. Figure 7 shows what the radio sky may look like at the nanojansky level. I wish to thank the LOC for their generous financial support during the meeting. I thank my collaborators T. Muxlow, K. Kellermann, E. Fomalont, B. Partridge, R. Windhorst and I. Waddington.
\section{Introduction} It is known since long that thermally pulsing Asymptotic Giant Branch (TP-AGB) stars provide a site for the so called s-process, i.e., the slow neutron capture process which forms neutron-rich isotopes heavier than iron (Clayton 1968). Heavy elements primarily produced by the s-process are overabundant at the surface of AGB stars (Smith \& Lambert 1990), including technetium (Little et al. 1987) which has no stable isotope and which is produced as $^{99}$Tc ($\tau_{\rm 1/2} = 2.1\times 10^5\,$yr) in the s-process. In particular the roughly solar magnesium isotopic pattern found in s-process enriched AGB stars has demonstrated that the $^{13}$C($\alpha$,n)$\,$ rather than the $^{22}$Ne($\alpha$,n)$\,$ neutron source is likely to operate the s-process in AGB stars (Gu\'elin et al. 1995, Lambert et al. 1995). Evidence for {\it in situ} s-processing is found exclusively in carbon stars (Smith \& Lambert 1990), which correspond to a late evolutionary stage on the TP-AGB where the stars have large $^{12}$C enrichments in their envelopes (Iben \& Renzini 1983, Wallerstein \& Knapp 1998). The $^{12}$C enrichment implies that these stars contain, at certain times, a region at the bottom of their hydrogen-rich envelope where $^{12}$C is abundant. This region where protons and $^{12}$C coexist may then perhaps form $^{13}$C through $^{12}$C(p,$\gamma$)$^{13}$N($\beta^+ \nu$)$^{13}$C. Although this scenario is unrivaled, the formation of a layer which is rich in protons {\em and} $^{12}$C in TP-AGB models has proven to be difficult, and its existence had hitherto to be assumed {\it ad hoc} in all s-process calculations (Gallino et al. 1998). Iben \& Renzini (1982) found a $^{13}$C layer in low metalicity ABG models. Recently, Herwig et al. (1997) have obtained a $^{13}$C-rich layer in TP-AGB models of solar metallicity, by invoking a diffusive overshoot layer at convective boundaries. Here, we investigate for the first time effects of rotationally induced mixing processes on the TP-AGB. \section{Numerical method and physical assumptions} Our calculations have been performed with a hydrodynamic stellar evolution code (cf., Langer 1998, and references therein), which has been upgraded to include angular momentum, the effect of the centrifugal force on the stellar structure, and rotationally induced transport of angular momentum and chemical species due to Eddington-Sweet circulations, the Solberg-H{\o}iland and Goldreich-Schubert-Fricke instability, and the dynamical and secular shear instability. We apply the rotational physics exactly as in Heger et al. (1999). In particular, the effects of gradients of the mean molecular weight~$\mu$, which pose barriers to any mixing process, have been included as in Heger et al. (i.e., $f_{\mu}=0.05$). As in Heger et al., we have also included the effects of $\mu$-barriers on convection by using the Ledoux-criterion for convection and semiconvection according to Langer et al. (1983), which is consistent with our treatment of the rotational mixing (Maeder 1997). Changes of the chemical composition and the nuclear energy generation rate are computed using nuclear networks for the three pp-chains, the four CNO-cycles, and the NeNa- and the MgAl-cycle. Further, the 3$\alpha$-reaction is included, and ($\alpha$,$\gamma$)-reactions on $^{12}$C, $^{14,15}$N, $^{16,18}$O, $^{19}$F, $^{20,21,22}$Ne, $^{24,25,26}$Mg, and ($\alpha$,n)-reactions on $^{13}$C, $^{17}$O, $^{21,22}$Ne, $^{25,26}$Mg. The inclusion of (n,$\gamma$)-reactions on $^{12}$C, $^{20,21}$Ne, $^{24,25}$Mg, $^{28,29}$Si allows an order of magnitude estimate of the neutron concentration. For more details see Heger et al. (1999) and Heger (1998). \section{Results} \subsection{Evolution towards the TP-AGB} \begin{figure}[t] \begin{centering} \epsfxsize=0.9\hsize \epsffile{hrd.ps} \caption{Evolutionary track of our rotating 3$\mso$ model (solid line) and of a non-rotating reference model (dotted line) in the HR diagram. The tracks start at the zero age main sequence and end on the TP-AGB. The beginning of the thermal pulses is marked. } \end{centering} \end{figure} \begin{figure}[ht] \begin{centering} \epsfxsize=0.7\hsize \epsffile{jm1u4u56.ps} \epsfxsize=0.7\hsize \epsffile{wm1u4u56.ps} \caption{{\bf Upper panel:} logarithm of the local specific angular momentum (in cm$^2\,$s$^{-1}$) as function of the mass coordinate for three models of our rotating 3$\mso$ sequence; at core hydrogen exhaustion ($t\simeq 3.3\times 10^8\,$yr, solid line), during core helium burning ($t\simeq 3.8\times 10^8\,$yr, dotted line), and between the 14th and the 15th thermal pulse ($t\simeq 4.4\times 10^8\,$yr, dashed line). {\bf Lower panel:} logarithm of the local angular velocity (in rad/s) as function of the mass coordinate for the same models which are displayed in the upper panel. } \end{centering} \end{figure} For our pilot study, we chose to compute the evolution of a 3$\mso$ star of roughly solar composition with an initial equatorial rotation velocity of 250$\kms$, which is typical for late~B main sequence stars (Fukuda 1982). This choice renders effects of magnetic braking and the core helium flash unimportant. The evolution of our model in the HR diagram, together with that of a non-rotating reference model, is shown in Fig.~1. As in massive stars (Heger et al. 1999), the dominant rotational mixing process on the main sequence is the Eddington-Sweet circulation. It leads to a $^{12}$C/$^{13}$C-ratio after the first dredge-up of~9.4, compared to a value of 19.4 in our non-rotating model (cf., Boothroyd \& Sackmann 1999). Figure~2 sketches the evolution of the angular momentum distribution in the innermost 0.8$\mso$ of our rotating 3$\mso$ model. It shows that the core specific angular momentum decreases continuously during the evolution. We can give a first quantitative prediction of a white dwarf rotation rate: $\log (j / {\rm cm}^2 {\rm s}^{-1} ) =15.3$ and $R_{\rm WD}=0.01\rso$ yields $v_{\rm rot} = 28\kms$. This is of the same order as current observational upper limits (Heber et al. 1997, Koester et al. 1998). \subsection{Mixing and nucleosynthesis on the TP-AGB} \begin{figure}[ht] \begin{centering} \epsfxsize=0.9\hsize \epsffile{mix1.ps} \caption{Section of the internal structure during and after the 25th thermal pulse of our rotating 3$\mso$ sequence. Diagonal hatching denotes convection. The convective envelope extends down to $M_{\rm r}\simeq 0.746\mso$. The pulse driven convective shell is located at $0.737\mso \lesssim M_{\rm r}\lesssim 0.746\mso$ and $30\,{\rm yr}\lesssim t \lesssim 120\,$yr. Vertical hatching denotes regions of significant nuclear energy generation, i.e., the hydrogen burning shell (at $M_{\rm r}\simeq 0.746$ and $t\lesssim 100\,$yr) and the helium burning shell ($0.734\mso \lesssim M_{\rm r}\lesssim 0.739\mso$ and $t\gtrsim 40\,$yr). Gray shading marks regions of significant rotationally induced mixing (see scale on the right side of the figure). Vertical marks at the bottom of the figure denote the time resolution of the calculation, where every fifth time step is indicated. Cf. also Fig.~4. During this thermal pulse, the maximum energy generation rate of the helium burning shell was $4\times 10^7\lso$. } \end{centering} \end{figure} \begin{figure}[ht] \begin{centering} \epsfxsize=0.9\hsize \epsffile{mix2.ps} \caption{Same as Figure~3, for the same time interval, but magnifying the dredge-up of the convective envelope. Note the hydrogen burning shell source at $M_{\rm r}\simeq 0.7461$ and $t\lesssim 100\,$yr, and the extension of the pulse driven convection zone up to $M_{\rm r}\simeq 0.74602\mso$ at $t\simeq 100\,$yr. } \end{centering} \end{figure} Figure~3 shows the evolution of the internal structure during and after the 25th thermal pulse of our rotating model. It shows that the tip of the pulse-driven convection zone leaves after its decay a region of strong rotational mixing. This mixing becomes even stronger when the convective envelope extends downward during the third dredge-up event (cf. also Figure~4). The reason is that convection enforces close-to-rigid rotation (cf. Heger et al. 1999), with an envelope rotation rate which is many orders of magnitude smaller than that of the core. The resulting strongly differential rotation (cf. also Figure~2) allows the Goldreich-Schubert-Fricke instability, and to a lesser extent the shear instability and Eddington-Sweet circulations, to produce a considerable amount of mixing between the carbon-rich layer and the hydrogen envelope. \begin{figure}[ht] \begin{centering} \epsfxsize=0.9\hsize \epsffile{c13.ps} \caption{Chemical profiles at the location of the maximum depth of the convective envelope during the 25th thermal pulse (cf. Figure~4) of our rotating 3$\mso$ sequence. The dotted and dashed lines mark the hydrogen and the $^{12}$C mass fractions at $t=1704\,$yr, with $t=0$ defined as in Figures~3 and~4. The fat solid line denotes the $^{13}$C mass fraction at the same time. The three thin solid lines represent the $^{13}$C mass fractions at $t=2016\,$yr, $t=4155\,$yr, and $t=5139\,$yr, with a later time corresponding to a smaller peak abundance. The maximum $^{13}$C mass fractions of 3.6\% occurs at $t=2016\,$yr. The $^{13}$C peak moved inwards in the time interval from $t=1704\,$yr to $t=2016\,$yr due to continued proton captures on both, $^{12}$C and $^{13}$C. } \end{centering} \end{figure} \begin{figure}[ht] \begin{centering} \epsfxsize=0.9\hsize \epsffile{ndens.ps} \caption{Neutron density as function of the mass coordinate for three models of our rotating $3\mso$ sequence after the 25th pulse. The fat solid line corresponds to $t=14\, 150\,$yr, the other two (thin solid lines) to $t=11\, 457\,$yr (peak at largest mass coordinate) and $t=16\, 727\,$yr, with $t=0$ defined as in Figures~3,~4 and~5. The time span between the 25th and the 26th thermal pulse is $\sim 30\, 000\,$yr. } \end{centering} \end{figure} Figure~5 depicts the resulting hydrogen and $^{12}$C abundance profiles after the convective envelope has receded. It shows a layer of several $10^{-5}\mso$ containing a large mass fraction of protons and $^{12}$C at the same time. Several 1000$\,$yr after the pulse, this layer heats up and $^{13}$C is formed through proton capture on $^{12}$C. Figure~5 shows the resulting $^{13}$C profiles for four different times. A maximum $^{13}$C mass fraction of almost~4\% is achieved. Starting some $10^4\,$yr after the pulse, the $^{13}$C-rich layer becomes hot enough for $\alpha$-captures on $^{13}$C to occur (Straniero et al. 1995). Figure~6 shows the resulting neutron densities $n_{\rm n}$ for three different times. Note that we did not include the reaction $^{14}$N(n,p)$^{14}$C in our network. Although most of the resulting protons may form new $^{13}$C, it may be an effective neutron sink (Jorissen \& Arnould 1989), in particular as also a large abundance of (primary) $^{14}$N is produced in the $^{13}$C-rich layer. Thus, our neutron densities can only be considered as an order of magnitude estimate. With $n_{\rm n}\simeq 10^7\,$cm$^{-3}$ for $\sim 10^4\,$yr, we obtain a neutron irradiation of $\tau\simeq 10^{27}\,$neutrons/cm$^2$ which results roughly in a number of neutron captures per iron seed of $n_{\rm c}\simeq 75$, i.e. a main component s-process (cf. Figures~7.22 and 7.23 of Clayton, 1968). \section{Discussion} By applying the concept of rotationally induced mixing as it has been developed for massive stars in our group during the last years without alteration to a $3\mso$ TP-AGB model sequence, we obtain conditions which appear favorable for the development of the s-process, i.e. a $^{13}$C-rich layer which produces a considerable neutron flux later-on. Although our model develops only a very late and weak third dredge-up we believe that the mechanism which diffuses the protons into the carbon layer and $^{12}$C into the envelope must occur with a similar magnitude in all TP-AGB stars which develop a third dredge-up. The reason is that the huge specific angular momentum jump at the hydrogen/carbon interface --- five orders of magnitude in our case --- is independent of the depth of the third dredge-up. The maximum $^{13}$C abundance and its distribution in our model is, at first, similar to that found due to diffusive convective overshooting by Herwig et al. (1997). However, in our case the rotational mixing spreads the $^{13}$C peak out before the neutrons are produced (cf. Figures~5 and~6), which is not the case in the models of Herwig et al. (1997). At the present time we can not discriminate which of these scenarios would agree better with empirical constraints. However, we want to stress that both mechanisms of $^{13}$C production, rotation and overshooting, do not exclude each other, and that it is possible that they act simultaneously in AGB stars. Finally, we want to emphasize that, although stars of less than $\sim 1.3\mso$ lose 99\% of their angular momentum due to a magnetic wind during their main sequence evolution, it can not be excluded that the proposed mechanism of $^{13}$C-production due to differential rotation also works for them. Certainly, the sun's core will spin-up and the envelope will further spin down during its post-main sequence evolution, which may result in a specific angular momentum jump of similar magnitude. The investigation of the mass and metallicity dependence of the production of $^{13}$C due to rotation is an exciting task for the near future. \begin{acknowledgements} We are grateful to Thomas Bl\"ocker for many fruitful discussions. This work has been supported by the Deutsche Forschungsgemeinschaft through grants La~587/15-1 and 16-1. \end{acknowledgements}
\section{Introduction} \def\nu_\alpha{\nu_\alpha} \def\nu_\beta{\nu_\beta} \def\nu_ \alpha \rightarrow \nu_\beta{\nu_ \alpha \rightarrow \nu_\beta} \def{\bf p}{{\bf p}} \def{\bf l}{{\bf l}} \def{\bf k}{{\bf k}} \def{\bf q}{{\bf q}} \def{\bf x}{{\bf x}} \def{\bf y}{{\bf y}} \def{\bf z}{{\bf z}} \def{\bf s}{{\bf s}} \def{\bf r}{{\bf r}} \def{\bf u}{{\bf u}} \def\bp#1{{\underline p_{#1} }} \def\bk#1{{\underline k_{#1} }} \defE_{\vq}{E_{{\bf q}}} \def\gamma^\mu{\gamma^\mu} \def\partial_\mu{\partial_\mu} \def\hat{\bf L}{\hat{\bf L}} Recent results from the Super-Kamiokande experiment \cite{sk} have confirmed, with high statistics, the reality of the solar \cite{solrev} and atmospheric \cite{atmrev} neutrino anomalies. In the case of the atmospheric neutrino anomaly, the Super-Kamiokande data can be interpreted as providing evidence that the anomaly is due to neutrino flavor mixing \cite{skatmint}, and ``long baseline'' accelerator and reactor neutrino experiments may be able to confirm this conclusion \cite{longbl}. Data from Super-Kamiokande and the Sudbury Neutrino Observatory (SNO) eventually should allow tests to determine if the solar neutrino anomaly is also due to neutrino oscillations \cite{solsig}. These developments are bringing to a critical test the possibility suggested by earlier observations of solar and atmospheric neutrinos---and by a persistent signal in an accelerator neutrino experiment, LSND \cite{lsnd}---that neutrino flavor mixing may provide one of the first experimental windows on physics beyond the Standard Model. In a generic neutrino oscillation experiment the event rate for the detection of $\nu_\beta$ from a flux of $\nu_\alpha$ from a point source at distance $L$ takes a form like the following: \begin{equation} d\Gamma_{\alpha\beta} = \int dE_{{\bf q}} \left(d\Gamma_{\alpha,\nu_\alpha} \over L^2\, d\Omega_{{\bf q}} \, dE_{{\bf q}} \right) \left(P_{\nu_ \alpha \rightarrow \nu_\beta}\right) \left(d\sigma_{\nu_\beta,\beta}\right), \label{rate} \end{equation} where the direction of the neutrino momentum ${\bf q}$ points from the source to the detector. The first factor in the integrand represents the flux of neutrinos of energy $E_{{\bf q}}$ from a process involving a charged lepton of flavor $\alpha$, and the third factor is the cross section for neutrino detection via a process involving a charged lepton of flavor $\beta$. These factors are computed by the standard techniques of quantum field theory (QFT), with the approximation of massless neutrinos. The middle factor---the so called ``oscillation probability''---is typically computed with a quantum mechanical model of the oscillation process. In this model the neutrino state inhabits a Hilbert space whose dimension is equal to the number of neutrino flavors. The Hilbert space is spanned by a mass basis and a flavor basis. These bases are connected by a mixing matrix, taken to be the same as that which connects neutrino ``flavor fields'' to neutrino ``mass eigenstate fields'' in a field theory Lagrangian. The Hamiltonian is simply the particle energy, and the phenomenon of neutrino flavor oscillations arises because the Hamiltonian is not diagonal in the flavor basis. It was pointed out by Wolfenstein \cite{wolf} that the parameters of flavor oscillations are altered in the presence of matter because of the effective mass induced by neutrino forward scattering off the background. The effective mass, which contributes to the Hamiltonian in this quantum mechanical model, can be computed by employing the famous formula relating the index of refraction to the forward scattering amplitude, where this amplitude is computed from QFT using standard interactions. Since the effective mass due to the background is diagonal in the flavor basis, in the limit of slowly varying background density the total Hamiltonian is diagonal in a new basis, the ``instantaneous mass'' basis. Mikheyev and Smirnov \cite{ms} subsequently noted that a level crossing of the ``instantaneous neutrino mass eigenstates in matter'' occurs in a background of monotonically varying density. The resulting ``MSW effect'' constitutes a new mechanism of flavor transformation that allows a parameter space of solution to, for example, the solar neutrino problem, that is very different from that provided by vacuum neutrino oscillations. The means just described of computing an experimental event rate for neutrino flavor transformations has the virtue of simplicity. However, being somewhat schizophrenic in its amalgamation of quantum field theoretical and quantum mechanical methods, it is not surprising that studies have appeared in which the neutrino production/oscillation/detection is examined as a single process in the context of QFT, with the neutrinos being virtual particles \cite{rich,grimusstock,camp,kier,cohere,ioapilaf}. These studies identify the conditions for which the amplitude for this overall process factorizes. However, by not showing the complete relationship between this amplitude and the neutrino production flux and detection cross section, these studies lack a firm justification (other than recognition from the usual quantum mechanical picture) for calling a particular factor an ``oscillation amplitude.''\footnote{A partial connection is made in Ref. \cite{grimusstock}, where it is shown, for example, how the $1/L^2$ flux factor and on shell momentum space neutrino spinors arise from the vacuum propagator. However, the overall event rate they arrive at by ``heuristic consideration,'' to use their words, contains a normalization constant. The relationship of this normalization constant to the coordinate space external particle spinors they employ is left unspecified.} In addition, previous works employing this ``scattering approach'' to flavor mixing have only considered vacuum flavor oscillations; it is our purpose here to consider the MSW effect in the context of QFT. Our approach is complementary to the work of Ref.\ \cite{sireraperez} where this is derived from QFT within the context of relativistic Wigner functions. In this scattering approach to neutrino oscillations, the derivation of the usual Schr\"{o}dinger equations for the MSW effect in the presence of a spatially varying background is a rather trivial consequence of the virtual neutrinos going on shell, since in that case the multiparticle nature of QFT becomes irrelevant. This property was noted in Ref.\ \cite{grimusstock} in the context of vacuum oscillations. If this property is invalid, then the usual quantum mechanical treatment must be modified and the quantum field theoretical treatment becomes useful. Hence, perhaps the most important point of this paper is that the usual quantum mechanical treatment should be modified suitably if one of the following conditions holds: the on-shell neutrino momentum is nonrelativistic, the neutrino production or detection vertices are non-chiral, the wave packets of the external production/detection particles are sensitive to momentum variations of the order of inverse source-detector distance, or the neutrino effective mass splittings (determined by the effective potential including the background matter contributions) are large compared to the spread in the momenta of the external particles. We begin in Sec. II with a discussion of general neutrino oscillations. While this ground has been partially explored previously, the discussion will serve to clarify some of the physics of the oscillation process and identify the extent to which the neutrino propagator determines the probability of neutrino oscillations. In particular, we do not assume the vacuum propagator at the outset; we find the precise form a generic propagator must take to allow a physically meaningful ``oscillation probability'' to be decoupled from neutrino production fluxes and detection cross-sections, and pinpoint the component that gives rise to the oscillation amplitude. In Sec. III, we discuss the effective Lagrangian approximation assumed in our framework; as specific examples we discuss $e^\pm$ and neutrino backgrounds, including an outline of how to obtain a self-consistent neutrino background in, for example, the supernova environment. Having identified the Green's function (or propagator) as that which determines the portion of the neutrino oscillation probability that is independent of the production/detection mechanisms, in Sec. IV we study the Green's function of an effective theory of neutrinos in a static, uniform background, finding a rich pole structure. Great simplification occurs in the relativistic limit, and we recover the same oscillation amplitude obtained with the usual quantum mechanical model. In Sec. V, we study the Green's function in a nonuniform background potential. Under appropriate conditions we recover the usual Schr\"{o}dinger-type equation for the oscillation amplitude. Sec. VI contains concluding remarks. An Appendix contains a quick and easy derivation in QFT of fully normalized neutrino oscillation event rates in the form of Eq. (\ref{rate}), which is justified by the more complete treatment in Sec. II. \section{GENERAL $\nu$ OSCILLATIONS} \label{sec:general} For the sake of completeness and to establish the setting of our calculation, we give a general overview of the neutrino oscillation calculation in field theory, going beyond previous works \cite{rich,grimusstock,camp,kier,cohere,ioapilaf} in a couple of ways. While earlier studies demonstrated the factorization of the amplitude under suitable conditions---enabling identification of the factor called the ``oscillation amplitude'' in the standard picture---they do not go all the way to a fully normalized expression like that of Eq. (\ref{rate}), leaving one without an unambiguous, physically meaningful reason to name this factor an ``oscillation amplitude.'' In addition, previous works exploring the conditions under which the neutrino propagator determines the oscillation probability, independent of the details of neutrino production and detection, have only considered the vacuum propagator. As a prelude to studying the MSW effect, we seek to elucidate the form a generic neutrino propagator must have to make it possible to disentangle the neutrino oscillation probability from the production/detection mechanisms. In this section we consider generalities without committing to a particular model (Lagrangian), establishing the connection between a physically clear definition of oscillation probability [given by \eqr{rate}] and the neutrino propagator. For a simple choice of external particle wave packets, we will give a fairly general expression for the neutrino oscillation event rate including a ``calculated'' normalization. For illustration, in the Appendix we give an explicit calculation leading to a fully normalized event rate like Eq. (\ref{rate}) in the context of a semirealistic model Lagrangian. In field theory and in physical situations, we distinguish a given flavor of neutrinos by their interactions with charged leptons. If we calculate the probability amplitude for the process involving flavor $\alpha$ at the source of the neutrinos and flavor $\beta$ at the detector of the neutrinos, we can calculate the probability of neutrino oscillations of $\alpha \rightarrow \beta$. The flavor of each interaction can be distinguished by measurable, on shell, external particles (i.e., the charged leptons). Hence, calculating neutrino oscillations is equivalent to calculating a scattering event where a neutrino propagator connects two flavor distinguishing vertices with external particles coming from them. In calculating scattering quantities such as cross sections, we usually calculate the plane wave scattering $S$-matrix \begin{equation} S(\{ k_i \},\{ p_j \} )-1 \equiv (2 \pi)^4 \delta^{4} \left(\sum_l (-1)^{d_l} p_l+\sum_l (-1)^{s_l} k_l\right) i {\cal M}, \label{eq:smatrix} \end{equation} where ${\cal M}$ is the usual invariant amplitude calculated with Feynman diagrams in momentum space and $s_l=d_l=1$ for incoming particles and $0$ for outgoing particles (the grouping of the momenta will be explained shortly). This is considered to be a good approximation in the case of calculating usual collider event rates since there the events of interest occur within a single volume element before the final state particles are detected, and the corrections arising from localization of the interactions usually are not important to the detection rate. However, because neutrino oscillations involve quantum interference effects over macroscopic distances which separate the production point and the detection point of the neutrinos, we must take more care to account for the localization of the interaction points to calculate the leading order observable quantity. This localization of the vertices necessarily requires that the incident and final states be spatially localized wave packets instead of plane wave states. (As we shall see, analyses which appear to use only plane wave external states while restricting spatial integrations in an ad hoc manner \cite{camp,ioapilaf} actually have complicated wave packets buried beneath the surface.) Hence, the probability amplitude\footnote{Our conventions for the metric, gamma matrices, and normalizations are the same as Ref. \cite{peskin}. The wave packet normalization is $\int d^3{{\bf p}}\; (2\pi)^{-3} \left|\psi({\bf p})\right|^2~=~1$. A plane wave packet $\psi({\bf p},{\bf p}')= {(2\pi)^3\over\sqrt{V}}\delta^3({\bf p}-{\bf p}')$, where $V$ is a volume factor, follows this normalization convention provided $\left[\delta^3({\bf p}-{\bf p}')\right]^2$ is interpreted as ${V\over(2\pi)^3}\delta^3({\bf p}-{\bf p}')$.} is a superposition of~\eqr{eq:smatrix}, \begin{equation} {\cal A} = \int \prod_j^{1 + F_D} [dp_j]\ \psi_{D j}(p_j, \bp{j}) \prod_i^{I_S+F_S} [dk_i]\ \psi_{S i}(k_i, \bk{i})\ \left[S(\{ k_m \}, \{ p_m \} )-1\right], \label{eq:amplitude} \end{equation} where $[dp_j]=d^3 {{\bf p}}_j/\left[ (2 \pi)^3 \sqrt{2 E_{{{\bf p}}_j}}\right]$, $\{ k_m \}$ are the external momenta of the vertex at the production region centered about ${\bf x}_S$, and $\{ p_m \}$ are the external momenta of the detection region centered about ${\bf y}_D$. Here the set of parameters $\bp{m}$ and $\bk{m}$ characterize the peak of the wave packets' distribution of momenta.\footnote{The 0th component of these ``parameters'' is taken to be on mass shell since we will choose the wave packets such that they behave like plane waves for large values of the spatial components of these ``parameters.''} We have also fixed the number of external particles to be $I_S$ incoming and $F_S$ outgoing particles at the source vertex, and 1 incoming and $F_D$ outgoing external particles at the detector vertex. Now, let us see how this is related to the usual quantum mechanical treatment. The procedure for calculating $P_{\nu_ \alpha \rightarrow \nu_\beta}$ in \eqr{rate} in a quantum mechanical model was described in Sec. I. In field theory, $P_{\nu_ \alpha \rightarrow \nu_\beta}$ defined by \eqr{rate} can be calculated by comparing \eqr{rate} with the event rate derived from \eqr{eq:amplitude}. We associate ${\cal A}_S$ with the amplitude for $\{I_S\} \rightarrow \{F_S\} + \stackrel{(-)}{\nu_\alpha}$ at the source and ${\cal A}_D$ with the amplitude for $D + \stackrel{(-)}{\nu_\beta} \rightarrow \{F_D\}$ at the detector (we specialize to one detector particle $D$), and choose plane wave packets for the source's final state (anti)neutrinos and the detector's initial state (anti)neutrinos: \begin{eqnarray} {\cal A}_S & =& {1\over\sqrt{2E_{\vq}}\sqrt{V}}\int \prod_i^{I_S+F_S} [dk_i]\ \psi_{S i}(k_i, \bk{i})\ \left[S_S(\{ k_m \}, q )-1\right], \nonumber\\ {\cal A}_D &=& {1\over\sqrt{2E_{\vq}}\sqrt{V}} \int \prod_j^{1 + F_D} [dp_j]\ \psi_{D j}(p_j, \bp{j}) \left[S_D(\{ p_m \},q )-1\right], \label{detamp} \end{eqnarray} where $q=(E_{\vq},\hat{\bf L}E_{\vq})$ is the neutrino momentum, and $\hat{\bf L}=({\bf y}_D - {\bf x}_S)/|{\bf y}_D -{\bf x}_S|$ points from the source to the detector. In ${\cal A}_S$ and ${\cal A}_D$ we have implicitly assumed that the neutrinos are massless, because massive flavor eigenstates cannot be asymptotic states. Standard kinematics then yields the relationship \begin{equation} \frac{\left| \cal A \right|^2}{T}=\int \frac{dE_{\vq}\, E_{\vq}^2}{(2 \pi)^3 L^2 v_{\nu D}} \frac{|{\cal A}_S|^2 V}{T_S} P_{ \stackrel{(-)}{\nu_\alpha}\rightarrow\stackrel{(-)}{\nu_\beta}} \frac{|{\cal A}_D|^2 V}{T_D}, \label{eq:relationship} \end{equation} where $V$ is the usual total volume factor associated with the phase space and normalization of plane wave packets; $T$, $T_S$, and $T_D$ are the usual time factors associated with stationary wave packets; $v_{\nu D}$ is the M$\o$ller speed (associated with the flux) between the detector particle and the neutrinos; and $L \equiv |{\bf y}_D -{\bf x}_S|$. As just indicated, we make the simplifying assumption of stationary wave packets. The main simplifying utility of this energy conservation approximation is to get rid of the neutrino momentum integral.\footnote{We refer the reader to Ref.\ \cite{cohere} and references therein for related discussions regarding coherence.} We encode our assumption of stationarity by defining spatially smeared functions \begin{eqnarray} g_S({\bf x},\{\bk{i}\},q)\; e^{i \sum_l (-1)^{s_l} {\underline k_l}\cdot x} & = & \int \prod_j^{I_S+F_S} [ dk_j]\; \psi_{S j}(k_j, \bk{j})\; e^{i \sum_l (-1)^{s_l} k_l \cdot x}\; i {\cal M}_S\left(\{k_i \}, q \right), \nonumber \\ g_D({\bf y},\{\bp{i}\},q)\; e^{i \sum_l (-1)^{d_l} {\underline p_l}\cdot y} & = & \int \prod_j^{1+F_D} [ dp_j]\; \psi_{D j}(p_j, \bp{j})\; e^{i \sum_l (-1)^{d_l} p_l \cdot y}\; i {\cal M}_D(\{p_i \}, q). \label{eq:stationary} \end{eqnarray} We note that ${\cal M}_S$ and ${\cal M}_D$ have the form (assuming V-A lepton currents) \begin{eqnarray} {\cal M}_S\left(\{k_i \}, q \right)&=& \bar u^-(q)P_R M_1(\{k_i\}), \ \ \ {\cal M}_D(\{p_i \}, q)= M_2(\{p_i\})P_L u^-(q)\ \ \ (\nu\ \rm{osc.}), \nonumber\\ {\cal M}_S\left(\{k_i \}, q \right)&=& M_2(\{k_i\})P_L v^+(q), \ \ \ {\cal M}_D(\{p_i \}, q)= \bar v^+(q)P_R M_1(\{p_i\}) \ \ \ (\bar\nu\ \rm{osc.}), \label{msmd} \end{eqnarray} where $M_1$ and $M_2$ are respectively column and row vectors in spinor space, and $P_L$ and $P_R$ are the left- and right-handed chiral projection operators. Following the conventions of \cite{peskin}, we represent the spinors $u$ and $v$ as \begin{equation} u^s(q) =\pmatrix{\sqrt{q\cdot\sigma}\;\xi^s \cr \sqrt{q\cdot\bar\sigma}\;\xi^s}, \ \ \ \ v^s(q)=\pmatrix{\sqrt{q\cdot\sigma}\;\eta^s \cr -\sqrt{q\cdot\bar\sigma}\;\eta^s}, \end{equation} where $\sigma^\mu=(1, \mbox{\boldmath$\sigma$} )$, $\bar\sigma^\mu=(1,-\mbox{\boldmath$\sigma$})$, and $\mbox{\boldmath$\sigma$}$ is the three-vector of Pauli matrices. Since these are spinors for massless particles, the spin index $s$ is associated with the spin component along the momentum axis; specifically, we have $\xi^- = \eta^+$, $\xi^+ = -\eta^-$, with $\mbox{\boldmath $\sigma$}\cdot\hat{\bf q}\; \xi^\pm(\hat{\bf q}) =\pm \xi^\pm(\hat{\bf q})$. Putting Eqs. (\ref{detamp},\ref{eq:stationary}) into \eqr{eq:relationship} yields \begin{equation} (2 \pi)\, \delta\left(\sum_l (-1)^{s_l} E_{\underline {{\bf k}}_l}+\sum_l (-1)^{d_l} E_{\underline {{\bf p}}_l}\right)\, P_{\stackrel{(-)}{\nu_\alpha}\rightarrow\stackrel{(-)}{\nu_\beta}} = 16 \pi^2 L^2 v_{\nu D}\, \frac{\left| \cal{A} \right|^2}{T}\, \left| \tilde{g}_S(-{\bf q}, \{ \bk{i} \}) \right|^{-2} \left| \tilde{g}_D({\bf q}, \{ \bp{i} \}) \right|^{-2}, \label{eq:almost} \end{equation} where we have defined \begin{equation} \tilde{g}_D({\bf q}, \{\bp{i}\})\equiv \int d^3{\bf y}\, g_D({\bf y}, \{ \bp{i}\},q ) \, e^{i \left({\bf q}-\sum_l(-1)^{d_l}{\underline{\bf p}_l}\right) \cdot {\bf y}}, \label{gtilde} \end{equation} and similarly for $\tilde{g}_S$. Stationarity constrains $|{\bf q}|\equivE_{\vq}=-\sum_l (-1)^{s_l} E_{\underline {{\bf k}}_l}$. Let us turn our attention to $\cal A$. Given that we have a neutrino propagator $G$ in our amplitude, and assuming V-A lepton currents, we can write \begin{equation} S(\{ k_i \}, \{ p_j \} )-1 = \int d^4y\,e^{i \sum_l (-1)^{d_l} p_l \cdot y} \int d^4x\,e^{i \sum_l (-1)^{s_l} k_l \cdot x}\, i\int {d^4s\over (2\pi)^4} e^{ \mp i s\cdot(y-x)} M_2\, P_L\, G(s)\, P_R\, M_1, \label{eq:factorit} \end{equation} where $M_1$ and $M_2$ are the same as in Eqs. (\ref{msmd}), and $s$ is the off-shell propagator momentum. The upper (lower) sign of $\mp$ in the exponential is for neutrino (antineutrino) oscillations; this arises from choosing $x$ ($y$) to always correspond to the source (detector). That is, for neutrino oscillations of flavor $\alpha$ to flavor $\beta$, the Green's function is $iG^{\beta\alpha}(y,x)=\langle T\{\nu^\beta(y)\bar\nu^\alpha(x)\}\rangle_0 =i\int{d^4s\over(2\pi)^4}\,e^{-is\cdot(y-x)}G^{\beta\alpha}(s)$ (with $T\{\}$ and $\langle \rangle_0$ denoting a time-ordered product and vacuum expectation value respectively), while for antineutrino oscillations $\alpha\rightarrow \beta$, the labeling is $iG^{\alpha\beta}(x,y)$. Insert the identity \begin{equation} {q^\mu \xi^\nu \{\gamma_\mu,\gamma_\nu\} \over 2q\cdot \xi}=1 \label{eq:ident1} \end{equation} on both sides of the Green's function in equation \eqr{eq:factorit}, where as before $q=(E_{\vq},\,\hat{\bf L}E_{\vq})$, and we define $\xi= (E_{\vq},\, \hat{\bf L}\sqrt{(E_{\vq})^2 - m^2})$ in which the parameter $m^2 < E_{\vq}^2$, though its precise value is unimportant in this context. We note that since $q$ is null, $q\cdot \gamma = \sum_s u^s(q)\bar u^s(q)$ (or $q\cdot \gamma = \sum_s v^s(q)\bar v^s(q)$, if one wishes to consider antineutrino oscillations). From the explicit form of $u$ and $v$ it is easy to see that \begin{eqnarray} P_L\, (q\cdot\gamma)(\xi\cdot\gamma)&=&P_L\, u^-(q)\bar u^-(q)\, (\xi\cdot\gamma), \ \ \ \ P_L\, (\xi\cdot\gamma)(q\cdot\gamma)= (\xi\cdot\gamma)\,P_R\, u^+(q)\bar u^+(q),\nonumber\\ (\xi\cdot\gamma)(q\cdot\gamma)\, P_R &=& (\xi\cdot\gamma)\, u^-(q)\bar u^-(q)\,P_R, \ \ \ \ (q\cdot\gamma)(\xi\cdot\gamma)\, P_R= u^+(q)\bar u^+(q)\, P_L\, (\xi\cdot\gamma). \label{spinors} \end{eqnarray} [The same relations hold for $u^\pm(q)$ replaced by $v^\mp(q)$.] Soon we will show the form that the Green's function must have, after localization by the source and detector, in order that the term with $u^-$ (or $v^+$) on both sides of $G$ be the only one to contribute. If more than one spin contributes, we will not recover the usual quantum mechanical treatment without spins taken into account. In that case, working with the full scattering picture of \eqr{eq:amplitude} is useful. Keeping only the term with the relevant spinor on either side of $G$, one can show that Eq. (\ref{eq:amplitude}) becomes \begin{equation} {\cal A}= -\int d^4y \; g_D({\bf y},\{\bp{i}\},q)\, e^{i \sum_l (-1)^{d_l} {\underline p_l}\cdot y} \int d^4x\; g_S({\bf x},\{\bk{i}\},q)\, e^{i \sum_l (-1)^{s_l} {\underline k_l}\cdot x}\, i\int {d^4s\over (2\pi)^4} e^{\mp i s\cdot(y-x)} \bar P\, G(s)\, P, \label{amplitude2} \end{equation} where $P=\gamma^0 u^-(q)/(2E_{\vq}) = \gamma^0 v^+(q)/(2E_{\vq}) $, $\bar P=P^\dagger \gamma^0$, and $g_S$ and $g_D$ are given by Eq. (\ref{eq:stationary}). In passing, we would like to remark that in \eqr{amplitude2}, we can always write (for neutrino oscillations, for example) \begin{equation} g_S({\bf x}, \{ \bk{i} \},q)\, e^{i \sum_l (-1)^{s_l} {\underline k_l}\cdot x} = i \bar{u}^-(q)\, M_1(\{ \bk{i} \})\, f_S(x, \{ \bk{i} \} ) \end{equation} (and similarly for $g_D$) where $f_S$ is a scalar function of ${\bf x}$. Then we obtain the form (again assuming the single-spin contribution is justified after spatial integration) \begin{equation} {\cal A} = \int d^4x\, d^4y\, f_S(x, \{\bk{i} \}) f_D(y, \{ \bp{i} \} ) \int \frac{d^4s}{(2 \pi)^4} e^{-i s \cdot (y-x) } i {\cal M}(\{ \bk{i} \}, s, \{ \bp{i} \} ), \label{eq:usual} \end{equation} which is a common starting point of analysis in the literature as in Refs. \cite{camp,ioapilaf}. However, with arbitrarily chosen smearing functions $f$, it is difficult to assess what actual scattering question the amplitude is an answer to, because the smearing functions are not the wave functions of the in-out particles, but are the wave functions smeared over the matrix elements. As a consequence, the normalization is usually ignored in this approach. We will return to the normalization later in this section. Returning to Eq. (\ref{amplitude2}), integration over $x^0$, $y^0$, and $s^0$ gives an overall energy-conserving delta function and sets $s^0 = E_{\vq}$ (for antineutrino oscillations, $s^0 = -E_{\vq}$). We also note that the chiral structure of $\bar P\, G P$ (as well as the original matrix element) picks out only the $G_{LR}$ block of the neutrino propagator, where $G_{LR}$ is the nonzero $2\times2$ submatrix left by $P_L G P_R$. In addition, the localization of $g_S$ and $g_D$ around ${\bf x}_S$ and ${\bf y}_D$ respectively ``clamps down'' on the coordinate space Green's function. In particular, if the characteristic widths $L_S$ and $L_D$ of $g_S$ and $g_D$ are much smaller than the source-detector distance $L=|{\bf y}_D-{\bf x}_S|$, we note that if the ``oscillation probability'' is to be disentangled from the details of neutrino production and detection, the relevant portion of the Green's function for oscillations $\alpha\rightarrow\beta$ must take the form \begin{eqnarray} G_{LR}^{\beta\alpha}(s^0=E_{\vq},{\bf y},{\bf x}) &=& \int {d^3{\bf s}\over (2\pi)^3}\, e^{i{\bf s} \cdot ({\bf y}-{\bf x})}\,G_{LR}^{\beta\alpha}(s^0=E_{\vq},{\bf s})\nonumber\\ & \simeq& -E_{\vq} \left(1-\mbox{\boldmath $\sigma$}\cdot \hat{\bf L}\right) {e^{iE_{\vq}\hat{\bf L}\cdot({\bf y}-{\bf x})} \over 4\pi |{\bf y}_D-{\bf x}_S|} H^{\beta\alpha}(E_{\vq},{\bf y}_D,{\bf x}_S)\ \ \ \ \ (\nu\ \rm{osc.}), \nonumber\\ & & \nonumber\\ G_{LR}^{\alpha\beta}(s^0=-E_{\vq},{\bf x},{\bf y}) & \simeq& +E_{\vq} \left(1-\mbox{\boldmath $\sigma$}\cdot \hat{\bf L}\right) {e^{iE_{\vq}\hat{\bf L}\cdot({\bf y}-{\bf x})} \over 4\pi |{\bf y}_D-{\bf x}_S|} \bar H^{\alpha\beta}(E_{\vq},{\bf y}_D,{\bf x}_S)\ \ \ \ \ (\bar\nu\ \rm{osc.}), \label{gform} \end{eqnarray} where $\hat{\bf L} = ({\bf y}_D-{\bf x}_S)/|{\bf y}_D-{\bf x}_S|$ points from the source towards the detector, and the quantities $H$ and $\bar H$ have only flavor indices. The factor $E_{\vq} \left(1-\mbox{\boldmath $\sigma$}\cdot \hat{\bf L}\right)$ arises from the kinetic term in the Lagrangian, and takes this form due to the relativistic limit. Another key ingredient is the factor $e^{iE_{\vq}|{\bf y}-{\bf x}|}$, which is the leading phase factor in the relativistic limit coming from $e^{ i {\bf s} \cdot ({\bf y}-{\bf x})}$ evaluated at the poles of $G_{LR}$. In addition, $1/|{\bf y}-{\bf x}|$ comes from the asymptotic expansion of the left hand side of \eqr{gform} in the limit that $|{\bf y}-{\bf x}| \rightarrow \infty$, and it can be considered to be the monopole term in a multipole expansion. We will discuss the validity of the factorization and the asymptotic expansion further below in momentum space. Before we talk about momentum space, let us give an example of \eqr{gform} by considering the vacuum propagator. In that case, it is straightforward to show that (anticipating the relativistic limit) \begin{eqnarray} G_{LR}(s^0,{\bf x},{\bf y}) & =& \left(s^0 + i\mbox{\boldmath $\sigma$}\cdot \mbox{\boldmath$\nabla$}\right)\left[M^{-1} G_{RR}(s^0,{\bf x},{\bf y})\right], \label{glrgrr}\\ \left[M^{-1} G_{RR}(s^0,{\bf x},{\bf y})\right]^{\alpha\beta}&=& - {e^{i|s^0||{\bf x}-{\bf y}|} \over 4\pi |{\bf x}-{\bf y}|} \sum_j U_{\alpha j}U_{\beta j}^* \exp\left(-i{m_j^2|{\bf x}-{\bf y}|\over 2|s^0|}\right), \label{grr} \end{eqnarray} where $G_{RR}$ is the nonzero $2\times2$ submatrix left by $P_R G P_R$, $M$ is the mass matrix appearing in the Lagrangian, and the $m_j$ are the mass eigenvalues. In Eq. (\ref{grr}) we have made the flavor indices explicit; the relationship between the flavor fields and mass eigenstate fields is $\nu_\alpha = \sum_i U_{\alpha i} \psi_i$, where the $U_{\alpha i}$ are elements of a unitary matrix. For $|s^0|L\gg 1$, \begin{equation} G_{LR}^{\alpha\beta}(s^0,{\bf x},{\bf y})\simeq - {e^{i|s^0||{\bf x}-{\bf y}|} \over 4\pi |{\bf x}-{\bf y}|} \left[s^0 -|s^0|\mbox{\boldmath $\sigma$}\cdot \hat{\bf r}({\bf x},{\bf y})\right] \sum_j U_{\alpha j}U_{\beta j}^* \exp\left(-i{m_j^2|{\bf x}-{\bf y}|\over 2|s^0|}\right), \label{glrfinal} \end{equation} where $\hat{\bf r}({\bf x},{\bf y})=({\bf x}-{\bf y})/|{\bf x}-{\bf y}|$. To apply Eq.\ (\ref{glrfinal}) to neutrino oscillations one takes $s^0\rightarrowE_{\vq}$ and ${\bf x},{\bf y} \rightarrow {\bf y},{\bf x}$. For antineutrino oscillations, $s^0\rightarrow -E_{\vq}$ and ${\bf x},{\bf y}\rightarrow{\bf x},{\bf y}$. After making these substitutions we will be integrating Eq. (\ref{glrfinal}) over localization functions of characteristic widths $L_S$ and $L_D$ centered on ${\bf x}={\bf x}_S$ and ${\bf y}={\bf y}_D$. This means that for $L_S,L_D \ll L$, we may replace ${\bf x}$ and ${\bf y}$ by ${\bf x}_S$ and ${\bf y}_D$ everywhere except the phase factors, in which we consider the first order variation, \begin{eqnarray} |{\bf x}-{\bf y}|&\simeq& | {\bf x}_S-{\bf y}_D | + \hat{\bf L} \cdot \left[ ({\bf y}-{\bf y}_D) - ({\bf x}-{\bf x}_S) \right] \nonumber\\ &= &\hat{\bf L} \cdot ({\bf y}-{\bf x}). \label{dapprox} \end{eqnarray} We see that a necessary mathematical condition (in addition to the relativistic assumption and $E_{\vq} L\gg 1$) for Eqs. (\ref{glrgrr}-\ref{glrfinal}) to reduce to the form of Eqs. (\ref{gform}) is $m_j^2\, L_{S,D} /(2E_{\vq}) \ll 1$ [which also implies the more familiar $(m_j^2-m_i^2)\, L_{S,D} /(2E_{\vq}) \ll 1$]. The physical basis of these conditions can also be inferred from Eqs. (\ref{glrgrr}-\ref{glrfinal}). $E_{\vq} L\gg 1$ allows the propagating neutrino to become an on-shell relativistic particle, and also allows appreciable oscillation phase to build up over the source-detector distance. $m_j^2\, L_{S,D} /(2E_{\vq}) \ll 1$ requires that no appreciable oscillation phase build up on length scales comparable to the width of the external particle wave packets. The necessity of these conditions for disentanglement of the flavor oscillations from the details of neutrino production and detection is evident. The origin of these conditions can also be understood in momentum space. First, rewrite \eqr{amplitude2} as \begin{eqnarray} {\cal A} & = & - (2 \pi) \delta\left(\sum_l (-1)^{s_l} E_{\underline {{\bf k}}_l} + \sum_l (-1)^{d_l} E_{\underline {{\bf p}}_l}\right) i \int \frac{d^3 s}{(2 \pi)^3} e^{ \pm i {\bf s} \cdot ({\bf y}_D - {\bf x}_S)} e^{ -i \left(\sum_l (-1)^{s_l} {\underline {\bf k}_l} \cdot {\bf x}_S + \sum_l (-1)^{d_l} {\underline {\bf p}_l} \cdot {\bf y}_D \right )}\nonumber\\ & &\times \tilde{h}_{D}({\bf s}, \{ \bp{i} \}, q) \tilde{h}_{S}(- {\bf s}, \{ \bk{i} \}, q) \bar{P}\; G\left(s^0 = \mp\sum_l (-1)^{s_l} E_{\underline {{\bf k}}_l} ,{\bf s} \right) P \label{eq:a} \end{eqnarray} where the functions $\tilde{h}_{S}$ and $\tilde{h}_{D}$ can be approximated to have no ${\bf x}_S$ or ${\bf y}_D$ dependence. This can easily be seen to be exactly true for the ideal case of isotropic smearing functions, e.g. $g_D({\bf y}, \{ \bp{i}\},q )= h_D(|{\bf y}-{\bf y}_D|)$; Eq. (\ref{gtilde}) then yields $\tilde{g}_D = e^{i{\bf u}\cdot{\bf y}_D} \tilde h_D(|{\bf u}|)$, where ${\bf u} \equiv {\bf s}-\sum_l(-1)^{d_l}{\underline{\bf p}_l}$. Because the propagator will in general have poles corresponding to the mass of the physical neutrino states, the dominant contribution to the integral in the asymptotic limit $ L E_{{\bf q}} \rightarrow \infty$ will be from a term that contains the integrand of \eqr{eq:a} as a factor evaluated at the poles and stationary phase points (critical points). For the vacuum, the constant potential, and the adiabatically spatially varying background potential cases, one can asymptotically expand the integral \eqr{eq:a} in the limit $LE_{{\bf q}} \rightarrow \infty$ (similarly as in Ref. \cite{grimusstock}) to find that to leading approximation, the term $\tilde{h}_{D}({\bf s}, \{ \bp{i} \}, q) \tilde{h}_{S}(-{\bf s}, \{ \bk{i} \},q)$ can be moved outside of the integral with the replacement ${\bf s} \rightarrow {\bf s}_*$ where ${\bf s}_*$ corresponds to one of the critical points. By factoring out $\tilde{h}$, we have implicitly assumed that $\tilde{h}_{D}({\bf s}, \{ \bp{i} \},q) \tilde{h}_{S}(-{\bf s}, \{ \bk{i} \},q)$ is not sensitive to the splittings in the critical points (otherwise different pole momenta ${\bf s}$ of $G_{LR}$ will cause $\tilde{h}_{D}({\bf s}, \{ \bp{i} \},q) \tilde{h}_{S}(-{\bf s}, \{ \bk{i} \},q)$ to have different values, preventing factorization). This means that the wave packets must be flat in momentum space at least within the range of pole momentum splitting. Also, if this is not the case, one of the poles will not contribute (because the amplitude of the wave packet has fallen off with respect to the amplitude at the other pole), and no neutrino oscillations will occur (or more accurately, the neutrino oscillations will be greatly suppressed relative to the background). We will refer to this flatness of the wave packet as insensitivity to ${\bf s}_*$ splitting. Also note that because of the presence of the exponential in Eq. (18), this leading term in the asymptotic expansion will not be a good approximation unless the inverse ``momentum scale height'' (i.e. logarithmic derivative) of $\tilde{h}_{D}\,\tilde{h}_{S}$ near the poles is much less than L. Hence, factoring out the wave packet dependence which is crucial for the validity of the usual quantum mechanical treatment requires the wave packet factor $\tilde{h}_{D}\, \tilde{h}_{S}$ to be insensitive under $1/L$ momenta variations as well as the ${\bf s}_{*}$ splitting variations. While localization is clearly necessary for the observation of oscillations, the source and detector localization scales $L_S,L_D$ implied by Eq. (\ref{eq:stationary}) cannot be smaller than the Compton wavelength of the lightest external particles. In the case that all the external particles connected to a given vertex are nonrelativistic, this gives rise to a constraint on the masses of these external particles. To see this, consider the ideal case mentioned above in which $g_D({\bf y}, \{ \bp{i}\},q )\approx h_D(|{\bf y}-{\bf y}_D|)$. Then $\tilde{g}_D = e^{i{\bf u}\cdot{\bf y}_D} \tilde h_D(|{\bf u}|)$, where $\tilde h_D(|{\bf u}|)$ is damped for $|{\bf u}| \equiv |{\bf s}_*-\sum_l(-1)^{d_l}{\underline{\bf p}_l}|$ larger than $1/L_D$. Hence the ${\bf s}_*$ splitting insensitivity condition can be written as \begin{eqnarray} \left| \sum_l (-1)^{s_l} {\underline {\bf k}_{l}} + {\bf s}_* \right| & \ll & \frac{1}{L_S} < M_{LS} \nonumber \\ \left| \sum_l (-1)^{d_l} {\underline {\bf p}_{l}} - {\bf s}_* \right| & \ll & \frac{1}{L_D} < M_{LD}. \label{eq:osccond1} \end{eqnarray} where we have denoted the lightest external particle masses to be $M_{LS}$ and $M_{LD}$ for source and detector, respectively. The critical momentum will be ${\bf s}_* \approx \hat{\bf L} \sqrt{(\sum_l (-1)^{s_l} E_{\underline {{\bf k}}_l})^2 - \tilde{m}_{j}^2}$ where $\tilde{m}_j$ is the effective pole mass of the particle. Hence, if $\tilde{m}_{j} \ll \left|\sum_l (-1)^{s_l} E_{\underline {{\bf k}}_l}\right|$ and the external particles are nonrelativistic, then \eqr{eq:osccond1} can be satisfied only if about equal mass of external particles enter and leave the source/detector vertices. If any of the the external particles connected to a given vertex are sufficiently relativistic, this severe constraint does not arise. We now show that Green's function must take the form found in Eqs. (\ref{gform}) after being spatially ``clamped'' by the source and detector if the terms projected by the spinors $u^-$ (or $v^+$) on both sides of $G$ are to be the only contributions to the amplitude. In Sec. \ref{sec:cback}, where we study the neutrino propagator in a uniform, static medium, we will find it convenient to identify $\hat{\bf L}$ of Eq. (\ref{gform}) with the positive third spatial direction. In that case it is straightforward to show, using Eq. (\ref{spinors}), that the terms with spinors of different spins on either side of $G$ pick out the off-diagonal spinor space elements of $G_{LR}$, while the terms with the same spins on both sides pick out the diagonal spinor space elements. The matrix $\left(1-\sigma^3\right)$ from Eqs. (\ref{gform}) confirms that only $G_{LR}^{22}$ is nonzero, and therefore only the term with $u^-(q)$ [$v^+(q)$] on both sides of $G$ survives for the neutrino (antineutrino) oscillations.\footnote{Note that with $\hat{\bf L}$ set to the third spatial direction, both $u^-(q)$ and $v^+(q)$ have 4-spinor components $(0,\sqrt{2 E_{\vq}}, 0, 0)$. } Recalling that ${\bf q}=\hat{\bf L}\,E_{\vq}$, and upon inserting Eqs. (\ref{gform}) into Eq. (\ref{amplitude2}) and Eq. (\ref{amplitude2}) into Eq. (\ref{eq:almost}), we finally arrive at the neutrino oscillation probability \begin{eqnarray} P_{\nu_ \alpha \rightarrow \nu_\beta} &=& \left|H^{\beta\alpha}(E_{\vq},{\bf y}_D,{\bf x}_S) \right|^{2},\ \ \ \ (\nu\ \rm{osc.}), \nonumber\\ P_{\bar\nu_\alpha\rightarrow\bar\nu_\beta} &=& \left|\bar H^{\alpha\beta}(E_{\vq},{\bf y}_D,{\bf x}_S) \right|^{2},\ \ \ \ (\bar\nu\ \rm{osc.}), \label{oscprob} \end{eqnarray} where $H$ and $\bar H$ are defined by Eqs. (\ref{gform}), and we have assumed that the detector particle $D$ is nonrelativistic such that $v_{\nu D}=1$. With the cancellation of the source and detector wave packets, one can see why employing a separate quantum mechanical model to compute the oscillation probability is possible. [Note that the standard vacuum oscillation probability is recovered here, as is clear from Eqs. (\ref{gform}-\ref{glrfinal})]. We emphasize that the Green's function here is the full propagator in any given theory and we have made no severe assumptions about the nature of the production and detection effective vertex.\footnote{The most significant assumptions leading to our final result were the V-A type of lepton currents; the stationary approximation in \eqr{eq:stationary}; relativistic neutrinos, i.e. $(E_{\vq} - |{\bf s}_*|)/E_{\vq} \ll 1$, where $|{\bf s}_*|$ is the magnitude of a pole in the momentum space propagator; sufficiently localized and separated source and detector, i.e. $(E_{\vq} - |{\bf s}_*|) L_{S,D} \ll 1$; and $E_{\vq} L \gg 1$. } Hence, as expected, the production/detection independent field theoretical effects on neutrino oscillations come from the coordinate space Green's function. What perhaps is less expected is the fact that unless the wave packets of the external particles satisfy specific properties, the transition probability will not just depend on the propagator, but the entire coherent scattering process which neutrino oscillation really is. Such tangled wave packet dependence is discussed for example in Ref.\ \cite{ioapilaf}. Before concluding this section, we would like to note that we can easily work out the neutrino oscillation detection rate including the normalization if we assume a particular class of wave packets. Let us define a box wave packet as a configuration such that the superposition integral gives, for each outgoing particle, for example,\footnote{We will write the wave packet centered about ${\bf x}_S$, since the one centered about ${\bf y}_D$ is analogous.} \begin{equation} \int \frac{d^3 k}{(2 \pi)^3 \sqrt{2 E_k}} \psi_{S}(k, {\underline k}) e^{i k \cdot x}\approx e^{iE_{\underline{{\bf k}}}x^0} \int \frac{d^3 k}{(2 \pi)^3 \sqrt{2 E_k}} \psi_{S}(k, {\underline k}) e^{-i {\bf k} \cdot {\bf x}} = N e^{i {\underline k} \cdot x} B({\bf x} -{\bf x}_S),\label{box} \end{equation} where $N$ is a constant independent of $x$ and $B({\bf z})$ is a function which vanishes if ${\bf z}$ is outside of a box centered about the origin with each dimension of length $L_S$ and is 1 everywhere else. The approximation in Eq. (\ref{box}) is valid for $x^0 \ll (E_{\underline{{\bf k}}} L_S)/(2\pi|\underline{{\bf k}}|)$. The normalization condition $\int \frac{d^3k}{(2 \pi)^3} |\psi(k)|^2 =1$ fixes $N$ and implies \begin{equation} \psi_{S}(k)= {1\over\sqrt{V}} \left[ 1- {\cal O}\left(\frac{1}{L^2 E_{\underline{{\bf k}}}^2}\right)\right] e^{i({\bf k}- {\underline{\bf k}}) \cdot {\bf x}_S} D_S({\bf k}-{\underline{\bf k}}), \label{eq:normal} \end{equation} where \begin{equation} D_S({\bf v}) = 8\, \frac{ \sin(v_x L_S/2) \sin(v_y L_S/2) \sin(v_z L_S/2)}{v_x v_y v_z}. \label{eq:boxtrans} \end{equation} Hence, since the box scale must be larger than the Compton wavelength scale, these external particles will generally have ``plane wave in a box'' type of normalization [up to a $(2 \pi)^3 \delta^{3}({\bf k}-{\underline{\bf k}})$ type of localization factor $D_S$]. This wave packet can be used to calculate the event rate using \eqr{eq:amplitude} in a standard way. Since $\psi_S$ will have a width $2 \pi/L_S$, smearing of any function that is proportional to momenta whose magnitude at the peak of the distribution is of the order $2 \pi/L_S$, (or less) will deviate significantly from the Dirac $\delta$ smearing of that function. Fortunately, because $M_1$ and $M_2$ do not depend on such small momenta, we can write the amplitude in the form of \eqr{eq:usual} with \begin{equation} f_{S}(x, \{\bk{i} \})= \prod_j^{I_S+F_S} \left[ \frac{e^{-i(-1)^{s_j} {\underline k_j} \cdot x}}{\sqrt{2 E_{{\underline{\bf k}}_j} V_S}} B({\bf x}-{\bf x}_S) \right] \label{eq:separate} \end{equation} (and similarly for $f_{D}$) which is what one would use to calculate scattering of particles confined to a box interacting with particles that can propagate outside of the box. Explicitly, the transition rate per source and detector particle is given by \begin{eqnarray} d \Gamma & = &(2 \pi)\, \delta\left(\sum_l (-1)^{s_l} E_{\underline {{\bf k}}_l} + \sum_l (-1)^{d_l} E_{\underline {{\bf p}}_l}\right) \prod_j^{I_S+F_S}\frac{1}{2 E_{{\underline{\bf k}_{j}}} V_S } \prod_i^{F_D+1}\frac{1}{ 2 E_{{\underline{\bf p}_{i}}} V_D} \prod_b^{F_S} \frac{d^3 {\underline{\bf k}}_{b} V_S}{(2 \pi)^3} \prod_a^{F_D} \frac{d^3 {\underline{\bf p}}_{a} V_D}{(2 \pi)^3} \nonumber \\ & &\times\left| \int \frac{d^3 {\bf s}}{(2 \pi)^3} e^{\pm i {\bf s} \cdot ({\bf y}_D -{\bf x}_S)} D_S\left(\sum_l (-1)^{s_l} {\underline{\bf k}}_l \pm {\bf s}\right) D_D\left(\sum_l (-1)^{d_l} {\underline{\bf p}_l} \mp {\bf s}\right) i {\cal M} \right|^2, \end{eqnarray} where $D_S$ is defined by \eqr{eq:boxtrans}, and $D_D$ is similarly defined. In this case, one can also use the usual heuristic box quantization formalism to calculate the event rates including the normalization. For pedagogical purposes we carry out this simple exercise explicitly using a fermion field toy model in the Appendix. To summarize this section, we have shown to what extent the neutrino ``oscillation probability'' is determined by the production/detection wave packet-independent propagator of the field theory. If wave packets for the production and detection events described in \eqr{rate} satisfy suitable localization properties and the effective mass splitting of the neutrinos is not large compared to the momentum width of the wave packets, the neutrino propagator determines the probability of transition as defined by \eqr{rate}. This factoring of the wave packets out of the transition amplitude is crucial to recover the usual quantum mechanical picture of neutrino oscillations. Furthermore, we see how the multiparticle nature of the field theory becomes irrelevant as the poles of the propagator are the only states to contribute in this limit. For this factorization to be possible, the source-detector separation $L$ must be large enough such that the wave packets do not vary over $1/L$ momentum perturbations about the pole momentum, and the pole momenta splitting must be small enough such that the wave packet amplitudes take on approximately the same value for the various pole momenta [as discussed between \eqr{eq:a} and \eqr{eq:osccond1}]. Furthermore, since the usual quantum mechanical treatment neglects the spin of the neutrinos, only one spin projection of the Green function must contribute to the amplitude to recover the usual treatment. We have seen in this section that the relativistic limit of the on shell neutrinos and the chiral nature of the interactions ensure this. Now that we see that wave packet dependence can be factored out (as is implicit in the usual simple quantum mechanical treatment), we shall concentrate on the wave packet-independent field theoretic calculation of the MSW effect, which is encoded in the propagator within a background medium. \section{Effective Lagrangian} In this section, we briefly explain the effective potential employed in our calculation of the MSW effect. Focusing on the physics well below the electroweak scale, we write the usual electroweak effective Hamiltonian density as (see for example \cite{kuorev1}) \begin{equation} {\cal H}_I= \frac{G_F}{\sqrt{2}} \left( J_c^{\mu} J_{c \mu}^{\dagger} + J_N^{\mu} J_{N \mu} \right), \label{ewham} \end{equation} where $J_c^{\mu}$ is the charged current and $J_N^{\mu}$ is the neutral current. Take for example the contribution to the neutrino-electron interaction of the form \begin{equation} {\cal H}_I = \frac{4 G_F}{\sqrt{2}} \bar{\nu}_e \gamma^\mu P_L \nu_e \bar{e} \gamma_\mu P_L e, \label{eq:enuint} \end{equation} which will be dominant for the MSW effect. We can distinguish two different types of scattering: forward scattering, for which the background particles do not change their momenta; and non-forward scattering. In calculating our transition rate, we will not account for non-forward scattering contributions because these can be considered to be separate production events. With this restriction, in expanding the S-matrix perturbatively, the main background contribution will come from the expectation values of \eqr{eq:enuint} taken with respect to the electron background states. The scattering amplitude will then receive contributions proportional to powers of \begin{equation} \langle V_\mu^e \rangle \equiv 2 \sqrt{2} G_F \langle n | \bar{e} \gamma_\mu P_L e| n \rangle \end{equation} where $n$ labels some many-body background electron state (not necessarily translationally invariant). Note that the right hand side is proportional to the left handed electron current of state $|n \rangle$. In an experimental setting, we are really interested in ensemble averages of the probabilities (not the averages of the $S$ matrix). However, for macroscopic numbers of electrons, we expect the main contribution to come from a set of degenerate states having the same spatial localization as the macroscopic distribution function. This approximation will break down if the density matrix is not sharply peaked about one set of states giving degenerate contributions to the scattering amplitude. We will assume that such a peaked distribution exists, and we will merely assign macroscopic currents to the expectation value of currents that will arise in the scattering amplitude calculation. This means that we will replace the interaction Hamiltonian density of \eqr{eq:enuint} with the effective density \begin{equation} {\cal H}_I^{\rm eff} = \frac{4 G_F}{\sqrt{2}} \bar{\nu}_e \gamma^\mu P_L \nu_e J_\mu^e \label{nueeff} \end{equation} where $J_\mu^e$ is the macroscopic left-handed electron current. For example, for an unpolarized $e^{\pm}$ background one would employ---based on consideration of the sum over spin states of single particle expectation values of $\bar e \gamma^\mu P_L e$, for example---the following expression: \begin{equation} J_\mu^e = {1\over 2}\int {d^3{\bf p}\over (2\pi)^3}\left[ f_{e^-}({\bf p})-f_{e^+}({\bf p})\right]{p_\mu\over E_{{\bf p}}}, \end{equation} where the $f_{e^\pm}({\bf p})$ are the usual distribution functions, including a factor of two for spin degeneracy. As usual, this procedure neglects higher order correlations. Note that for electrons in thermal equilibrium, our prescription, e.g. \begin{equation} \langle V_\mu^e \rangle = \sqrt{2}G_F(n_{e^-}-n_{e^+}) \delta_{\mu 0} \end{equation} gives the same mass shift as the real time thermal field formalism employed by \cite{notraf}. As another example, the effective potential due to background neutrinos is of interest in the envelope of a supernova/nascent neutron star, where the neutrino flavor composition can affect, for example, the explosion mechanism \cite{fuller} or the outcome of possible heavy element nucleosynthesis \cite{qf1,qf2}. In addition to the $e^\pm$ background, we must consider neutrino-neutrino forward scattering arising from another term in Eq. (\ref{ewham}), \begin{equation} {\cal H}_I = {G_F\over \sqrt{2}}\sum_{i,j}\bar{\nu}_i \gamma^\mu P_L \nu_i \bar{\nu}_j \gamma_\mu P_L \nu_j, \label{nunu} \end{equation} where the indices $i,j$ label the mass eigenstate fields. We work in the mass basis because the external neutrino background consists of on shell states, a point whose consequences were emphasized in Ref. \cite{sigl} (see also Ref. {\cite{qf2}, and references in these). As seen previously, in the perturbative expansion of the S-matrix we will have occasion to take a background expectation value of this interaction (this time with respect to a many-body background neutrino state). Two of the neutrino fields will be paired with fields in the ``production'' and ``detection'' interactions, leaving two other fields whose background expectation value is taken: \begin{equation} \langle {\cal H}_I \rangle = {G_F\over \sqrt{2}}\sum_{i,j}2 \bar{\nu}_i \gamma^\mu P_L \nu_i \langle \bar{\nu}_j \gamma_\mu P_L \nu_j\rangle + {G_F\over \sqrt{2}}\sum_{i,j} 2 \bar{\nu}_i \gamma^\mu P_L \langle\nu_i \bar{\nu}_j\rangle \gamma_\mu P_L \nu_j. \label{nunuave} \end{equation} While the correspondence of the expectation value in the first term of Eq. (\ref{nunuave}) with a macroscopic current is apparent, the meaning of the second term is less clear. We seek guidance by considering the expectation values with respect to the single particle neutrino states $|{\bf q}\,s\,\nu_k\rangle$ of momentum ${\bf q}$, spin $s$, and mass $m_k$. The expectation value in the first term of Eq. (\ref{nunuave}) is \begin{equation} \langle{\bf q}\,s\,\nu_k | \bar\nu_j\gamma^\mu P_L \nu_j |{\bf q}\,s\,\nu_k\rangle = {\delta_{jk}\over(2\pi)^3(2E_{\vq})} \bar u({\bf q}\,s\,\nu_k)\gamma^\mu P_L u({\bf q}\,s\,\nu_k).\label{eval1} \end{equation} We consider a relativistic neutrino background, so that to leading order there are only negative helicity states; then the momentum space spinors in Eq. (\ref{eval1}) are approximately \begin{equation} u({\bf q}\,s\,\nu_k) \simeq \pmatrix{\sqrt{2E_{\vq}}\xi^-(\hat{\bf q})\cr 0}, \ \ \ \ \ \xi^-(\hat{\bf q}) = \pmatrix{-\sin(\theta/2)e^{-i\phi} \cr \cos(\theta/2)}, \label{spinor} \end{equation} where $(\theta,\phi)$ denote the polar and azimuthal angles that define $\hat{\bf q}$. The first term in Eq. (\ref{nunuave}) becomes \begin{equation} {G_F\over \sqrt{2}}\sum_{i,j}2 \bar{\nu}_i \gamma^\mu P_L \nu_i \langle \bar{\nu}_j \gamma_\mu P_L \nu_j\rangle = \sqrt{2} G_F \sum_{i,j} \chi_i^\dagger {\delta_{jk}\over (2\pi)^3}{q_\mu \bar\sigma^\mu \over E_{\vq}}\chi_i,\label{term1} \end{equation} where $\chi_i$ denotes the upper two components of $P_L\nu_i$ and $\bar\sigma^\mu = (1,-\mbox{\boldmath $\sigma$})$. Turning to the second term in Eq. (\ref{nunuave}), one finds \begin{equation} \langle{\bf q}\,s\,\nu_k | \nu_i\bar \nu_j |{\bf q}\,s\,\nu_k\rangle = -{\delta_{ik}\delta_{jk} \over(2\pi)^3(2E_{\vq})} u({\bf q}\,s\,\nu_k)\bar u({\bf q}\,s\,\nu_k). \end{equation} Employing Eq. (\ref{spinor}), the second term in Eq. (\ref{nunuave}) becomes \begin{equation} {G_F\over \sqrt{2}}\sum_{i,j}2 \bar{\nu}_i \gamma^\mu P_L \langle \nu_i \bar{\nu}_j\rangle \gamma_\mu P_L \nu_j = \sqrt{2} G_F \chi_k^\dagger {\delta_{ik}\delta_{jk}\over (2\pi)^3}{q_\mu \bar\sigma^\mu \over E_{\vq}}\chi_k.\label{term2} \end{equation} Noting the similarity between Eqs. (\ref{term1}) and (\ref{term2}), in the relativistic limit we replace the Hamiltonian density of Eq. (\ref{nunu}) with the effective density \begin{equation} {\cal H}_I^{\rm eff} = \sqrt{2}G_F \sum_{i,j}\bar\nu_i J_{\nu_j}^\mu \gamma_\mu P_L \nu_i + \sqrt{2}G_F \sum_i \bar\nu_i J_{\nu_i}^\mu\gamma_\mu P_L \nu_i, \label{nunueff} \end{equation} where \begin{equation} J_{\nu_i}^\mu = \int {d^3{\bf p}\over (2\pi)^3}\left[ f_{\nu_i}({\bf p})-f_{\bar\nu_i}({\bf p})\right]{p^\mu\over E_{{\bf p}}}. \label{jnu} \end{equation} The flavor fields $\nu_\alpha$ are related to the mass eigenstate fields $\nu_i$ by $\nu_\alpha = \sum_i U_{\alpha i}\nu_i$, where the $U_{\alpha i}$ are elements of a unitary matrix. In the limit of vanishing mixing angles $(U_{\alpha i}= \delta_{\alpha i})$ and a thermal background, the effective interaction of Eq. (\ref{nunueff}) gives the same mass shifts as obtained in Ref. \cite{notraf}. For nontrivial mixing, however, in terms of the flavor fields Eq. (\ref{nunueff}) becomes \begin{equation} {\cal H}_I^{\rm eff} = \sqrt{2}G_F \sum_{\alpha,j} \bar\nu_\alpha J_{\nu_j}^\mu\gamma_\mu P_L \nu_\alpha + \sqrt{2}G_F \sum_{\alpha,\beta} \bar\nu_\alpha\left(U_{\alpha i} J_{\nu_i}^\mu U_{i\beta}\right)\gamma_\mu P_L \nu_\beta. \label{nunueff2} \end{equation} Considered as a matrix in flavor space, the quantity in parentheses in the second term of Eq. (\ref{nunueff2}) contains off-diagonal elements. It is clear from our derivation that the presence of these off-diagonal terms derives from the fact that the background neutrinos are mass eigenstates, since they must be on shell. This origin of off-diagonal flavor space terms in the background potential due to neutrino-neutrino scattering was pointed out in Ref. \cite{sigl}. In calculations of the effects of neutrino flavor oscillations in the supernova environment, the supernova core is typically treated as a stationary source of neutrinos free-streaming from a ``neutrinosphere,'' with the flux at the neutrinosphere being taken from large-scale numerical computations. Since the neutrinos forming the background also undergo flavor transformation, self-consistency between the oscillation probability and the background must be achieved. Following the framework of Ref. \cite{sigl}, such a calculation was carried out in Ref. \cite{qf2} in the quantum mechanical picture of neutrino oscillations. This involved a rather complicated procedure involving a flavor basis density matrix to describe the neutrinos above the neutrinosphere. Casual inspection of the form of the second term in Eq. (\ref{nunuave}) would seem to make this kind of approach necessary. However, having shown that this term can plausibly be written in terms of a macroscopic mass basis current, we see that the self-consistency between neutrino background and oscillation probability is most easily achieved by working in the mass eigenstate basis.\footnote{It would seem reasonable to define neutrino flavor distribution functions in the relativistic limit. While {\em at the emission point} (i.e. the neutrinosphere) one could argue that these would be related to the mass basis distribution functions by $f_{\nu_i}({\bf p}) = \sum_\alpha |U_{\alpha i}|^2 f_{\nu_\alpha}$, at points above the neutrinosphere the relation between these sets of distribution functions is rather complicated, due to the flavor/mass oscillations of free-streaming neutrinos in a background. The resulting absence of a simple connection between the macroscopic flavor and mass neutrino currents at arbitrary position to plug into Eq. (\ref{nunueff2}) makes working in the mass basis seem much more straightforward.} Given effective interaction Hamiltonians like Eqs. (\ref{nueeff}) and (\ref{nunueff}), the oscillation probability can be computed (in any basis) as described in Secs. \ref{sec:cback} and \ref{sec:vback}. We write Eq. (\ref{jnu}) as \begin{equation} J^\mu_{\nu_i}(r)= \int {E_{{\bf p}}^2\, d E_{{\bf p}}\, d(\cos\theta) d\phi \over (2\pi)^3}\,\left[ f_{\nu_i}(E_{{\bf p}},\cos\theta,r) - f_{\bar\nu_i}(E_{{\bf p}},\cos\theta,r)\right] {p^\mu\over E_{{\bf p}} }, \end{equation} where $p^\mu=(E_{{\bf p}},\,\hat{\bf p} E_{{\bf p}} )$, $r$ is the radial position of a point above the neutrinosphere, and $\theta$ is the angle between the neutrino momentum and the radial direction at the point with radial position $r$. Since the neutrinos are free-streaming, the distribution functions at $r$ can be expressed simply in terms of the ``known'' neutrino distribution functions at the neutrinosphere, e.g.: \begin{equation} f_{\nu_i}(E_{{\bf p}},\cos\theta,r)=\sum_{\nu_j} f_{\nu_j}(E_{{\bf p}},\cos\psi,R)\, P_{\nu_j \rightarrow \nu_i}(E_{{\bf p}},\cos\theta,r), \end{equation} where $R$ is the radius of the neutrinosphere, and $\psi$ is the angle of the neutrino emission with respect to the radial direction at the emission point; this angle is related to $\theta$ by $\cos\psi = \sqrt{1-[(r/R)\sin\theta]^2}$. The dependence of the oscillation probabilities on path length (and the background encountered on a particular path) are implicit in the $r,\theta$ dependence. With an iterative procedure, self-consistency between the macroscopic neutrino currents and the oscillation probabilities should be achieved. \section{Constant Background} \label{sec:cback} Having constructed effective interaction Hamiltonians as described in the last section, we employ the neutrino effective Lagrangian \begin{equation} {\cal L} = \bar\nu \left[\gamma^\mu(i\partial_\mu - V_\mu P_{L}) - M\right] \nu, \end{equation} where $M$ is the mass matrix and $V_\mu$ is a uniform background potential matrix. We make no assumptions about the number of neutrino generations or the structure of the potential matrix (other than to keep in mind that it might be singular). The canonical anticommutation relations yield the equation satisfied by the Green's function $G(x,y)$, \begin{equation} \left[\gamma^\mu(i\partial_\mu - V_\mu P_{L}) - M\right] G(x,y) = \delta^4(x-y), \label{green} \end{equation} where $i\,G(x,y)\equiv \langle T{\psi(x)\bar\psi(y)}\rangle_0$. With our convention for the $\gamma$ matrices it is convenient to define the $2\times 2$ (in spinor space) matrices $G_{IJ}$, the ``chiral blocks'' of the Green's function. Specifically, $G_{IJ}$ is the nonzero $2\times 2$ submatrix of $P_I G P_J$, where $I,J$ can take the values $L,R$. In Sec. II we noted that, with the assumption of V-A interactions, $G_{LR}$ is the object of interest. We also saw in Sec. II that with the assumption of stationarity it is natural to Fourier transform the time variable while maintaining interest in the spatial dependence of the Green's function. Defining $J = M^{-1} G_{RR}$, from Eq. (\ref{green}) we find \begin{equation} G_{LR}(\omega,{\bf x},{\bf y}) = (\omega + i\mbox{\boldmath $\sigma$} \cdot \mbox{\boldmath$\nabla$}) J(\omega,{\bf x},{\bf y}), \label{glr} \end{equation} where $f(x,y) = \int{d\omega\over 2\pi}\, e^{-i\omega(x^0-y^0)} f(\omega, {\bf x},{\bf y})$. In the context of neutrino oscillation experiments we are in interested in well-separated source and detector positions, so we ignore terms in $J$ with more than one factor of $|{\bf x}-{\bf y}|$ in the denominator, \begin{eqnarray} J(\omega,{\bf x},{\bf y}) &=& \int {d^3 p \over (2\pi)^3}\, e^{i{\bf p} \cdot ({\bf x}-{\bf y})} \,J(\omega, {\bf p}) \nonumber \\ &=& \int_0^{\infty} {du \, u^2 \over (2\pi)^3} \int_0^{2\pi} d\phi \, {1\over i u |{\bf x}-{\bf y}|} \left[ e^{iu|{\bf x}-{\bf y}|} J(\omega,u,\hat{\bf p}=+\hat{\bf r}) \right.\nonumber\\ & &\left. - e^{-iu|{\bf x}-{\bf y}|} J(\omega, u,\hat{\bf p}=-\hat{\bf r}) + {\cal O}\left(J\over u|{\bf x}-{\bf y}|\right)\right], \label{j1} \end{eqnarray} where we have integrated the $\cos\theta$ integral by parts, and defined $u\equiv |{\bf p}|$ and $\hat{\bf r}\equiv ({\bf x}-{\bf y})/|{\bf x}-{\bf y}|$. It is evident that the two leading terms are azimuthally symmetric, and that their the sum is even in $u$. Furthermore, the Feynman boundary conditions should ensure that the two leading terms give equal contributions. We then have \begin{equation} J(\omega,{\bf x},{\bf y}) \simeq {1\over (2\pi)^2 i |{\bf x}-{\bf y}|} \int_{-\infty}^{\infty} du\, u\, e^{iu|{\bf x}-{\bf y}|} J(\omega,u,\hat{\bf p}=+\hat{\bf r}). \label{jxy} \end{equation} From Eq. (\ref{green}), we find that $J$ obeys the momentum space equation \begin{equation} \left[\omega^2 - |{\bf p}|^2 - M^2 - \omega V^0 + {\bf p} \cdot {\bf V} + \mbox{\boldmath$\sigma$} \cdot \left(V^0{\bf p} - \omega{\bf V} + i {\bf V} \times {\bf p}\right)\right] J(\omega,{\bf p}) = 1, \label{jmom} \end{equation} or $D(\omega,u,\hat{\bf p}=+\hat{\bf r}) J(\omega,u,\hat{\bf p}=+\hat{\bf r}) = 1$. Detailed expressions for the spinor space elements of $D(\omega,u,\hat{\bf p}=+\hat{\bf r})$ for general orientation of ${\bf r}$ are not particularly illuminating. However, it is easy to formally express the spinor space elements of $J(\omega,u,\hat{\bf p}=+\hat{\bf r})$ in terms of the elements of $D$: \begin{eqnarray} J^{11}(\omega,u,\hat{\bf p}=+\hat{\bf r})&=& \left[ D^{11} - D^{12}(D^{22})^{-1} D^{21}\right]^{-1}, \label{j11}\\ J^{22}(\omega,u,\hat{\bf p}=+\hat{\bf r})&=& \left[ D^{22} - D^{21}(D^{11})^{-1} D^{12}\right]^{-1}, \label{j22}\\ J^{12}(\omega,u,\hat{\bf p}=+\hat{\bf r})&=& -(D^{11})^{-1}D^{12}J^{22},\\ J^{21}(\omega,u,\hat{\bf p}=+\hat{\bf r})&=& -(D^{22})^{-1}D^{21}J^{11}. \end{eqnarray} These results are valid without any relativistic limit assumptions. Given specific mass and potential matrices, one could solve explicitly for $J(\omega,u,\hat{\bf p}=+\hat{\bf r})$. To study neutrino oscillations, we then need $J(\omega,{\bf x},{\bf y})$, whose behavior is seen from Eq. (\ref{jxy}) to be determined by the poles of $J(\omega,u,\hat{\bf p}=+\hat{\bf r})$ with positive imaginary part (as determined by the Feynman boundary conditions). A few general comments regarding these poles are in order. Consider for example $J^{22}$, which can be expressed \begin{equation} J^{22}(\omega,u,\hat{\bf p}=+\hat{\bf r}) = ({\rm det\ } D^{11}) \left[({\rm det\ } D^{11}) D^{22} - D^{21}(C^{11})^T D^{12}\right]^{-1}, \end{equation} where $(C^{11})^T$ is the transpose of the matrix of cofactors of $D^{11}$. Since the diagonal elements of $D^{11}$ and $D^{22}$ are second order in $u$, $({\rm det\ } D^{11})$ is of order $2n$ in $u$, where $n$ is the number of neutrino generations; and overall the denominator of $J^{22}(\omega,u,\hat{\bf p}=+\hat{\bf r})$ will be a polynomial of order $4n$ in $u$. This is sensible in terms of a quasiparticle picture associated with the propagator: Each neutrino field, with two spin states each for particles and antiparticles, represents four states. For a single vacuum field the masses of these four states are degenerate; however, the presence of a parity and rotational invariance violating potential breaks this degeneracy. Let us examine the simplifications that occur in the relativistic limit and with source and detector localization. In the relativistic limit the poles contributing to the integral in Eq. (\ref{jxy}) take the form \begin{equation} u \simeq |\omega| - {\tilde m^2\over 2|\omega|} + i\epsilon, \end{equation} with the Feynman boundary conditions imposed by giving the ``masses'' $\tilde m^2$ a small negative imaginary part. (There are also negative poles, with negative imaginary parts, that do not contribute to the integral; these factors each become $\simeq 2|\omega|$ when evaluated at the positive poles.) Furthermore, following the discussion of Sec. \ref{sec:general} regarding spatial localization, Eq. (\ref{jxy}) takes the form \begin{equation} J(\omega,{\bf x},{\bf y}) \simeq {2|\omega|\,e^{i \omega \hat{\bf L} \cdot({\bf x}-{\bf y})} \over 4\pi |{\bf x}_S-{\bf y}_D|} \sum_j e^{-i{\tilde m_j^2\over2|\omega|}|{\bf x}_S-{\bf y}_D|} \left. \left[\left(u - |\omega| + {\tilde m_j^2\over 2|\omega|}\right) J\left(\omega,u,\hat{\bf p}=\frac{\omega}{|\omega|}\hat{\bf L}\right)\right] \right|_{u\rightarrow |\omega| - {\tilde m_j^2\over 2|\omega|}}, \label{jxy2} \end{equation} where the sum is over the poles with positive imaginary parts. We recall that $\omega$ is fixed by energy delta functions, to a positive value for neutrino oscillations and a negative value for antineutrino oscillations. Because the spatial localization sets $\hat{\bf r}=\pm \hat{\bf L}$ (with $\hat{\bf L}$ taken to be the third spatial direction), the matrix $D$ takes the relatively simple spinor space form \begin{equation} \left[D\left(\omega,u,\hat{\bf p}=\frac{\omega}{|\omega|} \hat{\bf L}\right)\right]= \pmatrix{d -\omega\left(1 - {u\over|\omega|}\right)V^0 + \omega \left({u\over|\omega|}-1\right)V^3 & -\omega \left(1 + {u\over|\omega|}\right)(V^1 - iV^2) \cr -\omega\left(1 - {u\over|\omega|}\right)(V^1 + iV^2) & d - \omega\left(1 + {u\over|\omega|}\right)V^0 + \omega\left( {u\over|\omega|} + 1\right)V^3}. \label{dmatrix} \end{equation} where $d \equiv \omega^2-u^2-M^2$. Next we examine the momentum space pole structure of $J^{22}$ in the relativistic limit. In \eqr{j22}, since the residues of only the positive poles contribute, we can replace $u$ by $|\omega|$ whenever it multiplies a component of $V^\mu$, committing errors of only ${\cal O}(\omega V^0 /\omega^2)$ or less with respect to other terms present. In that case, $D_{21}$ vanishes and $J^{22}\rightarrow (D^{22})^{-1}$ with \begin{equation} D^{22} \rightarrow \omega^2 - u^2 - M^2 - \frac{\omega}{|\omega|} 2 q \cdot V \label{d22}\\ \end{equation} where $q =(|\omega|,\hat{\bf L} |\omega|)$ is the same as the neutrino momentum defined just below \eqr{detamp}. We note that the denominator of $J^{22}$ is now only of order $2n$ in $u$; thus in the relativistic case, two of the quasiparticle propagating states are projected out. This is because in the relativistic limit the spin states naturally coincide with the chiral states. This is confirmed by noting that in Eq. (\ref{j22}), for example ($\omega >0$ case), as $D^{21}\rightarrow 0$ as $u\rightarrow \omega$, half of the poles contributing to $J^{22}$ come from (det $D^{11}) \rightarrow 0$. But for $u\rightarrow \omega$, $D^{11} \rightarrow \omega^2 - u^2 -M^2$. Thus these poles correspond to the vacuum masses; these are the right handed particle states and left handed antiparticle states whose masses are unaffected by the left handed effective potential. As $D^{21}$ reaches zero, the contribution of these poles vanishes completely. For $|\omega||{\bf x}-{\bf y}|\gg 1$, Eq. (\ref{glr}) becomes \begin{equation} G_{LR}(\omega,{\bf x},{\bf y}) \simeq \omega (1 - \mbox{\boldmath $\sigma$} \cdot \hat{\bf L}) J(\omega,{\bf x},{\bf y}). \label{glr2} \end{equation} Since we have chosen $\hat{\bf L} =({\bf y}_D-{\bf x}_S)/|{\bf y}_D-{\bf x}_S|$ to coincide with the third spatial dimension, the only nonzero component of $G_{LR}$ is $G_{LR}^{22}$ which for neutrino oscillations is $G_{LR}^{22}(|\omega|;{\bf y},{\bf x}) = 2|\omega| J^{22}(|\omega|,{\bf y},{\bf x})$, where the spinor space indices are exhibited and the mass/flavor indices are suppressed (note that $J^{21}$ vanishes), and for antineutrino oscillations is $G_{LR}^{22}(-|\omega|;{\bf x},{\bf y}) = -2 |\omega| J^{22}(-|\omega|,{\bf x},{\bf y})$. Thus we see that the Green's function takes the required form of Eqs. (\ref{gform}). The matrix $M^2 + 2q\cdot V$ (or $M^2 - 2 q\cdot V$) of \eqr{d22} is precisely the effective mass matrix $\tilde M^2$ appearing in the usual quantum mechanical model of neutrino (or antineutrino) oscillations. The effective mass matrix can be diagonalized by a unitary transformation, $\tilde U \tilde M^2 \tilde U^\dagger=1$. Thus for neutrino oscillations, for example, \begin{equation} (D^{22})^{-1}_{\beta\alpha} = \tilde U_{\beta j} \tilde U_{\alpha j}^* (\omega^2 - u^2 - \tilde m_j^2+i\epsilon)^{-1}, \end{equation} where $\tilde U_{\beta j}$ are the elements of $\tilde U$ and $\tilde m_j^2$ are the eigenvalues of $\tilde M^2$. Then Eq. (\ref{glr2}) becomes, using Eq. (\ref{jxy2}), \begin{equation} G_{LR}^{\beta\alpha} (\omega,{\bf y},{\bf x}) = -|\omega|(1 - \mbox{\boldmath $\sigma$} \cdot \hat{\bf L}) {e^{i|\omega|\hat{\bf L}\cdot({\bf y}-{\bf x})} \over 4\pi |{\bf y}_D-{\bf x}_S|} \sum_j \tilde U_{\beta j} \tilde U_{\alpha j}^* e^{-i{\tilde m_j^2\over2|\omega|}|{\bf y}_D-{\bf x}_S|} \label{glr3}, \end{equation} where we have assumed that the various conditions discussed in Sec. II are satisfied. Comparison of Eq. (\ref{glr3}) with Eqs. (\ref{gform},\ref{oscprob}) shows that the oscillation probability derived here is precisely the same as that found in the usual quantum mechanical model. The antineutrino case works out in a similar manner. \section{Nonuniform Background} \label{sec:vback} In this section we consider the case in which the effective potential $V^\mu=V^\mu({\bf x})$, that is, we allow it to vary in space (but not time). From Eqs. (\ref{glr3}),(\ref{gform}), and (\ref{oscprob}), it is clear that in the constant potential case the portion of the Green's function comprising the oscillation amplitude obeys a Schr\"{o}dinger-type equation, the same one used in the standard quantum mechanical picture. While one might think to simply replace the constant potential in this Schr\"odinger equation with a spatially varying one---thus arriving immediately at the standard result---we shall go back a little further in order to see what is being left out in the process. In allowing for spatial variation in $V^\mu$, Eqs. (\ref{glr}-\ref{jxy}) are unchanged; but Eq. (\ref{jmom}) becomes an integral equation, as $J(\omega,{\bf p})$ must be convolved with the momentum space dependence of $V^\mu$. Since such equations are difficult to deal with nonperturbatively, in this section, we take a different route of working with a partial differential equation in coordinate space. The coordinate space version of Eq. (\ref{jxy}) is \begin{equation} \left[\omega^2 +\nabla^2 - M^2 - \omega V^0({\bf x}) -i{\bf V}({\bf x})\cdot\right.${\boldmath $\nabla$}$ -i ${\boldmath $\sigma$}$ \cdot \left(V^0({\bf x})\right.${\boldmath $\nabla$}$ -i\omega{\bf V}({\bf x}) + i {\bf V}({\bf x}) \times ${\boldmath $\nabla$}$\left.\left.\right)\right] J(\omega,{\bf x},{\bf y}) = \delta^3({\bf x}-{\bf y}). \label{jxy3} \end{equation} Unlike in the case of an integral equation, to define the Green's function using this equation, we must also separately specify the boundary condition. We shall assume that the production region is localized to a region of adiabatically constant potential. Furthermore, since we expect the virtual particles to all have the same phase just after being produced, our boundary condition prescription will be that the the Green's function $J$ asymptote to the constant potential Green's function on an infinitesimal sphere centered about ${\bf y}$.\footnote{Note that one must match more than just the limiting singularity of the Green's function at ${\bf x}={\bf y}$ to define a unique solution.} The form of Eq. (\ref{jxy}), together with our experience in the vacuum and constant potential cases, suggests that in the relativistic limit [$M^2/2|\omega|^2 \ll 1, V|\omega|/|\omega|^2 \ll 1$ where we suppressed the matrix indices] we look for solutions of $J$ of the form \begin{equation} J(\omega,{\bf x},{\bf y})=-{e^{i|\omega||{\bf x}-{\bf y}|}\over 4\pi|{\bf x}-{\bf y}|} F(\omega,{\bf x},{\bf y}). \label{jf} \end{equation} With this substitution, \begin{eqnarray} (\nabla^2+\omega^2) J &=& \delta^3({\bf x}-{\bf y}) e^{i|\omega||{\bf x}-{\bf y}|} F - {2|\omega|\,e^{i|\omega||{\bf x}-{\bf y}|}\over 4\pi|{\bf x}-{\bf y}|} \left[{1\over 2|\omega|} \nabla^2 F + i (\hat{\bf r} \cdot \mbox{\boldmath $\nabla$} F) - {1 \over |\omega||{\bf x}-{\bf y}|}(\hat{\bf r} \cdot \mbox{\boldmath $\nabla$} F)\right], \label{ddj} \\ \mbox{\boldmath $\nabla$} J &=& - {2|\omega|\,e^{i|\omega||{\bf x}-{\bf y}|}\over 4\pi|{\bf x}-{\bf y}|} \left[ {i\hat{\bf r} \over 2}F + {1\over 2|\omega|} \mbox{\boldmath $\nabla$} F - {\hat{\bf r} \over 2|\omega||{\bf x}-{\bf y}|} F\right], \label{dj} \end{eqnarray} where as before $\hat{\bf r} \equiv {({\bf x}-{\bf y})/ |{\bf x}-{\bf y}|}$. Requiring the first term on the right hand side of Eq. (\ref{ddj}) to cancel the delta function in Eq. (\ref{jxy3}) gives a boundary condition on $F$, namely (restoring flavor indices) \begin{equation} F^{\beta\alpha}(\omega,{\bf x},{\bf y})|_{{\bf x}\rightarrow{\bf y}} = \delta^{\beta\alpha}. \label{bc} \end{equation} Aside from this boundary condition, we are interested in well-separated ${\bf x}$ and ${\bf y}$ (specifically $|\omega||{\bf x}-{\bf y}|\gg 1$), so that we may ignore the last term of Eqs. (\ref{ddj}) and (\ref{dj}). Then Eq. (\ref{jxy3}) becomes \begin{eqnarray} i (\hat{\bf r} \cdot \mbox{\boldmath $\nabla$} F) + {1\over 2|\omega|} \nabla^2 F - {1\over 2|\omega|} \left[M^2 + \omega V^0 - |\omega|(\hat{\bf r} \cdot {\bf V}) \right.& &\nonumber\\ \left.- \mbox{\boldmath $\sigma$} \cdot \left(V^0|\omega|\hat{\bf r} - \omega{\bf V} + i |\omega|{\bf V} \times \hat{\bf r} \right)\right] F + {\cal O}\left({V|\omega| \over |\omega|^2}|\mbox{\boldmath $\nabla$} F|\right)&=&0, \label{ddf} \end{eqnarray} where in accordance with the relativistic condition $V|\omega|/|\omega|^2 \ll 1$, we will neglect the terms represented by ${\cal O}(V|\omega||\mbox{\boldmath $\nabla$} F| / |\omega|^2)$ in comparison with the first term of Eq. (\ref{ddf}). One can distinguish three cases: (1) $|\mbox{\boldmath $\nabla$} F| \gg \epsilon F$, where $\epsilon = V^0 + \tilde{M}^2/(2|\omega|)$ and $\tilde{M}^2$ denotes the largest mass matrix eigenvalue squared; (2) $|\mbox{\boldmath $\nabla$} F| \sim \epsilon F$; and (3) $|\mbox{\boldmath $\nabla$} F| \ll \epsilon F$. One can argue that case (1) is not interesting since all terms leading to flavor mixing are rendered negligible. Case (3) is also not of present interest because one can argue using Eq. (\ref{ddf}) that it violates our relativistic assumption. In case (2), $|\nabla^2 F/(2|\omega|)|$ can be neglected compared with $| \mbox{\boldmath $\nabla$} F|$, provided that $|\mbox{\boldmath $\nabla$} V^0|/\epsilon^2 \lesssim 1$. Writing $F' \equiv \hat{\bf r}\cdot\mbox{\boldmath $\nabla$} F$, Eq. (\ref{ddf}) becomes \begin{equation} iF' + {1\over 2|\omega|} D(|\omega|,{\bf x})F = 0, \label{fprime} \end{equation} where the spinor space elements of $D(|\omega|,{\bf x})$ for the particular case when ${\bf x}$ lies along the third spatial axis $\hat{\bf L}$ are given by Eq. (\ref{dmatrix}) with $V\rightarrow V({\bf x})$ and $u\rightarrow |\omega|$. Since Eq. (\ref{glr2}) holds under the assumptions of case (2), together with the spatial localization of the source and detector, we see that $G_{LR}^{22}$ is the only nonvanishing component. Hence, from Eqs. (\ref{fprime}),(\ref{jf}),(\ref{gform}), and (\ref{oscprob}), we find that the neutrino oscillation amplitude $H$ obeys the Schr\"odinger equation \begin{equation} iH' = {1\over 2|\omega|}\left[M^2 + 2 q \cdot V({\bf x})\right]H, \end{equation} where $q = (|\omega|,\hat{\bf L}|\omega|)$. Similarly, in the case of antineutrino oscillations $(\omega < 0)$, the oscillation amplitude obeys \begin{equation} i\bar H' = {1\over 2|\omega|}\left[M^2 - 2 q \cdot V({\bf x})\right]\bar H. \end{equation} Before concluding this section, a remark regarding the boundary conditions is in order. Note that $F$ of \eqr{fprime} satisfying the boundary condition \eqr{bc} is in general different from $F$ satisfying \eqr{jxy3} with its associated boundary condition [described just below \eqr{jxy3}]. In particular, although \eqr{fprime} is valid naively only far away from ${\bf y}$, as we threw out the last terms of Eqs. (\ref{ddj}) and (\ref{dj}), we still insisted on the boundary condition \eqr{bc} at ${\bf y}$ to be the same as the boundary condition that would have been used for the exact equation. To justify this, we must show that the terms that we threw out are negligible even near the origin. We can argue this by noting that for case (2), we are already assuming $|{\boldmath \nabla} V^0|/\epsilon^2 \lesssim 1$ which turns out to imply (by expanding the potential to linear order in Taylor series about ${\bf y}$) that in the relativistic limit the fractional variation of $V^0$ is much smaller than 1 until $|\omega ({\bf x} - {\bf y})| \gg 1$ (after which the terms proportional to $1/|{\bf x} - {\bf y}|$ that we threw out are negligible). That means that the potential can be treated as a constant until the terms proportional to $1/|{\bf x} - {\bf y}|$ become negligible. This implies that one can place the boundary condition for the varying potential case on a sphere (centered about ${\bf y}$) on which \eqr{fprime} is valid using the solution to the constant potential case. As we saw in the last section, since the exact solution to the constant potential case on this sphere (with the appropriate boundary condition) is, up to relativistically suppressed terms, the same as the solution obtained by \eqr{fprime} with \eqr{bc}, we can just set the boundary condition for the varying potential case using \eqr{bc} as well. Note, however, that since our argument depends on the adiabaticity of the potential near the virtual particle production point, in other situations, one may need to be more cautious with the boundary conditions. Thus under appropriate conditions the results of the usual simplified picture are confirmed, including the boundary condition $H^{\beta\alpha}(\omega,{\bf x},{\bf y})|_{{\bf y}\rightarrow{\bf x}}= \delta^{\beta\alpha}$. \section{Conclusion} \label{sec:conc} Starting from quantum field theory (QFT), we have defined a physically meaningful flavor oscillation probability; determined that portion of the neutrino propagator that comprises the oscillation amplitude; and derived the ``Schr\"{o}dinger equation'' for that amplitude in the presence of spatially varying background matter. As expected, the ``Schr\"{o}dinger equation'' really corresponds to a time independent one since its derivation depends on the time-independence of the effective potential. In fact, the usual quantum mechanical approach is only really suited to problems in which oscillations occur in space only (that is, stationary systems like that studied here) or time only (e.g. the thermal bath in the early universe). While we have assumed the stationary case here, the basic framework could also be used to study oscillations in space in the presence of a time-dependent background. Ultimately, the description of flavor oscillating neutrinos in space and time in more general systems---i.e. those that do not lend themselves to interpretation in terms of a ``source'' and ``detector''---would require a formulation in terms of density matrices (cf. Ref. \cite{sigl}) or Wigner functions (cf. Ref. \cite{sireraperez}). For situations that can be interpreted in terms of a ``source'' and ``detector,'' we have also reviewed the conditions under which QFT will be useful in describing the neutrino oscillation process.\footnote{ Many of these conditions were noted in Refs. \cite{rich,grimusstock,camp,kier,cohere,ioapilaf}. } As long as we are looking in the regime in which the virtual neutrino goes on shell (rendering any propagator radiative corrections to be negligible or be merely a constant shift), the many body aspect of QFT is rendered irrelevant. In that case, only the production/detection vertex structure and the spins of the neutrinos are missing from the usual quantum mechanical treatment. These, in general, are less difficult to accommodate in the quantum mechanical treatment than the many body effects. Still, we believe them to be more straightforwardly accommodated in the quantum field theoretical treatment. For weak interactions, the chiral nature of the interactions combined with the relativistic nature of the on shell neutrinos suppresses all but one spin degree of freedom. Finally, the smallness of the neutrino mass splittings as well as the neutrinos going on shell allows one to factor out the production/detection part of the neutrino scattering process from the ``oscillation'' part, reducing the problem to the usual quantum mechanical system involving a single spinless particle. To state this another way, we have argued that in the context of stationary systems, the quantum field theoretic formulation used in this work is not of much use except when one or more of the following is true: the on-shell neutrino momentum is nonrelativistic, the production or detection vertices are non-chiral, the external particle wave packets vary appreciably about the pole value (value of the wave packet evaluated at the neutrino momenta) over momentum variations order of the inverse source-detector distance $1/L$, or the effective mass splittings (determined by the effective potential including the background matter contributions) are large compared to the spread in the momenta of the external production/detection particles. When the neutrinos are non-relativistic or the interactions are not chiral, more than one spin contributes per amplitude. In that case the usual quantum mechanical treatment must be modified to incorporate the effects due to the various spin components. In particular, in this case an oscillation probability cannot be determined apart from the production and detection processes since two neutrino spin contributions are summed before the amplitude is squared. This is, of course, automatically accounted for in a quantum field theoretic treatment. Also, if the wave packet varies appreciably about the pole value when the momentum is varied by ${\cal O}(1/L)$, the wave functions of the external particles creating and absorbing the virtual neutrino do not simply factorize out of the oscillation probability amplitude. Ref.\ \cite{ioapilaf} is a study of one such situation. They find that the oscillation amplitude can exhibit a novel ``plane-wave'' behavior. Just as most effects, this probably can also be accounted for in the quantum mechanical formulation, but it is much easier in the quantum field theoretical treatment used in this paper. The involvement of the details of the external particles' wave function in determining the oscillation probability also applies when the effective mass splitting is much larger than the momentum spread in the external wave packets, but in such cases, it is not clear whether any neutrino oscillations can be observed in general because of the strong suppression of the amplitudes (however, see for example Ref.\ \cite{ioapilaf}). We here make a few comments regarding Majorana vs. Dirac neutrinos. If the neutrinos were Majorana instead of Dirac, the only general arguments that would change are those dependent upon the existence of a right handed neutrino. Although we couched the mathematics in the Dirac spinor formalism, owing to the assumption of chiral nature of neutrino interactions and the relativistic limit, none of our general arguments depended upon the existence of a right handed neutrino. Hence, our conclusions for the relativistic limit are also valid for Majorana neutrinos. (Formulae for ``neutrinos'' apply to negative helicity Majorana neutrinos, while those for ``antineutrinos'' apply to positive helicity Majorana neutrinos.) From our justification of the ``Schr\"{o}dinger equation'' for stationary, flat spacetime systems, we expect that the heuristic ansatz used in Ref.\ \cite{cardall} for studying neutrino oscillations in a stationary curved spacetime to be valid to the extent that the flat spacetime treatment is valid. In a nonstationary curved spacetime, in addition to the time dependence of the potential, there may arise extra complications of using the S-matrix formalism due to the nontrivial Bogoliubov transformations of the asymptotic states. This also deserves further investigation. \acknowledgements{ CYC thanks Georg Raffelt for helpful conversations. We thank the warm hospitality of the Theoretical Astrophysics group at Fermilab where part of this work was carried out. This work was started at the GAAC sponsored meeting held at the Aspen Center for Physics during the summer of 1998. CYC is supported by DOE grant FG02-87ER40317.}
\section{Introduction} Globular clusters have long been recognized as excellent fossil records of the formation history of their host galaxies (Ashman \& Zepf 1998 and references therein). They also provide critical testbeds for the study of stellar evolution and stellar dynamics. However, the formation process of globular clusters themselves is not well understood. One hypothesis is that merger-induced starbursts are favorable environments for globular cluster formation (Schweizer 1987, Ashman \& Zepf 1992). Ashman \& Zepf (1992) specifically predicted that Hubble Space Telescope (HST) images of gas-rich mergers would reveal young globular clusters, readily identifiable through their very compact sizes, high luminosities, and blue colors. This prediction has been dramatically confirmed. Initial discoveries of compact, bright, blue star clusters in HST images of the peculiar galaxy NGC~1275 by Holtzman et al.\ (1992) and in the well-known galaxy merger NGC~7252 by Whitmore et al.\ (1993), have been followed up by the observations of similar objects with the characteristics of young globular clusters in a large number of starbursting and merging galaxies (see list in Ashman \& Zepf 1998). The identification of these compact, bright, blue objects as young star clusters has been confirmed by ground-based spectroscopy in several systems (e.g.\ Schweizer \& Seitzer 1998, Brodie et al.\ 1998, Zepf et al.\ 1995, Schweizer \& Seitzer 1993). There are even possible mass estimates from high resolution spectroscopy of a few nearby examples (e.g.\ Ho \& Filippenko 1997). These observations provide significant support for the idea that globular clusters form in galaxy mergers and strong starbursts. They also suggest that globular cluster formation may be a regular part of the starbursting process. The empirical evidence for globular cluster formation in these environments is broadly consistent with the hypothesis that globular clusters primarily form in mergers and starbursts rather than in other sites. Other globular cluster formation scenarios appear to have difficulties accounting for the observational properties of globular cluster systems (e.g.\ Ashman \& Zepf 1998, Harris 1996 and references therein). In particular, correlations between cluster and host galaxy properties and the absence of dark matter are problematic for primordial globular cluster formation models (e.g.\ Peebles \& Dicke 1968, Rosenblatt et al.\ 1988). Similarly, thermal instability models for globular cluster formation (e.g.\ Fall \& Rees 1985) appear to be unable to account for the absence of a correlation between globular cluster metallicity and mass along with the high metallicities of typical globular clusters ([Fe/H] $> -1$). The discovery of young globular cluster systems in nearby starbursts and galaxy mergers opens up the possibility of more detailed, empirical studies of the formation and evolution of globular clusters. One of the questions that remains to be answered is the efficiency with which globular clusters form in starbursts and mergers. This efficiency is critical for determining if most or all globular clusters can form in merger-like conditions. The efficiency can also constrain models of the formation the clusters themselves. For example, models in which globular clusters form as cores in much larger clouds predict low efficiencies, while models in which typical molecular clouds are compressed may be more efficient. A closely related question is the dynamical evolution of globular cluster systems. Most studies to date have concentrated on developing theoretical models and matching these to the properties of old globular cluster systems that have undergone evolution over most of a Hubble time. Attempts to infer the initial population and the effects of evolution from the remnant population are difficult. Observations of young cluster systems can provide valuable input into the initial conditions and early dynamical evolution of globular cluster systems. This is not only true of the mass (luminosity) function, but also of the radii and densities with which the clusters form. The efficiency of globular cluster formation in mergers and of the dynamical evolution of globular cluster populations also have significant implications for the use of globular cluster systems as fossil records of the formation history of their host galaxies. For example, Ashman \& Zepf (1992) predicted that if elliptical galaxies form by mergers, they should have two or more populations of globular clusters. One of these populations originates from the halos of the progenitor spirals and is therefore spatially extended and metal-poor, while the other forms during the merger and is thus more spatially concentrated and metal-rich. This prediction of multiple populations in the globular cluster systems of ellipticals formed by mergers has now been confirmed in many cases (e.g.\ Ashman \& Zepf 1998 and references therein). However, it has not yet been clearly demonstrated that the efficiency of cluster formation in galaxy mergers is sufficient to account for the metal-rich globular cluster population observed in elliptical galaxies. Furthermore, although there are strong theoretical arguments that the mass function of globular cluster systems evolves significantly over time to resemble the log-normal mass function of old globular cluster systems (e.g.\ Gnedin \& Ostriker 1997, Murali \& Weinberg 1997a), this evolution has not been demonstrated observationally. The goal of this paper is to address the questions of the formation and evolution of globular clusters through the study of the galaxy merger NGC~3256. The HST observations on which this study is based and the analysis of these data are presented in $\S 2$. The resulting sample of a large number of compact, bright, blue objects in NGC~3256 is examined in detail in $\S 3$. This section includes the determination of the relationship between the magnitudes, colors, and radii of the young cluster sample and the luminosity function. The implications of the results for the formation efficiency and dynamical evolution of globular cluster systems are discussed in $\S 4$, and the conclusions are given in $\S 5$. \section{Observations and Data Reduction} \subsection{Target Galaxy} We utilized HST and WFCP2 to obtain high resolution images of the galaxy NGC~3256. This galaxy was selected for our program because it has long been identified as a galaxy merger (e.g.\ Toomre 1977) and is fairly nearby, with $cz_{\odot} = 2820$ $\,{\rm km\,s^{\scriptscriptstyle -1}}$\ (English et al.\ 1999) which places NGC~3256 at a distance of 37 Mpc for H$_0$ = 75 km s$^{-1}$ Mpc$^{-1}$. As shown in Figures 1 and 2 (plates 1 and 2), the central region of NGC~3256 has star forming knots, dust lanes, and loops, along with a more extended, smoother component. In the radio continuum and at $2.2 \mu$ there appear to be two nuclei separated by about $5''$, or about 1 kpc (Norris \& Forbes 1995, Kotilainen et al.\ 1996, Doyon, Joseph, \& Wright 1994). Tidal tails can also be seen in the optical images in Figure 1, and have been shown to extend out to $\sim$ 50 kpc in HI (English et al.\ 1999). Toomre (1977) placed it in the middle of his sequence of disk galaxy mergers, suggesting that it is dynamically older than the NGC~4038/4039 system (the Antennae), but younger than NGC~7252. Of the eleven mergers on the Toomre list, NGC~3256 also has the most molecular gas ($1.5 \times 10^{10}$ \hbox{$\thinspace M_{\odot}$}, Casoli et al.\ 1991, Aalto et al.\ 1991, Mirabel et al.\ 1990), and is the brightest in the far-infrared ($L_{FIR} = 3 \times 10^{11}$ \hbox{$\thinspace L_{\odot}$}, Sargent et al.\ 1989). \subsection{HST Observations} The WFPC2 images of NGC~3256 were obtained with the Planetary Camera (PC) centered on the galaxy. At a distance of 37 Mpc, each PC pixel is 8 pc, and the PC covers a total of 7 kpc $\times$ 7 kpc, encompassing the starburst region identified in previous studies. The PC data centered on NGC~3256 are the subject of this paper. The data at larger radii will be discussed in future papers. We imaged NGC~3256 in the F450W and F814W filters. Two equal exposures were obtained through each filter, with total exposures times of 1800s in F450W and 1600s in F814W. For each filter, the two exposures were combined utilizing a cosmic-ray rejection routine kindly provided by Rick White. As a check on this procedure, we also performed the more standard CCREJECT task in STSDAS on the images in each filter, and then set a strict criterion for matching the object lists between the two filters. The final results were very similar to those produced by White's routine (c.f.\ Schweizer et al.\ 1996, Miller et al.\ 1997). A visual examination of the few differences favored the results of White's routine, so we used these combined images for further analysis. In any case, the number of compact sources observed is far greater than any possible residual defects. The resulting combined images are shown in Figure 2 (Plate 2). \subsection{Cluster Identification} A wealth of blue, compact objects is revealed in the HST images shown in Figure 2 (Plate 2). In order to determine the magnitudes and sizes of the compact objects discovered in the HST images, we first used the DAOFIND task in IRAF to identify objects. This task convolves the image with a Gaussian kernel, finds the best fitting Gaussian function at each point, and then searches for density enhancements which are both greater than a given threshold value and the brightest density enhancement in a localized region determined by the width of the Gaussian kernel. For this analysis, we set the FWHM of the Gaussian to be 2.8 pixels, which is the apparent width expected for an object with a true FWHM of roughly 2 pixels. We also applied broad cuts with the DAOFIND sharpness and roundness criteria to eliminate a few extremely diffuse or sharp features. There are two notable effects of identifying objects in this standard way. One is that it introduces a selection bias against objects significantly larger than the smoothing kernel. This is a direct result of the search for density enhancements on a given scale. Although not an issue if all of the objects are unresolved or marginally resolved, this selection effect needs to be accounted for in studies of the distribution of object sizes. A second aspect of DAOFIND is that the threshold is defined globally. Several other globular cluster searches have been performed using a local threshold, rather than a global one (e.g.\ Kundu et al. 1998, Carlson et al.\ 1998, Miller et al.\ 1997). Although this has the advantage of giving a uniform number of false detections over the image, it does so at the cost of producing a non-uniform magnitude limit across the frame. As this is critical for our purposes, we retain the global threshold, and simply set it high enough that the probability of a spurious source in regions of high background (noise) is negligible. Perhaps the most critical aspect is that the detection algorithm is well-understood, and can be run on a variety of artificial datasets to explore the success with which is recovers objects of various luminosities, colors, and sizes. \subsection{Cluster Photometry} The next step is to determine the magnitudes of the identified objects. Because of crowding, variable background, and signal to noise, it is not possible to determine the brightness profile of the objects out to large radius. Therefore we perform aperture photometry from one to several pixels in radius, and correct these modest apertures to total magnitudes. If the objects were unresolved, the aperture correction to total magnitude would be straightforward, as the HST point spread function (psf) is reasonably well-understood. Moreover, an aperture of several pixels incorporates the majority of the light from an unresolved source, even in the PC, so the overall correction is not a large one. However, the objects we detect in NGC~3256 are resolved, as expected for objects with sizes like those of Galactic globular clusters at the distance of NGC~3256. In this case, the aperture corrections depend on both the psf and the intrinsic radial profile of the object. There is a limited amount of spatial information in the surface brightness profile within the few pixels radius out to which it can be reliably measured. Therefore, if a form of the profile is assumed, the radial scale of that profile can be determined by the difference between magnitudes at small radii. A Gaussian shape for the intrinsic profile of the clusters has been adopted in most previous work (e.g.\ Whitmore et al.\ 1993, Whitmore \& Schweizer 1995, Schweizer et al.\ 1996). In order to facilitate comparison with these studies, we also adopt a Gaussian profile for the clusters, although we note that this will tend to underestimate the total magnitude if the clusters follow a more extended profile, such as a modified Hubble law (e.g.\ Holtzman et al.\ 1996, Carlson et al.\ 1998, Ostlin et al.\ 1998). In detail, we determine the size of each object by comparing the difference between magnitudes within apertures of one and three pixels ($m_3 - m_1$) to a table of values for a wide range of Gaussian profiles convolved with the HST psf given by the TINYTIM software (Krist 1993). This is done in both the B and the I filters, and the resulting intrinsic FWHM is taken as the average of the two. We tested this procedure by using I-band (F814W) observations of unsaturated stars in the globular cluster Omega Cen as the basis for the HST psf. This psf gives the same inferred sizes as the TINYTIM psf when the intrinsic input FWHM is greater than about 0.5 of a PC pixel. For objects with intrinsic sizes less than 0.5 of a PC pixel, a slightly ($\sim 10\%$) smaller size is inferred with the Omega Cen psf than with the TINYTIM psf. This difference in inferred size due to different psfs is much smaller than the uncertainty introduced by the assumption of a form for the intrinsic surface brightness profile of globular clusters with only a single free parameter. Therefore, we adopt the TINYTIM results for the remaining analysis. We note that although the absolute errors in total magnitudes and half-light radii due to the requirement of an assumed profile form for the clusters in the sample can be significant, they should give good {\it relative} sizes and total magnitudes, providing the clusters are roughly similar to each other in structural characteristics. Moreover, any systematic error is likely to affect both filters, so that the colors will be mostly unaffected. The total magnitude of each object is determined by applying the aperture correction from the magnitude within an aperture of 2 pixels in radius to a total magnitude, appropriate for the measured size for each object. This aperture correction for an object of typical size in our sample is roughly 0.85 magnitudes in B and 1.0 magnitudes in I. It compares to 0.25 magnitudes in B and 0.48 in I for completely unresolved objects. These differences emphasize both that our objects are significantly resolved, and that the colors are largely unaffected by the correction to total magnitudes (cf.\ Holtzman et al.\ 1996). We correct the magnitudes and colors of the objects for Galactic reddening using the dust maps of Schlegel, Finkbeiner, \& Davis (1998). These give $E(B-V)= 0.12$, and extinction in our HST filters of $A_{F450W} = 0.47$ and $A_{814W}=0.23$. The absolute photometric calibration to the standard B and I$_C$ system is achieved using the Holtzman et al.\ (1995) zero points for the F450W and F814W filters. \section {Analysis} The color, magnitude, and sizes of the objects detected in NGC~3256 are plotted in Figure 3 \footnote{The positions, magnitudes, colors, and sizes of the objects found in NGC 3256 are also given in Table 1, available either in the electronic journal or from the first author.}. Several features of the cluster system of this galaxy merger are apparent in this diagram. One is the very large number of bright star clusters in this galaxy, approximately 1,000 with $M_{B} < -9$ inside of the central 7 kpc $\times$ 7 kpc. These objects account for approximately $19\%$ of the B light and $7\%$ of the I light within this region. The star clusters generally have blue colors. The bright magnitudes and blue colors are indicative of massive star clusters at young ages. For reference, Figure 4 shows the prediction of two stellar population models for the color and magnitude of a $2 \times 10^5$ \hbox{$\thinspace M_{\odot}$}\ globular cluster as a function of age. The clusters are also compact like globular clusters, with typical sizes of $\ {\raise-0.5ex\hbox{$\buildrel<\over\sim$}}\ 10$ pc. Only a few of the 1,000 objects are expected to be compact background galaxies or foreground stars, based on similar analyses of blank fields, such as the Hubble Deep Field. In order to address the true distribution of the population in color, size, and magnitude, simulations of artificial datasets are required to calibrate the detection procedure. For example, the absence of objects with very faint B magnitudes and blue B-I colors may be caused by the detection limit in the I band. Similarly, the absence of large, faint clusters may also be possibly due to selection effects. Therefore, we created a grid of artificial objects with a range of magnitudes in each bandpass, and a range of sizes for each magnitude. By creating a full grid of artificial stars in both bandpasses, we can address the issue of any effect of incompleteness in B or I on the color distribution at faint magnitudes. The difference between input and output magnitudes also provides a calibration of the effect of ``bin jumping'' when constructing luminosity functions. Similarly, by incorporating a range of sizes in the artificial star tests, we can address the question of the intrinsic size distribution of the cluster population. This study is the first in which all of these effects have been modeled. We can then test the consistency of various models of the luminosity, color, and sizes of the candidate globular clusters in NGC~3256 against the observations. Specifically, we create model data sets with various combinations of luminosity functions, color-magnitude relations, and luminosity-size relations. For the luminosity function and luminosity size-relation, we adopt a power-law form, while we use a linear relation between color and magnitude. The intrinsic widths of the color and size distributions are drawn from the data at bright magnitudes where they are unaffected by selection. Predictions for observables for each model are made by convolving the model with the selection functions derived from the simulations described above. We compare these predictions of various models to the luminosity function, color-magnitude relation, and luminosity-size relation observed for the candidate young globular clusters in NGC~3256 (Figure 3). A model is considered to fit the color-magnitude and luminosity-size relation if the linear regression of these parameters is statistically consistent with the data. To insure that the results are not dependent on the use of a linear fit, we also compare the median colors and sizes as a function of magnitude of the models to the observations, with the uncertainties in the medians of the data determined via bootstrapping. For the luminosity function, the goodness of the fit is determined using the double-root-residual test (Ashman, Bird, \& Zepf 1994). We also test the effects of changes in any one of the underlying distributions on all of the observed properties, as they are not decoupled from each other. For example, an underlying luminosity function can be flat, but if the clusters are smaller at faint magnitudes, they will be easier to detect, and the luminosity function will appear to rise at faint magnitudes. Similarly, a trend of color with magnitude can also give rise to an apparent luminosity function different than the underlying one if different color clusters are detected with different efficiency at faint magnitudes. We find that the best fitting model cluster population has little or no correlation between luminosity and color (B-I independent of $M_{B}$), a shallow correlation between luminosity and radius ($r \propto L^{0.07}$), a power-law luminosity function $N(L) \propto L^{-1.8}$. This best fitting model is shown in Figure 5. The statistical uncertainties on the parameters of the underlying cluster population are roughly 0.05 in the slope of the magnitude-color relation, and 0.1 in the exponent of both the radius-luminosity relation and the luminosity function. We discuss the magnitude-color relation, luminosity-size relation, and luminosity function individually in more detail below. \subsection{Color-Magnitude Relation} The broad color distribution and absence of a strong relationship between color and luminosity places strong constraints on the nature and evolution of the young cluster system in NGC~3256. In order to produce the observed range of colors, either differential reddening by dust or a range of ages is required. However, both reddening and age generally produce fainter magnitudes for redder objects (see Figure 4). This is not observed in NGC~3256, as shown by the similar color-magnitude diagrams of the data (Figure 3) and a simulated data set with no relationship between color and magnitude (Figure 5). A quantitative measure of the close agreement between the observations and a simulated data set with no relationship between color and magnitude is the similar median (B-I) colors as a function of B magnitude, which are given in Table 2. In contrast, a simulated data set with an intrinsic slope of (B-I) $\propto 0.5$B, like that expected for a standard reddening law, gives a much steeper relation between (B-I) color and B magnitude, as shown in Table 2. Similar results are obtained using other robust measures of the average color as a function of magnitude. The fundamental result is that we are unable to account for the broad range of observed colors solely by differential reddening or a broad age distribution because there is no intrinsic trend of redder colors with fainter magnitudes. The observations can be accounted for in two ways. One possibility is that the young cluster system in NGC~3256 has an age distribution up to several hundred Myr {\it and} low mass clusters are preferentially destroyed over this timescale. In this way, the typical luminosity of the older, redder cluster population will not become much fainter than the younger bluer, cluster population because the younger population will have more low-mass clusters. A modest amount of reddening may also be required to produce the colors of the reddest clusters, but not so much that a strong color-luminosity trend is produced. The lack of a strong color-luminosity relation can also be accounted for if most of the clusters are very young. At ages up to $\sim 10$ Myr, stellar population models predict a fairly broad range of colors with little change in B luminosity (see Figure 4). This effect is due to red supergiants, and is stronger in the Leitherer et al.\ (1998) models than the Bruzual \& Charlot (1998) models because of the increased presence of red supergiants in the former models. As in the previous case, some reddening may be required to produce the reddest clusters, but reddening can not be the primary determinant of the cluster colors, or a color-magnitude relation would be introduced, contrary to the observations. The critical aspect of the possibility that red supergiants at young ages account for much of the observed color spread is the requirement of a very young age for the system as a whole because all of the models begin to produce a significant color-luminosity trend after $\sim 10$ Myrs. Both the destruction and red supergiant hypotheses can account for a broad color distribution without a strong color-luminosity relation. An additional constraint is that the range of ages in the young cluster population would not be expected to be less than the dynamical time of the region in which they formed. Adopting a radius of 3 kpc for the region in which the young clusters are found, and a typical velocity of $v \approx 150$ $\,{\rm km\,s^{\scriptscriptstyle -1}}$, we find a dynamical time of 20 Myr. Thus the very young age hypothesis is marginally inconsistent with the requirement that the cluster population not form on a timescale shorter than the dynamical time. Other explanations for the absence of a strong color-luminosity trend are strongly constrained by the large observed color spread. For example, a model in which younger clusters are more heavily reddened than older clusters can give a weak color-luminosity trend. However, it does so at the cost of narrowing the color spread, and therefore fails to account for the observations. \subsection{Luminosity-Radius Relation} A second observation of relevance for models of the formation and evolution of young star clusters is the luminosity-radius relationship. The NGC~3256 young star cluster system has a very shallow relationship between radius and luminosity, roughly $r \propto L^{0.07}$. The relationship of radius with mass is likely to be similar to that with luminosity. This follows from the absence of a correlation between color and luminosity which suggests the mass-to-light ratio is mostly independent of luminosity. Thus, cluster mass is likely to be fairly independent of radius. A weak correlation between mass and radius has significant implications for the formation and evolution of globular clusters. Clouds in hydrostatic equilibrium follow the relationship $ r \propto M^{1/2} P^{-1/4}$ (e.g.\ Ashman \& Zepf 1999, Elmegreen 1989). The shallow observed correlation between radius and mass therefore suggests that higher mass clusters form at higher pressure. If confirmed, this will play a significant role in developing models of globular cluster formation. A shallow relationship between mass and radius also suggests that on average low-mass clusters are formed with lower density and are less bound than higher mass clusters. Therefore, they will be more susceptible to destruction by mass loss at early ages, and through tidal shocks over time. This result is an important input into determinations of the effect of dynamical evolution on the mass function of clusters, as discussed in the previous and the following subsections. In particular, studies of the dynamical evolution of globular cluster systems must adopt some relation between radius and mass for the initial cluster population (e.g.\ Ostriker \& Gnedin 1997 and references therein). Without data from young clusters, this relation has been based on observations of old cluster populations, whose properties may have already been altered by dynamical evolution. Thus, observations of the radii of young cluster systems are an important part of the study of dynamical evolution of globular cluster systems. These conclusions regarding the mass-radius relationship require confirmation, as the clusters are only marginally resolved in the present data. We are able to recover differences in the sizes of the objects, given a form for the radial profile of the cluster. The inferred mass-radius relationship for the cluster population should not be sensitive to the form of the profile adopted because it is only based on relative values of cluster radii. Therefore, our result is not likely to be sensitive to the specific cluster profile chosen, although it will be affected by any systematic changes in the shape of the profile with cluster mass. \subsection{Luminosity Function} The luminosity function of the NGC~3256 cluster system has a best-fitting power with slope about $-1.8$, with tentative evidence that it flattens at faint magnitudes. This slope is similar to that found in other young globular cluster systems in galaxy mergers (e.g.\ Whitmore \& Schweizer 1995, Schweizer et al.\ 1996, Miller et al.\ 1997, Carlson et al.\ 1997) and also to that for populous clusters in the LMC (Elmegreen \& Efremov 1997, Elson \& Fall 1985). The most notable difference between the observed luminosity function and a power-law convolved with observational selection is that the data appear to be flatter at the faint end than the model. In order to assess the statistical significance of this difference, we utilized a double-root-residual test (Ashman et al.\ 1994). The test indicates the difference is significant at about the $2.5\sigma$ level. However, given the uncertainties modeling the selection at these faint levels, any deviation from a power-law slope is tentative. Although the luminosity function of the NGC~3256 cluster system is now roughly a power-law, both the luminosity-color and luminosity-radius relation suggest that it is likely to evolve significantly over time. This is also expected on theoretical grounds (e.g.\ Gnedin \& Ostriker 1997, Murali \& Weinberg 1997a). A comparison of the observations to these theoretical expectations is presented below. \section {Discussion} HST imaging of the galaxy merger NGC~3256 has revealed a large population of objects with the bright luminosities, blue colors, and compact sizes expected of objects like Galactic globular clusters at young ages. In this section, we explore in more detail the relation of young clusters observed in starbursting and merging galaxies like NGC~3256 to old globular clusters, and to implications these observations have for models of the formation of globular clusters. \subsection{Cluster Mass Function and Dynamical Evolution} The power-law luminosity function is the one feature of the NGC~3256 young cluster system that is decidedly different from that of old globular cluster systems around galaxies like M87 and the Milky Way, which have lognormal luminosity functions. Although the masses and sizes inferred for many of the young clusters in NGC~3256 are similar to those of old globular clusters (e.g.\ Figures 3 and 5), the difference in the shape of the luminosity function has long been used to argue that the clusters in mergers and starbursts are fundamentally different than in the old systems (e.g.\ van den Bergh 1995). However, it has also long been realized that dynamical evolution can significantly alter the mass function of clusters systems over a Hubble time (e.g.\ Fall \& Rees 1977, Gnedin \& Ostriker 1997, Murali \& Weinberg 1997a, 1997b, Ashman \& Zepf 1998 and many references therein). We therefore consider whether observations of the mass and luminosity functions of cluster systems over a range of ages can be consistently accounted for by dynamical evolution. The two long-term dynamical processes that may be relevant for determining the shape of the globular cluster mass function are evaporation and tidal shocking. These have the following scalings between the lifetime of a cluster and its mass and radius $ t_{evap} \propto M^{1/2} ~ r^{3/2} $, and $ t_{sh} \propto M ~ r^{-3} $. The timescale for tidal shocks also depends on the cluster orbit and galaxy potential. As long as these are similar in the galaxies being compared, the scaling is independent of these parameters. Therefore, we concentrate on the scaling of the destruction timescale with cluster mass and radius. In $\S 3.2$, we found that cluster mass and radius may be only weakly dependent on one another with $r \propto L^{\sim 0.07}$. With this result, the equations above give a timescale for destruction that scales as roughly $M^{2/3}$ for both evaporation and tidal shocking. These give relative timescales for destruction of clusters as a function of their mass. An absolute timescale can be placed on these scalings through the known turnover mass and age of the globular cluster systems of galaxies like the Milky Way and M87. This allows us to determine the cluster mass scale that is undergoing destruction at the ages corresponding to young cluster systems observed in galaxy mergers. In this way, we find that at an age of 100 Myr, the characteristic mass scale set by dynamical processes is $10^{-3}$ of that for the globular cluster systems of the Milky Way and M87. At 500 Myr the corresponding number is about $10^{-2}$ of the characteristic mass scale of old globular cluster systems. This calculation is also based on the simplifying assumption that the dynamical processes are independent of one another, which is not strictly true. However, the numbers given above are not likely to be dramatically wrong as long as the observation that $M$ and $r$ are mostly independent holds. If long-term dynamical processes are responsible for evolution of the mass function, the above calculation suggests the ``turnover'' in the mass function in young globular cluster systems will occur at much smaller masses than in older globular cluster systems like that in the Galaxy. These small masses make such a turnover very difficult to observe directly in young systems. Specifically, the above calculation suggests that clusters with masses of roughly $1\%$ of the current turnover mass are being destroyed at the ages typical of young globular cluster systems, such as the NGC~3256 system studied here, as well as the NGC~1275 system (Carlson et al.\ 1998) and the NGC~7252 system (Miller et al.\ 1997). This corresponds to clusters five magnitudes below the current turnover of the globular cluster luminosity function. In order to calculate the detectability of a dynamically induced turnover in young cluster systems, the expected brightening of young clusters must also be accounted for. Stellar populations models (e.g.\ Bruzual \& Charlot 1998) predict that old objects of this magnitude were about 5.0 magnitudes brighter in B and 3.7 in V at the ages inferred for NGC~3256 from their colors. The corresponding brightening for the slightly older NGC~1275 and NGC~7252 cluster systems we will also study is about 4.2 magnitudes in B and 3.1 in V. The predicted brightening in these young cluster systems relative to $\sim 13$ Gyr old populations is dependent primarily on the mass function between several \hbox{$\thinspace M_{\odot}$}\ and slightly less than one \hbox{$\thinspace M_{\odot}$}. The adopted numbers are for a Salpeter (1955) slope of $x =1.35$ are somewhat less for flatter mass functions. The net result is that in young cluster systems observed in the B-band, such as NGC~3256, the observations must reach absolute magnitudes around the current turnover luminosity in old globular cluster systems, roughly $B = -6.8$. As can been seen in Figure 3, the observations fail by several magnitudes to reach such faint levels. In fact, it will be difficult to obtain reliable cluster counts at such magnitudes because of potential confusion with individual bright stars, which can have similar luminosities. Analyses of the NGC~1275 and NGC~7252 datasets give similar conclusions that the highest mass scale at which dynamical evolution is expected to be effective is well below the observational limit. In particular, for these $\sim 500$ Myr old systems, the observations would have to reach limits of about $B \ {\raise-0.5ex\hbox{$\buildrel>\over\sim$}}\ -6$ and $V \ {\raise-0.5ex\hbox{$\buildrel>\over\sim$}}\ -5.5$, while the observations are limited to about two magnitudes brighter. The luminosity functions of known cluster systems are therefore consistent with a model in which globular clusters form with a power-law mass function, and evolve through well-known dynamical processes to have the log-normal mass function observed in old systems such as the Milky Way and M87. Testing this hypothesis by directly looking for the turnover in young cluster systems is difficult because the mass scale is predicted to be extremely small. This work suggests that searches for turnovers in globular cluster mass functions might be more profitable in more intermediate-age systems in which the predicted mass scale may be accessible to observation. In order to improve the predictions for the dynamical evolution of young globular cluster systems, better constraints on the initial mass-radius relation are required. Studies of young cluster systems are critical for providing these constraints because these systems have not undergone as much dynamical evolution and will more accurately reflect the relationship between mass and radius at the time the young clusters formed. This relation plays a major role in the dynamical calculations. Specifically, the timescales for evaporation and tidal shocking depend on both mass and radius. If the relationship between mass and radius is stronger than it appears to be in our observations of the NGC~3256 system, the timescales for disruption of clusters of different masses will be affected. For example, if $r \propto M^{1/2}$, then $t_{evap}$ will have a very steep dependence on mass, and $t_{sh}$ will be shorter for larger masses. In this case, the mass scale for evaporation at 100 Myr would be only about a factor of 5 less than that after 10 Gyr, and only a factor of 3 less at 500 Myr compared to 10 Gyr. However, the tidal shocking would then be inversely correlated with mass, so both high and low mass clusters would be destroyed and the observational predictions in this case are less clear. Perhaps the safest conclusion then from our analysis of the formation and destruction of globular clusters is that information on the radii of the clusters as well as their mass function is critical for determining the effects of dynamical evolution and comparing of mass functions of young and old cluster systems. The cluster radii and mass-radius relation also have significant implications for models of the formation of young globular clusters. Further exploration of observational constraints on the initial mass-radius relation is a promising route for future studies. \subsection {Efficiency of Compact, Massive Cluster Formation} The census of compact, young star clusters in NGC~3256 allows the determination of the efficiency with which clusters form. The simplest characterization of efficiency is the fraction of new stars formed that are in dense star clusters. As described in $\S 3.1$, this fraction is about $20\%$ based on the percentage of blue light that is in dense star clusters. This value is similar to that observed in smaller starburst regions by Meurer et al.\ (1995). It is somewhat larger than the fraction of blue light in dense star clusters observed in the older NGC~7252 system, suggesting that either the formation efficiency peaks near the peak of the starburst or the dynamical evolution removes some of the clusters on the timescale of the age of the NGC~7252 system. A critical aspect of these NGC~3256 data is that they demonstrate that high formation efficiency can occur over a large area (7 kpc $\times$ 7 kpc) encompassing much of a $\sim L_{*}$ galaxy. These data also indicate that the fraction of stars that form in these dense star clusters is not negligible, and suggest that this is an interesting mode of star formation. A second way to characterize the efficiency of cluster formation is to compare it to the total amount of gas available. This is less straightforward than comparing the fraction of light, but connects more directly to current theoretical models. The total amount of molecular gas inferred from CO observations and standard assumptions regarding the conversion to H$_2$ mass is $1.5 \times 10^{10}$ \hbox{$\thinspace M_{\odot}$}\ (Casoli et al.\ 1991, Aalto et al.\ 1991, Mirabel et al.\ 1990). The total mass in the young cluster system can be estimated from the color of each object, and a stellar populations model that allows color to be converted to age and mass-to-light ratio. The observed luminosity can then be converted into mass. Applying this procedure individually to each cluster in the NGC~3256 cluster sample and summing up gives a total mass in the young cluster system of $6 \times 10^{7}$ \hbox{$\thinspace M_{\odot}$}. This is based on taking all objects with $(B-I) < 1.5$ and inferred sizes less than 15 pc, and using the Charlot-Bruzual stellar population models with a Miller-Scalo initial mass function. A Salpeter initial mass function would increase the mass estimate by approximately $50\%$. Reddening of observed objects is less of a factor because its effects on the mass estimate cancel out to first order. Specifically, reddening makes the objects fainter, decreasing the apparent luminosity, but also redder, increasing the inferred mass-to-light ratio, so in the end the mass estimate is similar. For example, the mass estimate for the NGC~3256 cluster system is only about $30\%$ greater for an internal reddening of $A_B = 1.0$ compared to $A_B=0.0$. An additional effect is that some clusters may be extincted beyond detection. The resulting efficiency for the formation of massive, compact clusters computed in this way is then about $0.5\%$. This is a lower limit in the sense that cluster formation is ongoing and there is plenty of gas mass left from which to form more clusters. If clusters continue to make up $\sim 20\%$ of the stars formed, the total cluster mass fraction at the end of the starburst will be closer to this value. Therefore, the observations place the formation efficiency between $0.5-20\%$ depending on how exactly efficiency is defined and what happens in the future of the starburst in NGC~3256. It is interesting to compare our observed efficiency of globular cluster formation in NGC~3256 to the fraction of mass in old stellar populations that is in globular clusters. The highest observed fraction of mass in globular clusters to total stellar mass is about $1\%$, as seen in the Galactic halo and the richest extragalactic globular cluster systems such as M87 (e.g.\ Ashman \& Zepf 1998). Typical elliptical galaxies have ratios about five times lower, around $0.2\%$. Therefore, the globular cluster formation efficiency observed in NGC~3256 is more than sufficient to account for the globular cluster systems of elliptical galaxies. More specifically, if the mass fraction of stars that form in globular clusters is closer to $20\%$ over the full starburst/merger event, then mergers have an over-efficiency problem. This over-efficiency problem can be alleviated in several ways. One possibility is that the progenitor spirals are gas-poor, which may be true at the current epoch but seems unlikely at high redshift when most mergers are likely to occur. A more likely explanation is significant dynamical destruction of the cluster population. This destruction is predicted by theory and required by observation if the power-law luminosity functions of young cluster systems are to match the lognormal luminosity function of old globular cluster systems. Conversely, if the overall mass fraction of stars that form in globular clusters is closer to $0.5\%$, then gas-rich mergers will make about the right number of globular clusters for moderately rich systems, and gas-poor mergers will lead to moderately poor globular cluster systems. The reality is likely to be somewhere between these two extremes. It is unlikely that all globular cluster formation in NGC~3256 will immediately cease while the large reservoir of cold gas continues to form stars, as required for the overall efficiency to be $0.5\%$. It is also unlikely that all of the remaining gas will form stars in a starburst in which $20\%$ of the new stars are formed in dense, massive clusters, as is currently happening in NGC~3256. Moreover, many clusters are expected to be lost due to dynamical processes. The high efficiency of globular cluster formation observed in NGC~3256 also has significant implications for theoretical models of globular cluster formation. In particular, scenarios in which globular clusters form as cores within proposed super giant molecular clouds (SGMCs) predict that only about $\sim 0.2\%$ of the mass of the cloud forms stars in dense cores, as seen in Galactic giant molecular clouds (McLaughlin \& Pudritz 1996, Harris \& Pudritz 1994). Thus the observed globular cluster formation efficiency appears to pose a problem for the SGMC scenario. A second problem is the long timescale for the formation of SGMCs in these models compared to the apparently rapid formation of globular clusters in galaxy mergers and starbursts. The observations appear to be more consistent with models of globular cluster formation in which the clusters form from highly compressed giant molecular clouds of typical mass for spiral galaxies. These may either originate from the progenitor spirals or be newly made GMCs. \subsection{Conclusions} The primary conclusions of our study of HST images in B and I of the galaxy merger NGC~3256 are: \noindent 1. NGC~3256 has a very large population of compact, bright, blue objects. Many of these clusters have estimated sizes and masses like those of Galactic globular clusters. On this basis, we identify some fraction of these objects as young globular clusters. \noindent 2. The young cluster system has a broad range of colors, but little or no correlation between color and luminosity. This observation requires either destruction of low mass clusters over time or a very young age ($\ {\raise-0.5ex\hbox{$\buildrel<\over\sim$}}\ $ 20 Myr) for the young cluster system. \noindent 3. Dynamical evolution is likely to significantly affect the mass function of the young cluster system. If the system is not very young, this has already been observed. The mass-radius relation for the cluster population is an important constraint on the predictions for dynamical evolution of the cluster population. \noindent 4. The large number of candidate young globular clusters indicates a high efficiency of cluster formation. This is observed across the 7 kpc $\times$ 7 kpc region studied. The efficiency of cluster formation in the galaxy merger NGC~3256 is more than sufficient to account for the metal-rich globular cluster populations in elliptical galaxies if these form from gas-rich mergers. \acknowledgments We thank Richard Larson for useful discussions, Dave Carter for collaborating in obtaining the AAT image of NGC~3256, and Eddie Bergeron for making the color images frmo the HST data. We thank an anonymous referee for helpful suggestions. S.E.Z. and K.M.A. acknowledge support for this project from NASA through grants GO-05396-94A and AR-07542-96A awarded by the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., for NASA under contract NAS 5-26555. SEZ also acknowledges the support of a Hubble Fellowship and fruitful discussions with colleagues at UC Berkeley during the early stages of this work.
\section{Introduction} \label{sec:intro} It has been an extremely fruitful idea to study a conformal field theory by putting it on various surfaces, with or without boundaries. Apart from the sphere, that has been considered first, prime examples of non--trivial geometries include the torus \cite{cardy1} and the cylinder \cite{cardy2,cardy3}. They serve to probe different facets of a given conformal theory. However the data specific of these surfaces are inextricably related to each other, and this fact provides very stringent constraints on the theory itself, allowing for example to determine its field content. For minimal conformal theories, the problem on the torus for single--valued fields has been resolved in \cite{ciz}: consistent models have a periodic partition function that can be associated in a unique way with a pair $(A,G)$ of simple Lie algebras of ADE type. The solution of the analogous problem for the cylinder is much more recent, even if early calculations in either specific models or with specific boundary conditions have been carried out in \cite{cardy2,cardy3,sb}. The recent discovery in \cite{aos} of a new conformally invariant boundary condition in the 3--state Potts model triggered a renewal of interest in the problem. For minimal models, its solution was given in \cite{bpz,bppz}, and shown to be encoded in the same Dynkin graphs that specify the torus partition function. When a model has a symmetry, necessarily discrete in this context, fields can be multiple--valued on the torus, so that non--periodic sectors exist. Furthermore, the fields transform under the symmetry group, and, upon diagonalization, can be assigned charges. All this information is encoded in frustrated partition\break $$\mbox{}$$ \vspace{2.3cm} \mbox{} \noindent functions, covariant under the modular group of the torus, a fact that can be used to, first, detect the presence of a symmetry, and then to compute the various partition functions \cite{zuber,rv}. In this article, we address the question of the action of the symmetry group on the cylinder partition functions for the minimal models. We show how the symmetry group acts on the boundary conditions, and identify the invariant (or symmetric) ones. We then study the charge assignments of the fields that occur in the presence of those boundary conditions. Section II is a reminder about the minimal conformal models on a torus and on a cylinder. In Section III, we discuss the action of the symmetry group on the conformally invariant boundary conditions, which is then used in Section IV to compute frustrated partition functions on a cylinder, or equivalently the charge assignment of the boundary fields. Section V contains explicit formulae and computational details of a particular assignment. Its unicity (in fact non--unicity) is examined in Section VI, from which we conclude that, in general, a large number of distinct charge assignments are consistent. We also derive selection rules for the boundary fusion coefficients. We finish, in Section VII, with an analysis of the unitary models for which we propose an unambiguous charge assignment. Section VII contains the most interesting corollary of the previous sections. An analysis based on the expected consequences of the Perron--Frobenius theorem fixes a unique charge assignment in the unitary $(A,A)$ models, which we conjecture to be the correct one. This is in sharp contrast with the models of the $(A,D)$ and $(A,E_6)$ series. For those, there is no consistent charge assignment that is compatible with the Perron--Frobenius theorem, the reason being that there is no way to ensure an invariant ground state in all sectors. Motivated by the results obtained for the Potts model \cite{aos}, we will interpret this phenomenon as the non--existence of positive classical Boltzmann weights for some invariant boundary conditions. A simple characterization of them suggests itself in terms of their Dynkin graph labels. \section{Minimal models} \label{sec:rem} Minimal models are classified by a pair $(A,G)$ of simply--laced simple Lie algebras with coprime Coxeter numbers, $p$ and $q$. One may assume that $p$ is odd. Their periodic partition function on a torus of modulus $\tau$ is a sesquilinear form in the Virasoro characters \begin{equation} Z(A,G) = \sum_{i,j} \; M_{ij}\,\chi^*_i(\tau)\,\chi^{}_j(\tau), \qquad M_{ij} \in {\Bbb N}, \end{equation} where $i,j$ are labels for Virasoro highest weight representations. They lie in the Kac table $\{(r,s)\;:\;1 \leq r \leq p-1,\, 1 \leq s \leq q-1\}$, in which $(r,s)$ and $(p-r,q-s)$ must be identified. The connection with the Lie algebras is best brought out by writing the diagonal elements $M_{ii}$ as \cite{ciz} \begin{equation} Z(A,G) = {\textstyle {1 \over 2}} \sum_{r \in \text{Exp}\,A \atop s \in \text{Exp}\,G}\; |\chi_{r,s}|^2 + \text{off-diagonal}, \label{zpp} \end{equation} where $r$ and $s$ run over the exponents of $A$ and $G$. The full expressions of the partition functions are given in \cite{ciz}. The question of the symmetry group has been first addressed in \cite{z}, and solved in {rv} for the unitary models $|p-q|=1$. The analysis can however be easily extended to the non--unitary minimal models, with the following result. With the exception of the models $(A_{p-1},A_{q-1})$ with $p$ {\it and} $q$ odd, which have no symmetry at all, the other models $(A,G)$ have a finite symmetry group $\Gamma$, which is the group of automorphisms of the Dynkin graph of $G$, that is, $\Gamma(G)=Z_2$ except $\Gamma(D_4)=S_3$ and $\Gamma(E_7,E_8)=\{e\}$. When a model has a symmetry group, the fields may have a non--trivial monodromy along the two periods of the torus, transforming as $\phi(z+1) = {}^{g}\phi(z)$ and $\phi(z+\tau) = {}^{g'}\!\phi(z)$ for two commuting\footnote{That forces us to focus on Abelian subgroups of $\Gamma$. Thus in this paper we consider $Z_2$ and $Z_3$ (sub)groups only.} elements $g,g' \in \Gamma$. In the Hamiltonian formalism, this amounts to give a Hilbert space ${\cal H}_g$ of states with a $g$--monodromy along the first period, which are then acted on by $g'$ when transported along the second period. The latter action can be diagonalized, $\,^{g'}|\phi\rangle = e^{2i\pi Q/N} |\phi\rangle$, defining the charge $Q$ of the field $\phi$ under the action of $g'$, an element of order $N$. The field content of ${\cal H}_g$ as well as their charges can be read off from the frustrated partition functions $Z_{g,g'}(A,G)$. These are still sesquilinear forms but with coefficients in ${\Bbb Z}(e^{2i\pi/|\Gamma|})$: \begin{equation} Z_{g,g'} = {\rm Tr}_{{\cal H}_g} \;\Big[q^{L_0-c/24}\,\bar q^{\bar L_0-c/24}\,g'\Big]. \end{equation} Because a modular transformation mixes the two periods, it must be accompanied by a corresponding change of monodromies so that the net effect vanish (for a fixed pair $(A,G)$): \begin{equation} Z_{g,g'}(\tau) = Z_{g^ag'^c,g^bg'^d}({\textstyle{a\tau +b \over c\tau +d}}). \label{tor} \end{equation} All such functions are given explicitly in \cite{rv} (with a straightforward extension to the non--unitary case). The function (\ref{zpp}) corresponds to $g=g'=e$. On a cylinder, say of length $L$ and perimeter $T$, only one Virasoro algebra remains, so that the partition function is linear rather than sesquilinear in the characters \cite{cardy2}. Conformally invariant boundary conditions $\alpha,\beta$ must be prescribed on the two boundaries, and a monodromy condition $g$ must be imposed along the periodic coordinate, $\phi(z+T)={}^g\phi(z)$. We first consider a trivial monodromy, $g=e$. If the time variable is defined to run along the periodic direction, the partition function is the trace of the transfer matrix $e^{-TH_{\alpha,\beta}}$, \begin{equation} Z^e_{\alpha,\beta}(\tau) = \sum_i \; n^i_{\alpha,\beta} \, \chi_i(\tau), \qquad \tau=iT/2L. \label{form1} \end{equation} The integer $n^i_{\alpha,\beta}$ gives the multiplicity of the primary field with Kac label $i$ in the Hilbert space ${\cal H}_{\alpha,\beta}$. Alternatively, one may view the time evolution as going from one boundary to the other. In this case, the states on constant time surfaces belong to the bulk periodic Hilbert space ${\cal H}_e$, and are propagated in time from one boundary state $|\alpha\rangle$ to the other $|\beta\rangle$ (formally also in ${\cal H}_e$). The partition function is then \begin{equation} Z^e_{\alpha,\beta}(\tau) = \langle \beta| e^{-LH_e} |\alpha\rangle, \label{preform2} \end{equation} with $H_e$ the Hamiltonian corresponding to periodic bulk sector. The boundary states are conformally invariant, satisfying $(L_n-\bar L_{-n})|\alpha\rangle$ for all $n \in \Bbb Z$ \cite{cardy3}. The solutions to this equation are the Ishibashi states \cite{i}: every highest weight representation $[i \otimes \bar i]$ contains exactly one such state, which we denote by $|i\rangle\rangle$, while the other representations $[i \otimes \bar j]$, for $i \neq j$, do not contain any. In the present situation, the Ishibashi states must be taken from the space ${\cal H}_e$, and hence are labelled by ${\cal E}_e = \{i \;:\; [i \otimes \bar i] \in {\cal H}_e\}$. Expanding the boundary states in the basis of Ishibashi states, $|\alpha\rangle = \sum_i \; c^i_\alpha \, |i\rangle\rangle$, makes the partition function (\ref{preform2}) take the form \begin{equation} Z_{\alpha,\beta}(\tau) = \sum_{i \in {\cal E}_e} \; c^i_\alpha \, \bar c^{i}_\beta \, \chi_i({\textstyle{-1 \over \tau}}). \label{form2} \end{equation} The arguments of the characters in (\ref{form1}) and (\ref{form2}) are related by the modular transformation $\tau \mapsto {-1 \over \tau}$, under which the characters transform linearly through a unitary matrix $S$. Comparing the two formulae then yields Cardy's equation \cite{cardy3} \begin{equation} n^i_{\alpha,\beta} = \sum_{j \in {\cal E}_e} \; S_{i,j} \, c^j_\alpha \, \bar c^{j}_\beta. \label{cyl} \end{equation} The relations (\ref{cyl}) are overdetermined for the vectors $c^j$, and provide a means to classify the boundary conditions $|\alpha\rangle$, to compute the spectra of ${\cal H}_{\alpha,\beta}$, and in turn the surface scaling dimensions. Such calculations have been carried out in \cite{cardy2,sb,aos}, but the general answer appeared only very recently in \cite{bpz,bppz}. Let 1 be the label corresponding to the vacuum representation, namely to $(r,s)=(1,1)=(p-1,q-1)$. In \cite{bppz}, it was observed that, upon setting $c^i_\alpha = \psi^i_\alpha / \sqrt{S_{1,i}}$ for a set of complete and orthonornal vectors $\psi^i$, Cardy's equation appears as an explicit diagonalization \begin{equation} n^i_{\alpha,\beta} = \sum_{j \in {\cal E}_e} \; \psi^j_\alpha \, {S_{i,j} \over S_{1,j}} \, \bar\psi_\beta^{j}. \label{nie} \end{equation} The matrices $n^i$ have eigenvalues ${S_{i,j}/S_{1,j}}$, and a common eigenbasis is given by the vectors $\psi^j$. As a result, they satisfy the fusion rules \begin{equation} n^i \, n^j = \sum_k \; N_{ij}^k \, n^k. \end{equation} Reversing the argument, the authors of \cite{bppz} conclude that an $\Bbb N$--valued representation of the fusion algebra of dimension $|{\cal E}_e|$ provides a solution to Cardy's equation with $|{\cal E}_e|$ different boundary conditions. When $c^i_\alpha = \psi^i_\alpha / \sqrt{S_{1,i}}$ is an invertible matrix, this solution yields the maximal set of conformally invariant boundary conditions. Note that the boundary states $|\alpha\rangle$ are determined up to a phase, but the fact that the entries of $n^i$ are to be positive integers leaves only a global, unobservable, phase. For minimal models, this was all made explicit in \cite{bpz}. For the model $(A,G)$, it was shown that each node in the product Dynkin diagram $A \times G$, quotiented by an appropriate $Z_2$ automorphism, defines a boundary condition and vice--versa. Indeed, from (\ref{zpp}), the number of Ishibashi states in the periodic sector is $|{\cal E}_e| = {\textstyle {1 \over 2}} |\text{Exp} A \times \text{Exp} G|$, so that only half the nodes can define distinct boundary conditions. We will use the variables $\alpha,\beta$ and $(a_i,b_i)$ as labels for the nodes of $A \times G$. The letters $A$ and $G$ will denote at the same time the Lie algebras, the Dynkin diagrams or the corresponding adjacency matrices. As a result of the quotient of the product graph, the matrices $n^i$, for $i=(r,s)$, are given by \cite{bpz} \begin{eqnarray} n^i_{(a_1,b_1),(a_2,b_2)} &=& (\hat N_r)_{a_1,a_2}\,(V_s)_{b_1,b_2} + (\hat N_r)_{a_1,a^*_2}\,(V_s)_{b_1,b^*_2} \nonumber \\ &=& n^i_{(a^*_1,b^*_1),(a_2,b_2)} = n^i_{(a_1,b_1),(a^*_2,b^*_2)}. \label{ni} \end{eqnarray} In this formula, the $\hat N$'s and the $V$'s are the fused adjacency matrices of $A$ and $G$ respectively. They are defined recursively by $X_m = X_2X_{m-1} - X_{m-2}$, with $X_1={\bf 1}$ and $X_2=A$ if $X=\hat N$, and $X_2=G$ if $X=V$. Equivalently, \begin{equation} \hat N_r = U_{r-1}(A), \qquad V_s = U_{s-1}(G), \end{equation} where $U_m(2\cos{x}) = \sin{(m+1)x}/\sin{x}$ is the $m$--th Tchebychev polynomial of the second kind. The automorphism $(a,b) \mapsto (a^*,b^*)$ can be determined from the condition $n^{(r,s)}=n^{(p-r,q-s)}$ (necessary if the $n^i$ are to satisfy the fusion algebra). It yields $a^*$ and $b^*$ to be given\footnote{The automorphism $^*$ in $G$ thus coincides with the charge conjugation in the corresponding affine algebra $\hat G$.} by the non--trivial automorphism of $A$ and $G$, for $G \neq D_{\text{even}},E_7,E_8$, and $b^*=b$ for $G=D_{\text{even}},E_7,E_8$. Viewing the tensor products $F^i(A,G) = \hat N_r \otimes V_s$ as the fused adjacency matrices of $A \times G$, the above result may be summarized by saying that $n^i$ is a folded fused adjacency matrix of $A \times G$ \begin{equation} n^i_{\alpha,\beta} = F^i_{\alpha,\beta}(A,G) + F^i_{\alpha,\beta^*}(A,G). \label{nif} \end{equation} The eigendata for the matrices $A$ and $G$ make sure that the matrices in (\ref{ni}) satisfy the minimal model fusion algebra. For the $(A,A)$ models, the $a_i$ (resp. $b_i$) labels run over the same set as $r$ (resp. $s$), and the matrices $n^i$ are the fusion matrices $N^i$ themselves \cite{cardy3}. \section{Symmetric boundary conditions} \label{sec:sym} We now proceed to the analysis of the cylinder partition functions when there is a group of symmetry $\Gamma$. From now on, we thus take $q$ even, and $G \neq E_7,E_8$. The boundary states are combinations of periodic Ishibashi states, on which the action of $\Gamma$ is known from the torus partition functions $Z_{e,g}$. This induces an action on the boundary states which one can determine. That action must be by permutations. For the minimal models, a boundary state corresponds to a pair of nodes of $A$ and $G$, \begin{equation} |(a,b)\rangle = \sum_{i \in {\cal E}_e} \; {1 \over \sqrt{S_{1,i}}} \,\psi^i(a,b)\, |i\rangle\rangle. \label{exp} \end{equation} where the $\psi^i$ form an eigenbasis for the concrete matrices in (\ref{ni}). Let us denote by $\sigma$ the automorphisms of the Dynkin graph of $G$, so that every $\sigma$ has fixed points. (The automorphism of the $A$ factor has a free action, and is used to obtain a set of representatives under the $^*$ involution, see (\ref{ni}).) Each $\sigma$ has a diagonalizable action on the eigenvectors $\psi^i$. The action of $g \in \Gamma$ on a periodic Ishibashi state can be read off from the diagonal terms in the frustrated partition function $Z_{e,g}(A,G)$ \cite{rv}. These can be compactly presented as follows. If $g$ has order $N$, and if one writes the diagonal terms in $Z_{e,g}$ as \begin{equation} Z_{e,g} = \sum_{i \in {\cal E}_e} \; \zeta_{N}^{Q_{g}(i)} \: |\chi_i|^2 + \ldots, \end{equation} then, for a proper choice of the $\psi^i$, the phase is seen to be exactly equal to the eigenvalue of $\psi^i$ under an order $N$ automorphism $\sigma$: \begin{equation} \psi^i(a,\sigma(b)) = \zeta_{N}^{Q_{g}(i)} \, \psi^i(a,b). \label{auto} \end{equation} The $\sigma$ that is induced by $g$ through the previous formula is unambiguous in the models $(A,G)$ if $G$ is not $D_4$: the only non--trivial $g$ induces the only non--trivial $\sigma$. When the $D_4$ algebra is involved, exactly which $\sigma$ in $S_3$ arises from a set of charges $Q_g$ (univoquely given by $Z_{e,g}$) depends on the eigenbasis we choose. In particular, a same set of $Z_2$ charges can lead to the three different (but conjugate) order two $\sigma$'s. It quickly follows from (\ref{exp}) and (\ref{auto}) that an order $N$ group element $g$ acts on the boundary states as an order $N$ automorphism $\sigma$: \begin{equation} |(a,b)\rangle \longrightarrow |^g(a,b)\rangle = |(a,\sigma(b))\rangle. \end{equation} Therefore, for any subgroup $\gamma$ of $\Gamma$, the $\gamma$--symmetric boundary conditions correspond to the nodes of $A \times G$ that are fixed by a group $\gamma$ of automorphisms of $G$. This set of nodes form a graph which we call the fixed point graph and denote by $A \times G^\gamma$. In particular, the boundary conditions that are invariant under a group element $g$ correspond to the nodes in $A \times G^\sigma$, with $G^\sigma$ the part of $G$ that is fixed by the automorphism $\sigma$ induced by $g$. As before the pairs of nodes which are related by the $^*$ automorphism define the same invariant boundary conditions. In the minimal models, the fixed point diagrams that arise for the various choices of $g$ are \begin{eqnarray} (A_{p-1},A_{q-1})\;:\quad && T_{(p-1)/2} \times A_1,\nonumber \\ (A_{p-1},D_{q/2+1})\;:\quad && T_{(p-1)/2} \times A_{q/2-1}, \quad (g^2=e), \nonumber \\ (A_{p-1},D_4)\;:\quad && T_{(p-1)/2} \times A_1, \quad (g^3=e),\nonumber \\ (A_{p-1},E_6)\;:\quad && T_{(p-1)/2} \times A_2, \label{fixdyn} \end{eqnarray} where $T_{(p-1)/2}$ denotes the tadpole diagram obtained by quotienting $A_{p-1}$ by its automorphism $^*$. For instance, the fixed point graph of an element $g$ of order two in the $(A_{p-1},D_{q/2+1})$ model is graphically given by \hskip 1truecm \begin{picture}(180,35)(0,-10) \put(-4,0){\circle*{4}} \put(-10,-10){\makebox(0,0)[b]{\small $a\!=\!1$}} \put(-2,0){\line(1,0){15}} \put(15,0){\circle*{4}} \put(15,-10){\makebox(0,0)[b]{\small 2}} \put(17,0){\line(1,0){10}} \put(30,0){...} \put(40,0){\line(1,0){10}} \put(52,0){\circle*{4}} \put(52,-10){\makebox(0,0)[b]{\small $p-1 \over 2$}} \put(52,7){\circle{14}} \put(83,0){\makebox(0,0)[c]{$\times$}} \put(110,0){\circle*{4}} \put(104,-10){\makebox(0,0)[b]{\small $b\!=\!1$}} \put(112,0){\line(1,0){15}} \put(129,0){\circle*{4}} \put(129,-10){\makebox(0,0)[b]{\small 2}} \put(131,0){\line(1,0){10}} \put(144,0){...} \put(154,0){\line(1,0){10}} \put(166,0){\circle*{4}} \put(166,-10){\makebox(0,0)[b]{\small ${q \over 2}-1$}} \end{picture} \vskip 0.4truecm \section{Cylinder partition functions} \label{sec:cpf} The consequences of a symmetry can now be pursued at the level of the partition functions. Let us suppose that $\alpha$ and $\beta$ are two boundary conditions that are invariant under a group element $g$, of order $N$. It implies that the transfer matrix $e^{-H_{\alpha,\beta}}$ and $g$ commute, and can be diagonalized in the same basis. The effect, on the cylinder partition function, of the insertion of $g$ on a line connecting the two boundaries is to affect each Virasoro tower with a $N$--th root of unity, so that the first form (\ref{form1}) becomes \begin{equation} Z^g_{\alpha,\beta}(\tau) = \sum_i \; n^{(g)\,i}_{\alpha,\beta} \, \chi_i(\tau). \end{equation} This shows that $n^{(g)\,i}$ must be related in the following way to the restriction of $n^i$ to the $g$--symmetric boundary conditions: an entry of $n^i$ equal to $n$ becomes in $n^{(g)\,i}$ a sum of $n$ $N$--th roots of unity. In the second form, the boundary state $|\alpha\rangle$ is propagated to $|\beta\rangle$ by the Hamiltonian that acts on the bulk sector twisted by $g$, so that \begin{equation} Z^g_{\alpha,\beta}(\tau) = \langle \beta| e^{-LH_g} |\alpha\rangle. \end{equation} This formula makes it clear that the states $|\alpha\rangle$ and $|\beta\rangle$ should have a projection in the twisted Hilbert space ${\cal H}_g$, and being conformally invariant, must have expansions in Ishibashi states of the bulk $g$--sector, themselves labelled by ${\cal E}_g = \{i \;:\; [i \otimes \bar i] \in {\cal H}_g\}$. Setting $|\alpha\rangle = \sum_i \; c^{(g)\,i}_\alpha \, |i\rangle\rangle_g$, one obtains a Cardy equation \begin{equation} n^{(g)\,i}_{\alpha,\beta} = \sum_{j \in {\cal E}_g} \; S_{i,j} \, c^{(g)\,j}_\alpha \, \bar c^{(g)\,j}_\beta, \label{cylt} \end{equation} for all boundary conditions which are $g$--symmetric. By inspecting the torus partition functions $Z_{g,e}(A,G)$ \cite{rv} (see also the next section), one readily sees that the matrices $c^{(g)\,i}_\alpha$ are square, namely \begin{equation} |{\cal E}_g| = {\textstyle {1 \over 2}} \, |A \times G^\sigma| = |T \times G^\sigma|, \end{equation} where the factor ${\textstyle {1 \over 2}}$ accounts for the identification under $^*$. Let us also note that, since the $g$--Ishibashi states form a basis for boundary states that are invariant under $g$, they should themselves be all neutral for consistency. This is again easily checked from $Z_{g,g}$. The rest of this article is devoted to a discussion of the solutions to the Cardy equation (\ref{cylt}). We will suggest that the proper physical solution is a natural generalization to $g \neq e$ of the two formulae (\ref{nie}) and (\ref{nif}) for $n^i$. Our first statement is that a particular solution, compatible with $n^i \equiv n^{(e)\,i}$, is provided, modulo a sign $\delta_i$, by the folded fused adjacency matrices of the graph $A \times G^\sigma$: \begin{equation} \tilde n^{(g)\,i}_{\alpha,\beta} = \delta_i \, \big[F^i_{\alpha,\beta} (A,G^\sigma) + F^i_{\alpha,\beta^*}(A,G^\sigma)\big], \quad \delta_i=\pm 1. \label{fff} \end{equation} Here $\alpha=(a_1,b_1)$ and $\beta=(a_2,b_2)$ are pairs of nodes in $A \times G^\sigma$ (with the usual identification under $^*$), and the automorphism $^*$ is the same as before. When $g,\sigma \neq e$, this formula can be simplified because every $b_2$ in $G^\sigma$ is a fixed point of $^*$. Indeed since $\beta$ is a node of $A \times G^\sigma$, $b_2$ is a fixed point of $\sigma$. But $\sigma$ and $^*$ coincide, except for $G=D_{\text{even}}$ for which $^*$ is trivial. Thus the folding by $^*$ acts on $a_2$ only, resulting in an effective folding of the $A$ factor onto a $T$ graph (hence the graphs (\ref{fixdyn})). One also checks that the folded fused adjacency matrices of $A_{p-1}$ are the fused adjacency matrices of $T_{(p-1)/2}$. Thus the matrices in (\ref{fff}) are simply proportional to the fused adjacency matrices of the fixed point diagram \begin{equation} \tilde n^{(g)\,i}_{\alpha,\beta} = \delta_i \, F^i_{\alpha,\beta}(T,G^\sigma) = \delta_i \, U_{r-1}(T)_{a_1,a_2} U_{s-1}(G^\sigma)_{b_1,b_2}. \label{nigf} \end{equation} The matrices $F^i(T,G^\sigma)$ fall short of satisfying the minimal fusion algebra, but the factors $\delta_i$ can be adjusted so that the $\tilde n^{(g)\,i}$ do satisfy it. The fusion algebra of the minimal model ${\cal M}(p,q)$ is polynomially generated by two generators $X$ and $Y$, which one can associate with the representatives of $N^{(2,1)}$ and $N^{(1,2)}$ \cite{dms}. The other elements of the algebra are explicitly given by Tchebychev polynomials \begin{equation} N^i = U_{r-1}(X) \, U_{s-1}(Y), \label{pform} \end{equation} and the generators must satisfy three relations: \begin{equation} U_{p-1}(X) = U_{q-1}(Y) = U_{p-2}(X) - U_{q-2}(Y) = 0. \label{rel} \end{equation} The matrices $F^i(T,G^\sigma)$ have the proper form (\ref{pform}), and $T_{(p-1)/2}$ and $G^\sigma$ do indeed satisfy the first two relations in (\ref{rel}). This is most easily seen by verifying that all eigenvalues satisfy the relevant equation. For instance, the eigenvalues $\lambda_m$ of $T_{(p-1)/2}$ are in \begin{equation} \text{spec}(T_{{p-1\over 2}}) = \{2\cos{\textstyle{\pi m \over p}}\;:\; 1 \leq m\,\text{odd} \leq p-1\}, \label{spectrum} \end{equation} and clearly satisfy $U_{p-1}(\lambda_m)=0$. In the same way, one computes that \begin{equation} U_{p-2}(T_{{p-1\over 2}}) = {\bf 1}. \end{equation} The corresponding calculation for $G^\sigma$ yields\footnote{The adjacency matrix of $A_1$ is the number zero, so that its fused adjacency matrices are $U_{s-1}(0) = (-1)^{(s-1)/2}$ for $s$ odd, and 0 for $s$ even.}, in the same four cases as in (\ref{fixdyn}), \begin{eqnarray} G^\sigma=A_1\;: && \qquad U_{q-2}(G^\sigma) = (-1)^{{q \over 2}+1}\,{\bf 1}, \nonumber\\ G^\sigma=A_{{q \over 2}-1}\;: && \qquad U_{q-2}(G^\sigma) = -{\bf 1}, \nonumber\\ G^\sigma=A_1\;: && \qquad U_{q-2}(G^\sigma) = {\bf 1},\nonumber\\ G^\sigma=A_2\;: && \qquad U_{10}(G^\sigma) = -{\bf 1}, \label{ug} \end{eqnarray} where the last line refers to the models $(A_{p-1},E_6)$ for which $q=12$. Thus, except when $G^\sigma=A_1$ and when $q=2 \bmod 4$, the last condition in (\ref{rel}) is not fulfilled. Owing to the parity properties of the Tchebychev polynomials, $U_m(-x) = (-1)^m U_m(x)$, one easily sees that $X=(-1)^{{q \over 2}+1}\,T_{(p-1)/2}$ in the first and third cases of (\ref{ug}), and $X=-T_{(p-1)/2}$ in the second and fourth ones, together with $Y=G^\sigma$, do satisfy all three conditions and therefore generate the correct algebra. Correspondingly, one finds that the matrices $\tilde n^{(g)\,i} = F^i(X,Y) = \delta_i F^i(T,G^\sigma)$ with the following signs, \begin{eqnarray} (A_{p-1},A_{q-1}) \;: && \quad \delta_i = (-1)^{(r+1)({q \over 2}+1)},\nonumber\\ (A_{p-1},D_{{q \over 2}+1})\;: && \quad \delta_i = (-1)^{r+1}, \qquad (g^2=e),\nonumber\\ (A_{p-1},D_4)\;: && \quad \delta_i = 1, \qquad (g^3=e), \nonumber\\ (A_{p-1},E_6)\;: && \quad \delta_i = (-1)^{r+1}. \label{deltas} \end{eqnarray} obey the minimal fusion algebra. Because of the signs $\delta_i$ but also because the matrices $F^i(T,G^\sigma)$ are not positive for $\sigma \neq e$ (they are however of constant sign), the $\tilde n^{(g)\,i}$ provide $\Bbb Z$--representations\footnote{In case of a $Z_3$ symmetry group, one might expect ${\Bbb Z}(e^{2i\pi/3})$--valued representations. This is however excluded by the symmetry $Z^g_{\alpha,\beta} = Z^g_{\beta,\alpha}$ (time reversal invariance), which implies the reality of $n^{(g)\,i}_{\alpha,\beta}$.} of the minimal fusion algebra. It remains to prove our earlier assertion that the so--defined $\tilde n^{(g)\,i}$ are solutions to Cardy's equation (\ref{cylt}). Since they satisfy the fusion algebra, the $\tilde n^{(g)\,i}$ must have eigenvalues given by ratios ${S_{i,j} \over S_{1,j}}$ of $S$ matrix elements. It is not difficult to see, by looking first at the partition functions $Z_{g,e}$ to get ${\cal E}_g$ and then by computing the ratios explicitly, that the eigenvalues of $\tilde n^{(g)\,i}$ are precisely the above ratios for $j \in {\cal E}_g$ (see next section). Thus the following diagonalization formulae hold \begin{equation} \tilde n^{(g)\,i}_{\alpha,\beta} = \sum_{j \in {\cal E}_g} \; \psi^{(g)\,j}_\alpha \, {S_{i,j} \over S_{1,j}} \, \bar\psi_\beta^{(g)\,j}, \label{nig} \end{equation} where the vectors $\psi^{(g)\,j}$ form a common orthonormal eigenbasis (also common to all fused adjacency matrices $F^i(T,G^\sigma)$ of the fixed point diagram). This yields the value of the coefficients in (\ref{cylt}) \begin{equation} c^{(g)\,j}_\alpha = {1 \over \sqrt{S_{1,j}}}\,\psi^{(g)\,j}_\alpha. \label{coeff} \end{equation} To complete the proof, it is enough to show that they are compatible with the $n^i$, in the sense that has been explained in Section \ref{sec:sym}: an entry in $n^i$ equal to $n$ goes over, in $\tilde n^{(g)\,i}$, to a sum of $n$ roots of unity, and moreover $\tilde n^{(g)\,1} = {\bf 1}$. One may verify that this is indeed the case. We omit the proof here since, to a large extent, it is given in the next section. The formulae (\ref{fff}) and (\ref{nig}) bear much resemblance with the corresponding ones for $n^i$, of which they constitute a natural extension. Like the $n^i$, the matrices $\tilde n^{(g)\,i}$ have a graph theoretic description derived from that of $n^i$ through the action of $g$, they satisfy the minimal fusion algebra, and their eigenvalues are exactly labelled by the set ${\cal E}_g$ which specifies the diagonal terms of the twisted partition functions $Z_{g,e}$. In a sense, this set ${\cal E}_g$ can also be viewed as the set of exponents of the fixed point graph that serves to define $\tilde n^{(g)\,i}$. \section{More explicit formulae} \label{sec:expl} We give here the computational details and the proofs that were missing in the previous section. We begin by recalling the formula giving the $S$ matrix elements in the minimal model ${\cal M}(p,q)$, for $i=(r,s)$ and $j=(r',s')$, \begin{equation} S_{i,j} = \sqrt{8 \over pq} \, (-1)^{rs'+r's+1} \, \sin{\pi qrr' \over p} \, \sin{\pi pss' \over q}. \end{equation} We examine in turn each of the three infinite series. \subsection{The series (A,A)} The models $(A_{p-1},A_{q-1})$, $p$ odd and $q$ even, have the symmetry group $Z_2$. The invariant boundary conditions $\alpha=(a,b)$ are controlled by the tadpole graph $T_{(p-1)/2} \times A_1$, i.e. $a$ runs from 1 to $(p-1)/2$ and $b=q/2$. The frustrated partition functions are \cite{rv}, \begin{equation} Z_{g,e}(A,A) = {\textstyle {1 \over 2}} \sum_{r,s} \chi_{r,s}^* \, \chi^{}_{r,q-s} = \sum_{1 \leq r\;\text{odd} \leq p-1 \atop 1 \leq s \leq q-1} \; \chi_{r,s}^* \, \chi^{}_{r,q-s}, \end{equation} from which it follows that the twisted Ishibashi states $|j\rangle\rangle_g$ can be labelled by \begin{equation} {\cal E}_g(A,A) = \{j=(m,{\textstyle {q \over 2}}) \;:\; 1 \leq m \;{\rm odd} \leq p-1\}. \end{equation} (Which representative $(r,s)$ or $(p-r,q-s)$ we take does not matter, since the $S$ matrix elements are the same.) For these values of $j$, an easy calculation yields \begin{equation} {S_{i,j} \over S_{1,j}} = (-1)^{(r+1)({q \over 2}+1)}\,U_{r-1}(-2\cos{\textstyle {\pi qm \over p}})\,U_{s-1}(0). \label{ratios1} \end{equation} Since $q$ is even, the numbers which appear as arguments of $U_{r-1}$ coincide with the set (\ref{spectrum}) of eigenvalues of the incidence matrix $T_{(p-1)/2}$. A simple comparison with the matrices $\tilde n^{(g)\,i}$, as computed from (\ref{nigf}) and (\ref{deltas}), \begin{equation} \tilde n^{(g)\,i} = (-1)^{(r+1)({q \over 2}+1)} \, U_{r-1}(T_{p-1 \over 2}) \, U_{s-1}(0). \end{equation} shows that the eigenvalues of $\tilde n^{(g)\,i}$ are indeed the numbers in (\ref{ratios1}) for $j \in {\cal E}_g$. As mentioned before, the matrices $n^i$ are the fusion matrices $N^i$ themselves \cite{cardy3}, equal, from (\ref{ni}), to \begin{equation} n^i_{(a_1,{q \over 2}),(a_2,{q \over 2})} = N^i_{(a_1,{q \over 2}),(a_2,{q \over 2})} = U_{r-1}(T_{{p-1\over 2}})_{a_1,a_2}, \end{equation} for all odd $s$, and identically equal to zero for $s$ even. This then leads to \begin{equation} \tilde n^{(g)\,i}_{(a_1,{q \over 2}),(a_2,{q \over 2})} = (-1)^{(r+1)({q \over 2}+1)+{s-1 \over 2}} \, N^i_{(a_1,{q \over 2}),(a_2,{q \over 2})}. \label{aa} \end{equation} This equation shows clearly that $\tilde n^{(g)\,i}$ is compatible with $n^i$ in the sense explained before. \subsection{The series (A,D)} All models $(A_{p-1},D_{q/2+1})$, with two coprime integers $p,q$ and $p$ odd as before, have also a $Z_2$ symmetry. The non--trivial group element $g$ induces the automorphism $\sigma$ of $D_{q/2+1}$ which exchanges the last two nodes. Therefore the symmetric boundary states correspond to the nodes $(a,b)$ of the fixed point diagram $T_{(p-1)/2} \times A_{q/2-1}$, pictured in Section \ref{sec:sym}, so that $a$ is between 1 and $(p-1)/2$, and $b$ is between 1 and $q/2-1$. The eigenvalues of $T_{(p-1)/2}$ have been recalled earlier, while those of $A_{q/2-1}$ are well known: \begin{eqnarray} && \text{spec}(T_{{p-1\over 2}}) = \{2\cos{\textstyle{\pi m \over p}}\;:\; 1 \leq m\,\text{odd} \leq p-1\}, \label{spect} \\ && \text{spec}(A_{{q \over 2}-1}) = \{2\cos{\textstyle{\pi m' \over q}}\;:\; 1 \leq m'\,\text{even} \leq q-1\}. \label{speca} \end{eqnarray} The frustrated (antiperiodic) partition function on the torus is (the double sums run over $[1,p-1]\times[1,q-1]$) \cite{rv} \begin{equation} Z_{g,e}(A,D) = \sum_{r \, {\rm odd} \atop s \,{\rm even}} |\chi^{}_{r,s}|^2 + \sum_{r \, {\rm odd} \atop s=1+{q \over 2} \bmod 2} \chi_{r,s}^*\chi^{}_{r,q-s}. \end{equation} Thus the Kac labels of the $g$--Ishibashi states $|j\rangle\rangle_g$ can be chosen in the set \begin{eqnarray} {\cal E}_g(A,D) &=& \{j=(m,m') \;:\; 1 \leq m \;{\rm odd} \leq p-1, \nonumber\\ && \hskip 0.6cm 1 \leq m' \;{\rm even} \leq q-1\}. \end{eqnarray} {}From this, one computes \begin{equation} {S_{i,j} \over S_{1,j}} = (-1)^{r+1} \, U_{r-1}(-2\cos{\textstyle{\pi qm \over p}}) \, U_{s-1}(-2\cos{\textstyle{\pi pm' \over q}}), \label{rat} \end{equation} which coincide, in view of (\ref{spect}) and (\ref{speca}), with the eigenvalues of \begin{equation} \tilde n^{(g)\,i}_{\alpha,\beta} = (-1)^{r+1} \: U_{r-1}(T_{{p-1\over 2}})_{a_1,a_2} \, U_{s-1}(A_{{q \over 2}-1})_{b_1,b_2}. \label{ad} \end{equation} The numbers in the set $\{2\cos{\textstyle{\pi pm' \over q}}\}$ come by pairs of opposite sign, so that the set of ratios (\ref{rat}), for fixed $i$, is the same whether or not there is a minus sign in the argument of $U_{s-1}$. Each individual ratio however differs by a factor $(-1)^{s+1}$, which then leads to an alternative solution $(-1)^{s+1} \tilde n^{(g)\,i}$. Finally the compatibility of $\tilde n^{(g)\,i}$ with the original matrices $n^i$ can be established. In the sector of invariant boundary conditions, the latter read \begin{equation} n^i_{\alpha,\beta} = U_{r-1}(T_{{p-1\over 2}})_{a_1,a_2} \, U_{s-1}(D_{{q \over 2}+1})_{b_1,b_2}, \end{equation} where $b_1,b_2$ are restricted to lie in $[1,q/2-1]$. One may simply notice the following modular identity (same values of the indices) \begin{equation} U_{s-1}(D_{{q \over 2}+1}) = U_{s-1}(A_{{q \over 2}-1}) \bmod 2. \end{equation} It has the immediate consequence that \begin{equation} \tilde n^{(g)\,i}_{\alpha,\beta} = n^i_{\alpha,\beta} \bmod 2, \end{equation} which shows the required compatibility. One may note that all the entries of $\tilde n^{(g)\,i}$ are in $\{0,+1,-1\}$, and that those of $n^i$ are in $\{0,1,2\}$, which implies that all doubled primary fields have opposite $Z_2$ charges within each pair. When $q=6$, i.e. for the $(A_{p-1},D_4)$ models, $Z_3$ invariant boundary conditions can be investigated. They are labelled by nodes $(a,2)$ with $a$ in $T_{(p-1)/2}$. The $Z_3$ frustrated partition functions on the torus are \cite{rv} \begin{equation} Z_{g,e}(A,D_4) = \sum_{r\,{\rm odd}} |\chi^{}_{r,3}|^2 + \sum_{r\,{\rm odd}} \chi^*_{r,3}[\chi^{}_{r,1}+\chi^{}_{r,5}]+{\rm c.c.}, \end{equation} so that the Ishibashi states in the $Z_3$--twisted sector have labels $j=(m,3)$ for $m$ odd between 1 and $p-1$. The matrices $\tilde n^{(g)\,i}$ in (\ref{nigf}) can be compared with the restriction of $n^i$ to the sector of invariant boundary conditions, given by $U_{r-1}(T_{(p-1)/2})_{a_1,a_2} U_{s-1}(D_4)_{2,2}$. All matrices are identically zero for $s$ even, while for $s$ odd: \begin{eqnarray} && n^i = U_{r-1}(T_{{p-1\over 2}}) = \tilde n^{(g)\,i}, \quad \text{for}\;s=1,5,\nonumber\\ && n^i = 2U_{r-1}(T_{{p-1\over 2}}), \quad \tilde n^{(g)\,i} = -U_{r-1}(T_{{p-1\over 2}}), \quad \text{for}\;s=3. \nonumber\\ \label{z3} \end{eqnarray} As in the $Z_2$ case, the second line shows that the doubled fields have opposite and non--zero $Z_3$ charge (if $\omega \neq 1$ is a third root of unity, $\omega +\omega^2=-1$). \subsection{The series (A,E$_{\bf 6}$)} The models $(A_{p-1},E_6)$ are similar to the $(A,D)$ models. In particular the formula for the matrices $\tilde n^{(g)\,i}$ is the same as for the $(A,D)$ models (with $A_{q/2-1}$ replaced by $A_2$). A unique feature of the models based on $E_6$ however is that some of the fields occur tripled in some boundary conditions (in addition to some others being doubled). One finds that these are the fields $(r,s)$ with $s=5$ and 7, in the boundary conditions corresponding to the nodes $(a,3)$, for $a$ in $T_{(p-1)/2}$ (with $b=3$ the intersection of the three branches of $E_6$). This follows from the fused adjacency matrices $U_4(E_6)$ and $U_6(E_6)$, which, when restricted to the nodes $b=3,6$ corresponding to the symmetric boundary conditions, read \begin{equation} U_4(E_6) = U_6(E_6) = \pmatrix{3 & 0 \cr 0 & 1}. \end{equation} \section{Unicity} The boundary conditions that are invariant under a group element $g$ correspond to boundary states which have expansions in $g$--Ishibashi states \footnote{The full expansion of $|\alpha\rangle$ involves Ishibashi states from the $g$--twisted bulk sectors for all $g$ which leave $\alpha$ invariant.} \begin{equation} |\alpha\rangle = \sum_{i \in {\cal E}_g} \; c^{(g)\,i}_\alpha \, |i\rangle\rangle_g. \end{equation} The coefficients in (\ref{coeff}) provide a specific solution $\tilde n^{(g)\,i}$ to Cardy's equation (\ref{cylt}). As for the $n^i$, one may raise the question of the unicity of this solution. For every $g$, the symmetric boundary conditions exhaust the $g$--Ishibashi states. It means that every other symmetric boundary state must be a linear combination of those we already have, and so must be one of them. However, since the boundary states $|\alpha \rangle$ enter Cardy's formula through scalar products, it is the boundary rays more than the boundary states which matter. Thus the basic question is whether one keeps a sensible solution if one multiplies the boundary states by phases. Clearly if the symmetric boundary states are multiplied by phases, $|\alpha\rangle \rightarrow \varphi_\alpha |\alpha\rangle$, the matrices change according to $\tilde n^{(g)\,i}_{\alpha,\beta} \rightarrow \varphi_\alpha \varphi^*_\beta \tilde n^{(g)\,i}_{\alpha,\beta}$, which satisfy the minimal fusion algebra for any choice of phases. Whereas for $g=e$, the positivity of $n^{(e)\,i}=n^i$ forces all the phases to be equal, this is no longer the case when $g \neq e$. Since the matrices $n^{(g)\,i}$ are $\Bbb Z$--valued, the only condition one has is that the phases must be equal up to signs, $\varphi_\alpha=\epsilon_\alpha \, \varphi$. For a $Z_2$ symmetry (or subgroup), the new matrices $\epsilon_\alpha \epsilon_\beta \tilde n^{(g)\,i}_{\alpha,\beta}$ are also solutions of the Cardy equation, because they too are compatible with the $n^i$. Indeed the compatibility amounts to check that $n^i$ and $\tilde n^{(g)\,i}$ coincide modulo 2, which obviously remains true if signs are inserted. Moreover, the identity occurs in the diagonal boundary conditions only, $\alpha = \beta$, for which the signs cancel out. On the contrary, in the case of a $Z_3$ symmetry, the insertion of signs $\epsilon_\alpha$ does not yield sensible solutions (as far as the minimal models are concerned). The reason is that some of the fields occur with multiplicity two. Since the corresponding entries in $n^{(g)\,i}$ must be real combinations of two third roots of unity, they can only be 2 or $-1$. Therefore, changing their sign by inserting some $\epsilon_\alpha$ is not consistent. Thus when the symmetry group is $Z_2$, there is a vast number of seemingly acceptable solutions. These various solutions differ by the charges which are assigned to the primary fields in mixed boundary conditions ($\alpha \neq \beta$). The freedom we have in choosing the $\epsilon_\alpha$ reflects the fact that the charge normalization in mixed boundary conditions cannot be fixed a priori, unlike what happens for diagonal boundary conditions, in which an identity occurs. One may try to derive more constraints on the charge normalizations by requiring that the boundary charge assignments be compatible with {\it (i)} the charge assignments in the bulk, and {\it (ii)} the boundary field operator product coefficients. The first requirement is a condition on the way bulk fields close to a boundary (taken to be $y=0$) expand in boundary fields \cite{cl,l} \begin{equation} \phi_j(x+iy) \sim \sum_{{\rm b.c.}\;\alpha} \sum_k \; ^{(\alpha)}\!B_j^k \, (2y)^{h_k-2h_j} \, \phi_k^{\alpha\alpha}(x), \label{bbe} \end{equation} where the summation on $\alpha$ is over all boundary conditions, not just the invariant ones. The $Z_2$ symmetry implies selection rules on the coefficients since a bulk field of a given parity should expand in a combination of boundary fields that transforms the same way. It means that the parity of the field $\phi_k^{\alpha \alpha}$ must match that of $\phi_j$ for all invariant boundary conditions $\alpha$ such that $^{(\alpha)}\!B_j^k \neq 0$. Since these expansions involve fields in diagonal boundary conditions only, the selection rules that follow are the same no matter what the signs $\epsilon_\alpha$ are. This does not prove however that the selection rules are indeed satisfied. For the diagonal models $(A,A)$, the coefficients $^{(\alpha)}\!B_j^k$ are known explicitly \cite{r}, and it would be interesting to check directly that their values are consistent with the boundary charge assignment found here. The second check concerns the operator algebra of the boundary fields themselves \cite{cl,l} \begin{equation} \phi^{\alpha\beta}_i(x) \, \phi^{\beta\gamma}_j(x') \sim \sum_k \; C^{(\alpha\beta\gamma)\,k}_{ij}(x-x')^{h_k-h_i-h_j}\,\phi^{\alpha\gamma}_k(x'). \label{bope} \end{equation} Restricting to invariant boundary conditions $\alpha,\beta,\gamma$, the discrete symmetry implies again selection rules which require that the charges given by the matrices $n^{(g)\,i}$ provide a grading of the boundary fusion algebra\footnote{We leave aside the cases where some matrix elements $n^{(g)\,i}_{\alpha,\beta}$ are zero without having the corresponding elements in $n^i$ equal to zero. This happens when primary fields come in pairs of opposite charge.}: \begin{equation} C^{(\alpha\beta\gamma)\,k}_{ij} \neq 0 \quad \Longrightarrow \quad n^{(g)\,i}_{\alpha,\beta} n^{(g)\,j}_{\beta,\gamma} = n^{(g)\,k}_{\alpha,\gamma}. \label{bbf} \end{equation} It is obvious that if the matrix coefficients $\tilde n^{(g)\,i}_{\alpha,\beta}$ satisfy (\ref{bbf}), the same will be true of $\epsilon_\alpha \epsilon_\beta \tilde n^{(g)\,i}_{\alpha,\beta}$, so that here too, these matrices are all consistent with the boundary operator product expansion (\ref{bope}), or else none of them is. As the discrete symmetry is expected to occur, one can be confident in the fact that the selection rules will be satisfied. We give below examples of selection rules in the most explicit case, namely the diagonal models. We have not shown in general that they are indeed satisfied, and as before, a proof not based on symmetry arguments would be valuable. In the diagonal models $(A,A)$, the boundary conditions are in one--to--one correspondence with the chiral primary fields through their labelling by two nodes $(a,b)$ taken in $A_{p-1}$ and $A_{q-1}$. As before, the boundary conditions $(a^*,b^*) = (p-a,q-b)$ and $(a,b)$ are to be identified. Without loss of generality, one may thus assume that the first label (the ``$r$--label'') is odd. The boundary operator product coefficients are known explicitly from \cite{r}, where they were proved to be equal to coefficients of the crossing matrices (in a suitable normalization) \begin{equation} C^{(\alpha\beta\gamma)\,k}_{ij} = F_{\beta,k}\left[\matrix{\alpha & \gamma \cr i & j}\right]. \end{equation} Since for instance, an odd boundary field $\phi^{\alpha\alpha}_i$ cannot occur in its fusion with itself, the corresponding crossing coefficient must vanish. The verification that it does is non--trivial only when the chiral field $i$ indeed occurs in its own bulk fusion (namely $N_{ii}^i \neq 0$), when the primary field $i$ indeed occurs in the diagonal boundary conditions ($n^i_{\alpha,\alpha} \neq 0$ for $\alpha$ invariant under $Z_2$), and when it is an odd field ($\tilde n^{(g)\,i}_{\alpha,\alpha} = -1$). All three conditions can be easily worked out, and yield \begin{equation} F_{\alpha,i}\left[\matrix{\alpha & \alpha \cr i & i}\right] = 0 \label{f} \end{equation} for all $i=(r,s)$ such that $r,s$ are odd, $s=3 \bmod 4$, $r \leq (2p-1)/3$, $s \leq (2q-1)/3$, and for all $\alpha=(a,q/2)$ such that $(r+1)/2 \leq a \leq p/2$. The simplest example where such constraints arise is the tetracritical Ising model $(A_4,A_5)$, in which (\ref{f}) implies (in terms of conformal weights) \begin{equation} F_{{1 \over 15},{1 \over 15}}\left[{\scriptsize \matrix{{1 \over 15} & {1 \over 15} \cr {1 \over 15} & {1 \over 15}}}\right] = F_{{1 \over 15},{2 \over 3}}\left[{\scriptsize \matrix{{1 \over 15} & {1 \over 15} \cr {2 \over 3} & {2 \over 3}}}\right] = F_{{2 \over 3},{2 \over 3}}\left[{\scriptsize \matrix{{2 \over 3} & {2 \over 3} \cr {2 \over 3} & {2 \over 3}}}\right] = 0. \end{equation} More conditions can be derived in a generic diagonal model. To summarize, the matrices $\tilde n^{(g)\,i}$ displayed in (\ref{nigf}) and (\ref{deltas}) yield but a particular solution to Cardy's equation. For a $Z_3$ symmetry, they form the only consistent solution, \begin{equation} n^{(g)\,i}_{\alpha,\beta} = \tilde n^{(g)\,i}_{\alpha,\beta}, \qquad (g^3=e), \end{equation} whereas, in the case of a $Z_2$ symmetry, there are many more given by \begin{equation} n^{(g)\,i}_{\alpha,\beta} = \epsilon_\alpha \epsilon_\beta \tilde n^{(g)\,i}_{\alpha,\beta}, \quad \epsilon_\alpha = \pm 1 \qquad (g^2=e), \label{nnn} \end{equation} for arbitrary signs. The effect of these signs is to reverse (or to maintain, depending to the value of $\epsilon_\alpha \epsilon_\beta$) the parity of all the fields that occur in the sector of boundary conditions $\alpha,\beta$. The ambiguity in the normalization of the $Z_2$ charges that arises due to these signs must be resolved on physical grounds. As the interpretation of the boundary fields is lacking in the general non--unitary model, it is not clear to the author what the correct requirement should be. In this context, the specific choice $\epsilon_\alpha=+1$ for all $\alpha$ is a minimal and natural one, as it extends nicely the corresponding formula for $g=e$, and retains much of the graph theoretic description. It also has the distinctive feature of producing matrices $\tilde n^{(g)\,i}$ of constant sign, either totally positive or totally negative\footnote{There is another solution in terms of matrices of constant sign, which is obtained by substituting $-G^\sigma$ for $G^\sigma$ in the formula (\ref{nigf}) giving $\tilde n^{(g)\,i}$. The substitution has no effect when $G^\sigma=A_1$, since the associated adjacency matrix is the number zero, while in the other cases, it causes the matrices $\tilde n^{(g)\,i}$ to be multiplied by $(-1)^{s+1}$. This sign can be seen to be in the line of the previous discussion, because it is equal to $(-1)^{s+1} = \epsilon_\alpha \epsilon_\beta$ with $\epsilon_\alpha = (-1)^{b+1}$ if $\alpha=(a,b)$. The existence of this solution is a consequence of a non--trivial automorphism of the graph $G^\sigma$.}. However in view of what follows, this may not be the correct choice. In a unitary model, the ground state of every sector is expected to be invariant under the symmetry group, on account of the Perron--Frobenius theorem applied to the transfer matrix. This provides a well--defined criterion to fix the normalization of the charges, and therefore the physical value of the signs $\epsilon_\alpha$. We will use this criterion as a guide, in order to see if a particular set of values $\epsilon_\alpha$ emerges from this point of view. \section{Unitary models} In this last section, we explore the possibility of fixing the value of the signs $\epsilon_\alpha$ by using the criterion we have just mentioned: if the continuum limit is smooth enough, it is expected that the consequences of the Perron--Frobenius theorem on the finite--dimensional transfer matrix be maintained in the corresponding conformal field theory. In particular, for all invariant boundary conditions, the ground state of the Hamiltonian $H_{\alpha,\beta}$ (the primary field of lowest conformal dimension in ${\cal H}_{\alpha,\beta}$) should be non--degenerate and (hence) invariant under the symmetry group. In short, we will call this the Perron--Frobenius (PF) criterion. As already said, it is automatically satisfied in the diagonal boundary conditions. Thus we look for a set of $\epsilon_\alpha$ such that the $Z_2$ charge assignment meet the PF criterion. Incidentally, when the symmetry group is $Z_3$, there is only one consistent charge assignment (see the previous section). In that case, we will merely check whether the PF criterion is satisfied. The outcome of this investigation is somewhat surprising. The unitary diagonal models are the only ones where the PF criterion can be met, for a unique choice of the $\epsilon_\alpha$'s. In all other unitary models, there is no way in which it can be fulfilled, if one insists that it be valid in all sectors. A physical interpretation of this will be proposed\footnote{I am indebted to Gerard Watts for a clarifying discussion about this issue.}. Nonetheless, for all those models but two, we will see that a unique set of $\epsilon_\alpha$'s is singled out by demanding a minimal violation of the PF criterion. We recall that the conformal weight of a primary field labelled by $i=(r,s)$ is equal to \begin{equation} h_{r,s} = {(qr-ps)^2-(p-q)^2 \over 4pq}. \end{equation} Throughout this section, we will take $p$ odd and $q=p \pm 1$ even. Then the smallest conformal weights correspond, in ascending order, to $i=(1,1), (2,2), (3,3), \ldots$. \subsection{The unitary series (A,A)} The only boundary primary fields that occur in the diagonal models have their $s$--label odd (see (\ref{aa})). Since the identity (1,1) does not appear in mixed boundary conditions, the primary with the lowest weight that can possibly occur in mixed boundary conditions corresponds to (3,3), and consequently, the off--diagonal entries of \begin{equation} n^{(g) \, (3,3)}_{\alpha,\beta} = n^{(g) \, (3,3)}_{(a_1,{q \over 2}),(a_2,{q \over 2})} = -\epsilon_{a_1}\,\epsilon_{a_2}\,U_2(T_{p-1 \over 2})_{a_1,a_2} \end{equation} must be positive. The off--diagonal matrix coefficients $U_2(T)_{a_1,a_2}$ equal 1 if $|a_1-a_2|=2$ or if $\{a_1,a_2\}=\{(p-3)/2,(p-1)/2\}$, and 0 otherwise (it counts the number of paths of length 2 going from $a_1$ to $a_2$ on the graph $T_{(p-1)/2}$). Thus one obtains the condition $\epsilon_{a_1} \epsilon_{a_2}=-1$ for all these pairs. This fixes the vector $\epsilon_a$ in a unique way (up to a global sign that does not matter) as \begin{equation} \epsilon_a = (\ldots,+1,+1,-1,-1,+1,+1,-1,-1,+1). \label{epa} \end{equation} For these specific signs, one may then verify that in the remaining mixed boundary sectors (those for which (3,3) does not occur), the field of lowest weight has indeed a parity +1 (zero charge). To do that, one can first observe that any mixed boundary sector has its field of lowest weight in $\{(r,s) \;:\; 3 \leq r=s \;{\rm odd} \leq p-2\}$. Next point is to note that $U_{r-1}(T)_{a_1,a_2}$ is zero unless the nodes $a_1,a_2$ can be related by a path of length $r-1$. If the two nodes cannot be connected by a shorter path, it follows from (\ref{epa}) that $\epsilon_{a_1} \epsilon_{a_2} = (-1)^{(r-1)/2}$, so that the numbers \begin{equation} n^{(g) \, (r,r)}_{(a_1,{q \over 2}),(a_2,{q \over 2})} = \epsilon_{a_1} \epsilon_{a_2} (-1)^{(r-1)/2} U_{r-1}(T)_{a_1,a_2} \end{equation} are positive (or zero). That $a_1$ and $a_2$ can be connected by a shorter path means that the field $(r,r)$ is not the primary with the lowest weight in that sector, and we are back to the first case. Since the PF criterion can be satisfied in all sectors for a unique set of $\epsilon_\alpha$'s, it is tempting to conjecture that these are the correct physical values. The charge content in the various sectors of the unitary diagonal models would then be given by \begin{equation} n^{(g) \, i}_{(a_1,{q \over 2}),(a_2,{q \over 2})} = \epsilon_{a_1} \, \epsilon_{a_2} \, \tilde n^{(g) \, i}_{(a_1,{q \over 2}),(a_2,{q \over 2})}, \end{equation} with the signs (\ref{epa}), and the $\tilde n^{(g)\,i}$ as in (\ref{aa}). \subsection{The unitary series (A,D)} The same calculations can be carried out for the unitary models of the $(A,D)$ series, with however different results. To illustrate it most clearly, we will start with the simplest case, namely $(A_4,D_4)$, corresponding to the critical 3--Potts model ($p=5$, $q=6$). A set of $Z_2$--symmetric boundary conditions is provided\footnote{The model has eight conformally invariant boundary conditions which are invariant under a $Z_2$, but not under the same $Z_2$. One finds three groups of four boundary conditions that are simultaneously invariant under the same $Z_2$. They clearly correspond to the three conjugate $Z_2$ subgroups of $S_3$, the automorphism group of $D_4$.} by the so--called A, BC, Free and New \cite{aos}. They correspond respectively to the nodes (1,1), (2,1), (1,2) and (2,2). (Free and New, being fully invariant under $S_3$, must correspond to $b=2$, which is the only node of $D_4$ invariant under $S_3$.) Together they define 10 different sectors. It is not difficult to find the field with lowest weight in each of these sectors, and then compute the parity they are assigned by the matrices $\tilde n^{(g)\,i}$ computed in Section \ref{sec:expl}. Writing these in a matrix $\tilde M$, one obtains (indices are A, BC, Free, New) \begin{eqnarray} \tilde M_{\alpha,\beta} &=& \Big(\tilde n^{(g)\,imin}_{\alpha,\beta} \;:\; \min_{i \in {\cal H}_{\alpha,\beta}} h_i = h_{imin}\Big)_{\alpha,\beta} \nonumber\\ &=& {\footnotesize \pmatrix{+1 & -1 & +1 & -1 \cr -1 & +1 & -1 & -1 \cr +1 & -1 & +1 & 0 \cr -1 & -1 & 0 & +1}}. \end{eqnarray} The zeros are due to the partition function (superscripts are the conformal weights) \begin{equation} Z_{\rm Free,New} = 2\,\chi_{3,3}^{1/15} + \chi_{3,5}^{2/5} + \chi_{3,1}^{7/5}, \end{equation} which shows that the ground state in that sector is doubly degenerate, the two states having opposite parities. The above matrix makes clear that the charge assignment implied by $\tilde n^{(g)\,i}$ does not satisfy the PF criterion in all sectors, either because the ground state is not invariant, or because it is degenerate. One may try to find values for $\epsilon_\alpha$ that render the non--degenerate ground states invariant, but one easily sees that it is not possible: no values for $\epsilon_\alpha$ can be found so that $\tilde M_{\alpha,\beta} \geq 0$ for all $\alpha,\beta$. One can relax our demands by looking for a set of $\epsilon_\alpha$ which minimizes the number of sectors that violate the PF criterion. One then finds that the minimal number of such sectors, which we call non--PF, is equal to \begin{equation} N_{\rm non-PF} = 2. \end{equation} This number is realized for $\epsilon_\alpha = (+1,-1,+1,-1) = (+1,-1)_a \otimes (1,1)_b$, the two non--PF sectors being BC,New and Free,New. Indeed for these $\epsilon_\alpha$, one obtains \begin{equation} \epsilon_\alpha \epsilon_\beta \tilde M_{\alpha,\beta} = {\footnotesize \pmatrix{+1 & +1 & +1 & +1 \cr +1 & +1 & +1 & -1 \cr +1 & +1 & +1 & 0 \cr +1 & -1 & 0 & +1}}. \end{equation} Let us also notice that if one excludes just one boundary condition, namely ``New'', the expected consequences of the PF theorem are indeed verified. Thus in this case, the minimal number of boundary conditions that have to be excluded for this to be true is equal to 1. Finally one may note that $\epsilon_\alpha = (+1,-1,+1,+1)$ share the same properties, the two non--PF sectors being now A,New and Free,New. In any case, one must conclude that the transfer matrix, in certain sectors of boundary conditions, does not satisfy the conditions of the PF theorem. There can be only two reasons for that: either the transfer matrix is not irreducible\footnote{The unicity of the largest eigenvalue is guaranteed only for non--negative primitive matrices \cite{ms}. Under mild assumptions on the transfer matrix, its irreducibility is sufficient.}, or else it contains negative entries, implying that some of the boundary Boltzmann weights are negative (or both). That the first condition fails is unlikely because the periodic transfer matrix is irreducible and because the boundary conditions are undecomposable. So one should favour the second alternative, which points to the unphysical nature of some of the boundary conditions, their classical description requiring negative Boltzmann weights. We note that a boundary condition $\alpha$ which is described by negative Boltzmann weights does not necessarily lead to unphysical (negative, non--PF) partition functions. Whether or not this is the case depends on which other boundary condition is associated with $\alpha$. The appearance of negative classical boundary Boltzmann weights to describe the New boundary condition in the critical 3--Potts model has been discussed in \cite{aos}, and is confirmed by the explicit calculation of the critical boundary weights \cite{p}. As we shall see, what is true in the 3--Potts model is true in all unitary models of the $(A,D)$ series. No values for the $\epsilon_\alpha$'s exist which make all sectors to satisfy the PF criterion, but a suitable choice, unique, contrary to the above case, of $\epsilon_\alpha$ minimizes the number of sectors which do not satisfy it. As above we will take the point of view that these features are the signal that a certain number of boundary conditions are unphysical, because they require negative Boltzmann weights for their classical description. We have not carried out the analysis of the whole series, but instead we have investigated the first eight models, up to $p=13$ and $q=12$, with the following results. In each of these models, we have determined the minimal number $N_{\rm unphys}$ of boundary conditions that must be disregarded in order to satisfy the PF criterion in all the sectors involving the remaining ones. This uniquely singles out a set of boundary conditions, which naturally qualifies as the set of unphysical boundary conditions. This also determines unique values of the $\epsilon_\alpha$ for the physical ones. The values of $\epsilon_\alpha$ for the unphysical $\alpha$ are then fixed (uniquely, except in the 3--Potts model) by requiring a minimal number of non--PF sectors (which necessarily correspond to one or two unphysical boundary conditions). That minimal number is denoted by $N_{\rm non-PF}$. The results are as follows. In the model $(A_{p-1},D_{q/2+1})$ (we have looked at the eight models corresponding to $6 \leq q \leq 12$), the number $N_{\rm unphys}(p,q)$ only depends on the rank of the $D$ factor. It increases rather quickly since it is equal to 1, 3, 6 and 10 for the two models involving the algebra $D_4$, $D_5$, $D_6$ and $D_7$ respectively. We found that the unphysical boundary conditions form the set (the labelling of the nodes is as in the figure of Section \ref{sec:sym}) \begin{equation} \{\alpha = (a,b) \in T_{p-1 \over 2} \times A_{{q \over 2}-1} \;:\; a+b \geq {\textstyle {p+3 \over 2}}\}. \label{excl} \end{equation} Moreover, the signs which make the number of non--PF sectors minimal are unique and given by \begin{eqnarray} \epsilon_\alpha &=& (+1,-1,+1,-1,\ldots)_a \otimes (1,1,1,\ldots)_b \nonumber\\ &=& (-1)^{a+1}, \qquad \alpha=(a,b). \label{epd} \end{eqnarray} As pointed out above, in the model $(A_4,D_4)$, there is another solution $\epsilon_\alpha = (+1,-1,+1,+1)$, which however appears to contradict the duality relations (see below). We have determined $N_{\rm non-PF}$ by mere counting, and found that it equals 2, 3, 11, 15, 36, 46, 89, 109 for the first eight models, ordered as $(A_4,D_4)$, $(A_6,D_4)$, $(A_6,D_5)$, ... (By symmetry, the sectors $(\alpha,\beta)$ and $(\beta,\alpha)$ are identical and count for one.) These results strongly suggest the general pattern in which the number of boundary conditions in (\ref{excl}) equals a binomial coefficient \begin{equation} N_{\rm unphys}(A,D_{{q \over 2}+1}) = {{q \over 2}-1 \choose 2}. \end{equation} This is a large number since essentially half the invariant boundary conditions would have to be discarded as classically unphysical. A bit more of numerology also shows that the number of non--PF sectors fits the simple formula \begin{eqnarray} N_{\rm non-PF}(A_{q\mp 1-1},D_{{q \over 2}+1}) &=& \nonumber\\ && \hspace{-1.5cm} \big\{({\textstyle {q-2 \over 4}})^4\big\} + {\textstyle {q(q-2)(q \mp 2)(q-4) \over 192}}, \label{unphys} \end{eqnarray} where $\{x\}$ is the integer closest to $x$. The two numbers in the r.h.s. of the previous equation have separately a well--defined meaning: the first one is the number of sectors where the ground state is non--degenerate but odd under the $Z_2$ symmetry, while the second one gives the number of sectors where the ground state is doubly degenerate. The reader may wish to check the above assertions in a less simple instance than the Potts model. A good example is to consider the $(A_6,D_5)$ model, for which one computes (in the tensor product basis) \begin{equation} \tilde M_{\alpha,\beta} = {\footnotesize \pmatrix{ +1 & -1 & +1 & +1 & -1 & +1 & +1 & -1 & +1 \cr -1 & +1 & -1 & -1 & +1 & -1 & -1 & +1 & +1 \cr +1 & -1 & +1 & +1 & -1 & -1 & +1 & +1 & +1 \cr +1 & -1 & +1 & +1 & -1 & +1 & +1 & -1 & 0 \cr -1 & +1 & -1 & -1 & +1 & +1 & -1 & 0 & -1 \cr +1 & -1 & -1 & +1 & +1 & +1 & 0 & -1 & -1 \cr +1 & -1 & +1 & +1 & -1 & 0 & +1 & 0 & 0 \cr -1 & +1 & +1 & -1 & 0 & -1 & 0 & +1 & 0 \cr +1 & +1 & +1 & 0 & -1 & -1 & 0 & 0 & +1}}. \end{equation} The values of $\epsilon_\alpha$ mentioned in (\ref{epd}) are nothing but the first line of $\tilde M_{\alpha,\beta}$, and the boundary conditions to discard label the rows and columns 6, 8 and 9, which correspond, in terms of the fixed point graph $T_3 \times A_3$, to the pairs of nodes $(a,b) = (3,2),(2,3)$ and $(3,3)$, as announced in (\ref{excl}). There are 6 zeros in the upper triangular part of $\tilde M_{\alpha,\beta}$, which is the value of the second summand of (\ref{unphys}). All this leads to the reasonable guess that (\ref{epd}) might give the correct physical values of the $\epsilon_\alpha$'s. Inserted in (\ref{nnn}), it not only determines the parities of all primaries in the sectors where the PF criterion is satisfied, but it also points to the boundary conditions that can have a problematic lattice interpretation. These conjectural statements must be confirmed or dismissed by the explicit calculation of the boundary Boltzmann weights. The results obtained so far seem to give some support to our conjecture \cite{betal}. Assuming this conjecture, it is not difficult to give an explicit formula for the parities. From (\ref{nigf}), (\ref{deltas}) and (\ref{epd}), they are determined from \begin{eqnarray} n^{(g)\,i}_{(a_1,b_1),(a_2,b_2)} &=& (-1)^{a_1+a_2+r+1} \times \nonumber\\ && \hspace{-7mm} U_{r-1}(T_{(p-1) \over 2})_{a_1,a_2} \, U_{s-1}(A_{{q \over 2}-1})_{b_1,b_2}. \end{eqnarray} The matrices $U_{r-1}(T_{(p-1)/2})$ are all positive, unlike the $U_{s-1}(A_{q/2-1})$, which are positive for $s<q/2$, negative for $s>q/2$, and identically zero for $s=q/2$, on account of $U_{q-s-1}(A_{q/2-1}) = - U_{s-1}(A_{q/2-1})$. Putting all these observations together, one can conclude that the paired fields have opposite $Z_2$ parities within each pair (as already pointed out), and that the parity of an unpaired field in the sector of boundary conditions $\alpha,\beta$ is equal to \begin{equation} g(\phi_i^{\alpha\beta}) = \cases{ (-1)^{a_1+a_2+r+1} \, \phi_i^{\alpha\beta} & if $s<q/2$, \cr \noalign{\smallskip} (-1)^{a_1+a_2+r} \, \phi_i^{\alpha\beta} & if $s>q/2$.} \end{equation} In the critical 3--Potts model for instance, one finds the following frustrated partition functions (in terms of the conformal weights) \begin{eqnarray} && Z^g_{\rm A,A} = \chi_{0} - \chi_{3}, \\ && Z^g_{\rm A,BC} = \chi_{2/5} - \chi_{7/5}, \\ && Z^g_{\rm A,Free} = \chi_{1/8} - \chi_{13/8}, \\ && Z^g_{\rm BC,BC} = \chi_{0} - \chi_{3} - \chi_{2/5} + \chi_{7/5}, \\ && Z^g_{\rm BC,Free} = \chi_{1/40} - \chi_{21/40},\\ && Z^g_{\rm Free,Free} = \chi_{0} - \chi_{3} + \chi_{2/3} - \chi_{2/3^+},\\ && Z^g_{\rm A,New} = \chi_{1/40} - \chi_{21/40}, \label{anew} \\ && Z^g_{\rm New,New} = \chi_{0} - \chi_{3} - \chi_{2/5} + \chi_{7/5} \nonumber\\ && \hspace{1.5cm} + \chi_{2/3} - \chi_{2/3^+} + \chi_{1/15} - \chi_{1/15^+}. \end{eqnarray} These functions are computed using the $\epsilon_\alpha$'s given in (\ref{epd}), and appear to be consistent with the duality of the model \cite{aos}. For instance, the equality \begin{equation} Z_{\rm BC,Free} = Z_{\rm A,New} \label{dual} \end{equation} is maintained for the frustrated partition functions, while \begin{equation} Z_{\rm Free,Free} = Z_{\rm A,A} + Z_{\rm A,B} + Z_{\rm A,C} \end{equation} becomes $Z^g_{\rm Free,Free} = Z^g_{\rm A,A}$ since $Z^g_{\rm A,B} = Z^g_{\rm A,C} = 0$. The use of the other solution $\epsilon_\alpha = (+1,-1,+1,+1)$ has the effect of multiplying by $-1$ the partition functions of all sectors with one ``New'', so that $Z^g_{\rm A,New}$ would be minus the expression in (\ref{anew}), contradicting the duality relation (\ref{dual}). There is a $Z_3$ symmetry in two models only, namely the critical and tricritical 3--Potts models $(A_4,D_4)$ and $(A_6,D_4)$. They possess respectively 2 (``Free'' and ``New'') and 3 invariant boundary conditions, namely $\alpha=(a,2)$ for $a$ a node of $T_2$ and $T_3$. The relevant $\tilde M$ matrices are equal to \begin{equation} \tilde M_{\alpha,\beta} = \footnotesize{\pmatrix{+1 & -1 \cr -1 & +1}} \quad {\rm and} \quad \footnotesize{\pmatrix{+1 & +1 & -1 \cr +1 & +1 & -1 \cr -1 & -1 & +1}}, \end{equation} where a $-1$ sign indicates that the corresponding sector has two degenerate ground states, of opposite and non--zero charge (none of them is invariant under the $Z_3$). In the first case (the $(A_4,D_4)$ model), it is the second boundary condition $(2,2)$ (i.e. ``New'') that appears to be unphysical, while in the second case, it is the third boundary condition $(3,2)$. This should not be surprising since they are precisely the boundary conditions which were unphysical from the $Z_2$ point of view: from (\ref{excl}), $\alpha=(a,2)$ was to be discarded if $a+2 \geq (p+3)/2$, that is, if $a=(p-1)/2$. Therefore, the boundary conditions which were causing problems for the $Z_2$ charges also cause problems for the $Z_3$ charges. \subsection{The unitary models (A,E$_{\bf 6}$)} We will content ourselves with making a few comments on the two unitary models $(A_{10},E_6)$ and $(A_{12},E_6)$ ($p=11$ or 13, and $q=12$). As we have said above, the models involving the $E_6$ algebra have the peculiarity of possessing primary fields that occur with multiplicity 1, 2 and 3. It turns out that the same is true of the ground state in various sectors. Let us examine in some detail the simplest model $(A_{10},E_6)$. That model possesses 10 invariant boundary conditions, labelled as $\alpha=(a,b)$ with $a=1,2,\ldots,5$ a node of $T_5$, and $b=3,6$ a node of the $A_2$ subgraph of $E_6$, fixed by its non--trivial automorphism. One can compute as before the matrix $\tilde M_{\alpha,\beta}$ which collects those entries of $\tilde n^{(g)\,i}_{\alpha,\beta}$ for which $i$ is the lowest weight primary in the sector $\alpha,\beta$. The result is \begin{eqnarray} && \tilde M_{\alpha,\beta} = \nonumber\\ && {\footnotesize \pmatrix{ +1 & 0 & 0 & +1^* & -1^* & +1 & -1 & -1 & +1 & 0 \cr 0 & +1 & +1^* & 0 & -1^* & -1 & -1 & -1 & 0 & +1 \cr 0 & +1^* & +1 & -1^* & 0 & -1 & -1 & 0 & -1 & +1 \cr +1^* & 0 & -1^* & +1 & 0 & +1 & 0 & -1 & +1 & -1 \cr -1^* & -1^* & 0 & 0 & +1 & 0 & +1 & +1 & -1 & -1 \cr +1 & -1 & -1 & +1 & 0 & +1 & -1 & -1 & +1 & -1 \cr -1 & -1 & -1 & 0 & +1 & -1 & +1 & +1 & -1 & -1 \cr -1 & -1 & 0 & -1 & +1 & -1 & +1 & +1 & -1 & -1 \cr +1 & 0 & -1 & +1 & -1 & +1 & -1 & -1 & +1 & -1 \cr 0 & +1 & +1 & -1 & -1 & -1 & -1 & -1 & -1 & +1}}, \nonumber\\ \end{eqnarray} where the stars mean that the ground state in the corresponding sector is three times degenerate, the number $\pm 1$ being the sum of the three parities. As before, a zero indicates that there are two degenerate ground states with opposite parity. We can repeat what we did for the $(A,D)$ series, and look for a set of $\epsilon_\alpha$ which minimizes the violation of the PF criterion. By varying the $\epsilon_\alpha$, one finds that the minimal number of non--PF sectors is equal to 21, and that the non--PF sectors have at least one boundary condition in the set \begin{equation} \{(2,3), (3,3), (4,3), (5,3), (5,6)\} \end{equation} in terms of the nodes of $T_5 \times A_2$ (they correspond to the rows and columns 2, 3, 4, 5, 10). So these five boundary conditions can presumably be called unphysical in the sense of the previous subsection. Hence \begin{equation} N_{\rm unphys}(A_{10},E_6) = 5, \quad N_{\rm non-PF}(A_{10},E_6) = 21. \end{equation} There are four solutions for the $\epsilon_\alpha$'s for which these values can be realized. Among them, the most symmetrical one is $\epsilon_\alpha = (+1,-1,-1,+1,-1) \otimes (1,1)$. The other model $(A_{12},E_6)$ is similar. One finds \begin{equation} N_{\rm unphys}(A_{12},E_6) = 5, \quad N_{\rm non-PF}(A_{12},E_6) = 27. \end{equation} The presumably unphysical boundary conditions correspond to the nodes (3,3), (4,3), (5,3), (6,3), (6,6) of $T_6 \times A_2$. The signs for which these numbers are reached are unique and given by $\epsilon_\alpha = (+1,-1,+1,+1,-1,+1) \otimes (1,1)$. \acknowledgments It is a pleasure to thank Jean-Bernard Zuber for an encouraging and stimulating exchange, and for reading the manuscript. I also wish to thank Paul Pearce who has informed me of the preliminary results that he and his collaborators have obtained on boundary Boltzmann weights, Ingo Runkel for pointing out misprints in the draft, and Gerard Watts for a very fruitful discussion.
\section{Introduction} \vspace{-13.cm} \hfill {\Large STAR Note SN388} \vspace{13.cm} The study of collective flow in nuclear collisions at high energies has been attracting increasing attention from experimentalists~\cite{H-G,olli98}. This is partly because recent progress has been made in the development of new techniques suitable for flow studies at high energies~\cite{olli92,lvzh,ollimeth,olli98,posk98,meth}. Instead of studying $\mpx$, in these new methods a Fourier expansion of the azimuthal distribution of particles is used in which the first harmonic coefficient, $v_1$, quantifies the directed flow and the second harmonic coefficient, $v_2$, quantifies the elliptic flow. In some cases $A1$ and $A2$ were reported, which in modern terminology, are twice the square of the sub-event resolution. Using these new techniques anisotropic flow has now been observed for heavy symmetric systems at both the AGS and SPS. At the AGS the E877 Collaboration pioneered the use of the Fourier expansion method to measure $v_1$ and $v_2$. They studied these quantities (as well as $v_4$) from a calorimeter as a function of centrality in different pseudorapidity windows~\cite{l877flow1}. Then they studied nucleons as well as pions as a function of pseudorapidity for different centralities~\cite{l877flow3}. Using their spectrometer to identify particles while still obtaining the event plane from the calorimeter, they measured $v_1$ and $v_2$ as a function of $p_t$ for different rapidities and centralities~\cite{l877flow2}. They also reported $\mpx$ as a function of rapidity~\cite{l877flow2}. In their latest papers they extended this study to light nuclei~\cite{volo98,l877flow4}. The E802 Collaboration studied $\mpx$ for light nuclei in the target rapidity region using a forward hodoscope to determine the event plane~\cite{ahle98}. At the SPS NA49 first observed elliptic flow in a calorimeter study which reported $A2$ as a function of centrality~\cite{lna49}. WA98 reported $A1$ as a function of centrality for protons and $\pi^+$ in the target rapidity region~\cite{lwa98,wa9898}. They also studied $\mpx$ in the target rapidity region~\cite{wa9898}. NA45 used silicon drift detectors to study $v_1$ and $v_2$ as a function of pseudorapidity~\cite{lna45}. NA49 has presented a differential study of $v_1$ and $v_2$ as a function of $p_t$ and $y$~\cite{posk98} and has also started to study the centrality dependence~\cite{posk_annu99}. Also, the importance of flow for other measurements has just begun to be studied. For two particle correlations relative to the event plane the mathematical scheme has been worked out~\cite{lvc,wied97,heisel2,heisel}. Some first results have been given by WA98~\cite{lwa98}. Also, for non-identical particles the correlation relative to the event plane has been discussed~\cite{volo97}. \section{Physics Motivation} Anisotropic flow, in particular elliptic flow, in spite of the relatively small absolute value of the effect, contains very rich physics. In general words, it is very sensitive to the equation of state which governs the evolution of the system created in the nuclear collision. Being such, anisotropic flow provides important information on the state of matter under the extreme conditions of the nuclear collision. The anticipated phase transition to QGP should have a dramatic effect on elliptic flow due to the softening of the equation of state. First it was pointed out in the pioneering work of Ollitrault\cite{olli92}, who suggested elliptic anisotropy as a possible signature of transverse collective flow. Within the hydro-dynamical model Ollitrault analyzed the role of different equations of state and phase transitions on the final anisotropy. Hung and Shuryak~\cite{lshur} suggested scanning with beam energy in order to look for the QCD phase transition. Using their idea of the softest point in the equation of state combined with hydro-dynamical calculations, Rischke~\cite{risc96} predicted a dramatic drop in the elliptic flow signal at the corresponding beam energies (in the original calculations this was at AGS energies). Sorge has shown\cite{lsorge} that the elliptic flow is very sensitive to the pressure at maximum compression, which is the most interesting time in the system evolution. Recent studies~\cite{zhang99} within the parton cascade model yield similar conclusions providing also the relation between the strength of the elliptic flow and parton-parton cross sections. Recently, Sorge also tried~\cite{sorge98} to combine the early system evolution in accordance to a QGP equation of state with a later hadron cascade. He looked at the centrality dependence of the elliptic flow in order to detect QGP production. Summarizing this part, we would conclude that the effect of QGP should be seen in the anisotropic flow dependence on the energy of the colliding nuclei, or in the dependence on the centrality of the collision. If the situation would be such that a QGP is produced only in a small fraction of the collisions than fluctuations in flow would be one of the best observables for this effect. The formation of DCC in nuclear collisions could also result in an event anisotropy. It could be due to the anisotropic shape of the DCC domains~\cite{wang98} or just to local fluctuations in the charged multiplicity, which should result in ``orthogonal'' flow in charged and neutral sectors~\cite{nayak}. The very magnitude of anisotropic flow is sensitive to the degree of equilibration in the system. Note that at present there is no calculation based on the hydro-dynamical picture which accounts for the experimentally observed values of the effect. This could have its origin in the obvious difficulties of hydrodynamic model calculations, but it could also indicate a non-applicability of the picture to nuclear collisions. The cascade models such as RQMD describe the data much better. From this point of view the analysis of elliptic flow in the collision-less and hydrodynamic limits performed in~\cite{heisel} is very interesting. The HBT interferometry performed relative to the event plane~\cite{lvc,wied97,heisel2,heisel,volo_ann98_hbt} becomes also extremely important at this point. Does the system really expand in the reaction plane as prescribed by hydrodynamics? Simultaneous measurements of the anisotropic flow and the two-particle, identical as well as non identical~\cite{volo97}, correlations in principle should answer this question. We must also mention the importance of anisotropic flow measurements to the vast variety of other measurements, which from first look have nothing to do with anisotropic flow. Let us consider high $p_t$ particle production. It could be that the production mechanism (hard parton scattering) is very insensitive to the in-plane expansion, but that the rescattering of high $p_t$ partons is different in the different directions of particle emission due to the anisotropic geometry of the collision zone. This would lead to anisotropy in high $p_t$ particle production and gives another opportunity to study how it develops~\cite{snell_annu99,volo_ww99}. Another example is HBT measurements averaged over all orientations of particle emission. One would think that this does not require reaction plane measurements, but this is not really true. The mixed pair distribution usually used in the correlation function calculation can strongly depend on the relative orientation of the reaction plane of the events used to create the mixed pair. Therefore one should have this information even in the case where the dependence of the HBT parameters on the reaction plane is not studied. \section{Technical Requirements} The study of azimuthal anisotropy of unidentified charged hadrons needs the momenta of the particles but does not have any unusual requirements for calibrations, momentum resolution, acceptance, efficiency, two-track resolution, or two-track efficiency. However, for future analyses it would be good to have particle identification. \section{Directed and Elliptic Flow at RHIC} The anisotropy in the azimuthal distribution of particles is often characterized by $v_1$, $v_2$ and called directed and elliptic flow respectively. This anisotropy, especially $v_2$, plays an important role in high energy nuclear collisions and is expected to be even more important at RHIC energies~\cite{lsorge}. The azimuthal distribution of particles is described by a Fourier expansion~\cite{lvzh} \begin{equation} E\frac{{\mathrm d}^3N}{{\mathrm d}^3p} = \frac{1}{2\pi} \frac{{\mathrm d}^2N}{p_t{\mathrm d}p_t{\mathrm d}y} \left( 1+\sum_{n=1}^{\infty} 2 v_n \cos [n(\phi-\Psi_r)]\right), \end{equation} where $\Psi_r$ is the true reaction plane angle. The reaction plane is defined by the beam direction and the impact parameter vector ${\bf b}$. In a given rapidity ($y$) and $p_t$ interval the coefficients are determined by \begin{equation} v_{n} = \langle \cos [n(\phi-\Psi_r)] \rangle. \end{equation} Similarly this Fourier expansion can be done in coordinate space, where for a given rapidity and $p_t$ interval the coefficients are determined by \begin{equation} r_{n} = \langle \cos [n(\arctan (\frac{y}{x})-\Psi_r)] \rangle \end{equation} where $x,y$ are the particle space coordinates at freeze-out. Of course, these equations only apply to simulations where one knows $\Psi_r$. Comparing the anisotropy coefficients in momentum space ($v_n$) with the anisotropy coefficients in coordinate space ($r_n$) as a function of $p_t$ helps us to understand the space-time evolution of nucleus-nucleus collisions~\cite{lvc,nxu_vspace}. To study this space-time evolution at RHIC, $40\;000$ Au+Au collisions at $\sqrt s$~=~200~$A$GeV have been analyzed using the RQMD v2.4 model. \begin{figure}[t] \centering \mbox{ \mfigure[ {\it Anisotropy coefficients for nucleons and charged pions in RQMD for collisions in the impact parameter range of 5 $\leq b \leq$ 10 fm.} ] {\epsfig{figure=vxrx_all_c.eps,width=.49\textwidth} \label{vxrx_rqmd} } \mfigure[ {\it Anisotropy coefficients for nucleons and charged pions in HIJING for collisions in the impact parameter range of 5 $\leq b \leq$ 10 fm.} ] {\epsfig{figure=vx_all_hijing.eps,width=.49\textwidth} \label{vx_all_hijing} } } \end{figure} Figs.~\ref{vxrx_rqmd}a-d show the first harmonic both in momentum and coordinate space for nucleons and pions. For nucleons at mid-rapidity note the similarity in shape of $v_1$ versus $y$ and r$_1$ versus $y$. Here (Fig. 1a) both the slopes of $v_1$ versus $y$ and r$_1$ versus $y$ show a reversal of sign. This finds an explanation in a picture with strong (positive) space-momentum correlations, taking into account the correlation between nucleon stopping and the original position of the nucleons in the transverse plane~\cite{snell2_annu99}. For pions, the rapidity dependence of $v_1$ is predominantly governed by rescattering on comoving nucleons. Figs.~\ref{vxrx_rqmd}e-h show $v_2$ for nucleons and pions. For both nucleons and pions $v_2$ is positive and is larger for particles with $p_t \ge 1.5$ GeV. Particles acquire a large $p_t$ when they are produced by a hard collision (which should not produce an event anisotropy) or when they have a large number of soft collisions (rescattering). The latter would explain the increase in $v_2$ and it explains why r$_2$ goes from negative for nucleons integrated over all $p_t$ to positive for nucleons with large $p_t$. Collective flow and the coefficients $v_1$ and $v_2$ are usually associated with soft processes. However, the coefficients describe the event anisotropy and are not limited to only soft physics. At RHIC energies hard processes become important. They happen early in the reaction and thus can be used to probe the early stage of the evolution of a dense system. During this time a quark-gluon plasma (QGP) could exist. Associated with hard processes are jets. However, when the transverse energy of the jets becomes smaller it becomes increasingly difficult to resolve them from the ``soft'' particles. These jets with $E_T <$ 5 GeV are usually refered to as mini-jets. At RHIC energies it has been estimated that 50\% of the transverse energy is produced by mini-jets~\cite{jets}. Medium induced radiative energy loss of high $p_t$ partons (jet quenching) could be very different in a hadronic medium and a partonic medium. Recently it was shown that this energy loss per unit distance, $dE/dx$, grows linearly with the total length of the medium~\cite{baieretall}. For non central collisions the hot and dense overlap region has an almond shape. This implies different path lengths and therefore different energy loss for particles moving in the in-plane versus the out-plane direction. To study this anisotropy with respect to the reaction plane~\cite{snell_annu99}, $100\;000$ Au+Au collisions at $\sqrt s$~=~200~$A$GeV have been generated using HIJING~\cite{hijing} v1.35. Figs.~\ref{vx_all_hijing}a-d show $v_1$ and $v_2$ for nucleons and charged pions. The coefficient $v_1$ shows a small negative slope around mid-rapidity for both nucleons and pions and this becomes more pronounced for particles with $p_t \ge 1.5$ GeV. The coefficient $v_2$ is slightly negative over the whole rapidity range for both charged pions and nucleons. For particles with $p_t \ge 1.5$ GeV, $v_2$ becomes more negative especially at forward and backward rapidity. Figs.~\ref{vx_all_hijing}e-f show that without jet quenching the anisotropy coefficients become zero. This indicates that interactions among particles, either quenching or rescattering, are important for producing the anisotropy. \section{Event Plane Resolutions} Within event generators the true reaction plane angle $\Psi_r$ is known. This is not the case experimentally and the reaction plane has to be estimated from the data. This is done using the anisotropy in the azimuthal distribution of particles itself. The estimated reaction plane angle for the $n^{th}$ harmonic is called $\Psi_n$. The magnitude of the anisotropy and the finite number of particles available to determine this event plane leads to a finite resolution. Therefore, the measured $v_n^{obs}$ coefficients with respect to the event plane have to be corrected for this event plane resolution \begin{equation} v_n = \frac{v_{n}^{obs}}{\langle \cos [n(\Psi_n-\Psi_r)] \rangle} \;. \label{vn} \end{equation} However, eq.~\ref{vn} uses the true reaction plane which is not known experimentally. Following Ref.~\cite{meth}, if one constructs the event plane from two random subevents one can relate the resolution of the subevents to the full event plane resolution, \begin{equation} \langle \cos [n(\Psi_{n}-\Psi_r)] \rangle = C \times \sqrt{\langle \cos [n(\Psi_{n}^{a}-\Psi_{n}^{b})] \rangle }, \end{equation} where $C$ is a correction~\cite{meth} for the difference in subevent multiplicity compared to the full event and $\Psi_{n}^{a}, \Psi_{n}^{b}$ are the angles of the event planes determined in the subevents. \begin{figure}[t] \centering \mbox{ \mfigure[ {\it RQMD v2.4 prediction for $v_2$ using $\pi^+ +\pi^-$ within -1.5~$\leq y \leq$~1.5. The multiplicity and event plane resolution are also shown.} ] { \epsfig{figure=resolution_v2_pions_tpc.eps,width=.68\textwidth} \label{restpc} } } \end{figure} To calculate how well the event plane can be determined in STAR, we considered the TPC (-1.5~$\leq y \leq$~1.5) and the FTPCs (2.5~$\leq |y| \leq$~4.). For this the RQMD v2.4 model predictions for Au+Au at $\sqrt s$~=~200~$A$GeV have been used. In Fig.~\ref{restpc}a, $v_2$ for charged pions integrated over the TPC rapidity region is shown versus the impact parameter $b$. Fig.~\ref{restpc}b shows the corresponding multiplicity as a function of $b$. These quantities lead to a resolution for $v_2$, calculated using the true reaction plane, as shown in Fig.~\ref{restpc}c. The resolution for $v_2$ which can be obtained in the STAR TPC using subevents is shown in Fig.~\ref{restpc}d. For $v_2$ charged pions and protons both contribute positively and therefore do not need to be identified. However, the multiplicity of protons at mid-rapidity is small compared to that of pions and, therefore, including protons does not significantly change the resolution. \begin{figure}[t] \centering \mbox{ \mfigure[ {\it RQMD v2.4 prediction for elliptic flow using $\pi^+, \pi^-$ and protons within 2.5~$\leq |y| \leq$~4.0.} ] { \epsfig{figure=resolution_v2_all_ftpc.eps,width=.48\textwidth} \label{resv2ftpc} } \mfigure[ {\it RQMD v2.4 prediction for directed flow using $\pi^+, \pi^-$ and protons within 2.5~$\leq |y| \leq$~4.0.} ] { \epsfig{figure=resolution_v1_all_ftpc_s.eps,width=.48\textwidth} \label{resv1ftpc} } } \end{figure} In Fig.~\ref{resv2ftpc}a, $v_2$ integrated over the FTPC rapidity region is shown versus the impact parameter $b$. For the FTPCs the $\pi^+, \pi^-$ and protons are combined. It was shown in Fig.~\ref{vxrx_rqmd}e that $v_2$ is relatively flat as a function of rapidity and its magnitude is therefore comparable in the FTPC and TPC regions. Fig.~\ref{resv2ftpc}b shows the corresponding multiplicity as a function of $b$ for the combined FTPCs. These quantities lead to a resolution for $v_2$, calculated using the true reaction plane, as shown in Fig.~\ref{resv2ftpc}c. The resolution for $v_2$ which can be obtained in the STAR FTPCs using subevents is shown in Fig.~\ref{resv2ftpc}d. If only one FTPC would be used this resolution would be approximately $\sqrt 2$ smaller. Using $v_2$ the event plane can be determined, however the sign of $v_2$ is not determined relative to ${\bf b}$. This sign could be determined from $v_2$ relative to $\Psi_1$. Fig.~\ref{vxrx_rqmd}c shows that around mid rapidity $v_1$ is maximally 0.5\% which makes $\Psi_1$ extremely hard to measure. From Fig.~\ref{vxrx_rqmd}a and~\ref{vxrx_rqmd}c it is clear that the best region to measure $v_1$ is at forward rapidity. Fig.~\ref{resv1ftpc}a shows $v_1$ integrated over the FTPC rapidity region, versus $b$. As for $v_2$, the $\pi^+, \pi^-$ and protons are combined. This decreases the magnitude of $v_1$ because their signs are opposite but the FTPCs are not able to separate these particles. At large $b$ the magnitude of $v_1$ becomes $\approx$ 1\% and, although this is already hard to measure, also the multiplicity decreases rapidly at large $b$. This leads to negligible resolution for $v_1$ at all values of $b$, which is shown in Fig.~\ref{resv1ftpc}c. \section{Conclusion} We have investigated the feasibility of reconstructing the event plane. Both Fig.~\ref{restpc} and Fig.~\ref{resv2ftpc} show that it is possible to determine the second harmonic event plane and calculate $v_2$ within STAR, assuming the RQMD predictions (multiplicity distribution, magnitude of $v_2$) are correct. For $v_2$ both the TPC or the FTPCs can be used. This would initially provide a cross check and later combining both detectors would increase the resolution. For this study we only need the momenta of the charged hadrons and thus anisotropic flow could be one of the first results from STAR. For future analyses it would be good to have particle identification. Because it is important to study the dependence of $v_2$ as a function of $b$~\cite{sorge98} we would like to have 10 centrality bins, which would be possible with $1\;000\;000$ minimum bias events. \section{Acknowledgments} We would like to thank the other members of the STAR LBNL Soft Hadron Group and in particular the group leader Nu Xu for help with this work.
\section{Introduction}\label{sec-Intro} There are more than 150 different objects enumerated by Catalan numbers; \cite{EC2} contains an extensive list of such combinatorial objects and their properties. Two of the most carefully studied ones are noncrossing partitions and permutations avoiding a 3-letter pattern. A partition $\pi$ of the set $[n] \colon =\{1,2,\cdots, n\}$, having blocks $\beta_1,\beta_2,\cdots ,\beta_k$, is called {\em noncrossing} if there are no four elements $1 \leq a<b<c<d \leq n$ so that $a,c\in \beta_i$ and $b,d \in \beta_j$ for some distinct blocks $\beta_i$ and $\beta_j$. The set of noncrossing partitions of $[n]$ constitutes a lattice under the refinement order (where $\pi < \nu$ if each block of $\nu$ is a union of blocks of $\pi$). An investigation of structural and enumerative properties of this lattice was initiated by Kreweras \cite{kreweras}, and continued by several authors, e.g., \cite{Ed1}, \cite{Ed2}, \cite{EdSi}, \cite{Montenegro}, \cite{NicaSpei}, \cite{SiU}, \cite{St-NCact}. We denote the lattice of noncrossing partitions of $[n]$ as $NC^A_n$, since it is a subposet (indeed, a sub-meet-semilattice) of the intersection lattice associated with the type-A hyperplane arrangement in ${\bf R}^n$ (which consists of the hyperplanes with equations $x_i = x_j$, for $1 \leq i < j \leq n$). For our purposes, recall from \cite{kreweras} that the poset $NC^A_n$ is ranked with rank function ${\rm rk}(\pi) = n - {\rm bk}(\pi)$ (where ${\rm bk}(\pi)$ denotes the number of blocks of the partition $\pi$), rank-symmetric and rank-unimodal with rank sizes given by the Narayana numbers $\left( {1 \over n} {n \choose k} {n \choose {k+1}} \right)_{0 \leq k <n}$. Furthermore, it is self-dual (see \cite{kreweras}, \cite{SiU}) and has the strong Sperner property (see \cite{SiU}; that is, for every $k$, the maximum cardinality of the union of $k$ antichains is the sum of the $k$ largest rank-sizes). A permutation $\sigma = \sigma_1 \sigma_2 \cdots \sigma_n$ of $[n]$, or, in what follows, an $n$-permutation, is called {\em 132-avoiding} if there are no three positions $1 \leq a<b<c \leq n$ so that $\sigma_a < \sigma_c < \sigma_b$. Classes of restricted permutations avoiding other patterns are defined similarly. Such classes of permutations arise naturally in theoretical computer science in connection with sorting problems (e.g., \cite{Knuth}, \cite{Tarjan}), as well as in the context of combinatorics related to geometry (e.g., the theory of Kazhdan-Lusztig polynomials \cite{Brenti-KL} and Schubert varieties \cite{Billey}). The investigation of classes of pattern-avoiding permutations from an enumerative and algorithmic point of view includes \cite{Barc}, \cite{Bona}, \cite{ChowWest}, \cite{DuGiWe}, \cite{NoonanZeil}, \cite{schmidt}, \cite{West}, to name a few. In Section \ref{sec-A} of this paper we introduce the partially ordered set $P^A_n$ whose elements are the 132-avoiding $n$-permutations, ordered by $\sigma < \rho$ if ${\rm Des}(\sigma) \subset {\rm Des}(\rho)$, where ${\rm Des}$ denotes the descent set of a permutation. One can think of $P^A_n$ as a Boolean algebra of rank ${n-1}$ in which each element $S$ is replicated as many times as there are 132-avoiding permutations with $S$ as the descent set. We show that this poset of restricted permutations is an extension of the lattice of noncrossing partitions $NC^A_n$ by exhibiting a natural order-preserving bijection from the dual order $(NC^A_n)^*$ to the poset $P^A_n$. This yields the fact that $P^A_n$ has the same rank-generating function as $NC^A_n$ (implicit in \cite{Si-ncstats}, where the joint distribution of the descent and major index statistics on 132-avoiding permutations is shown to agree with the joint distribution of the block and {\rm rb} statistic on noncrossing partitions). It then follows that $P^A_n$ is rank-unimodal, rank-symmetric and strongly Sperner. We also prove that $P^A_n$ is itself a self-dual poset. We also present type-B analogues of these results. These constitute Section \ref{sec-B} of the paper. The notion of a type-B noncrossing partition of $[n]$ is that first considered by Montenegro \cite{Montenegro}, systematically studied by Reiner \cite{Reiner-NCB}, and further investigated by Hersh \cite{Hersh-NC}. These authors show that the type-B noncrossing partitions of $[n]$ form a lattice, $NC^B_n$, which shares naturally a variety of properties of $NC^A_n$. In particular, $NC^B_n$ is a rank-unimodal, self-dual, strongly Sperner poset. We define a poset $P^B_n$ into which $NC^B_n$ can be embedded via an order-preserving bijection, with properties analogous to those obtained for type-A. The parallel between the type-A and type-B cases includes the fact that the poset $P^B_n$ is defined in terms of pattern-avoiding elements of the hyperoctahedral group (or signed permutations), ordered by containment of the descent set. The relevant pattern restriction is the simultaneous avoidance of the patterns $21$ and ${\overline{2}}\ {\overline{1}}$. This class of restricted signed permutations was considered in \cite{Si-Bstats}, where B-analogues are proposed for type-A results in \cite{Si-ncstats} concerning combinatorial statistics for noncrossing partitions and restricted permutations. In brief, a class of partitions and one of permutations are equinumerous, and further, the count of the partitions by number of blocks agrees with the count of permutations by number of descents. A similar situation arises for certain type-B analogues of these objects. Our results show that these enumerative relations are manifestations of structural relations between partial orders which can be defined naturally on the objects under consideration. We also discuss posets $Q^A_n$ and $Q^B_n$ of restricted permutations and signed permutations ordered by containment of their sets of excedences. The final section of the paper consists of remarks and problems for further investigation. \section{The type-A case}\label{sec-A} \subsection{A bijection and its properties} It is not difficult to find a bijection from the set of noncrossing partitions of $[n]$ onto that of 132-avoiding $n$-permutations. Here we exhibit and analyze the structure of such a bijection, $f$, which will serve as the main tool in proving the results of this section. To avoid confusion, integers belonging to a partition will be called {\em elements}, while integers belonging to a permutation will be called {\em entries}. An $n$-permutation will always be written in the one-line notation, $p=p_1p_2\cdots p_n$, with $p_i = p(i)$ denoting its $i$th entry. Let $\pi \in NC^A_n$. We construct the 132-avoiding permutation $p=f(\pi)$ corresponding to it as follows. Let $k$ be the largest element of $\pi$ which is in the same block of $\pi$ as 1. Put the entry $n$ of $p$ in the $k$th position, i.e., set $p_k=n$. As $p$ is to be 132-avoiding, this implies that the entries larger than $n-k$ are on the left of $n$ in $p$, and the entries smaller than or equal to $n-k$ are on the right of $n$. Delete $k$ from $\pi$ and apply this procedure recursively, with obvious minor adjustments, to the restrictions of $\pi$ to the sets $\{ 1, \dots, k-1 \}$ and $\{ k+1, \dots , n \}$, which are also noncrossing partitions. Namely, if $j$ is the largest element in the same block as $k+1$, we set $p_{j} = n-k$, so that the restriction $\pi_1$ of $\pi$ to $\{k+1,k+2,\dots ,n\}$ yields a 132-avoiding permutation of $\{1,2,\dots, n-k\}$ placed on the right of $n$ in $p = f(\pi)$. Similarly, if in the restriction $\pi_2$ of $\pi$ to the set $\{ 1, 2, \dots, k-1 \}$ the largest element in the same block as 1 is equal to $j$, we set $p_j = n-1$. Thus, recursively, $\pi_2$ yields a 132-avoiding permutation which we realize on the set $\{ n-k+1, n-k+2, \dots, n-1 \}$ and we place it to the left of $n$ in $p = f(\pi)$. In other words, with a slight abuse of notation, $f(\pi)$ is the concatenation of $f(\pi_2)$, $n$, and $f(\pi_1)$, where $f(\pi_2)$ permutes the set $\{n-k+1,n-k+2,\cdots ,n-1\}$ and $f(\pi_1)$ permutes the set $[n-k]$. To see that this is a bijection note that we can recover the maximum of the block containing the element 1 from the position of the entry $n$ in $p$, and then proceed recursively. \begin{example} {\em If $\pi=(\{1,4,6\}, \{2,3\}, \{5\}, \{7,8\})$, then $f(\pi)=64573812$. }\end{example} \begin{example} {\em If $p=(\{1,2,\cdots ,n\})$, then $f(p)=12\cdots n $. }\end{example} \begin{example} {\em If $p=(\{1\},\{2\},\cdots ,\{n\})$, then $f(p)=n\cdots 21 $. }\end{example} The following definition is widely used in the literature. \begin{definition} Let $p=p_1p_2\cdots p_n$ be an $n$-permutation. We say that $i \in [n-1]$ is a {\em descent} of $p$ if $p_i>p_{i+1}$. The set of all descents of $p$ is called the {\em descent set} of $p$ and is denoted ${\rm Des}(p)$. \end{definition} Now we are in a position to define the poset $P^A_n$ of 132-avoiding permutations we want to study. \begin{definition} Let $p$ and $q$ be two 132-avoiding $n$-permutations. We say that $p<q$ in $P^A_n$ if ${\rm Des}(p)\subset {\rm Des}(q)$. \end{definition} Clearly, $P^A_n$ is a poset as inclusion is transitive. The Hasse diagram of $P^A_4$ is shown in Figure 1. \begin{figure}[ht] \begin{center} \epsfig{file=poset4.eps} \caption{The Hasse diagram of $P^A_4$ } \label{hasse} \end{center} \end{figure} \begin{observation}\label{obs-lrminA} In a 132-avoiding permutation, $i$ is a descent if and only if $p_{i+1}$ is smaller than every entry on its left. Such an element is called a {\em left-to-right minimum}. So $p<q$ in $P^A_n$ if and only if the set of positions of left-to-right minima in $p$ is a proper subset of the set of positions of left-to-right minima in $q$. \end{observation} The following proposition describes the relation between the blocks of $\pi \in NC^A_n$ and the descent set of the 132-avoiding permutation $f(\pi)$. \begin{proposition} \label{prop-block} The bijection $f$ has the following property: Let $i \ge 1$. Then $i\in {\rm Des}(f(\pi))$ if and only if $i+1$ is the smallest element of its block in $\pi \in NC^A_n$. \end{proposition} \begin{proof} For $n=1$ and $n=2$ the statement is clearly true and we use induction on $n$. Suppose we know the statement for all positive integers smaller than $n$. Then we distinguish two cases: \begin{enumerate} \item If 1 and $n$ are in the same block of $\pi$, then the construction of $f(\pi)$ starts by putting the entry $n$ in the last slot of $f(\pi)$, then deleting the element $n$ from $\pi$. This does not alter either the set of minimum elements of the blocks nor the set of descents. Therefore, this case reduces to the general case for $n-1$, and is settled by the inductive hypothesis. \item If the largest element $k$ of the block containing 1 is smaller than $n$, then as we have seen above, $f(\pi)$ is the concatenation of $f(\pi_2), n, f(\pi_1)$, and $f(\pi_1)$ is not empty. Clearly, by the definition of $f(\pi)$, $k \in {\rm Des}(f(\pi))$ and the element $k+1$ is the minimum of its block. From this and the inductive hypothesis applied to $f(\pi_1)$ and $f(\pi_2)$, the proof follows. \end{enumerate} \end{proof} \subsection{Properties of $P^A_n$} Proposition \ref{prop-block} implies that $P^A_n$ is isomorphic to the order on noncrossing partitions in which $\pi<\pi'$ if the set of minima of the blocks of $\pi'$ is contained in the set of minima of the blocks of $\pi$. This yields the first result of this section. \begin{theorem}\label{thm-embPA} The lattice of noncrossing partitions $NC^A_n$ is a subposet of $P^A_n$. \end{theorem} \begin{proof} We show that our bijection $f$ is an order-reversing map $NC^A_n \to P^A_n$. The conclusion then follows from the self-duality of the lattice of noncrossing partitions. Suppose $\pi < \tau$ in $NC^A_n$. This means $\pi$ is a finer partition than $\tau$, so every element which is the minimum of its block in $\tau$ is also the minimum of its block in $\pi$. By Proposition \ref{prop-block} this implies ${\rm Des}(f(\tau))\subset {\rm Des}(f(\pi))$, so $f(\pi) > f(\tau)$ in $P^A_n$. \end{proof} Clearly, $P^A_n$ is a ranked poset (with rank function ${\rm rk}_{P^A_n}(p) = \# {\rm Des}(p)$), and we have ${\rm rk}_{NC^A_n}(\pi) = n-1- {\rm rk}_{P^A_n}(f(\pi))$. \begin{corollary}\label{cor-symA} The poset $P^A_n$ is rank-symmetric, rank-unimodal and strongly Sperner, and its rank generating function is equal to that of $NC^A_n$. \end{corollary} \begin{proof} The properties of the rank sizes of $P^A_n$ are immediate consequences of Proposition \ref{prop-block} and the corresponding properties known to hold for $NC^A_n$. Moreover, every antichain of $P^A_n$ is, via the bijection $f$, an antichain of $NC^A_n$, and the strong Sperner property of $P^A_n$ follows from the strong Sperner property of $NC^A_n$. \end{proof} We now turn to showing that $P^A_n$ is self-dual, based on the next lemma. For $S \subseteq [n-1]$, let $Perm_n(S)$ denote the number of 132-avoiding $n$-permutations with descent set $S$. \begin{lemma}\label{lemma-rev-compl} Let $S$ be any subset of $[n-1]$ and let $\alpha (S)$ denote its ``reverse complement,'' that is, $i\in \alpha(S) \Longleftrightarrow n-i \notin S$. Then $Perm_n(S)=Perm_n(\alpha(S))$. \end{lemma} \begin{proof} We use induction on $n$. For $n=1,2,3$ the statement is true. Now suppose we know it for all positive integers smaller than $n$. Denote by $t$ the smallest element of $S$, and let $p$ be a 132-avoiding $n$-permutation whose descent set is $S$. \begin{enumerate} \item Suppose that $t>1$. Then we have $p_1<p_2<\cdots <p_t$ and, because $p$ avoids the pattern 132, the values of $p_1,p_2,\dots,p_t$ are {\em consecutive} integers. So, for given values of $p_1$ and $t$, we have only one choice for $p_2,p_3,\dots ,p_t$. This implies \begin{equation} \label{1a} Perm_n(S)=Perm_{n-(t-1)}(S-(t-1)),\end{equation} where $S-(t-1)$ is the set obtained from $S$ by subtracting $t-1$ from each of its elements. On the other hand, we have $n-t+1,n-t+2,\cdots ,n-1 \in \alpha(S)$, meaning that in any permutation $q$ counted by $Perm_n(\alpha(S))$ the chain of inequalities $q_{n-t+1}>q_{n-t+2}>\cdots >q_n$ holds. To avoid forming a 132-pattern in $q$, we must have $(q_{n-t+2},\dots ,q_n)=(t-1,t-2,\dots 1)$. Therefore, \begin{equation}\label{1b} Perm_n(\alpha(S))=Perm_{n-(t-1)}(\alpha(S)|n-(t-1)) \end{equation} where $\alpha(S)|n-(t-1)$ denotes the set obtained from $\alpha(S)$ by removing its last $t-1$ elements. Clearly, $Perm_{n-(t-1)}(S-(t-1))=Perm_{n-(t-1)}(\alpha(S)|n-(t-1))$ by the induction hypothesis, so equations (\ref{1a}) and (\ref{1b}) imply $Perm_n(S)=Perm_n(\alpha(S))$. \item If $t=1$, but $S\neq [n-1]$, then let $u$ be the smallest index which is {\em not} in $S$. Then again, to avoid forming a 132-pattern, the value of $p_u$ must be the smallest positive integer $a$ which is larger than $p_{u-1}$ and is not equal to any $p_i$ for $i\leq u-1$. So again, we have only one choice for $p_u$. On the other hand, the largest index in $\alpha(S)$ will be $n-u$. Therefore, in permutations $q$ counted by $Perm_n(\alpha(S))$, we must have $q_{n-u}=1$ as Observation 1 implies that $q_{n-u}$ must be the rightmost left-to-right minimum in such permutations, and that is always the entry 1. In order to use this information to reduce our permutations in size, we define $S'\subset [n-2]$ as follows: $i\in S'$ if and only if either $i<u$ and then, by the definition of $u$, $i\in S$, or $i>u$ and $i+1\in S$. In other words, we decrease elements larger than $u$ by 1; intuitively, we remove $u$ from $[n-1]$, and translate the interval on its right one notch to the left. If we now take $\alpha(S')$, that will consist of entries $j$ so that $j<n-u$ and $(n-1)-(j-1)=n-j\notin S$. So in other words, we simply remove $n-u$ from $[n-1]$ (there has been nothing on the right of $n-u$ in $\alpha(S)$ to translate). Note that the size of $\alpha(S)$ decreases with this operation as $n-u\in \alpha(S)$. As we have seen in the previous paragraph, we had only one choice for $p_u$ and $p_{n-u}$, so removing them this way does not change the number of permutations with a given descent set. Thus we have $Perm_n(S)=Perm_{n-1}(S')$, and also $Perm_n(\alpha(S))= Perm_{n-1}(\alpha(S'))$. By induction hypothesis, the right hand sides of thes two equations agree, and therefore the left hand sides must agree, too. \begin{example} {\em If $n=8$ and $S=\{1,6\}$, and so $\alpha(S)= \{1,3,4,5,6\}$, then $u=2$, $n-u=6$, and indeed, $S'=\{1,5\}$ and $\alpha(S')= \{1,3,4,5\}$. } \end{example} \item Finally, if $S=[n-1]$, then the statement is trivially true as $Perm_n(S)=Perm_n(\alpha(S))=1.$ \end{enumerate} So we have seen that $Perm_n(S)=Perm_n(\alpha(S))$ in all cases. \end{proof} It is now easy to verify that the reverse complementation of the descent set can be used to construct an anti-automorphism of $P^A_n$. \begin{theorem} The poset $P^A_n$ is self-dual. \end{theorem} \begin{proof} It is clear that, in $P^A_n$, permutations which have the same descent set will cover the same elements and they will be covered by the same elements. The permutations with a prescribed descent set $S$ form an orbit of $Aut(P^A_n)$ and they can be permuted among themselves arbitrarily by elements of $Aut(P^A_n)$. On the other hand, Lemma \ref{lemma-rev-compl} shows that the orbits corresponding to $S \subseteq [n-1]$ and to its reverse-complement $\alpha(S)$ are equinumerous. Hence, a map $P^A_n \to P^A_n$ which establishes a bijection between $\{ p \in P^A_n \ \colon \ {\rm Des}(p) = S \}$ and $\{ q \in P^A_n \ \colon \ {\rm Des}(q) = \alpha(S) \}$ for each $S \subseteq [n-1]$ provides an order-reversing bijection on $P^A_n$. \end{proof} \subsection{A poset derived from excedences}\label{subsec-excA} It is shown in \cite{Si-ncstats} that the joint distribution of the excedence and Denert statistics on 321-avoiding permutations agrees with the joint distribution of the block and ${\rm rb}$ statistics on noncrossing partitions. This suggests the definition of the poset $Q^A_n$ consisting of the 321-avoiding $n$-permutations ordered by containment of the set of excedences, and invites the question of how $Q^A_n$ compares with the poset $P^A_n$. A permutation $\sigma$ has an {\em excedence} at $i$ if $\sigma(i) > i$. For example, the {\em excedence set} of $\sigma = 3 2 5 1 4$ is ${\rm Exc}(\sigma) = \{ 1, 3 \}$. Let ${\rm exc}(\sigma)$ denote the number of excedences of $\sigma$. Following \cite{Si-ncstats}, there is a bijection $\theta$ from $NC^A_n $ to 321-avoiding $n$-permutations such that ${\rm exc}(\theta(\pi)) = {\rm bk}( \pi ) - 1$. Namely, if the set of minima of the blocks of $\pi \in NC^A_n$, omitting the block containing 1, is $\{ f_2 < \cdots < f_k \}$ and the set of maxima of the blocks, again, omitting the block containing 1, is $\{ l_2 < \cdots < l_k \}$, then let $\theta(\pi)$ be the permutation whose value at $f_i - 1$ is $l_{i} $ for $i = 2, 3, \dots, k$, and whose other values constitute an increasing subsequence in the remaining positions. For instance, if $\pi = \{ 1, 5, 7 \} \{ 2 \} \{ 3, 4 \} \{ 6 \} \{ 8, 10 \} \{ 9 \} \in NC^A_{10}$, then we have $( f_2, \dots, f_6) = ( 2, 3, 6, 8, 9)$ and $( l_2, \dots, l_6) = (2, 4, 6, 9, 10)$, and we obtain $\theta(\pi) = 2\ 4\ 1\ 3\ 6\ 5\ 9\ 10\ 7\ 8$. Recall from \cite{Si-ncstats} that the set of excedences of $\theta(\pi)$ is precisely $\{ f_2 -1, f_3 -1, \dots, f_k -1 \}$. Similarly to the case of descents discussed for 132-avoiding permutations, a covering relation $\pi < \pi'$ in $NC^A_n$ corresponds to the deletion of an excedence: ${\rm Exc}(\theta(\pi')) = {\rm Exc}(\theta(\pi)) - \{ i \}$, for a suitable $i \in {\rm Exc}(\theta(\pi))$. Hence, taking advantage of the self-duality of $NC^A_n$, one can establish directly that the poset $Q^A_n$ enjoys the same properties as $P^A_n$: There is an embedding of $NC^A_n$ into the poset $Q^A_n$ of 321-avoiding $n$-permutations ordered by containment of the set of excedences; the embedding is rank-preserving and $Q^A_n$ is a strongly Sperner poset. The fact that the posets $P^A_n$ and $Q^A_n$ have strongly similar properties is not accidental. \begin{proposition}\label{prop-isoA} The posets $P^A_n$ and $Q^A_n$ are isomorphic. \end{proposition} \begin{proof} For each $S \subseteq [n-1]$, let $E^{321}_n(S)$ be the set of 321-avoiding $n$-permutations with excedence set $S\subseteq [n-1]$. Let also $D^{132}_n(\alpha(S))$ be the set of 132-avoiding $n$-permutations with descent set equal to $\alpha(S)$, the reverse-complement of $S$. Thus, in the notation of the previous subsection, the cardinality of $D^{132}_n(\alpha(S))$ is $Perm_n(\alpha(S))$. We construct a bijection $s \colon E^{321}_n(S) \to D^{132}_n(\alpha(S))$ (illustrated by example \ref{eg-sbij}). If $p\in E^{321}_n(S)$, then, as seen earlier in the definition of $\theta$, the entries $p_j$ with $j\notin S$ form an increasing subsequence. This, and the definition of excedence imply that $p_j$ is a {\em right-to-left minimum} (that is, smaller than all entries on its right) if and only if $j\notin {\rm Exc}(p)=S$. Now let $p'= p_n p_{n-1} \cdots p_1$ be the reverse of $p$. Then $p'$ is a 123-avoiding permutation having a left-to-right minimum at position $i\leq n$ exactly if $n+1-i \notin S$. There is exactly one 132-avoiding permutation $p''$ which has this same set of left-to-right minima at these same positions \cite{schmidt}. Namely, $p''$ is obtained by keeping the left-to-right minima of $p'$ fixed, and successively placing in the remaining positions, from left to right, the smallest available element which does not alter the left-to-right minima. We set $s(p)=p''$. Observation 1 then tells us that $i\in {\rm Des} (p'')$ if and only if $n-i\notin S$, in other words, when $i\in \alpha(S)$, and so $p''$ belongs indeed to $D^{132}_n(\alpha(S))$. It is easy to see that $s$ is invertible. Clearly, $p'$ can be recovered from $p''$ as the only 123-avoiding permutation with the same values and positions of its left-to-right minima as $p''$. (All entries which are not left-to-right minima are to be written in decreasing order). Then $p$ can be recovered as the reverse of $p'$. The bijections $s \colon E^{321}_n(S) \to D^{132}_n(\alpha(S))$ for all the choices of $S \subseteq [n-1]$ produce an order-reversing bijection from $Q^A_n$ to $P^A_n$. But $P^A_n$ is self-dual, so the proof is complete. \end{proof} \begin{example}\label{eg-sbij} {\em Take $p=2\ 4\ 1\ 6\ 3\ 5\ 9\ 10\ 7\ 8 \in E^{321}_{10}(S)$ for $S= \{ 1, 2, 4, 7, 8 \}$. Then its reversal $p'=8\ 7\ 10\ 9\ 5\ 3\ 6\ 1\ 4\ 2$ has left-to-right minima 8, 7, 5, 3, 1 in positions 1, 2, 5, 6, 8. We obtain $s(p) = p''= 8\ 7\ 9\ 10\ 5\ 3\ 4\ 1\ 2\ 6$, a permutation in $D^{132}_{10}( \{ 1, 4, 5, 7 \})$. } \end{example} \section{The type-B case}\label{sec-B} \subsection{The type-B noncrossing partitions}\label{subsec-NCB} The hyperplane arrangement of the root system of type $B_n$ consists of the hyperplanes with equations $x_i = \pm x_j$ for $1 \leq i < j \leq n$ and the coordinate hyperplanes $x_i = 0$, for $1 \leq i \leq n$. The subspaces of ${\bf R}^n$ arising as intersections of hyperplanes from among these can be encoded by partitions of $\{ 1, 2, \dots, n , {\overline{1}}, {\overline{2}}, \dots, {\overline{n}} \}$ satisfying the following properties: i) if $\{ a_1, \dots, a_k \}$ is a block, then $\{ {\overline{a_1}}, \dots , {\overline{a_k}} \}$ is also a block, where the bar operation is an involution; and ii) there is at most one block, called {\em the zero-block}, which is invariant under the bar operation. The collection of such partitions are the {\em type-B partitions of $[n]$}. If $1, 2, \dots, n, {\overline{1}}, {\overline{2}}, \dots, {\overline{n}}$ are placed around a circle, clockwise in this order, and if cyclically successive elements of the same block are joined by chords drawn inside the circle, then, following \cite{Reiner-NCB}, the class of {\em type-B noncrossing partitions}, denoted $NC^B_n$, is the class of type-B partitions of $[n]$ which admit a circular diagram with no crossing chords. Alternatively, a type-B partition is noncrossing if there are no four elements $a, b, c, d$ in clockwise order around the circle, so that $a,c$ lie in one block and $b,d$ lie in another block of the partition. The total number of type-B noncrossing partitions of $[n]$ is ${ {2n} \choose n}$ (see \cite{Reiner-NCB}). As in the case of type A, the refinement order on type-B partitions yields a geometric lattice (in fact, isomorphic to a Dowling lattice with an order-2 group), and the noncrossing partitions constitute a sub-meet-semilattice as well as a lattice in its own right. As a poset under the refinement order, $NC^B_n$ is ranked, with ${\rm rk}(\pi) = n - \# ($ of pairs of non-zero blocks $)$. For example, $\pi = \{ 1, {\overline{3}}, {\overline{5}} \}, \{ {\overline{1}}, 3, 5 \}, \{ 4 \}, \{ {\overline{4}} \}, \{ 2, {\overline{2}} \}$ is an element of $NC^B_5$ having 2 pairs of non-zero blocks and its rank is equal to 3. The rank-sizes in $NC^B_n$ are given by $\left( {n \choose k}^2 \right)_{0 \leq k \leq n}$ (see \cite{Reiner-NCB}). The numerous properties of $NC^A_n$ which also hold for $NC^B_n$ (as shown in \cite{Reiner-NCB}, \cite{Hersh-NC}), establish the latter as a natural B-analogue. In particular, $NC^B_n$ is a self-dual, rank-unimodal, strongly Sperner poset, analogously to the properties of $NC^A_n$ of concern in Section \ref{sec-A}. We now turn to a type-B counterpart of the restricted permutations considered in the preceding section. \subsection{A class of pattern-avoiding signed permutations}\label{subsec-Brestr} We will view the elements of the hyperoctahedral group $B_n$ as signed permutations written as words of the form $b = b_1 b_2 \dots b_n$ in which each of the symbols $1, 2, \dots, n$ appears, and may or may not be barred. Thus, the cardinality of $B_n$ is $n! 2^n$. To find a B-analogue of the poset $P^A_n$, we need a subset of $B_n$ whose cardinality is $\# NC^B_n = { {2n} \choose n }$, which is characterized via pattern-avoidance, and over which the distribution of the descent statistic agrees with the distribution across ranks of the type-B noncrossing partitions of $[n]$. Such a class of signed permutations is $B_n(12, {\overline{2}}\ {\overline{1}})$ which appears in \cite{Si-Bstats}. We include its description for the reader's convenience. Consider the elements of $B_n$ which avoid simultaneously the patterns $21$ and ${\overline{2}}\ {\overline{1}}$. That is, the set of elements $b = b_1 b_2 \cdots b_n \in B_n$ such that there are no indices $1 \leq i < j \leq n$ for which i) either both $b_i, b_j$ are barred, or neither is barred, and ii) $|b_i| > |b_j|$ (the absolute value of a symbol means $|a| = a$ if $a$ is not barred, and $|a| = {\overline{a}}$ if $a$ is barred; effectively, the absolute value removes the bar from a barred symbol). The following is immediate: a $(21, {\overline{2}}\ {\overline{1}})$-avoiding permutation in $B_n$ is a shuffle of an increasingly ordered subset $L$ of $[n]$ whose elements we then bar, with its increasingly ordered complement in $[n]$. For example, $b = {\overline{ 2}} 1 3 {\overline{5}} 4 {\overline{ 6}} 7$ is one of ${7 \choose 3}$ elements of $B_7(21, {\overline{2}}\ {\overline{1}})$ associated with the subset $L = \{ 2, 5, 6 \} \subseteq [7]$. Obviously, summing over the choices of $L$ of cardinality ranging from zero to $n$ and over the shuffles, it follows that \begin{equation}\label{eq-Brestr-card} \# B_n(21, {\overline{2}}\ {\overline{1}}) = \sum_{k=0}^n { {n \choose k}^2 } = { {2n} \choose n} = \# NC^B_n, \end{equation} as desired. Furthermore, the distribution of descents over $B_n(21, {\overline{2}}\ {\overline{1}})$ is as desired. We say that $b = b_1 b_2 \dots b_n \in B_n$ has a {\em descent} at $i$, for $1 \leq i \leq n-1$, if $b_i > b_{i+1}$ with respect to the total ordering $1 < 2 < \cdots < n < {\overline{n}} < \cdots < {\overline{2}} < {\overline{1}}$, and that it has a descent at $n$ if $b_n$ is barred. As usual, the {\em descent set} of $b$, denoted ${\rm Des}(b)$, is the set of all $i \in [n]$ such that $b$ has a descent at $i$. For example, for $b = {\overline{ 2}} 1 {\overline{3}} {\overline{5}} 4 7 {\overline{ 6}}$ we have ${\rm Des}(b) = \{ 1, 3, 4, 7 \}$. It is then transparent that if $b \in B_n(21, {\overline{2}}\ {\overline{1}})$, then its descent set is precisely the set of positions occupied by barred symbols. In conclusion, \begin{observation}\label{obs-bij-B2121} For an element $b$ of the hyperoctahedral group $B_n$, let $L(b)$ denote the set of symbols which are barred in $b$, and ${\rm Des}(b)$ denote the descent set of $b$. Then the map $b \mapsto (L(b), {\rm Des}(b))$ gives a bijection between the class of restricted signed permutations $B_n(21, {\overline{2}}\ {\overline{1}})$ and ordered pairs of subsets of $[n]$ of equal cardinality. \end{observation} \subsection{The poset $P^B_n$}\label{subsec-PBn} As the B-analogue of the poset of 132-avoiding permutations $P^A_n$ of the preceding section, we consider the poset $P^B_n$ consisting of the $(21, {\overline{2}}\ {\overline{1}})$-avoiding elements of the hyperoctahedral group $B_n$, with the order relation given by $b < b'$ if and only if ${\rm Des}(b) \subset {\rm Des}(b')$. Based on the preceding discussion and an encoding of type-B noncrossing partitions appearing in \cite{Reiner-NCB}, one readily obtains the properties of $P^B_n$ which parallel those of $P^A_n$. \begin{theorem} The poset $P^B_n$ of $(12,{\overline{2}}\ {\overline{1}})$-avoiding elements of the hyperoctahedral group $B_n$, ordered by containment of the descent set, is an extension of the refinement order on the type-B noncrossing partition lattice $NC^B_n$. The poset $P^B_n$ has the same rank-generating-function as $NC^B_n$, therefore it is rank-symmetric and rank-unimodal, and it is a self-dual and strongly Sperner poset. \end{theorem} \begin{proof} It is immediate from its definition and Observation \ref{obs-bij-B2121} that $P^B_n$ is a ranked poset (namely, ${\rm rk}(b) = \# {\rm Des}(b)$) and has rank-sizes given by $\left( {n \choose k}^2 \right)_{0 \leq k \leq n}$, equal to the rank-sizes in $NC^B_n$. Also, $P^B_n$ is a self-dual poset: clearly, if $b'$ is the $(21, {\overline{2}}\ {\overline{1}})$-avoiding signed permutation which corresponds to the pair $( [n] - L(b), [n] - {\rm Des}(b) )$, then the mapping $b \leftrightarrow b'$ is an order-reversing involution on $P^B_n$. Toward checking that there is an order-preserving bijection from $NC^B_n$ to $P^B_n$, we first recall a fact from \cite{Reiner-NCB}: every partition $\pi \in NC^B_n$ can be encoded by a pair $(L(\pi), R(\pi))$ of subsets of $[n]$ whose cardinality is the number of pairs of non-zero blocks of $\pi$. Informally, these sets consist of the Left and Right delimiters of non-zero blocks when the elements are read in clockwise order (in the circular diagram of $\pi$). More precisely, if $n=0$ or if $\pi$ has only a zero-block, we set $L=R=\emptyset$. Otherwise, $\pi \in NC^B_n$ has some non-zero block consisting of cyclically consecutive elements in its diagram. If such a block consists of $j_1, j_2, \dots, j_k$ in clockwise order, then $|j_1|$ belongs to $L(\pi)$ and $|j_k|$ belongs to $R(\pi)$. By deleting this block and its image under barring, a type-B noncrossing partition of $[n-k]$ is obtained and the construction of the sets $L(\pi)$ and $R(\pi)$ is completed by repeating this process as long as non-zero blocks arise. For instance, if $\pi = \{ 1, {\overline{3}}, {\overline{5}} \}, \{ {\overline{1}}, 3, 5 \}, \{ 4 \}, \{ {\overline{4}} \}, \{ 2, {\overline{2}} \}$, then $L(\pi) = \{ 3, 4 \}$ and $R(\pi) = \{ 1, 4 \}$. Now suppose that $\pi < \pi'$ in $NC^B_n$, and that this is a covering relation (i.e., ${\rm rk}(\pi') = {\rm rk}(\pi) + 1$). Then there exist $l \in L(\pi)$ and $r \in R(\pi)$ such that $L(\pi') = L(\pi) - \{ l \}$ and $R(\pi') = R(\pi) - \{ r \}$, as a result of the merging of blocks entailed by the covering relation. Thus it is clear that if $\pi \in NC^B_n$ is mapped to the signed permutation $b \in P^B_n$ with the property that $(L(b), {\rm Des}(b)) = (L(\pi), R(\pi))$, then one obtains an order-reversing embedding of $NC^B_n$ into $P^B_n$. Combining this with the self-duality of $P^B_n$ we obtain the desired embedding of $NC^B_n$ into $P^B_n$. Finally, the strong Sperner property of $P^B_n$ follows as in type A, from the strong Sperner property of $NC^B_n$ (see \cite{Reiner-NCB}) and the rank-preserving embedding of $NC^B_n$ into $P^B_n$. \end{proof} \subsection{A poset based on type-B excedences}\label{subsec-excB} As in the type-A case, there is a self-dual poset of $\# NC^B_n$ restricted signed permutations ordered by containment of the set of excedences. In fact, there is more than one definition of the excedence statistic in the literature, in the case of the hyperoctahedral group. We briefly mention two possibilities considered in \cite{Steing}. Given a signed permutation $b$, let $k$ be the number of symbols which are {\em not} barred in $b$. We associate to $b$ an $(n+1)$-permutation $\sigma(b)$ by setting $\sigma(b)_{n+1} = k+1$ and, for $1 \leq i \leq n$, letting $\sigma(b)_i = j$ if $b_i$ is the $j$th smallest among the symbols $b_1, b_2, \dots, b_n, n+1$ with respect to the linear ordering $1 < 2 < \cdots < n < n+1 < {\overline{1}} < {\overline{2}} < \cdots < {\overline{n}}$. For example, if $b = 1\ {\overline{3}}\ 2\ 4\ 5\ {\overline{6}}\ {\overline{8}}\ 7$, then $\sigma(b) = 1\ 7\ 2\ 3\ 4\ 8\ 9\ 5\ 6$. Now, the excedence set of $b$ is defined to be that of $\sigma(b)$. It turns out \cite{Gal-pers} that for $b \in B_n(21, {\overline{2}} \ {\overline{1}})$ this definition makes the excedence set coincide with the descent set for each $b$. Therefore, this leads to the poset $P^B_n$ again. An alternative definition for excedences of ``indexed permutations'' appears in \cite{Steing}. Specialized to the hyperoctahedral group it is the following. \begin{definition}\label{def-Bexc2} If $b \in B_n$, its {\em excedence set} is the union of the sets $S(b)$ and $F(b)$, where $S(b)$ is the set of excedences computed in the symmetric group for the permutation $|b_1| |b_2| \dots |b_n|$ obtained by removing all bars from the symbols in $b$, and $F(b) = \{ i \ \colon \ b_i = {\overline{i}} \}$, the set of barred fixed points of $b$. \end{definition} Thus, for $b = 1\ {\overline{3}}\ 2\ 4\ 5\ {\overline{6}}\ {\overline{8}}\ 7$ we obtain excedences at $\{ 2, 6, 7 \}$ by either of the two definitions. But $b = {\overline{1}} 3 {\overline{2}}$ has excedences at $\{ 1, 3 \} $ by the first definition (based on $\sigma(b) = 3\ 1\ 4\ 2$), and $\{ 1, 2 \}$ if the second definition is adopted. For the remainder of this section, we work with the notion of excedence as in Definition \ref{def-Bexc2}. \begin{proposition}\label{prop-QBn-sdual} Let $Q^B_n$ denote the poset of $(21, {\overline{2}}\ {\overline{1}})$-avoiding signed permutations in $B_n$, ordered by containment of their excedence set. The poset $Q^B_n$ is self-dual. \end{proposition} \begin{proof} Let $b\in B_n$ and $b'$ be the reverse of $b$. Let $b''$ be the ``barred complement'' of $b'$, that is, $|b''_i| = n+1- |b'_i|$, and $b''_i$ is barred if and only if $b'_i$ is not barred. Then it is straightforward to verify that $i\in S(p'')\cup F(p'')$ if and only if $i\notin S(p)\cup F(p)$. Therefore, the reverse complement operation reverses the inclusion of excedence sets for signed permutations. (Thus, the entire hyperoctahedral group $B_n$ ordered by containment of the excedence set is a self-dual poset.) But, clearly, this involution preserves the $(21, {\overline{2}}\ {\overline{1}})$-avoidance property, and thus $Q^B_n$ is self-dual. \end{proof} By \cite{Steing}, the rank generating function of $Q^B_n$ is equal to that of $P^B_n$. Therefore it is natural to ask whether the posets $P^B_n$ and $Q^B_n$ are isomorphic, just as their type-A counterparts are (Proposition \ref{prop-isoA}). The answer in this case is negative. Indeed, if $n=3$ it is straightforward to verify that all atoms of $P^B_3$ are covered by six elements, while the atom ${\overline{1}}\ 2\ 3$ of $Q^B_3$ is covered by seven elements (namely, ${\overline{1}}\ {\overline{2}}\ 3, {\overline{1}} \ 2 \ {\overline{3}}, {\overline{1}} \ 3 \ {\overline{2}}, 2 \ 3 \ {\overline{1}}, {\overline{2}} \ {\overline{3}} \ 1, 2 \ {\overline{1}} \ {\overline{3}}$, and ${\overline{2}} \ 1 \ {\overline{3}}$). \section{Remarks and questions for further investigation}\label{sec-Further} \begin{enumerate} \item{ {\em Is $NC^B_n$ a subposet of $Q^B_n$?} We do not know whether the lattice of type-B noncrossing partitions can be embedded in the poset $Q^B_n$ of $(21, {\overline{2}}\ {\overline{1}})$-avoiding signed permutations ordered by their excedence set of definition \ref{def-Bexc2}. } \item{ {\em Self-duality of $NC^A_n$ and $NC^B_n$ extending to self-duality for $P^A_n$ and $P^B_n$.} We have seen that {\em each} of the posets $NC^A_n$ and $P^A_n$ is self-dual and that $NC^A_n$ is a subposet of $P^A_n$. The same is true for the pair $NC^B_n$, $P^B_n$. Both for type A and for type B one can exhibit an order-reversing involution on the larger poset which restricts to an order-reversing involution on the smaller one. We first construct such an involution $g$ for $NC^A_n$ which will be similar, though not identical, to the involution defined in \cite{SiU}. Write the elements $1,2,\cdots ,n$ clockwise around a circle, and write elements $1',2',\cdots ,n'$ interlaced in counterclockwise order, so that $1'$ is between $1$ and $n$, $2'$ is between $n$ and $n-1$, and so on, $i'$ is between $n+2-i$ and $n+1-i$. For $\pi \in NC^A_n$, join by chords -- as usual -- cyclically successive (unprimed) elements belonging to the same block of $\pi$. Then define $g(\pi)$ to be the coarsest noncrossing partition on the elements $1',2',\cdots ,n'$ so that the chords joining primed elements of the same block do not intersect the chords of $\pi$. See Figure 2 for an example. \begin{figure}[ht] \begin{center} \epsfig{file=duality.eps} \caption{The partition $\pi=(\{1\},\{2,3,8\},\{4,5,7\},\{6\})$ and its image $g(\pi)$} \label{duality} \end{center} \end{figure} The map $g$ is certainly a bijection, and it is order-reversing in $NC^A_n$ since merging two blocks of $\pi$ subdivides a block of $g(\pi)$. We claim that $g$ is also order-reversing on $P^A_n$. To see this, observe that for any $i>1$, the element $i$ is the smallest in its block in $\pi$ if and only if the element $(n+2-i)'$ is {\em not} the smallest in its block in $g(\pi)$. Indeed, the definition of $g$ implies that exactly one of $i$ and $(n+2-i)'$ can be connected to smaller elements by a chord. Therefore, $g$ takes the set of block-minima (not equal to 1) of $\pi$ into its reverse complement in $[n+2]$, so $g$ is indeed order-reversing on $P^A_n$. In the type-$B$ case, one can obtain an analogous bijection $h$ in a similar way: take the circular (clockwise) representation of $\pi$, then write the elements $1',2',\cdots, n',\bar{1}',\bar{2}',\cdots ,\bar{n}'$ so that the primed numbers interlace the unprimed, placing $1'$ between $1$ and $\bar{n}$ and continuing counterclockwise. For $\pi \in NC^B_n$, define $h(\pi)$ as above, that is, as the unique coarsest partition on the primed set whose chords do not cross those of $\pi$. Then $h$ is certainly an order-reversing bijection of $NC^B_n$, and as above, it reverses the containment of the sets $L(\pi)$ and $R(\pi)$, so it does extend to an order-reversing bijection of $P^B_n$. } \item{ {\em The M\"obius function and order complexes of $P^A_n$ and $P^B_n$.} It is easy to write an expression for the number $c_m(P^B_n )$ of chains ${\hat 0} < b^1 < b^2 < \cdots < b^m < {\hat 1}$ of length $m+1$ in $P^B_n$, for $m \ge 0$. Of course, $c_0(P^B_n) = 1$, and \begin{equation}\label{eq-PBchains} c_m(P^B_n) = \sum_{0 <k_1 <k_2 < \cdots < k_m < n} { { n \choose {k_1}} {n \choose {k_2}} \cdots {n \choose {k_m}} { {n!} \over { k_1! (k_2 - k_1)! \cdots (k_m - k_{m-1})! (n - k_m)! }}} \end{equation} since under the correspondence $b^i \leftrightarrow (L(b^i), {\rm Des}(b^i))$ a chain in $P^B_n$ corresponds to an $m$-tuple of subsets $(L(b^i))$ and a chain of subsets ${\rm Des}(b^1) \subset {\rm Des}(b^2) \subset \cdots \subset {\rm Des}(b^m)$ of $[n]$, with $\# L(b^i) = \# {\rm Des}(b^i) = k_i$. In turn, this leads to an expression for the M\"obius function of $P^B_n$, $\mu_{P^B_n}({\hat 0}, {\hat 1}) = \sum_{m \ge 0}{ (-1)^{m-1} c_m(P^B_n)}$. These expressions can be regarded as partial success with the computation of the zeta polynomial and the M\"obius function. It would be interesting to elucidate further the question of these invariants for $P^A_n$ and $P^B_n$, and to describe the order complexes of these posets. } \item{ {\em Other posets of combinatorial objects with similar properties.} The behavior of noncrossing partitions and restricted permutations suggests the following question: what other combinatorial objects admit a natural partial order which is self-dual and possibly, has other nice properties? A natural candidate is the class of {\em two-stack sortable permutations} \cite{doron}. It is known \cite{schaeffer} that there are as many of them with $k$ descents as with $n-1-k$ descents. However, the poset obtained by the descent ordering is not self-dual, even for $n=4$, so another ordering is needed. Similarly, the type-D noncrossing partitions and the interpolating BD-noncrossing partitions do not, in general, form self-dual posets when ordered by refinement (see \cite{Reiner-NCB}). However, it may be interesting to find corresponding classes of pattern-avoiding elements in the Weyl group for type D, along with an order-preserving embedding $NC^D_n \to P^D_n$ analogous to the type-A and B cases.} \end{enumerate}
\section{Introduction} In quantum field theory the structure of superselection sectors is entirely encoded in the set of localized endomorphisms of the algebra of local observables \cite{DHR3,H}. In the case of main physical interest, viz.\ in four dimensional Minkowski space, the set of (equivalence classes of) localized endomorphisms can be identified with the representation category of a unique compact group, the \emph{global gauge group} of the theory \cite{DR89,DR90}. This gauge group acts on a larger field algebra containing besides the observables charge carrying fields with normal commutation relations which reach all superselection sectors from the vacuum \cite{DR89a,DR90}. Gauge group and field algebra are intrinsically determined by the observable data. The relation between localized endomorphisms and representations of the gauge group is made concrete in the following way \cite{DR72}. There is a functor which assigns to a localized endomorphism $\varrho$ the \emph{Hilbert space of isometries} $H_\varrho$ consisting of all local fields $\Psi$ which induce $\varrho$: $$H_\varrho\DEF\SET{\Psi}{\Psi a=\varrho(a)\Psi,\text{ for all local observables }a}.$$ The action of the gauge group on the field algebra restricts to a unitary representation $D_\varrho$ on $H_\varrho$ relative to the inner product $\langle\Psi,\Psi'\rangle\1\DEF\Psi^*\Psi'$. This representation of the gauge group determines the \emph{charge} of the endomorphism $\varrho$; it is customary to refer to any label which characterizes the representation $D_\varrho$ as the \emph{charge quantum numbers} of $\varrho$. It will be used below that the representation $D_\varrho$ is in a canonical way unitarily equivalent to the representation on the Hilbert space $H_\varrho\Omega$ generated by applying the field operators in $H_\varrho$ to the vacuum vector $\Omega$. Any orthonormal basis $\Psi_1,\dots,\Psi_d$ in $H_\varrho$ generates a representation of the Cuntz algebra $\mathcal{O}_d$ and implements the endomorphism $\varrho$ as follows: \begin{equation} \label{IMP} \varrho(a)=\sum_{j=1}^d\Psi_ja\Psi_j^*. \end{equation} Using this formula, $\varrho$ can be canonically extended to an endomorphism of the field algebra. This extension is \emph{gauge invariant}, i.e.\ commutes with all gauge automorphisms: \begin{Prp} \label{prp:GI} Let $\varrho$ be an endomorphism of the field algebra which is implemented by a Hilbert space of isometries $H_\varrho$ as in \eqref{IMP}. Then $H_\varrho$ is gauge invariant if and only if $\varrho$ is gauge invariant. \end{Prp} \begin{proof} Assume first that $H_\varrho$ is invariant under gauge automorphisms $\gamma$. Since the representation $D_\varrho$ of the gauge group on $H_\varrho$, given by $D_\varrho(\gamma)\DEF\gamma|_{H_\varrho}$, is unitary, the $\gamma(\Psi_j)$ also form an orthonormal basis in $H_\varrho$. Since the endomorphism associated with a Hilbert space of isometries as in \eqref{IMP} is independent of the choice of an orthonormal basis in $H_\varrho$, it follows that, for any field operator $f$, \begin{equation*} \gamma(\varrho(f))=\sum_j\gamma(\Psi_j)\gamma(f) \gamma(\Psi_j)^*=\varrho(\gamma(f)). \end{equation*} Conversely, assume that $\varrho$ is gauge invariant. Let $\Psi\in H_\varrho$ and let $\gamma$ be a gauge transformation. Then one has for any field operator $f$ $$\gamma(\Psi)f=\gamma\bigl(\Psi\gamma^{-1}(f)\bigr) =\gamma\bigl(\varrho(\gamma^{-1}(f))\Psi\bigr)=\varrho(f)\gamma(\Psi)$$ so that $\gamma(\Psi)\in H_\varrho$. \end{proof} The existence of localized endomorphisms and associated Hilbert spaces of isometries follows from first principles of local quantum field theory. But it is by no means obvious how to obtain them explicitly in concrete models. In previous work we have developed a general theory of quasi-free endomorphisms of the CAR and CCR algebras which can be implemented by Hilbert spaces of isometries on Fock space \cite{CB1,CB3}. Among the results are implementability conditions for endomorphisms, which generalize the well-known criteria of Shale and Stinespring for automorphisms \cite{S,SS65}, and detailed constructions of field operators which implement endomorphisms according to \eqref{IMP}. In the present paper we are interested in the possible charge quantum numbers of such endomorphisms. The CAR resp.\ CCR algebra will play the role of the field algebra. Therefore quasi-free endomorphisms have to be viewed as endomorphisms of the field algebra and, by Proposition~\ref{prp:GI}, we have to restrict attention to endomorphisms which are gauge invariant under an appropriate group action. We will consider quasi-free actions of arbitrary groups which leave the Fock vacuum invariant. We show that the charge quantum numbers are then determined by the natural Fock space structure found in \cite{CB1,CB3} of the Hilbert spaces of isometries $H_\varrho$ implementing gauge invariant quasi-free endomorphisms $\varrho$: The representation $D_\varrho$ is unitarily equivalent to the representation $\Lambda_{\mathfrak{k}_\varrho}$ on the antisymmetric Fock space over an auxiliary unitary $G$-module $\mathfrak{k}_\varrho$, tensored with a certain character ${\det}_{\mathfrak{h}_\varrho}$: $$D_{\varrho}\simeq{\det}_{\mathfrak{h}_\varrho}\otimes \Lambda_{\mathfrak{k}_\varrho}\quad\text{(CAR),}$$ resp.\ to the representation $\mathfrak{S}_{\mathfrak{k}_\varrho}$ on the symmetric Fock space over $\mathfrak{k}_\varrho$: $$D_{\varrho}\simeq\mathfrak{S}_{\mathfrak{k}_\varrho} \quad\text{(CCR).}$$ This is our main result, contained in Theorems~\ref{thm:CHARGE} and \ref{ccr:thm:CHARGE}. It follows that $D_{\varrho}$ is reducible if $\varrho$ is non-surjective. Any $\varrho$ has a quasi-free conjugate $\varrho^c$ such that $D_{\varrho^c}$ is equivalent to the complex conjugate of $D_{\varrho}$, provided that the single-particle space has a particle-antiparticle symmetry. The analysis of the representations $D_\varrho$ is completely independent of localization properties of endomorphisms. In order to show that localization and implementability are not in conflict, we give in Section~\ref{sec:EXA} an explicit example of a \emph{localized} implementable gauge invariant endomorphism $\varrho$, with $\dim H_\varrho=2^N$, of the free massless Dirac field with $U(N)$ gauge symmetry in two dimensions. The construction rests on the use of ``local'' Fourier bases for the chiral components, and is in this respect similar to the known examples \cite{JMB1,JMB2} of localized endomorphisms in conformal field theory. The present investigations are taken from the author's Ph.\,D. thesis in physics \cite{CB4}, to which we refer for further results and discussions. \section{The Fermionic Case} \label{sec:FER} First of all we need some formalism and some results from \cite{CB1,CB4}. \subsection{Preliminaries on the implementation of quasi-free endomorphisms (CAR)} \label{sec:PRECAR} Recall Araki's approach to the canonical anticommutation relations \cite{A70,A87}: Let \KK be an infinite dimensional separable complex Hilbert space, endowed with a complex conjugation $f\mapsto f^*$. The {\em(selfdual) CAR algebra} \CK over \KK is the unique (simple) $C^*$-algebra generated by \1 and the elements of \KK, subject to the anticommutation relation $$f^*g+gf^*=\langle f,g\rangle\1,\quad f,g\in\KK.$$ Let $P_1$ be a fixed \emph{basis projection} on \KK, i.\,e.\ an orthogonal projection such that $$\4{P_1}=\1-P_1.$$ Here the bar denotes the complex conjugate of an operator $A$: $$\4{A}(f)\DEF A(f^*)^*,\quad f\in\KK.$$ Let $P_2$ be the complementary (basis) projection: $$P_2\DEF \1-P_1.$$ The components of an operator $A$ on \KK with respect to the decomposition $$\KK=\KK_1\oplus\KK_2$$ given by $P_1$ and $P_2$ will be denoted by $$A_{mn}\DEF P_mAP_n,\quad m,n=1,2$$ and will be regarded as operators from $\KK_n$ to $\KK_m$. To the basis projection $P_1$ there corresponds a unique (pure, quasi-free) \emph{Fock state} $\omega_{P_1}$ which is completely determined by the condition that \begin{equation} \label{FS} \omega_{P_1}(f^*f)=0\quad\text{if}\quad P_1f=0. \end{equation} The GNS representation associated with $\omega_{P_1}$ will be denoted by $(\FF_{P_1},\pi_{P_1},\Omega_{P_1})$. The Hilbert space $\FF_{P_1}$ can be identified with the antisymmetric Fock space over $\KK_1$. The elements of \KK are then represented by sums of creation and annihilation operators: \begin{equation} \label{PP1} \pi_{P_1}(f)= a^*(P_1f)+a(P_1f^*),\quad f\in\KK, \end{equation} and the cyclic vector $\Omega_{P_1}$ is the Fock vacuum vector. Every isometry $V$ on \KK which commutes with complex conjugation extends to a unique \emph{quasi-free endomorphism} $\varrho_V$ of \CK: $$\varrho_V(f)=V(f),\quad f\in\KK.$$ As shown in \cite{CB1}, an endomorphism $\varrho_V$ can be implemented by a Hilbert space of isometries $H_{\varrho_V}$ (cf.~\eqref{IMP}) in the Fock state $\omega_{P_1}$ if and only if \begin{equation} \label{IMPCOND} [P_1,V]\text{ is Hilbert--Schmidt (HS)}. \end{equation} These isometries form a semigroup $$\END_{P_1}(\KK)\DEF\SET{V\in\BB(\KK)}{V^*V=\1,\ \4{V}=V,\ [P_1,V] \text{ is HS}}$$ isomorphic to the semigroup of all implementable quasi-free endomorphisms of \CK. The \emph{statistics dimension} $d_{\varrho_V}$ of $\varrho_V$ is given by $$d_{\varrho_V}\DEF\dim H_{\varrho_V}=2^{\2\IND V},$$ where $\IND V$ is, up to the sign, the Fredholm index of $V$: $$\IND V=\dim\ker V^*\in\{0,2,4,\dots,\infty\},\quad V\in\END_{P_1}(\KK).$$ The grading automorphism of \CK is equal to the quasi-free automorphism $\varrho_{-\1}$. Let $\Gamma(-\1)$ be the (self-adjoint, unitary) second quantization of $\varrho_{-\1}$, given by $$\Gamma(-\1)\pi_{P_1}(a)\Omega_{P_1}=\pi_{P_1}(\varrho_{-\1}(a)) \Omega_{P_1},\quad a\in\CK,$$ and let $\theta(-\1)$ be the unitary operator \begin{equation} \label{THETA} \theta(-\1)\DEF\tfrac{1}{\sqrt{2}}\bigl(\1-i\Gamma(-\1)\bigr). \end{equation} Then the \emph{twisted Fock representation} $\psi_{P_1}$ induced by $P_{1}$ is defined by \begin{equation} \label{TWIST} \psi_{P_1}(a)\DEF\theta(-\1)\pi_{P_1}(a)\theta(-\1)^*. \end{equation} It can be used to describe the commutants of ``local'' subalgebras: If $\HH\subset\KK$ is a subspace invariant under complex conjugation, and $\mathfrak{C}(\HH)$ the $C^*$-subalgebra of \CK generated by $\HH$, then $$\pi_{P_1}(\mathfrak{C}(\HH))'=\psi_{P_1}(\mathfrak{C}(\HH^\bot))'' \qquad\text{(\emph{twisted duality})}.$$ Let $V\in\END_{P_1}(\KK)$ be given. As mentioned in the introduction, it suffices for our purposes to consider group actions on the Hilbert space $H_{\varrho_V}\Omega_{P_1}$. An orthonormal basis in this space can be obtained as follows. Define a finite dimensional subspace $\mathfrak{h}\subset\KK_1$ by \begin{equation} \label{H} \mathfrak{h}\DEF V_{12}(\ker V_{22}), \end{equation} and an antisymmetric\footnote{An operator $A$ on \KK is \emph{antisymmetric} if $\4{A^*}=-A$.} Hilbert--Schmidt operator $T$ from $\KK_1$ to $\KK_2$ by \begin{equation} \label{T} T\DEF V_{21}{V_{11}}^{-1}-{V_{22}}^{-1*}{V_{12}}^*[\ker {V_{11}}^*]. \end{equation} Here the bounded operator ${V_{11}}^{-1}$ is defined to be zero on $\ker{V_{11}}^*$, and $[\HH]$ denotes the orthogonal projection onto a closed subspace $\HH\subset\KK$. Then one has $T(\mathfrak{h})=0$; such pairs $(\mathfrak{h},T)$ parameterize the class of all Fock states which are unitarily equivalent to the given Fock state $\omega_{P_1}$ \cite {CB4}. The basis projection $P$ corresponding to the pair $(\mathfrak{h},T)$ is explicitly given by \begin{equation} \label{P} P\DEF(P_1+T)(P_1+T^*T)^{-1}(P_1+T^*)-[\mathfrak{h}]+ [\mathfrak{h}^*], \end{equation} and $\mathfrak{h}$ and $T$ can be recovered from $P$ by \begin{align} \mathfrak{h} &=\ker P_{11},\label{HDEF}\\ T &=P_{21}{P_{11}}^{-1}\label{TDEF} \end{align} (${P_{11}}^{-1}$ is defined in a similar way as ${V_{11}}^{-1}$ above). The Fock state $\omega_P$ associated with $P$ is an extension of the partial Fock state $\omega_{P_1}\0\varrho_V^{-1}|_{\RAN\varrho_V}$, and is induced by the cyclic vector $$\Omega_P=\bigl({\det}(P_1+T^*T)\bigr)^{-1/4}\psi_{P_1}(e_1\dotsm e_L)\exp(\3\4{T}a^*a^*)\Omega_{P_1}\in\FF_{P_1}$$ where the determinant has to be computed on $\KK_1$, $\{e_1,\dots,e_L\}$ is an orthonormal basis in $\mathfrak{h}$, and the exponential term has a well-defined meaning as a strongly convergent series on a dense domain containing $\Omega_{P_1}$. The vector $\Omega_P$ belongs to $H_{\varrho_V}\Omega_{P_1}$; in fact, the latter Hilbert space consists precisely of the vectors in $\FF_{P_1}$ which induce extensions of the partial Fock state $\omega_{P_1}\0\varrho_V^{-1}|_{\RAN\varrho_V}$. A complete orthonormal basis in $H_{\varrho_V}\Omega_{P_1}$ can be obtained by applying suitable partial isometries from the commutant of $\RAN\varrho_V$ to $\Omega_P$. The basis projection $P$ leaves $\ker V^*$ invariant. Let \begin{equation} \label{K} \mathfrak{k}\DEF P(\ker V^*),\quad\text{with } \dim\mathfrak{k}=\3\IND V, \end{equation} and let $\{g_j\}$ be an orthonormal basis in $\mathfrak{k}$. For any multi-index $\alpha=(\alpha_1,\dots,\alpha_l),\ 1\leq\alpha_1<\dots<\alpha_l\leq\2\IND V$ (resp.\ $\alpha=0$ if $l=0$), set \begin{equation} \label{OA} \Omega_\alpha\DEF\psi_{P_1}(g_{\alpha_1}\dotsm g_{\alpha_l})\Omega_P. \end{equation} Then the $\Omega_\alpha$ constitute an orthonormal basis in $H_{\varrho_V}\Omega_{P_1}$, and they determine an orthonormal basis $\{\Psi_\alpha\}$ in $H_{\varrho_V}$ via \begin{equation} \label{HRV} \Psi_\alpha\pi_{P_1}(a)\Omega_{P_1}= \pi_{P_1}(\varrho_V(a))\Omega_\alpha. \end{equation} Since $\psi_{P_1}$ is a representation of the canonical anticommutation relations and since $\Omega_P$ is annihilated by the operators $\psi_{P_1}(g_j)^*$, the spaces $H_{\varrho_V}$ and $H_{\varrho_V}\Omega_{P_1}$ can both be identified with the antisymmetric Fock space over $\mathfrak{k}$ \cite{CB1,CB4}. \subsection{Gauge invariant quasi-free endomorphisms (CAR)} \label{sec:GICAR} Let $G\subset\END_{P_1}(\KK)$ be a group\footnote{The results in this section hold also for non-compact groups (relative to the strong topology). However, the close relationship between representations of $G$ and superselection sectors is then lost. We also do not require that $-\1\in G$, which would be necessary if the $G$-invariant elements of \CK were to be interpreted as physical observables.} consisting of unitary operators which commute with $P_1$. The usual second quantization of $U\in G$ (or, more precisely, of $U_{11}$) will be denoted by $\Gamma(U)$; the map $U\mapsto\Gamma(U)$ is strongly continuous. The corresponding gauge automorphisms, which leave $\omega_{P_1}$ invariant, will be denoted by $\gamma_U$. We are interested in the representation $D_{\varrho_V}$ of $G$ on the Hilbert space $H_{\varrho_V}$ which implements a gauge invariant quasi-free endomorphism $\varrho_V$. \emph{Gauge invariant} implementable quasi-free endomorphisms are given by the elements of the semigroup $$\END_{P_1}(\KK)^G\DEF\SET{V\in\END_{P_1}(\KK)}{[V,G]=0}.$$ To determine $D_{\varrho_V}$ up to unitary equivalence, it suffices to calculate the transformed vectors $\Gamma(U)\Omega_\alpha$, $U\in G$, where the $\Omega_\alpha$ are the basis elements in $H_{\varrho_V}\Omega_{P_1}$ defined in \eqref{OA}. The basic observation is that the objects entering the construction of the $\Omega_\alpha$, namely the spaces $\mathfrak{h}$ and $\mathfrak{k}$ and the operator $T$, are all gauge invariant. \begin{Lem} \label{lem:GP} Let $U\in G$. Then $\Gamma(U)$ implements the gauge automorphism $\gamma_U$ in the twisted Fock representation $\psi_{P_1}$: $$\Gamma(U)\psi_{P_1}(a)=\psi_{P_1}(\gamma_U(a))\Gamma(U), \quad a\in\CK.$$ \end{Lem} \begin{proof} Since $U$ commutes with $P_1$, the implementer $\Gamma(U)$ is even: $$[\Gamma(U),\Gamma(-\1)]=0.$$ This implies that $[\Gamma(U),\theta(-\1)]=0$ (see \eqref{THETA}) so that, by \eqref{TWIST}, \begin{equation*} \Gamma(U)\psi_{P_1}(a)\Gamma(U)^*=\theta(-\1)\Gamma(U) \pi_{P_1}(a)\Gamma(U)^*\theta(-\1)^*=\psi_{P_1}(\gamma_U(a)). \end{equation*} \end{proof} \begin{Lem} \label{lem:EXP} Let $V\in\END_{P_1}(\KK)^G$. Then $\exp(\3\4{T}a^*a^*)\Omega_{P_1}$, with $T$ defined by \eqref{T}, is invariant under all gauge transformations $\Gamma(U),\ U\in G$. \end{Lem} \begin{proof} If $S$ is an antisymmetric operator from $\KK_1$ to $\KK_2$ of finite rank, then one readily verifies that $$\Gamma(U)(\3\4{S}a^*a^*)\Gamma(U)^*=\3(U\4{S}U^*)a^*a^*$$ (here the expressions of the form $\4{S}a^*a^*$ are defined by expanding $S=\sum f_j^*\langle g_j,.\rangle$ with $f_j,g_j\in\KK_1$, and by setting $\4{S}a^*a^*=\sum a^*(f_j)a^*(g_j)$). Approximating $T$ by such finite rank operators in the Hilbert--Schmidt norm (cf.~\cite{CR}), one finds that $$\Gamma(U)(\3\4{T}a^*a^*)^n\Omega_{P_1}= \bigl(\3(U\4{T}U^*)a^*a^*\bigr)^n\Omega_{P_1},\quad n\in\NN,$$ because $\Gamma(U)\Omega_{P_1}=\Omega_{P_1}$. It follows that \begin{equation*} \begin{split} \Gamma(U)\exp(\3\4{T}a^*a^*)\Omega_{P_1}&=\sum_{n=0}^\infty \frac{1}{n!}\Gamma(U)(\3\4{T}a^*a^*)^n\Omega_{P_1}\\ &=\exp\bigl(\3(U\4{T}U^*)a^*a^*\bigr)\Omega_{P_1}. \end{split} \end{equation*} Since $U$ commutes with $P_1,\ P_2$ and $V$, it also commutes with all components of $V$ and $V^*$, including the operators ${V_{11}}^{-1},\ [\ker{V_{11}}^*]$ etc. The operator $T$ is by \eqref{T} a bounded function of these components, so that \begin{equation} \label{GT} [U,T]=0,\quad U\in G. \end{equation} Hence we get $\Gamma(U)\exp(\3\4{T}a^*a^*)\Omega_{P_1}= \exp(\3\4{T}a^*a^*)\Omega_{P_1}$ as claimed. \end{proof} Setting $\delta\DEF\bigl({\det}(P_1+T^*T)\bigr)^{-1/4}$, we thus arrive at the following formula: \begin{equation} \label{GPSIOM} \Gamma(U)\Omega_\alpha=\delta\cdot\psi_{P_1}\bigl(U(g_{\alpha_1}) \dotsm U(g_{\alpha_l})U(e_1)\dotsm U(e_L)\bigr)\exp(\3\4{T}a^*a^*) \Omega_{P_1}. \end{equation} \begin{Thm} \label{thm:CHARGE} Let $P_1$ be a basis projection of \KK, let $G$ be a group of unitary operators on \KK commuting with $P_1$ and with complex conjugation, and let $V\in\END_{P_1}(\KK)^G$. Then the finite dimensional subspace $\mathfrak{h}\subset\KK_1$ and the $(\2\IND V)$ dimensional subspace $\mathfrak{k}\subset\KK$ associated with $V$ by \eqref{H} and \eqref{K} are both invariant under $G$. Let $\Lambda_{\mathfrak{k}}$ be the unitary representation of $G$ on the antisymmetric Fock space over $\mathfrak{k}$ that is obtained by taking antisymmetric tensor powers of the representation on $\mathfrak{k}$. Then the unitary representation $D_{\varrho_V}$ of $G$ on the Hilbert space of isometries $H_{\varrho_V}$ which implements $\varrho_V$ in the Fock state $\omega_{P_1}$ is unitarily equivalent to $\Lambda_{\mathfrak{k}}$, tensored with the one dimensional representation ${\det}_{\mathfrak{h}}(U)\DEF\det(U|_{\mathfrak{h}})$\textup{:} \begin{equation} \label{REPG} D_{\varrho_V}\simeq{\det}_{\mathfrak{h}}\otimes \Lambda_{\mathfrak{k}}. \end{equation} \end{Thm} \begin{proof} The subspace $\mathfrak{h}=V_{12}(\ker V_{22})$ is invariant under $G$ because $G$ commutes with the components of $V$ (cf.\ the proof of Lemma~\ref{lem:EXP}). Since $\{e_1,\dots,e_L\}$ is an orthonormal basis in $\mathfrak{h}$, it follows from the canonical anticommutation relations that $$U(e_1)\dotsm U(e_{L})=\det\bigl(U|_{\mathfrak{h}}\bigr)\cdot e_1\dotsm e_{L},\qquad U\in G.$$ Similarly, $G$ commutes with the basis projection $P$ (cf.\ \eqref{P} and \eqref{GT}) and leaves $\ker V^*$ invariant, so that $\mathfrak{k}=P(\ker V^*)$ is also left invariant. It then follows from the canonical anticommutation relations that $g_{\alpha_1}\dotsm g_{\alpha_l}$ transforms like the $l$-fold antisymmetric tensor product of $g_{\alpha_1},\dots,g_{\alpha_l}$ under $G$. Thus we see from \eqref{GPSIOM} that the representation of $G$ on $H_{\varrho_V}\Omega_{P_1}$ is unitarily equivalent to ${\det}_{\mathfrak{h}}\otimes\Lambda_{\mathfrak{k}}$, and the same holds true for the representation $D_{\varrho_V}$. \end{proof} \begin{Rms} (i) Theorem~\ref{thm:CHARGE} shows that non-surjective quasi-free endomorphisms $\varrho_V$ are always \emph{reducible} in the sense that the representation $D_{\varrho_V}$ (or, if $G$ is compact, the representation induced by $\varrho_V$ of the pointwise gauge invariant ``observable'' algebra $\CK^G$ on the subspace of $\Gamma(G)$-invariant vectors in $\FF_{P_1}$) is reducible. In fact, each ``$n$-particle'' subspace of $H_{\varrho_V}$, i.e.\ the closed linear span of all $\Psi_\alpha$ with $\alpha$ of length $n$, is invariant under $G$, and may decompose further. Let $D_{\varrho_V}^{(n)}$ be the restriction of $D_{\varrho_V}$ to this subspace. Closest to irreducibility is the case that at least $D_{\varrho_V}^{(1)}$ is irreducible. In typical situations, the remaining representations $D_{\varrho_V}^{(n)}$ will then also be irreducible. This happens for instance if $G\cong U(N)$ or $G\cong SU(N)$, and $\mathfrak{k}$ carries the defining representation of $G$. In the $U(N)$ case, the $D_{\varrho_V}^{(n)}$ are not only irreducible, but also mutually inequivalent. In the $SU(N)$ case, the representations $D_{\varrho_V}^{(0)},\dots,D_{\varrho_V}^{(N-1)}$ are mutually inequivalent, but $D_{\varrho_V}^{(N)}$ is equivalent to $D_{\varrho_V}^{(0)}$. In general, it can nevertheless happen that $D_{\varrho_V}^{(1)}$ is irreducible but some $D_{\varrho_V}^{(n)}$ are not, as is the case if $G\cong SO(N)$ ($N>2$ even) and $\mathfrak{k}$ carries the defining representation of $G$ (cf.\ \cite{Wey,Boe}). If already $D_{\varrho_V}^{(1)}$ is reducible, then one has an additional Clebsch--Gordan type splitting. (ii) Theorem~\ref{thm:CHARGE} characterizes the representation $D_{\varrho_V}$ associated with a fixed gauge invariant endomorphism $\varrho_V$ in terms of the representations on $\mathfrak{h}$ and $\mathfrak{k}$. The question which representations of $G$ can be realized on the spaces $\mathfrak{h}$ and $\mathfrak{k}$ by letting $V$ vary through $\END_{P_1}(\KK)^G$ is studied in \cite{CB4}. In typical field theoretic situations, where the single-particle space $\KK_1$ carries an irreducible representation of a compact group together with its complex conjugate, both with infinite multiplicity, as e.\,g.\ in Section~\ref{sec:EXA} below, one finds that any irreducible representation of $G$ realized on Fock space $\FF_{P_1}$ is equivalent to a subrepresentation of some $D_{\varrho_V}$. (iii) A special case worth mentioning is the case $G\cong U(1)$ and $\IND V=0$, i.e.\ the case of the \emph{restricted unitary group.} It is well-known from the work on the external field problem (see e.\,g.\ \cite{CHOB}) that the charge of elements of the restricted unitary group is given by a certain Fredholm index $\IND V_{++}$ (which has nothing to do with the index of $V$, but refers to a finer decomposition $V_{11}=V_{++}\oplus V_{--}$). This fact can be easily derived from our general result: The factor $\Lambda_{\mathfrak{k}}$ in \eqref{REPG} becomes trivial, whereas $${\det}_{\mathfrak{h}}(U_\lambda)=\exp(i\lambda\IND V_{++})$$ if $U_\lambda\in G$ corresponds to $e^{i\lambda}\in U(1)$ \cite{CB4}. Similarly, in the case $G=\{\pm\1\}\cong\ZZ_2$ and $\IND V=0$, the factor $\Lambda_{\mathfrak{k}}$ is trivial, but $${\det}_{\mathfrak{h}}(-\1)=(-1)^{\dim\mathfrak{h}}= (-1)^{\dim\ker V_{11}}$$ yields the $\ZZ_2$-index of Araki and Evans \cite{AE,A87,EK}. (iv) If the single-particle space $\KK_1$ decomposes into the direct sum of two anti\-unitarily equivalent $G$-modules (``particle-antiparticle symmetry''), then there exists an involutive automorphism $V\mapsto V^c$ of $\END_{P_1}(\KK)^G$ such that the spaces $\mathfrak{h}^c$ and $\mathfrak{k}^c$ corresponding by \eqref{H} and \eqref{K} to $V^c$ are, as $G$-modules, antiunitarily equivalent to the spaces $\mathfrak{h}$ and $\mathfrak{k}$ corresponding to $V$. That is, the representation $D_{\varrho_{V^c}}$ is unitarily equivalent to the complex conjugate of $D_{\varrho_{V}}$ (\emph{charge conjugation}). \end{Rms} \subsection{An example: Localized endomorphisms of the chiral Dirac field} \label{sec:EXA} In Section~\ref{sec:GICAR} we have analyzed the charge quantum numbers of gauge invariant implementable quasi-free endomorphisms in complete generality. In particular, and in sharp contrast to the field theoretic situation, it was not necessary to assume any localization properties of endomorphisms. If one could find, in a specific model, a localized implementable quasi-free endomorphism, then our methods would apply and could be used to determine its charge and to construct the corresponding local fields. It is however not clear from the outset whether localization and implementability are compatible with each other\footnote{Known results concerning this question are restricted to the case of automorphisms. Building on the work of Carey and Ruijsenaars \cite{CR} and others, we constructed in \cite{CB0} a family of (implementable and transportable) localized automorphisms, carrying arbitrary $U(1)$-charges, of the free Dirac field in two spacetime dimensions with arbitrary mass. The operators $V\in\END_{P_1}(\KK)^{U(1)}$ belonging to these automorphisms are given by two $U(1)$-valued functions which are equal to $1$ at spacelike infinity, and the charge $\IND V_{++}$ (cf.\ Remark~(iii) in Section~\ref{sec:GICAR}) of $\varrho_V$ is equal to the difference of the winding numbers of these functions. However, unlike in two dimensions, there seem to be no known examples of implementable charge-carrying automorphisms in the case $G\cong U(1)$ in four spacetime dimensions.}. To show that this is in fact the case, we will present below an explicit example of a non-surjective implementable localized quasi-free endomorphism of the free massless Dirac field in two spacetime dimensions. Let us first introduce the free Dirac field with global $U(N)$ symmetry. Let \begin{equation} \label{HL2} \HH\DEF L^2(\RR^{2n-1},\CC^{2^n}) \end{equation} be the single-particle space of the time-zero Dirac field in $2n$ spacetime dimensions. Let $H=-i\vec\alpha\vec\nabla+\beta m$ be the free Dirac Hamiltonian, with spectral projections $E_\pm$ corresponding to the positive resp.\ negative part of the spectrum of $H$. Tensored with $\1_N$, these operators act on the space \begin{equation} \label{CN} \HH'\DEF\HH\otimes\CC^N. \end{equation} The gauge group $U(N)$ also acts naturally on $\HH'$. In the selfdual CAR formalism, one sets \begin{equation} \label{H'} \KK\DEF\HH'\oplus{\HH'}^*, \end{equation} where ${\HH'}^*$ is the Hilbert space conjugate to $\HH'$. There is a natural conjugation $f\mapsto f^*$ on \KK which is inherited from the antiunitary identification map $\HH'\to{\HH'}^*$. The basis projection $P_1$ corresponding to the vacuum representation of the field is given by $$P_1\DEF E_+'\oplus\4{E_-'}$$ with $E_\pm'=E_\pm\otimes\1_N$. Gauge transformations act like $U=(\1_\HH\otimes u)\oplus(\1_{\HH^*}\otimes\4{u}),\ u\in U(N)$, on \KK. They commute with $P_1$. The field operators $\varphi_t$ at time $t$ are given by $$\varphi_t(f)\DEF\pi_{P_1}(e^{itH'}f)= a(E_+'e^{itH'}f)^*+a(\4{E_-'}e^{-it\4{H'}}f^*)$$ with $H'\DEF H\otimes\1_N,\ f\in\HH'$. They are solutions of the Dirac--Schr\"odinger equation $$-i\frac{d}{dt}\varphi_t(f)=\varphi_t(H'f),\quad f\in D(H').$$ If $O$ is a double cone in Minkowski space with base $B\subset\RR^{2n-1}$ at time $t$, then the local field algebra associated with $O$ is the von Neumann algebra generated by all $\varphi_t(f)$ with $\SUPP f\subset B$. The local observable algebras are the fixed point subalgebras of the local field algebras under the gauge action. A whole net of local algebras is generated from these special ones by applying Lorentz transformations. Gauge invariant implementable localized endomorphisms of the $N$-component Dirac field can be characterized as follows. A quasi-free endomorphism $\varrho_V$ is gauge invariant if and only if $V$ has the form \begin{equation} \label{VFORM} V=(v\otimes\1_N)\oplus(\4{v}\otimes\4{\1_N}) \end{equation} with respect to the decomposition \eqref{H'}, where $v$ is an isometry of \HH. This follows from the fact that the defining representation of $U(N)$ and its complex conjugate are disjoint, so that the commutant of $G$ on \KK is given by $$G'=\bigl(\BB(\HH)\otimes\1_N\bigr)\oplus\bigl(\BB(\HH^*)\otimes \4{\1_N}\bigr).$$ For $V$ of the form \eqref{VFORM} one has $$[P_1,V]=\Bigl([E_+,v]\otimes\1_N\Bigr)\oplus\Bigl(\4{[E_-,v]} \otimes\4{\1_N}\Bigr)$$ so that the implementability condition \eqref{IMPCOND} holds if and only if \begin{equation} \label{Ev} [E_+,v]\text{ and }[E_-,v]\text{ are Hilbert--Schmidt}. \end{equation} Therefore $\END_{P_1}(\KK)^{U(N)}$ is isomorphic to the semigroup of all isometries $v$ of \HH which fulfill \eqref{Ev}. An endomorphism of the algebra of all local observables is \emph{localized} in a bounded region $O$ in Minkowski space if it acts like the identity on observables which are localized in bounded regions contained in the spacelike complement $O'$ of $O$ \cite{H}. Localized elements of $\END_{P_1}(\KK)^{U(N)}$ (at time zero) can be characterized as follows. \begin{Prp} \label{prop:LOC} Let $O$ be a double cone with base $B\subset\RR^{2n-1}$ at time zero. Let $V$ be an element of $\END_{P_1}(\KK)^{U(N)}$, and let $v$ be the isometry of \HH associated with $V$ by \eqref{VFORM}. Then $\varrho_V$ is localized\footnote{More precisely, the normal extension of $\varrho_V$ in the representation $\pi_{P_1}$ is localized in $O$.} in $O$ if and only if there exists, for each connected component $\Delta$ of $\RR^{2n-1}\setminus B$, a phase factor $\tau_\Delta\in{U(1)}$ such that \begin{equation} \label{TAUDELTA} v(f)=\tau_\Delta f\quad\text{for all }f\in\HH\text { with } \SUPP f\subset\Delta. \end{equation} \end{Prp} \begin{proof} Assume that $\varrho_V$ is localized in $O$. Let $b_1,\dots,b_N$ be the standard basis in $\CC^N$, let $\Delta$ be a component of the complement of $B$, and let $f,g\in\HH$ with $\SUPP f,g\subset\Delta$. Then $$a(f,g)\DEF\sum_{j=1}^N(f\otimes b_j)(g\otimes b_j)^*$$ is gauge invariant, and $\pi_{P_1}(a(f,g))$ is an observable localized in $O'$. Since $\varrho_V$ is localized in $O$, one has $a(f,g)=\varrho_V(a(f,g))=\sum_j(v(f)\otimes b_j)(v(g)\otimes b_j)^*$. Since the $b_j$ are linearly independent, it follows that \begin{equation} \label{FG} (f\otimes b_j)(g\otimes b_j)^*=(v(f)\otimes b_j)(v(g)\otimes b_j)^*, \qquad j=1,\dots,N. \end{equation} Now let $P'$ be the (basis) projection onto $\HH'\subset\KK$, and let $\omega_{P'}$ be the corresponding Fock state. One has $$\omega_{P'}\bigl((f\otimes b_j)^*(f\otimes b_j)(f\otimes b_j)^* (f\otimes b_j)\bigr)=\NORM{f}^4,$$ and, since $(v(f)\otimes b_j)^*$ belongs to the annihilator ideal of $\omega_{P'}$, \begin{equation*} \omega_{P'}\bigl((f\otimes b_j)^*(v(f)\otimes b_j)(v(f) \otimes b_j)^*(f\otimes b_j)\bigr)=\ABS{\langle v(f),f\rangle}^2. \end{equation*} Therefore one gets from \eqref{FG}, in the special case $f=g$, that $\NORM{f}^2=\ABS{\langle v(f),f\rangle}$. It follows that there exists $\tau_f\in{U(1)}$ such that $v(f)=\tau_f f$. By the same argument, $v(g)=\tau_g g$ for some $\tau_g\in{U(1)}$. Then \eqref{FG} yields that $\tau_f=\tau_g$. Therefore these phase factors depend only on $\Delta$ and not on the functions. Conversely, assume that \eqref{TAUDELTA} holds. Then $\varrho_V$ acts on fields localized in bounded regions in $O'$ like a gauge transformation, and therefore like the identity on observables localized in $O'$. It follows that $\varrho_V$ is localized in $O$. \end{proof} Of course, $\RR^{2n-1}\setminus B$ is connected if $n>1$, but it has two connected components if $n=1$. This is the basic reason for the possible occurrence of braid group statistics and soliton sectors in two dimensional Minkowski space. Next let us demonstrate that at least the free massless Dirac field in two spacetime dimensions possesses non-surjective implementable localized quasi-free endomorphisms. It suffices to consider one chiral component of the field. Thus consider the Hilbert space $\HH=L^2(\RR)$ with Dirac Hamiltonian $-i\tfrac{d}{dx}$. It is convenient to transform to the Hilbert space $L^2({S^1})$ via the Cayley transform $\vartheta$ $$\vartheta: \RR\cup\{\infty\}\to{S^1},\qquad x\mapsto -e^{2i\arctan x}=\frac{x-i}{x+i}$$ (cf.\ \cite{CR}). $\vartheta$ induces a unitary transformation $\tilde\vartheta$ $$\tilde\vartheta:L^2({S^1})\to L^2(\RR),\qquad(\tilde\vartheta f)(x)= \pi^{-\2}\frac{f(\vartheta(x))}{x+i}.$$ The important point is that the spectral projections $E_\pm$ of $-i\tfrac{d}{dx}$ are transformed into the Hardy space projections: Set $\tilde E_\pm\DEF\tilde\vartheta^{-1}E_\pm\tilde\vartheta$, then $$\tilde E_+=\sum_{n\geq0}e_n\langle e_n,.\,\rangle,\quad \tilde E_-=\sum_{n<0}e_n\langle e_n,.\,\rangle,\quad e_n(z)\DEF z^n\ (z\in{S^1},n\in\ZZ).$$ We want to construct an isometry $v$ of $L^2({S^1})$ with $[\tilde E_\pm,v]$ Hilbert--Schmidt (implementability), with $\IND v=1$ (close to irreducibility, cf.\ Rem.~(i) in Section~\ref{sec:GICAR}), and such that $v(f)=f$ for all $f\in L^2({S^1})$ with $\SUPP f\subset{S^1}\setminus I$, where $I\subset{S^1}$ is a fixed interval (localization). As localization region we shall choose the interval $$I\DEF\SET{e^{i\lambda}}{\tfrac{\pi}{2}\leq\lambda\leq\tfrac{3\pi}{2}}$$ which corresponds, by the inverse Cayley transform, to the interval $\vartheta^{-1}(I)=[-1,1]$ in \RR. We need the following orthonormal basis $(f_m)_{m\in\ZZ}$ in $L^2(I)\subset L^2({S^1})$ $$f_m(z)\DEF\sqrt{2}(-1)^mz^{2m}\chi_I(z),\quad z\in{S^1},$$ where $\chi_I$ is the characteristic function of $I$. We now define the isometry $v$ by \begin{equation} \label{v} v\DEF\1+\sum_{m\geq0}(f_{m+1}-f_m)\langle f_m,.\,\rangle. \end{equation} Note that $v$ acts like the identity on functions with support in ${S^1}\setminus I$, that $v(f_m)=f_m$ if $m<0$, and that $v$ acts like the unilateral shift on the remaining $f_m$: $v(f_m)=f_{m+1}$ if $m\geq0$. \begin{Lem} \label{lem:HS} The commutators $[\tilde E_+,v]$ and $[\tilde E_-,v]$ are Hilbert--Schmidt. \end{Lem} \begin{proof} The rather lengthy estimates of the Hilbert--Schmidt norms of these commutators, which are essentially due to P.~Grinevich, can be found in \cite{CB4}. \end{proof} Thus the operator $V\in\END_{P_1}(\KK)^{U(N)}$ induced by $v$ via \eqref{VFORM} yields a localized endomorphism $\varrho_V$ of the chiral Dirac field. Since by construction $$\3\IND V=N,$$ it is clear that the space $\mathfrak{k}$ associated with $V$ by \eqref{K} carries either the defining representation of $U(N)$ or its complex conjugate. By Remark~(i) in Section~\ref{sec:GICAR}, the irreducible constituents of $\varrho_V$ correspond to the irreducible, mutually inequivalent representations $D_{\varrho_V}^{(n)},\ n=0,\dots,N$. Note that the same isometry $v$ gives rise to localized gauge invariant implementable endomorphisms for \emph{arbitrary} symmetry groups $G$, by replacing the defining representation of $U(N)$ in Eq.~\eqref{CN} with a suitable finite dimensional representation of $G$. \section{The Bosonic Case} We need some preparations from \cite{CB3}. The exposition will closely follow the lines of the Fermionic case considered in Section~\ref{sec:FER}. \subsection{Preliminaries on the implementation of quasi-free endomorphisms (CCR)} \label{sec:PRECCR} We start with a Fock representation of the canonical commutation relations. Thus we may assume as above that \KK is an infinite dimensional separable complex Hilbert space with a complex conjugation $f\mapsto f^*$, and that $P_1$ is a fixed basis projection. Let $P_2$ be the complementary projection. Define a self-adjoint unitary operator $$C\DEF P_1-P_2$$ so that $\4{C}=-C$, and a nondegenerate hermitian sesquilinear form $$\kappa(f,g)\DEF\langle{f,Cg}\rangle$$ so that $$\kappa(f^*,g^*)=-\kappa(g,f),\quad f,g\in\KK.$$ It must be emphasized that the basic form on \KK which determines the canonical commutation relations is $\kappa$ and not the Hilbert space inner product. In fact, the latter depends on the choice of the Fock state, i.\,e.\ on the choice of $P_1$. The \emph{(selfdual) CCR algebra} \CKK over $(\KK,\kappa)$ is the simple *-algebra which is generated by \1 and elements $f\in\KK$, subject to the commutation relation \cite{AS71,A71} \begin{equation*} f^*g-gf^*=\kappa(f,g)\1,\quad f,g\in\KK. \end{equation*} The \emph{Weyl algebra} $\WK$ over $(\KK,\kappa)$ is the simple $C^*$-algebra generated by unitary operators $w(f),\ f\in\KK$ with $f=f^*$, subject to the relations $$w(f)^*=w(-f),\quad w(f)w(g)=e^{-\2\kappa(f,g)}w(f+g).$$ The \emph{Fock state} $\omega_{P_1}$ over \CKK induced by $P_1$ is again determined by condition \eqref{FS}. The GNS representation $(\FF_{P_1},\pi_{P_1},\Omega_{P_1})$ of $\omega_{P_1}$ can be identified with the representation $\pi_{P_1}$ given by formula \eqref{PP1}, where $a^*(f)$ and $a(f)$, $f\in\KK_1$, now are Bosonic creation and annihilation operators, acting on the symmetric Fock space $\FF_{P_1}$ over $\KK_1$ with Fock vacuum vector $\Omega_{P_1}$. All operators $\pi_{P_1}(a)$, $a\in\CKK$, are defined on the invariant dense domain $\DD\subset\FF_{P_1}$ of algebraic tensors, are closable, and fulfill $\pi_{P_1}(a^*)\subset\pi_{P_1}(a)^*$. The irreducible Fock representation of the Weyl algebra $\WK$ induced by $P_1$ is obtained by identifying the Weyl operator $w(f)$, $f=f^*\in\KK$, with the exponential of the closure of $i\pi_{P_1}(f)$. The vacuum expectation value of $w(f)$ is $$\langle{\Omega_{P_1}, w(f)\Omega_{P_1}} \rangle=e^{-\frac{1}{4}\NORM{f}^2},\quad f=f^*.$$ Let \HH be a subspace of \KK with $\HH=\HH^*$, and let $\WW(\HH)$ be the $C^*$-algebra generated by all $w(f)$ with $f=f^*\in\HH$. If $\HH\SC$ is the orthogonal complement of \HH with respect to $\kappa$, then \begin{equation*} \WW(\HH)'=\WW(\HH\SC)''\qquad\text{(\emph{duality})}. \end{equation*} Every operator $V$ on \KK which preserves the form $\kappa$ and which commutes with complex conjugation extends to a unique \emph{quasi-free endomorphism} $\varrho_V$ of \CKK: $$\varrho_V(f)=V(f),\quad f\in\KK,$$ and to a unique *-endomorphism, denoted by the same symbol, of $\WK$: $$\varrho_V(w(f))=w(V(f)),\quad f=f^*.$$ As shown in \cite{CB3}, an endomorphism $\varrho_V$ of $\WK$ can be implemented by a Hilbert space of isometries $H_{\varrho_V}$ on $\FF_{P_1}$ as in \eqref{IMP} if and only if the Hilbert--Schmidt condition \eqref{IMPCOND} holds. Such operators $V$ form a semigroup $$\END_{P_1}(\KK,\kappa)\DEF\SET{V\in\BB(\KK)}{V\+V=\1,\ \4{V}=V,\ [P_1,V] \text{ is HS}}$$ isomorphic to the semigroup of all implementable quasi-free endomorphisms of $\WK$. Here we use the notation $$A\+\DEF CA^*C$$ for the adjoint of an operator $A$ on \KK relative to the form $\kappa$. Every $V\in\END_{P_1}(\KK,\kappa)$ has a well-defined Fredholm index: $$\IND V=\dim\ker V\+\in\{0,2,4,\dots,\infty\}.$$ The \emph{statistics dimension} $d_{\varrho_V}$ of $\varrho_V$ is in the Bosonic case given by \begin{equation} \label{ccr:DRV} d_{\varrho_V}\DEF\dim H_{\varrho_V}= \begin{cases} 1, & \IND V=0,\\ \infty, & \IND V\neq 0. \end{cases} \end{equation} In order to obtain an orthonormal basis in the Hilbert space $H_{\varrho_V}\Omega_{P_1}$ associated with a fixed $V\in\END_{P_1}(\KK)$, it is again convenient to extend the partial Fock state $\omega_{P_1}\0\varrho_V^{-1}|_{\RAN\varrho_V}$ to a proper Fock state $\omega_P$ which is unitarily equivalent to $\omega_{P_1}$. The basis projection\footnote{In the Bosonic case, a basis projection $P$ is an operator on \KK such that $P=P^2=P\+=\1-\4{P}$, and such that $CP$ is positive definite on $\RAN P$.} $P$ corresponding to this new Fock state has the form $$P\DEF VP_1V\++p$$ where the basis projection $p$ of $\ker V\+$ is defined as follows \cite{CB3}. Let $E$ be the orthogonal projection onto $\ker V\+$, and let $A\DEF ECE$ be the operator describing the restriction of $\kappa$ to $\ker V\+$ with respect to the scalar product. Let $A_+$ be the positive part of $A$, i.\,e.\ the unique positive operator such that $A=A_+-\4{A_+}$ and $A_+\4{A_+}=0$. Let $A_+^{-1}$ be defined on $\RAN A_+$ as the inverse of $A_+$, and on $(\RAN A_+)^\bot$ as zero. Then $p$ is defined as $$p\DEF A_+^{-1}C.$$ The class of all Fock states over $\WK$ which are unitarily equivalent to $\omega_{P_1}$ is parameterized by symmetric\footnote{I.\,e.\ $T=\4{T^*}$.} Hilbert--Schmidt operators $T$ from $\KK_1$ to $\KK_2$ with $\NORM{T}<1$. The operator $T$ corresponding to $P$ is again given by \eqref{TDEF}: \begin{equation} \label{ccr:TDEF} T\DEF P_{21}{P_{11}}^{-1}, \end{equation} whereas $P$ can be recovered from $T$ by $$P=(P_1+T)(P_1+T\+T)^{-1}(P_1+T\+).$$ The cyclic vector $\Omega_P$ in $\FF_{P_1}$, unique up to a phase, which induces the state $\omega_P$, is given by $$\Omega_P=\bigl({\det}(P_1+T\+T)\bigr)^{1/4}\exp(-\3\4{T}a^*a^*) \Omega_{P_1}.$$ It belongs to $H_{\varrho_V}\Omega_{P_1}$. A complete orthonormal basis in $H_{\varrho_V}\Omega_{P_1}$ is obtained by applying certain isometries from the commutant of $\RAN\varrho_V$ to $\Omega_P$. The basis projection $P$ leaves $\ker V\+$ invariant, and $\kappa$ is positive definite on $\RAN P$. Let \begin{equation} \label{ccr:K} \mathfrak{k}\DEF P(\ker V\+),\quad\text{with } \dim\mathfrak{k}=\3\IND V, \end{equation} and let $g_1,g_2,\dotsc$ be a basis in $\mathfrak{k}$ such that $\kappa(g_j,g_k)=\delta_{jk}$. Let $\psi_j$ be the isometry obtained by polar decomposition of the closure of $\pi_{P_1}(g_j)$. For any multi-index $\alpha=(\alpha_1,\dots,\alpha_l)$ with $1\leq\alpha_j\leq\alpha_{j+1}\leq\3\IND V$ ($\alpha=0$ if $l=0$), set \begin{equation} \label{ccr:OA} \Omega_\alpha\DEF\psi_{\alpha_1}\dotsm\psi_{\alpha_l}\Omega_P. \end{equation} Then the $\Omega_\alpha$ form an orthonormal basis in $H_{\varrho_V}\Omega_{P_1}$ and, by \eqref{HRV}, induce an orthonormal basis in $H_{\varrho_V}$. The spaces $H_{\varrho_V}$ and $H_{\varrho_V}\Omega_{P_1}$ can both be identified with the symmetric Fock space over $\mathfrak{k}$ \cite{CB3}. \subsection{Gauge invariant quasi-free endomorphisms (CCR)} \label{sec:GICCR} We assume again that a group $G\subset\END_{P_1}(\KK,\kappa)$ consisting of unitary operators $U$ which commute with $P_1$ (so that $G$ can be identified with a subgroup of $U(\KK_1)$) acts by second quantization $\Gamma(U)$ on $\FF_{P_1}$. \emph{Gauge invariant} implementable quasi-free endomorphisms correspond to the elements of the semigroup $$\END_{P_1}(\KK,\kappa)^G\DEF\SET{V\in\END_{P_1} (\KK,\kappa)}{[V,G]=0}.$$ To determine the representation $D_{\varrho_V}$ of $G$ on the Hilbert space $H_{\varrho_V}$ associated with $V\in\END_{P_1}(\KK,\kappa)^G$, it suffices to consider the action of $\Gamma(U)$ on the vectors $\Omega_\alpha$ defined in \eqref{ccr:OA}. In contrast to the Fermionic case (cf.\ Lemma~\ref{lem:GP}), there is no simple transformation law for the $\psi_j$ under $G$. They obey however a linear transformation law when applied to $\Omega_P$; in fact, one can show that $\Omega_\alpha$ is proportional to \begin{equation} \label{ccr:OA'} \pi_{P_1}(g_{\alpha_1})\dotsm\pi_{P_1}(g_{\alpha_l}) \exp(-\3\4{T}a^*a^*)\Omega_{P_1} \end{equation} (cf.\ \cite{CB4}; taking the closures of the $\pi_{P_1}(g_j)$ is tacitly assumed here). The behavior of the $\pi_{P_1}(g_j)$ under gauge transformations is obvious. \begin{Lem} \label{ccr:lem:GEXP} Let $V\in\END_{P_1}(\KK,\kappa)^G$ be given, and let $T$ be defined by \eqref{ccr:TDEF}. Then $\exp(-\3\4{T}a^*a^*)\Omega_{P_1}$ is invariant under all gauge transformations $\Gamma(U)$, $U\in G$. \end{Lem} \begin{proof} Let $E$ be the orthogonal projection onto $\ker V\+=C\ker V^*$, and let $A\DEF ECE$ as in Section~\ref{sec:PRECCR}. Then $E$ and $A$ commute with $G$ because $V$ and $P_1$ do so. Therefore the positive part $A_+$ of $A$ and the operator ${A_+}^{-1}$ defined in Section~\ref{sec:PRECCR} also commute with $G$. It follows that $P=VP_1V\++{A_+}^{-1}C$ and $T\DEF P_{21}{P_{11}}^{-1}$ commute with $G$ as well. Arguing as in the proof of Lemma~\ref{lem:EXP}, one finds for $U\in G$ $$\Gamma(U)\bigl(-\3\4{T}a^*a^*\bigr)^n\Omega_{P_1} =\bigl(-\3(U\4{T}U\+)a^*a^*\bigr)^n\Omega_{P_1}$$ and finally $$\Gamma(U)\exp\bigl(-\3\4{T}a^*a^*\bigr)\Omega_{P_1}= \exp\bigl(-\3(U\4{T}U\+)a^*a^*\bigr)\Omega_{P_1}= \exp\bigl(-\3\4{T}a^*a^*\bigr)\Omega_{P_1}.$$ \end{proof} \begin{Thm} \label{ccr:thm:CHARGE} Let $P_1$ be a basis projection of $(\KK,\kappa)$, let $G$ be a group of unitary operators on \KK commuting with $P_1$ and with complex conjugation, and let $V\in\END_{P_1}(\KK,\kappa)^G$. Then the subspace $\mathfrak{k}$ defined in \eqref{ccr:K} is invariant under $G$, and the unitary representation $D_{\varrho_V}$ of $G$ on the Hilbert space of isometries $H_{\varrho_V}$ which implements $\varrho_V$ in the Fock representation determined by $P_1$ is unitarily equivalent to the representation $\mathfrak{S}_\mathfrak{k}$ on the symmetric Fock space over $\mathfrak{k}$ that is obtained by taking symmetric tensor powers of the representation on $\mathfrak{k}$: \begin{equation} \label{ccr:REPG} D_{\varrho_V}\simeq\mathfrak{S}_{\mathfrak{k}}. \end{equation} \end{Thm} \begin{proof} $\mathfrak{k}$ is invariant under $G$ because $\ker V\+$ is invariant and because $P$ commutes with $G$ (see the proof of Lemma~\ref{ccr:lem:GEXP}). The assertion hence follows from \eqref{ccr:OA'} and Lemma~\ref{ccr:lem:GEXP}. \end{proof} \begin{Rms} (i) Theorem~\ref{ccr:thm:CHARGE} shows that non-surjective quasi-free endomorphisms of the CCR algebra are even ``more reducible'' than endomorphisms of the CAR algebra in that they are always \emph{infinite} direct sums, a fact which explains the generic occurrence of infinite statistics in the CCR case (cf.\ \eqref{ccr:DRV}). Again, each closed subspace of $H_{\varrho_V}\Omega_{P_1}$ spanned by the $\Omega_\alpha$ with length of $\alpha$ fixed is invariant under $G$. (ii) Any representation of $G$ which is contained in $\KK_1$ with infinite multiplicity is realized on some space $\mathfrak{k}$ belonging to a $V\in\END_{P_1}(\KK,\kappa)^G$. (iii) Quasi-free automorphisms are less interesting in the CCR case because they are all neutral: $D_{\varrho_V}$ is the trivial representation of $G$ if $\IND V=0$. (iv) If the single-particle space $\KK_1$ has a particle-antiparticle symmetry, then every $V\in\END_{P_1}(\KK,\kappa)^G$ has a conjugate in $\END_{P_1}(\KK,\kappa)^G$, just as in the Fermionic case. \end{Rms} \section{Concluding Remarks} As we have seen, gauge invariant implementable quasi-free endomorphisms of the CAR and CCR algebras with statistics dimension $d\neq1$ restrict to reducible endomorphisms of the observable algebra. In typical cases, e.g.\ if $G$ is isomorphic to one of the classical compact Lie groups, any irreducible representation of the group is equivalent to a subrepresentation of some tensor power of the defining representation. In such cases there will exist quasi-free endomorphisms, behaving like ``master endomorphisms'', which contain each superselection sector as a subrepresentation. It is an interesting problem how to obtain the irreducible ``subobjects'' of a quasi-free endomorphism $\varrho$. Suppose that $\{\Psi_j\}$ is an (incomplete) orthonormal set in $H_\varrho$ which transforms irreducibly under $G$. According to the general theory \cite{DHR3}, there should exist a gauge invariant isometry $\Phi$ on Fock space with $\RAN\Phi=\oplus_j\RAN\Psi_j$. The corresponding irreducible endomorphism $\varrho_\Phi$ (which is not quasi-free) would then be given by $$\varrho_\Phi(a)\DEF\Phi^*\Bigl(\sum_j\Psi_ja\Psi_j^*\Bigr)\Phi.$$ Collections of gauge invariant isometries $\{\Phi_j\}$ fulfilling the Cuntz relations would permit to define direct sums of quasi-free endomorphisms $\{\varrho_j\}$: \begin{equation*} (\oplus_j\varrho_j)(a)\DEF\sum_j\Phi_j\varrho_j(a)\Phi_j^*, \end{equation*} so that one would get the whole Doplicher--Roberts category generated by quasi-free endomorphisms. Another important question is how to find basis-independent examples of, say, localized isometries $v$ with index one on the single-particle space \HH (see \eqref{HL2}) of the time-zero Dirac field, such that the implementability condition \eqref{Ev} holds. Recall that our construction of such an operator in Eq.~\eqref{v} made essential use of the existence of local Fourier bases on the circle. Of particular interest would be the massive case in two dimensions, where one might hope to find localized quasi-free endomorphisms obeying non-Abelian braid group statistics. However, preliminary calculations based on the explicit formulas in \cite{CB1,CB3} indicate that the commutation relations of implementers corresponding to irreducible subobjects of quasi-free endomorphisms only admit Abelian braid group statistics. \begin{acknowledgements} The author profited from discussions with P.~Grinevich on the estimates mentioned in the proof of Lemma~\ref{lem:HS}. \end{acknowledgements} \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
\section{Introduction} The recent atmospheric neutrino data from the Super-Kamiokande (SK) experiment \cite{Su99} are in excellent agreement with the hypothesis of flavor oscillations generated by nonzero neutrino mass and mixing \cite{Po67} in the $\nu_\mu\leftrightarrow\nu_\tau$ channel \cite{SKEV}. Such hypothesis is consistent with all the SK data, including sub-GeV $e$-like and $\mu$-like events (SG$e,\mu$) \cite{SKSG}, multi-GeV $e$-like and $\mu$-like events (MG$e,\mu$) \cite{SKMG}, and upward-going muon events (UP$\mu$) \cite{SKUP}, and is also corroborated by independent atmospheric neutrino results from the MACRO \cite{MACR} and Soudan-2 \cite{SOUD} experiments, as well as by the finalized upward-going muon data sample from the pioneering Kamiokande experiment \cite{KAUP}. Oscillations in the $\nu_\mu\leftrightarrow\nu_\tau$ channel are also compatible with the negative results of the reactor experiment CHOOZ in the $\nu_\mu\leftrightarrow\nu_e$ channel \cite{CHOO}. However, it has been realized that {\em dominant\/} $\nu_\mu\leftrightarrow\nu_\tau$ oscillations plus {\em subdominant\/} $\nu_\mu\leftrightarrow\nu_e$ oscillations are also consistent with SK+CHOOZ data, and lead to a much richer three-flavor oscillation phenomenology \cite{Fo99}. A detailed $3\nu$ analysis of the SK observations, including the full 33 kTy data sample, can be found in Ref.~\cite{Fo99}. Here we report and comment briefly the results of our updated analysis, based on the recent 45 kTy SK data \cite{Me99,Ha99}. The theoretical framework is based on the so-called one mass scale dominance \cite{Fo95}, which has been used also for three-flavor oscillation studies of pre-SK atmospheric and reactor neutrino data in Refs.~\cite{Fo95,Fo97,Fo98}. \section{$3\nu$ analysis of SK phenomenology (45 kTy)} In the hypothesis that the two lightest neutrinos $(\nu_1,\nu_2)$ are effectively degenerate $(m^2_1\simeq m^2_2)$ (one mass scale dominance), it can be shown \cite{Fo95,Fo99} that atmospheric neutrinos probe only $m^2\equiv m^2_3-m^2_{1,2}$, together with mixing matrix elements $U_{\alpha i}$ related to $\nu_3$, namely: \begin{equation} {\rm Parameter\ space} \;\equiv\;(m^2,U^2_{e1},U^2_{e2},U^2_{e3})\ , \end{equation} where $U^2_{e1}+U^2_{e2}+U^2_{e3}=1$ for unitarity. The unitarity constraint can be conveniently embedded in a triangle plot \cite{Fo95,Fo97,Fo99}, whose corners represent the flavor eigenstates, while the heights projected from any inner point represent the $U^2_{\alpha3}$'s. Within this framework, we analyze 30 data points, related to the zenith distributions of sub-GeV events (SG $e$-like and $\mu$-like, 5+5 bins), multi-GeV events (MG$e$,$\mu$ 5+5 bins) and upward-going muons (UP$\mu$, 10 bins), using the latest 45 kTy SK sample \cite{Me99,Ha99}. We also consider the rate of events in the CHOOZ reactor experiment \cite{CHOO} (one bin), which constrains the $\nu_e$ disappearance channel. Constraints are obtained through a $\chi^2$ statistic, as described in Ref.~\cite{Fo99}. Figure~1 shows the regions favored at 90\% and 99\% C.L.\ in the triangle plot, for representative values of $m^2$. The CHOOZ data, which exclude a large horizontal strip in the triangle, appear to be crucial in constraining three-flavor mixing. Pure $\nu_\mu\leftrightarrow\nu_e$ oscillations (right side of the triangles) are excluded by SK and CHOOZ independently. The center of the lower side, corresponding to pure $\nu_\mu\leftrightarrow\nu_\tau$ oscillations with maximal mixing, is allowed in each triangle both by SK and SK+CHOOZ data. However, deviations from maximal $(\nu_\mu\leftrightarrow\nu_\tau)$ mixing, as well as subdominant mixing with $\nu_e$, are also allowed to some extent. Such deviations from maximal $2\nu$ mixing are slightly more constrained than in the previous analysis of the 33 kTy SK data \cite{Fo99}. Figure~2 shows the constraints on the mass parameter $m^2$ for unconstrained three-flavor mixing. The best fit value is reached at $m^2\sim3\times 10^{-3}$ eV$^2$, and is only slightly influenced by the inclusion of CHOOZ data. However, the upper bound on $m^2$ is significantly improved by including CHOOZ. As compared with the corresponding plot in Ref.~\cite{Fo99} (33 kTy), this figure shows that the 45 kTy data are in better agreement with the oscillation hypothesis (lower $\chi^2$). Moreover, the favored range of $m^2$ is restricted by $\sim 10\%$ with respect to Ref.~\cite{Fo99}. Figures~1 and 2 clearly show the tremendous impact of the SK experiment in constraining the neutrino oscillation parameter space. Prior to SK, the data could not significantly favor $\nu_\mu\leftrightarrow\nu_\tau$ over $\nu_\mu\leftrightarrow\nu_e$ oscillations, and could only put weak bounds on $m^2$ (see Refs.~\cite{Fo97,Fo98}). \begin{figure} \epsfysize=14.5truecm \hspace*{0.7truecm} \epsfbox{fig1.ps} \vskip-1.3cm \caption{Three-flavor analysis in the triangle plot, for five representative values of $m^2$. Left and middle column: separate analyses of Super-Kamiokande (45 kTy) and CHOOZ data, respectively. Right column: combined SK+CHOOZ allowed regions. Although the SK+CHOOZ solutions are close to pure $\nu_\mu\leftrightarrow\nu_\tau$ oscillations, the allowed values of $U^2_{e3}$ are never negligible, especially in the lower range of $m^2$.} \label{fig1} \end{figure} \newpage \begin{figure} \epsfysize=9.truecm \phantom{.} \vspace*{-.5truecm} \hspace*{2.3truecm} \epsfbox{fig2.ps} \vskip-2.5cm \caption{Values of $\chi^2$ as a function of $m^2$, for unconstrained three-flavor mixing. Dashed curve: fit to SK data only (45 kTy). Solid curve: fit to SK+CHOOZ. The minimum of $\chi^2$ is reached for $m^2\simeq 2.8\times 10^{-3} {\rm\ eV}^2$.} \label{fig2} \end{figure} Finally, Fig.~3 shows the best-fit zenith distributions of SG$e,\mu$, MG$e,\mu$, and UP$\mu$ events, normalized to the no-oscillation rates in each bin. There is excellent agreement between data and theory, especially for the $\mu$ distributions. The nonzero value of $U^2_{e3}$ at best fit leads to a slight expected electron excess in the MG$e$ sample for $\cos\theta\to -1$. The observed electron excess is, however, somewhat larger than expected, both for SG$e$'s and for MG$e$'s. A significant reduction of the statistical error is needed to probe possible MG$e$ distortions, which would be unmistakable signals of subdominant $\nu_\mu\to\nu_e$ oscillations. \section{Outlook} The Super-Kamiokande data are compatible with three-flavor oscillations dominated by $\nu_\mu\leftrightarrow\nu_\tau$ transitions. The amplitude of the $\nu_\mu\leftrightarrow\nu_e$ channel is not necessarily zero, although being strongly constrained by both SK and CHOOZ. A contribution from the $\nu_\mu\leftrightarrow\nu_e$ channel might explain part of the electron excess observed in SK, especially for multi-GeV $e$-like events. Higher statistics is needed to test such interpretation. A definite progress in confirming the oscillation hypothesis, and in constraining the mass-mixing parameters, emerges from a comparison of the 33 kTy and 45 kTy SK data analyses.\\[2mm] {\bf Acknowledgments.} G.L.F.\ thanks the organizers of the workshop WIN'99 for kind hospitality. \newpage \begin{figure} \epsfysize=12.truecm \phantom{.} \vspace*{-2.8truecm} \hspace*{1.3truecm} \epsfbox{fig3.ps} \vskip-4.0cm \caption{SK zenith distributions of leptons at best fit (dashed lines), also including CHOOZ (solid lines), as compared with the 45 kTy experimental data (dots with error bars). The $3\nu$ mass-mixing values at best fit are indicated in the upper right corner.} \label{fig3} \end{figure} \section*{References}
\section{Introduction} \setcounter{equation}{0} In a recent paper \cite{BL} two of the present authors proposed a new superparticle model with tensorial central charges \cite{HP}--\cite{RS} and auxiliary fundamental spinor variables. An interesting peculiar feature of this model is that it describes a superparticle whose presence breaks only one of target--space supersymmetries. In all previously known cases superparticles and (in general) superbranes break half or more supersymmetries of a target superspace vacuum. As we shall show in this paper the model of \cite{BL} describes an infinite tower of massless particles of arbitrary (half)integer helicities. The model can be regarded as an extension of a Ferber--Shirafuji formulation \cite{F78,S83} of $D=4$ superparticle mechanics. In the framework of the $N=1$ Ferber--Shirafuji model one performs, at the classical level, the twistor transform \cite{F78,S83} from the $N=1$, $D=4$ superspace description of massless superfields to their description in terms of supertwistors forming a fundamental representation of a superconformal group $SU(2,2|1)$. In an analogous way the superparticle model of \cite{BL} admits the description in terms of $OSp(1|8)$ supertwistors \cite{Lstw,8}. The supergroups $SU(2,2|1)$ and $OSp(1|8)$ are not subgroups of each other, but they are different subgroups of the supergroup $OSp(2|8)$. Hence, one can assume that the Ferber--Shirafuji model and the model \cite{BL} are different reductions of an $OSp(2|8)$ supertwistor model. In this paper we construct such a generic $N=1$, $D=4$ superparticle action which depends on a numerical non--negative real parameter $a$. When the value of $a$ varies within the interval $0<a<1$ the model admits an $OSp(2|8)$ supertwistor description, while $a=0$ and $a=1$ are two critical points. At $a=0$ the model reduces to the Ferber--Shirafuji superparticle. And at $a=1$ one arrives at the $OSp(1|8)$ supertwistor model of ref. \cite{BL}. For all values of $a$ except for $a=1$ the superparticle breaks half of the target--space supersymmetries, while at $a=1$ only one supersymmetry is broken. When $a>1$ the supertwistor group becomes $OSp(1,1|8)$ which contains a noncompact group $SO(1,1)$ as a subgroup instead of $SO(2)$ in the case of $OSp(2|8)$. The (super)twistor formulation of relativistic (super)particle dynamics is useful in many aspects. Let us recall that, since the relativistic (super)particle is a constrained dynamical system not all its dynamical variables are independent. By performing (super)twistor transform we deal directly with independent physical degrees of freedom of the (super)particle in a covariant way. This, for instance, simplifies the quantization procedure and the analysis of the spectrum of quantum states of the model. We perform the quantization of the generic superparticle model for arbitrary values of the parameter $a$ and find that in $D=4$ first--quantized states of the superparticle form an infinite tower of massless states of a (half)integer helicity. We thus demonstrate that an extra dynamical (central charge) coordinate in the model under consideration has the physical meaning of a spin variable. This allows one to admit that the model considered might be related to the higher--spin field theory of Vasiliev \cite{Vasiliev} (see also relevant papers \cite{rel}). We first quantize the superparticle in the supertwistor formulation where the quantization is almost straightforward since at $a\not =0$ the supertwistor model is unconstrained and at $a=0$ there is only one first--class constraint. We then quantize the model in the $N=1$, $D=4$ superspace extended with the tensorial central charge coordinates, and show that the resulting spectrum of the quantum physical states coincides with that of the supertwistor formulation. Since in this formulation the model contains second class constraints our main tool in carrying out the quantization procedure will be the extension of the model in such a way that all the constraints of the initial model become first class constraints. This method, which can be traced back to the papers by Faddeev \& Shatashvili \cite{9,10}, Batalin, Fradkin \& Fradkina \cite{12,13} and Egorian \& Manvelian \cite{11} has already been applied to the quantization of `standard' massless superparticles by Moshe \cite{14} and Eisenberg \& Solomon \cite{15,ES,17}. The main advantage of this method is that it allows one to avoid problems with covariant splitting fermionic constraints into first and second class ones. The initial formulation of the model is recovered when we gauge fix additional gauge symmetries (associated with new first--class constraints) by putting the conversion variables to zero. In its nature the conversion method is related to an old Stueckelberg formalism \cite{18} which extends the theory of massive vector fields with an auxiliary scalar gauge degree of freedom. The quantization of the model at $a=1$ has additional peculiarities. In this case superparticle dynamics is subject to only one second--class constraint, which is quite unusual. Dynamical systems with the odd number of fermionic second class constraints are rather rare. One of few known examples is a superparticle in $D=2$ superspace with a single chiral fermion direction \cite{Sorokin87}. So the quantization of such systems is an interesting exercise by itself which requires one to deal with a single Clifford--like variable. In the case under consideration we shall use an auxiliary Clifford variable of Grassmann--odd parity to convert the single second class fermionic constraint into the one of a first class. Further we present two methods for quantizing the model with the single Clifford variable, both producing the same spectrum of first--quantized physical states. The paper is organized as follows. In Section 2 we consider the one-parameter family of actions describing the generalized $D=4$ superparticle models labelled by the real positive parameter $a\ge 0$ where the case $a=0$ corresponds to the Ferber--Shirafuji model, while at $a=1$ the action describes the superparticle model of \cite{BL}. We demonstrate that in the target space with four supersymmetries the $a\not=1$ models possess two fermionic $\kappa$-symmetries and, hence, corresponding superparticle configurations preserve 1/2 of the supersymmetries, while the $a=1$ generalized superparticle has three $\kappa$--symmetries and, hence, preserves $3\over 4$ of the supersymmetries. We also find that the $U(1)$ symmetry is inherent to the $a=0$ case only. In Section 3 we describe the transform to a supertwistor form of the action. We show that $0<a<1$ models are described by a free $OSp(2|8)$ supertwistor action and thus link the Ferber-Shirafuji $SU(2,2|1)$ supertwistor model and the free $OSp(1|8)$ supertwistor model of \cite{BL}. The model with $a>1$ is transformed into a free $OSp(1,1|8)$ supertwistor action. We perform the quantization of the supertwistor models and find that the `supertwistor' wave function describes an infinite tower of short supersymmetric multiplets of massless fields of all possible helicities. In Section 4 we extend the initial phase space of the superparticle model with auxiliary variables and perform the conversion of the initial set of first and second class constraints into the first--class constraints generating new gauge symmetries. We then carry out the quantization of the extended model and find that the dependence of the wave functions on Grassmann--odd conversion variables is inessential and can be ignored. We show that the wave function of the first--quantized model of Section 2 can be identified with the supertwistor wave function of Section 3 if Cartan--Penrose twistor formulae relating superspace and supertwistor coordinates are imposed. Thus, we find that the infinite spectrum of the first--quantized states of the superparticle consists of massless fields of an arbitrary (half)integer helicity. In Section 5 we consider a multidimensional generalization of the $a=1$ model and its quantization. It appears that after quantization the superwave function depends on only one Grassmann variable, and all other fermionic degrees of freedom can be eliminated by $N-1$ $\kappa$--transformations, where $N$ is the total number of supersymmetries. Thus, the corresponding superparticle configuration preserves the $(N-1)/N$ fraction of target--space supersymmetry. In the Appendix we analyze in detail the quantization of a supersymmetric system with one real Grassmann variable, which after quantization becomes a single Clifford variable. \section{$D=4$ model with fundamental spinor and tensorial central charge coordinates} \setcounter{equation}{0} Let us consider the following $D=4$ superparticle action \begin{equation}\plabel{ac(a)} S = \int d \tau \left( \lambda_{A}\bar{\lambda}_{\dot{B}} \Pi_\tau^{A\dot{B}} + a \lambda_{A}\lambda_{B} \Pi_\tau^{AB} + \bar a \bar{\lambda}_{\dot{A}} \bar{\lambda}_{\dot{B}} \, \Pi_\tau^{\dot{A}\dot{B}} \right)\, , \end{equation} where \begin{equation}\plabel{vielbeine} \begin{array}{l} \displaystyle \Pi^{A\dot{B}} \equiv d\tau \Pi_\tau^{A\dot{B}} = dx^{A\dot{B}} + i \left( d\Theta^{A} \bar{\Theta}^{\dot{B}} - \Theta^{A} d\bar{\Theta}^{\dot{B}}\right) \, , \\ \nonumber \displaystyle \Pi^{AB} \equiv d\tau \Pi_\tau^{AB} = d{y}^{AB} - ~i~ \Theta^{(A}~d{\Theta}^{B)}\, , \\ \displaystyle \bar{\Pi}^{\dot{A}\dot{B}} \equiv d\tau \bar{\Pi}_\tau^{\dot{A}\dot{B}} = d{\bar{y}}^{\dot{A}\dot{B}} - ~i ~\bar{\Theta}^{(\dot{A}} ~d{\bar{\Theta}}\,{}^{\dot{B})}\, , \end{array} \end{equation} $A, B=1,2$, $\dot{A},\dot{B} =1,2$ are Weyl spinor indices, and the spin--tensors $x^{A\dot{B}}$ and ${y}^{AB}$ are related to $D=4$ vector coordinates $x^m$ and antisymmetric tensorial coordinates $y^{mn}$ through the Pauli matrices \begin{equation}\label{vt} x^{A\dot{B}}=x^m\sigma_m^{A\dot{B}}, \qquad {y}^{AB}= {1 \over 2} y^{mn}(\sigma_{[m}\tilde\sigma_{n]})^{A{B}} = ( \bar{y}^{\dot{A}\dot{B}})^* \end{equation} $a$ is a numerical parameter which, without the loss of generality, can be taken to be real and positive definite $a=\bar a \in [0, \infty )$. Indeed, if $a$ is complex its phase can always be absorbed by the bosonic spinor $\l_A$ redefined in an appropriate way ($\l_A \rightarrow (\bar{a}/|a|)^{1/2} \l_A$). The action \p{ac(a)} describes a superparticle propagating in the extended superspace \begin{equation}\plabel{superspace} M^{(4+6|4)} = \{ Y^M \} \equiv \{ (x^{A\dot{A}}, y^{AB}, \bar{y}^{\dot{A}\dot{B}}; \Theta^{A}, \bar{\Theta}^{\dot{A}})\} \end{equation} {\sl with tensorial central charge coordinates} $y^{AB}, \bar{y}^{\dot{A}\dot{B}}$ \footnote{For previous consideration of different models of superparticles and p-branes in superspaces with tensorial central charges see \cite{ES,Cz,RS}.}. The configuration space of the system \begin{equation}\plabel{confsuperspace} {\cal M}^{(4+6+4|4)} = \{ q^{{\cal M}} \} \equiv \{ (Y^M; \l^A, \bar{\l}^{\dot{A}}) \} =\{ (x^{A\dot{A}}, y^{AB}, \bar{y}^{\dot{A}\dot{B}}; \l^A, \bar{\l}^{\dot{A}}; \Theta^{A}, \bar{\Theta}^{\dot{A}})\} \end{equation} contains in addition four bosonic spinor coordinates $\l^A, \bar{\l}^{\dot{A}}$. The presence of the parameter $a$ in the action \p{ac(a)} reflects the property that each of its three terms is separately invariant under global supersymmetry transformations acting on $M^{(4+6|4)}$ as follows $$ \d \Theta^A = \e^A, \qquad \d \bar{\Theta}^{\dot{B}}= \bar{\e}^{\dot{B}}, $$ \begin{equation}\plabel{susy} \d x^{A\dot{B}} = i \e^A \bar{\Theta}^{\dot{B}} - i \Theta^A \bar{\e}^{\dot{B}}, \qquad \d {y}^{AB}= ~i~ \e^{(A}~{\Theta}^{B)}, \qquad d{\bar{y}}^{\dot{A}\dot{B}}=~i ~\bar{\e}^{(\dot{A}} ~{\bar{\Theta}}\,{}^{\dot{B})}, \qquad \end{equation} $$ \d\l_A =0, \qquad \d \bar{\l}_{\dot{A}}=0. $$ The generators of the transformations \p{susy} $$ \d Y^M = i (\e^AQ_A +\bar{Q}_{\dot{A}}\bar{\e}^{\dot{A}}) Y^M $$ satisfy the supersymmetry algebra with central charges \begin{equation}\plabel{alg} \{ Q_A, Q_B\} = Z_{AB}, \qquad \{ Q_A, {\bar{Q}}_{\dot{B}} \} = - 2 P_{A\dot{B}}, \qquad \{\bar{Q}_{\dot{A}}, \bar{Q}_{\dot{B}}\} = \bar{Z}_{\dot{A}\dot{B}}, \qquad \end{equation} and all other commutators of the generators vanish. The superalgebra \p{alg} has the following realization in the superspace \p{superspace} \begin{equation}\plabel{genera} Q_A = i \partial_A - \partial_{A\dot{B}} \bar{\Theta}^{\dot{B}} - {1 \over 2} \Theta^B \partial_{AB}, \qquad \bar{ Q}_{\dot{A}} = - i \bar{\partial}_{\dot{A}} + \partial_{\dot{A}{B}} \Theta^B + {1 \over 2} \bar{\partial}_{\dot{A}\dot{B}} \bar{\Theta}^{\dot{B}}, \qquad \end{equation} \begin{equation}\plabel{gener} P_{A\dot{B}} = - i \partial_{A\dot{B}}, \qquad Z_{AB} = - {i}\partial_{AB} , \qquad \bar{Z}_{\dot{A}\dot{B}}= - {i} \bar{\partial}_{\dot{A}\dot{B}} \end{equation} $$ \partial_{A\dot{B}}\equiv { \partial \over \partial x^{A\dot{B}}}, \qquad \partial_{AB}\equiv { \partial \over \partial y^{AB}}, \qquad \bar{\partial}_{\dot{A}\dot{B}}\equiv { \partial \over \partial \bar{y}^{\dot{A}\dot{B}}}, $$ $$ \partial_{A} \equiv { \partial \over \partial \Theta^A }, \qquad \bar{\partial}_{\dot{A}}\equiv { \partial \over \partial\bar{\Theta}^{\dot{A}}}. $$ If $a=0$ the action \p{ac(a)} reduces to the Ferber--Shirafuji action \cite{F78,S83}, and if $a=1$ the action becomes the one considered in \cite{BL}. We shall see that $a=0$ and $a=1$ are `critical' values of the parameter where symmetries of the action \p{ac(a)} as well as the physical content of the model are modified. \subsection{Critical points $a=0$ and $a=1$} In order to analyze symmetry properties of the action \p{ac(a)} at different values of $a$ we consider the general variation of \p{ac(a)} which (modulo boundary terms) has the form \begin{equation}\plabel{dac(a)} \delta S = \int \left[ \d\lambda_{A} \Big(\bar{\lambda}_{\dot{B}} \Pi^{A\dot{B}} + 2a \lambda_{B} \Pi^{AB}\Big) + \d \bar{\lambda}_{\dot{A}} \Big(\Pi^{B\dot{A}}\lambda_{B} + 2{a} \bar{\lambda}_{\dot{B}} \Pi^{\dot{A}\dot{B}} \Big) \right]\, - \end{equation} $$ - \int \left[ d \Big(\lambda_{A} \bar{\lambda}_{\dot{B}} \Big) i_\d \Pi^{A\dot{B}} + a ~d \Big(\lambda_{A} \lambda_{B}\Big) i_\d \Pi^{AB} + {a} ~ d \Big(\bar{\lambda}_{\dot{A}}\bar{\lambda}_{\dot{B}}\Big) i_\d\Pi^{\dot{A}\dot{B}} \right]\, + $$ $$ + \int \left[2i \Big( d\bar{\Theta}^{\dot{B}} \bar{\lambda}_{\dot{B}} + a d{\Theta}^{{B}} \l_B \Big) \d{\Theta}^{A} \l_A + 2i \Big(d{\Theta}^{{B}} \l_B +{a} d\bar{\Theta}^{\dot{B}} \bar{\lambda}_{\dot{B}}\Big) \d\bar{\Theta}^{\dot{B}} {\bar{\l}}_{\dot{B}} \right], $$ where the basis in the space of variations of $x$ and $y$ is chosen in the form \begin{equation}\plabel{var} \begin{array}{l} \displaystyle i_\d \Pi^{A\dot{B}} \equiv \d{x}^{A\dot{B}} + i \left( \d\Theta^{A} \bar{\Theta}^{\dot{B}} - \Theta^{A} \d\bar{\Theta}^{\dot{B}}\right) \, , \\ \nonumber \displaystyle i_\d \Pi^{AB} \equiv \d{y}^{AB} - ~i~ \Theta^{(A}~\d {\Theta}^{B)}\, , \\ \displaystyle i_\d \bar{\Pi}^{\dot{A}\dot{B}} \equiv \d {\bar{y}}^{\dot{A}\dot{B}} - ~i ~\bar{\Theta}^{(\dot{A}} ~d{\bar{\Theta}}\,{}^{\dot{B})}\, . \end{array} \end{equation} \subsubsection{$U(1)$ gauge symmetry of $a=0$ model} Let us consider the variation of the action when the variations of all fields except for $\lambda_A$, $\bar\lambda_{\dot A}$ are zero \begin{equation}\plabel{lv} \delta S = \int \left[ \d\lambda_{A} \Big(\bar{\lambda}_{\dot{B}} \Pi^{A\dot{B}} + 2a \lambda_{B} \Pi^{AB}\Big) + \d \bar{\lambda}_{\dot{A}} \Big(\Pi^{B\dot{A}}\lambda_{B} + 2{a} \bar{\lambda}_{\dot{B}} \Pi^{\dot{A}\dot{B}} \Big) \right]. \end{equation} If the variation of $\lambda$ corresponds to local infinitesimal $U(1)$ rotations \begin{equation}\plabel{eqdllss} \l^\prime_A (\tau )= \l_A e^{i\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon (\tau )} , \qquad \bar{\l}^\prime_{\dot{A}}=\bar{\l}_{\dot{A}} e^{-i\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon (\tau )} \end{equation} the Eq. \p{lv} takes the form $$ \delta S = \int \left[ i\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon(\tau)\lambda_{A} \Big(\bar{\lambda}_{\dot{B}} \Pi^{A\dot{B}} + 2a \lambda_{B} \Pi^{AB}\Big) -i\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon(\tau) \bar{\lambda}_{\dot{A}} \Big(\Pi^{B\dot{A}}\lambda_{B} + 2{a} \bar{\lambda}_{\dot{B}} \Pi^{\dot{A}\dot{B}} \Big) \right]. $$ Such a variation vanishes at $a=0$. Hence at this value of $a$ the $U(1)$ transformations \p{eqdllss} describe local symmetry of the model which is inherent to the Ferber-Shirafuji formulation \cite{F78,S83} of the massless superparticle. Note that for all values of the parameter $a$ the spinors $\lambda$ are constants on the mass shell. Indeed, the equations of motion $$ {\d S \over \d {x}^{A\dot{B}}}=0, \qquad {\d S \over \d {y}^{AB}}=0, \qquad {\d S \over \d {\bar{y}}^{\dot{A}\dot{B}}}=0 $$ which follow from \p{dac(a)} have the form \begin{equation}\plabel{eqdll} d(\l_A\bar{\l}_{\dot{B}})=0 , \qquad a~d(\l_A \l_B)=0 , \qquad a~d(\bar{\l}_{\dot{A}} \bar{\l}_{\dot{B}})=0. \end{equation} In the framework of any twistor or twistor-like approach \cite{Penrose} one assumes that the bosonic spinors parametrize a projective space. This requirement does not allow $\l$ to have all its components equal to zero simultaneously. Then in the generic case $a\not=0$ eqs. \p{eqdll} imply \begin{equation}\plabel{eqdll1} d(\l_A)=0 , \qquad d(\bar{\l}_{\dot{A}})=0, \end{equation} i.e. the bosonic spinor is constant on the mass shell \begin{equation}\plabel{eqdlls} \l_A (\tau )= \l_A^0=const , \qquad \bar{\l}_{\dot{A}}=\bar{\l}_{\dot{A}}^0=const. \end{equation} When $a=0$ only one equation is left in \p{eqdll} \begin{equation}\plabel{eqdll0} a=0: \qquad d(\l_A\bar{\l}_{\dot{B}})=0 . \qquad \end{equation} The general solution of \p{eqdll0} is \begin{equation}\plabel{eqdlls0} \l_A (\tau )= \l^0_A e^{i\tilde\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon (\tau )} , \qquad \bar{\l}_{\dot{A}}=\bar{\l}^0_{\dot{A}} e^{-i\tilde\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon (\tau )}. \end{equation} The arbitrary function $\tilde\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon (\tau )$ (whose presence in \p{eqdlls0} reflects the $U(1)$ gauge symmetry of the $a=1$ model) can be gauged away by the local $U(1)$ transformation \p{eqdllss}, and we are again left with constant $\l$ on the mass shell. \bigskip \subsubsection{Fermionic variations and $\kappa$--symmetry} Let us consider now the formula \p{dac(a)} with the variations of fermionic coordinates accompanied by the following variations of $x$ and $y$ \begin{eqnarray}\plabel{varXZ} & \d{x}^{A\dot{B}}= - i \left( \d\Theta^{A} \bar{\Theta}^{\dot{B}} - \Theta^{A} \d\bar{\Theta}^{\dot{B}}\right) ~\Rightarrow & i_\d \Pi^{A\dot{B}}=0 , \nonumber \\ & \d{y}^{AB} = ~i~ \Theta^{(A}~\d {\Theta}^{B)} ~~~~~~ \Rightarrow & i_\d \Pi^{AB} =0 , \\ & \d {\bar{y}}^{\dot{A}\dot{B}} = ~i ~\bar{\Theta}^{(\dot{A}} ~d{\bar{\Theta}}^{\dot{B})} ~~~~~~ \Rightarrow & i_\d \bar{\Pi}^{\dot{A}\dot{B}} = 0 . \nonumber \end{eqnarray} The bosonic spinor $\lambda_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon$ remains unchanged. In such a case eq. \p{dac(a)} takes the form \begin{equation}\plabel{dfac(a)} \delta S = \int \left[2i d{\Theta}^{A} \l_A \Big( \d\bar{\Theta}^{\dot{B}} \bar{\lambda}_{\dot{B}} + a \d{\Theta}^{{B}} \l_B \Big) + 2i d{\bar{\Theta}}^{\dot{B}}{\bar{\l}}_{\dot{B}} \Big(\d{\Theta}^{{B}} \l_B + {a} \d\bar{\Theta}^{\dot{B}} \bar{\lambda}_{\dot{B}} \Big) \right]. \end{equation} We see that for $a\not=1$ only two out of four variations of the fermionic coordinates $\d{\Theta}^{{A}}, \d\bar{\Theta}^{\dot{A}} $ are effectively involved into the variation \p{dfac(a)} of the action \p{ac(a)}. This reflects the presence of local fermionic $\kappa$-- symmetry \cite{AL} with {\sl two} independent parameters $\kappa=(\kappa_1 + i\kappa}\def\l{\lambda}\def\L{\Lambda}\def\s{\sigma}\def\S{\Sigma_2)$, $\bar{\kappa}=(\kappa_1 - i\kappa}\def\l{\lambda}\def\L{\Lambda}\def\s{\sigma}\def\S{\Sigma_2)$. The $\kappa$--transformations of the coordinates are given by eq. \p{varXZ} and \begin{equation}\plabel{kappag} \d{\Theta}^{A} = \kappa \l^A = (\kappa_1 + i\kappa}\def\l{\lambda}\def\L{\Lambda}\def\s{\sigma}\def\S{\Sigma_2) \l^A, \qquad \d\bar{\Theta}^{\dot{B}} = \bar{\kappa}\bar{\lambda}^{\dot{B}} = (\kappa_1 - i\kappa}\def\l{\lambda}\def\L{\Lambda}\def\s{\sigma}\def\S{\Sigma_2) \bar{\lambda}^{\dot{B}}. \end{equation} At the critical point $a=1$ the number of independent $\kappa$--symmetries increases from two to {\sl three}, since in this case only one linear combination ($\d{\Theta}^{{B}} \l_B + \d\bar{\Theta}^{\dot{B}} \bar{\lambda}_{\dot{B}}$) of four real fermionic variations enters into the variation of the action \begin{equation}\plabel{dfac(1)} a= 1~:~ \qquad \delta S = \int 2i \Big(d{\Theta}^{{A}} \l_A + d\bar{\Theta}^{\dot{A}} \bar{\lambda}_{\dot{A}}\Big) ~ \Big( \d{\Theta}^{{B}} \l_B + d\bar{\Theta}^{\dot{B}} \bar{\lambda}_{\dot{B}} \Big). \end{equation} Thus remaining three fermionic variations correspond to the local fermionic symmetries of the $a=1$ model \cite{BL}. In order to present an explicit form of these {\sl three $\kappa$ symmetries} we should introduce an additional bosonic spinor $u_A$ such that \begin{equation}\plabel{hatu} \l^A {u}_A = \bar{\l}^{\dot{A}}{\bar{u}}_{\dot{A}} =1. \end{equation} Then one can perform the decomposition of the unit matrix in the spinor space \footnote{The pair of Weyl spinors $\lambda^{A}, u_{A}$ is analogous to the Newman--Penrose dyad \cite{NewmanPenrose63} widely used in General Relativity.} \begin{equation}\plabel{ddec} \d_B^{~A} = \l^A{u}_B -{u}^A \l_B, \qquad \d_{\dot{B}}^{\dot{A}}= \bar{\l}^{\dot{A}} {\bar{u}}_{\dot{B}} - {\bar{u}}^{\dot{A}} \bar{\l}_{\dot{B}} \end{equation} and use it to decompose the fermionic variation of $\Theta$. As a result we find that the $\kappa$--symmetry transformations of the $a=1$ model are given by eqs. \p{varXZ} and \begin{equation}\plabel{kappa1} \d{\Theta}^{A} = (\kappa_1 + i\kappa}\def\l{\lambda}\def\L{\Lambda}\def\s{\sigma}\def\S{\Sigma_2) \l^A + i \kappa}\def\l{\lambda}\def\L{\Lambda}\def\s{\sigma}\def\S{\Sigma_3{u}^A, \qquad \d\bar{\Theta}^{\dot{B}} = (\kappa_1 - i\kappa}\def\l{\lambda}\def\L{\Lambda}\def\s{\sigma}\def\S{\Sigma_2) \bar{\lambda}^{\dot{B}} - i \kappa}\def\l{\lambda}\def\L{\Lambda}\def\s{\sigma}\def\S{\Sigma_3{\bar{u}}^{\dot{A}}. \end{equation} \subsection{Hamiltonian analysis} We now turn to the Hamiltonian analysis of the model with the purpose of getting all the constraints on the dynamics of the system, classifying them {\it a la} Dirac and thus identifying all local symmetries of the model. For the case $a=1$ the analysis has been performed in \cite{BL}. The generic model ($a\not =0$) has the same total number of constraints as the $a=1$ model, the only difference being that when the parameter $a$ takes the value $a=1$ one of the fermionic second--class constraints becomes the first--class constraint generating the third $\kappa$--symmetry. So what we should do is just to adapt the results of the Hamiltonian analysis of \cite{BL} to the generic case. The constraints corresponding to the case $a=0$ of the Ferber--Shirafuji superparticle are obtained from the generic set of constraints by putting the canonical momenta for the central charge coordinates identically equal to zero. The canonical momenta of the generic system are \begin{equation}\label{momenta} {\cal P}_{{\cal M}} = { \partial L \over \partial \dot{q}^{{\cal M}} } = (P_{A\dot{A}}, Z_{AB}, \bar{Z}_{\dot{A}\dot{B}}; P^A, \bar{P}^{\dot{A}}; \pi_{A}, \bar{\pi}_{\dot{A}}) , \end{equation} \begin{equation}\label{brackets} [ {\cal P}_{{\cal M}}, {q}^{{\cal N}} \}_P = - (-1)^{{\cal M}{\cal N}} [{q}^{{\cal N}}, {\cal P}_{{\cal M}} \}_P= \d _{{\cal M}}^{{\cal N}} ~: \end{equation} $$ [P_{A\dot{A}}, x^{B\dot{B}}]_P = \d_{A}^{B}\d_{\dot{A}}^{\dot{B}}, \qquad [Z_{AB}, y^{CD}]_P = 2 \d_{[A}^{C}\d_{B]}^{D}, \qquad [\bar{Z}_{\dot{A}\dot{B}}, \bar{y}^{\dot{C}\dot{D}}]_P = 2 \d_{[\dot{A}}^{\dot{C}} \d_{\dot{B}]}^{\dot{D}}, \qquad $$ $$ [P^{A}, \l_{B}]_P = \d_{B}^{A}, \qquad [\bar{P}^{\dot{A}}, \bar{\l}_{\dot{B}}]_P = \d^{\dot{A}}_{\dot{B}}, \qquad $$ $$ \{\pi_{A}, \Theta^B \}_P= \d_{A}^{B}, \qquad \{ \bar{\pi}_{\dot{A}}, \bar{\Theta}^{\dot{B}} \}_P= \d_{\dot{A}}^{\dot{B}}. \qquad $$ They satisfy the following set of constraints \begin{equation}\label{Phi1} \Phi_{A\dot{B}} \equiv P_{A\dot{B}} - \lambda_{A} \bar{\lambda}_{\dot{B}} = 0 , \end{equation} \begin{equation}\label{Phi2} \Phi_{A{B}} \equiv Z_{A{B}} - a \lambda_{A} {\lambda}_{{B}} = 0 , \end{equation} \begin{equation}\label{Phi3} \bar{\Phi}_{\dot{A}\dot{B}} \equiv \bar{Z}_{\dot{A}\dot{B}} - a \bar{\lambda}_{\dot{A}} \bar{\lambda}_{\dot{B}} = 0 , \end{equation} \begin{equation}\label{PA=0} P_{A}=0, \qquad \bar{P}_{\dot{A}} = 0 , \end{equation} \begin{equation}\label{DA=0} D_{A}\equiv - \pi_A + i P_{A\dot{B}} \bar{\Theta}^{\dot{B}} + i Z_{AB} \Theta^B=0, \qquad \end{equation} \begin{equation}\label{bDA=0} \bar{D}_{\dot{A}}\equiv \bar{\pi} _{\dot{A}} - i \Theta^B P_{B\dot{A}} - i \bar{Z}_{\dot{A}\dot{B}} \bar{\Theta}^{\dot{B}} = 0. \end{equation} To separate the constraints \p{Phi1}--\p{bDA=0} into the first and second class let us project them on the bosonic spinors $\lambda$ and $u$ (Eqs. \p{hatu}, \p{ddec}). We get \begin{equation}\label{B1} B_1 = {\l}^{A} {\bar{\l}}^{\dot{B}} P_{A\dot{B}} = 0, \end{equation} \begin{equation}\label{B2} B_2 = {\l}^{A} {\bar u}^{\dot{B}} P_{A\dot{B}} - \l^A{u}^B Z_{AB} = 0, \end{equation} \begin{equation}\label{B3} B_3 \equiv (B_2)^*= {u}^{A} {\bar{\l}}^{\dot{B}} P_{A\dot{B}} - \bar{\l}^{\dot{A}} {\bar{u}}^{\dot{B}} \bar{Z}_{\dot{A}\dot{B}} = 0, \end{equation} \begin{equation}\label{B4} B_4 = 2 {u}^{A} \hat{{u}}^{\dot{B}} P_{A\dot{B}} - {1\over a}{u}^A {u}^B Z_{AB} - {1\over a}{\bar{u}}^{\dot{A}} {\bar{u}}^{\dot{B}} \bar{Z}_{\dot{A}\dot{B}} = 0, \end{equation} \begin{equation}\label{B5} B_5 = {\l}^{A} {\bar{\l}}^{{B}} Z_{AB} = 0, \end{equation} \begin{equation}\label{B6} B_6 \equiv (B_5)^*= \bar{\l}^{\dot{A}} {\bar{\l}}^{\dot{B}} \bar{Z}_{\dot{A}\dot{B}} = 0, \end{equation} \begin{equation}\label{SB12} B_7\equiv {\l}^{A}{\bar{u}}^{\dot{B}} P_{A\dot{B}} + \l^A {u}^B Z_{AB} = 0, \qquad B_8\equiv {u}^{A} {\bar{\l}}^{\dot{B}} P_{A\dot{B}} + \bar{\l}^{\dot{A}}{\bar{u}}^{\dot{B}} \bar{Z}_{\dot{A}\dot{B}} = 0, \end{equation} \begin{equation}\label{SB34} B_9\equiv {u}^A {u}^B Z_{AB} - a = 0, \qquad B_{10}\equiv {\bar{u}}^{\dot{A}} {\bar{u}}^{\dot{B}} \bar{Z}_{\dot{A}\dot{B}} -a = 0, \end{equation} \begin{equation}\label{SB56} B_{11}\equiv i(\lambda^AP_{A}-\bar\l^{\dot A}\bar{P}_{\dot{A}}) =0, \end{equation} \begin{equation}\label{SB57} B_{12}\equiv \lambda^AP_{A}+\bar\l^{\dot A}\bar{P}_{\dot{A}}=0, \end{equation} $$ B_{13}\equiv u^AP_{A}=0,\qquad B_{14}\equiv \bar{u}^{\dot A}\bar{P}_{\dot{A}} = 0. $$ \begin{equation}\label{F1} F_1 = {\l}^{A} D_{A} = 0, \end{equation} \begin{equation}\label{F2} F_2 \equiv (F_1)^*= \bar{\l}^{\dot{A}} {\bar{D}}_{\dot{A}}= 0, \end{equation} \begin{equation}\label{F3} F_3 = {u}^{A} D_{A} + {\bar{u}}^{\dot{A}} {\bar{D}}_{\dot{A}}= 0, \end{equation} \begin{equation}\label{SF1} F_4\equiv {u}^{A} D_{A} - {\bar{u}}^{\dot{A}} {\bar{D}}_{\dot{A}}= 0. \end{equation} For arbitrary $a\not =0,1$ it can be checked that the bosonic constraints \p{B1}--\p{B6} and the fermionic constraints \p{F1} and \p{F2} belong to the first class, i.e. their Poisson brackets with all constraints vanish on the constraint surface, and the constraints \p{SB12}--\p{SB57}, \p{F3} and \p{SF1} are second--class. When computing the Poisson brackets of the constraints one should take into account that, because of the normalization condition \p{hatu}, the spinor $u^A$ should be regarded as a variable depending on $\l_A$. The simplest way of taking this into account is to assume the following Poisson (actually Dirac) brackets of $u^A$ with the $\lambda$--momentum $P_A$ \footnote{These brackets appear as Dirac brackets with respect to the pair of the second class constraints \p{hatu} and $u^AP^{(u)}_A = 0$, (and their complex conjugate pair), when the bosonic spinor $u$ is considered as an independent variable whose momentum is constrained to be zero $P_A^{(u)} = 0$. Then it is not hard to verify that the new phase space variables $u^A, P^{(u)}_A$ do not introduce new redundant degrees of freedom into the system under consideration. } $$ [P_A,u^B]_P=- u^Bu_A. $$ Thus in the case $a\not = 0,1$ among 14 bosonic and 4 fermionic constraints 6 bosonic and 2 fermionic constraints are of the first class and 8 bosonic and 2 fermionic constraints are of the second class. The first class constraints generate local symmetries of the dynamical system. For instance, the constraints \p{B1}, and $(B_5+B_6)$ of \p{B5}, \p{B6} generate worldline reparametrizations of the coordinates $x$ and $y$. The fermionic constraints \p{F1} and \p{F2} generate the $\kappa$--symmetry transformations \p{varXZ} and \p{kappag}. Each first class constraint reduces the number of independent phase space variables by two, while each second class constraint eliminates only one degree of freedom. Hence, in the case $a\not =0,1$ the phase space of $2\times (4+6+4)=28$ bosonic and $2\times 4=8$ fermionic canonical variables of the system is reduced to \begin{equation}\label{H(a)} a\not= 0, 1: \qquad n_{ph}=\hbox{{\bf 8$_b$ + 2$_f$}} \end{equation} i.e. we get eight bosonic and two fermionic physical degrees of freedom. In order to see how at $a=1$ the fermionic second--class constraint \p{F3} transforms into the first--class constraint generating the third $\kappa$--symmetry \p{kappa1} let us consider the Poisson bracket of the constraint \p{F3} with itself \begin{equation}\label{F3PB} \{F_3,F_3\}_P=2(a-1). \end{equation} When $a\not =1$ the r.h.s. of \p{F3PB} is nonzero and hence this constraint is second class, but at $a=1$ the r.h.s. of \p{F3PB} vanishes. Since $F_3$ weakly commutes with all other constraints, at this critical value of $a$ we obtain one more fermionic first class constraint, and we achieve the reduction of the number of independent fermionic physical degrees of freedom from two to one \begin{equation}\label{H(a1)} a=1: \qquad n_{ph}=\hbox{{\bf $8_b$ + $1_f$}}. \end{equation} Finally, when $a=0$ the tensorial coordinates $y$ disappear from the action \p{ac(a)}, and in Eqs. \p{B1}--\p{SF1} we must put to zero their canonical momenta $Z$. The remaining set of the constraints takes the following form \begin{equation}\label{B10} B_1 = {\l}^{A} {\bar{\l}}^{\dot{B}} P_{A\dot{B}} = 0, \end{equation} \begin{equation}\label{B20} B_2 = {\l}^{A} {\bar u}^{\dot{B}} P_{A\dot{B}} = 0, \end{equation} \begin{equation}\label{B30} B_3 \equiv (B_2)^*= {u}^{A} {\bar{\l}}^{\dot{B}} P_{A\dot{B}} = 0, \end{equation} \begin{equation}\label{B40} B_4 = {u}^{A} \hat{{u}}^{\dot{B}} P_{A\dot{B}} - 2 = 0, \end{equation} \begin{equation}\label{SB560} B_{5}\equiv i(\lambda^AP_{A}-\bar\l^{\dot A}\bar{P}_{\dot{A}}) =0, \end{equation} \begin{equation}\label{SB570} B_{6}\equiv \lambda^AP_{A}+\bar\l^{\dot A}\bar{P}_{\dot{A}}=0, \end{equation} $$ B_{7}\equiv u^AP_{A}=0,\qquad B_{8}\equiv \bar{u}^{\dot A}\bar{P}_{\dot{A}} = 0. $$ \begin{equation}\label{F10} F_1 = {\l}^{A} D_{A} = 0, \end{equation} \begin{equation}\label{F20} F_2 \equiv (F_1)^*= \bar{\l}^{\dot{A}} {\bar{D}}_{\dot{A}}= 0, \end{equation} \begin{equation}\label{F30} F_3 = {u}^{A} D_{A} + {\bar{u}}^{\dot{A}} {\bar{D}}_{\dot{A}}= 0, \end{equation} \begin{equation}\label{SF10} F_4\equiv {u}^{A} D_{A} - {\bar{u}}^{\dot{A}} {\bar{D}}_{\dot{A}}= 0. \end{equation} These are the constraints of the Ferber--Shirafuji formulation of the superparticle which has been analyzed in detail in a number of papers \cite{S83,14,15,ES,17}. Now two bosonic constraints \p{B10} and \p{SB560} are first--class and other six are second--class, while, as in the generic case $a\not =1$, two of the fermionic constraints are first--class \p{F10}, \p{F20} and two are second--class \p{F30}, \p{SF10}. Therefore, the number of independent phase--space physical degrees of freedom of the standard $N=1$, $D=4$ superparticle consists of six bosonic and two fermionic variables \begin{equation}\label{H(a0)} a= 0: \qquad n_{ph}=\hbox{{\bf $6_b$ + $2_f$}}. \end{equation} In the next section we shall show that the independent phase--space physical degrees of freedom \p{H(a)}, \p{H(a1)} and \p{H(a0)} of the generic superparticle model can be covariantly described by $OSp(2|8)$ (or $OSp(1,1|8)$), $OSp(1|8)$ and $SU(2,2|1)$ supertwistors, respectively. \section{Supertwistor transform. $OSp(2|8)$, $OSp(1|8)$ and $SU(2,2|1)$ supertwistors} \setcounter{equation}{0} Let us integrate the action \p{ac(a)} by parts and neglect the boundary term. The result is \begin{equation}\plabel{ac(a)tw} S = - \int \left( \mu^A d\l_A + \bar{\mu}^{\dot{A}}d\bar{\l}_{\dot{A}}\right) - i\int \left( \chi d \bar{\chi}+ \bar{\chi}d\chi +a \chi d\chi + {a} \bar{\chi}d\bar{\chi} \right) \end{equation} or $$ S = - \int \left( \mu^A d\l_A + \bar{\mu}^{\dot{A}}d\bar{\l}_{\dot{A}}\right) - 2i\int \left( (1+a)\chi_1 d{\chi}_1+ (1-a){\chi}_2d\chi_2 \right) $$ \begin{equation}\plabel{ac(a)tw2} = - \int \left( \mu^A d\l_A + \bar{\mu}^{\dot{A}}d\bar{\l}_{\dot{A}}\right) - 2i\int \bar\chi(a)d{\chi}(a), \end{equation} where \begin{eqnarray}\plabel{Pcorr(a)b} \mu^A= x^{A\dot{B}}\bar{\l}_{\dot{B}} + 2a y^{AB}\l_B + i \Theta^A [(\bar{\Theta}\bar{\l}) + a ({\Theta}{\l})], \qquad \\ \nonumber \bar{\mu}^{\dot{A}} = \l_B x^{B\dot{A}} + 2\bar{a} \bar{y}^{\dot{A}\dot{B}} \bar{\l}_{\dot{B}} + i \bar{\Theta}^{\dot{A}} [({\Theta}{\l})+ \bar{a} (\bar{\Theta}\bar{\l})], \end{eqnarray} \begin{equation}\plabel{Pcorr(a)f} \chi = ({\Theta}{\l})\equiv {\Theta}^A {\l}_A, \qquad \bar{\chi}= (\bar{\Theta}\bar{\l})\equiv \bar{\Theta}^{\dot{A}} \bar{\l}_{\dot{A}}, \end{equation} \begin{equation}\plabel{chi} \chi_1 = {1 \over 2} ( \chi + \bar{\chi}),\qquad \chi_2 = {i \over 2} ( \bar{\chi}- \chi). \end{equation} \begin{equation}\plabel{chia} \chi(a)=\sqrt{(1+a)}\chi_1+i\sqrt{(1-a)}\chi_2,\quad \bar\chi(a)=\sqrt{(1+a)}\chi_1-i\sqrt{(1-a)}\chi_2. \end{equation} (Note that $\chi(a)$ and $\bar\chi(a)$ are complex conjugate to each other only for $a<1$, while for $a\ge 1$ they are real spinors). Thus, in the generic case $a\not=0,1$ one can reformulate the dynamical system in terms of 8 bosonic variables $\l_A, \mu^A ; \bar{\l}_{\dot{B}}, \bar{\mu}^{\dot{B}}$ and two real fermionic variables $\chi_1,~ \chi_2$. These variables can be regarded as components of a real $(8,2)$ component supertwistor (cf. with \cite{BL}) \begin{equation}\plabel{stw(a)} {Y}_{{\cal A}} = (y_1, \ldots , y_8; \chi_1,\chi_2 ) = (\l^\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon, \mu^\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon, \chi_1,\chi_2) \end{equation} where $\l^\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon$ and $\mu^\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon$ are real Majorana spinors formed of the Weyl spinors $$ \l^\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon ~\Leftrightarrow ~ \left(\matrix{\l_A \cr \bar{\l}^{\dot{A}}}\right), \qquad \mu^\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon ~\Leftrightarrow ~ \left(\matrix{\mu_A \cr \bar{\mu}^{\dot{A}}} \right) $$ One can write the action \p{ac(a)tw2} in the form \begin{equation} \plabel{actw(a)2} S = - {1 \over 2} \int d \tau {Y}_{{\cal A}} G^{{\cal A} {\cal B}} \dot{Y}_{{\cal B}} \end{equation} where \begin{equation} \plabel{OSpm(a)} G^{{\cal A}{\cal B}} = \left( \matrix{ \om^{(8)} & 0 \cr 0 & i \om^{(2)} \cr} \right) = \left( \matrix{ \left({\matrix{ 0_2 & I_2 & 0_2 & 0_2 \cr -I_2 & 0_2 & 0_2 & 0_2 \cr 0_2 & 0_2 & 0_2 & I_2 \cr 0_2 & 0_2 & {-I}_2 & 0_2 \cr }}\right) & 0 \cr 0 & i \left( \matrix{ 2(1+a) & 0 \cr 0 & 2(1-a) }\right) \cr} \right)\, . \end{equation} $\om^{(8)}$ is the $Sp(8)$ invariant simplectic metric. When $a\not = 1$ we can rescale the fermionic variables ${\chi}_1$ and ${\chi}_2$, i.e. multiply them, respectively, by $\sqrt{1+a}$ and $\sqrt{|1-a|}$. This results in the following form of the metric $\om^{(2)}$ in \p{OSpm(a)} \begin{equation} \plabel{mfer(a)+} \om^{(2)} \rightarrow ~~~ \om^{(2)}= 2 \left( \matrix{ 1 & 0 \cr 0 & 1 }\right)\qquad for~a<1 \end{equation} or \begin{equation} \label{mfer(a)-} \om^{(2)} \rightarrow ~~~ \om^{(2)} = 2 \left( \matrix{ 1 & 0 \cr 0 & -1 }\right)\qquad for~a>1. \end{equation} We see that the symmetry group of the fermionic sector of the metric \p{OSpm(a)} is $SO(2)=U(1)$ when $a<1$ and $SO(1,1)$ when $a>1$. Hence, when $a<1$ the complete symmetry group of the metric \p{OSpm(a)} is the supergroup $OSp(2|8)$, while in the case $a>1$ the symmetry group becomes $OSp(1,1|8)$. The supertwistors \p{stw(a)} transform under the fundamental representations of these supergroups. When $a=1$ the metric becomes degenerate \begin{equation} \plabel{mfer(a)1} \om^{(2)} \rightarrow ~~~ \om^{(2)}_{a=1} = 2 \left( \matrix{ 1 & 0 \cr 0 & 0 }\right). \end{equation} This reflects the absence of the second fermionic variable $\chi_2$ from the action \p{ac(a)tw2}. Thus at $a=1$ the supergroups $OSp(2|8)$ and $OSp(1,1|8)$ reduce to $OSp(1|8)$ with the corresponding supertwistor representation having one real fermionic component (cf. with \cite{BL}) \begin{equation}\plabel{stw(1)} {Y}_{{\cal A}} = (y_1, \ldots , y_8; \chi_1). \end{equation} Consider now the case $a=0$. At $a=0$ the action has the same form as for $a<1$ and hence is formally $OSp(2|8)$ invariant. But, as we have seen in the previous section, at this critical point the model acquires additional local $U(1)$ symmetry, which must have its counterpart in the supertwistor description, i.e. there should appear a first--class constraint on the supertwistor variables which generates this symmetry. In order to identify this constraint we use the defining relations \p{Pcorr(a)b} and consider the following bilinear combination of supertwistor components for an arbitrary value of $a$ \begin{equation}\plabel{U(1)(a)} \mu^A(a) \l_A - \bar{\mu}^{\dot{A}}(a) \bar{\l}_{\dot{A}} + 2i \bar{\chi} \chi= 2a \l y \l - 2\bar{a} \bar{\l} \bar{y}\bar{\l}. \end{equation} At $a=0$ eq. \p{U(1)(a)} does not involve central charge coordinates $y$ and thus we obtain the pure supertwistor constraint \begin{equation}\plabel{U(1)(0)} \mu^A(0) \l_A - \bar{\mu}^{\dot{A}}(0) \bar{\l}_{\dot{A}} + 2i \bar{\chi} \chi= 0. \end{equation} Hence, at $a=0$ for the action \p{ac(a)tw} to be equivalent to \p{ac(a)} it must be supplemented with the (first--class) constraint \p{U(1)(0)} introduced through a Lagrange multiplier term \begin{equation}\plabel{ac(0)tw} a=0: \qquad S = - \int \left( \mu^A d\l_A + \bar{\mu}^{\dot{A}}d\bar{\l}_{\dot{A}}\right) - i\int \left( \chi d \bar{\chi}+ \bar{\chi}d\chi \right) + i \int d\tau \L (\mu \l - \bar{\mu} \bar{\l} + 2i \bar{\chi} \chi) \end{equation} The constraint \p{U(1)(0)} generates the $U(1)$ gauge symmetry appearing only in the $a=0$ case. This constraint introduces the complex structure and thus breaks the $OSp(2|8)$ symmetry of the $a<1$ model down to $SU(2,2|1)$. As a result one gets the Ferber-Shirafuji formulation \cite{F78,S83} of a conventional massless superparticle \cite{casal,vt,BS} in terms of $SU(2,2|1)$ supertwistors $$ {\cal Z}_{{\cal A}} = (\l_A, \bar{\mu}^{\dot{A}}, \chi ), \qquad \bar{{\cal Z}}^{{\cal A}} = ( \mu^A, \bar{\l}_{\dot{A}}, \bar{\chi} ), $$ \begin{equation}\plabel{ac(0)tw1} a=0: \qquad S = - \int \left( \bar{{\cal Z}}^{{\cal A}}d{\cal Z}_{{\cal A}} \right) + i \int d\tau \L (\bar{{\cal Z}}^{{\cal A}}{\cal Z}_{{\cal A}} - s), \end{equation} where the constant $s$ has been introduced in order to have the possibility of describing massless superparticles with nonzero (super)helicity \cite{Penrose} (see \cite{ES,B90} and references therein for details). \subsection{Quantization of the supertwistor model} \subsubsection{Canonical supertwistor quantization} The quantization of the dynamical system \p{ac(a)tw2} with $a\not=0,1$ is quite straightforward. The action is of the first order form, therefore $\mu, \bar{\mu}$ should be identified with the canonical momenta conjugate to $\l_A, \bar{\l}_{\dot{A}},$ and $2i\bar{\chi}(a)$ is the momentum conjugate to $\chi(a)$ (remember that $\chi(a)$ and $\bar{\chi}(a)$ are defined by \p{chia}). The canonical Poisson brackets are \begin{equation}\plabel{twbrack} [\mu_A, \l^B ]_P = \d_A^{~B}, \qquad [\bar{\mu}_{\dot{A}}, \bar{\l}^{\dot{B}}]_P = \d_{\dot{A}}^{\dot{B}}, \qquad \{\bar\chi(a),\chi(a)\}_P=-{i \over 2}. \end{equation} At the quantum level the dynamical variables become operators, and the Poisson brackets are replaced by (anti)commutators ($[..., ...]_P~\rightarrow i[...,...]$, $\{ , \}_P ~\rightarrow -i\{ ,\}$).). For instance, in the `coordinate' representation the momenta are the derivatives of corresponding coordinates \begin{equation}\plabel{twqu} \mu^A = i {\partial \over \partial \l_A} , \qquad \bar{\mu}^{\dot{A}} = i {\partial \over \partial \bar{\l}_{\dot{A}}}, \qquad \bar{\chi}(a) = - {1 \over 2} {\partial \over {\partial \chi(a)}}. \end{equation} The canonical Hamiltonian of the system vanishes identically. The wave function of the system in the supertwistor `coordinate' representation is \begin{equation}\plabel{twvfa<1} a\not =1: \Phi (\l_A, \l_{\dot{A}}, \chi(a) )= \phi (\l_A, \l_{\dot{A}}) + i \chi(a) \psi (\l_A, \l_{\dot{A}}) \end{equation} and the spectrum of quantum states is described by one bosonic and one fermionic function depending on Weyl spinor variables. At $a=1$ $\chi(1)$ becomes a real Clifford variable and the field \p{twvfa<1} becomes a Clifford algebra valued function. We shall discuss this case in detail in Subsections 3.1.4, 5.2. and Appendix. To understand what kind of physical states are described by the function \p{twvfa<1} in the case $a\not=0,1$ we shall first consider the well known case $a=0$. \subsubsection{ $a=0$} At $a=0$ the dynamics of the system is subject to the first--class constraint \p{U(1)(0)} which at the quantum level is imposed on the wave function \p{twvfa<1} \footnote{In this section we basically follow the quantization procedure of references \cite{15}. The operator $D=D^{(0)} + \chi {\partial \over \partial \chi} $ is the superhelicity operator.} \begin{equation}\plabel{twD0} (D^{(0)} + \chi {\partial \over {\partial \chi} }- s) \Phi (\l_A, \bar\l_{\dot{A}}; \chi ) =0, \end{equation} where $$ D^{(0)}= \l_A {\partial \over \partial \l_A} - \bar{\l}_{\dot{A}} {\partial \over \partial \bar{\l}_{\dot{A}}} $$ is the supertwistor representation of the bosonic part of the $U(1)$ generator, $\chi=\chi(0)$ (see Eqs. \p{Pcorr(a)f}--\p{chia}) and $s$ is an integer constant which appears due to the ambiguity in ordering the operators in \p{twD0} (see \cite{Penrose,ES,B90} and refs. therein for details, here we only note that the quantization of $s$ follows from the requirement for the wave function to be single valued). The (half)integer values of $s/2$ describe helicities of massless quantum states. Let us consider first the case $s=0$. Eq. \p{twD0} requires the bosonic and fermionic components of the superfield \p{twvfa<1} to be homogeneous functions of $\l, \bar{\l}$ of the degree $0$ and $-1$, respectively, \begin{equation}\plabel{twD0c} D^{(0)} \phi (\l_A, \l_{\dot{A}}) =0, \qquad D^{(0)} \psi (\l_A, \l_{\dot{A}}) = - \psi (\l_A, \l_{\dot{A}}) \qquad \end{equation} The solution is \footnote{ A rigorous approach \cite{B90,ZF} consists in the consideration of the decomposition of the wave function $\phi (\l_A , \bar{\l}_{\dot{A}})$ in the basis of the functions on {\bf C}$^2- \{ 0\}$ formed by homogeneous infinite-differentiable functions $ \phi_{\nu_1, \nu_2} (z\l_A , \bar{z}\bar{\l}_{\dot{A}}) = z^{\nu_1}\bar{z}^{\nu_2} \phi_{\nu_1, \nu_2} (\l_A , \bar{\l}_{\dot{A}}) $ of a homogeneity index $\chi = (\nu_1 , \nu_2 )$ \cite{Gelfand}. The homogeneous functions are defined by the Mellin transformation $$ \phi_{\nu_1, \nu_2} (\l_A ,\bar{\l}_{\dot{A}}) = {i \over 2} \int dz d\bar{z}~z^{\nu_1+1}\bar{z}^{\nu_2+1} \phi (z\l_A , \bar{z}\bar{\l}_{\dot{A}}). $$ The decomposition $$ \phi (\l_A , \bar{\l}_{\dot{A}}) = \Sigma^{{}^{+\infty}}_{{}^{-\infty}} \int\limits^{+\infty}_{-\infty} d\rho \phi_{(n+i\rho )/2, (-n+i\rho )/2} (\l_A , \bar{\l}_{\dot{A}}) $$ can be substituted into Eq. \p{twD0c} instead of the power series in $\l_A$, $\l_{\dot{A}}$ to obtain the general solution. We refer the reader to \cite{B90,ZF} for further details and to \cite{Gelfand} for an excellent presentation of related mathematics, and, for simplicity, use a physical 'shortcut' of the rigorous approach. } \begin{equation}\plabel{twD0s} \phi= \phi_0 (p_m) , \qquad \psi = \bar{\l}^{\dot{A}}{\bar\psi}_{\dot{A}} (p_m) \end{equation} where, by definition, $p_m$ is a light-like vector composed of $\l$ and $\bar\l$ (see also \p{Phi1}) \begin{equation}\plabel{twqCP} p_{A\dot{A}} = p_m \s^m_{A\dot{A}}= \l_A \bar{\l}_{\dot{A}}. \qquad \end{equation} We see that the spectrum of the Ferber--Shirafuji model at $s=0$ consists of a massless $N=1$, $D=4$ (anti)chiral supermultiplet containing a complex scalar field of zero helicity and a Weyl fermion field of helicity $-{1\over 2}$. This supermultiplet can be described either by the set of bosonic and fermionic wave functions depending on the bosonic Weyl spinor variables $\lambda_{A}, \lambda_{\dot{A}}$ in accordance with the formula \p{twD0s}-\p{twqCP}, or as a set of the unrestricted bosonic scalar function $\phi_0 (p_m)$ and the fermionic spinor function $\bar{\psi}_{\dot{A}}(p_m)$ depending on the light-like vector $p_mp^m=0$ which we identify with the momentum of the massless superparticle. In such a way we establish the relation of the supertwistor formulation with the space--time description of the massless superparticle, and this dual description can be extended to the case of more general model with nonvanishing central charge coordinates. Finally, let us consider the case of the nonvanishing operator ordering constant $s$ in \p{twD0} which we shall call the superhelicity parameter, characterizing the helicity properties of the superfield solutions. The component form of the constraint \p{twD0} now reads \begin{equation}\plabel{twD0cs} D^{(0)} \phi (\l_A, \l_{\dot{A}}) = s \phi (\l_A, \l_{\dot{A}}), \qquad D^{(0)} \psi (\l_A, \l_{\dot{A}}) = (s-1) \psi (\l_A, \l_{\dot{A}}) \qquad \end{equation} For integer $s>0$ the solution of \p{twD0cs} is \begin{equation}\plabel{twD0ss} \phi= \l^{A_1} \ldots \l^{A_s}\phi_{A_1 \ldots A_s} (p_m) , \qquad \psi = \l^{A_1} \ldots \l^{A_{s-1}} {\psi}_{A_1 \ldots A_{s-1}} (p_m) \end{equation} We thus obtain supermultiplets whose components have the helicities $s/2$ and $s/2-1/2$, respectively. The choice of the statistics of the superfields \p{twD0ss} should be made in accordance with the general spin--statistics theorem, such that for the even values of $s$ (integer superhelicities) the superfields \p{twD0ss} are bosonic and for odd $s$ (half-integer superhelicities) they are fermionic. Notice that the Grassmann parity of the superfield $\Phi$ \p{twD0} (and its components $\phi$ and $\psi$) is related to the parity of $\Phi(\lambda,\bar\l,\chi)$ under the change $\lambda ~\rightarrow~-\l$ ($\lambda$--parity) which implies $\chi~\rightarrow~-\chi$. If $\Phi(- \lambda,- \bar\l,- \chi)= \Phi(\lambda,\bar\l,\chi)$ then from \p{twD0ss} follows that $s$ is even (integer superhelicities) and such a superfield is Grassmann--even ($\phi$ is bosonic and $\psi$ is fermionic). Analogously if the superfield $\Phi(\lambda,\bar\l,\chi)$ changes the sign under the $\l$--parity, then $s$ is odd (half-integer superhelicities) and the superfield is Grassmann--odd ($\phi$ is fermionic, and $\psi$ is bosonic). For integer $s<0$ the solution of \p{twD0cs} is \begin{equation}\plabel{twD0ss1} \phi= \l^{\dot{A}_1} \ldots \l^{\dot{A}_{-s}} \bar{\phi}_{\dot{A}_1\ldots \dot{A}_{-s}} \qquad \psi = \l^{\dot{A}_1} \ldots \l^{\dot{A}_{-s+1}} \bar{\psi}_{\dot{A}_1\ldots \dot{A}_{-s+1}} \end{equation} and thus the spectrum of the quantum states of the model is represented by a supermultiplet of helicity $(-s/2,(-s+1)/2)$. \subsubsection{$a\not = 0,1$} Let us return to the generic $a\not= 0,1$ models. Their spectrum is defined by arbitrary scalar bosonic and fermionic functions of the Weyl bosonic spinors $\phi (\l_A, \l_{\dot{A}})$ and $\psi (\l_A, \l_{\dot{A}})$ which, in contrast to the $a=0$ case, are not subject to any constraints. The bosonic spinor components can be regarded to be defined through the components of the light-like vector $p_mp^m=0$ \p{twqCP} up to the phase transformations \begin{equation}\plabel{phase} \l_A ~\rightarrow~ e^{i\alpha (\tau)}\l_A, \qquad \l_{\dot{A}}~ \rightarrow~ e^{-i\alpha (\tau)}\l_{\dot{A}}. \end{equation} Thus for $a\not=0$ we can consider the bosonic and fermionic wave functions to depend on the light-like vector and a $U(1)$ angle variable $\alpha \sim \alpha + 2\pi k$ \begin{equation}\plabel{twa<1rep} \phi (\l_A, \l_{\dot{A}})= \phi (p_m, \alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon ), \qquad \psi (\l_A, \l_{\dot{A}})= \psi (p_m, \alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon ). \end{equation} Hence, in contrast to the Ferber--Shirafuji model, the wave function of the generic dynamical system \p{ac(a)tw} with $a\not=0$ depends on {\sl one additional variable} which {\sl parametrizes a compact manifold} $U(1)=S^1$. This means that the functions $\phi$ and $\psi$, as the single valued functions, can be expanded in the Fourier series \begin{equation}\plabel{twa<1ser0} \phi (p_m, \alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon )= \S_{k \in {\hbox{{\bf Z}}} } e^{ik\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon } \phi_k (p_m), \qquad \psi (p_m, \alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon )= \S_{k \in {\hbox{{\bf Z}}}} e^{ik\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon } \psi_k (p_m). \end{equation} The meaning of this series expansion becomes clear if we use the Lorentz--covariant representation of $\phi$ and $\psi$ as single valued functions of $\l_A$ \p{twvfa<1} where $\l_A\bar\l_{\dot A}$ are replaced by $p_m$. Then the series \p{twa<1ser0} acquires the form \begin{equation}\plabel{sphi} \phi (\l, \bar{\l} )= \phi^0 (p_m)+ \S_{k \in {\hbox{{\bf Z}}_+}} \left(\l^{A_1}\ldots \l^{A_k} \phi_{A_1\ldots A_k} (p_m) + \bar{\l}^{\dot{A}_1} \ldots \bar{\l}^{\dot{A}_k} \bar{\phi}_{\dot{A}_1 \ldots \dot{A}_k} (p_m) \right), \end{equation} $$ \psi (\l, \bar{\l} )= \psi^0 (p_m)+ \S_{k \in {\hbox{{\bf Z}}_+}} \left(\l^{A_1}\ldots \l^{A_k} \psi_{A_1\ldots A_k} (p_m) + \bar{\l}^{\dot{A}_1} \ldots \bar{\l}^{\dot{A}_k} \bar{\psi}_{\dot{A}_1 \ldots \dot{A}_k} (p_m) \right). $$ We therefore conclude that the most general solution of the model with $a\not = 0$ describes an infinite doubly degenerate spectrum of massless fields of an arbitrary helicity, with the additional compact $S^1$--coordinate in the momentum space conjugate to the discrete helicity variable. \bigskip If we assume the validity of spin--statistics theorem the bosonic fields should have positive $\lambda$--parity, and fermionic fields should have odd $\lambda$--parity. Thus, the $\l$-even part $ \Phi_+ (\l, \bar{\lambda}, \chi) \equiv \Phi_+ (-\l, -\bar{\lambda}, -\chi) = \phi_+ (\lambda, \bar{\lambda}) + i \chi \psi_-(\lambda, \bar{\lambda}) $ of the general superfield solution $ \Phi (\l, \bar{\lambda}, \chi) = \phi(\lambda, \bar{\lambda}) + i \chi \psi (\lambda, \bar{\lambda}) $ (see \p{twvfa<1} and \p{sphi}) should be regarded as bosonic (i.e. Grassmann--even). Consequently this implies that the wave function $\phi_{+}(\lambda, \bar{\lambda})$ has positive $\lambda$--parity (even powers of $\lambda$) i.e. it is bosonic, and the fermionic wave function $\psi_{-}(\lambda, \bar{\lambda})$ has negative $\lambda$--parity (odd powers of $\lambda$). Another sector of the full quantum state spectrum is described by the fermionic $\l$--odd superfield $ \Phi_- (\l, \bar{\lambda}, \chi) \equiv - \Phi_- (-\l, -\bar{\lambda}, -\chi) = \phi_- (\lambda, \bar{\lambda}) + \chi \psi_+(\lambda, \bar{\lambda}) $ which is composed of the fermionic $\l$-odd field $\phi_- (\lambda, \bar{\lambda})$ and the bosonic $\l$--even field $\psi_+ (\lambda, \bar{\lambda})$. We, therefore, see that in order to obtain physically meaningful solutions described by the superfields $\Phi_{+} (\lambda,\bar{\lambda}, \chi)$ and $\Phi_{-} (\lambda, \bar{\lambda},\chi)$ with definite Grassmann parity one should divide the general solution \p{twvfa<1} into two parts with even and odd $\lambda$--parity. Note that the superfield solutions with definite even/odd $\lambda$--parity have the even/odd superhelicities, but each of them contains a complete nondegenerate spectrum of states with both even and odd helicities. It is instructive to compare the consequences of the presence of the ``internal'' compact coordinate in our case and in Kaluza--Klein theories. In the Kaluza--Klein theories the compact variables arise in an extension of space--time with extra directions and lead to the quantization of corresponding ``internal'' momenta in the extended momentum space. The ``internal'' quantized momenta describe masses and gauge charges of Kaluza--Klein fields in the dimensionally reduced theory \footnote{It is worth mentioning that `usual' Lorentz--scalar central charges can be interpreted as Kaluza-Klein momenta \cite{ALcch}}. In our case the compactification is achieved by expressing the generalized momenta in terms of bosonic spinor (twistor) componenta. Thus, we have the opposite situation: the compact ``internal'' manifold is in the extended {\sl momentum} (twistor) space and a {\sl quantized (discrete)} central charge coordinate is in the extended {\it coordinate} space (space--time + central charge coordinates). The Fourier transform of the compact ``internal" momentum results in the discrete values of the conjugate coordinate, which are described by an integer $s$. From the physical point of view the (half)integer number $s/2$ describes the possible helicities of the massless quantum states. The quantum states of our model form a reducible (infinite--dimensional) representation of target space supersymmetry. Indeed, as the bosonic spinor is inert under global supersymmetry, the fields \p{sphi} can be collected into the superfield series expansion with each term having definite superhelicity \begin{equation}\plabel{sPhi} \Phi (\l^{A}, \bar{\l}^{\dot{A}}, \chi(a) )= \Phi^0 (p_m,\chi (a) ) + \S_{k \in {\hbox{{\bf Z}}_+}} \l^{A_1}\ldots \l^{A_k} \Phi_{A_1 \ldots A_k} (p_m,\chi(a)) + \end{equation} $$ + \S_{k \in {\hbox{{\bf Z}}_+}} \bar{\l}^{\dot{A}_1} \ldots \bar{\l}^{\dot{A}_k} \bar{\Phi}_{\dot{A}_1 \ldots \dot{A}_k} (p_m, \chi(a)) . $$ It is easy to see that each term is separately invariant under supersymmetry \footnote{Remember that the supersymmetry transformations of the superfield \p{sPhi} are generated by the transformations of $\chi$ as functions of $\Theta$ (eqs. \p{Pcorr(a)f}).}. In the case $a=0$ the additional $U(1)$ constraint \p{twD0} appears. It singles out one irreducible superfield with a definite superhelicity out of the infinite series \p{sPhi}. \subsubsection{$a=1$} Consider now a peculiarity of the model at $a=1$. In this case the action \p{ac(a)tw}, \p{ac(a)tw2} contains only one real fermionic variable $\chi_1$. The corresponding term in the action is \begin{equation}\plabel{chi1} S_\chi=-4i\int \chi_1d\chi_1 \qquad \chi_1 = {1\over 2}(\Theta \l+ \bar{\Theta}\bar{\l}). \end{equation} From \p{chi1} we conclude that the odd momentum of $\chi$ is proportional to $\chi$ itself \begin{equation}\plabel{S1} S_{\chi} = \pi_{\chi_1} -4i\chi_1 =0. \end{equation} Eq. \p{S1} is the second--class constraint being typical of any free fermion theory \begin{equation}\plabel{S1S1} \{ S_\chi, S_\chi \}_P = -8i. \end{equation} It can be regarded to be satisfied in the strong sense \cite{Dirac} after we pass from the Poisson brackets to the Dirac brackets \begin{equation}\plabel{DB} [ f, g\}_D = [ f, g\}_P -{i \over 8}[ f , S_\chi \}_P [ S_\chi, g \}_P \end{equation} which imply \begin{equation}\plabel{xixiD} \{ \chi_1, \chi_1 \}_D = 2i. \end{equation} Hence, upon quantization $\xi$ becomes a Clifford variable of odd Grassmann parity \begin{equation}\label{cliff} (\hat{\chi_1})^2=1. \end{equation} The Clifford algebra generated by this variable consists of two elements, the unit element and $\hat{\chi_1}$. Hence, all functions of $\hat{\chi_1}$ can be written as a `Clifford algebra valued superfield' having two components \cite{Sorokin87} \begin{equation}\plabel{csf} \Phi(\hat{\xi})=\phi + i\hat{\chi_1}\psi, \end{equation} where $\phi$ and $\psi$ do not depend on $\hat{\chi_1}$. We conclude that {\sl at $a=1$ the wave functions \p{twvfa<1}, \p{sPhi} become Clifford `superfields' whose components again (as in the case $a\not=0,1$) describe an infinite tower of fields of all possible helicities}. We can decompose the superfields \p{csf} into the even and odd parts with respect to $\l$-parity and thus have the wave functions with definite Grassmann parity (bosonic and fermionic superfields). We see that at $a=1$ the model has the same spectrum of quantum physical states as in the generic case. The difference with the generic case ($a\neq 1$) is only in the transformation properties of the field components with respect to target space supersymmetry -- the $a=1$ supersymmetry multiplets are shortened (see [1], Section 2). Note also that in the models with $a\ge 1$ one can impose additional reality condition on the quantum wave functions. In the Appendix we shall present another way of quantizing a single classical fermionic variable $\chi_{1}$ (see \p{chi1}) which allows to treat it as a usual Grassmann variable and quantum superwave functions as standard superfields. \section{Quantization by using the conversion method} \setcounter{equation}{0} In order to justify the results of the supertwistor quantization of the model presented in Sect. 3 and to clarify the space--time structure of the quantum wave functions, in this section we shall perform the quantization directly in the coordinate representation. Because of the appearance of a particular mixture of fermionic first and second class constraints there appears a problem of quantizing the system covariantly. However, there exists a powerful method to handle this problem \cite{11,12,13}, which is based on the conversion of the second class constraints into the first class ones. The quantization of the Ferber-Shirafuji model by the conversion method was considered in \cite{14,15,ES,17}. In \cite{ES} a $D=10$ supersymmetric particle with extra tensorial coordinates has been also discussed. In the present paper, however, the relation between spinor variables and the tensorial central charges, as well as their physical interpretation, goes far beyond the results presented in \cite{ES}. \bigskip \subsection{Conversion degrees of freedom} To convert the second class constraints into the first class ones we introduce additional (conversion) phase space degrees of freedom, whose number is equal to the the number of the second class constraints. Thus, for the $a\not= 0,1 $ models we need {\bf 8}$_b$ $+$ {\bf 2}$_f$ conversion degrees of freedom. For this purpose we introduce bosonic spinors $\rho_A$, $\bar{\rho}_{\dot{A}}$ plus its canonical momenta $P_{(\rho )}^A$, $\bar{P}_{\bar{(\rho})}^{\dot{A}}$ \begin{equation}\plabel{brcvb} [P_{(\rho )}^A, \rho_{B}]_P = \d_{B}^{~A}, \qquad [\bar{P}_{\bar{\rho}}^{\dot{A}}, \bar{\rho}_{\dot{B}}]_P = \d^{~\dot{A}}_{\dot{B}}, \qquad \end{equation} and two real fermionic variables $f_1$ and $f_2$ whose Poisson brackets form a Clifford algebra \begin{equation} \plabel{SS1,2} \{f_1, f_1 \}_P = -i (1-a), \qquad \{f_1, f_2 \}_P = 0, \qquad \{f_2, f_2 \}_P = -i (1+a). \qquad \end{equation} Instead of $f_1$ and $f_2$ we shall also use two conjugate fermionic variables \begin{equation} \plabel{tS} {S}= \sqrt{(1+a)} f_1 - i\sqrt{(1-a)}f_2 , \qquad \bar{{S}}= \sqrt{(1+a)} f_1 + i\sqrt{(1-a)}f_2, \qquad \end{equation} \begin{equation} \plabel{tStS} \{ {S},{S}\}_P = 0, \qquad \{\bar{{S}}, \bar{{S}} \}_P = 0, \qquad \{ {{S}}, \bar{{S}} \}_P = -2i (1-a^2). \end{equation} Note, that $\bar{{S}}$ is complex conjugate of ${{S}}$ only for the case $0<a<1$, while for $a>1$, where $ \sqrt{(1-a)}= i\sqrt{|1-a|}$ is imaginary, both ${S}$ and $\bar{{S}}$ are independent real variables. For $a\not=1$ $\bar{{S}}$ can be regarded as the momentum conjugate to ${{S}}$. \bigskip \subsection{Conversion of the constraints} We use the additional degrees of freedom \p{brcvb} and \p{SS1,2} in order to convert the mixture of first and second class constraints \p{Phi1}--\p{bDA=0} into the first class ones. As it was shown in \cite{BMRS}, in twistor-like formulations of particle mechanics it is convenient to perform conversion of the whole set of primary constraints, without dividing them into the sets of first and second class constraints. For any $a$ the first class constraints obtained as the result of conversion are \footnote{We denote the converted constraints with the same letters as the original ones.} \begin{equation}\plabel{tPhi1} {\Phi}_{A\dot{B}} \equiv {P}_{A\dot{B}} - (\lambda_{A} + \rho_A ) (\bar{\lambda}_{\dot{B}} + \bar{\rho}_{\dot{B}} ) = 0 , \end{equation} \begin{equation}\plabel{tPhi2} {\Phi}_{A{B}} \equiv Z_{A{B}} - a (\lambda_{A} + \rho_A ) (\lambda_{B} + \rho_B ) = 0 , \end{equation} \begin{equation}\plabel{tPhi3} {\bar{\Phi}}_{\dot{A}\dot{B}} \equiv \bar{Z}_{\dot{A}\dot{B}} - a (\bar{\lambda}_{\dot{A}} + \bar{\rho}_{\dot{A}} ) (\bar{\lambda}_{\dot{B}} + \bar{\rho}_{\dot{B}} ) = 0 , \end{equation} \begin{equation}\plabel{tPA=0} {{\Phi}}_{{A}}\equiv P^{A}_{(\lambda )} + P^{A}_{(\rho )} =0, \qquad {\bar{\Phi}}_{\dot{A}}\equiv \bar{P}^{\dot{A}}_{(\bar{\lambda})}+\bar{P}^{\dot{A}}_{(\bar{\rho})} = 0 , \end{equation} \begin{equation}\plabel{tPrho=0} P^{A}_{(\lambda )}- P^{A}_{(\rho )} =0, \qquad \bar{P}^{\dot{A}}_{(\bar{\lambda})} - \bar{P}^{\dot{A}}_{(\bar{\rho})} = 0 , \end{equation} \begin{equation}\plabel{tDA=0} {D}_{A}\equiv - \pi_A + i P_{A\dot{B}} \bar{\Theta}^{\dot{B}} + i Z_{AB} \Theta^B + (f_1+if_2) ~(\lambda_{A} + \rho_A ) =0, \qquad \end{equation} \begin{equation}\plabel{tbDA=0} {\bar{D}}_{\dot{A}}\equiv \bar{\pi} _{\dot{A}} - i \Theta^B P_{B\dot{A}} - i \bar{Z}_{\dot{A}\dot{B}} \bar{\Theta}^{\dot{B}} + (f_1-if_2)~ (\bar{\lambda}_{\dot{A}} + \bar{\rho}_{\dot{A}} )= 0. \end{equation} The algebra of the first class constraints \p{tPhi1}--\p{tbDA=0} is quite simple. The only nonvanishing brackets (in the strong sense) appear in the fermionic sector and have the form \begin{equation} \label{DDD} {{\cal D}}_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} \equiv \left( \matrix{ \{{ D}_{A},{ D}_{B} \}_P & \{ { D}_{A},{\bar{D}}_{\dot{B}} \}_P \cr \{ {\bar{D}}_{\dot{A}},{D}_{{B}} \}_P & \{ {\bar{D}}_{\dot{A}},{\bar{D}}_{\dot{B}} \}_P \cr }\right) = 2i \left( \matrix{ -{\Phi}_{AB} & {{\Phi}}_{A\dot{B}} \cr {{\Phi}}_{B\dot{A}} & - { \bar{\Phi}}_{\dot{A}\dot{B}} \cr }\right). \end{equation} The r.h.s. of \p{DDD} vanishes weakly, i.e. on the constraint surface \p{tPhi1}--\p{tPhi3}. Note that the expressions \p{tPhi1}--\p{tPhi3}, \p{tDA=0} and \p{tbDA=0} contain only the combination ($\l+\rho$) of the commuting spinors. We denote this combination by $\tilde\l$ \begin{equation} \label{tlambda} \tilde{{\l}}_{A}= \lambda_{A} + \rho_A , \qquad \tilde{\bar{\l}}_{\dot{A}} = \bar{\lambda}_{\dot{A}} + \bar{\rho}_{\dot{A}} , \end{equation} while the linearly independent variables \begin{equation} \label{trho} \tilde{{\rho}}_{A}= \lambda_{A} - \rho_A, \qquad \tilde{\bar{\rho}}_{\dot{A}} = \bar{\lambda}_{\dot{A}} - \bar{\rho}_{\dot{A}} \end{equation} completely decouple and have vanishing canonical momenta \p{tPrho=0}. Hence, the variables \p{trho} can be excluded from the consideration, since the wave functions will not depend on these variables. \subsection{Quantization of the converted system: equations for the wave function} Now it is straightforward to quantize the system by using the Dirac method \cite{Dirac}. For this purpose let us choose the (super-)Shr\"{o}dinger representation for the superspace coordinates and the bosonic spinor variables \begin{equation}\plabel{hatP4D} \hat{P}_{A\dot{B}} = - i {\partial \over \partial X^{A\dot{B}}}, \qquad \hat{Z}_{A{B}} = - { i} {\partial \over \partial y^{A{B}}}, \qquad \hat{\bar{Z}}_{\dot{A}\dot{B}} = - i {\partial \over \partial \bar{y}^{\dot{A}\dot{B}} }, \qquad \end{equation} $$ \hat{P}^{A}_{\tilde{\l} }= - i {\partial \over \partial \tilde{\l}^{A}}, \qquad \hat{\bar{P}}^{\dot{A}}_{\tilde{\bar{\l}} }= - i {\partial \over \partial \tilde{\bar{\l}}^{\dot{A}}}, \qquad $$ \begin{equation}\plabel{hatpi4D} \hat{\pi}_{A} = + i {\partial \over \partial \Theta^{A}}, \qquad \hat{\bar{\pi}}_{\dot{A}} = + i {\partial \over \partial \bar{\Theta}^{\dot{A}}}. \qquad \end{equation} The fermionic variables $f_1$ and $f_2$ become Clifford algebra operators \begin{equation} \plabel{brcvfq} ({\hat{f}_1})^2 = {1\over 2}(1-a) , \qquad ({\hat{f}_2})^2 = {1\over 2}(1+a) , \qquad \{\hat{f}_1,\hat{f}_2 \} = 0. \qquad \end{equation} The Grassmann parity of $f_1$ and $f_2$ must be odd because the constraints \p{tDA=0} and \p{tbDA=0} should have definite parity. Note that the linear combinations \p{tS} of fermionic quantum variables $f_1$ and $f_2$ satisfy the commutation relations \begin{equation} \plabel{qtStS} \{ \hat{S}, \hat{S} \} = 0, \qquad \{ \bar{\hat{S}}, \bar{\hat{S}} \} = 0, \qquad \{ {\hat{S}}, \bar{\hat{S}} \} = 2 (1-a^2). \end{equation} So one can choose $S$ as an odd coordinate and $\bar{\hat{S}}$ as its momentum operator $$\hat{\bar{S}}=2(1-a^2){\partial\over{\partial {S}}}, \qquad \hat{S}=S$$ Despite of the fact that such a representation makes hermiticity condition nonmanifest, it is convenient since it simplifies the calculations and provides the possibility of treating the cases $0<a<1$ and $a>1$ on an equal footing. \bigskip After quantization the first--class constraints \p{tPhi1}-\p{tPA=0}, \p{tDA=0} and \p{tbDA=0} are imposed on the wave function \begin{equation} \plabel{wf4D} \Psi = \Psi (x^{A\dot{A}}; y^{AB}, \bar{y}^{\dot{A}\dot{B}}; \tilde{\l}^A, \tilde{\bar{\l}}^{\dot{A}}; \Theta^A, \bar{\Theta}^{\dot{A}}, S) \end{equation} (we recall that we have consistently removed the variables \p{trho} from the consideration). Thus, the wave function of the system satisfies the first order differential equations \begin{equation}\plabel{tPhi1q} ({\partial \over \partial x^{A\dot{B}}} - i \tilde{\lambda}_{A} \tilde{\bar{\lambda}}_{\dot{B}} ) \Psi = 0 , \end{equation} \begin{equation}\plabel{tPhi2q} ({1 \over 2} {\partial \over \partial y^{A{B}}} - ia\tilde{\lambda}_{A}\tilde{\lambda}_{B}) \Psi = 0 , \end{equation} \begin{equation}\plabel{tPhi3q} ({1 \over 2} {\partial \over \partial \bar{y}^{\dot{A}\dot{B}}} - i a \tilde{\bar{\lambda}}_{\dot{A}} \tilde{\bar{\lambda}}_{\dot{B}}) \Psi = 0 , \end{equation} \begin{equation}\plabel{tDA=0q} \left({\partial \over \partial \Theta^A} + i {\partial \over \partial x^{A\dot{B}}} \bar{\Theta}^{\dot{B}} + {i \over 2} a {\partial \over \partial y^{A{B}}}\Theta^B - i (f_1+if_2) \tilde{\lambda}_{A}\right) \Psi = 0 , \end{equation} \begin{equation}\plabel{tbDA=0q} \left({\partial \over \partial \bar{\Theta}^{\dot{A}}} + i {\partial \over \partial x^{B\dot{A}}} \Theta^B + {i \over 2} a {\partial \over \partial \bar{y}^{\dot{A}\dot{B}}} \bar{\Theta}^{\dot{B}} +i (f_1-if_2)\bar{\lambda}_{\dot{A}}\right) \Psi = 0 , \end{equation} The solution of eqs. \p{tPhi1q}--\p{tPhi3q} is \begin{equation} \plabel{wf4Ds1} \Psi = e^{ i \l_A \bar{\lambda}_{\dot{A}} x^{A\dot{A}} + ia \l_A\l_By^{AB} + ia \bar{\lambda}_{\dot{A}}\bar{\lambda}_{\dot{B}} \bar{y}^{\dot{A}\dot{B}}} ~~~ g( \tilde{\l}^A, \tilde{\bar{\l}}^{\dot{A}}; \Theta^A, \bar{\Theta}^{\dot{A}}, S). \end{equation} Because of the constraints \p{tDA=0q} and \p{tbDA=0q} the function $g$ \p{wf4Ds1} satisfies the conditions \begin{equation}\plabel{tDA=0q1} \left({\partial \over \partial \Theta^A} - \tilde{\lambda}_{A} [(\bar{\Theta}\bar{\lambda}) - a ({\Theta}{\lambda}) - i(f_1+if_2) ]\right) g = 0 , \end{equation} \begin{equation}\plabel{tbDA=0q1} \left({\partial \over \partial \bar{\Theta}^{\dot{A}}} - \bar{\lambda}_{\dot{A}} [({\Theta}{\lambda}) - a (\bar{\Theta}\bar{\lambda}) + i (f_1-if_2)]\right) g = 0 , \end{equation} An evident consequence of eqs. \p{tDA=0q1} and \p{tbDA=0q1} is that $g$ depends only on the composite Grassmann variables $\chi=\Theta^B\l_B$ and $\bar\chi= \bar{\Theta}^{\dot{B}}\bar{\l}_{\dot{B}}$ introduced in \p{Pcorr(a)f} \begin{equation} \plabel{wf4Ds2} g( \tilde{\l}^A, \tilde{\bar{\l}}^{\dot{A}}; \Theta^A , \bar{\Theta}^{\dot{A}}, S)= g( \tilde{\l}^A, \tilde{\bar{\l}}^{\dot{A}}; \Theta^B\l_B , \bar{\Theta}^{\dot{B}}\bar{\l}_{\dot{B}}, S). \end{equation} Then the eqs. \p{tDA=0q1} and \p{tbDA=0q1} reduce to \begin{equation}\plabel{tDA=0q2} \left({\partial \over \partial \chi} - \bar{\chi } - a \chi - i ({f_1}+if_2) \right) ~g(\l_A, \bar{\l}_{\dot{A}}; \chi , \bar{\chi }, S) = 0 , \end{equation} \begin{equation}\plabel{tbDA=0q2} \left({\partial \over {\partial \bar{\chi}}} - {\chi}- a\bar{\chi} + i ({f_1}-if_2) \right) ~g (\l_A, \bar{\l}_{\dot{A}}; \chi , \bar{\chi }; S) = 0. \end{equation} \subsection{Quantization of the converted system: dependence on the fermionic variables} To find the solution of equations \p{tDA=0q2} and \p{tbDA=0q2} we take their linear combinations and rewrite them in the following form \begin{equation}\plabel{dchiS} \left[\sqrt{1-a^2}({\partial\over{\partial\bar\chi(a)}} -\chi(a))-S\right]g=0, \end{equation} \begin{equation}\plabel{dchibS} \left[\sqrt{1-a^2}({\partial\over{\partial{\chi(a)}}}-\bar\chi(a))-2(1-a^2) {\partial\over{\partial S}}\right]g=0, \end{equation} where $\chi(a)$ were introduced in \p{chi}, \p{chia} and $S$ are defined in \p{tS}. The equations \p{dchiS} and \p{dchibS} are easily solved in terms of the components of the superfunction $g(S)$ \begin{equation}\plabel{gs} g(S)=g_0(\chi,\bar\chi)+iSg_1(\chi,\bar\chi) \end{equation} which satisfies the conditions \begin{equation}\plabel{g0} ({\partial\over{\partial\bar\chi(a)}} -\chi(a))g_0=0\quad \Rightarrow g_0(\chi,\bar\chi)=e^{-\bar\chi(a)\chi(a)} \Phi(\l_A, \bar{\l}_{\dot{A}}; \chi(a)), \end{equation} \begin{equation}\label{g1} 2\sqrt{1-a^2}g_1(\chi,\bar\chi)= -i({\partial\over{\partial{\chi(a)}}}-\bar\chi(a))g_0. \end{equation} We see that when $a\not =1$ the component $g_1$ of the superfunction \p{gs} is expressed in terms of $g_0$ which is specified by the condition \p{g0} in terms of a single independent superfield $\Phi(\l_A, \bar{\l}_{\dot{A}}; \chi(a))$. Hence, the independent wave function \p{wf4D} which describes the general solution of the equations \p{tPhi1q}--\p{tbDA=0q} is \begin{equation} \plabel{wf4Ds3} \Psi = e^{ i \l_A \bar{\lambda}_{\dot{A}} x^{A\dot{A}} + ia \l_A\l_By^{AB} + ia \bar{\lambda}_{\dot{A}}\bar{\lambda}_{\dot{B}} \bar{y}^{\dot{A}\dot{B}} - \bar\chi(a)\chi(a)} ~~~ \Phi( \tilde{\l}^A, \tilde{\bar{\l}}^{\dot{A}};\chi(a)). \end{equation} At the critical value $a=1$ the result is the same though the proof is slightly changed since in such a case (as in Subsection 3.1.4) we should deal with a single conversion Clifford variable $f$ instead of $f_1$ and $f_2$ (and/or $S$ and $\bar S$). More precisely, in eqs. \p{tDA=0q2} and \p{tbDA=0q2} one should put a=1, $\bar\chi=\chi$ and $f_1=S=0$, and then follow the quantization prescription described in the Appendix. One should notice that the wave function $\Phi(\tilde{\l}^A, \tilde{\bar{\l}}^{\dot{A}};\chi(a))$ in \p{wf4Ds3} has exactly the same structure as in the supertwistor case \p{twvfa<1}, but where now $\chi(a)$ are the composite Grassmann coordinates defined by eqs. \p{chi}, \p{chia}. We therefore conclude that the direct supertwistor quantization and the quantization with the use of conversion of the superparticle model based on the generic action \p{ac(a)} result in the same supersymmetric spectrum of the quantum states. The supersymmetry transformations of the components of $\Phi( \tilde{\l}^A, \tilde{\bar{\l}}^{\dot{A}};\chi(a))$ are easily derived from \p{wf4Ds3} using the supersymmetric variations \p{susy} of the coordinates. The higher dimensional generalization of the $a=1$ model and its quantization will be the subject of the next section. \section{The $a=1$ model in higher dimensions and internal degrees of freedom} \setcounter{equation}{0} A generalization of the $D=4$, $a=1$ superparticle model has been proposed in \cite{BL}. In higher space--time dimensions $D$ we consider an extension of an $N=1$ supersymmetry algebra by tensorial central charges \begin{equation}\plabel{n1cc} \{Q_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon,Q_\b\}=P_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}, \quad [P_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b},Q_\g]=0, \end{equation} where, depending on space--time dimension $D$, the supercharges $Q_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon$ $(\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon = 1,\ldots 2^k)$ are real Majorana, or Majorana--Weyl spinors\footnote{\label{lufoot} In the general case one can also consider the cases of pseudo--Majorana, simplectic--Majorana and Dirac (complex) supercharges \cite{west}. Technical details of the extension of the results of Section 5 to arbitrary type of supercharges will be considered in another publication.} and $P_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}$ is a symmetric generalized `momentum' generator conjugate to $2^k (2^k +1)$ symmetric spin-tensor coordinates $X^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}$, which can be split into the usual space--time coordinates and tensorial central charge coordinates, as we shall demonstrate below. We assume that $P_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}$ is defined by the Cartan--Penrose relation \begin{equation}\plabel{CPrepD} P_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} = \l_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon} \l_{\b}, \end{equation} where the real bosonic spinor $\l_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}$ has the same spinor properties as the supercharge $Q_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon$. The expression \p{CPrepD} implies the BPS condition $det P_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}=0$ and can be obtained as a primary constraint from the action functional \cite{BL} \begin{equation}\plabel{actionD} S = \int_{{\cal M}^1} \l_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon} \l_{\b} \Pi^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} \end{equation} $$ \Pi^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} = dX^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} - i d\Theta^{(\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}\Theta^{\b )} =d\tau \Pi^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}_\tau , \qquad $$ $$ \alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon =1,...,2^k. $$ \bigskip For any value of $k$ the model possesses $2^k$ global target space supersymmetries generated by \p{n1cc} $$ \d_{susy} X^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} = i \Theta^{(\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}\e^{\b )}, \qquad \d_{susy} \Theta^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}=\e^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon }, \qquad \d_{susy} \l_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon} = 0, \qquad $$ as well as $2^k-1$ $\kappa$--symmetries. To show the presence of the $2^k-1$ $\kappa$--symmetries let us write the variation of the action \p{actionD} \begin{equation}\plabel{actionDv} \d S = \int_{{\cal M}^1} \left(2 \d \l_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon} ~\l_{\b} \Pi^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} + d (\l_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon} \l_{\b}) i_\d \Pi^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} - 2i d\Theta^\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon \l_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon} \l_{\b}\d\Theta^\b \right) + \left(\l_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon} \l_{\b} i_\d \Pi^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} \right)\vert_{\tau_i}^{\tau_f}, \end{equation} where $$ i_\d \Pi^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} = \d X^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} - i \d \Theta^{(\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}\Theta^{\b )}. \qquad $$ One can see that only one linear combination $\l_{\b}\d\Theta^\b$ of $2^k$ independent variations of Grassmann coordinates $\d\Theta^\b$ is effectively involved into the variation of the action. Hence, other $2^k-1$ Grassmann coordinate variations (which do not appear in \p{actionDv}) can be identified with the parameters of local fermionic $\kappa$--symmetry. They can be written in the form $$\kappa^I= u^I_{\b}\d\Theta^\b, $$ where $u^I_{\b}$ ($I=1,\ldots, 2^k-1$) and $\l_\b$ form a set of $2^k$ linearly independent bosonic spinors. \bigskip Identifying the $\kappa$ symmetry with the part of target space supersymmetry which is preserved by the particle or brane configuration, we claim that the model \p{actionD} describes the dynamics of BPS states preserving all but one target space supersymmetries in space--time of a dimension $D$. \bigskip {\bf Examples.} \bigskip In $D=3$ (where $k=1$) the action \p{actionD} describes the standard massless superparticle $$ k=1~\leftrightarrow ~D=3: \quad X^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} = X^m \g_m^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}, \quad P^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} = P^m \g_m^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}. $$ On the other hand the case $k=1$ can be regarded as a model in a `minimal' $D=2+2$ superspace with self-dual tensorial central charge coordinates $X^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}= y^{mn}\s_{mn}^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}$, $y^{mn} = {1\over 2} \e^{mnkl} y_{kl}$. \bigskip The case of $k=2$ corresponds to the $D=4$, $a=1$ model considered in Sections 2--4 but written in the Majorana representation. \bigskip The construction also holds in $D=6$ where $k=3$, but here we should use the ($SU(2)$--Majorana--Weyl) `reality' conditions. In addition to the 4--dimensional spinor index $\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon$ the complex 8--component spinors $Q_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon^i$ and $\l_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon^i$ carry the $SU(2)$ index $i=1,2$ and they are the $SU(2)$ Majorana--Weyl spinors (see for details \cite{west}). The number of tensorial central charges in this model is $30$. The case $k=4$ can be regarded as describing a $D=10$ massless superparticle with $126$ composite (self--dual) tensorial central charges $Z_{m_1 \ldots m_5}$ \cite{BL} (cf. with \cite{HP,ES}). The real supercharges $Q_{\alpha}$ satisfy the Majorana--Weyl reality condition. \bigskip The action \p{actionD} with $k=5$ corresponds to a $0$--superbrane model in $D=11$ superspace with $517$ tensorial central charges composed from $32$ components of one real bosonic Majorana spinor. In contrast to the cases of $D=3,4,6$ and 10, in such a model the superparticle is not massless, the mass of the 0-brane being generated dynamically in a way similar to the mechanism generating the tension of superstrings and superbranes \cite{generation} (see \cite{BL} for some details). \bigskip On the other hand it is possible to use the twelve dimensional $D=2+10$ $32\times 32$ gamma matrices to treat the $k=5$ model from the point of view of two--time physics \cite{Bars}. The bosonic coordinates $X^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}$ are decomposed into two--index and self--dual 6--index central charge coordinates $y^{mn}, y^{m_1\ldots m_6}= 1/6! \epsilon^{m_1\ldots m_6 n_1\ldots n_6} y_{n_1\ldots n_6}$. \bigskip \subsection{$OSp(1|2^k)$ supertwistor representation of the D-dimensional model} Performing the integration by parts we can rewrite the action \p{actionD} in the $OSp(1|2^k)$ invariant form (i.e. $OSp(1|16)$ for $D=10$ and $OSp(1|32)$ for $D=11$) in terms of a supertwistor $Y^{{\cal A}}= (\mu^\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon, \zeta )$ \begin{equation}\plabel{actiontwD} S= - \int (\mu^\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon d\l_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon + i d\zeta ~\zeta ), \qquad \alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon = 1,\ldots , 2^k. \end{equation} The generalized Penrose--Ferber correspondence between the supertwistors and the generalized superspace looks as follows \begin{equation}\plabel{muXZD} P_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} = \l_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon} \l_{\b},\quad \mu^\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon = X^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} \l_\b - i \Theta^\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon (\Theta^\b \l_\b), \quad \zeta = \Theta^\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon \l_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon . \end{equation} and does not imply other constraints. \subsection{Quantization of the higher dimensional model with the use of conversion} The quantization of the supertwistor formulation \p{actionD} is straightforward and is completely analogous to the quantization of the $D=4$ model considered in Subsection {\bf 3.1.4} and Appendix. The spectrum of quantum states is described by the superfield \begin{equation}\plabel{wftwD} \Phi = \Phi ( \l_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon, \zeta ) = \phi ( \l_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon) + i \zeta \psi ( \l_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon) \end{equation} depending on the bosonic spinor $\l_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon$ and one Grassmann (or, equivalently, Clifford) variable $\zeta $. \bigskip For completeness we briefly describe the quantization of the higher dimensional model \p{actionD} with the use of conversion. The primary constraints of the model \p{actionD} are \begin{equation}\plabel{Phi} \Phi_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} \equiv P_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} - \l_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\l_\b = 0 \end{equation} \begin{equation}\plabel{D} D_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon} \equiv \pi_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon} + i \Theta}\def\th{\theta}\def\om{\omega}\def\Om{\Omega}\def\G{\Gamma^\b P_{\b\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon} = 0 \end{equation} \begin{equation}\plabel{Pl} P^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}_{(\l )} = 0 \end{equation} where the momenta are defined in such a way that the nonvanishing Poisson brackets have the following form \begin{equation}\plabel{Pbr} [P_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}, X^{\g\d}]_P= 2\d_{(\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}^{(\g}\d_{\b )}^{\d) } , \qquad \{ \pi_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}, \Theta}\def\th{\theta}\def\om{\omega}\def\Om{\Omega}\def\G{\Gamma^{\b} \}_P = \d_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}^{\b} , \qquad [P^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}_{(\l )}, \l_\b ]_P = \d^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}_{\b} . \qquad \end{equation} This set of $2^k (2^k + 1) / 2$ bosonic and $2^k$ fermionic constraints obeys the algebra \begin{equation}\plabel{algD} [\Phi_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}, P^{\g}_{(\l )}]_P= 2\l_{(\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon} \d_{\b )}^{\d} , \qquad \{ D_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}, D_{\b} \}_P = 2i P_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} \equiv 2i (\Phi_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} + \l_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\l_\b), \end{equation} $$ all~other~brackets~= 0, $$ and thus contains $2^k$ bosonic and $1$ fermionic second class constraints. Therefore, our system with $2^k (2^k + 1) / 2$ bosonic and $2^k$ fermionic configuration space variables contain $2^k$ bosonic and $1$ fermionic physical degrees of freedom which can be identified with the components of $OSp(1|2^k)$ supertwistor. \bigskip Exactly as in the $D=4$ case, to perform the conversion (see \cite{9}--\cite{13}) of the second class constraints into the first class ones we introduce additional 'conversion' degrees of freedom (two for each pair of the bosonic second class constraints and one self--conjugate fermionic variable for each fermionic second class constraint) \begin{equation}\plabel{Pbrmu} [P^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}_{(\rho )}, \rho_\b ]_P = \d^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}_{~\b}, \qquad \{ \xi , \xi \} = - {i \over 2}, \qquad \end{equation} and transform the second--class constraints into first--class ones extending the former with the new coordinates and momenta \begin{equation}\plabel{tPhi} \tilde{\Phi}_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} \equiv P_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} - \tilde{\l}_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon \tilde{\l}_\b= 0 , \end{equation} \begin{equation}\plabel{tD} \tilde{D}_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon} \equiv D_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon + 2 \tilde{\l}_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon \xi \equiv \pi_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon} + i \Theta}\def\th{\theta}\def\om{\omega}\def\Om{\Omega}\def\G{\Gamma^\b P_{\b\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon} + 2 \tilde{\l}_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon \xi = 0, \end{equation} \begin{equation}\plabel{tPl} P^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}_{\tilde{\rho} } = 0. \end{equation} where \begin{equation}\plabel{tildel} \tilde{\l}_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}=\l_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon + \rho_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}, \qquad \tilde{\rho}_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}=\l_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon - \rho_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}. \end{equation} Following the Appendix we obtain the superwave function describing the first--quantized states of the model determined by the single superfield \p{wftwD} depending on $\tilde{\l}_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon$ and one Grassmann variable $\chi= (\Theta}\def\th{\theta}\def\om{\omega}\def\Om{\Omega}\def\G{\Gamma\l)$. We have \begin{equation}\plabel{Psisol} \Psi(\tilde{\l}_a, (\Theta}\def\th{\theta}\def\om{\omega}\def\Om{\Omega}\def\G{\Gamma\tilde\l))= e^{{i \over 2} \tilde{\l}_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\tilde{\l}_\b X^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon \b}} \left[\phi(\tilde{\l}_a) + i(\Theta}\def\th{\theta}\def\om{\omega}\def\Om{\Omega}\def\G{\Gamma\tilde\l )~ \psi (\tilde{\l}_a)\right]. \end{equation} In the sector with even $\lambda$--parity of the wave function ($\Psi=\Psi_{+}$) the spectrum of the quantum states of the model \p{actionD} is described by one bosonic $\phi_+(\tilde{\l}_a)$ and one fermionic $ \psi_-(\tilde{\l}_a)$ function, while in the $\l$--odd sector ($\Psi=\Psi_{-}$) we have the fermionic field $\phi_-(\tilde{\l}_a)$ and the bosonic field $ \psi_+(\tilde{\l}_a)$. This is in complete correspondence with the result of the quantization of the free supertwistor model \p{actiontwD}. \subsection{Properties of the wave function with arbitrary helicity spectrum} To clarify the meaning of the wave function \p{wftwD} (or \p{Psisol}), let us consider its bosonic limit at $\Theta^\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon=0$ \begin{equation}\plabel{Psisol0} \Psi= e^{{i \over 2} \tilde{\l}_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\tilde{\l}_\b X^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon \b}} \phi(\tilde{\l}_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon), \qquad \alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon=1,\ldots , 2^k, \end{equation} and use the decomposition of the product of the spinor representations in the basis of $D$-dimensional gamma-matrices. \bigskip For the simplest case $k=1$ ($\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon =1,2$), where our model coincides with a $D=3$ counterpart of the usual (Ferber-Shirafuji) model \cite{F78,S83}, the Fierz identity reads \footnote{ We use the matrices $\gamma^{a}_{\alpha \beta}$ which are symmetric and obtained from standard Dirac matrices $(\gamma^{a})_{\alpha}^{~\beta}$ by lowering one of the indices with the charge conjugation matrix $C=\gamma_{0}= i \tau_2$.} $$ D=3: \qquad \l_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon \l_\b = {1 \over 2} \gamma^a_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} (\l \g_a\l)= {1 \over 2} \gamma^a_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} P_a, $$ and we can identify the matrix coordinates $X^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}$ and their momenta $P_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}$ with the usual vector coordinates and momenta $$ D=3: \quad X^a = {1 \over 2} \gamma^a_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}X^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}, \quad X^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}= X^a \gamma_{a\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}; \quad P^a = {1 \over 2} \gamma^a_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}P^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}, \quad P_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}= P^a \gamma_{a\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}. \qquad $$ Thus, in $D=3$ eq. \p{Psisol0} describes a plane wave solution\footnote{More precisely, in $D=3$ $\phi( \l_a ) = \psi (p_m, sign(\l ))$, where $sign(\l )$ denotes the sign factor $(\pm 1)$ of the bosonic spinor. This is a `parameter' of the residual {\bf Z}$_2$ symmetry, whose action on $\l$ does not change $p_m$.} \begin{equation}\plabel{Psisol3D0} D=3: \qquad \Psi= e^{i p_m X^{m}}\phi (p_m). \end{equation} \bigskip The case $k=2$, $D=4$ has been analized in detail in Sections 2--4. To transform the wave function \p{Psisol0} to the wave function \p{wf4Ds3} (at $a=1$, and $\Theta=0$) one should perform the similarity transformation from the real Majorana to the complex Weyl representation of the $D=4$ gamma--matrices and replace the Majorana spinor by the pair of complex conjugate Weyl spinors \begin{equation}\plabel{sdec4D} \tilde\l_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon ~~ \leftrightarrow \pmatrix{ \tilde{\l}_A \cr {\bar{\tilde{\lambda}}}_{\dot{A} }\cr }. \end{equation} In the momentum representation the wave function $\phi (\l_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon )$ differs from the usual one given by $\phi_0 (p_m)$ by the presence of additional dependence on the angle variable $\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon$ which describes the common phase factor of the Weyl spinor $\lambda_{A}$ $(\lambda_{1} = e^{i(\alpha +\beta)} |\lambda_{1}|$, $\lambda_{2}=e^{i(\alpha - \beta)} |\lambda_{2}|$) and parametrizes the 1-dimensional sphere $S^1$ $$ D=4: \qquad \phi (\l_a)=\phi (\l_A, \bar{\l}_{\dot{B}}) = \phi (p_m, \alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon), \qquad \alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon \in [0,2\pi) $$ where $ p_m = {1 \over 4} \l \g_m \l = {1 \over 2} \l^A {\s}_{A\dot{A}} \bar{\l}^{\dot{A}}$ (see also $^{10}$). The additional internal momentum variable $\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon$ is the only independent degree of freedom contained in the $D=4$ tensorial central charges composed of the bosonic spinor $$ \alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon \in [0,2\pi) \qquad \Leftrightarrow \qquad Z_{mn} = {1 \over 4} \l\g_{mn} \l \qquad $$ It describes the $D=4$ helicity spectrum of the quantum states. \bigskip In the general case of $k > 2$ with $2^k$ equal to the dimension of an irreducible spinor representation of $SO(1,D-1)$ in $D= 3,4,6, ~10 ~(mod~8)$ (i.e. $k= 3,4,\ldots$) the discussion is similar. For example, in the case $k=4$, $D=10$ we can use the basis of symmetric $\sigma$ matrices $\s_m, \s_{m_1 \ldots m_5}$ to make the decomposition \begin{equation}\plabel{ll10D} \l_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon} \l_{\b} \equiv P_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} = P_m \s^m_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}+ Z_{m_1...m_5} \s^{m_1...m_5}_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} , \end{equation} where \begin{equation}\plabel{P10D} P_m = {1 \over 16} \l_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon} \s_m^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}\l_{\b} \qquad \Rightarrow \qquad P_mP^m = 0 \end{equation} is an ordinary light--like momentum vector in $D=10$ and \begin{equation}\plabel{Z10D} Z_{m_1...m_5} = {1 \over 16\cdot 5! } \l_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon} \s_{m_1...m_5}^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}\l_{\b} \qquad \end{equation} is the momenta canonically conjugate to the {\bf 126} tensorial central charge coordinates $y^{m_1...m_5}$. It was demonstrated that the $D=10$ model contains the local symmetries (first class constraints) and second--class constraints which reduce the number of the classical bosonic degrees of freedom to the ones described by the $16$--component bosonic spinor $\l_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon$ and its momentum plus one Grassmann degree of freedom. In the quantum theory this is reflected in the dependence of the `momentum space representation' of the wave function on 16 bosonic spinor variables and one Grassmann variable only. Due to the identities $(\s_m)_{(\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} (\s^m)_{\g ) \d}=0$ the $D=10$ momentum \p{P10D} is light--like. Hence, the tensorial central charge momenta $Z_{m_1...m_5}$ contain $16-9= 7$ additional degrees of freedom which are not determined by the light-like momentum. \bigskip We now show that these additional internal degrees of freedom parametrize an $S^7$ sphere. For this purpose we perform a Lorentz transformation to a frame where the light--like momentum \p{P10D} acquires the form \begin{equation}\plabel{P10Dst} P_m= (p, 0,0,0,0,0,0,0,0, p). \end{equation} Then in this frame we make an $SO(8)$ invariant split of the bosonic spinor $\tilde{\l}_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon$ \begin{equation}\plabel{sdec10D} \tilde{\l}_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon =\left(\matrix{\L_q \cr \S_{\dot{q}}\cr }\right), \qquad q=1, \ldots 8, \qquad \dot{q} = 1, \ldots 8 \end{equation} and choose the $SO(8)\times SO(1,1)$ covariant representation for the D=10 $\sigma$-matrices $$ \s^{\underline{ 0}}_{\underline{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}\underline{\b}}= \hbox{ {\it diag}}(\delta _{{ qp}}, \delta_{{\dot q}{\dot p}}) = \tilde{\s }^{\underline{0}~\underline{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}\underline{\b}} , \qquad \s^{\underline{9}}_{\underline{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}\underline{\b}}= \hbox{ {\it diag}} (\delta _{qp}, -\delta _{{\dot q}{\dot p}}) = -\tilde{\s }^{\underline{9}~\underline{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}\underline{\b}} , $$ \begin{equation}\plabel{gammarep} \s^{i}_{\underline{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}\underline{\b}} = \left(\matrix{0 & \gamma ^{i}_{q\dot p}\cr \tilde{\gamma}^{i}_{{\dot q} p} & 0\cr} \right) = - \tilde{\s }^{i~\underline{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}\underline{\b}} , \end{equation} $$ \s^{{++}}_{\underline{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}\underline{\b}} \equiv (\s^{\underline{ 0}}+ \s^{\underline{ 9}})_{\underline{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}\underline{\b}}= \hbox{ {\it diag}}(~2\delta _{qp},~ 0) = -(\tilde{\s }^{\underline{ 0}}- \tilde{\s }^{\underline{ 9}})^{\underline{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}\underline{\b}} = \tilde{\s }^{{--}~\underline{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}\underline{\b}} , \qquad $$ $$ \s ^{{--}}_{\underline{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}\underline{\b}}\equiv (\s ^{\underline{ 0}}-\G ^{\underline{ 9}} )_{\underline{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}\underline{\b}}= \hbox{ {\it diag}}(~0, ~2\delta _{{\dot q}{\dot p}}) = (\tilde{\s }^{\underline{ 0}}+ \tilde{\s }^{\underline{ 9}})^{\underline{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}\underline{\b}} = \tilde{\s }^{{++}~\underline{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}\underline{\b}}. $$ In the frame \p{P10Dst} the Cartan-Penrose representation \p{P10D} looks as follows \begin{equation}\plabel{CP10Dst} \L_q~\L_q = p, \qquad 2 \S_{\dot q} \S_{\dot q} =0, \qquad \L_{q} \g^i_{q\dot{p}} \S_{\dot{p}}=0 \end{equation} The general solution of eqs. \p{CP10Dst} is \begin{equation}\plabel{sdec10Dsol} \S_{\dot q} =0, \qquad \Rightarrow \qquad \tilde{\l}_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon =\left(\matrix{\L_q \cr 0\cr }\right), \qquad \end{equation} and the only nonvanishing component of the momentum \p{P10Dst} is given by the norm of the $SO(8)$ spinor $\L_q$ \begin{equation}\plabel{CP10Dst0} p=\L_q~\L_q . \qquad \end{equation} The expression \p{CP10Dst0} is invariant under the $SO(8)$ rotations $$ \L_q ~\rightarrow~\L_p S_{pq}, \qquad S~S^T=I. $$ But not all $SO(8)$ transformations act on $\L_q$ effectively. Indeed, if one fixes the $SO(8)$ gauge \begin{equation}\plabel{sga10D} \L_q = \left(\matrix{ \pm \sqrt{p} \cr 0 \cr \ldots \cr \ldots \cr \ldots \cr 0 \cr }\right) \end{equation} one finds that i) this gauge is invariant under the $SO(7)$ transformations, ii) any form of the spinor $\L_q$ can be obtained from \p{sga10D} by a transformation from the coset space $SO(8)/SO(7)$ isomorphic to the sphere $S^7$. \bigskip Thus, the 16 components of the bosonic spinor $\l_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon$ in $D=10$ can be split into (double covering ($\lambda \simeq - \lambda$)) \\ i) degrees of freedom which characterize the light-like momentum $P_m$, \\ ii) 7 coordinates of the sphere $S^7$. The variables parametrizing the sphere $S^7$ correspond to `helicity' degrees of freedom of the quantum states of the massless $D=10$ superparticle. \bigskip It is worth mentioning that the appearance of extra compact dimensions in the momentum spaces of the superparticle models considered above is related to the well known fact that in $D=3,4,6$ and 10 the commuting spinors (twistors) with $n=2(D-2)=2,4,8$ and 16 components parametrize, modulo scale transformations, $S^1$, $S^3$, $S^7$ and $S^{15}$ spheres, respectively. These spheres are Hopf fibrations (fiber bundles) which are associated with the division algebras {\bf R,~C,~H} and {\bf O}. Their bases are the spheres $S^1$, $S^2$, $S^4$ and $S^8$, and the fibers are $Z^2$, $S^1=U(1)$, $S^3 =SU(2)$ and $S^7$, respectively. The base spheres are parametrized (up to a scaling factor which, due to the Cartan-Penrose representation, is identified with the square of spinor components) by the light--like vectors (massless particle momenta) in $D=3,4,6$ and $10$, respectively. We see that the fibers are extra ``momentum dimensions" which we have in our models with central charges (at $a\not = 0$). This is the geometrical ground for the appearance of $S^{1}$ in $D=4$ and $S^7$ in $D=10$. \section{Conclusion and discussion} We have performed the detailed analysis and quantization of the massless superparticle model with tensorial central charges associated with twistor--like commuting spinors in space--times of dimension $D=3,4,6$ and 10. The physical phase space degrees of freedom of this model have a natural description in terms of supertwistors which form a fundamental representation of a corresponding maximal supergroup of conformal type underlying the dynamics of the superparticle. A peculiarity of the $a=1$ model is that it possesses $n=2^{[{D\over 2}]}-1$ $\kappa$--symmetries, while the standard massless superparticles have $n=2^{[{D\over 2}]-1}$ $\kappa$--symmetries. The presence of such a large number of $\kappa$--symmetries in the $a=1$ models means that the superparticle breaks only one of the $2^{[{D\over 2}]}$ supersymmetries of the target space vacuum. This results in very short two--component supermultiplets describing the quantum states of the $a=1$ superparticle, since the corresponding target space superfields depend only on one Grassmann coordinate. The existence of these short Lorentz--covariant superfields is made possible because the target superspace has been enlarged by commuting spinor coordinates, whose role is in singling out a `small' covariant subsuperspace in the extended target superspace. Let us compare this situation with well known cases. In the case of the ordinary massless superparticle in N=1, D=4 superspace the quantum states of the superparticle are described by a chiral scalar superfield \cite{n1}. The chirality constraint is a consequence of first--class fermionic constraints generating $\kappa$--symmetries. Consequently the chiral superfield effectively depends on only two Grassmann coordinates, which reflects the fact that the ordinary superparticle preserves half (i.e. two out of four) supersymmetries of $N=1$, $D=4$ superspace. In the case of an $N=2$, $D=4$ superparticle in harmonic superspace \cite{n2} which also breaks half of the target space supersymmetries, $SU(2)$--harmonic variables allow one to pick a harmonic analytic subsuperspace out of the general $N=2$, $D=4$ superspace \cite{harm}, and quantum states of the superparticle are described by analytic superfields which depend on four Grassmann coordinates singled out from the original eight Grassmann coordinates by the use of the harmonic variables. In the analogous way, in the case of the generalized superparticle model \p{ac(a)}, \p{actionD} at $a=1$, when only one target space supersymmetry is broken, one finds a Lorentz covariant subsuperspace of the target superspace, which has only one Grassmann direction parametrized by the Lorentz scalar $\theta^\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\lambda_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon$. The ``short'' superfields \p{wf4Ds3}, \p{Psisol}, which exist only due to the presence of the auxiliary spinor variable, describe the quantum states of the generalized superparticle. We have shown that in contrast to standard superparticles the considered model possesses additional compact phase--space variables which describe helicity degrees of freedom of the superparticle and which upon quantization parametrize infinite tower of free states with arbitrary (half)integer helicities. Due to this property it would be interesting to consider the possibility of treating our generalized superparticle model as a classical mechanics counterpart of the theory of higher--spin fields developed by M. Vasiliev \cite{Vasiliev}. Since the nontrivially interacting higher spin fields should live in a space--time of (anti)-de--Sitter geometry a natural generalization of the results of this paper would be to consider a superparticle model on supergroup manifolds describing isometries of corresponding AdS superspaces. For $D=4$ the supergroup $OSp(1|4)$ is the isometry of a $D=4$ AdS superspace ${OSp(1|4)}\over{SO(1,3)}$ which in addition to 4 bosonic directions has 4 Grassmann fermionic directions. Six bosonic coordinates corresponding to the group $SO(1,3)$ (which extends the coset superspace ${OSp(1|4)}\over{SO(1,3)}$ to the supergroup manifold $OSp(1|4)$) are a non--Abelian generalization of the central charge coordinates of the $D=4$ model considered above. It appears that our model with central charges can be regarded as an appropriate truncation of the $OSp(1|4)$ model. Work in this direction is now in progress. \bigskip We should also remark that tensorial central charges are usually associated with brane charges, which are topological and take discrete (quantum) values. In contrast, in the superparticle models considered in this paper the central charges take continuous values and parametrize compact manifolds, while their Fourier conjugate coordinates are quantized. \bigskip {\bf Acknowledgements}. The authors would like to thank J. de Azcarraga, K. Landsteiner, C. Preitschopf and M. Tonin for interest to this work and valuable discussions. Work of I.B. and D.S. was partially supported by the INTAS Grants N96--308 and N 93--493--ext. \newpage \noindent \section*{APPENDIX} \setcounter{equation}{0} \subsection*{A1. Quantization of one fermionic degree of freedom by a `half-conversion' prescription} Here we shall present a method of quantizing a single fermionic variable alternative to that used in Subsection 3.1.4, but which leads to the same spectrum of quantum states. Let us convert the second--class constraint \p{S1} into the first--class constraint by introducing one more Clifford-like variable $\xi^\prime$ \begin{equation}\plabel{Pbrxi'} \{ \xi^\prime , \xi^\prime \}_P = - {i \over 2}. \qquad \end{equation} Using $\xi^\prime$ we replace \p{S1} by the first class constraint \begin{equation}\plabel{Azeta} {\cal A} \equiv \xi -i \pi_{(\xi)} + 2 \xi^\prime =0, \qquad \{ {\cal A} , {\cal A} \}_P= 0, \end{equation} where instead of $\chi_1$ of Eq. \p{S1} we have introduced ${\xi =2\chi_1}$. \bigskip Let us quantize the model using the coordinate representation for the original Grassmann variable $\xi$ $$ \hat{\xi}=\xi, \qquad \hat{\pi}_\xi = i {\partial\over \partial \xi } $$ and a real $2 \times 2$ matrix representation \begin{equation}\plabel{xitau} \hat{\xi^\prime}= {1 \over 2} \tau = {1 \over 2} \left(\matrix{ * & * \cr * & * \cr}\right) \qquad \tau^2=I \end{equation} for the new Clifford algebra valued variable $\hat{\xi^\prime}$ $$ \hat{\xi^\prime}^2 = 1/4 . $$ Then the wave function is regarded as a column \begin{equation}\plabel{Psicol} \Psi_a= \left(\matrix{ \phi (\xi) \cr \psi (\xi) \cr}\right) \end{equation} and the quantum counterpart of the first class constraint \p{Azeta} \begin{equation}\plabel{hatA} \hat{{\cal A}} \equiv (\xi - {\partial \over \partial \xi }) \tau^\prime + \tau \qquad \end{equation} should be imposed on the wave function \begin{equation}\plabel{APsi=0} \hat{{\cal A}}_{ab}\Psi_b \equiv [(\xi - {\partial \over \partial \xi })~ \tau^\prime_{ab} + \tau_{ab}]\Phi_b (\xi) = 0 . \qquad \end{equation} In \p{hatA} and \p{APsi=0} the second $2 \times 2$ matrix \begin{equation}\plabel{tau'} \tau^\prime_{ab} = \left(\matrix{ * & * \cr * & * \cr}\right), \qquad (\tau^\prime)^2=I \end{equation} was introduced. It is required to ensure the anticommutativity of the Grassmann and Clifford part of the first class constraint \p{APsi=0}. Indeed, let us calculate the square of the quantum constraint \p{hatA} \begin{equation}\plabel{hatA20} \hat{{\cal A}}^2 \equiv {1 \over 2}\{ \hat{{\cal A}}, \hat{{\cal A}}\} = \tau ^2 + (\xi - {\partial \over \partial \xi })^2 (\tau^\prime)^2 + (\xi - {\partial \over \partial \xi }) \{ \tau^\prime , \tau\} . \qquad \end{equation} Since \begin{equation}\plabel{op2} (\xi - {\partial \over \partial \xi })^2= {1 \over 2} \{ (\xi - {\partial \over \partial \xi }), (\xi - {\partial \over \partial \xi }) \} = -1 \end{equation} and $$ \tau^2 = I = (\tau^\prime)^2, $$ one easily finds that the first two terms in \p{hatA20} cancel and arrives at \begin{equation}\plabel{hatA2} \hat{{\cal A}}^2 = (\xi - {\partial \over \partial \xi }) {1 \over 2} \{ \tau^\prime , \tau\} . \qquad \end{equation} The last input vanishes if and only if $\{ \tau^\prime , \tau\}=0$. This result can not be reached if one chose $ \tau^\prime$ to be the unit matrix. $ \tau$ and $ \tau^\prime$ can be chosen to be two Pauli matrices. Let us stress that the necessity to introduce the second matrix $ \tau^\prime$ is a peculiarity of the quantization of the odd number of Clifford variables. $ \tau^\prime$ can be the unit matrix in the case of even number of Clifford variables (see e.g. \cite{even} and references therein). To fix the representation for the matrices $\tau$ and $\tau^\prime$ one has to note that the conservation of the Grassmann parity in the form of the first class constraint \p{APsi=0} requires that \begin{itemize} \item The components $\phi (\xi)$ and $ \psi (\xi)$ of \p{xitau} must have {\sl different Grassmann parity}. For instance, if we choose $\phi (\xi)$ to be bosonic superfield then $\psi (\xi)$ is fermionic. \item If the diagonal representation is chosen for one of the matrices, say $\tau^\prime$, then another matrix $\tau$ is antidiagonal. \end{itemize} Taking these in mind we choose \begin{equation}\plabel{taurep} \tau^\prime = \tau_3 \equiv \left(\matrix{ 1 & 0 \cr 0 & -1 \cr}\right), \qquad \tau = \tau_1 \equiv \left(\matrix{ 0 & 1 \cr 1 & 0 \cr}\right). \qquad \end{equation} Then the quantum constraints \p{APsi=0} acquire the form \begin{equation}\plabel{hatAPsi=0} \hat{{\cal A}}_{ab}\Psi_b \equiv \left(\matrix{ (\xi - {\partial \over \partial \xi })~ & 1 \cr 1 & - (\xi - {\partial \over \partial \xi }) \cr}\right) \left(\matrix{ \phi (\xi) &\cr \psi (\xi)& \cr}\right)_b = \left(\matrix{ (\xi - {\partial \over \partial \xi }) \phi (\xi) + \psi (\xi) \cr \phi (\xi) - (\xi - {\partial \over \partial \xi }) \psi (\xi) \cr}\right) = 0 , \qquad \end{equation} which splits into two equations \begin{equation}\plabel{eqs} (\xi - {\partial \over \partial \xi }) \phi (\xi) =- \psi (\xi) , \qquad \end{equation} $$ \phi (\xi) = (\xi - {\partial \over \partial \xi }) \psi (\xi) $$ Using \p{op2}, we notice that the second equation is a consequence of the first one. The first equation \begin{equation}\plabel{eq} (\xi - {\partial \over \partial \xi }) \phi (\xi) =- \psi (\xi) \qquad \end{equation} expresses the fermionic superfield through the bosonic one. I.e. if we write $\phi (\xi)$ in components $$ \phi (\xi) = \phi_0 + i \xi \psi_1, $$ then from eq. \p{eq} it follows that $$ \psi (\xi) = i \psi_1 - \xi \phi_0. $$ Thus, we can represent the spectrum of states carrying one Clifford degree of freedom by one (either bosonic or fermionic) superfield $\phi (\xi)$ depending on the single {\sl Grassmann} variable $\xi$ $$ \xi^* = \xi, \qquad \xi^2 = 0. \qquad $$ This result is in accordance with that of Subsection 3.1.4 (see also \cite{Sorokin87}), and both methods of quantizing a single fermionic variable result in the same field content of quantum states (one boson and one fermion). \subsection*{A2. Quantization of the high-dimensional model with the use of conversion} Here we present some details of getting the wave function \p{Psisol} from the converted system of constraints \p{tPhi}, \p{tD} and \p{tPl} describing the high-dimensional generalization of the first--quantized $a=1$ model. Let us choose the (super)coordinate representation for supercoordinates and bosonic spinors \begin{equation}\plabel{hatP} \hat{P}_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} = - i {\partial \over \partial X^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}}, \qquad \hat{P}^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}_{(\tilde{\l}) }= - i {\partial \over \partial \tilde{\l}^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}}, \end{equation} \begin{equation}\plabel{hatpi} \hat{\pi}_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon} = i {\partial \over \partial \Theta^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}} \end{equation} and use the $2\times 2$ matrix representation \begin{equation}\plabel{xitauD} \hat{\xi}= {1 \over 2} \tau_2 = \left(\matrix{ 0 & 1 \cr 1 & 0 \cr}\right), \quad \{ \hat{\xi}, \hat{\xi}\} = {1 \over 2}. \end{equation} for the Clifford variable $\xi$. Then the the wave function is a column \begin{equation}\plabel{wfD} \Psi_a = \Psi_a (X^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}, \tilde{\l}_\b) = \left(\matrix{ \phi (X^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}, \tilde{\l}_\b, \tilde{\rho}_\g) \cr \psi (X^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}, \tilde{\l}_\b) \cr}\right) \end{equation} with the elements carrying opposite Grassmann parity (e.g. $\phi$ is bosonic and $\psi$ is fermionic) and the quantum first class constraints \p{tPhi}, \p{tD}, \p{tPl} should be taken in the form \begin{equation}\plabel{qtildePhi} \hat{\tilde{\Phi}}_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b} = - i (\partial_{\b\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}- i \tilde{\l}_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon \tilde{\l}_\b)~I \end{equation} \begin{equation}\plabel{qtildeD} \hat{\tilde{D}}_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon = \hat{{D}}_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon + \tilde{\l}_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon \hat{\xi} = i (\partial_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon - i \Theta^\b \partial_{\b\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon}) \tau_3 + \tilde{\l}_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon \tau_2). \end{equation} The incorporation of the $\tau_3 $ matrix is necessary to provide the properties of the first class constraints to form the closed algebra \begin{equation}\plabel{qDbr} \{ \hat{\tilde{D}}_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon, \hat{\tilde{D}}_\b \} = - 2i \hat{\tilde{\Phi}}_{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\b}. \end{equation} This is a peculiarity of the quantization of the models with odd number of phase space Grassmann variables (see section A1. of this Appendix). \bigskip The further steps of the quantization procedure exactly repeat the steps of the $D=4$ case (see Section 4). The wave function describing the spectrum of the quantum states is \begin{equation}\plabel{Psisolc} \tilde{\Psi_a}= e^{{i \over 2} \tilde{\l}_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon\tilde{\l}_\b X^{\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon \b}} \left(\matrix{ \Psi (\tilde{\l}_a, \Theta}\def\th{\theta}\def\om{\omega}\def\Om{\Omega}\def\G{\Gamma\l ) \cr -i \left(\partial_\chi + \chi\right) \Psi (\tilde{\l}_a, \Theta}\def\th{\theta}\def\om{\omega}\def\Om{\Omega}\def\G{\Gamma\l ). \cr}\right)\end{equation} As the second element in the column is expressed through the first one, we can describe the spectrum of the quantum states by the single superfield \p{Psisol} depending on bosonic $\tilde{\l}_\alpha}\def\b{\beta}\def\g{\gamma}\def\d{\delta}\def\e{\epsilon$ and fermionic $\chi= (\Theta}\def\th{\theta}\def\om{\omega}\def\Om{\Omega}\def\G{\Gamma\l)$.
\section{Introduction} The observation of ultra-high-energy cosmic rays (UHECR) with energies above $10^{11} \, \text{GeV}$\cite{Haya94,Bird94} poses a serious challenge to the particle acceleration mechanisms so far proposed. This fact has motivated the search for non-acceleration models, in which the high energy cosmic rays are produced by the decay of a very heavy particle. Topological defects are attractive candidates for this scenario. Due to their topological stability these objects can retain their energy for very long times and release quanta of their constituents, typically with GUT scale masses, which in turn decay to produce the UHECR. Various topological defect models and mechanisms have been studied by numerous authors\cite{Pijus98}. In this paper we investigate two different scenarios involving the annihilation of monopole-antimonopole pairs. We first discuss standard magnetic monopole pair annihilation \cite{Hill83,Bhatta95}, paying particular attention to the kinetics of monopolonium formation. We find that, due to the inefficiency of the pairing process, the density of monopolonium states formed is many orders of magnitude less than the value required to explain the UHECR events. We then present a different scenario in which very massive monopoles ($ m \sim 10^{14} \, \text{GeV}$) are bound by a light string formed at approximately $100\, \text{GeV}$. These monopoles do not have the usual magnetic charge, or in fact any unconfined flux. Gravitational radiation is the only significant energy-loss mechanism for the bound systems.\footnote{Such systems were studied in a different context by Martin and Vilenkin\cite{Martin:1997cp}.} Their lifetimes can then be comparable with the age of the universe, and their final annihilation will then contribute to the high energy end of the cosmic ray spectrum. \section{Required monopolonium abundance.} What density of decaying monopolonium states is required to produce the observed cosmic rays? The monopolonium will behave as a cold dark matter (CDM) component and will cluster in the galactic halo, producing a high energy spectrum of cosmic rays without the Greisen-Zatsepin-Kuzmin (GZK) cutoff\cite{Greisen,Zatsepin}. Since the observational data does not seem to show any such cutoff, this is an advantage of such topological defect models \cite{Bere97,Bere98}. For a given monopole mass, we can set the lifetime of the monopolonium at least equal to the age of the universe, and obtain the required density of monopolonium in the halo by normalizing the flux to the observed high energy spectrum\cite{Bere97}. The required number density decreases with the monopole mass, so as a lower limit we can take the required density corresponding to $m_M =10^{17}\, \text{GeV}$\cite{Bere97}, \begin{equation} N^{h} _{M \bar M} (T_0) > 6\times 10^{-27} \, \text{cm}^{-3}\,. \end{equation} Since the different components of the CDM cluster in the same way we can use this halo density to get the mean density in the universe, by computing, \begin{equation} N_{M \bar M} = {{N^h_{M \bar M} \, \Omega_{CDM} \, \rho_{cr}}\over {\rho^{h}_{CDM}}}. \end{equation} For $\Omega_{CDM} h^2 = 0.2$, $\rho^{h}_{CDM}= 0.3\, \text{GeV}$ $\text{cm}^{-3}$, and $\rho_{\text{cr}} = 10^4 h^2\,\text{eV cm}^{-3}$, we get \begin{equation} N_{M \bar M} (T_0) > 10^{-32} \, \text{cm}^{-3}\,. \end{equation} \label{sec:needed} We will work with a comoving monopolonium density $\Gamma = N_{M\bar M}/s$ where $s$ is the entropy density, currently $s \approx 3 \times 10^{3} \, \text{cm}^{-3}$, so that we require \begin{equation}\label{eqn:Gammaneeded} \Gamma > 10^{-35} \end{equation} to explain the observed UHECR. \section{Magnetic monopole states} \subsection{Introduction} Monopolonium states are expected to have been formed by radiative capture if there was a non-zero density of free monopoles in the early universe. They will typically be bound in an orbit with a large quantum number, so we can treat them as classical objects emitting electromagnetic radiation as they spiral down to deeper and deeper orbits, until they annihilate in a final burst of very high energy particles. The electromagnetic decay of monopolonium was analyzed by Hill\cite{Hill83} using the dipole radiation formula. The rate of energy loss is\footnote{Here and throughout we use units where $\hbar = c = k_B = 1$.} \begin{equation} {{dE}\over{dt}} = {{64 \, E^4}\over{3 \, g_M^2 m_M^2}}\,, \end{equation} where $g_M$ is the magnetic charge. From this expression, the lifetime of monopolonium with radius $r$ and binding energy $E = g_M^2/2r$ is\cite{Hill83} \begin{equation} \tau_E \sim {m_M^2 \, r^3\over {8 g_M^4}}\,. \end{equation} For $m_M = 10^{16}\, \text{GeV}$, $g_M = 1/(2e) \approx\sqrt {34}$, and an initial radius of $r= 10^{-9} \, \text{cm}$, this gives $\tau_E\sim 10^{18} \, \text{sec}$, comparable to the age of the universe. Bhattacharjee and Sigl\cite{Bhatta95} used a thermodynamic equilibrium approximation to estimate the monopolonium density and argued that the late annihilation of very massive magnetic monopoles could explain the UHECR events observed. Here we recalculate the density of monopolonium states, taking into account the kinematics of formation and the frictional energy loss of monopolonium formed at early times. \subsection{Friction} Before electron-positron annihilation, monopoles interact with a background of relativistic charged particles. These interactions produce a force which, for a non-relativistic monopole is given by \cite{Kolb81} \begin{equation}\label{eqn:F} F = {\pi\over 18}N_c T^2 v \int^{b_{\text{max}}}_{b_{\text{min}}} {db\over b} \end{equation} where $N_c$ is the number of species of charged particles, $v$ the velocity of the monopole with respect to the background gas of charged particles and $b$ the impact parameter of the incident particles. Since we are interested in the friction that a monopole feels in a bound state orbit of monopolonium, we will not consider the interaction of charged particles with impact parameter greater than the radius of the monopolonium, so $b_{\text{max}} \approx g_M^2 E^{-1}$. Initially, the monopoles are bound with energy $E\sim T$, so $b_{\text{max}}\approx g_M^2T^{-1}$. Equation (\ref{eqn:F}) is derived using the approximation that each charged particle is only slightly deflected. This approximation breaks down for impact parameters that are too small, so we should cut off our integration at \cite{Kolb81} $b_{min} \approx T^{-1}$. Using $N_c = 2$ and $g_M^2 \approx 34$, we get \begin{equation} F \approx 1.22 \, T^2 v \end{equation} so the energy loss rate due to interactions with charged particles in the background is \begin{equation} {{dE}\over{dt}} \approx 1.22 \, T^2 v^2. \end{equation} Taking the system to be bound in a circular orbit, we have \begin{equation} m_M v^2 \sim E \label{eqn:e} \end{equation} so we can write \begin{equation} {{dE}\over{dt}} \approx 1.22 \, T^2 \, {E\over {m_M}} \label{eqn:dedtf1} \end{equation} The time scale for this process is \begin{equation} \tau_F={E\over dE/dt} \approx {{m_M}\over {1.22 \, T^2}} \end{equation} If we compare it with the Hubble time, \begin{equation} \tau_H =\sqrt {90\over 8\pi^3 g_*} m_{pl} T^{-2}\approx 0.184 m_{pl} T^{-2}\,, \label{eqn:time} \end{equation} where $m_{pl}$ is the Planck mass, and $g_*$ is the number of effectively massless degrees of freedom, $g_* = 10.75$, we get \begin{equation} {{\tau_F}\over {\tau_H}} \approx 0.15 \, {{m_M}\over {m_{pl}}} \ll 1. \end{equation} Thus, we see that the damping of the monopolonium energy due to friction is very effective in this regime, and the monopoles spiral down very quickly. When the distance between monopoles becomes small as compared to $T^{-1}$, the effect of friction is reduced and Eq.\ (\ref{eqn:F}) is no longer accurate. However, even for $T = 1 \, \text{MeV}$, the radius has been reduced about two orders of magnitude to $r\sim 2\times 10^{-11}\, \text{cm}$, and the electromagnetic lifetime has been reduced by about six orders of magnitude. Thus only monopolonium states formed after electron-positive annihilation can live to decay in the present era. After electron-positron annihilation the number of charged particles in the thermal background has decreased by a factor $\sim 10^{-9}$ so $\tau_F/\tau_H\gg 1$ and the monopolonium is little affected by friction. \subsection{Formation rate} We can obtain an upper limit for the monopolonium density by solving the Boltzmann equation, \begin{equation} {dN_{M \bar M}\over{dt}} = \langle \sigma_b v \rangle n_{M}^2 - 3 \, H\, N_{M \bar M}, \end{equation} where $n_M$ denotes the free monopole density, $N_{M \bar M}$ the monopolonium density, $H$ the Hubble constant, and $\langle \sigma_b v \rangle$ the average product of the binding cross section times the thermal velocity of the monopoles. With the comoving monopole density $\gamma = n_{M}/s$, we can rewrite the equation above as \begin{equation} {{d \Gamma}\over{dt}} = \langle \sigma_b v \rangle {\gamma} \, n_{M} = \langle \sigma_b v \rangle \gamma^2 s \,. \end{equation} Using the approximation for the classical radiative capture cross section of monopoles with thermal velocities given by \cite{Dicus82}, \begin{equation} \langle \sigma_b v \rangle \approx {{\pi^{7/5}}\over 2} \, {{{g_{M}}^4}\over {{m_M}^2}} {\left({{m_M}\over T}\right)}^{9/10}\,, \label{eqn:radiative} \end{equation} and with \begin{equation} s ={2\pi^2\over 45} g_{*S} T^3\,, \end{equation} where $g_{*S}$ is the number of degrees of freedom contributing to the entropy, we get \begin{equation} {{d \Gamma}\over{dt}} = {{\pi^{17/5}}\over 45} {g_{M}^4 \gamma^2 \over m_M^2} {\left({{m_M}\over T}\right)}^{9/10} g_{*S} T^3\,. \end{equation} Since we are interested in the evolution of the monopolonium density after electron-positron annihilation, we will take a constant value $g_{*S}\approx 3.91$ to get \begin{equation} {{d \Gamma}\over{dt}} \approx 4.25 {g_{M}^4 \gamma^2 \over m_M^2} {\left({{m_M}\over T}\right)}^{9/10} T^3\,. \label{eqn:dgdt} \end{equation} As we will see, only a tiny fraction of the monopoles will ever be bound, so we can consider the comoving number of monopoles $\gamma$ to be constant. To integrate Eq (\ref{eqn:dgdt}), we will make a change of variable \begin{equation} t =\sqrt {90\over 32\pi^3 g_*} m_{pl} T^{-2}\approx 0.164m_{pl} T^{-2}\,, \end{equation} appropriate to times after electron-positron annihilation, to get \begin{equation} {{d \Gamma}\over{dT}} \approx -1.34 {g_{M}^4 m_{pl} \gamma^2 \over m_M^2} {\left({{m_M}\over T}\right)}^{9/10}\,. \end{equation} and thus \begin{equation} \Gamma_f \approx 13.4 g_M^4 \left({{m_{pl}}\over {m_M}}\right) {\left({{T_i}\over {m_M}}\right)}^{1/10} {\gamma}^2\,. \end{equation} We now take $T_i \sim 1 \, \text{MeV}$ and $g_M^2=34$, and note that to produce the observed UHECR, we must have $m_M > 10^{11}\, \text{GeV}$, so that for a fixed monopole comoving density $\gamma$, we have the bound, \begin{equation} \Gamma_f < 4\times 10^6 {\gamma}^2\,. \label{eqn:Gammaf} \end{equation} \subsection{Monopole density bound} The formation of magnetic monopoles via the Kibble mechanism\cite{Kibble76} is inevitable in all GUT models of the early universe, and annihilation mechanisms are not efficient in a rapidly expanding background\cite{Zeldovich78,Preskill79}, so that the typical initial density of monopoles produced at a GUT phase transition will very soon dominate the energy density of the universe. The most attractive solution for this problem is the inflationary scenario\cite{Guth81}. In standard inflation, the exponential expansion of the universe reduces the monopole density to a completely negligible value. However, it is possible for new monopoles to be formed at the end of inflation\cite{Turner82,Shafi:1984tt,Shafi:1984bd,Kuzmin99,Tkachev:1998dc}. The exact relic abundance of monopoles created in this period is very model dependent, but its value is constrained by the Parker limit \cite{Parker70}: To prevent the acceleration of monopoles from eliminating the galactic magnetic field, the monopole flux into the galaxy must be limited by \begin{equation} F < 10^{-16}\, \text{cm}^{-2} \, \text{s}^{-1} \, \text{sr}^{-1}\,. \end{equation} Assuming a monopole velocity with respect to the galaxy of $ \sim 10^{-3}c$, we can translate this bound into a limit on the monopole density, \begin{equation} n_{M} < 10^{-23} \, \text{cm}^{-3}\,, \end{equation} and thus $\gamma < 10^{-26}$.Then, from Eq.\ (\ref{eqn:Gammaf}) we have \begin{equation} \Gamma_f < 10^{-45}\,. \end{equation} Since this conflicts with Eq.\ (\ref{eqn:Gammaneeded}) by 10 orders of magnitude, we conclude that primordial bound states of magnetic monopoles cannot explain the UHECR. We note that we have used several approximations which overstate the possible value of $\Gamma_f$: First, we have considered the total classical radiative capture cross section. This takes into account not only the monopolonium formed with the right energy to decay at present, but all the possible binding energies, clearly overestimating the value of ${\Gamma}_f$. Second, it has been argued that the classical cross section given in Eq (\ref{eqn:radiative}) overestimates its real value due to photon discreteness effects\cite{Dicus82}. Finally, some of the monopolonium will have decayed before the present time, reducing the value of $\Gamma$. All of these effects make the conflict above more serious. \section{Monopoles connected by strings.} We present now a different scenario for the formation and annihilation of monopole-antimonopole bound states. The main problem in explaining the UHECR by the conventional magnetic monopolonium system is the inefficiency of the binding mechanism. This can be solved if we assume that all the monopoles get connected by strings in a later phase transition. Since the U(1) symmetry of the monopoles would be broken by the second phase transition, this U(1) must be a field other than the usual electromagnetism.\footnote{This is different from the Langacker-Pi scenario\cite{Langacker:1980kd}, where electromagnetism is broken and then restored at a lower temperature, and monopoles do feel large frictional forces.} We furthermore assume that these monopoles will not have any other unconfined charge, so that they will feel almost no frictional force moving in a background of particles. We take the comoving density of bound monopole systems $\Gamma$ to be constant. With a monopole mass of $10^{14}\, \text{GeV}$ the calculation of Sec.\ \ref{sec:needed} gives $\Gamma\sim 10^{-33}$, and with all monopoles bound, $\gamma = 2\Gamma$. The proper density at the time of string formation is then \begin{equation} n_M (T_s) =\gamma s ={2\pi^2\over 45} g_{*S} T_s^3 \gamma\sim 10^{-32} T_s^3\,. \end{equation} We can then compute the mean separation between monopoles at the time the string is formed, \begin{equation} L_i \sim {\left[n_{M}(T_s)\right]}^{-1/3}\,. \end{equation} If we take $T\sim 100\, \text{GeV}$, we obtain \begin{equation} L_i \sim 10^{-6} \, \text{cm}, \end{equation} which is much smaller than the horizon distance, $d_H \sim 3$ cm at $T\sim 100\, \text{GeV}$. We will assume that there are no light ($m\sim T_s$ or less) particles that are charged under the string flux. This means that there will be no charged particles that interact with the monopoles and cause the system to lose energy, so that gravitational radiation will be the only energy loss mechanism. When the strings are formed they may have excitations on scales smaller than the distance between monopoles, but these will be quickly smoothed out by gravitational radiation, leaving a straight string. The energy stored in the string is then $\mu L_i$, where $\mu \sim {T_s}^2$ is the energy per unit length of the string. This is smaller than the monopole mass by the ratio \begin{equation} {{\mu L_i}\over {m_M}} \sim 10^{-2} \end{equation} so the monopoles will move non-relativistically. In order to estimate the radiation rate we can assume that the monopoles are moving in straight lines. In fact, at the time of string formation the monopoles will have thermal velocities, so that in general the system will be formed with some non-zero angular momentum. However, in general this will be small compared to the linear motion due to the string tension, so we will ignore it, except to note that the monopoles will pass by without collision. The half oscillation of one monopole is parameterized by \begin{equation} x(t) = (2 a L)^{1/2} t - {1\over 2} a t^2 \end{equation} with $a = \mu / m_M$ and $0 < t < (8L/a)^{1/2}$. Using the quadrupole approximation,\footnote{The fully relativistic situation was considered in \cite{Martin:1997cp}.} the rate of energy loss of the system is \begin{equation} {{dE}\over{dt}} = {288 \over 45} G \mu^2 \left({{\mu L}\over {m_M}}\right) \end{equation} Since $\mu L$ is the energy in the string, we can integrate this equation to obtain \begin{equation} L = L_i e^{- t/\tau_g} \end{equation} with \begin{equation}\label{eqn:tau} \tau_g = {45 \over 288} {m_M \over G \mu^2 } = {45\over 288}{m_{pl}^2 m_M \over T_s^4}\,. \end{equation} The lifetime of the state will thus be $\tau_g\ln (L_i/r_M)$, where $r_M\sim m_M^{-1}$ is the radius of the monopole core. For $T\sim 100 \, \text{GeV}$ and $m_M\sim 10^{14} \, \text{GeV}$, Eq.\ (\ref{eqn:tau}) gives $\tau_g\sim 10^{17}\, \text{sec}$, comparable with the age of the universe. This suggests that the bound system formed by a monopole-antimonopole pair connected by a string can slowly decay gravitationally, and release the energy stored in the monopole in a final annihilation when the two monopole cores become close enough. \section{Conclusions} We have shown that is not possible to construct a consistent model for the origin of the UHECR based on the electromagnetic decay and final annihilation of magnetic monopole-antimonopole bound states formed in the early universe. We have obtained an upper limit for the monopolonium density today, taking into account its enhancement in the galactic halo and the maximum average free monopole density consistent with the Parker limit. Due to the small radiative capture cross section for the monopoles and the rapid expansion of the universe, the maximum density of monopolonium is many orders of magnitude below the concentration required to explain the highest energy cosmic ray events. We then proposed a different scenario in which the monopoles are connected by strings that form at a relatively low energy. This mechanism solves the problem of the inefficiency of the binding process, since every monopole will be attached to an antimonopole at the other end of the string. Due to the confinement of the monopole flux inside of the string , the main source of energy lose for these bound systems will be gravitational radiation. If we assume a monopole mass of $10^{14}\, \text{GeV}$ and a string energy scale of the order of $100 \, \text{GeV}$, the lifetime of the bound states would be comparable with the age of the universe, making them a possible candidate for the origin of the ultra-high-energy cosmic rays. \section{Acknowledgments} We would like to thank Alex Vilenkin for suggesting this line of work, and Xavier Siemens and Alex Vilenkin for helpful conversations. This work was supported in part by funding provided by the National Science Foundation. J. J. B. P. is supported in part by the Fundaci\'on Pedro Barrie de la Maza.
\section{Correction for PSF Anisotropy} \label{sec:anisotropy} In the foregoing we have computed how the shape polarization $q_\alpha$ responds to a shear. This allows us to correctly calibrate the circularizing effect of seeing. In general, asymmetry of the PSF will also introduce spurious systematic shape polarization $\langle q_\alpha \rangle$ in (\ref{eq:gammaestimator}), and it is crucial that this be measured and corrected for. As discussed in \S\ref{sec:introduction}, for unweighted second moments, the effect of PSF anisotropy is rather simple since the (flux normalized) final second moment $q_{lm}$ is the sum of the intrinsic second moment and that of the instrument, so the anisotropic parts of the instrumental second moment $q_\alpha = M_{\alpha l m} q_{lm}$ can be measured from stars, and subtracted from those observed. For weighted or isophotal moments things are more complicated, and depend on how the anisotropy is generated. KSB considered a simple model in which the final PSF is the convolution of some circularly symmetric PSF with a small, but highly asymmetric `anisotropizing kernel' $k(r)$. In \S\ref{subsec:convolutionmodel} we will further explore this `convolution model', the attempts that have been made to implement this perturbative approach, and estimate the error in this method. In \S\ref{subsec:recircularisation} we show that the convolution model is quite inappropriate when applied to diffraction limited seeing and we develop a more general technique for correcting PSF anisotropy. Finally, in \S\ref{subsec:noisebias} we draw attention to a type of noise related bias that affects shear estimators, but which has hitherto been overlooked, and we show how this can be dealt with. The artificial polarization produced by PSF anisotropy is an effect which is present in the absence of any gravitational shear. Thus at leading order in $\gamma$ we can set $\gamma = 0$. This effectively decouples the computation and correction of the PSF anisotropy from the shear-polarization calibration problem considered above. It also means that in this section we can assume that the intrinsic shapes of galaxies are statistically isotropic. \subsection{The Convolution Model} \label{subsec:convolutionmodel} KSB computed the effect of a PSF anisotropy under the assumption that this can be modeled as the convolution of a perfect circular PSF with some compact kernel $k(r)$. This would be a good model, for instance, for observations primarily limited by atmospheric turbulence with seeing disk of radius $r_g$, but with very small guiding or registrations errors or optical aberrations with extent $\delta r \ll r_g$, and the KSB analysis gives a correction to lowest order in $\delta r$. In this model, the effect of `switching on' the anisotropy is the transformation on the observed image, which is smooth on scale $\sim r_g$, \begin{equation} \label{eq:fobstaylorexpansion} f_{\rm o}(r) \rightarrow f_{\rm o}'(r) = \int d^2 r' \; k(r') f_{\rm o}(r - r') = f_{\rm o}(r) - k_i \partial_i f_{\rm o}(r) + {1 \over 2} k_{ij} \partial_i \partial_j f_{\rm o}(r) - {1 \over 6} k_{ijk} \partial_i \partial_j \partial_k f_{\rm o}(r) + \ldots \end{equation} where we have Taylor expanded $f_{\rm o}(r - r')$ and where $k_i \equiv \int d^2 r r_i k(r)$, $k_{ij} \equiv \int d^2 r r_i r_j k(r)$ etc. Taking the kernel to be centered, we have $k_i = 0$, and to lowest non-vanishing order the effect on the polarization is \begin{equation} \label{eq:ksbpolarisability1} q_\alpha' = \int d^2 r \; w_\alpha(r) f_{\rm o}'(r) = q_\alpha + {1\over 2} k_{ij} \int d^2 r\; f_{\rm o}(r) \partial_i \partial_j w_\alpha(r) \end{equation} where $w_\alpha \equiv M_{\alpha l m} w(r) r_l r_m$ and we have integrated by parts. The induced stellar polarization is linear in $k_{ij}$ and therefore scales as the square of the extent of the convolving kernel since $k_{ij} \sim \delta r^2$. Performing the decomposition $k_{ij} = k_A M_{Aij}$ we find that an individual galaxy's polarization $q_\alpha$ will depend on both the trace $k_0$ and the trace-free parts $k_\alpha$ of $k_{ij}$. The {\sl average} induced polarization, however, only depends on $k_\alpha$ and we have $\langle q_\alpha \rangle = k_\beta P^{\rm sm}_{\alpha \beta} $ with \begin{equation} \label{eq:ksbpolarisability2} P^{\rm sm}_{\alpha \beta} = {1\over 2} M_{\alpha l m} M_{\beta ij} \int d^2 r\; \langle f_{\rm o} \rangle \partial_i \partial_j (w(r) r_l r_m). \end{equation} To lowest order in the PSF anisotropy we can use the KSB polarizability (\ref{eq:ksbpolarisability2}) to infer $k_\alpha$ from the shapes of stars, and then correct the weighted second moments or the galaxies. At the same level of precision one can convolve one's image with a small kernel designed to nullify the anisotropy. \citeN{ft97} have presented a $3 \times 3$ smoothing kernel which does this. In the typical situation, the shapes are measured from an average of numerous images taken with a pattern of shifts, and which are co-registered to fractional pixel precision, and a simpler but equally effective approach is then to average over pairs of images which have been deliberately displaced from the true solution by the small displacement vector $\pm \delta r = \sqrt{|k_A|}\{\cos(\theta), \sin(\theta)\}$ with $\theta = \tan^{-1}(k_2/k_1)/2 + \pi / 2$. What is the error in this linearized approximation? To answer this we must consider higher terms in the expansion (\ref{eq:fobstaylorexpansion}). The next order correction to $q_\alpha$ involves $k_{ijk}$ which, unlike the centroid $k_i$ cannot be set to zero, so for a given object the fractional error in the KSB correction is on the order of $\delta r / r_0 \sim \sqrt{e_\alpha}$. Galaxies, however, are randomly oriented on the sky, so the average change in $q_\alpha$ for a galaxy of some arbitrary morphology but averaged over all position angles at this order in $\delta r / r_0$ is \begin{equation} \langle \delta q_\alpha \rangle \sim k_{ijk} \int d^2 r\; \langle f_{\rm o}(r) \rangle \partial_i \partial_j \partial_k w_\alpha(r). \end{equation} This vanishes since $\langle f_{\rm o}(r) \rangle$ is an even function while $\partial_i \partial_j \partial_k w_\alpha(r)$ is odd, provided we take $w(r)$ to be circularly symmetric at least, which we will assume is the case. The effective net fractional error in the KSB approximation is therefore on the order of $(\delta r / r_0)^2$, or typically on the order of the induced stellar ellipticity. \subsection{General PSF Anisotropy Correction} \label{subsec:recircularisation} In principle, one can develop a higher order correction scheme within the context of this model, but this does not seem to be particularly promising; it is not clear that the result will be sufficiently robust and accurate for even for ground based observations --- in very good seeing conditions the PSF anisotropy from aberrations becomes large, and the perturbative approach will break down. Moreover, for observations with telescopes in space, the model of the PSF as a convolution of a perfect circular PSF with a kernel, small or otherwise, is wholly unfounded. The instantaneous OTF is \begin{equation} \label{eq:diffotf} \tilde g(k) = \int d^2r \; A(r) A(r + k D \lambda / 2 \pi) \exp(i[\varphi(r) - \varphi(r + k D \lambda / 2 \pi)]) \end{equation} where $A$ is the real transmission function of the telescope input pupil, $D$ is the focal length, and $\varphi(r)$ is the phase error due to mirror aberrations and the atmosphere. It is often said that the general OTF factorizes into a set of terms describing the atmosphere; the telescope aperture; and aberrations, and this would seem to justify the convolution model discussed above. For atmospheric turbulence, and for aberrations arising from random small scale mirror roughness this is correct. This is because the average of the complex exponential term in (\ref{eq:diffotf}) depends only on $\delta r = k D \lambda / 2 \pi$ and is independent of $r$, and (\ref{eq:diffotf}) then factorizes into two independent terms, and the PSF is then the convolution of two completely independent and non-negative functions, but this is not the case in general. To be sure, one can write the combined aperture and aberration OTF as a product of some `perfect' OTF with some other function (the true OTF divided by the perfect one) but quite unlike the case for random turbulence and mirror roughness, this function is neither independent of the shape of the pupil, nor is it positive; if you compute this function for a telescope subject to a low order aberration from figure error, for instance, you will find that the this function is strongly oscillating and is just as extended as the true PSF. Consider the situation in WFPC2 observations, for example, where the PSF anisotropy has important contributions from both from asymmetry of the pupil $A(r)$ and from phase errors, though with the former tending to dominate (though not enormously so) for long wavelengths and far off axis with the WFPC2. The aperture function for the lower left corner of chip 2 computed from the Tiny-Tim model \cite{krist95} is shown in figure \ref{fig:wfpc2pupilandphase}. The off-axis pupil function is approximately that on-axis minus a disk of some radius $r_1$ at some distance $\Delta r$ off-axis. If we let the radius of the primary be $r_0$ and define a disk function $D_{r_0}(r) = \Theta(r / r_0 - 1)$ then on computing the electric field amplitude $a(x)$ and squaring we find that the on-axis PSF is of course given by the Airy disk: $g \simeq \tilde D_{r_0}^2$ while the off axis PSF $g'$ is given by $g'(x) \simeq g(x) - \tilde D_{r_1} \tilde D_{r_0} \cos(2 \pi x \Delta r / D \lambda)$. For $r_1 \ll r_0$, $\tilde D_{r_1}$ is relatively slowly varying and close to unity, in which case the extra off-axis obscuration introduces a perturbation which is proportional to the product of $D_0$ and a planar wave, or essentially an asymmetric modulation of the side lobes of the on-axis PSF (see LH panel of figure \ref{fig:gcircplot}). \begin{figure}[htbp!] \centering\epsfig{file=pupilphase.ps,width={1.0 \linewidth},angle=0} \caption{The panel on left shows the WFPC2 pupil function and phase error from Krist's Tiny Tim model for an off-axis point on the focal plane. The phase is given in radians for a wavelength of 800 nm. The panel on the right shows the modulation transfer function (the real part of the OTF). } \label{fig:wfpc2pupilandphase} \end{figure} In general then, and especially for diffraction limited observations, the perturbative convolution model cannot be trusted. Luckily, at least in the context of the shear estimators discussed here, a simple solution clearly presents itself: Since to compute the polarizability requires that one generate a rather detailed model of the PSF, and that we probably want to apply some kind of re-convolution $f_{\rm o} \rightarrow f_s = g^\dagger \otimes f_{\rm o}$ to obtain a well behaved shear response, we might as well use the opportunity to choose $g^\dagger(r)$ in order to re-circularize the PSF exactly. For example, from the observed PSF, one can compute the OTF $\tilde g(k)$ and then form the circularly symmetric function $\tilde g_{\rm min}(k)$ being the greatest real function which lies everywhere below $|\tilde g(k)|$. The function $g^\dagger(r)$ with transform \begin{equation} \tilde g^\dagger(k) = \tilde g_{\rm min}(k)^2 / \tilde g(k) \end{equation} both guarantees a well defined shear polarizability and gives a perfectly circular resulting total PSF. This is illustrated in figure \ref{fig:gcircplot}. \begin{figure}[htbp!] \centering\epsfig{file=gcircplot.ps,width={1.0 \linewidth},angle=0} \caption{HST WFPC2 PSF from Tiny Tim and re-circularizing filter computed as described in the text.} \label{fig:gcircplot} \end{figure} In some situations the PSF may have near 180 degree symmetry, in which case a good approximation is to re-convolve with a PSF which has been rotated through 90 degrees. Since the PSF is real, the OTF must satisfy the symmetry $\tilde g(k) = \tilde g^*(-k)$. In the absence of aberrations, a diffraction limited telescope has OTF $\tilde g = A \oplus A$ which is real and so has exact symmetry under rotation by $\pi$, a property which is shared the PSF, so $R_\pi g = g$. If one convolves with $g^\dagger = R_{\pi/2} g$ then the resulting total PSF $g^\dagger \otimes g$ is symmetric under 90 degree rotations (proof: $R_{\pi/2} \tilde g_{\rm tot} = R_{\pi/2} (R_{\pi/2} \tilde g \tilde g) = R_\pi \tilde g R_{\pi/2} \tilde g = \tilde g R_{\pi/2} \tilde g = \tilde g_{\rm tot}$). For a galaxy of given form, the average PSF induced polarization is \begin{equation} \langle q_\alpha \rangle = \int d^2 r \; \langle g_{\rm tot} \otimes f \rangle w_\alpha = [\langle g_{\rm tot} \otimes f\rangle \otimes w_\alpha]_{r = 0} = \int {d^2 k \over (2 \pi)^2} \langle \tilde f \rangle \tilde g_{\rm tot} \tilde w_\alpha \end{equation} where $w_\alpha = M_{\alpha l m} w(r) r_l r_m$. But $ \langle \tilde f \rangle $ is circularly symmetric and $\tilde w_\alpha$ is anti-symmetric under 90 degree rotations, so $\langle q_\alpha \rangle = 0$. Other situations where this would work would be ground based observations but with large amplitude telescope oscillations, and in fast guiding. In general, however, aberrations from figure errors will introduce both real and imaginary contributions to the OTF and the latter will destroy the exact 180 degree symmetry of the PSF. This is certainly the case for the WFPC2 PSF at optical wavelengths. In the linearized approximate schemes described earlier, the PSF anisotropy is entirely characterized by the two coefficients $k_\alpha$. In general, these will vary substantially with position on the image, but this can be treated by modeling $k_\alpha$ as some smoothly varying function of image position $R$ such as a low order polynomial. In the scheme proposed here we need to be able to generate not just these two coefficients, but the full two dimensional PSF $g(r, R)$ (being the intensity at distance $r$ from the centroid of a star at position $R$). A simple practical approach is to solve for a model $g(r,R) = \sum_I g_I(r) f_I(R)$, with $f_I(R)$ being some set of polynomial or other basis functions. A least squares fit for the image valued mode coefficients is $g_I(r) = m_{IJ}^{-1} G_J(r)$ with $m_{IJ} = \sum_{\rm stars} f_I(R) f_J(R)$; $G_I(r) = \sum_{\rm stars} f_I(R) g_{\rm obs}(r)$. Armed with the image valued coefficients one can then, for example, compute the convolution $f_s = g \otimes f_{\rm o}$ as $f_s(r) = \sum_I f_I(r) (g_I \otimes f_{\rm o})$, that is we convolve the source image with each of components $g_I(r)$ and then combine with spatially varying coefficients $f_I(R)$. \hide{This is only valid if the scale length for changes of the PSF greatly exceeds its own width, but that is usually an excellent approximation.} Non-linear functions of the PSF such as the re-circularizing kernel $g^\dagger$ are easily generated by making realizations of the PSF models on say a coarse grid of points covering the source image, and from each of these computing the desired function, and then fitting the results to a low order polynomial just as for $g(r,R)$. The approach described here lends itself very nicely to observations with large mosaic CCD cameras, in which misalignment of the chip surfaces and the focal plane coupled with telescope aberrations can give rise to a PSF which varies smoothly across each chip, but which changes discontinuously across the chip boundaries. This results in a very complicated PSF pattern on the final image obtained by averaging over many dithered images. It is a good deal simpler to re-circularize each of the contributing images. We still need to generate a spatially varying model of the final $g(r)$, $g^\dagger(r)$ in order to compute the polarizability, but if we fail to model these exactly, it will only introduce a relatively minor error in the shear-polarization calibration. \input noisebias.tex \section{Discussion} \label{sec:discussion} We have considered the problem of how to estimate weak gravitational shear from observations which have been degraded by atmospheric and/or instrumental effects. Previous analyses of this problem have made simplifying assumptions which render the results inaccurate. A major result of the paper is the finite resolution shear operator (\ref{eq:fobsoperator}) which gives the response of an observed image to a gravitational shear applied before smearing with the PSF. This result can be used to properly calibrate the effect of any shear estimator, and is valid for arbitrary PSF, be it turbulence or diffraction limited. We then focused on the application to weighted moment shear estimators. We have computed the response of individual objects to a shear in \S\ref{subsec:individualresponse}, and the response of the population of background galaxies with given photometric properties in \S\ref{subsec:populationresponse}, and from this we have devised an optimal weighting scheme \S\ref{sec:optimisation}. In the last section we have considered the correction for PSF anisotropy. While there are still some approximations in the present analysis, we feel that they place the techniques of shear measurement on a much firmer footing than before. \section{Acknowledgements} I wish to acknowledge helpful discussions with Malcolm Northcott, Francois and Claude Roddier, John Tonry, Ger Luppino, Pat Henry, Ken Chambers, Jeff Kuhn and Christ Ftaclas. \subsection{Atmospheric Dispersion} Variation of the refractive index of the atmosphere with wavelength will cause images to be dispersed into spectra, and the consequent elongation of images was estimated by KSB, who pointed out that this could potentially cause problems when using PSF's of stars to correct the shapes of galaxies, since the elongation depends on the spectrum of the object; an object with a line-like spectrum, or one with a sharp-edged absorption band which falls within the bandpass, will of course be less dispersed than a continuum object. Here we will make this more quantitative. Weak-shear measurements are usually made on images composed from multiple exposures taken at a range of air masses and which are co-registered using centroids of foreground stars. The astrometric solution obtained using stars of a variety of spectra will be a compromise. Assuming, as is usually the case, that one has many more stars at ones disposal than the number of coefficients of the transformation one is trying to solve for (typically a low order polynomial), then the solution obtained by minimising least squared residuals will be one which correctly maps surface brightness at some wavelength $\lambda_0$ determined by the average properties of the foreground stars, but with photons at other wavelengths displaced by an amount proportional to $\lambda - \lambda_0$. The sum of a set of $N$ such images is \begin{equation} f_{\rm tot}(r, \lambda) = \sum\limits_{I=1}^N f(r + \alpha_I (\lambda - \lambda_0), \lambda) \end{equation} where $f(r, \lambda)$ is the image seen at zenith. The 2-vector $\alpha_{Ii}$ has $|\alpha_I| = (\Theta_0 / \lambda_0) (\partial \ln \Theta / \partial \ln \lambda) \tan z_I$, with $z_I$ the zenith distance, and $\hat \alpha_I$ is directed towards the horizon. Tables of the refraction angle $\Theta(\lambda)$ are given by \citeN{allen73}. A star with SED $f_\lambda$, and therefore with photon distribution function $S(\lambda) d \lambda \propto \lambda f_\lambda d \lambda$, gives, in a photon counting system, a response at zenith $f(r,\lambda) = \delta(r) S(\lambda)$, and the summed image of such a star is then \begin{equation} f_{\rm tot}(r) = \int d \lambda f_{\rm tot}(r, \lambda) = \sum\limits_{I=1}^N\int d\lambda \delta(r - \alpha_I (\lambda - \lambda_0)) S(\lambda) \end{equation} The centroid of a star in the image $f_{\rm tot}(r)$ is an average of the centroids that would be obtained from each of the contributing images. Consider the $I$'th image, and rotate the coordinate system so that $\alpha_I$ lies along the $x$-axis. The centroid for this component is $\overline r_I = ({\overline x}_I, 0)$ with \begin{equation} {\overline x}_I = { \int dx \int dy \int d\lambda \; x \delta(x - \alpha_I(\lambda - \lambda_0)) \delta(y) S(\lambda) \over \int dx \int dy \int d\lambda \; \delta(x - \alpha_I(\lambda - \lambda_0)) \delta(y) S(\lambda) } = {\alpha_I \int d\lambda\; (\lambda - \lambda_0) S(\lambda) \over \int d\lambda\; S(\lambda)} = \alpha_I ({\overline \lambda} - \lambda_0) \end{equation} from which it follows that the $\lambda_0$ which minimises $\langle \overline x_I^2 \rangle$, being the average over the stars used for registration of the squared displacement, is $\lambda_0 = \langle \overline \lambda \rangle = {1 \over n_{\rm stars}} \sum \overline \lambda$, i.e.~simply the average over the registration stars of their mean wavelength. The centroid for the summed image is \begin{equation} \overline r_i = ({\overline \lambda} - \lambda_0) \langle \alpha_i \rangle_I \end{equation} where $\langle \alpha_i \rangle_I \equiv {1 \over N} \sum \alpha_{Ii}$. Similarly, the central second moment is \begin{equation} \label{eq:pijdispersion} p_{ij} = {\int d^2 r \; (r_i - \overline r_i)(r_j - \overline r_j)f_{\rm tot}(r) \over \int d^2 r f_{\rm tot}(r)} = \overline{(\lambda - \lambda_0)^2} \langle \alpha_i \alpha_j \rangle_I - (\overline \lambda - \lambda_0)^2 \langle \alpha_i \rangle_I \langle \alpha_j \rangle_I \end{equation} Note that since, for any sensible zenith angle, the width of the spectrum is tiny compared to that of the PSF, this unweighted second moment fully characterises the effect of dispersion. Consider, for illustration, the case of an equatorial field observed with a telescope near the equator, and for a range of zenith angle $|z| < z_{\rm max}$. In this case $\sum \alpha_I = 0$, so the second term in (\ref{eq:pijdispersion}) vanishes (this is not a particularly unusual special case; for a number of observations spread over a range of zenith angles, one would generally expect the second term here to become relatively small). When we solve for the PSF from the shapes of foreground stars of various types we are effectively averaging over a mix of SED's appropriate for a low redshift spiral galaxy. The average PSF moment obtained from the stars can, for the equatorial case, then be written as \begin{equation} \langle p_{xx} \rangle = (\langle \overline{\lambda^2} \rangle - \langle \overline \lambda \rangle^2) \langle \alpha^2 \rangle_I \end{equation} where by $\langle \alpha^2 \rangle_I = (\Theta_0 / \lambda_0)^2 (\partial \ln \Theta / \partial \ln \lambda)^2 {1\over N} \sum_I \tan^2 z_I$, and the SED dependent coefficient of $\langle \alpha^2 \rangle_I$ is the mean of the dispersions of the stars plus the dispersion of the means of the stars. For the equatorial case this is the same as the second moment for a single point like object with an SED like the average SED of the registration stars. This is not strictly true in general, but the difference is probably a minor one, and it then follows that if the faint galaxies have SED's such that their $\overline{(\lambda - \lambda_0)^2}$ differs systematically from that for a low redshift spiral then the PSF correction will be systematically in error. Figure \ref{fig:dispersion} quantifies the importance of this effect, by using redshifted galaxy SED's of various types from \citeN{cww80}. The panel on the left shows the displacement of the centroid of galaxies of the various indicated types as their spectra are red-shifted, and the right hand panel shows the variation of the second moment of the PSF. We find that the I-band second moment is very stable, with peak fluctuations $\delta p_{lm} \sim 300 {\rm mas}^2$ for these parameters, and with little systematic difference from a low spiral redshift galaxy if we integrate over a range of redshifts. For illustration, if we take the systematic change in polarisation to be say $\delta p_{lm} \sim 100 {\rm mas}^2$ and a Gaussian profile galaxy, this corresponds to a shear of $\gamma \simeq 1.38 \delta p_{lm} / {\rm FWHM}^2 \simeq 6 \times 10^{-4} (0''.5 / {\rm FWHM})^2$. In the V-band the polarisation fluctuations are larger by a factor 2-3, but even then, and even for marginally resolved objects in excellent seeing the effect is at or below the sensitivity for current and near future surveys. The effect scales as $\langle \tan^2 z \rangle$ however, so observations at $z \gg 1 $ should be avoided. This would also become more of an issue with fast guiding, where we can hope to resolve much smaller objects. The effect is also dependent on the details of the system response function; the CFH12K system having a particularly broad I-band response which exacerbates the effect. \begin{figure}[htbp!] \centering\epsfig{file=disperse.ps,width={1.0 \linewidth},angle=0} \caption{Atmospheric dispersion. Left hand panels show the angular displacement due to changes in the SED for galaxies at a range of redshifts. A zenith angle of 45 degrees was assumed, and upper and lower panels show the deflection for the I-band and V-band system response functions for the CFH12K camera. More concretely, the quantity plotted is the mean wavelength $\overline \lambda - \lambda_0 = \int d \lambda \; (\lambda - \lambda_0) \lambda f_\lambda R(\lambda) / \int d \lambda \; \lambda f_\lambda R(\lambda)$ where $f_\lambda$ is the SED, $R(\lambda)$ the system response, $\lambda_0$ was taken to be $820$nm, and $\overline \lambda - \lambda_0$ was converted to angle as described in the text, and assuming a 4000m observatory altitude. The V-band plot has been limited to $z < 1.5$ since the CWW SED's are limited to proper wavelength $\lambda > 200$nm and redshift past the $V$-band at around $z = 1.5$. The plots shows that the shifts due to SED variation are typically on the order of $20$mas in the I-band, but somewhat larger in V. The elliptical/S0 SED shows a somewhat greater excursion at high redshift, but this it caused by the SED red-shifting right out of the filter band, so one would not expect to find many such objects in flux limited samples. The panels on the right show how the width of the spectrum, and hence the PSF polarization, varies with redshift and galaxy type. Here the quantity plotted is $ \overline{(\lambda - \lambda_0)^2} = \int d \lambda \; (\lambda - \lambda_0)^2 \lambda f_\lambda R(\lambda) / \int d \lambda \; \lambda f_\lambda R(\lambda)$, again converted to angle, here assuming $\langle \tan^2 z \rangle = 1$. } \label{fig:dispersion} \end{figure} In the foregoing we assumed that the foreground stars used for registration were numerous and well mixed. With a finite number of stars there will be some additional color dependent error in the registration which, if uncorrected, would give rise to artificial field distortion. However, we have found from simulations \hide{The relative shift of centroids of stars according to their colors can also cause artificial distortion in the image registration. At a zenith distance of 45 degrees (air-mass of 1.414), for instance, and for standard broadband colors we find that centroids shift by about 10mas per magnitude in $V-I$ color, and } using the USNOA catalog as the astrometric reference system that this introduces artificial shear at around the $10^{-4}$ level, which is negligibly small for present and near future surveys. \section{Introduction} \label{sec:introduction} Weak lensing provides a probe of the dark matter distribution on a range of scales from galaxy halos, through clusters of galaxies, to large-scale-structure (see e.g.~the recent review of \citeN{mellier99}, and references therein). In the weak lensing or thin lens approximation, the effect of gravitational lensing on the image of a distant object is a mapping of the surface brightness: \begin{equation} \label{eq:lensmapping} f'(r_i) = f((\delta_{ij} - \psi_{ij}) r_j) \end{equation} where the 2-vector $r_i$ is the angular position on the sky measured relative to the center of the image, $f$ is the intrinsic surface brightness that would be seen in the absence of lensing, and the symmetric `distortion tensor' $\psi_{ij}$ is an integral along the line of sight of the transverse second derivatives of the gravitational potential $\Phi$ \cite{gunn67}. In an open cosmology, for instance, the distortion for an object at conformal distance $\omega_s$ can be written as \begin{equation} \label{eq:psifromphi} \psi_{lm} = 2 \int d \omega {\sinh \omega \sinh (\omega_s - \omega) \over \sinh \omega_s} \partial_l \partial_m \Phi \end{equation} (\citeNP{barkana96}; \citeNP{k98}) where $\partial_i \equiv \partial / \partial x_i$ and $x_i = \theta_i \sinh \omega$ with $\theta_i$ being a 2-component Cartesian vector in the plane of the sky and where the potential is related to the density contrast by $\nabla^2 \Phi = 4 \pi G \delta \rho$, the Laplacian here being take with respect to proper spatial coordinates. Equation (\ref{eq:psifromphi}) can be generalized to deal with sources at a range of distances, and with either accurately known redshifts or partial redshift information from broadband colors. The distortion will, in general, change both the shapes and sizes, and hence luminosities, of distant objects. Any component of the distortion which is coherent over large scales --- larger than the typical angular separation of background galaxies --- is therefore potentially observable as a relative modulation of the counts of objects or as a statistical anisotropy of the galaxy shapes, and this allows one to constrain the fluctuations in the total density $\delta \rho$ on the corresponding scales. Here we will focus on analysis of shape anisotropy, or `image shear', though the methodology is readily extendible to include the effects of magnification. While the effect of a shear on the sky surface brightness (\ref{eq:lensmapping}) is rather simply stated, no completely satisfactory method for estimating the distortion has yet emerged. Perhaps ideally one would attack this problem using likelihood; that is, one would ask: what is the probability to observe a given set of background galaxy images given that they are drawn from some statistically isotropic unlensed parent distribution, as a function of the parameters $\psi_{lm}$. Unfortunately, this does not seem to be a particularly tractable problem. Also, the problem is further complicated by the finite resolution of real observations, and by noise in the images. Instead, what has been done is to adopt some plausible shape statistic --- typically some kind of central second moment --- and then compute how this responds to a gravitational shear. We will now review the various different shear estimators that have been proposed, how they relate to one another, and what their limitations are. \subsection{Projection Matrices and the Shear Operator} As a preliminary, we now introduce some mathematical formalism which will simplify the analysis. Symmetric $2 \times 2$ tensors like $\psi_{lm}$ feature prominently in what follows. Such tensors have 3 real degrees of freedom. For instance, it is conventional to parameterize the three real degrees of freedom of $\psi_{lm}$ by the triplet comprising the convergence $\kappa$ and the shear $\gamma_\alpha$, $\alpha = 1,\;2$ with \begin{equation} \psi_{ij} = \left[\matrix{ \kappa + \gamma_1 & \gamma_2 \cr \gamma_2 & \kappa - \gamma_1 }\right] \end{equation} A simple way to convert between triplet and tensor components is to use the three constant $2 \times 2$ `projection matrices': \begin{equation} \label{eq:Mdefinition} M_{0ij} = \left[ \matrix{ 1 & 0 \cr 0 & 1 } \right] \quad\quad\quad\quad M_{1ij} = \left[ \matrix{ 1 & 0 \cr 0 & -1 } \right] \quad\quad\quad\quad M_{2ij} = \left[ \matrix{ 0 & 1 \cr 1 & 0 } \right] \end{equation} The symmetrized products of pairs of these matrices $[M_A M_B] \equiv {1\over 2} (M_{A l m} M_{B m n} + M_{A n m} M_{B m l})$ have multiplication table \begin{equation} [M_A M_B] = \left[\matrix{ M_0 & M_1 & M_2 \cr M_1 & M_0 & 0 \cr M_2 & 0 & M_0 }\right] = \delta_{AB} M_0 + (\delta_{A0} \delta_{B\alpha} + \delta_{A\alpha}\delta_{B0}) M_\alpha \end{equation} from which follow the useful identities that the contractions of products and triple products are \begin{eqnarray} \label{eq:traceMAMB} M_{Alm} M_{Bml} & = & 2 \delta_{AB} \\ \label{eq:traceMAMBMC} M_{Alm} M_{Bmn} M_{Cnl} & = & 2(\delta_{BC} \delta_{A0} + \delta_{AC} \delta_{B0} + \delta_{AB} \delta_{C0}). \end{eqnarray} Any symmetric tensor $t_{lm}$ can be written as a linear combination of projection matrices with coefficients $t_A$, that is, $t_{lm} = t_A M_{Alm}$, and using (\ref{eq:traceMAMB}) we have $M_{Blm}t_{lm} = M_{Blm} M_{Alm} t_A = 2 \delta_{AB} t_A$, so $t_A = {1\over 2} M_{Alm} t_{lm}$. In this language the convergence and the shear are the three components of the triplet representation of the distortion tensor: $\kappa = {1 \over 2} M_{0lm} \psi_{lm} = \psi_0$ and $\gamma_\alpha = {1 \over 2} M_{\alpha lm} \psi_{lm} = \psi_\alpha$, $\alpha = 1,2$. We will adopt the convention that upper case Roman indices range over $0,1,2$ while lower case Greek symbols range over $1,2$ and that repeated indices are to be summed over. A two component `polar' $t_\alpha$ transforms under rotations as $t_\alpha \rightarrow R_{\alpha \beta}(2 \theta) t_\beta$ while $t_0$ transforms as a scalar under rotations. An alternative and widely used formalism \cite{sef92} is to regard $\gamma_1,\;\gamma_2$ as the real and imaginary parts of a complex shear, but we shall not adopt that approach here. It is also convenient to define a `shear operator' $S_\gamma$, which generates the mapping (\ref{eq:lensmapping}). \begin{equation} f' = S_\gamma f \end{equation} At linear order in $\gamma$, which will be valid for sufficiently weak shear, one can perform a first order Taylor expansion of the RHS of (\ref{eq:lensmapping}) and $S_\gamma$ becomes the differential operator \begin{equation} \label{eq:linearshearoperator} S_\gamma = 1 - \gamma_\alpha M_{\alpha i j} r_i \partial_j \end{equation} where $\partial_j \equiv \partial / \partial r_j$. This operator is rather similar to a rotation operator. An important question is what is the domain of validity of (\ref{eq:linearshearoperator}). The answer depends on the content of the image to which it is applied. For an image containing information only at spatial frequencies below some upper limit $k_{\rm max}$ this will be a good approximation provided $r \ll 1 / (\gamma k_{\rm max})$, so for finite shear (\ref{eq:linearshearoperator}) will not apply for spatial frequencies $\gtorder 1 / (\gamma r)$. The limit here is that combination of frequency and distance from the origin that a shear of strength $\gamma$ corresponds to a local translation of about an inverse wavenumber. We would however expect (\ref{eq:linearshearoperator}) to correctly describe the effect of shear on the low frequency behavior of an image, in the sense that if applied to an image which has had the high frequency information removed, then the result will be essentially identical to the low-frequency content of an exactly sheared image. \subsection{Second Moment Shear Estimators for Perfect Seeing} A pure shear will cause a circular object to become elliptical, and will change the ellipticities of non-circular objects. Following \citeN{vjt83} a natural choice of statistic to measure such a distortion is the second central moment or quadrupole moment \begin{equation} \label{eq:secondmomentdefinition} q_{lm} = \int d^2 r \; r_l r_m f(r) \end{equation} where we will assume that the total flux, which is unaffected by a pure shear, has been normalized such that $\int d^2 r \; f(r) = 1$, and where the origin of coordinates has been taken such that the dipole moment $d_l = \int d^2 r \; r_l f(r)$ vanishes. The triplet coefficients of the symmetric matrix $q_{lm}$ are \begin{equation} \label{eq:pdefinition} q_A = {1 \over 2} M_{Aij} q_{ij} =\left[\begin{matrix}{ (q_{xx} + q_{yy})/2 \cr (q_{xx} - q_{yy})/2 \cr q_{xy} }\end{matrix}\right] \end{equation} The first component $q_0$ is a measure of the size, or area, of the object, while the latter two $q_\alpha$ are a measure of the eccentricity of the object --- they vanish for a circular object --- which we will refer to as the `polarization'. We can compute how the quadrupole moment is affected by a shear by applying (\ref{eq:linearshearoperator}) to $f$ in (\ref{eq:secondmomentdefinition}) to find \begin{equation} q'_{lm} = q_{lm} - \gamma_\alpha M_{\alpha i j} \int d^2 r \; r_l r_m r_i \partial_j f = q_{lm} + 2 \gamma_\alpha M_{\alpha l i} q_{im} \end{equation} where we have integrated by parts and invoked the symmetry ($M_{\alpha i j} = M_{\alpha j i}$) and tracelessness ($M_{\alpha i i} = 0$) of the matrices $M_1$, $M_2$. With $q_{im} = M_{Bim} q_B$ etc., and using(\ref{eq:traceMAMBMC}), we find \begin{equation} \delta q_A = {1\over 2} M_{\alpha l m} (q'_{lm} - q_{lm}) = \gamma_\alpha M_{Alm} M_{\alpha l i} M_{Bim} q_B = 2 \gamma_\beta(\delta_{A0} q_\beta + \delta_{A\beta} q_0) \end{equation} or equivalently \begin{equation} \label{eq:deltap} \begin{matrix}{ q'_\alpha = q_\alpha + 2 \gamma_\alpha q_0 \cr q'_0 = q_0 + 2 \gamma_\alpha q_\alpha }\end{matrix} \end{equation} A weak gravitational shear therefore causes a change in the the polarization $\delta q_\alpha = 2 q_0 \gamma_\alpha$, which is proportional to the area $q_0$, and it also induces a change in the area $\delta q_0$ which is proportional to the eccentricity. For intrinsically randomly oriented galaxies the average intrinsic polarization $\langle q_\alpha \rangle$ vanishes by symmetry, so $\langle q'_\alpha \rangle = 2 \langle q_0 \rangle \gamma_\alpha $ and $\langle q'_0 \rangle = \langle q_0 \rangle$ and therefore \begin{equation} \label{eq:simpleshearestimate} \gamma_\alpha = \langle q'_\alpha \rangle / 2 \langle q'_0 \rangle \end{equation} which which one can use to estimate the shear on a patch of the sky by replacing the averaging operator $\langle \ldots \rangle$ by a summation: $\sum \ldots / N$. This type of shear estimator was first introduced and used by \citeN{vjt83} in their search for large scale shear. The averages of $q_\alpha$ and $q_0$ here are heavily weighted towards larger galaxies, which is not ideal. More commonly, what has been done (e.g.~\citeNP{tvjm84}) is to normalize the polarization by the trace of the second moments and define the `ellipticity vector' \begin{equation} \label{eq:naiveshearestimator} e_\alpha = q_\alpha / q_0 \end{equation} which depends only on the galaxy shape, and whose expectation value is \begin{equation} \langle e'_\alpha \rangle = \left\langle {q_\alpha + 2 \gamma_\alpha q_0 \over q_0 + \gamma_\beta q_\beta} \right\rangle \simeq 2 \gamma_\alpha -2 \gamma_\beta \langle q_\alpha q_\beta / q_0^2 \rangle = 2 \gamma_\alpha (1 - \langle e^2 \rangle / 2) \end{equation} and where we have kept only terms up to linear order in the shear (\citeNP{ksb95}, hereafter KSB). An alternative \cite{bm95} is to normalize the second moments by the square-root of the determinant of $q_{lm}$ rather than by the trace: \begin{equation} \label{eq:BMshearestimator} e^{\rm BM}_\alpha = {q_\alpha \over \sqrt{|q_{lm}|}} = {q_\alpha \over \sqrt{q_0^2 - q_\alpha q_\alpha}} \end{equation} which, again at first order in $\gamma$, has expectation value \begin{equation} \langle e^{\rm BM}_\alpha \rangle = 2 \gamma_\alpha \left\langle \sqrt{1 + ( e^{\rm BM}_\alpha )^2}\right\rangle. \end{equation} Yet another possibility is to use only the information in the position angle $\theta = (\tan^{-1} q_2 / q_1) / 2$ \cite{kochanek90}, or equivalently in the unit shear vector $|e_\alpha|$, and there are numerous other similar statistics one could use. These formulae for the response of the polarization statistics should be used with caution. While formally correct, the averages here must be taken in an unweighted manner over all galaxies. This is generally neither possible nor particularly desirable, as there are detection limits, and one would ideally like to estimate the shear as some optimally weighted combination of polarizations of galaxies of different fluxes and sizes, and these simple relations no longer hold. We will return to this later. First, however, let us consider the effect on these estimators of a finite point spread function, which is well illustrated by the simple estimator (\ref{eq:simpleshearestimate}). \subsection{Second Moment Shear Estimators for Finite Seeing} In real observations, some or all of atmospheric turbulence, optical aberrations; aperture size; guiding or registration errors; atmospheric dispersion; finite pixel size; scattering etc.~will combine to give an observed surface brightness \begin{equation} \label{eq:fobsdefinition} f_{\rm o}(r) = \int d^2 r'\; g(r') f(r - r') \equiv g \otimes f \end{equation} where $g(r)$ is the point spread function (PSF). The combined effect of gravitational shearing followed by instrumental and atmospheric seeing is the transformation \begin{equation} \label{eq:fobsprimedefinition} f_{\rm o}' = g \otimes S_\gamma f \end{equation} Circular seeing will tend to reduce the ellipticity while departures from circularity will introduce an artificial polarization. To make accurate shear measurements we need to correct for the latter and calibrate the former. For estimators like (\ref{eq:simpleshearestimate}), which are computed from moments $q_{lm}$ as defined in (\ref{eq:secondmomentdefinition}), the effect of the PSF is rather simple since, as noted by \citeN{vjt83}, the second central moment of a convolution of two normalized functions is just the sum of the second central moments: \begin{equation} \label{eq:momentsum} q_{lm}(f_{\rm o}) = q_{lm}(f) + q_{lm}(g) \end{equation} and this additive property is shared by the independent components $q_A$. Thus one can recover the second moments $q_A(f)$ that would be measured by a large perfect telescope in space from the observed moments $q_A(f_{\rm o})$ simply by subtracting the moments of the PSF $q_A(f) = q_A(f_{\rm o}) - q_A(g)$. These may be measured from shapes of foreground stars in the image, or, in the case of diffraction limited seeing, computed from the telescope design. In terms of observed moments, the shear estimator (\ref{eq:simpleshearestimate}) becomes \begin{equation} \label{eq:calibratedsimpleestimator} \hat \gamma_\alpha = {\langle q_\alpha - q^\ast_\alpha \rangle \over 2 \langle q_0 - q_0^\ast \rangle } \end{equation} where superscript $\ast$ denotes the value for a stellar object. This procedure --- using (\ref{eq:momentsum}) to correct the measured moments of the PSF --- simply and rather elegantly compensates for both the circularizing and distorting effects of realistic seeing. \subsection{Weighted Moment Shear Estimators} Unfortunately, while very simple to analyze, the shear estimator constructed from the second moments as defined in (\ref{eq:secondmomentdefinition}) is not at all useful when applied to real data. For one thing, photon counting noise introduces an uncertainty in the moments which diverges as the square of the radius out to which one integrates, and the effect of neighboring objects will similarly grossly corrupt the signal. There is also the problem that for the kinds of PSFs that arise in real telescopes, the second moment is not well defined. To obtain a practical estimator, it is necessary to truncate the integral in (\ref{eq:secondmomentdefinition}). This can be done in various ways; the approach implemented in the FOCAS software package \cite{jt81} and also the Sextractor package \cite{ba96} is to truncate the integral at some isophotal threshold $f_{\rm crit}$ and compute moments of the non-linear function $F(f_{\rm o}) = \Theta(f_{\rm crit} - f_{\rm o}) f_{\rm o}$, where $\Theta(f)$ is the Heaviside function. In fact, isophotal moments are most commonly computed from an image which has been smoothed with some kernel --- usually an approximate model of the PSF itself --- in order that the isophotal boundary be well defined. An alternative, as advocated by \citeN{bm95} and by KSB, is to limit the range of integration with a user supplied weight function $w(r)$ and define \begin{equation} \label{eq:weightedmomentdefinition} q_{lm} \propto \int d^2 r\; w(r) r_l r_m f_{\rm o}(r). \end{equation} Another possibility is to define a polarization vector in terms of the second derivatives of a smoothed image $f_s = w \otimes f$ as \begin{equation} q_{lm} \propto \partial_l \partial_m \int f_s \end{equation} evaluated at the peak of $f_s$. For a Gaussian weight function $w(r)$, however, this is essentially equivalent to the weighted second moment (\ref{eq:weightedmomentdefinition}). As we shall see, the weighted moment statistics, being linear in the surface brightness, offer significant advantages over the isophotal threshold method, but in either case, the simple relation (\ref{eq:momentsum}) between observed and intrinsic second moments no longer holds, and compensating for the effects of the PSF becomes considerably more complicated than equation (\ref{eq:calibratedsimpleestimator}). A partial solution to this problem has been offered by KSB, who computed the response of shear estimators like (\ref{eq:naiveshearestimator}) to an anisotropy of the PSF under that assumption that this can be modeled as the convolution of a circular PSF $g_{\rm circ}(r)$ with some compact but possibly highly anisotropic function $k(r)$. This would be a reasonable approximation for example for the case of atmospheric seeing in the presence of small amplitude guiding errors. They found that such a PSF anisotropy would introduce an artificial ellipticity $\delta e_\alpha = P^{\rm sm}_{\alpha\beta}(f_{\rm o}) p_\beta$ where $p_\beta \equiv M_{\beta l m} \int d^2r \; r_l r_m k(r)$ is the unweighted polarization of $k(r)$ (though see \citeN{hfks98} who corrected a minor error in the analysis). The `smear polarizability' $P_{\alpha\beta}^{\rm sm}$ is a combination of weighted moments of $f_{\rm o}$, and is essentially a measure of the inverse area of the object. An interesting feature of this analysis is that only the second moment of the convolving kernel appears here; all other details of $k(r)$ are irrelevant, and it is relatively straightforward to determine $p_\alpha$ from observed stars, set up a model for how this two-vector field $p_\alpha(r)$ varies across the field, and then correct, at linear order at least, the ellipticities to what would have been measured by a telescope with a perfectly circular PSF. KSB also computed the response of the ellipticity to a shear applied to $f_{\rm o}$ {\sl after\/} smearing with the PSF and found $\delta e_\alpha = P^{\gamma}_{\alpha\beta}(f_{\rm o}) \gamma_\beta(q)$ where $P^\gamma_{\alpha\beta}$ is another combination of moments of $f_{\rm o}$ . This of course is not what one wants, as one really needs the response to a shear applied before smearing with the PSF, and they suggested that one use deep images from the Hubble Space Telescope to empirically deduce a correction for finite seeing. \hide{ There are various ways one can proceed. One possible approach for calibrating ground based observations is to take , subject these to an artificial shear and then degrade the observations to simulate ground-based seeing, as was done by \citeN{ksb95}. By measuring how the shapes respond to the input shear one can empirically determine a response function or `shear polarizability' $P^\gamma = \partial \langle q \rangle / \partial \gamma$ which gives the desired calibration factor. The main limitation of this approach is the paucity of deep HST fields, which makes it hard to obtain a truly representative sample of objects. Another problem is that this procedure cannot be used to calibrate HST observations themselves, and additionally, there may be slight problems in calibrating ground-based observations due to differences in the filter transmission functions. } A related approach, suggested by \citeN{wcf96a} is to iteratively deconvolve the images, apply a shear and then re-convolve and again use the change in the polarization with applied shear to calibrate the relation between $\gamma$ and the polarization measured from the original images. \citeN{lk97} have used a somewhat different approach. They noted that the real operation (\ref{eq:fobsprimedefinition}) can be written as $g \otimes S_\gamma f = S_\gamma((S_\gamma^{-1} g) \otimes f)$ i.e.~applying a shear before smearing is equivalent to smearing with an anti-sheared PSF $S_\gamma^{-1} g$ and then shearing. Now if the PSF is Gaussian, applying a weak shear to it is precisely equivalent to smearing it with another small but anisotropic Gaussian, so the effect of this can therefore be computed using the smear polarizability of KSB, and it then follows that for a nearly circular Gaussian PSF, the operation (\ref{eq:fobsprimedefinition}) will cause a response \begin{equation} e_\alpha \rightarrow e_\alpha' = e_\alpha + \delta e_\alpha = e_\alpha + P^\gamma_{\alpha\beta} \gamma_\beta + P^{\rm sm}_{\alpha\beta} p_\beta(S_\gamma^{-1} g) \end{equation} where $p_\beta(S_\gamma^{-1} g)$ is the unweighted polarization of the anti-sheared PSF and is of first order in $\gamma$. One can most easily infer the value of $p_\beta(S_\gamma^{-1} g)$ from the values of the shear and smear polarizabilities for a stellar object: shearing a point source has no effect, so for a star we must have $p_\alpha = - (P^\gamma(g) / P^{\rm sm}(g)) \gamma_\alpha$ where we have suppressed the indices on the stellar polarizabilities --- for a nearly circular PSF these are approximately diagonal --- and hence the net effect of a real shear in this approximation is \begin{equation} \label{eq:lkpolarisability} \delta e_\alpha = (P_{\alpha\beta}^\gamma(f_{\rm o}) - (P^\gamma(g) / P^{\rm sm}(g)) P_{\alpha\beta}^{\rm sm}(f_{\rm o})) \gamma_\beta. \end{equation} This is nice, as it expressed the linear response of the polarization $e_\alpha$ to a shear entirely in terms of the observables $f_{\rm o}$ and $g$, but it rests somewhat shakily on the assumption that shearing a realistic PSF can indeed be modeled as smearing it with some compact kernel. For a Gaussian PSF this is exact, but this is a rather special, and unfortunately unrealistic, case. One indication that (\ref{eq:lkpolarisability}) cannot apply for a general PSF comes from considering the factor $(P^\gamma(g) / P^{\rm sm}(g))$. At no point have we specified the scale of the weighting function $w(r)$, so this factor must be invariant to choice of scale length. For a Gaussian PSF this is indeed the case, but for a PSF generated by atmospheric turbulence for instance this is not the case, and (\ref{eq:lkpolarisability}) is then inconsistent. Similarly, the factor $q_\alpha(g) / P^{\rm sm}(g)$ appearing in the KSB correction for PSF anisotropy is, in general, dependent on the scale of the window used to measure the PSF properties. \citeN{hfks98} have found from analysis of simulated images that for the HST WFPC2 instrument one can adjust the scale of the weight function for the stars to render the calibration (\ref{eq:lkpolarisability}) and the KSB anisotropy correction reasonably accurate, but it is not clear that this will apply in general. Indeed, for diffraction limited seeing, the inadequacy of the KSB formalism has a deeper root. While the assumption that the real PSF can be modeled as a convolution of a circular PSF with some compact kernel $k(r)$ may be a good approximation for atmospheric turbulence seeing with small amplitude guiding errors and such-like, as we shall see, this is not the case in general. The current situation is therefore somewhat unsatisfactory. In this paper we will develop an improved method of shear estimation which does not suffer from the inadequacies noted above and works for a PSF of essentially arbitrary form. The layout of the paper is as follows: In \S\ref{sec:operators} we construct an operator which generalizes (\ref{eq:linearshearoperator}) to finite size PSF and which generates the effect of a gravitational shear on the observed (i.e.~post-seeing) surface brightness $f_{\rm o}$. We first show that, quite generally, the effect on $f_{\rm o} = g \otimes f$ of a shear applied to $f$ is equivalent to a shear applied directly to $f_{\rm o}$ plus a `commutator' term which is a convolution of $f_{\rm o}$ with a kernel $\gamma_\alpha H_\alpha(r)$ where $H_\alpha(r)$ may be computed from the PSF $g$. We explore the properties of this kernel for various types and PSF. \hide{ and find, for instance, that for seeing arising from atmospheric turbulence the kernel is not compact, thus invalidating the calibration procedure using (\ref{eq:lkpolarisability}). } The kernel $H_\alpha$ involves $\ln \tilde g$ (where the tilde denotes the Fourier transform), which appears, particularly in the case of diffraction limited seeing, to be formally ill defined since $\tilde g$ vanishes at finite radius. We show however, that the operator which generates the effect of a shear on a filtered image which has had frequencies close to the diffraction limit attenuated does not suffer from this problem. In \S\ref{sec:estimators} we specialize to weighted second moments, and compute their response to shear, both for individual objects \S\ref{subsec:individualresponse} and for the population of galaxies of a given flux, size and shape \S\ref{subsec:populationresponse}. In \S\ref{sec:optimisation} we show how to optimally combine estimates of the shear from galaxies of various different types. The method is tested using simulated mock images. In \S\ref{sec:anisotropy} we discuss the correction for PSF anisotropy, and propose a new technique involving measuring shapes from images which have been convolved with a re-circularizing PSF. We draw attention to a hitherto ignored noise related bias and show how this can be analyzed and corrected for. \hide{In \S\ref{sec:acf} we show how these results can be extended to calibrate shear estimators derived from the auto-correlation function of the sky brightness.} In \S\ref{sec:discussion} we summarize the main results, and outline how they can be applied to real data. In the analysis here we will draw heavily on the properties of real physical PSFs, and we include as an appendix a brief review of basic PSF theory and discussion of PSF properties as they relate to weak lensing observations. \subsection{Noise Bias} \label{subsec:noisebias} In the absence of noise, the recircularization procedure exactly annuls any PSF anisotropy; the {\sl signal\/} content of the image is exactly statistically isotropic. The {\sl noise\/} component in the images (which in the original images is incoherent Poisson noise) will however be correlated with an anisotropic two-point function -- the peaks and troughs of the noise will appear as ellipses with correlated position angles. For a given object, the noise is equally likely to produce a positive or negative fluctuation in the polarization $q_\alpha$, and the effects cancel out on average. Due to the nature of faint galaxy counts, however, objects of a given observed flux are more likely to be intrinsically fainter objects which have been scattered upward in flux than intrinsically brighter objects which have been scattered down, and there will therefore be a tendency for faint objects to be aligned like the re-circularizing PSF; i.e.~oriented opposite to the PSF. We will refer to this as `noise-bias'. It should noted that an analogous effect is present in the old KSB anisotropy correction scheme. In that case the noise is isotropic, but objects are detected as peaks of a smoothed image and there is then a tendency for the galaxies close to the threshold for detection to be aligned like the PSF. In the context of perturbative schemes as described in \S\ref{subsec:convolutionmodel} there are other unevaluated errors in the correction which are typically of similar order to the noise bias effect, which goes some way to explain why it has been ignored in the past. In the improved PSF anisotropy correction method proposed here it is the leading source of error and it behooves us to analyze and correct for it. Noise will affect the photometric parameters like $q_\alpha$ in two ways; in addition to a straightforward additive noise term, noise will also affect the object detection, and will shift the centroid about which we measure the centered second moments, for instance. In fact, the de-centering effect is quite weak. To see why, consider the following simple (though actually quite practical and realistic) model for the detection and measurement process in which we take the re-circularized image $f_s$ and define objects as peaks of the field $F(r) = w \oplus f_s$ where $w$ is a some smooth, circularly symmetric weight function (we typically use a compensated `Mexican hat' filter), and the object moments $q_A$ are the value of the field $q_A(r) = w_A \oplus f_s$ evaluated at the peak location. Let us assume that there exists some object which, in the absence of noise, would lie at the origin, so for this object we would have \begin{equation} \begin{matrix}{ F = F(0) = \int d^2r \; w(r) f_s(r) \cr 0 = d_i(0) = \int d^2r \; (\partial_i w) f_s(r) \cr q_A = q_A(0) = \int d^2r \; w_A(r) f_s(r) }\end{matrix} \end{equation} where the condition that the object be a peak of $F$ is expressed as the vanishing of the dipole-like quantity $d_i(r) = \partial_i F$. If we now add noise, and let $f_s \rightarrow f_s + f_n$, where $f_n(r)$ is the noise component of the re-circularized image, all of the photometric fields $F(r)$, $d_i(r)$, $q_A(r)$ will change to $F' = F + \delta F = F + w \oplus f_n$ etc. The dipole $d_i$ will no longer vanish at the origin, but will have some value $d_i(0) = \int d^2r\; (\partial_i w) f_n$, and the peak of $F(r)$ will have moved to some position $r_{\rm pk} \ne 0$. In the vicinity of the peak we have $d_i(r) \simeq 0 + (r - r_{\rm pk})_j \times \partial_j d_i(r)$ and hence, to first order in the noise amplitude, the peak location is \begin{equation} r_{i,\rm pk} = - (\partial_j d_i)^{-1} \delta d_j = - \left[\int d^2 r\; \partial_i \partial_j w f_s \right]^{-1} \int d^2r\; \partial_j w f_n \end{equation} The shift of the peak has no effect on $F$ at first order (because $F$ is stationary) but is does affect the central moments $q_A$ and we have \begin{equation} \label{eq:deltaFqnoise} \begin{matrix}{ \delta F = \int d^2r \; w(r) f_n(r) \cr \delta q_A = \int d^2 r \left[w_A(r) - v_{Aj} \partial_j w\right] f_n }\end{matrix} \end{equation} where \begin{equation} v_{Ai} = \left[ \int d^2r \; (\partial_i \partial_j w) f_s \right]^{-1} \int d^2 r \; (\partial_i w_A) f_s \end{equation} Thus, in general, the first order change in the second moments $q_A$ depends not just on the noise and the kernel $w_A$, but on the form of the underlying noise-free image $f_s$ through the term involving $v_{Aj}$. There is good reason to think that this form dependence is weak for real objects; the function $\partial_i w_A$ is an odd function, so the extra term vanishes if the galaxy is symmetric under 180 degree rotations. Also, it is hard to see why the presence of this term would cause a systematic polarization; the change in the polarization $\delta q_A$ associated with the centroid shift is proportional to the vector $v_{Aj}$ which, being a function of the re-circularised image $f_s$ is equally likely to be positive or negative. For each galaxy that, for a given realisation of noise, suffers a certain $\delta q_A$ there is a 180 degree rotated clone which has precisely the same weighted flux, polarization etc.~(these being even functions) but which suffers a $\delta q_A$ of opposite sign. The kind of effect envisioned in the leading paragraph of this sub-section arises if there is a correlation between fluctuations in the polarisation and the flux $F$, coupled with a gradient of the density of objects $n(F)$. The first term in $\delta q_A$ in (\ref{eq:deltaFqnoise}) does, as we shall see, indeed correlate with $\delta F$, but, when we average over orientation of the underlying noise-free galaxies, the second does not. Unfortunately, we have not been able to rigorously demonstrate that the centroid shift term vanishes. Nonetheless, in what follows we will assume that this term is negligible. This should certainly be adequate in order to estimate the magnitude of the effect, and is probably sufficiently accurate to give a useful correction, though the results should strictly be regarded as only approximate. Let us now assume one has measured a set of $n$ photometric parameters $p_i$ for a galaxy which, like $F$, $q_0$, $q_\alpha$, are linear functions of the brightness $f_{\rm o}$ and of the form $p_i = \int d^2r \; K_i(r) f_{\rm o}(r)$. The effect of noise in the images will be to introduce a perturbation in these parameters $\delta p_i$. These perturbations have zero mean, $\langle \delta p_i \rangle = 0$, and since there are typically a very large number of photons (noise plus signal) contributing to the parameters $p_i$, the central limit theorem dictates that the $\delta p_i$ will have a multivariate Gaussian distribution: \begin{equation} \sigma(\delta p) d^n\delta p = ((2 \pi)^n |C|)^{-1/2} \exp(- \delta p_i C^{-1}_{ij} \delta p_j / 2) d^n \delta p \end{equation} where the covariance matrix is $C_{ij} = \langle \delta p_i \delta p_j \rangle$ and is given by \begin{equation} \label{eq:covariance} C_{ij} = \int d^2 r \; K_i (K_j \otimes \xi_n) \end{equation} where $\xi_n(r) = \langle f_n(r') f_n(r' + r) \rangle$ is the 2-point function of the noise $f_n(r)$, and which may be computed as described in appendix \ref{sec:PSFtheory}. For objects which are detected at a fairly high level of significance $\nu \gg 1$, the noise will cause a small modification of the underlying distribution function: $n(p) \rightarrow n'(p)$, which we can compute in a perturbative manner: \begin{equation} n'(p) = \int d^n \delta p \; n(p - \delta p) \sigma(\delta p) = n(p) - {\partial n \over \partial p_i} \langle \delta p_i \rangle + {1\over 2} {\partial^2 n \over \partial p_i \partial p_j} \langle \delta p_i \delta p_j \rangle + \ldots \end{equation} Since $\sigma(p)$ is an even function, all the odd terms in this expansion vanish, and we have $n'(p) = n(p) + {1\over 2} C_{ij} \partial^2 n / \partial p_i \partial p_j$ at leading order. Let us specialize now to the $n=4$ dimensional case: $p_i = F, q_0, q_\alpha$ and compute the mean polarization for galaxies of given $F$, $q_0$ with some weight function $W(q = |q_\alpha|)$: \begin{equation} \langle q_\alpha \rangle_{F, q_0} = {\int d^2q \; W(q) q_\alpha n'(F, q_0, q_\alpha) \over \int d^2q \; W(q) n'(F, q_0, q_\alpha)} \simeq {\int d^2q \; W(q) q_\alpha \delta n(F, q_0, q_\alpha) \over \int d^2q \; W(q) n(F, q_0, q_\alpha)} \end{equation} where $\delta n(F, q_0, q_\alpha) = {1\over 2} C_{ij} \partial^2 n / \partial p_i \partial p_j$. Since $q_\alpha$ is an odd function and $W(q)$, $n(F, q_0, q_\alpha)$ are even, the only terms which contribute to the integral in the numerator are those involving a single derivative with respect to polarization: $\partial^2 n /\partial F \partial q_\beta$ and $\partial^2 n /\partial q_0 \partial q_\beta$ and, on integrating by parts, we have \begin{equation} \langle q_\alpha \rangle = {- {1\over 2} \int d^2q \; \left(\langle \delta F \delta q_\beta \rangle {\partial n \over \partial F} + \langle \delta q_0 \delta q_\beta \rangle {\partial n \over \partial q_0}\right) {\partial W(q) q_\alpha \over \partial q_\beta} \over \int d^2q \; W(q) n(F, q_0, q_\alpha)} \end{equation} Using $\partial(W(q) q_\alpha) / \partial q_\beta = \delta_{\alpha \beta} W(q) + q_\alpha \hat q_\beta dW / dq$, integrating over angle and converting from $n(F, q_0, q_\alpha)$ to $n'(F, q_0, q) = 2 \pi q n(F, q_0, q_\alpha)$ and dropping the prime we have \begin{equation} \langle q_\alpha \rangle = {-{1\over 2} \int dq \; \left(\langle \delta F \delta q_\alpha \rangle {\partial n \over \partial F} + \langle \delta q_0 \delta q_\alpha \rangle {\partial n \over \partial q_0}\right) (W(q) + {1\over 2} q W'(q)) \over \int dq \; n W(q)} \end{equation} and letting $W(q)$ become a delta-function to isolate a single value of $q$ we have \begin{equation} \label{eq:noisebias} n \langle q_\alpha \rangle = - {1\over 2} \left(\langle \delta F \delta q_\alpha \rangle {\partial n \over \partial F} + \langle \delta q_0 \delta q_\alpha \rangle {\partial n \over \partial q_0}\right) \end{equation} Thus, as anticipated, there is a net induced polarization if there is a non-zero correlation between the polarization fluctuation $\delta q_\alpha$ and $F$ and/or $q_0$, and also significant gradients of the distribution function with respect to $F$ or $q_0$. The effect on the re-scaled polarization of the first term in (\ref{eq:noisebias}) is $\langle q'_\alpha \rangle = \langle q_\alpha \rangle / F = -({1\over 2} \langle \delta F \delta q_\alpha \rangle / F^2) \partial \ln n / \partial /ln F$ and is second order in the inverse significance: $\langle q'_\alpha \rangle \propto \nu^{-2}$, where $\nu^2 = F^2 / \langle \delta F^2 \rangle$ and rapidly becomes small for well observed objects. Thus it should be possible to set a sensible limit on the significance of object detection, and, if necessary, use (\ref{eq:noisebias}) together with (\ref{eq:covariance}) for $\langle \delta F \delta q_\alpha \rangle$ etc.~to correct for this. The error in this linearized approximation is of fourth order in the inverse significance. As a simple illustrative example consider a Gaussian galaxy of scale $r_G$, a Gaussian window function $w$ with scale $r_w$, and a Gaussian ellipsoid PSF $g = \exp(-(x^2 / r_a^2 + y^2 / r_b^2)/2)$ with $r_a = r_g(1 + \epsilon / 2)$, $r_b = r_g (1 - \epsilon / 2)$. A suitable recircularising kernel is then $g^\dagger = \exp(-(x^2 / r_b^2 + y^2 / r_a^2)/2)$ and the normalised two point function of the noise is then \begin{equation} \xi_n(r) = {1 \over \pi r_a r_b} e^{-(x^2 / 2 r_b^2 + y^2 / 2 r_a^2)/2} \end{equation} The expectation averages are, to first order in $\epsilon$, $\langle \delta F^2 \rangle = 2 (r_w^2 + r_g^2)$; $\langle \delta q_1 \delta F \rangle = r_w^4 (r_a^2 - r_b^2) / (r_w^2 + r_g^2) = 2 \epsilon r_w^4 r_g^2 / (r_w^2 + r_g^2)$, and so taking only the first term in (\ref{eq:noisebias}) for simplicity \begin{equation} \langle \delta q_1' \rangle \equiv {\langle \delta q_1 \rangle \over F} = - {1\over 2} {\partial \ln n \over \partial \ln F} {1 \over \nu^2} {\langle \delta q_1 \delta F \rangle \over \langle \delta F^2 \rangle} \simeq {1\over 2} {\epsilon \over \nu^2} r_w^4 r_g^2 / (r_w^2 + r_g^2) \end{equation} using $\partial \ln n / \partial \ln F \simeq 1$ as observed. A shear of strength $\gamma$ applied to the Gaussian galaxy produces a polarisation \begin{equation} \langle \delta q_1' \rangle = 4 \gamma r_w^4 r_G^2 / (r_G^2 + r_w^2 + r_g^2) \end{equation} therefore a PSF anisotropy of strength $\epsilon$ is equivalent to an effective shear \begin{equation} \gamma = {\epsilon \over 8 \nu^2} {r_g^2 \over r_G^2} {(r_w^2 + r_g^2 + r_G^2)^2 \over (r_w^2 + r_g^2)^2} \end{equation} or, for $r_w = r_G = \sqrt{2} r_g$ say, $\gamma = 25 \epsilon / 144 \nu^2$, which for $\nu = 6$ and PSF asymmetry $\epsilon = 0.3$, which is a reasonable value for off-axis points on the CFHT in good seeing, gives a shear of around $1.4\%$. This is a sizable effect, and should therefore be corrected for. \section{Finite Resolution Shear Operator} \label{sec:operators} We now consider how the shear operator (\ref{eq:linearshearoperator}) is modified by finite resolution arising either in the atmosphere, telescope optics or detector. Let the unperturbed surface brightness (i.e.~that which would have been observed in the absence of lensing) be \begin{equation} f_{\rm o} = g \otimes f \end{equation} so the perturbed surface brightness is \begin{equation} \label{eq:perturbedsb} f_{\rm o}' = g \otimes S_\gamma f \end{equation} Fourier transforming, and using the result that, at linear order in $\gamma$, applying a shear in real space is equivalent to applying a shear of the opposite sign in Fourier space (i.e.~if $a = S_\gamma b$ then $\tilde a = S_{-\gamma} \tilde b$, where $\tilde F (k)\equiv \int d^2 r \; F(r) e^{ik\cdot r}$), we have \begin{equation} \tilde f_{\rm o}' = \tilde g S_{-\gamma} \tilde f = \tilde f_{\rm o} - \tilde g \delta S_{\gamma}(\tilde f_{\rm o} / \tilde g) \end{equation} where $\delta S_\gamma \equiv S_\gamma - 1$. Since $\delta S_\gamma$ is a 1st order differential operator we have $\delta S_\gamma(\tilde f_{\rm o} / \tilde g) = \tilde g^{-1} \delta S_\gamma f_{\rm o} - \tilde g^{-2} \tilde f_{\rm o} \delta S_\gamma \tilde g$, and $\tilde g^{-1} \delta S_\gamma \tilde g = \delta S_\gamma \ln \tilde g$. Consequently, to first order in $\gamma$ \begin{equation} \label{eq:deltatildefobs} \tilde f_{\rm o}' = \tilde f_{\rm o} - \delta S_\gamma \tilde f_{\rm o} + \tilde f_{\rm o} \delta S_\gamma \ln \tilde g \end{equation} or, in real space, \begin{equation} f_{\rm o}' = f_{\rm o} + \delta S_\gamma f_{\rm o} - (\delta S_\gamma h) \otimes f_{\rm o} \end{equation} where $h$ is the inverse transform of $\tilde h \equiv \ln \tilde g$, i.e.~the logarithm of the optical transfer function (OTF). Invoking the definition of the shear operator (\ref{eq:linearshearoperator}), we have \begin{equation} \label{eq:fobsoperator} f_{\rm o}' = f_{\rm o} - \gamma_\alpha M_{\alpha i j} (r_i \partial_j f_{\rm o} - (r_i \partial_j h) \otimes f_{\rm o}). \end{equation} Note that the PSF is a real function so $\tilde g(-k) = \tilde g^*(k)$, and this symmetry is shared by $\ln \tilde g$, so $h$ is also real. The finite resolution shear operator (\ref{eq:fobsoperator}) is then $f_{\rm o}' = S_\gamma f_{\rm o} - (\delta S_\gamma h) \otimes f_{\rm o}$, that is, it is the regular shear operator $S_\gamma$ applied to the post-seeing image $f_{\rm o}$ plus a `commutator term' $g\otimes S_\gamma f - S_\gamma(g \otimes f)$ which is a correction for finite PSF size and which is a convolution of $f_{\rm o}$ with a kernel $\gamma_\alpha H_\alpha(r) = \gamma_\alpha M_{\alpha i j} r_i \partial_j h = \delta S_\gamma h$ which one can compute from the PSF $g$. This seems quite promising; the effect of the first term on the polarization of an object is just $\gamma_\alpha$ times the KSB post-seeing shear polarizability. The response of the polarization to the second term should also be calculable; if the kernel is very compact then the response will be given by the KSB linearized smear polarizability, but even if it is not, (\ref{eq:fobsoperator}) should still allow one to compute the polarization response since it expresses the change in $f_{\rm o}$, and therefore in the polarization, or indeed any other statistic computable from $f_{\rm o}$, directly in terms of the observed surface brightness $f_{\rm o}$ itself. The finite resolution shear operator involves the function $h(r)$ which is the transform of the logarithm of the OTF. Since the OTF becomes exponentially small or may vanish, two questions immediately arise: Is the function $h(r)$ mathematically well defined? and can it be reliably computed from PSFs measured from real stellar images? To address these questions we now explore the form of this `commutator kernel' $H_\alpha(r)$ for various types of PSF. \subsection{Gaussian Ellipsoid PSF} \label{subsec:gaussianshearoperator} Consider first the simple though unrealistic case of a Gaussian ellipsoid PSF: $g(r) = (2 \pi)^{-1} |m|^{-1/2} \exp(-r_i m_{ij}^{-1} r_j / 2)$, where $m_{ij} = \langle r_i r_j \rangle$ is the matrix of central second moments. In this case, $\ln \tilde g(k) = -k_i m_{ij} k_j/ 2$, and so $\delta S_\gamma \ln \tilde g(k) = \gamma_\alpha M_{\alpha i m} m_{mj} k_i k_j$ and on transforming this we find that the kernel is the operator $H_\alpha = M_{\alpha i m} m_{mj} \partial_i \partial_j$ (in which case the function $H_\alpha(r)$ can be realized, for example, as the contraction of the constant matrix $\gamma_\alpha M_{\alpha i m} m_{m j}$ with the limit as $\sigma \rightarrow 0$ of the matrix of second partial derivatives of a Gaussian ball $(2\pi \sigma^2)^{-1} \exp(- r^2 / 2 \sigma^2)$), and therefore \begin{equation} \label{eq:GaussianPSFoperator} f_{\rm o}' = f_{\rm o} - \gamma_\alpha M_{\alpha i j} (r_i \partial_j f_{\rm o} + m_{il} \partial_l \partial_j f_{\rm o}). \end{equation} So, for a Gaussian PSF, the finite resolution shear operator is well defined and is a purely local differential operator. Gaussian PSF's are, however unphysical, and do not arise in real instruments. \subsection{Atmospheric Turbulence} \label{subsec:atmosphericshearoperator} Now consider atmospheric turbulence limited seeing. As reviewed in appendix \ref{sec:PSFtheory}, in that case $\tilde g(k) = \exp(- S(k D \lambda / 2 \pi)/2)$ where for fully developed Kolmogorov turbulence the structure function is $S(r) = 6.88 (r / r_0)^{5/3}$ with $r_0$ the Fried length, so in this case $\tilde h \propto \ln \tilde g \propto - k^{5/3}$. The commutator kernel therefore involves the transform of this power-law, which diverges strongly at high $k$, so how do we make sense of this? Dimensional analysis would suggest that $h(r) = \int d^2k \; k^{5/3} e^{ik \cdot r}$ be a power law with $h(r) \propto r^{-11/3}$. The same argument, however, applied to a Gaussian would suggest $h(r) \propto r^{-4}$, which we know to be false as we have just shown that in that case $h(r)$ is just the second derivative of a $\delta$-function and has no extended tail. To clarify the situation, and to verify the validity of the power-law form for $h(r)$ for atmospheric seeing, consider the function $h(r;R)$ which is the transform of $k^{5/3}$ times an exponential cut-off function $\exp(-kR)$, that is \begin{equation} h(r; R) \propto \int d^2 k k^{5/3} e^{-kR} e^{ik \cdot r} \end{equation} so $h(r)$ is the limit as $R \rightarrow 0$ of $h(r;R)$. If we write this as an integral with respect to rescaled wavenumber $y = kR$ it becomes clear that $h(r;R)$ obeys a self-similar scaling with respect to choice of $R$ and can be expressed in terms of some universal function $F(y)$ such that $h(r;R) = R^{-11/3} F(r/R)$. If we postulate that $h(r;R)$ tends to some $R$-independent limit for finite $r$ as $R \rightarrow 0$ then that limit must be a power law with $F(z) \propto z^{-11/3}$ so $h(r)$ must be proportional to $r^{-11/3}$. This argument does not indicate the coefficient multiplying the power law which, for a Gaussian, happens to vanish. The value of $h$ at the origin is $h(0;R) = 2 \pi R^{-11/3} \Gamma(8/3)$, i.e.~on the order of the value of the $r \gg R$ power law asymptote extrapolated to $r \sim R$ (note that for the Gaussian the analogous gamma function is not defined). A numerical integration for various values of $R$ is shown in figure \ref{fig:turbulenthkernel}. This shows that as we decrease $R$, the function $h(r;R)$ does indeed tend to a $r^{-11/3}$ power law with finite non-zero $R$-independent amplitude, but that the power law breaks (with a change of sign) at $r \sim R$ and becomes asymptotically flat for $r \ll R$. For finite $R$ this function is well characterized as a positive `softened $\delta$-function' core with width $\sim R$, central value $\sim R^{-11/3}$ and therefore with weight proportional to $R^{-5/3}$, surrounded by a negative power law halo with $h \propto r^{-11/3}$. In the limit $R \rightarrow 0$, the core shrinks and becomes negligible, leaving only the power law halo. This power law has the same slope as the well known large angle $g(r) \propto r^{-11/3}$ form of the PSF, but the PSF departs from this law at angular scale $r_g \sim \lambda / r_0$ corresponding to the Fried length, whereas $h(r)$ is a perfect power law and shows no features at the Fried scale. \begin{figure}[htbp!] \centering\epsfig{file=hkernel.ps,width={0.8 \linewidth},angle=0} \caption{Numerical calculation of $h(r; R) = \int d^2 k \; k^{5/3} \exp(-kR + i k \cdot r)$ for a range of $R$ values increasing by factors of 2. Also shown is a pure $r^{-11/3}$ power law. } \label{fig:turbulenthkernel} \end{figure} The function $h(r)$ diverges strongly at the origin and the same is true of the kernel $H_\alpha = M_{\alpha i j} r_i \partial_j h$, so $H_\alpha = - 11/3 r^{-11/3} \{\cos 2 \varphi, \sin 2 \varphi\}$. This does not give rise to any inconsistency with (\ref{eq:fobsoperator}) however. The observed sky $f_{\rm o}$ is coherent on the scale of the PSF $r_g$, so in computing the contribution to the commutator term \begin{equation} \delta f_{\rm o} = \gamma_\alpha M_{\alpha i j} \int d^2r' \; r'_i r'_j (r')^{-17/3} f_{\rm o}(r - r') \end{equation} at small $r' \ll r_0$, we can perform a Taylor expansion of $f_{\rm o}$, and we find that the first non-vanishing term is the second order term $\delta f_{\rm o} \sim (\partial_i \partial_j f_{\rm o})_r \int d^2r'\; r'_i r'_j (r')^{-11/3}$ which has no physical divergence at $r' = 0$. Somewhat unfortunately perhaps, the same line of argument shows that the commutator term cannot in this case be assumed to be a convolution with some compact function $k(r)$. A necessary condition for this to be valid is that the unweighted second moment of the kernel $p_\alpha = M_{\alpha l m} \gamma_\beta \int d^2r\; r_l r_m H_\beta(r)$ should be well defined and tend to a finite limit within some small radius $\ll r_g$, the characteristic width of the PSF. But this is the same integral as above, which does not tend to any well defined value, but rather diverges as the $1/3$ power of the upper limit on the integration radius. Thus while the $h(r) \propto r^{-11/3}$ asymptote is quite steep, it is not sufficiently steep to render the $p_\alpha$ value well defined. This strictly invalidates the approximation of \citeN{lk97} for turbulence limited seeing. The Kolmogorov law is only expected to apply over a finite range of scales. The structure function will fall below the $r^{5/3}$ form at the `outer scale', which recent measurements at La Silla suggest to be typically on the order of 20m (\citeN{mtz+98}, though see also the review of earlier results in \citeN{azb+97}). Fast guiding (often referred to as `tip-tilt' correction) would also effectively reduce the structure function at these frequencies, and such effects act in a similar manner to the simple exponential cut-off we have assumed. To compute the OFT properly, one should include the effect of the aperture. The detailed form of $h(r)$ at very small $r$ will be sensitive to the detailed form of the outer scale cut-off and/or aperture, but provided these lie at scales much greater than the Fried length (which is the case for large aperture telescopes at good sites) the effect on the commutator term should be almost independent of the cut-off because any signal at the relevant spatial frequencies will have been attenuated by a large factor. There will also be deviations from Kolmogorov structure function law at small separation; diffusion will damp out fine scale turbulence, and mirror roughness will add additional high spatial frequency phase errors. These effects will modify the $h \propto r^{-11/3}$ halo at large $r$, just as they modify the large-angle $g \propto r^{-11/3}$ form of the PSF. These effects may profoundly influence the behavior of $h$, $H_\alpha$ at large angle, but have little impact on the type of shape statistics considered here, where the large-angle contribution to the polarization is suppressed by the weight function or the isophotal cut-off. \subsection{Diffraction Limited Seeing} \label{subsec:diffractionshearoperator} Let us now consider diffraction limited observations. In this case the OTF $\tilde g$ falls to zero at finite spatial frequency --- the diffraction limit --- and so $\ln \tilde g$ diverges as one approaches the diffraction limit from below and is not defined for higher frequencies. \hide{This divergence is similar to that which afflicts the deconvolution operator $\tilde g^{-1}$, and immediately raises worries about whether convolution with $H$ would be well defined and the possibility of amplification of noise.} It is easy to understand how this divergence arises physically. As noted above, applying a weak shear in real space is equivalent to applying a shear of the opposite sign in Fourier space. Thus the information in some Fourier mode of the sheared image comes from a slightly displaced mode in the unsheared image. If the OTF $\tilde g$ is finite and continuous, as is the case for atmospheric turbulence limited seeing, then the image $f_{\rm o}$ contains all of the information required to predict $f_{\rm o}'$. For diffraction limited seeing however, and for observations through a narrow band filter, the OTF has a well defined edge, so the information contained in $f_{\rm o}$ for spatial frequencies just inside the cut-off may lie outside the cut-off in $f_{\rm o}'$ and so will be missing. For a finite band pass the form of the OTF will be modified, but must still fall to zero at the diffraction limit for the highest frequencies passed by the filter and $\ln \tilde g$ is still formally divergent. Thus, in general, from knowledge of $f_{\rm o} = g \otimes f$, it is strictly speaking not possible to say how $f_{\rm o}$ would change in response to a small but finite shear applied before seeing; there are Fourier modes within a distance $\delta k \sim \gamma k_{\rm max}$ of the diffraction limit that one cannot predict. As an extreme example, imagine we have a pure sinusoidal ripple on the sky which lies just outside the diffraction limit and which is therefore invisible. Applying an appropriate shear can bring that mode inside the limit and the ripple will appear as if from nowhere. As applied to real signals, however, this formally divergent behavior does not present a serious problem. First, if we observe galaxies of overall extent $r_G$, which is typically not much larger than the seeing disk, then the transform must vary smoothly with coherence scale $\delta k \sim 1/ r_G$ which, for sufficiently small $\gamma$ will greatly exceed $\gamma k_{\rm max}$; this rules out the possibility of isolated spikes lurking just beyond the diffraction limit as in the example. Also, for filled aperture optical telescopes the marginal modes are quite strongly attenuated, and, for any finite measurement error from e.g.~photon counting statistics, will contain very little information. The situation here is similar to that encountered in analyzing atmospheric turbulence PSF; there, while the very small scale details of $h(r)$ are sensitive to the aperture or outer scale cut-off, when applied to real data they have essentially no effect. Similarly, if one can generate a function which accurately coincides with $\tilde h = \log \tilde g$ where $g(k)$ exceeds some small value, but which tapers off smoothly at larger frequencies (rather than diverging at $k_{\rm max}$), then this should have a well defined transform and should give what is, for all practical purposes, a good approximation to the true finite resolution shear operator. One way to explicitly remove the divergence is to compute shapes from an image which has had the marginally detectable modes attenuated. If one re-convolves the observed field $f_{\rm o}$ with some filter function $g^\dagger(r)$ to make a smoothed image $f_{\rm s} = g^\dagger \otimes f_{\rm o}$ then in Fourier space we have \begin{equation} \tilde f_{\rm s}' = \tilde f_{\rm s} - \tilde g^\dagger \delta S_\gamma \tilde f_{\rm o} + \tilde f_{\rm o} \tilde g^\dagger \tilde g^{-1} \delta S_\gamma \tilde g \end{equation} so provided $\tilde g^\dagger$ falls off at least as fast as $\tilde g$ as one approaches the diffraction limit this operator is well defined. In real space the corresponding operator is \begin{equation} \label{eq:fsoperator0} f_{\rm s}' = f_{\rm s} - \gamma_\alpha M_{\alpha i j} (g^\dagger \otimes ( r_i \partial_j f_{\rm o}) - g^\dagger \otimes (r_i \partial_j h) \otimes f_{\rm o}) \end{equation} \hide{ \begin{equation} f_{\rm s}' = f_{\rm s} - \gamma_\alpha (g^\dagger \otimes (M_{\alpha i j} r_i \partial_j f_{\rm o}) - H_\alpha \otimes f_{\rm o}) \end{equation} } so now the commutator term is the convolution of $f_{\rm o}$ with $ H_\alpha = M_{\alpha i j} g^\dagger \otimes (r_i \partial_j h)$. In principle one can design $g^\dagger$ such that $f_{\rm s}$ is essentially identical to $f_{\rm o}$; simply set $\tilde g^\dagger$ to unity for all modes within some tiny distance of the diffraction limit and zero otherwise. However, this may not be a very good idea; the sharp edge of such a filter function in $k$-space will result in ringing in real space, both in the filter $g^\dagger(r)$ and especially in $H_\alpha(r)$. These extended wings have no effect on real signal, but will couple to the incoherent noise in the images, whose spectrum does not fall to zero as one approaches the diffraction limit. \hide{ There are two possible resolutions to this obstacle. First, rather than trying to compute the polarizability of every object individually, one can average the polarization of galaxies binned by by flux, size and eccentricity say, and also compute the polarizability for the mean profile, thus effectively avoiding problems with correlated noise. } To ameliorate these problems one can choose $\tilde g^\dagger$ to have a soft roll-off as one approaches $k_{\rm max}$ to counteract the divergence of $1/\tilde g$. \hide{The optimal choice of filter $\tilde g^\dagger$ must necessarily depend on the type of signal one is dealing with, but we suspect that in most cases it corresponds to a smoothing quite similar to the PSF itself. } One simple option is to set $g^\dagger = g$; i.e.~to re-convolve with the PSF itself. In this case $H_\alpha(r)$ will be about as compact as the PSF, and we then have \begin{equation} \label{eq:fsoperator1} f_{\rm s}' = f_{\rm s} - \gamma_\alpha M_{\alpha i j} (g \otimes (r_i \partial_j f_{\rm o}) - (f_{\rm o} \otimes (r_i \partial_j g)) \end{equation} or equivalently \begin{equation} \label{eq:fsoperator2} f_{\rm s}' = f_{\rm s} + \gamma_\alpha M_{\alpha i j} (2 (r_i \partial_j g) \otimes f_{\rm o} - r_i (\partial_j g \otimes f_{\rm o})) \end{equation} where we have integrated by parts to avoid explicitly differentiating the image $f_{\rm o}$. To summarize, we have shown in (\ref{eq:fobsoperator}) that the effect on the observed sky $f_{\rm o}$ of a weak shear applied before seeing can be written as a shear applied after seeing plus a convolution with some kernel which is the sheared transform of the logarithm of the OTF, which can be computed from the PSF. We have explored the form of this kernel for both Gaussian and more realistic models for the PSF. For atmospheric turbulence limited seeing the kernel is a power law $H \sim r^{-11/3}$. For diffraction limited seeing the shear operator appears formally ill-defined but this is not a serious problem when applied to real data and that one can, for example, compute the operator for a slightly smoothed sky $f_s = g^\dagger \otimes f_{\rm o}$ in a divergence free manner. We have presented the shear operator for the case where $g^\dagger = g$. This could potentially be used to compute the response of the shape statistics measured by the FOCAS and/or Sextractor packages, as these are measured from just such a re-convolved image, but we will not pursue this here. \section{Properties of Optical Point Spread Functions} \label{sec:PSFtheory} Here we shall briefly review and derive some properties of telescope point spread functions which are used above. For more detailed background the reader should consult (\citeNP{roddier81}; \citeNP{beckers93}) and references therein. We will highlight the wavelength or color dependence of the various sources of PSF anisotropy, which may be crucially important for weak lensing searches for large-scale structure and galaxy-galaxy lensing. According to elementary diffraction theory \cite{bw64} the complex electromagnetic field amplitude $a(x)$ due to a distant source at position $x_{\rm phys}$ on the focal plane (we will suppress polarization subscripts for clarity) is given as an integral over the input pupil \begin{equation} \label{eq:fresnelintegral} a(x_{\rm phys}) = \int d^2r \; A(r) C(r) e^{2 \pi i x_{\rm phys} r / L \lambda} \end{equation} where $A(r)$ is the `pupil function' describing the aperture transmission, $C(r)$ is the complex electric field amplitude of the incoming wave, $\lambda$ is the wavelength of the radiation and $L$ is the focal length. The field amplitude $C(r)$ will incorporate any random amplitude and phase variations of the incoming wavefronts due to atmospheric turbulence, whereas constant wavefront distortions due to aberrations in the optical elements of the telescope are incorporated as a complex factor in $A$. Thus $a(x)$ is the Fourier transform of $AC$, evaluated at wave-number $k = 2 \pi i x / D \lambda$. The PSF $g$ is the square of the field amplitude and in rescaled coordinates $x = 2 \pi x_{\rm phys} / L \lambda$, is \begin{equation} \label{eq:PSF1} g(x) = |a(x)|^2 = \int {d^2z \over (2 \pi)^2} e^{-ix\cdot z} \tilde g(z) \end{equation} where the OTF is \begin{equation} \label{eq:OTF1} \tilde g(z) = \int d^2r C(r) C^*(r+z) A(r) A^*(r + z). \end{equation} For very short ground-based observations the atmospheric rippling is frozen and the PSF consists of speckles. For long exposures we are taking the time average of the OTF and can replace $C(r) C^*(r+z)$ by its time average $\langle C(r) C^*(r+z) \rangle = \xi_C(z)$. This is independent of $r$, so the OTF factorizes into two independent functions \begin{equation} \label{eq:OTF2} \tilde g(z) = \xi_C(z) \int d^2 r A(r) A^*(r + z) \end{equation} and the same is true for random small scale amplitude or phase fluctuations introduced by e.g.~random fine scale mirror roughness. \subsection{Atmospheric Turbulence} Ground based observations on large telescopes are usually limited by atmospheric seeing arising from inhomogeneous random turbulence, and it is a good approximation to ignore the finite size of the entrance aperture and set the factor involving $A$ in (\ref{eq:OTF2}) to unity. In the `near field' immediately behind the turbulent layer, the effect on the incoming wave is a pure phase shift $C(r) = e^{i\varphi(r)}$ where $\varphi = 2 \pi i d(r) / \lambda$ and $d(r)$ is the vertical displacement of the wavefront due to turbulence. The displacement $d$ is nearly independent of wavelength, so $\varphi(r) \propto 1/ \lambda$. At greater depths this phase shift evolves into a combination of amplitude and phase variations, but the 2-point function $\langle C(r) C^*(r+z) \rangle$ remains invariant (\citeNP{fried66}; \citeNP{roddier81})and the `natural seeing' OTF is \begin{equation} \tilde g(z) = \langle e^{i(\varphi(r) - \varphi(r + z))} \rangle. \end{equation} For steady turbulence and long integrations the central limit theorem guarantees that the phase error $\psi = \varphi(r) - \varphi(r + z)$ will have a Gaussian probability distribution $p(\psi) = (2 \pi \langle \psi^2 \rangle)^{-1/2} \exp(- \psi^2 / 2 \langle \psi^2 \rangle)$ and so the time average of the complex exponential is \begin{equation} \label{eq:expavg} \langle e^{i \psi} \rangle = \int d\psi \; p(\psi) e^{i \psi} = \exp(-\langle \psi^2 \rangle / 2) = \exp(-S_\varphi(r) / 2) \end{equation} where the `phase structure function' is \begin{equation} S_\varphi(\Delta r) \equiv \langle (\varphi_1 - \varphi_2)^2 \rangle. \end{equation} There are strong theoretical \cite{tatarski61} and empirical reasons to believe that on scales much less than some `outer scale' the turbulence will have the Kolmogorov $n = -11/3$ spectrum, for which $S_d(r) \propto r^{5/3}$. The structure function for the phase is conventionally written as $S_\phi(r) = 6.88 (r/r_0)^{5/3}$ where $r_0$ is the `Fried length' \cite{fried66} being on the order of tens of cm for typical observing conditions (an $r_0$ of 20cm gives a FWHM = $0''.5$ at $\lambda = 550$ nm). The rms phase difference rises with separation as $r^{5/6}$ in the `inertial range' delimited at the upper end by the outer scale, set by the width of the mixing layer, which recent estimates (\citeNP{azb+97}; \citeNP{mtz+98}) find to be around $10-20$m and much larger than $r_0$. The on-axis OTF computed from stars in deep CFHT imaging agrees quite well with the theoretical expectation. The inertial range is limited at the low end by diffusion, but at scales much smaller than $r_0$, so little error is incurred in ignoring this; in real telescopes mirror roughness and other effects modify the OTF at small scales. These effects dominate the PSF at very large radii, but are unimportant for weak lensing observations. The optical transfer function is $\tilde g(k) = \exp(-S_\phi(k D \lambda / 2 \pi) / 2)$ and is real and positive. The PSF is the transform of $\exp(-6.88(z/r_0)^{5.3})$ with width which scales as $R_{\rm FWHM} \propto \lambda^{-1/5}$ which again is found to apply quite well in practice. This very weak dependence on wavelength is a blessing in weak lensing since one uses stars to measure the PSF for the galaxies, yet the stars and the galaxies may have different colors. The atmospheric PSF is expected to be isotropic. At large angles the PSF has profile $g \propto x^{-11/3}$ so the unweighted second moment of the PSF is not well defined. For Kolmogorov turbulence the log of the OTF is just proportional to $k^{5/3}$ and is well defined for all $k$. \subsection{Fast Guiding} According to the Kolmogorov law, the rms wave-front tilt, averaged over scale $r$ varies as $r^{-1/6}$, which suggests that even for telescopes with reasonably large $D/r_0$ there may be useful gain in image quality from fast guiding, and experience with HRCAM on CFHT \cite{mgr+89} would seem to support this, though part of the dramatic improvement is likely due to inadequacy of the existing slow guiding system. A technological advance which may have implications for weak lensing is the advent of on-chip fast guiding \cite{tbs97} with OTCCD chips. With a mosaic camera composed of such devices it should be possible to obtain partial image compensation over a large angular scale, with one or more guide stars for each `isokinetic' patch. The theoretical fast guiding PSF was first explored by \citeN{fried66} who argued that the OTF should take the form of the natural seeing or uncompensated OTF times an `inverse Gaussian' $\exp(+\alpha k^2)$, with scale factor $\alpha$ given in terms of $D$ and $r_0$. This is a physically reasonable picture, since it implies that the natural PSF is the convolution of the corrected PSF with a Gaussian to describe the distribution of tilt, but is only an approximate result. Modified forms of the `Fried approximation' have been explored by \citeN{young74} and \citeN{jenkins98}, and the fast guiding OTF has been simulated by \citeN{christou91}. It can be shown that in the near-field limit the exact fast-guiding OTF is given by \begin{equation} \label{eq:fastguiding1} \tilde g(z) = \int d^2 r\; A(r) A(r + z) \exp(- \langle \psi(r,z)^2 \rangle /2) \end{equation} where \begin{equation} \label{eq:fastguiding2} \langle \psi(r,z)^2 \rangle = S({\bf z}) + z_i [(W_i \otimes S)_{\bf r} - (W_i \otimes S)_{{\bf r} + {\bf z}}] - {1\over 2} z_i z_j \int d^2r' \; W_i({\bf r}') (W_j \otimes S)_{{\bf r}'} \end{equation} and where $W_i \equiv \partial_i A^2$. Examples are shown in figure \ref{fig:fastguiding}. These plots show that the impact of fast guiding on the atmospheric PSF for large telescopes will be rather modest, at least if current, rather low, estimates of the outer scale are correct. Fast guiding may, however, yield dramatic improvements for small ($\sim 1$m diameter) telescopes. How well this would work depends largely on the altitude of the turbulent layer. The isokinetic patch size (over which stars move coherently) is $\sim D / h$ so for $h = 10$km and $D=1$, say, this is on the order of $20''$, the motion needs to be sampled at a rate $\gtorder v / D$ reflecting the relatively high wind speed at high altitude, and it may then be difficult to find bright enough guide stars. There are strong indications (\citeNP{cr85}; \citeNP{tbs97}; \citeNP{mfa+91}) that centroid motions are coherent over much larger angular scales than this, indicating that much of the image degradation arises from low-altitude turbulence, and this greatly improves the outlook as one can determine the local motion by averaging a number of stars, and one can afford to sample at a lower rate. A collection of small telescopes equipped with wide angle OTCCD cameras could be a formidable instrument for weak lensing or other projects requiring high resolution imaging over wide fields. Fast guiding, while offerering important resolution gains, will also present its own challenges since one expects the PSF to become sytematically anisotropic depending on location with respect to the guide stars for the reasons described by \citeN{mfa+91}. Also, fast guiding does not cure PSF anisotropies from telescope aberrations. \begin{figure}[htbp!] \centering\epsfig{file=fastguiding.ps,width={1.0 \linewidth},angle=0} \caption{Optical transfer functions and corresponding PSFs given by equations (\ref{eq:fastguiding1}, \ref{eq:fastguiding2}) for a range of telescope aperture diameters: $1.0$m solid; $1.5$m dash; $2.2$m dot-dash; $3.6$m dotted; $\infty$ dot-dot-dot-dash. A von Karman turbulence spectrum with outer scale of 20m and Fried length $r_0 = 0.2$m were assumed, and the telescope was assumed to be operating at a wavelength $\lambda = 550$nm. } \label{fig:fastguiding} \end{figure} \input dispersion.tex \subsection{Aberrations} Aberrations of the optical elements of the telescope can be a significant contribution to the anisotropy of the PSF. These can be analyzed in the same manner as the wavefront deformation due to the atmosphere, but with a couple of distinctive features: First, for low-order `classical' aberrations where the phase error varies smoothly, and for ground based observing conditions, if an aberration is an important factor then it's contribution to the OTF will be nearly achromatic. This is because if there is a smooth variation of the wavefront error (rather than a Gaussian random field with power at all scales as in atmospheric turbulence) amounting to $N \gg 1 $ wavelengths, then the PSF will be very well approximated by its geometric optics limit with shape defined by the pattern of caustics (though for a narrow band filter, the PSF would actually be found on close examination to be composed of a set of speckle sized patches concentrated along the lines where the classical caustics form \cite{bu80}). The wavefront deformation can be measured directly from out of focus images \cite{rr93} so this contribution to the PSF can be directly predicted. \subsection{Diffraction Limited Seeing} Truly diffraction limited seeing arises when the rms phase error across the aperture (due to the atmosphere and/or aberration of the optical elements of the telescope) is much less than unity. In this case the optical transfer function $\tilde g(k)$ is, to a good approximation, just the auto-correlation of the aperture $A\oplus A$ at lag $\Delta r = k D \lambda / 2 \pi$, and must therefore vanish for spatial frequencies $k > 2 \pi / (f \lambda)$, where $f$ is the ratio of the aperture diameter to the focal length. The log of the OTF becomes ill defined as one approaches the diffraction limit. From (\ref{eq:OTF2}) we see that this cut-off is also present in the case of turbulence dominated seeing, but it occurs at a high frequency where the the optical transfer function has already become exponentially small due to atmospheric effects, and has little impact. In the absence of aberrations, the OTF for diffraction limited seeing is real and non-negative. The OTF is symmetric under rotations of 180 degrees, and so any quadrupole anisotropy of the PSF can be anulled simply by re-convolving one's image with a 90 degree rotated PSF. For diffraction limited seeing the size of the PSF scales as the inverse of the wavelength, a fact which can be incorporated in empirical or theoretical \cite{krist95} modeling of the PSF. In the HST WFPC2, figure errors are not negligible. The imaginary part of the OTF is excited to the degree that re-convolution with a 90-degree PSF still leaves non-negligible PSF anisotropy. The phase errors are not large --- the telescope is nearly diffraction limited --- so this means that the wavelength dependence could be quite complicated. The systematic error arising from differences between faint galaxy and foreground star SED's can be estimated much as we did for atmospheric dispersion. \subsection{Guiding Errors, Pixellisation, and Detector Effects} \label{subsec:pixellisation} So far we have considered the continuous distribution of intensity on the focal plane $f_{\rm o}(r)$. In real detectors we sample the image with a grid of pixels. The response of a pixel in not uniform and has been directly measured for front-illuminated EEV devices by scanning a small spot of light across the CCD (\citeNP{jdo93}, \citeNP{jdo94}). They found very little `leakage' of electrons across pixel boundaries, but according to the \citeN{krist95} this is a substantial effect for the WFPC2 instrument on HST; a photon landing in one pixel has a non-negligible probability of being detected by a neighboring pixel instead. Such effects are expected, and found, to depend on wavelength. The value of a pixel is a sample of the convolution of the sky surface brightness with the `pixel function' $p(r)$. In many real systems the pixel spacing $d$ is not much less than the instrumental resolution and images from single exposures are quite badly under-sampled. For this and other reasons, images are typically constructed from a series of exposures with either systematically (in the case of HST) or randomly (for terrestrial observations) staggered positions on the sky. Each image gives a 2-dimensional grid of samples of $p \otimes f_{\rm o}$, and a piecewise continuous function $f_I$ can be constructed by shifting the grid of delta-functions into an absolute astrometric coordinate system and convolving with some interpolation function $p_{\rm interp}$. One can incorporate the effect of guiding errors on a single exposure as a convolution with the pixel function. If we average a set of $N$ such images the result is \begin{equation} F(r) = {1\over N}\sum_I [(f_{\rm o} \otimes p). c_{\Delta_I}] \otimes p_{\rm interp} \end{equation} where $c_\Delta(r)$ is a 2-dimensional comb function with spatial offset (in units of the pixel spacing) $\Delta$: \begin{equation} c_\Delta(r) = \sum\limits_{i_x, i_y = -\infty}^{\infty} \delta (r - (i + \Delta)) \end{equation} The form of $p_{\rm interp}$ depends on the type of interpolation used. For `nearest pixel' interpolation $p_{\rm interp}$ is just a uniform box of side $d$, but if one linearly interpolates between the pixel samples, for example, then $p_{\rm interp}$ will be a more extended, but again readily computable, function. The transform of $F(r)$ is \begin{equation} \tilde F(k) = \tilde p_{\rm interp} \sum \limits_{m_x, m_y = -\infty}^{\infty} (\tilde f_{\rm o} \tilde p)_{k - 2 \pi m / d} {1\over N} \sum_I \exp(2 \pi i m\cdot \Delta_I) \end{equation} The $m_x=m_y=0$ term in the first sum here is just the ideal image $f_{\rm o}$ convolved with $p$ and with $p_{\rm interp}$, while the higher order terms represent aliasing. The transform $\tilde F(k)$ being $p_{\rm interp}$ times the superposition of a grid of images of $\tilde p \tilde f_{\rm o}$ with spacing $2 \pi / d$. Since $\tilde p$ is a fairly compact function the dominant aliasing comes from the low-order images $m = \pm 1$. Aliasing is most severe for a single exposure since the low-order aliased images contribute with unit weight. If we average $N$ randomly shifted images then the strength of the $m \ne 0$ terms is reduced by a factor $\sim 1 / \sqrt{N}$, aliasing is greatly reduced, and to a good approximation the field $F(r)$ is simply the convolution of $f_{\rm o}$ with $p \otimes p_{\rm interp}$. With systematically staggered images, as is possible with HST and potentially with fast on-chip guiding, one can do even better; with a uniform $M \times M$ grid of offsets covering the unit pixel, the nearest, and therefore most problematic, modes $m_x, m_y = \pm 1$ are then zero and the modes remain small until we get to a multiple of $2 M$ times the Nyquist frequency. This assumes that the transformation from detector to sky coordinates is determined and applied accurately. If we make finite errors $\delta \Delta$ in registration, the resulting image will be the convolution of the ideal PSF with a highly compact cluster of delta-functions, and the optical transfer function $\tilde g$ will be the product of the atmospheric, telescope transfer functions with the Fourier transform of this pattern. In practice, one can typically register images to a small fraction of a pixel (say $\ltorder 0.05$ pixels), and the effect of inaccuracy at this level will have negligible effect on the final PSF. Noise in the images, assumed to be incoherent Poisson noise in the source images, can be analyzed in a similar manner and we find that the two-point function of the noise is just the convolution of $p_{\rm interp}$ with itself, and the two-point function of the noise in re-circularized images can be obtained by convolving the raw noise ACF with $g^\dagger$ twice. \section{Weighted Moment Shear Estimators} \label{sec:estimators} We now specialize to weighted quadrupole moments as defined in (\ref{eq:weightedmomentdefinition}) and compute how these respond to shear. We first compute the response of the moments of an individual object, and we then compute the conditional mean response for a population of objects having given flux, size etc. This will allow us to compute an optimal weight function for combining shear estimates from galaxies of a different fluxes, sizes and shapes. \subsection{Response of Weighted Moments for Individual Objects} \label{subsec:individualresponse} Consider again for illustration the case of a Gaussian ellipsoid PSF $g(r) \propto \exp(-r_i m^{-1}_{ij} r_j/2)$ and moments $q_{lm} = \int d^2r \; f_{\rm o}(r) w(r) r_l r_m$. From (\ref{eq:GaussianPSFoperator}), (\ref{eq:weightedmomentdefinition}) we have $q'_{lm} = q_{lm} + \delta q_{lm}$ with \begin{equation} \delta q_{lm} = - \gamma_\alpha M_{\alpha i j} \int d^2r\; w r_l r_m (r_i \partial_j f_{\rm o} + m_{ip} \partial_p \partial_j f_{\rm o}) \end{equation} Integrating by parts to replace derivatives of $f_{\rm o}$ with derivatives of $wr_l r_m$ we find the linear response of $q_A \equiv {1 \over 2} M_{Alm} q_{lm}$ to a shear can be written \begin{equation} \label{eq:weighteddeltap} \delta q_A = P_{A\beta} \gamma_\beta \end{equation} with `shear polarizability' \begin{equation} \label{eq:polarisability} P_{A\beta} = \int d^2 r \; {\cal P}_{A\beta}(r) f_{\rm o}(r) \end{equation} and where \begin{equation} \label{eq:Pdefinition} {\cal P}_{A\beta}(r) = {1\over 2} M_{Alm} M_{\beta ij} [r_i \partial_j w r_l r_m - m_{ip} \partial_p \partial_j w r_l r_m] \end{equation} For the special case of $w = 1$, i.e.~unweighted moments, we find ${\cal P}_{\alpha\beta} = \delta_{\alpha\beta}(r_i r_i - m_{ii})$ and hence $\delta q_\alpha = \gamma_\alpha (q_{ii} - m_{ii}) = 2 \gamma_\alpha (q_0 - m_0)$ in accord with the result obtained in the Introduction. In this case, $P_{\alpha\beta}$ is a combination of zeroth and second moments of $f_{\rm o}$. For a general PSF and for moments measured from a filtered field $f_s = g^\dagger \otimes f_{\rm o}$ as a weighted moment \begin{equation} \label{eq:polarisationfromfs} q_A = {1\over 2} M_{Alm} \int d^2r \; w(r) r_l r_m f_s(r) \end{equation} we find using (\ref{eq:fobsoperator}) that the response $\delta q_A$ can be cast in the same form, but now with \begin{equation} \label{eq:Pab1} {\cal P}_{A \beta}(r) ={1\over 2} M_{Alm} M_{\beta ij} [r_i ( g^\dagger \oplus (\partial_j \omega r_l r_m)) - (r_i h) \oplus g^\dagger \oplus (\partial_j \omega r_l r_m)] \end{equation} where we have defined the correlation operator $\oplus$ such that $(a \oplus b)_r \equiv \int d^2 r' a(r') b(r' + r)$. If the moments are measured directly from the unfiltered images $f_{\rm o}$ then one can replace $g^\dagger(r)$ with a Dirac $\delta$-function to obtain \begin{equation} \label{eq:Pab2} {\cal P}_{A \beta}(r) ={1\over 2} M_{Alm} M_{\beta ij} [r_i \partial_j w r_l r_m + (hr_i) \oplus (\partial_j w r_l r_m)] \end{equation} whereas for the special case $g^\dagger = g$ \begin{equation} \label{eq:Pab3} {\cal P}_{A \beta}(r) = {1 \over 2} M_{Alm} M_{\beta ij} [2 r_i (g \oplus \partial_j(w r_l r_m)) - g \oplus (r_i \partial_j(w r_l r_m)) ]. \end{equation} The function ${\cal P}_{\alpha\beta}(r)$ is shown in figure \ref{fig:PSFkernel} for a turbulence limited PSF and for a Gaussian window function $w(r)$. Note that (\ref{eq:Pab1}) and (\ref{eq:Pab3}) are well defined continuous functions even in the limit that the weight function becomes arbitrarily small; i.e.~$w(r) \rightarrow \delta(r)$. \begin{figure}[htbp!] \centering\epsfig{file=kernel_turb.ps,width={1.0 \linewidth},angle=0} \caption{The panels on the left show the pair of functions $w_1$, $w_2$ with $w_\alpha(r) = {1 \over 2} M_{\alpha l m} w(r) r_l r_m$ which when multiplied by $f_s$ and integrated give the polarization statistic $q_\alpha$. The four panels on the right show the components of the polarizability kernel ${\cal P}_{\alpha\beta}(r)$ which when multiplied by $f_{\rm o}$ and integrated yields the polarizability $P_{\alpha\beta}$. The PSF was computed from a turbulence limited OTF $\tilde g = \exp(-0.5(kr_*)^{5/3})$, and the smoothing kernel was $w(r) = \exp(-0.5(r/r_*)^{2})$ with scale length $r_*$ equal to one eighth of the box side. } \label{fig:PSFkernel} \end{figure} Equation (\ref{eq:weighteddeltap}), along with the appropriate expression for $P_{A \beta}(r)$ tells us how the polarization statistic for an individual object formed from weighted quadrupole moments responds to a gravitational shear. This is an essential ingredient in calibrating the shear-polarization relation for a population of galaxies. As we shall see in the next section, there are some subtleties involved, but for now we note that if we simply average (\ref{eq:weighteddeltap}) over all galaxies on a patch of sky we have \begin{equation} \langle q'_\alpha \rangle - \langle q_\alpha \rangle = \langle P_{\alpha \beta} \rangle \gamma_\beta \end{equation} so an estimate of the net shear is given by \begin{equation} \label{eq:gammaestimator} \hat \gamma_\alpha = \langle P_{\alpha \beta} \rangle^{-1} (\langle q'_\beta \rangle - \langle q_\beta \rangle) \end{equation} Which is the generalization of (\ref{eq:calibratedsimpleestimator}) to weighted moments. The term $\langle q'_\beta \rangle$ in (\ref{eq:gammaestimator}) is the averaged observed polarization. The term $\langle q_\alpha \rangle$ is the mean polarization generated by anisotropy of the PSF, and we will show how this can be dealt with below. Ideally, in averaging polarizations, one should apply weight proportional to the square of the signal to noise ratio, which one would expect to be a function of the flux, size, eccentricity etc.~of the objects. If the shape is measured with some fairly compact window function $w(r)$, then the total flux, which may be dominated by the profile of the object at considerably larger radius, is probably not ideal and one will likely obtain better performance if one takes the weight function to be a function of $q_0$, $q^2 \equiv q_\alpha q_\alpha$ and a weighted flux \begin{equation} F = \int d^2r \; w(r) f_s(r). \end{equation} The response of $F$ can be computed in much the same way as $q_A$, and we find from (\ref{eq:fsoperator0}) $\delta F = R_\alpha \gamma_\alpha$ with $R_\alpha = \int d^2 r \; {\cal R}_\alpha f_{\rm o}(r)$ with \begin{equation} {\cal R}_\alpha(r) = M_{\alpha i j} [r_i (g^\dagger \oplus \partial_j w) - (r_i h) \oplus g^\dagger \oplus \partial_j w)] \end{equation} or, for the case $g^\dagger = g$, \begin{equation} {\cal R}_\alpha(r) = M_{\alpha i j} [2 r_i (g \oplus \partial_j w) - g \oplus \partial_j w r_i)]. \end{equation} In the foregoing we have implicitly assumed that applying a shear does not affect the location of an object. This is not necessarily the case. If objects are detected as peaks of the surface brightness $f_{\rm o}$ smoothed with some detection filter $w_d$, that is as peaks of $f_d = w_d \otimes f_{\rm o}$, then for an object which in the absence of shear lies at the origin, we have $d_i(0) = \partial_i (w_d \otimes f_{\rm o})_0 = \int d^2 r \; \partial_i w_d(r) f_{\rm o}(r) = 0$ while after applying a shear we have $d_i(0) = \gamma_\alpha M_{\alpha l m} \int d^2 r \; \partial_i w_d [r_l \partial_m f_{\rm o} - (r_l \partial_m h) \otimes f_{\rm o}]$. This will not, in general, vanish, implying that the peak location will have shifted, and consequently the central second moments should be measured about the shifted peak location, whereas in the above formulae we have computed the change in the moments without taking the shift into account. One could incorporate this effect, but at the expense of considerable complication of the results. There is some reason to think that this effect is rather weak. In particular, if the galaxy is symmetric under rotation by 180 degrees, so $f_{\rm o}(r) = f_{\rm o}(-r)$, then the shift in the centroid vanishes. In general this is not the case, and the formulae above should be considered only an approximation. \subsection{Response of the Population} \label{subsec:populationresponse} Equation (\ref{eq:gammaestimator}) above gives a properly calibrated estimate of the gravitational shear. It is, however, less than ideal as the polarization average is taken over all galaxies with equal weight. This is neither desirable nor is it achievable in practice due to selection limits, and what one would rather have is an expression for the average induced polarization for all galaxies in some cell of flux, size and shape space, which we parameterize by $F$, $q_0$, and $q^2 \equiv q_\alpha q_\alpha$. One can then average appropriately weighted combinations of the average shear estimate for each cell. The mean induced polarization for such a cell depends not only on the polarizabilities of the objects contained therein, but also on the gradient of the mean density of objects as a function of the photometric parameters $F$, $q_0$, $q_\alpha$. Consider a slice through this 4-space at constant $F$, $q_0$. A shear will induce a general flow of particles in this space in the direction $\hat q_\alpha = \hat \gamma_\alpha$. The mean polarisation for a cell in $F-q_0-q^2$ is the average around an annulus in $q_\alpha$ space, and depends quite sensitively on the local slope of the distribution of particles in $|q_\alpha$. These factors can have a profound influence on the weighting scheme; for a distribution which is flat near the origin in $q_\alpha$ space, like a Gaussian for example, nearly circular objects have no response and should therefore receive no weight. For a randomly oriented distribution of circular disk galaxies, in contrast, the distribution in $q_\alpha$-space has a cusp at the origin; the response becomes asymptotically infinite, and these objects dominate the optimally weighted combination. These examples are both idealized, but underline the importance of computing the population response in order to obtain optimal signal to noise. Let us first compute the conditional mean polarization for galaxies of a given flux and size: $\langle q_\alpha \rangle_{F,q_0}$. The mapping of the photometric parameters $F$, $q_0$, $q_\alpha$ is \begin{equation} \label{eq:Fq0qmapping} \begin{matrix}{ F' = F + R_\beta \gamma_\beta\cr q'_0 = q_0 + P_{0\beta} \gamma_\beta \cr q'_\alpha = q_\alpha + P_{\alpha \beta} \gamma_\beta }\end{matrix} \end{equation} so we need to consider the distribution of galaxies in $F,\;q_0,\;q_\alpha,\;R_\alpha,\;P_{0\alpha},\;P_{\alpha \beta}$ with lensed and unlensed distribution functions related by \begin{equation} n'(F', q'_0, q'_\alpha, R'_\alpha, P'_{0\alpha}, P'_{\alpha \beta}) dF' dq'_0 d^2 q' d^2 R d^2 P'_0 d^4 P' = n(F, q_0, q_\alpha, R_\alpha , P_{0\alpha}, P_{\alpha \beta}) dF dq_0 d^2 q d^2 R d^2 P_0 d^4 P. \end{equation} Multiplying by $W(F', q'_0) q'_\alpha$, where $W$ is some arbitrary function, and integrating over all variables we have \begin{equation} \int dF' dq'_0 d^2 q' d^2 R d^2 P'_0 d^4 P' n' W(F', q'_0) q'_\alpha = \int dF dq_0 d^2 q d^2 R d^2 P_0 d^4 P n W(F + \delta F, q_0 + \delta q_0) (q_\alpha + \delta q_\alpha). \end{equation} Now to zeroth order in $\gamma$ this vanishes because of statistical anisotropy of the unlensed population, so using (\ref{eq:Fq0qmapping}) for $\delta F$ etc.~and performing a Taylor expansion of $W(F + \delta F, q_0 + \delta q_0)$ and integrating by parts we have \begin{equation} \int dF' \hide{dq'_0 d^2 q' d^2 R' d^2 P'_0} \ldots d^4 P' \; n' W(F', q'_0) q'_\alpha = \gamma_\beta \int dF \hide{dq_0 d^2 q d^2 R d^2 P_0} \ldots d^4 P \; W(q_0) \left(n P_{\alpha \beta} - P_{0\beta} q_\alpha {\partial n \over \partial q_0} - R_\beta q_\alpha {\partial n \over \partial F} \right). \end{equation} To first order in $\gamma$ we can replace unprimed by primed quantities throughout the integral on the RHS since we need to compute this only to zeroth order accuracy. We now have a relation between the mean value of the observed polarization on the LHS and some other observable, again integrated over the distribution of observed galaxy properties (rather than of the unlensed parent distribution). Since $W(q_0)$ is arbitrary, and dropping primes, this implies \begin{equation} \int d^2 q d^2 R d^2 P_0 d^4 P n q_\alpha = \gamma_\beta \int \cdots \int d^2 q d^2 R d^2 P_0 d^4 P \left(n P_{\alpha \beta} - P_{0\beta} q_\alpha {\partial n \over \partial q_0} - R_\beta q_\alpha {\partial n \over \partial F} \right) \end{equation} or equivalently, that the conditional average polarization is \begin{equation} \langle q_\alpha \rangle_{F, q_0} = \gamma_\beta \left[\langle P_{\alpha \beta} \rangle - {1 \over n} {\partial n \langle P_{0\beta} q_\alpha \rangle \over \partial q_0} - {1 \over n} {\partial n \langle R_{\beta} q_\alpha \rangle \over \partial F} \right]_{F, q_0} \end{equation} where $n = n(F, q_0)$. Thus, as expected, the shear induced shift in the mean polarization for galaxies of a given size and flux differs from the mean of the shift $\delta q_\alpha = P_{\alpha \beta} \gamma_\beta$ for an individual object. This is because a shear changes the weighted flux and size of an object in a way which is correlated with its shape. When we average the shear for galaxies in some cell in flux-size space we are averaging over galaxies which have been scattered in size and flux and we obtain a bias in the net polarization which depends on the gradients of the distribution function. We can generalize this to compute the response for galaxies of given flux $F$, size $q_0$, and rotationally invariant shape parameter $q^2 = q_\alpha q_\alpha$. To do this, we set $q_\alpha = q \hat q_\alpha$ with the unit polarization vector $\hat q_\alpha = \{\cos \varphi, \sin \varphi\}$, so $d^2q = q dq d\varphi$. We then have, now for some arbitrary function $W(F, q_0, q^2)$, \begin{equation} \int dF' dq'_0 d{q^2}' d\varphi' d^2 R d^2 P'_0 d^4 P' n' W(F', q'_0, {q^2}') q'_\alpha = \int dF dq_0 dq^2 d\varphi d^2 R d^2 P_0 d^4 P n W(F + \delta F, q_0 + \delta q_0, q^2 + \delta q^2) (q_\alpha + \delta q_\alpha) \end{equation} where now $n = n(F, q_0, q^2, \varphi, R_\alpha, P_{0\alpha}, P_{\alpha \beta})$. Using $\delta F$ etc.~from (\ref{eq:Fq0qmapping}) and $\delta q^2 = 2 q_\eta P_{\eta \beta} \gamma_\beta$ we have \begin{equation} \int d\varphi d^2 R d^2 P_0 d^4 P n q_\alpha = \gamma_\beta \int d\varphi d^2 R d^2 P_0 d^4 P (n P_{\alpha \beta} -R_\beta q_\alpha \partial n / \partial F - P_{0\beta} q_\alpha \partial n / \partial q_0 - 2 P_{\eta \beta} \partial n q_\eta q_\alpha / \partial q^2) \end{equation} or equivalently $\langle q_\alpha \rangle_{F, q_0, q^2} = {\overline P}_{\alpha \beta} \gamma_\beta$ with effective polarizability \begin{equation} \label{eq:Peff1} {\overline P}_{\alpha \beta} = \langle P_{\alpha \beta} \rangle - {2 \over n} {\partial n \langle q_\eta P_{\eta \beta} q_\alpha \rangle \over \partial q^2} - {1 \over n} {\partial n \langle P_{0\beta} q_\alpha \rangle \over \partial q_0} - {1 \over n} {\partial n \langle R_{\beta} q_\alpha \rangle \over \partial F} \end{equation} where now $n = n(F, q_0, q^2)$, and all averages are at fixed $F, \; q_0, \; q^2$. The photometric parameters $q_0, q_\alpha$ appearing here are unnormalized. This is convenient for computing the linear response functions. It is, however, somewhat awkward here since the distribution function $n(F, q_0, q^2)$ is highly skewed since $q_0$ and $q^2$ correlate very strongly with the flux. In computing the effective polarizability it is more convenient to work with rescaled variables $q_0' = q_0 / F$, and ${q^2}' = q^2 / F^2$. The distribution function in rescaled variables is \begin{equation} n'(F, q_0', {q^2}') = F^3 n(F, q_0, q^2). \end{equation} If we also re-scale the polarizabilities $R_\alpha' = R_\alpha / F$, $P_{A\beta}' = P_{A\beta} / F$, ${\overline P}_{A\beta}' = {\overline P}_{A\beta} / F$, re-express (\ref{eq:Peff1}) entirely in terms of primed quantities and then drop the primes we find \begin{equation} \label{eq:Peff2} {\overline P}_{\alpha \beta} = \langle P_{\alpha \beta} \rangle - {2 \over n} {\partial n \langle q_\eta P_{\eta \beta} q_\alpha \rangle \over \partial q^2} - {1 \over n} {\partial n \langle P_{0\beta} q_\alpha \rangle \over \partial q_0} + {1 \over n} \left[1 - {\partial \over \partial \ln F} + {\partial \over \partial \ln q_0} + 2 {\partial \over \partial \ln q^2} \right] n \langle R_\beta q_\alpha \rangle \end{equation} where we have used the result that for any function $X(F, q_0', {q^2}')$ the partial derivative WRT $F$ at constant $q_0$, $q^2$ is $(\partial X(F, q_0', {q^2}') / \partial F)_{q_0,q^2} = \partial X / \partial F - (q_0'/F) \partial X / \partial q_0' - (2 {q^2}'/F) \partial X /\partial {q^2}'$ to compute the term involving $R_\alpha$. The rather cumbersome expression (\ref{eq:Peff2}) calibrates the relation between the shear and the mean polarization for galaxies in a small cell in flux-size-shape space. To compute it we need to bin galaxies in this space to obtain the mean density $n$ and the various averages appearing here, and then perform the indicated partial differentiation. The form (\ref{eq:Peff2}) is somewhat inconvenient as the density $n(F, q_0, q^2)$ is asymptotically constant as $q^2 \rightarrow 0$ and one has to properly deal with the discontinuous derivative at this boundary. A computationally more convenient approach is to make one final transformation from $q^2 \rightarrow q$; since $n(F, q_0, q) = 2 q n(F, q_0, q^2)$ falls to zero as $q \rightarrow 0$, and there is then no need for any special treatment of the derivatives at the boundary. With this transformation we have \begin{equation} \label{eq:Peff3} {\overline P}_{\alpha \beta} = \langle P_{\alpha \beta} \rangle - {1 \over n} {\partial n \langle q_\eta P_{\eta \beta} \hat q_\alpha \rangle \over \partial q} - {1 \over n} {\partial n \langle P_{0\beta} q_\alpha \rangle \over \partial q_0} + {q \over n} \left[1 - {\partial \over \partial \ln F} + {\partial \over \partial \ln q_0} + {\partial \over \partial \ln q} \right] n \langle R_\beta \hat q_\alpha \rangle \end{equation} where now $n = n(F, q_0, q)$. What about measurement noise? Let us assume that one has been given an image containing signal and measurement noise, and that one has detected objects, and measured quantities like $F$, $q_A$. How would these photometric parameters change under the influence of a gravitational shear? The answer is given by (\ref{eq:Fq0qmapping}), but with the understanding that $R_\beta$, $P_{0\beta}$ etc.~be the response functions one would measure in the absence of noise. This means that (\ref{eq:Peff3}) is also applicable with the same proviso. The major terms in (\ref{eq:Peff3}) are however invariant of additive noise. The exceptions are the terms invoving $\langle P_{0\beta} q_\alpha \rangle$ and $\langle R_\beta q_\alpha \rangle$ which are quadratic in the sky surface brightness. The measured expectation values $\langle P_{0\beta} q_\alpha \rangle$ etc.~therefore exceed the true values, but by an amount one can calculate from the known properties of the measurement noise. Another implicit assumption in the above analysis is that the objects are actually detected, which restricts applicability to objects which are detected at a reasonable level of significance. Aside from this, the results above should be applicable in the presence of measurement noise. To test these claims we have made extensive simulations with mock data, the details of which are described in appendix \ref{sec:simulation}. Figure \ref{fig:Pplot} shows the results of one of these. The actual polarization agrees quite closely with the effective polarizability, even for very faint objects. While the differences between the effective polarizability and that for individual objects is not very large - typically on the order of 20\% or so - the effective polarizability clearly describes the true response more faithfully. We shall now use (\ref{eq:Peff3}) to construct a minimum variance weighting scheme for combining shear estimates. \begin{figure}[htbp!] \centering\epsfig{file=Pplot.ps,width={1.0 \linewidth},angle=0} \caption{The top panel shows the density of galaxies detected in the simulations described in appendix \ref{sec:simulation} as a function of $F$, $q_0$, $|q|$. Each of the sub-plots shows the density at fixed $F$ with abscissa $|q|$ and ordinate $q_0$. These sub plots are arranged with flux increasing from left to right with flux $F$ increasing by a factor 1.43 at each step. All particles detected with significance level greater than 4-sigma are shown. The panel below this shows the density of particles weighted by the individual object polarizability, and the panel below that shows the density weighted by the effective polarizability given by (\ref{eq:Peff3}). The lower plot shows the density of objects weighted by the actual polarization (divided by the shear applied in the simulation, or $\gamma = 0.1$ in this case). } \label{fig:Pplot} \end{figure} \subsection{Optimal Weighting with Flux, Size and Shape} \label{sec:optimisation} Armed with the conditional mean polarization $\langle q_\alpha \rangle_{F, q_0, q^2}$ we can now compute an optimal weight (as a function of, for example, flux $F$, size $q_0$, and eccentricity $q^2)$ for combining the estimates of the shear from galaxies of different types. Let us assume that one has measured fluxes etc.~for a very large number of galaxies --- from an entire survey say --- and that from these data one has determined the mean number density of galaxies $n(F, q_0, q^2)$ and also the various conditional averages and gradients that appear in (\ref{eq:Peff3}). Now consider a relatively small spatial subsample of these galaxies and bin these into cells in $F, q_0, q^2$ space, and for each bin compute the occupation number $N$ and the summed polarization $\sum q_\alpha$. A shear estimator for a cell which happens to have occupation number $N$ is \begin{equation} \hat \gamma_\beta = {\overline P}^{-1}_{\alpha \beta} \sum q_\alpha / N. \end{equation} To simplify matters, let us neglect for now any anisotropy of the point spread function, in which case we can write ${\overline P}_{\alpha \beta} = {\overline P} \delta_{\alpha \beta}$ where ${\overline P} \equiv {\overline P}_{\eta\eta} / 2$ and we then have \begin{equation} \hat \gamma_\beta = {1\over N {\overline P}} \sum q_\beta . \end{equation} The expectation value for the variance in $\hat \gamma$ (for a cell which happens to contain $N$ galaxies) is \begin{equation} \langle \hat \gamma^2 \rangle = \langle \sum q^2 \rangle / (N^2 {\overline P}^2) = q^2 / N{\overline P}^2 \end{equation} where we have used $\langle (\sum q_\alpha) (\sum q_\alpha) \rangle = \langle \sum q^2 \rangle$ since, in the limit of weak shear, the galaxy polarizations are uncorrelated. As different cells give shear estimates whose fluctuations are mutually uncorrelated, the optimal way to combine the shear estimates from all the cells is to average them with weight per cell $W_{\rm cell} \propto 1 / \langle \hat \gamma^2 \rangle = N{\overline P}^2 / q^2$ to obtain a final optimized total shear estimate \begin{equation} \hat \gamma_\alpha^{\rm total} = {\sum\limits_{\rm cells} (N {\overline P}^2 / q^2) ({1\over N {\overline P}} \sum q_\alpha ) \over \sum\limits_{\rm cells} N{\overline P}^2 / q^2} = {\sum\limits_{\rm galaxies} {\overline P} q_\alpha / q^2 \over \sum\limits_{\rm galaxies} {\overline P}^2 / q^2} = {\sum\limits_{\rm galaxies} Q \hat q_\alpha \over \sum\limits_{\rm galaxies} Q^2} \end{equation} where $Q \equiv {\overline P} / q$. Thus the optimized cell weighting scheme corresponds to averaging the shear estimates from individual galaxies $\hat \gamma_\alpha^{\rm galaxy} = q_\alpha / {\overline P}$ (for galaxies of given $F, q_0, q^2$) with weights $W_{\rm galaxy} = {\overline P}^2 / q^2$. \hide{Note that while the estimator (....) for the shear for an individual cell appears to be ill defined for empty cells, the final result is well defined since the empty cells receive zero weight.} The variance in the total shear estimator is \begin{equation} \label{eq:totalvariance} \langle \gamma^2 \rangle = {\sum Q^2 \langle \hat q^2 \rangle \over (\sum Q^2)^2} = (\sum Q^2)^{-1}. \end{equation} The quantity $\sum Q^2$ is extensive with the number of galaxies and its value, per unit solid angle of sky, provides a useful figure of merit for weak lensing data. From a 2.75hr I-band integrations of solid angle $d\Omega = 0.165$ square degrees at CFHT taken in good seeing ($0''.60$ FWHM) we obtained $\sum Q^2 \simeq 4.7\times 10^4$ \cite{kwl+98} or $\sum Q^2 / d \Omega \simeq 2.85 \times 10^5 / {\rm sq\; degree}$, so with data of this quality, the statistical uncertainty in the net shear (per component) measured over one square degree would be around $\sigma_\gamma \simeq (2 \times 2.85 \times 10^5)^{-1/2} = 1.32 \times 10^{-3}$. This figure of merit allows one to tune the parameters of one's shape measurement scheme, such as the weight function scale size, in an unbiased and objective manner. The weighting scheme derived above is appropriate if the shear is independent of the measured flux etc. This is the case for lensing by low redshift clusters where for all relevant values of source redshift the sources are effectively `at infinity' and the shear has saturated at its value for an infinitely distant source $\gamma^{\infty}_\alpha$. For high redshift lenses the shear will vary with source redshift, and if one has some, perhaps probabilistic, distance information at one's disposal, then the weighting scheme should be modified. Let us assume that the measured photometric properties $p_i$ of some object indicate it has a probability distribution to be at distance $z$ of $p(z|p_i)$; with high resolution spectroscopy this would be a delta-function at the measured redshift whereas with only broadband colors the conditional probability would be smeared out and perhaps multimodal. The conditional mean shear for this object, for a given a foreground lens, is proportional to the mean inverse critical critical surface density, and the optimal estimate for the shear at infinity is \begin{equation} \hat \gamma_\beta^\infty = {1 \over \Sigma_{\rm crit}(\infty)} {\sum\limits_{\rm galaxies} \langle 1 / \Sigma_{\rm crit}(z) \rangle Q q_\alpha \over \sum\limits_{\rm galaxies} \langle 1 / \Sigma_{\rm crit}(z) \rangle^2 Q^2 } \end{equation} \section{Simulated Data} \label{sec:simulation} To test the procedures described here we have generated simulated mock data and then analysed these. The simulations were made to match as closely as possible observations of $\sim 3$hr integration on the CFHT with $0''.6$ seeing. We first generated a set of 200 mock catalogue of galaxies each corresponding to a patch of sky of size $2'.56$ on a side. Galaxies were drawn from a Schechter style luminosity function laid down in a Poissonian manner in an Einstein de Sitter cosmology. Images with pixel scale $0''.075$ were then generated by realising the galaxies as exponential disks with random orientations and scale lengths corresponding to fixed rest frame central surface brightness. The galaxies were modelled as optically thick, since the optically thin model looks unrealistic as it has too many very bright edge-on systems as compared to real images. A number of point-like stars were added to the images, which were then sheared with $\gamma = 0.1$, convolved with a Kolmogorov turbulence PSF with $0''.6$ FWHM, and then rebinned to $0''.15$ pixel scale. When the real data are analysed they are interpolated from the original $0''.2$ pixel scale to the final $0''.15$ scale with bi-linear interpolation. This results in a further convolution of the signal and the noise in the real images, but with slightly different smoothing kernels. These kernels were computed by modelling the image shifts as a uniform distribution within the final pixel size; the noise-free mock images were convolved with the appropriate kernel and then Gaussian white-noise images were generated to model the sky noise, and were convolved with the appropriate kernel and then added to the images. A sample image is shown in figure \ref{fig:cfhvssim_images}, and the corresponding size-magnitude diagram is shown in figure \ref{fig:cfhvssim_cats}, from which it is apparent that the simulated objects have properties very similar to those detected in the real data. \begin{figure}[htbp!] \centering\epsfig{file=cfhvssim_images.ps,width={1.0 \linewidth},angle=0} \caption{Simulated and real image sections $2'.56$ on a side.} \label{fig:cfhvssim_images} \end{figure} \begin{figure}[htbp!] \centering\epsfig{file=cfhvssim_cats.ps,width={1.0 \linewidth},angle=0} \caption{Catalogs from image sections shown in figure \ref{fig:cfhvssim_images}. } \label{fig:cfhvssim_cats} \end{figure} These data were analysed exactly like the real data. That is, the objects were detected as peaks of a smoothed image. The stars were extracted and their shapes fit in the manner described to obtain the PSF. A smoothed image $f_s = g \otimes f_{\rm o}$ was generated and from this the polarisation $q_\alpha$ was computed using (\ref{eq:polarisationfromfs}), and the polarisability for each object was computed using (\ref{eq:Pab3}).
\section{Introduction} The existence of a galactic population of $\gamma$-ray sources is known since the days of the COS B experiment (Swanenburg et al. 1981). Montmerle (1979) showed that about 50 \% of the unidentified COS B detections lie in regions containing young objects, like supernova remnants (SNRs) and OB massive stars. He suggested that the $\gamma$-ray emission could stem from $\pi^0$-decays resulting from hadronic interactions of high-energy protons (or nuclei) and ambient matter. These protons would be locally injected by young stars in the SNR shocks where they would be diffusively accelerated up to high energies by Fermi mechanism. Cass\'e \& Paul (1980) argued that particle acceleration at the terminal shock of strong stellar winds alone could be responsible for the $\gamma$-ray sources without the mediation of the SNR shock waves advocated by Montmerle. Gamma-ray production in shocks generated by massive stars has been discussed since then, and from different points of view, by V\"olk \& Forman (1982), White (1985), Chen \& White (1991a,b), and White \& Chen (1992), among others. Since 1991, with the advent of the Energetic Gamma Ray Experiment Telescope (EGRET) onboard the Compton satellite, the observational data on galactic $\gamma$-ray sources have been dramatically improved. Two of the previously unidentified COS-B sources, Geminga and PSR 1706-44, are now known to be pulsars. The detection of pulsed high-energy emission from other sources (there are seven $\gamma$-ray pulsars so far, see Thompson 1996 for a review) and the identification of Geminga as a radio quiet object have prompted several authors to explore the possibility that all unidentified low latitude sources in the Second EGRET (2EG) catalog (Thompson et al. 1995, 1996) are pulsars (with the exception of a small extragalactic component which is seen through the Galaxy). In particular, Kaaret \& Cottam (1996) have used OB associations as pulsar tracers finding out a significant positional correlation with 2EG unidentified sources. A similar study, including SNRs and HII regions (these latter considered as tracers of star forming regions and, consequently, of possible pulsar concentrations), has been carried out by Yadigaroglu \& Romani (1997), who concluded that the pulsar hypothesis for the 2EG sources is consistent with the available information. However, recent spectral analyses made by Merck et al. (1996) and Zhang \& Cheng (1998) clearly show that several 2EG sources are quite at odds with the pulsar explanation. Time variability in the $\gamma$-ray flux of many sources also argues against a unique population behind the unidentified galactic $\gamma$-ray detections (McLaughlin et al. 1996, Mukherjee et al. 1997). Sturner \& Dermer (1995) and Sturner et al. (1996) have investigated the possible association of $\gamma$-sources with SNRs, finding significant statistical support for the idea that some remnants could be $\gamma$-ray emitters. Esposito et al. (1996) have shown that five 2EG sources are coincident with well known SNRs and, more recently, Combi et al. (1998a) have detected a new shell-type SNR at the position of 2EGS J1703-6302, as well as an interacting compact HI cloud, through multiple radio observations, clearly demonstrating, in this way, that at least some EGRET detections are physically related to SNRs. With the publication of the Third EGRET (3EG) catalog of high-energy gamma-ray sources (Hartman et al. 1999), which includes data from Cycles 1 to 4 of the space mission, new and valuable elements become available to deepen the quest for the nature of the unidentified $\gamma$-ray sources. The new catalog lists 271 point sources, including 170 detections with no conclusive counterparts at other wavelengths. Of the unidentified sources, 74 are located at $|b|<10^o$ (this number can be extended to 81 if we include sources with their 95 \% confidence contours reaching latitudes $|b|<10^o$). This means that the number of possible galactic unidentified sources is now nearly doubled respect to the 2EG catalog. Can these new sources be associated with pulsars? How many sources could be ascribed to known SNRs? Is there new statistical evidence for the identification of some detections in the 3EG catalog with massive stars that generate very strong winds? In the present paper we investigate these questions in the light of the new $\gamma$-ray data of the 3EG catalog. We use numerical simulations (constrained by adequate boundary conditions) of $\gamma$-ray source populations to weight the statistical significance of the different levels of positional coincidences determined for diverse types of candidates such as individual massive stars (Wolf-Rayet and Of stars with strong stellar winds), SNRs, and OB associations. The contents of the paper are as follows. In the next section we describe the numerical procedure implemented for the analyses. Sections 3, 4, and 5 deal with the possible association of unidentified 3EG sources with stars, SNRs, and star-forming regions considered as pulsar tracers, respectively. In Section 6 we present some further comments and, finally, in Section 7, we draw our conclusions. \section{Numerical simulations and statistical results} With the aim of finding the positional coincidences between 3EG unidentified sources at $|b|<10^o$ and different populations of galactic objects, we have developed a computer code that determines the angular distance between two points in the sky, taking into account the positional uncertainties in each of them. The code can be used to obtain a list of $\gamma$-ray sources with error boxes (here assumed as the 95 \% confidence contours given by the 3EG catalog) overlapping different kinds of objects, both extended (like SNRs or OB associations) and punctual (like stars). We run the code with the 81 unidentified EGRET sources at galactic latitudes $|b|<10^o$ and complete lists of Wolf-Rayet (WR) stars, Of stars, SNRs, and OB associations. These lists were obtained from van der Hucht et al. (1988), Cruz-Gonz\'alez et al. (1974), Green (1998), and Mel'nick \& Efremov (1995), respectively. We have found that 6 $\gamma$-ray sources of the 3EG catalog are positionally coincident with WR stars, 4 with Of stars, 22 with SNRs, and 26 with OB associations. In order to estimate the statistical significance of these coincidences, we have simulated a large number of sets of EGRET detections, retaining for each simulated position the original uncertainty in its galactic coordinates. Specifically, in each case we have generated by computer 1500 populations of 81 $\gamma$-ray sources through rotations on the celestial sphere, displacing a source with original coordinates $(l,b)$ to a new position $(l^\prime,b^\prime)$. The new pair of coordinates is obtained from the previous one by setting $l^\prime=l+ R_1 \times 360^o$. Here, $R_1$ is a random number between 0 and 1, which never repeats neither from source to source nor from set to set. Since we are simulating a galactic source population and not arbitrary sets at $|b|<10^o$, we impose that the new distribution (i.e. each of the simulated sets) retains the form of the actual histogram in latitude of the unidentified 3EG sources, with 1$^o$ or 2$^o$-binning. The histogram, for 1$^o$-binning, is shown in Figure ~\ref{fig.1}. In order to accomplish the mentioned constraint, we make $b^\prime=b + R_2 \times 1^o$, and then, if the integer part of $b^\prime$ is greater than the integer part of $b$ or if the sign of $b^\prime$ and $b$ are different, we replace $b^\prime$ by $b^\prime - 1^o$. Here, again, $R_2$ is a random number between 0 and 1. This ensures that the new set of artificial positions preserves the actual histogram in latitude at 1$^o$-binning. Similarly, a 2$^o$-binning distribution can be maintained. Both sets of simulations provide comparable results. The unidentified 3EG sources have, additionally, a non-uniform distribution in galactic longitude, showing a concentration towards the galactic center. However, when doing the simulations, we imposed no constraints in longitude because we wanted to consider any kind of possible galactic populations. Once we performed 1500 simulations for each type of counterparts (a larger number of simulations do not significantly modify the results), we estimated the level of positional coincidences between each simulated set and the different galactic populations under consideration. From these results we obtained an average expected value of chance associations and a corresponding standard deviation. The probability that the observed association level had happened by chance was then evaluated assuming a Gaussian distribution of the outputs. The results of this study are shown in Table 1, where we list, from left to right, the type of object under study, the number of actual positional coincidences, the number of expected chance coincidences according to 1$^o$-binning simulations, the probabilities that the actual coincidences can be due to chance, and the similar results for simulations with 2$^o$-binning. From Table 1, it can be seen that there is a strong statistical correlation between unidentified $\gamma$-ray sources of the 3EG catalog and SNRs (at $\sim6\sigma$ level) as well as with OB associations (at $\sim4\sigma$ level). Regarding the stars, we find that there is a marginally significant correlation with WR and Of stars ($\sim3\sigma$). Remarkably, the probability of a pure chance association for SNRs is as low as 5.4$\times10^{-10}$ according to the 2$^o$-binning simulations ($1.6\times10^{-8}$ for 1$^o$-binning). For the stars, we obtain probabilities in the range $10^{-2}-10^{-3}$, which are suggestive but not overwhelming. In the next sections we explore these results in more detail. \section{Massive stars} The case for possible association of unidentified EGRET sources with WR stars was previously presented --using data from the 2EG catalog-- by Raul \& Mitra (1997). In the former catalog, there are 37 unidentified sources at $|b|<10^o$. Raul and Mitra proposed, on the basis of positional correlation, that 8 of these sources could be produced by WR stars. Their analysis of the possible chance occurrence of these associations, which was purely analytic and assumed equiprobability for each position on the sky, yielded an a priori expectation of $\sim 10^{-4}$. Their results are notably modified when the 3EG catalog is considered. Changes in position and smaller positional uncertainties reduce the number of positional coincidences despite the remarkable increment in the number of sources. Additionally, a more rigorous treatment in the probability analysis has the effect of significantly enhance the possibility of chance association (see Table 1). In Tables 2 and 3 we list the 3EG sources positionally coincident with WR and Of stars, respectively. As far as we are aware this is the first time that a statistical study of the correlation between Of stars and EGRET detections is carried out, despite that the possibility of $\gamma$-ray production in this kind of objects has been extensively discussed in the literature (e.g. V\"olk \& Forman 1982). In the tables we provide, from left to right, the 3EG source name, the measured (summed over Cycles 1 to 4) $\gamma$-ray flux, the photon spectral index $\Gamma$ ($N(E)\propto E^{-\Gamma}$), the star name, the angular distance from the star to the $\gamma$-ray source best position, the distance to the star, the terminal wind velocity, the mass loss rate, the expected intrinsic $\gamma$-ray luminosity assuming the star's distance (the minimum one when there are more than one star in the field) and isotropic emission with average index $\Gamma=2$, and, in the last column, any other positional coincidence revealed in our study. From Table 2, it can be seen that most of the possible associations claimed by Raul \& Mitra (1997) are no longer viable ones. Just WR stars 37-39, 138, and 142 of their list stay after our analysis. In order to compare Raul and Mitra's results with our's, it is worth remembering that when testing against positional coincidences they assumed an angular uncertainty of 1$^o$ for all EGRET sources. If we would make such an assumption, we would have found 20 positional coincidences in the 3EG catalog (i.e. 24.7 \% of the unidentified low-latitude sources). However, due to the new reduced EGRET errors, just 7 \% of these sources are now positionally consistent with WR stars, with a priori probability of $\sim10^{-3}$ of being by chance. In addition, there are 4 sources with Of stars within their error boxes. The probability that these latter associations result just by chance is $\sim10^{-2}$. Several mechanisms have been proposed to generate $\gamma$-rays in the vicinity of massive stars with strong winds. A compact $\gamma$-ray source could be the result of $\pi^0$-decays which occur as a consequence of hadronic interactions between relativistic protons or nuclei, locally accelerated by shocks arising from line-driven instabilities in the star wind, and thermal ions (White \& Chen 1992). The same embedded shocks can also accelerate electrons that could provide an additional source of (inverse Compton) $\gamma$-ray emission through the upscattering of stellar UV photons (Chen \& Withe 1991a). Synchrotron losses of these energetic electrons can produce observable nonthermal radio emission, as detected in several massive stars (e.g. Abbott et al. 1984). A different region where the $\gamma$-rays might be generated is at the interface between the supersonic wind flow and the interstellar medium. There, the terminal shock can reaccelerate ions up to high energies and, if sufficient concentration of ambient matter is available (e.g. small clouds or swept-up material), nuclear $\gamma$-rays copious enough to be detected could be produced (Cass\'e \& Paul 1980). V\"olk \& Forman (1982) have argued that stellar energetic particles lose too much energy in the expanding wind to be efficiently accelerated at the terminal shock, in such a way that local injection (e.g. from a nearby star) is required. However, White (1985) showed that the shocks embedded in the highly unstable radiatively driven winds can be responsible for much higher initial energies and partial reacceleration of the particles during the adiabatic expansion, so isolated massive stars could be also efficient $\gamma$-ray emitters if they present sufficiently strong winds. The a posteriori analysis of our association results show that three stars are of especial interest as possible counterparts of EGRET sources: WR 140, WR 142, and Cyg OB2 No.5. The first one is a binary system composed of a WC 7 plus an O4-5 star. The region of stellar wind collision seems to be particularly suitable for producing high energy emission. Eichler \& Usov (1993) have studied the particle acceleration in this system concluding that it should be a strong $\gamma$-ray source. Based on observational data on WR 140, they predicted a $\gamma$-ray luminosity in the range $5\times10^{32}-2.5\times10^{35}$ erg s$^{-1}$, in well agreement with the measured EGRET flux from 3EG J2022+4317 and the distance to the system (see Table 2). The second promising star, WR 142, is one of the five WR stars which present strong OVI lines without being associated with planetary nebulae. The large Doppler broadening of all spectral lines reveals the existence of a very high wind velocity of $\sim5200$ km s$^{-1}$, which doubles what is usually observed in WR stars (Polcaro et al. 1991). The identification of WR 142 with the COS-B source 2CG 075+00 was proposed in Polcaro et al.'s (1991) paper, where they considered the $\gamma$-ray production in the strong stellar wind. In the 3EG catalog the star position is consistent with the source 3EG J2021+3716. If the star is responsible for the observed $\gamma$-ray flux, its intrinsic luminosity would be $\sim3\times10^{34}$ erg s$^{-1}$, which is of the order of what is expected from White \& Chen's (1992) hadronic model for isolated stars. Finally, the binary system Cyg OB2 No.5 seems to be another interesting candidate for producing $\gamma$-rays. Usually considered as a contact binary formed by two O7 I stars (e.g. Torres-Dodgen et al. 1991), recent observations suggest that the secondary star in this system would be of spectral type B0 V--B2 V (Contreras et al. 1997). Variable radio emission was detected by several authors (e.g. Persi et al. 1990), with timescales of $\sim7$ years. A weak radio component of nonthermal nature has been observed with the VLA at a separation of $\sim0.8''$ from the main radio source, which is thermal and coincident with the primary optical component (Contreras et al. 1997). The radio variability in Cyg OB2 No.5 has been interpreted in terms of a colliding wind model by Contreras et al. (1997), who suggested that the weaker radio component is not a star but a bow shock produced by the wind collision. In this shock, electrons can be locally accelerated up to relativistic energies, yielding the synchrotron radiation that constitutes the secondary nonthermal source. Additionally, $\gamma$-rays are generated through inverse Compton losses in the UV radiation field of the secondary (Eichler \& Usov 1993). The same Fermi mechanism that accelerates the electrons should also operate on protons, providing a source of energetic ions that could contribute with higher energy $\gamma$-ray emission, as in the case involving WR stars. For strong shocks, the test particle theory predicts that the relativistic protons will have a differential energy spectrum given by $N(E)\propto E^2$. The $\pi^0$-decay $\gamma$-rays resulting from $p-p$ collisions should conserve the shape of the original proton spectrum, in such a way that at energies above 100 MeV the photon spectral index would be $\Gamma\sim2$, as observed by EGRET. \section{Supernova remnants} Possible correlation between SNRs and unidentified EGRET sources, on the basis of two dimensional positional coincidence, has been proposed since the release of the first EGRET (1EG) catalog. Sturner \& Dermer (1995) suggested that some of the unidentified sources lying at galactic latitudes $|b|<10^o$ might be associated with SNRs: of 37 detections, 13 overlapped SNR positions in the 1EG catalog. However, their own analysis showed that the statistical significance was not too high as to provide a strong confidence. Chance association was just 1.8$\sigma$ away from the obtained result. Using the 2EG catalog, Sturner et al. (1996) repeated the analysis, and showed that 95\% confidence contours of 7 unidentified EGRET sources overlapped SNRs, some of them appearing to be in interaction with molecular clouds. Similar results were independently reported by Esposito et al. (1996), although neither of them assessed the chance probablility of these 2EG-catalog findings. Considering the 1EG catalog, 35\% of the unidentified sources were positional related to SNRs. This drop to 21.8\% in the 2EG catalog, and is currently about 27\%. One important point to take into account when evaluating these differences is not only to consider the evolution of the EGRET catalog but also that of the supernova remnant Green's catalog. At the time of the first studies by Sturner \& Dermer (1995), the supernova catalog contained 182 SNRs. This grew up to 194 in 1996, and currently it lists 220 remnants. In Table 4 we show the 3EG sources that are positionally consistent with SNRs listed in the latest version of Green's catalog. From left to right we provide the $\gamma$-ray source name, the measured flux, the photon spectral index $\Gamma$, the SNR identification, the angular distance between the best $\gamma$-ray source position and the center of the remnant, the size of the remnant in arcminutes, the SNR type (S for shell, F for filled-centre, and C for composite), and other positional coincidences found in our study. The table contains 22 possible associations with an a priori probability of being purely by chance completely negligible ($\leq 10^{-8}$). It is important to remind that this list is formed entirely of positional coincidences with currently catalogued SNRs. However, the diffuse galactic disk nonthermal emission, originated in the interaction of the leptonic component of cosmic rays with the galactic magnetic field, is veiling many remnants of low surface brigthness. Recent observational studies using filtering techniques in the analysis of radio data have revealed many new SNR candidates that are not included in Green's catalog (e.g. Duncan et al. 1995, Combi \& Romero 1998, Combi et al. 1998b, 1999). If these candidates were included in our analysis a larger number of associations would have resulted. The intrinsic $\gamma$-ray luminosity of SNRs, stemmed from interactions between cosmic rays reaccelerated at the supernova shock front and swept-up material, is expected to be rather low (Drury et al. 1994). However, if a cloud is near the particle acceleration site, the enhanced nuclear cosmic rays from the shock can ``illuminate'' the cloud through $\pi^0$-decays yielding a compact $\gamma$-ray source (Aharonian et al. 1994). Such scenario has been recently study by Combi et al (1998a) in relation with the source 3EG J1659-6251 (previously 2EGS J1703-6302). \section{OB associations} In Table 5 we list the unidentified 3EG sources that are positionally coincident with the OB associations in the catalog by Mel'nik \& Efremov (1995). Our results can be compared with the similar work by Kaaret \& Cottam (1996). Using the 2EG catalog, they have already found a statistically significant correlation: 9 of the unidentified 2EG sources have position contours overlapping an OB association and other 7 lie within $1^o$ angular distance. These results are totally compatible with our's. Here, we find 26 superpositions out of 81 unidentified 3EG sources (32\%), 5$\sigma$ away from what is expected from pure chance association. The mean angular separation between the centroid of the OB association and the EGRET source is 1.5$^o$, although most sources are at angular distances of less than $1^o$. The differences between both methods of analysis are worth commenting. In particular, we decided, for completitude, to keep the nearby association Sco 2A despite its proximity. Any pulsar traced by it must have a negligible proper velocity in order to be consistent with its angular size, but its existence cannot be ruled out only on a priori grounds. To calculate the chance superposition probability Kaaret and Cottam studied EGRET sources just within [-5$^o$, 5$^o$] in galactic latitude (only 25 sources of the total 129 unidentified ones present in the 2EG catalog), and generated sample locations using two Gaussian distributions, in longitude and latitude, with central value and deviation provided by the actual positions of the unidentified sources. They also used a galactic model to map the gas distribution. This procedure yields almost the same results than the method we follow (chance association probability around $10^{-5}$). Interestingly, despite all EGRET sources changed their positions from the 2EG to the 3EG catalog and a significant number of new detections has been added, the percentage and the confidence level of the positional coincidences remains almost the same in both studies. All known $\gamma$-ray pulsars are young objects ($\leq 10^{6}$ yr) with spectral indices smaller than 2.15 (Crab's) and a trend for spectral hardening with characteristic age (Fierro et al. 1993). From Table 5, if we consider just sources coincident {\em only} with OB associations and exclude the three sources with very steep indices (3EG J 1308-6112, 3EG J 1718-3313, and 3EG J 1823-1314), we get $<\Gamma>=2.07$ and $1\sigma=0.12$ for the 8 remaining EGRET sources. These are the most promising candidates for pulsar associations. We have marked them with a star symbol in Table 5. \section{Further comments} In Figure ~\ref{fig.2} we show a plot of the $\gamma$-ray luminosity (assuming isotropic emission) of the unidentified sources coincident with OB associations against the estimated distance to the associations. By using different symbols we indicate whether there are additional positionally coincident objects for each $\gamma$-ray source. The solid horizontal line represents the luminosity of Vela pulsar. A similar plot of luminosity versus photon spectral index $\Gamma$ is shown in Figure ~\ref{fig.3}. The first plot shows that the luminosity distribution of this subset of 3EG sources is consistent with the observed distribution for $\gamma$-ray pulsars when emission into $4\pi$ sr is assumed (see Kareet \& Cottam 1996). Figure ~\ref{fig.3} shows, however, that not all sources superimposed to OB associations present the spectral signature expected from pulsars: they should concentrate in the left-upper corner of the frame. There, two sources clearly differentiate from the rest: 3EG J1027-5817 and 3EG J1048-5840. They have luminosities similar to Vela's and hard spectra with $\Gamma<2$, which make them good candidates for $\gamma$-ray pulsars. The identification of 3EG J1048-5840 (formerly 2EG J1049-5847) with a pulsar (PSR B1046-58) was already proposed by Zhang \& Cheng (1998), who showed that its $\gamma$-ray spectrum is consistent with the predictions of outer gap models. In addition, these authors also suggested that 3EG J1823-1314 (2EG J1825-1307) could be the pulsar PSR B1823-13. This latter identification must be now rejected in the light of the new determination of the spectral index of the $\gamma$-ray source in the 3EG catalog, $\Gamma=2.69\pm0.19$, which is too steep for a pulsar. Regarding 3EG J1027-5817, no known radio pulsar is found within its 95 \% confidence contour. It could be a Geminga-like object or the effect of the combined emission of a pulsar and a weak SNR in Car 1A-B (the 3EG catalog notes that it is a possible case of multiple or extended source). Some of the low luminosity sources in Fig. ~\ref{fig.3} might be yet undetected SNRs, whereas the sources with the steepest indices could be background AGNs. A simple extrapolation of the high latitude population of $\gamma$-ray blazars shows that about 10 of this sources should be detected throughout the Galaxy within $|b|<10^o$ (Yadigaroglu \& Romani 1997). Most of them, however, should belong to the group of 43 3EG sources for which we have found not positional coincidences with any known galactic object. This set of sources has an average value of galactic latitude $<|b|>=5.8\pm3.3$, which suggests a significant extragalactic contribution. Finally, we want to mention two interesting additional possibilities to explain some 3EG sources: isolated Kerr-Newman black holes (Punsly 1998a,b) and isolated standard black holes accreting from the diffuse interstellar medium (Dermer 1997). In Punsly's model, a bipolar magnetically dominated MHD wind is driven by a charged black hole located in a low density region (otherwise it would discharge rapidly). The wind forms two leptonic jets which propagate along the rotation axis in opposite directions, as it occurs in AGNs. Self-Compton losses provide $\gamma$-ray luminosity in the range $10^{32}-10^{33}$ erg s$^{-1}$ for a 7-$M_{\odot}$ black hole with a polar magnetic field of $\sim10^{10}$ G. If such an object is relatively close ($\sim300$ pc), it could appear as a typical unidentified EGRET source with $\Gamma\sim2.5$. In the case of isolated black holes accreting from a diffuse medium, a hole with mass of 10 $M_{\odot}$ and a velocity of 10 km s$^{-1}$ can produce a $\gamma$-luminosity $\sim7\times10^{33}$ erg s$^{-1}$ in a medium with density of 0.1 cm$^{-3}$ (Dermer 1997). Changes in the particle density can result in $\gamma$-ray flux variability, as observed in several unidentified sources. None of these $\gamma$-sources based on black holes can be ruled out at present, and their observational signatures at other wavelengths seem to be worth of careful search. \section{Conclusions} We have studied the level of two-dimensional positional coincidences between unidentified EGRET sources at low galactic latitudes in the 3EG catalog and different populations of galactic objects, finding out that there is overwhelming statistical evidence for the association of $\gamma$-ray sources with SNRs and OB star forming regions (these latter considered as pulsar tracers). Additionally, there is marginally significant evidence for the association with early-type stars endowed with very strong winds, like Wolf-Rayet stars and Of stars. A posteriori analyses of the star candidates show that there are at least three systems (WR 140, WR 142, and Cyg OB2 No. 5) which are likely $\gamma$-ray sources. Several sources positionally coincident with OB associations are probably pulsars, like 3EG J1048-5840 and similar sources with hard spectra. Besides, there are 43 3EG sources for which we have not found any positional coincidence with known objects. This set of sources could include undetected low-brightness SNRs in interaction with dense and compact clouds, some Geminga-like pulsars, and, perhaps, a new kind of galactic $\gamma$-ray sources, like Kerr-Newman black holes or isolated black holes accreting from the interstellar medium. The main conclusion to be drawn is that there seems to exist more than a single population of galactic $\gamma$-ray sources. Pulsars constitute a well established class of sources, and there is no doubt that under certain conditions some SNRs are also responsible for significant $\gamma$-ray emission in the EGRET scope. Both isolated and binary early-type stars are likely to present high-energy radiation strong enough to be detected by EGRET in some special cases. We propose that, in addition to the well-known WR stars 140 and 142, the Cyg OB2 No. 5 binary system could be a strong $\gamma$-ray source, the first one to be detected involving no WR stars. The large number of unidentified EGRET sources free of any positional coincidence with luminous objects also encourage further studies to find whether there exist a population of exotic objects yet undetected at lower wavelengths. \begin{acknowledgements} This work has been partially supported by the Argentine agencies CONICET and ANPCT. \end{acknowledgements}
\section{Introduction} \setcounter{equation}{0} The problem of defining meaningful conserved charges in gauge theories is notoriously subtle. It has been addressed in the literature along various lines. One approach relies on the use of Noether identities and conserved currents \cite{No,BH2,Fl,Bg,Ju,BCJ,JS}. Another is based on Hamiltonian methods and asymptotic symmetries \cite{ADM,RT,Asht,AD}. The first approach is probably the most familiar and emphasizes locality. It has been recognized, however, that it suffers from ambiguities, which, if improperly resolved, may lead to incorrect results. It was recalled in \cite{JS} that in modern language, Noether already showed that on-shell the conserved charge associated to one-parameter subgroup of a gauge group is topological and hence lives at any infinity (ignoring singularities). More than $40$ years later \cite{ADM,RT}, it was understood that charges can indeed be defined at infinity for a good choice of boundary conditions, and in one to one correspondence with their (asymptotic) symmetries. The later may be finite or infinite in number. The purpose of this paper is to derive the correct superpotential for the supercharges in supergravity theories. We show that a ``natural'' application of the Noether identities yields an incorrect supercharge. We then derive a correct superpotential by adopting the criterion proposed in \cite{Si} and verify equivalence with the Hamiltonian approach. Finally, we explain in a first appendix why the first-order and second-order formulations yield the same superpotentials for local supersymmetries. The second appendix analyses in some detail the case of ${\cal N}_4 =2$ supergravity and provides explicit boundary conditions for the fields that enable one to meaningfully compute the charges and their algebra that contain a central charge as in the rigid supersymmetry case. \section{Noether superpotential} \setcounter{equation}{0} As shown by E. Noether, any continuous one-parameter invariance of the action leads to a conservation law $\partial_\mu j^\mu \approx 0$, where $\approx$ means ``{\it equal to} when the equations of motion hold''. The conserved Noether current is defined through \begin{equation} j^\mu = S^\mu - \frac{\partial L}{\partial \partial_\mu \phi} \delta \phi \label{NC} \end{equation} where $\delta \phi$ is the infinitesimal variation under the (local or global) symmetry and $\delta L = \partial_\mu S^\mu$. The fields $\phi$ may carry an index over which one sums in (\ref{NC}), but this will not be explicitly indicated. In the class of Lagrangians having the same (bulk) Euler-Lagrange derivative ($L \rightarrow L + \partial_\mu k^\mu$), one must adjust the surface term in such a way that the action $\int d^Dx L = 0$ is truly stationary on-shell. This surface term is fixed by a choice of boundary conditions. However, even for a given $L$, there is some ambiguity in the choice of $S^\mu$ since the addition to $S^\mu$ of the divergence $\partial_\nu k^{\mu \nu}$ of an antisymmetric tensor $k^{\mu \nu} = - k^{\nu \mu}$ does not modify $\partial_\mu S^\mu$. Expressed in terms of the currents, this (topological) ambiguity reads \begin{equation} j^\mu \rightarrow j^\mu + \partial_\nu k^{\mu \nu} \label{redef}\end{equation} and is particularly relevant in the case of gauge symmetries. Indeed, in this case, conserved Noether currents $j^\mu$ derive from superpotentials, \begin{equation} j^\mu \approx \partial_\nu U^{ \mu\nu} , \; U^{\mu \nu} = - U^{\nu \mu} \end{equation} This has been proved in many references (see \cite{No,BH2,Fl,Bg,Ju,BCJ,JS} and also \cite{BBH} for a cohomological interpretation). The fact that $j^\mu$ derives from a superpotential implies that one can set it equal to zero by means of the redefinitions (\ref{redef}). In other words, ``everything is in the superpotential", which indicates how crucial it is to resolve correctly the above-mentioned ambiguities. A wrong choice would lead to an incorrect integrated charge, \begin{equation} Q = \int_V d^{D-1}x \, j^0 \approx \int_{\partial V} d^{D-2} S_i\, U^{0i} \end{equation} which would e.g. not generate the appropriate transformations through the Poisson bracket. A ``natural'' choice for the superpotential may seem to be \begin{equation} U^{\mu \nu} = -\frac{1}{2} (M^{\mu \nu}_A - M^{\nu \mu}_A) \xi^A \label{wrongsup} \end{equation} with \begin{equation} M^{\mu \nu}_A = \frac{\partial L}{\partial \partial_\mu \phi}\Delta_A^\nu \label{w2} \end{equation} One finds \begin{equation} j^\mu = \partial_\nu U^{\mu\nu} + \frac{\delta L}{\delta \phi} \Delta_A^\mu \xi^A \label{current} \end{equation} where the $\Delta_A^\mu$'s are the coefficients of the derivatives of the gauge parameters in the variations of the fields, \begin{equation} \delta_\xi \phi = \xi^A \Delta_A + \partial_\mu \xi^A \Delta_A^\mu \label{deltafi}\end{equation} (we assume for simplicity that only the first-order derivatives appear). The choice (\ref{wrongsup}) corresponds to taking the coefficient of $\partial_\nu \xi^A$ in $S^\mu= \xi^A \Sigma_{A}^{\mu }+\partial_{\nu} \xi^A \Sigma_{A}^{\mu \nu }$ to be symmetric in $\mu$, $\nu$ and may be regarded as being ``natural" on this ground. However, this choice is not always correct and does in fact give an incorrect supercharge in supergravity for instance. To see this, note that only the fields that transform into derivatives of the gauge parameter contribute to the superpotential, and only the piece of the action containing derivatives of these fields is relevant. In all supergravity theories, the relevant part of the supersymmetry transformations is thus \begin{equation} \delta_\epsilon \psi^A_\sigma = \partial_\sigma \epsilon^A + \hbox{ ``more"} \end{equation} and the relevant piece in the action is the kinetic term for the gravitini\footnote{The conventions are the following: $\eta^{ab}=\{ -,+,\dots,+\}$, $\epsilon_{01\ldots (D-1)}=1$, $\gamma^a$ are $D$ real Majorana matrices and $\gamma^{a_1\ldots a_i}:=\gamma^{[a_1}\ldots\gamma^{a_i]}$ . The gravitini are described by Majorana spinors.}, \begin{equation} \frac{i}{2} \bar{\psi}^A_\lambda \gamma^{\lambda \mu \nu} \partial_\mu \psi^A_\nu. \end{equation} One finds by application of formulas (\ref{wrongsup}-\ref{w2}) \begin{equation} U^{ \mu\nu}_{\bar{\epsilon}} =- \frac{i}{2} \bar{\epsilon}^A \gamma^{ \mu \nu \lambda} \psi^A_\lambda. \label{incorrect}\end{equation} As discussed in the first appendix, one obtains (\ref{incorrect}) by working either in first or second order formalism. Although simple, the formula (\ref{incorrect}) is incorrect. It gives only half of the supercharge as can be seen by comparing with the Hamiltonian formalism \cite{T} or by computing the variation of the supercharge under a supersymmetry transformation, where one finds only half of the $4$-momentum $P^\mu$ instead of $P^\mu$ itself. The correct supercharge is \cite{T} \begin{equation} Q_{\bar{\epsilon}} =- i \int_{S_\infty} d^{D-2}S_i \; \bar{\epsilon}^A \gamma^{0ik} \psi^A_ k \label{supercharges}\end{equation} while the integral of $j^0$ given by (\ref{current}), (\ref{incorrect}) is clearly only half of this expression. That (\ref{incorrect}) is incorrect is perhaps not surprising since it is well appreciated that there exist in general relativity a plethora of superpotentials, many of which yield incorrect energy, momentum, or angular momentum \cite{Bg,Ko}. For a recent and informative discussion, see \cite{KBL}. We have just pointed out the supersymmetric extension of this problem. What is needed is a criterion that selects among the many candidate superpotentials the correct one. Such a criterion has been proposed in \cite{Si} and tested with success in many models. We apply below this criterion and show that it yields the correct supercharge. \section{Construction of the correct surface integrals at infinity} \setcounter{equation}{0} The approach proposed in \cite{Si} is a ``superpotential-based generalization" of the Hamiltonian approach of \cite{RT}. It may not always be equivalent to it, but in the case of theories like supergravity where one can write the Lagrangian in terms of forms and exterior products (in the sense specified after equation (\ref{defvm})), it does yield the same supercharge. The starting point of \cite{Si} is the relationship between the superpotential and the conserved current associated with a given one parameter group of gauge transformations of the fields through the Noether identities. For any choice of surface terms, this relationship reads \cite{JS} \begin{equation} j_U^\mu = \partial_\nu U^{ \mu\nu} + \frac{\delta L}{\delta \phi} \Delta_A^\mu \xi^A \label{currentbis} \end{equation} with the term proportional to the equations of motion being independent of the choice. The idea is then to find a criterion which, given the term proportional to the equations of motion, completes it in a definite way in equation (\ref{currentbis})\footnote{The current in (\ref{currentbis}) is conserved on-shell due to the antisymmetry of $U^{\mu\nu}$ and the so-called Noether identities, $\partial_\mu \left(\frac{\delta L}{\delta \phi} \Delta_A^\mu \xi^A \right) = \frac{\delta L}{\delta \phi} \delta_{\xi} \varphi$.}. In the case of supergravity, the supersymmetry current identity is \begin{eqnarray} j_{\bar{\epsilon}}^\mu &=& \partial_\nu U^{ \mu\nu}_{\bar{\epsilon}} + \bar{\epsilon}^A \frac{\delta L}{\delta \bar{\psi}_\mu^A} \nonumber \\ &=& \partial_\nu U^{ \mu\nu}_{\bar{\epsilon}} + i \bar{\epsilon}^A \gamma^{\mu \rho \sigma} (\partial_\rho \psi^A_\sigma + \Lambda_{\rho \sigma}^A) \label{jm} \end{eqnarray} where $\Lambda_{\rho \sigma}^A $ denotes terms containing undifferentiated gravitini fields. If one varies this equation with respect to the gravitini fields\footnote{In computing the superpotential associated with supersymetries, one may assume $\delta$(other fields)$=0$, since the terms proportional to $\delta$(other fields) die off faster at infinity, where the superpotential is actually defined. For instance, in four dimensions, this condition is verified in the asymptotically flat case, for which we adopt the boundary conditions of appendix B, or in the asymptotically anti-de Sitter case, for which we take the precise boundary conditions of \cite{HTads}. Thus the expression (\ref{superpotential}) for the superpotential associated with supersymmetries is correct in both cases.}, one gets, upon integration by parts, \begin{equation} \delta j_{\bar{\epsilon}}^\mu = \partial_\nu \delta U^{ \mu\nu}_{\bar{\epsilon}} + \partial_\nu V^{ \mu\nu} - i \partial_\nu \bar{\epsilon}^A \gamma^{\mu \nu \sigma} \delta \psi^A_\sigma + i \bar{\epsilon}^A \gamma^{\mu\rho \sigma} \delta \Lambda_{\rho \sigma}^A \label{divergence} \end{equation} with \begin{equation} V^{\mu\nu} (\epsilon, \phi, \partial \phi) = i \bar{\epsilon}^A \gamma^{\mu \nu \sigma} \delta \psi^A_\sigma \label{defvm}\label{defV} \end{equation} Note that here $V^{\mu \nu}$ is antisymmetric in $\mu$ and $\nu$. In fact, $V^{\mu\nu}$ is defined by (\ref{divergence}) up to a total divergence, $\tilde{V}^{\mu \nu}=V^{\mu \nu}+\partial_\rho X^{\mu[\nu\rho]}$, which can in the present case be adjusted so that $V^{\mu\nu}$ is antisymmetric in $\mu$ and $\nu$. Quite generally, the antisymmetry is guaranteed if the theory can be written in a first order formulation (see appendix of \cite{Si}). What is proposed in \cite{Si} is to take $U^{ \mu\nu}$ in such a way that the divergence terms cancel in (\ref{divergence}), i.e., such that \begin{equation} \delta U^{ \mu\nu}_{\bar{\epsilon}} + V^{ \mu\nu} = 0 \label{criterion} \end{equation} The integration in field space of this equation is straightforward and leads to \begin{equation} U^{\mu\nu}_{\bar{\epsilon}} = - i \bar{\epsilon}^A \gamma^{\mu \nu \sigma}\psi^A_\sigma \label{superpotential}\end{equation} without the factor one-half (recall that $\delta$(other fields) $=0$ - as stated in \cite{Si}, (\ref{criterion}) must be imposed only at infinity). As in the Hamiltonian formalism of \cite{RT}, equation (\ref{criterion}) defines the charge up to a constant which can be adjusted so that the superpotential vanishes for the vacuum. This superpotential is correct since, contrary to (\ref{incorrect}), it yields the correct supercharges (\ref{supercharges}). The last two terms of (\ref{jm}) are not relevant for charge evaluation. Actually the bulk charge has not yet been defined for singular solutions even in the presence of horizons. That the supercharge that follows from (\ref{criterion}) is the same as the one obtainable by Hamiltonian methods is easy to understand. Indeed, one may identify $\partial_i V^{0i}$ as the surface term that one picks up at {\em the spatial boundary} when breaking the variation of the spatial integral of the field-equation term in (\ref{currentbis}) \begin{equation} \delta \int_{x^0=C^{t}} d^{D-1} x \,\frac{\delta L}{\delta \phi} \Delta_A^0 \xi^A := \delta \int_{x^0= C^{t}} d^{D-1} x \, \bar{\epsilon}^A \frac{\delta L} {\delta \bar{\psi}_0^A} \end{equation} into a bulk term and a surface term, \begin{equation} \delta \int_{x^0= C^{t}} d^{D-1} x \,\bar{\epsilon}^A \frac{\delta L} {\delta \bar{\psi}_0^A} = \hbox{ ``bulk" } + \int_{S_\infty} d^{D-2} S_i \;V^{0i}. \label{varbulk} \end{equation} where ``bulk'' contains only undifferentiated variations of the canonical fields. This identification manifestly holds in supergravity. The only field equation that appears in (\ref{varbulk}) is the equation associated with the field that transforms into the time derivative of the gauge parameter, i.e. $\bar{\psi}^A_0$. Now, it is well-known that the temporal component $\bar{\psi}^A_0$ of the gravitino field is the Lagrange multiplier for the supersymmety constraint-generator, namely $\frac{\delta L} {\delta \bar{\psi}_0^A}$. Therefore, the term being varied in (\ref{varbulk}) is of the form ``Lagrange multipliers" times ``Hamiltonian constraints". This property, verified here for supergravity, is actually generic since the Lagrange multipliers transform always into the time derivatives of the gauge parameters (see e.g. \cite{HT}). Thus, the zeroth component of the criterion (\ref{criterion}) \begin{equation} \delta U^{0i}_{\bar{\epsilon}} + V^{0i} = 0 \end{equation} precisely guarantees that the charge \begin{equation} Q_{\bar{\epsilon}} = \int_{x^0= C^{t}} d^{D-1}x \; \bar{\epsilon}^A \frac{\delta L} {\delta \bar{\psi}_0^A} + \int_{S_\infty} d^{D-2}S_i \,U_{\bar{\epsilon}}^{0i} \end{equation} has well-defined functional derivatives in the sense of \cite{RT}, i.e., has a variation that contains only a bulk part. Since the asymptotic equality $\delta U^{0i}_{\bar{\epsilon}} + V^{0i} = 0$, together with a covariance argument at infinity, implies $\delta U^{\mu\nu}_{\bar{\epsilon}} + V^{ \mu\nu} = 0$ (asymptotically), one may conclude that in the case of supergravity, the superpotential method supplemented by the criterion of \cite{Si} and the Hamiltonian method are equivalent. \section{Conclusions} \setcounter{equation}{0} In this paper, we have shown how superpotential methods apply to local supersymmetry transformations in supergravity. We have rederived the correct supercharges. Once the supercharges are known, one may compute their algebra \cite{T,GH}. The case of ${\cal{N}}_4=2$ supergravity is particularly interesting because the algebra of the supercharges contains central charges. One of these, the magnetic central charge, arises in exactly the same way as the conformal central charge in $2+1$ anti-de Sitter gravity \cite{BH}: the central charge is not seen in the algebra of the asymptotic symmetries, but does appear in the algebra of their canonical generators. The calculation is direct. Because it lies somewhat outside the main line of this paper, it is discussed in the second appendix where precise boundary conditions that include magnetic sources are displayed. \section*{Acknowledgements} \bigskip We thank Eug{\`e}ne Cremmer and Thomas Materna for useful conversations. MH is grateful to the ``Laboratoire de Physique Th{\'e}orique de l'Ecole Normale Sup{\'e}rieure" for warm hospitality extended to him in May of 1998 while this work began. The work of MH has been partly supported by the ``Actions de Recherche Concert{\'e}es" of the ``Direction de la Recherche Scientifique - Communaut{\'e} Fran{\c c}aise de Belgique", by IISN - Belgium (convention 4.4505.86) and by Proyectos FONDECYT 1970151 and 7960001 (Chile). This work has been partly supported by the EU TMR contract ERBFMRXCT96-0012. \section*{Appendix A: Superpotential in first-order and second-order formalisms} \renewcommand{\theequation}{A.\arabic{equation}} \setcounter{equation}{0} The superpotentials for pure gravity are better understood in a pure $1^{st}$ order formalism, i.e. where the connection $\omega_{\mu b}^a$ is varied independently \cite{JS}. However, supergravities are usually given in a $2^{nd}$-order formulation of the supersymmetry transformations laws. That is the (spin)-connection is not treated as an independent field, and then, its supersymmetry transformation law is not required. The point of this appendix is to check in the special case of ${\cal{N}}_4=2$ supergravity that the above computation of the superpotential does not depend on which formalism ($1^{st}$ or $2^{nd}$ order) we are using. This provides a consistency check for \cite{Si}. The same consistency holds for pure gravity \cite{JS2}. First of all, the starting point for computing superpotentials is equation (\ref{currentbis}). In the special case of supergravity, the result was given in (\ref{jm}). Now, the only difference between a $1^{st}$ or $2^{nd}$ order computation would come from a term proportional to the equation of motion of the connection when computing $V^{\mu\nu}$ using the second term of the rhs of equation (\ref{currentbis}). However, if the $1^{st}$ order formulation of supergravity is such that the supersymmetry variation of the connection does not contain any term proportional to $\partial_\mu \xi^A$, no additional contribution is expected (recall the definitions (\ref{deltafi}) and (\ref{currentbis})). In a parallel publication \cite{HJS2}, we will present a general scheme to derive a pure $1^{st}$ order formulation from a $1.5$ one. That simply requires to compute a supersymmetry transformation law for the connection. Then, the simplest way to show that the superpotential computation for $1^{st}$ order formulation is equivalent to the $2^{nd}$ order one is to show that this supersymmetry transformation law does not depend on the derivative of the gauge parameter. This holds for the ${\cal{N}}_4=1$ supergravity, where the supersymmetry transformation law for the connection is \cite{DZ}: \begin{equation} \delta_1 \omega_{\mu}^{ab} = 2 i\kappa^2 \bar{\epsilon} \gamma^5 \left(\gamma_\mu \tilde{\psi}^{ab} - \frac{1}{2} e_\mu^a \gamma_c \tilde{\psi}^{cb} + \frac{1}{2} e_\mu^{b} \gamma_c \tilde{\psi}^{ca} \right) \label{varomf} \end{equation} where $ \tilde{\psi}^{ab} :=\frac{1}{2} \epsilon^{abcd} {\cal D}_c \psi_d$, with the covariant derivative acting on spinors defined as usual by ${\cal D}_{\sigma }:=\partial_{\sigma } + \frac{\gamma _{ab}}{4}\omega _{\sigma }^{ab}$. In the ${\cal{N}}_4=2$ supergravity case, it is also possible to find a $1^{st}$ order formulation, that is a supersymmetry transformation law for the connection. The result is \cite{HJS2}: $$ \delta_1 \omega_\mu^{ab} = 2 i\kappa^2 \bar{\epsilon}^A \gamma^5 \left(\gamma_\mu \tilde{\hat{\psi}}^{ab}_A -\frac{1}{2} e_\mu^a \gamma_c \tilde{\hat{\psi}}^{cb}_A + \frac{1}{2} e_\mu^b \gamma_c \tilde{\hat{\psi}}^{ca}_A \right) $$ \begin{equation} + i \kappa^3 \bar{\epsilon}_A \varepsilon^A_{\ B} \left(\hat{F}^{ab}- \tilde{\hat{F}}^{ab} \gamma^5 \right) \psi^B_\mu \label{tranw}\end{equation} Here $\tilde{\hat{\psi}}_{ab}^A := \frac{1}{2} \varepsilon_{abcd} \hat{{\cal D}}^c \psi^{dA}$. The $( {\cal N}_{4}={2})$-``hatted'' covariant derivative is defined by $\hat{{\cal D}}_{\sigma } :={\cal D}_{\sigma } +\frac{\kappa }{2}\varepsilon^A_{\ B}\left(\hat{F}_{\rho \sigma } \gamma ^{\rho }+\tilde{\hat {F}}_{\rho \sigma }\gamma^{\rho } \gamma ^{5}\right)$ together with (super-covariant) ``hatted'' field strength $\hat {F}_{\mu \nu }:= F_{\mu\nu }+i\kappa\varepsilon_{AB} \bar{\psi}^{A}_{\mu }\psi^{B}_{\nu }$ and its Hodge-dual $\tilde{\hat{F}}_{\mu \nu }:=\frac{1}{2}\varepsilon_{\mu \nu \rho \sigma } \hat{F}^{\rho \sigma}$. Again, there is no term in $\partial_\mu \bar{\epsilon}^A$ in the above transformation law. Another example is the eleven dimensional supergravity, which will be presented also in \cite{HJS2}. \section*{Appendix B: Boundary conditions, asymptotic symmetry algebra and algebra of charges for ${\cal N}_4 = 2$ supergravity} \renewcommand{\theequation}{B.\arabic{equation}} \setcounter{equation}{0} To compute the algebra of the charges of ${\cal N}_4 = 2$ supergravity, one first needs to give boundary conditions for the fields at spatial infinity that ensure that the surface integrals yielding the corresponding charges be all finite. The boundary conditions should also be invariant under the asymptotic symmetries, i.e., under the asymptotic rigid $N=2$ SUSY algebra. In the second order formalism, the independent fields of ${\cal N}_4 = 2$ supergravity are the tetrads $e^a_\mu$, the abelian connection $A_\mu$ and the two gravitini $\psi_\mu^A$. The boundary conditions for the tetrads and the gravitini are given in \cite{RT,T} (the tetrads approach their Minkowskian values up to terms falling off like $1/r$ with definite parity conditions, while the gravitini fall off like $1/r^2$). For this reason, we shall focus here only on the abelian connection. Our treatment includes magnetic monopoles. As in \cite{RT,T}, we shall express the boundary conditions in terms of the canonical variables, which are $A_k$ and $\pi^k$ (electric field). To enforce the Lorentz invariance properties of the boundary conditions, it is convenient to use the ``improved" form of the transformation of the connection under diffeomorphisms, which is \begin{equation} \delta_\xi A_\mu = \xi^\nu F_{\nu \mu} \label{potential} \end{equation} This is in fact the transformation that arises in the Hamiltonian formalism \cite{CT78}. Indeed, the constraints associated with diffeomorphisms are, for $N=2$ supergravity, \begin{eqnarray} {\cal H} &=& {\cal H}^G + {\cal H}^{em} + {\cal H}^{\frac{3}{2}} \approx 0 , \label{constraint1}\\ {\cal H}_k &=& {\cal H}^G_k + {\cal H}^{em}_k + {\cal H}^{\frac{3}{2}}_k \approx 0 \label{constraint2} \end{eqnarray} where ${\cal H}^G$ and ${\cal H}^G_k$ are the metric contributions, while ${\cal H}^{em}$ and ${\cal H}^{em}_k$ (respectively, ${\cal H}^{\frac{3}{2}}$ and ${\cal H}^{\frac{3}{2}}_k$) are the electromagnetic contributions (respectively, the gravitini contributions). We shall only need here the electromagnetic terms (the gravitational terms do not contribute to the transformations of the electromagnetic variables, while the spin $3/2$ terms can be neglected asymptotically). One has \begin{eqnarray} {\cal H}^{em}&=& \frac{g_{ij}}{2 \sqrt{g}} (\pi^i \pi^j + {\cal B}^i {\cal B}^j) \\ {\cal H}^{em}_k &=& F_{km} \pi^m \end{eqnarray} (electromagnetic energy density and Poynting vector). Here, ${\cal B}^i$ is the magnetic field and $g_{ij}$ the spatial metric induced on the surfaces $x^0 = C^{t}$; $g$ is its determinant. It is straightforward to check that $\int d^3x (\xi {\cal H} + \xi^k {\cal H}_k)$, where $\xi$ is the displacement normal to the hypersurface $x^0 = C^{t}$ and $\xi^k$ the tangential displacement, generates (\ref{potential}) when acting on $A_k$ through the Poisson bracket. The transformation law (\ref{potential}) for $A_0$ follows then from the general variation of the Lagrange multiplier under transformations generated by the constraints (see e.g. \cite{HT} equation (3.26b) - note that the brackets of the constraints (\ref{constraint1}) and (\ref{constraint2}) involve the Gauss' law constraint \cite{CT78}). The transformation of the electric field will also be evaluated through its Poisson bracket with the Hamiltonian constraints and is \begin{equation} \delta_{\xi} \pi^k = D_m(\xi F^{km} \sqrt{g}) + (\xi^m \pi^k),_m - \xi^k,_m \pi^m \label{electric} \end{equation} where $D_m$ is the spatial covariant derivative. The transformation (\ref{electric}) coincides on-shell with the standard transformation. To motivate the boundary conditions, consider first the zero-monopole sector. In Minkowski space, the electromagnetic potential for an electric charge at rest at the origin is given by \begin{equation} A_0 \sim \frac{e}{r}, \; \; A_k = 0 \end{equation} while the electromagnetic field behaves as \begin{equation} F_{0i} \sim \frac{e n^i}{r^2}, \; \; F_{ij} = 0 \end{equation} where $n^i$ is the unit radial vector. By boosting this solution, one generates non-vanishing $r^{-1}$-order terms for $A_k$; these terms have the interesting property of being even under the parity ${\bf n} \rightarrow -{\bf n}$. Similarly, the leading-order term in the electromagnetic field has the property of being odd under the same parity transformation. These features remain valid if one superposes various charges. This suggests taking as boundary conditions at spatial infinity (at each given time) \begin{equation} A_k = \frac{a_k^{(1)}({\bf n})}{r} + \frac{a_k^{(2)}({\bf n})}{r^2} + o(r^{-2}) \label{bcp} \end{equation} for the spatial components of the connexion and \begin{equation} \pi^k = \frac{p^k_{(1)}({\bf n})}{r^2} + \frac{p^k_{(2)}({\bf n})}{r^3} + o(r^{-3}) \label{bcef} \end{equation} for the electric field. These conditions imply \begin{equation} F_{ik} = \frac{f_{ik}^{(1)}({\bf n})}{r^2} + \frac{f_{ik}^{(2)}({\bf n})}{r^3} + o(r^{-3}) \label{bcmf}\end{equation} for the magnetic field. We impose also that the first coefficients in $A_k$ and $\pi^k$ has definite parity properties: \begin{equation} a_k^{(1)}( - {\bf n}) = a_k^{(1)}({\bf n}), \; \; p^k_{(1)}( -{\bf n}) = - p^k_{(1)}({\bf n}) . \end{equation} It follows that \begin{equation} f_{ik}^{(1)}(-{\bf n}) = - f_{ik}^{(1)}({\bf n}) \end{equation} One easily verifies that these conditions are invariant under Lorentz transformations and futhermore guarantee the vanishing of the boundary term of the variational principle, namely $\delta A_{0} \pi^{i}+\delta A_{j} F^{ji}$, when integrated at infinity. In fact, the combined asymptotic conditions of \cite{RT,T} and those given here are invariant under the full rigid $N=2$ SUSY algebra acting at infinity. We shall verify for example the asymptotic Lorentz invariance of the above boundary conditions. The invariance of (\ref{bcp}) follows from (\ref{potential}) and (\ref{bcef}), (\ref{bcmf}) and the fact that for boosts and rotations, the leading orders of $\xi$ and $\xi^k$ are both parity-odd. Similarly, the invariance of (\ref{bcef}) follows from (\ref{electric}) and the same observation on the parity of the leading orders of $\xi$ and $\xi^k$. One should stress that the boundary conditions adopted here can probably be somewhat relaxed. However, our goal is not to provide here the most flexible admissible boundary behaviour, but to give a consistent and complete set of boundary conditions that enforce asymptotically the $N=2$ supersymmetry algebra. The asymptotic symmetry algebra may actually involve also a central $U(1)$. This issue is somewhat subtle because there is no charged field in ${\cal{N}}_4=2$ (ungauged) supergravity. Whether there is a non trivial $U(1)$ at infinity depends on the topology of the spatial sections. To see this, one notes that the boundary conditions on the potential are invariant under gauge transformations that behave asymptotically as \begin{equation} \delta_{\Lambda } A_\mu = \partial_\mu \Lambda \label{reducible} \end{equation} with \begin{equation} \Lambda = \lambda_0 + \lambda({\bf n}) + o(r^0) \label{asympt} \end{equation} where $ \lambda_0$ is a constant and $\lambda({\bf n})$ is parity-odd, \begin{equation} \lambda(- {\bf n}) = - \lambda({\bf n}). \end{equation} However, the transformations associated with $\lambda({\bf n})$ are really irrelevant because their corresponding charges, given by the flux of the electric field times $\lambda({\bf n})$ at infinity (up to non-written Gauss' law constraint terms) \begin{equation} Q[\lambda({\bf n})] = \int_{S_\infty} \lambda({\bf n}) \mbox{\boldmath$\pi$} \cdot {\bf dS} \label{irrel} \end{equation} all vanish identically for the field configurations allowed here. Thus, one can factor them out. These transformations are somewhat the analogs of the supertranslations of the BMS group, which are eliminated in the same way through the parity conditions \cite{RT}. In fact, the main motivation for adopting similar parity conditions on the electromagnetic potential is to effectively remove the charges (\ref{irrel}), which do not seem to have a direct physical interpretation (these charges are {\em not} associated with higher multipole moments). Thus, only the constant piece $\lambda_0$ in (\ref{asympt}) is relevant. The charge associated with gauge transformation that tend to a constant at infinity reads (again up to unwritten Gauss' law terms) \begin{equation} Q[\lambda_0] = \lambda_0 \int_{S_\infty} \mbox{\boldmath$\pi$} \cdot {\bf dS} \label{electriccharge}. \end{equation} There may be additional surface terms if there are other boundaries (other asymptotic regions, or other surfaces on which boundary conditions are required: horizons, surfaces surrounding singularities), but since we are interested here only in the asymptotic symmetry generators in a single asymptotic region (called ``infinity''), we shall consider gauge transformations that are zero except in a vicinity of this asymptotic region. These gauge transformations are correctly generated by (\ref{electriccharge}), without any other surface contribution. Being zero at the other boundaries (if any) they certainly preserve any set of specific boundary conditions given there. Now, whether (\ref{electriccharge}) identically vanishes or not depends on the topology of the spatial sections. If these are homeomorphic to $R^3$, then (\ref{electriccharge}) is zero by Gauss' law and there is no non trivial $U(1)$ asymptotic generator. However, if the $2$-sphere at infinity is not contractible, (\ref{electriccharge}) needs not vanish (``charge without charge"). Furthermore, as a generator, (\ref{electriccharge}) acts non trivially on the gauge-invariant Mandelstam variables \begin{equation} M= \int_\gamma dx^k A_k \end{equation} where $\gamma$ is any path joining the asymptotic region being considered to any other asymptotic region. The variable $M$ is invariant under ``proper" gauge transformations, i.e., under gauge transformations that vanish on all boundaries. In that sense, it is an observable. However, it is not invariant under global transformations characterized by a non-vanishing $\lambda_0$ provided it ends outside the component of infinity under consideration. One gets $\delta_{\Lambda } M = \lambda_0$, which implies that $M$ is conjugate to the electric charge measured at infinity, \begin{equation} [M, \int_{S_\infty} \mbox{\boldmath$\pi$} \cdot {\bf dS}] = 1 \end{equation} (again, we do not write explicitly the constraint terms that accompany the surface integral at infinity). There is thus a non-trivial $U(1)$ generator at infinity, which generates asymptotic symmetries. Note that one would also find $U(1)$ symmetries on the other boundaries. The sum over all boundaries of the $U(1)$ charges is zero by Gauss' law (the flux lines can only end on the boundaries). The fact that the charges add up to zero just reflects the fact that the generator associated with everywhere constant gauge transformations vanishes, as it should (such transformations have no action on the fields). The inclusion of magnetic charges is direct. Instead of imposing (\ref{bcp}) on the potential $A_k$, one requires that it behaves asymptotically as \begin{equation} A_k = A_k^{mon} + A'_k \end{equation} where $A_k^{mon}$ is the potential for the magnetic monopole(s), living on a non-trivial $U(1)$ bundle over $S_2$ (and defined over patches), while $A'_k$ is globally defined and subject to the same boundary conditions (\ref{bcp}). The boundary conditions on the electric field are unchanged. The magnetic field has also the same asymptotic behaviour because the field of a monopole obeys (\ref{bcmf}). These more general boundary conditions are still Lorentz invariant because the variation of $A_k^{mon}$, which is globally defined, behaves as $A'_k$. They are in fact also invariant under supersymmetry transformations and constant gauge transformations. Furthermore, the (graded) commutator of two local supersymmetry transformations that tend asymptotically to constants $\epsilon^A$ and $\zeta^A$ is the sum of a diffeomorphism that becomes asymptotically a translation with parameter $-i \kappa^2 \bar{\epsilon}_A \gamma^\mu \zeta^A$ and a gauge transformation that behaves at infinity as a constant gauge transformation with parameter $\Phi = -i \kappa \bar{\epsilon}_A \varepsilon^A_B \zeta^B$ (the asymptotic value of the gauge transformation is defined up to a constant, but the ambiguity disappears if one imposes that $\Phi$ vanishes when $\epsilon$ or $\zeta$ is equal to zero). There is no sign of the magnetic charge in the asymptotic algebra of the asymptotic symmetry transformations. Let us turn now to the algebra of the charges defined, in the Lagrangian context, through the variations of the charges under the transformations generated by one another. The variation of the superpotential was given in equation (\ref{criterion}), together with the result (\ref{defV}) for supergravity theories. Now in the special case where $\delta$ is another supersymmetry gauge transformation, that is $\delta:=\delta_\zeta$, the variation becomes: \begin{eqnarray}\label{algebre} \delta_\zeta U_{\bar{\epsilon}}^{\mu \nu } &=& -i \bar{\epsilon}_A\gamma ^{\mu \nu \sigma } \hat{\cal D}_{\sigma }\zeta^A\nonumber\\ &=& -i \bar{\epsilon}_A\gamma ^{\mu \nu \sigma } \left( {\cal D}_{\sigma } +\frac{\kappa }{2}\varepsilon^A_{\ B}\left(\hat {F}_{\rho \sigma } \gamma ^{\rho }+\tilde{\hat {F}}_{\rho \sigma }\gamma ^{\rho } \gamma ^{5}\right)\right)\zeta^B \nonumber\\ &=& -i \bar{\epsilon}_A\gamma ^{\mu \nu \sigma }{\cal D}_{\sigma }\zeta^A -i\kappa \bar{\epsilon}_A \varepsilon^A_{\ B} \left(\hat {F}^{\mu \nu } + \gamma ^{5}\tilde{\hat {F}}^{\mu \nu } \right) \zeta^B \end{eqnarray} Where we used the $( {\cal N}_{4}={2})$-``hatted'' covariant derivative defined in appendix A, after equation (\ref{tranw}). Let us now integrate (\ref{algebre} ) at spatial infinity: \begin{enumerate} \item The first term in the rhs of (\ref{algebre}) is nothing but the Nester-Witten superpotential \cite{NW}, and so the covariant version of ADM mass (and momentum). We will denote it by $ H_{\xi }$, where $\xi^{\mu }$ is an asymptotic displacement and is given by $\xi^{\mu }=-i \kappa ^{2}\bar{\epsilon}_{A}\gamma ^{\mu }\zeta^{A}$. The idea of relating the positivity of gravitational mass and local supersymmetry can be found in \cite{DTG}. The supergravity origin and positivity of the Nester-Witten superpotential \cite{NW} was already established for the ${\cal{N}}_{4}=1$ supergravity in \cite{Hu}. \item The second term gives electric and magnetic contributions (``central'' charges) to the algebra of supersymmetry. Due to the asymptotic behavior of the gravitino (falling off like $1/r^{2}$), the hatted field strength can be replaced by the ordinary field strength, the difference between both falling off like $1/r^{4}$. \end{enumerate} Let us define: \[ Q :=- \int_{S_{\infty}} \left( i \kappa \bar{\epsilon}_A \varepsilon^A_{\ B}\zeta^B \right) F^{0i}\hspace{.15in} \mbox{and}\hspace{.15in} P :=-\int_{S_{\infty}} \left(i \kappa\bar{\epsilon}_A \varepsilon^A_{\ B}\gamma^5\zeta^B \right) \tilde{F}^{0i}\nonumber \] Then we find that the algebra of the SUSY charges (which generate the asymptotic supersymmetric invariance of ${\cal{N}}_{4}=2$ supergravity) has a magnetic central charge besides the expected $U(1)$ generator. These are familiar from rigid supersymmetry. Explicitly, we find the algebra \begin{equation} \delta_\zeta Q_{\bar{\epsilon}}:=[Q_{\bar{\epsilon}},Q_{\zeta}] = H_{\xi} + Q + P \label{susyAlgebra}\end{equation} in agreement with \cite{T,GH}. The electric and magnetic charges appear as central charges for the $N=2$ supersymmetry algebra. There is, however, an apparent difference between the two in the present (electric) formulation of ${\cal N}_4 = 2$ supergravity. While the Noetherian electric charge is a non trivial generator which does act on some canonical fields (when it does not vanish identically), the magnetic charge commutes with all of them and not just with the other generators of the asymptotic symmetry algebra, even when it is non zero. It is (in this limited treatment) a fully central charge. It is customary to adjust the constants in the generators of the asymptotic symmetries so that these vanish on some specified background. When specialized to that background, (\ref{susyAlgebra}) becomes \begin{equation} \delta_\zeta Q_{\bar{\epsilon}} = P \hbox{ on background}\end{equation} ($H_{\xi} =0$, $Q =0$). The central charge is the variation of the background generator of asymptotic supersymmetries. The magnetic central charge can (and does) arise because the electromagnetic field on a non-trivial $U(1)$-bundle is non zero and breaks some of the supersymmetries ($\delta_\zeta Q_{\bar{\epsilon}} \not= 0$). Note that in a dual (magnetic) formulation the role of electric and magnetic charges would be exchanged, a fully duality invariant treatment should restore the symmetry between the two. The algebra of the asymptotic symmetry transformations (once more defined on the electric canonical variables) and the algebra of their generators are different. The algebra (\ref{susyAlgebra}) gives another example of the general theorem of \cite{BH1} which establishes that extra central charges are allowed in the asymptotic charge algebras. This has been already used in \cite{BH} and generalises the classical ambiguity of the charge algebra relative to the symmetry algebra in Hamiltonian dynamics.
\section{Introduction} The Blume-Emery-Griffiths (BEG) model is a spin-1 Ising model, originally proposed to study $^3$He-$^4$He mixtures\cite{BEG}. Later, it was used in the description of a variety of different physical phenomena: multicomponent fluids \cite{exp1}, microemulsions \cite{exp2}, semiconductor alloys \cite{exp3}, electronic conduction models \cite{exp4}, etc. Its hamiltonian reads: \begin{equation} {\cal H} = - J \sum_{<i,j>} S_i S_j - K \sum_{<i,j>} S_i^2 S_j^2 + \sum_i \Delta_i S_i^2, \;\; \label{hamil} \end{equation} where the first two sums are over all nearest-neighbor pairs on a lattice, the last one is over all sites and $S_i = \pm 1,0$. $J$ is the exchange parameter, $K$ is the biquadratic interaction and $\Delta_i$ is a site-dependent crystal field ($\Delta_i=\Delta$ for the pure model). The phase diagram of the model presents first-order and continuous phase transitions and, for $K<0$, a rich variety of multicritical points \cite{ber1,ber2}. Nevertheless, some systems were found to be better described by a disordered BEG model, as, for instance, conventional shape memory alloys \cite{barce} and fluid mixtures on disorder materials (like aerogel, for example) \cite{mar1,mar2}. From the theoretical point of view, on the other hand, it has been argued that randomness may have drastic consequences on multicritical behavior \cite{Imry,berker}. In two dimensions, for instance, any infinitesimal amount of disorder supresses non-symmetry-breaking first-order phase transitions and replaces symmetry-breaking first-order phase transitions by continuous ones. The effect of disorder on three-dimensional systems is different: first-order phase transitions only disappear at a finite amount of randomness \cite{berker}. This behavior has been observed in some models \cite{berker,branco,cardy,falicov}. In order to study the effects of disorder on its phase-diagram, we study the BEG model in a random crystal field (henceforth called RBEG model) given by the probability distribution: \begin{equation} {\cal P}(\Delta_i) = r \; \delta(\Delta_i+\Delta) + (1-r) \; \delta(\Delta_i-\Delta) \label{distri} \end{equation} It is worthy stressing that the exact form of the disorder is not relevant to the overall {\it qualitative} consequences on the phase diagram. If randomness is chosen to be in the interactions $J$ or $K$, the qualitative effects will be the same (in what concerns first-order phase transitions). This is due to the fact that, even if the initial disorder is on the bonds (interactions $J$ or $K$), a scale transformation will propagate this disorder to the crystal field term, which will act just like field randomness on the coexistence boundary. Moreover, the exact form of the probability distribution is not relevant, either; we have performed calculations with other distributions and they lead to the same qualitative picture as the one found in this work. Finally, we would like to mention that, to the best of our knowledge, the BEG model in a random crystal field has not been studied so far. Previous studies concentrated on the random Blume-Capel model \cite{mar1,mar2,branco,mf1,mf2}, which has a simpler phase-diagram than the BEG model's. The remainder of this paper is organized as follows. In section II we outline the mean-field approximation we use and discuss the results, in section III we present the real-space renormalization-group (RSRG) calculation (expected to hold for two-dimensional systems), and in the last section we summarize our main conclusions and comment on the influence of a degeneracy parameter $p$ on the critical behavior. \section{Mean-field Calculation} We chose an ordinary mean-field approximation to study the three-dimensional system. The procedure is rather usual and we refer the reader to Ref. \onlinecite{bachmann}, where a detailed discussion of the method is done. However, we would like to stress that the mean-field approximation we use is equivalent to a model where the interaction is of infinite-range, i.e., each spin interacts with every other spin in the system. This will have explicit consequences on the phase diagram and we will return to this point later. Most of the information about the phase diagram is numerically calculated but some analytical results can be obtained. Among them, we can find the ground state for any values of $J$, $K$, $\Delta$ and $r$. It is possible to show that the ground state magnetization, $m_0$, for $d\equiv \Delta/zJ > 0$, where $z$ is the coordination number of the lattice, is given by [results for $\Delta<0$ can be inferred from the mapping $(r,\Delta) \leftrightarrow (1-r,-\Delta)$]: \begin{equation} m_0 = 1 - (1-r) \; \theta\left[ d - \left( k+1 \right) \left( \frac{1+r}{2} \right) \right] , \end{equation} where $k \equiv K/J$ and $\theta[x]$ is the step function, such that $\theta[x]=0$ or $1$ for $x<0$ or $x>0$ respectively. The ferromagnetic phase $O_1$ (see figures in this subsection), with $m_0=1$, is stable for $d \leq d_c = \left( k+1 \right) \left( \frac{1+r}{2} \right)$, while for $d \geq d_c$ the ground state is such that $m_0=r$ (denoted $O_2$ in our figures). Note that, except for $r=0$, the ground state is always ordered; this is a consequence of the simple mean-field approximation we used (we will return to this point below). One can obtain the continuous transition line exactly, by expanding $\Phi_{min}$ in powers of the magnetization $m$ and taking the coefficient of $m^2$ equal to zero: \begin{equation} t_c = 2 \left( \frac{1-r}{2 + e^{-k} e^{d/t_c}} + \frac{r}{2 + e^{-k} e^{-d/t_c}} \right), \end{equation} where $t_c \equiv k_B T_c / z J$. More specifically, note that, for $d \gg 1$, the value of the critical temperature is $t_c = r$. So, for any value of $r \neq 0$, the critical line between the paramagnet and the $O_2$ phases extends to $d = \infty$ (see figures in this subsection). This is not the expected behavior for a cubic lattice, for the following reason. The RBEG model for $d = \infty$ is equivalent to the site-diluted spin-1/2 Ising model, since for $\Delta=\infty$, a $+\Delta$ crystal field acting on a given site forces that site to be in the $S=0$ state (absent), while a $-\Delta$ field forces the site to be either in the state $S=1$ or in the state $S=-1$ (both represent a present site). Thus, only for high enough $r$ an infinite cluster of $S=\pm 1$ states will form and will be able to sustain order. Exactly at $r=r_c$, there is such an infinite cluster but its critical temperature is zero. Therefore, the critical parameter $d_c$ should only reach infinity for $r \geq r_c$. However, the simple mean-field analysis we made leads to $r_c=0$, since it is equivalent to a model with infinite-range interactions. In some cases \cite{mar1,mf2}, more elaborated mean-field-like procedures were applied to the Blume-Capel model in a random crystal field. Briefly, the consequence of these approaches is that the transition line between $O_2$ and $D$ phases does not extend to $d = \infty$ for all values of $r$. All other results are similar to the ones obtained with our simple mean-field approximation. We note in advance that the approach we used for the two-dimenisonal model leads to a finite value of $r_c$, as expected on physical grounds. We have already pointed out that $t_c(\Delta=\infty)$ does not depend on $K$; this comes from the mapping between the RDBEG model and the site-diluted spin-1/2 Ising model. The $S=0$ states (absent sites) play no role in the dynamics of the model and the present sites can only be in the states $S=1$ or $S=-1$; thus, the biquadractic interaction, $K$, is irrelevant in this limit. If, for instance, the probability distribution ${\cal P}(\Delta_i) = r \; \delta(\Delta_i) + (1-r) \; \delta(\Delta_i-\Delta) $ is used, the $\Delta = \infty$ limit will be equivalent to the site-diluted {\it spin-1} Ising model; then, $t_c(\Delta=\infty)$ will depend on $K$. Note that the discussion in this paragraph applies to the two-dimensional case as well. We now turn to the discussion of the $k_B T / z J \times \Delta / z J $ phase diagrams. In Figs.~\ref{k5r01}, ~\ref{k5r03}, ~\ref{k5r05} and \ref{k5r07} we depict sections of constant $K/J=5$, for many values of $r$. The phase diagram for $r=0$ (pure BEG model) is qualitatively the same as for $r=0.1$ (Fig.~\ref{k5r01}), except that the $O_2$ phase is not present. \begin{figure} \epsfxsize=6.5cm \begin{center} \leavevmode \epsffile{k5r01L.eps} \caption{Mean-field phase diagram for $K/J=5$ and $r=0.1$. Filled circles and the open square stand for critical end points and a critical point, respectively. Continuous (dashed) lines represent continuous (first-order) transitions. The phases are: ordered with $m=1$ $(O_1)$, ordered with $m=r$ $(O_2)$, disordered with $q>1/2$ $(D_1)$, and disordered with $q<1/2$ $(D_2)$.} \label{k5r01} \end{center} \end{figure} Note that the size of the ordered phases increases with $r$. This is expected, since $r$ is the fraction of sites which feel a $-\Delta$ crystal field (we have already commented on the ``tail'' which separates the $O_2$ and $D_2$ phases, given by $t_c=r$). Another important feature is the presence of a first-order line between two disordered phases, for $r=0.1$ and $r=0.3$. In both of these phases $m=0$ but $q>1/2$ for $D_1$, while $q<1/2$ for $D_2$. We would like to call attention for the phase diagram for $r=0.3$ (Fig.~\ref{k5r03}); this type of diagram is not present in the Blume-Capel model. \begin{figure} \epsfxsize=6.5cm \begin{center} \leavevmode \epsffile{k5r03L.eps} \caption{Mean-field phase diagram for $K/J=5$ and $r=0.3$. Same conventions as in Fig.~\ref{k5r01}; $T$ stands for tricritical points.} \label{k5r03} \end{center} \end{figure} \begin{figure} \epsfxsize=6.5cm \begin{center} \leavevmode \epsffile{k5r05L.eps} \caption{Mean-field phase diagram for $K/J=5$ and $r=0.5$. Same conventions as in Fig.~\ref{k5r03}.} \label{k5r05} \end{center} \end{figure} \begin{figure} \epsfxsize=6.5cm \begin{center} \leavevmode \epsffile{k5r07L.eps} \caption{Mean-field phase diagram for $K/J=5$ and $r=0.7$. Same conventions as in Fig.~\ref{k5r03}.} \label{k5r07} \end{center} \end{figure} The corresponding phase diagrams for $K/J=3$ show only three types of critical behavior: for $r$ near zero, they are similar to the phase diagram for $K/J=5$ and $r=0.3$ (Fig.~\ref{k5r03}); for intermediate values of $r$, the behavior is the same as for $K/J=5$ and $r=0.5$ (Fig.~\ref{k5r05}); and for $r$ close to one, the equivalence is with the diagrams for $K/J=5$ and $r=0.7$ (Fig.~\ref{k5r07}). The Blume-Capel model ($K/J=0$) has already been studied within mean-field approximations \cite{mar1,mar2,mf1,mf2}, although for different probability distributions; the results we find in this case are in qualitative agreement with those of Refs.\ \onlinecite{mar2} and \onlinecite{mf1} and we shall not depict all of them here. The only exception is the diagram for $r=0.1$ (Fig.~\ref{k0r01}), which is not present for higher values of $K/J$. \begin{figure} \epsfxsize=6.5cm \begin{center} \leavevmode \epsffile{k0r01L.eps} \caption{Mean-field phase diagram for $K/J=0$ and $r=0.1$. Same conventions as in Fig.~\ref{k5r03}.} \label{k0r01} \end{center} \end{figure} On general grounds, one should note that the mean-field approximation we employed suggests that the random crystal field does not destroy the first order transitions between disordered phases and between an ordered and a disordered phase. Even first order lines between ordered phases (like the one in Fig.~\ref{k0r01}) survive the introduction of randomness. \section{Two-dimensional Renormalization-group} It is well known that mean-field-like approximations are not suitable to describe low-dimensional systems. We have then to resort to a different technique, in order to study the RBEG model in two dimensions. RSRG procedures, on the other hand, have been successfully applied to two-dimensional systems. Note, however, that RSRG approximations, in general, do not lead to results as precise as those obtained with Monte Carlo simulations, phenomenological renormalization or conformal invariance techniques. Nevertheless, they allow for a correct description of universality classes, order of the transitions, crossover phenomena, etc. The procedure is the same as the one thouroughly discussed in Ref. \onlinecite{branco}. There is just one technical point we would like to stress. Although we start with a uniform distribution for $J$ and $K$, the renormalization procedure will introduce randomness in all renormalized quantities ($J', K'$ and $\Delta'$). A possible approach is to follow the successive renormalized distributions of these parameters in order to study the phase diagram. We adopted an alternative way, which forces the renormalized distributions to be the same as the initial ones, but with renormalized parameters, namely, $ {\cal P}'_{ap}(J) = \delta(J-J')$, ${\cal P}'_{ap}(K) = \delta(K-K')$ and ${\cal P}'_{ap}(\Delta_i) = r' \; \delta(\Delta_i+\Delta') + (1-r') \; \delta(\Delta_i-\Delta')$. The values of $J'$ and $K'$ are obtained imposing that the first moment of the actual distributions for $J$ and $K$ and of ${\cal P}'_{ap}(J)$ and ${\cal P}'_{ap}(K)$ are equal, respectively. The values $r'$ and $\Delta'$ are calculated imposing that the two lowest moments of ${\cal P}'_{ap}(\Delta)$ match those of the real distribution. This procedure has to be used with some care: in certain systems, where the random-field mechanism is important and the initial randomness is on the interaction ($J$, for instance), forcing the field back into a uniform distribution leads to incorrect results. In Ref. \onlinecite{yeo}, for instance, the crystal field probability distribution is maintained uniform throughout the renormalization procedure. Consequently, the critical behavior of the random model is characteristic of a high-dimensional system: the critical temperature of the tricritical point diminishes as randomness is increased but only reaches the zero temperature axis at a finite value of the disorder. As discussed in Ref. \onlinecite{berker}, the mechanism responsible for the lack of first-order phase transitions in two-dimensional random systems is the disorder in the crystal field, which is not taken into account by approximations such as the one used in Ref. \onlinecite{yeo}. In the model we study in this paper, however, the disorder in the field is not approximated away by our RSRG procedure. Our results for $K/J=2$ are presented in Fig.~\ref{k2r}, where we depict the $kT/zJ \times \Delta/zJ$ phase diagram for $r=0$ (pure BEG model), $r=0.2$, $r=0.45$, and $r=0.5$. Let us first comment on the pure BEG model (curve (a) of Fig.~\ref{k2r}). As for $K/J=5$ in three dimensions, there are two types of disordered phases: both have $m=0$ but $q>1/2$ for phase $D_1$ and $q<1/2$ for phase $D_2$. The continuous line between phases $O$ and $D_1$ belongs to the universality class of the Ising model: this line is attracted to the Ising fixed point, $C^* \equiv \left( J=0.4407, K=-0.07308, \Delta=-\infty \right)$. The dashed line between phases $O$ and $D_2$ is attracted to the fixed point $F_1 \equiv \left( J=\infty, K=\infty, \Delta=2(J+K) \right) $, which represents a first-order transition in both $m$ and $q$, i.e., the largest eigenvalue of the even and the odd sectors of the linearized RGT matrix are equal to $b^d$ (see Ref. \onlinecite{fisher}). On the other hand, the dashed line between phases $D_1$ and $D_2$ is attracted to the fixed point $F_2 \equiv \left( J=0, K=\infty, \Delta=2K+\ln 2 \right)$; in this fixed point only the largest eigenvalue of the {\it even} sector of the linearized RGT matrix is equal to $b^d$; this is a sign of a discontinuity in $q$ (but not in $m$) when the line is crossed \cite{fisher}. \begin{figure} \epsfxsize=6.5cm \begin{center} \leavevmode \epsffile{k2rL.eps} \caption{Renormalization-group phase diagram for $K/J=2$ and (a) $r=0$, (b) $r=0.2$, (c) $r=0.45$, and (d) $r=0.5$. Filled circles stand for critical end points, $O$ for the ordered phase and $D_1$ and $D_2$ for the two disordered phases (see text). Continuous (dashed) lines represent continuous (first-order) transitions. The transition lines extend to $\Delta \rightarrow \infty$ only for $r \geq 0.5$. Note that the critical end point (filled circle) is present only for $r=0$.} \label{k2r} \end{center} \end{figure} In curves $(b)$, $(c)$, and $(d)$ of Fig.~\ref{k2r} we depict the $kT/zJ \times \Delta/zJ$ phase diagram for $r \neq 0$. We note that the first-order line is either replaced by a line of continuous transitions (between $O$ and $D_2$ phases) or is eliminated (between $D_1$ and $D_2$ phases), for any infinitesimal amount of randomness. In fact, the first-order fixed point attractors, $F_1$ and $F_2$, are unstable along the $r$ direction. There is still a line separating the two disordered phases (not depicted in Fig.~\ref{k2r}), $D_1$ and $D_2$, for $r \neq 0$, but this line is attracted to the $\left( r=1/2, J=0, K=0, \Delta=\infty \right)$ fixed point. This point represents a model with independent spins, in which no phase transition can take place. We note that our results are in accordance with general arguments on the effects of randomness on multicritical phase diagrams \cite{berker}, although, to the best of our knowledge, some features of these arguments have never been tested so far. On the other hand, the whole line of continuous transitions for $r \neq 0$ belongs to the pure Ising model universality class, i.e., $C^*$ is a stable fixed-point along the $r$ direction. This is expected, since, for the hierarchical lattice used in this work, the specific heat critical exponent of the pure Ising model, $\alpha$, is negative and disorder is irrelevant, according to the Harris criterion \cite{Har}. For the corresponding model on a two-dimensional Bravais lattice, where $\alpha = 0$, the Harris criterion is inconclusive. The accepted behavior, when disorder is present, is the following: critical exponents of the random model retain the same values as their pure conterparts but logarithmic corrections are introduced by randomness \cite{Fabio}. Experimental results also indicate the same critical exponents for pure and random two-dimensional Ising model \cite{exper}. We would like to call attention to the behavior of the critical point which separates the $O$ and $D_2$ phases at $T=0$. For $r < 0.5$, the transition at zero temperature takes place at a finite value of $\Delta/zJ$. For $r \geq 0.5$, the critical line between the ordered and the disordered phases extends to $\Delta/zJ = \infty$ in the diagram. In fact, for $\Delta/zJ = \infty$ the RBEG model is equivalent to the site-dilute spin-1/2 Ising model, as discussed above. Thus, only for high enough $r$ an infinite cluster of $S=\pm 1$ states will form and will be able to sustain order. There is a critical value of $r$, $r_c$, such that the critical line between the ordered and disordered phases only reaches $\Delta/zJ = \infty$ for $r \geq r_c$. Our evaluation of $r_c$ is 1/2, while the accepted value for the site percolation critical probability on the square lattice is $r_c=0.5927$ \cite{perco}. This difference is due to the small cell we use in this work; nevertheless, the correct qualitative picture is obtained, i.e., a finite value of $r_c$. Finally, we would like to stress that there are only two types of phase diagrams for the BEG model; for high values of $K/J$ these diagrams have the same structure as for $K/J=2$. For small values of $K/J$, the structure is the same as for the Blume-Capel ($K=0$) model. As this model has been studied elsewhere \cite{branco}, we will not discuss it here. \section{Summary} We studied the BEG model in two and three dimensions within a RSRG framework and a mean-field approximation, respectively. The disorder is on the crystal field term, which follows a probability distribution given by: ${\cal P}(\Delta_i) = r \; \delta(\Delta_i+\Delta) + (1-r) \; \delta(\Delta_i-\Delta)$. For the mean-field approximation (expected to represent the qualitative behavior of three-dimensional systems), the presence of randomness increases the ordered phase and brings qualitative changes to the $kT/zJ \times \Delta/zJ$ phase diagram. More specifically, first-order transitions are present in the disordered model, but new multicritical points emerge, depending on the value of $r$. In two dimensions, the RSRG approach we use shows that randomness has a drastic effect on critical behavior: it supresses non-symmetry-breaking first-order transitions and replaces symmetry-breaking discontinuous transitions by continuous ones. These results are in accordance with general arguments concerning the effects of quenched impurities on multicritical behavior (but, to the best of our knowledge, the disappearance of the first-order line between disordered phases or between ordered phases has never been seen in an actual calculation). The line of continuous transition, present for the disordered $(r \neq 0)$ model, belongs to the universality class of the two-dimensional {\it pure} Ising model; this results agrees with the Harris criterion, since the specific heat critical exponent, $\alpha$, is negative for the hierarchical lattice used in this work. It has been conjectured that a new unstable critical point, at finite temperatures, might be present for the disordered system \cite{mar1}; we found no evidence of this point, for any value of $K$. We have also studied the so-called degenerate Blume-Emery-Griffiths (DBEG) model, introduced in the study of martensitic transitions \cite{barce}. In the DBEG model, the $S=0$ states are assumed to have a degeneracy $p$, which mimics the effects of vibrational degrees of freedom. It has been shown in Ref.\onlinecite{barce} that the effect of increasing $p$ is to shrink the ordered phase and to increase the region where the transition is of first-order. Using the same probability distribution for the crystal field as in the RBEG model, we were able to show that the parameter $p$ may bring only {\it quantitative} changes to the phase diagrams, for any $K/J$, $r$, and $p$. This is expected, since the DBEG model is equivalent to the usual $BEG$ model with all crystal fields displaced by $\ln(p)$ \cite{referee}. In particular, any infinitesimal amount of randomness in two dimensions destroys the first order transitions, no matter the value of $p$. Finally, we would like to stress that our approximation does not allow for a study of the BEG model with negative $K$, where new and interesting critical behavior emerges \cite{ber2}. Work is now being made to discuss this model in the presence of a random crystal field. \section{Acknowledgments} We would like to thank Prof. J. F. Stilck for a critical reading of the manuscript and Prof. Anna Chame for calling our attention to Ref. \onlinecite{barce} while this work was in progress.
\section{Introduction} The nearby spiral galaxy NGC~7331 the main parameters of which are given in Table 1 represents a challenge for the observers interested in structure and dynamics of disk galaxies. It has been observed many times, both photometrically and spectroscopically; but results of each new approach contradicted often to results of previous ones. Many years ago Bosma (\cite{bosma}) trying to give a general description of NGC~7331 noted that according to Sandage (\cite{ha}) who saw spiral arms as close to the center as at $r\approx 6\arcsec$ NGC~7331 is a disk-dominated galaxy; however the deep photograph reported by Arp and Kormendy (\cite{ak72}) revealed a presence of prominent extended bulge. The problem of the bulge role in NGC~7331 is not solved yet despite numerous photometric studies. Boroson (\cite{bor}) estimated a bulge-to-disk ratio as 1.10 by analysing a major-axis surface brightness profile. Kent (\cite{kent}) analysed two-dimensional CCD images and proposed a method of bulge-disk decomposition based on different ellipticities of the bulge and disk isophotes; but he noted that this method is inapplicable to NGC~7331 because in this galaxy the isophotes of the bulge and of the disk demonstrate the same (!) ellipticity. As a result, he decomposed only the major-axis profile and derived $B/D=0.66$ and the disk with a central hole. Among recent studies, von Linden et al. (\cite{germ}) have published a surface brightness profile in the $I$--band extended up to $R\approx 100\arcsec$ and have concluded that a de Vaucouleurs' bulge dominates over the whole radius range under consideration, whereas Prada et al. (\cite{espls}) analysing together $I$-- and $K$--images have reached the best fit with the compact bulge, having effective radius of $\sim 10\arcsec$, and two exponential disk components with different characteristic scales which meet at $R\approx 100\arcsec$. So the situation with the morphological characteristics of the bulge in NGC~7331 remains to be uncertain. A similar uncertainty exists relating to a dynamical status of the bulge in this galaxy: Prada et al. (\cite{espls}) have found the bulge to counter-rotate with respect to the stellar disk basing on the long-slit observations in the \ion{Ca}{2}\,IR triplet spectral range, but Mediavilla et al. (\cite{esptig}) who observed the central part of NGC~7331 with a panoramic fiber spectrograph in two spectral ranges, near Mgb and near \ion{Ca}{2}\,IR, do not agree stating that the bulge in NGC~7331 corotates the stellar and gaseous disks. \begin{table} \caption[ ] {Global parameters of NGC~7331} \begin{flushleft} \begin{tabular}{lc} \hline\noalign{\smallskip} Type (NED) & SA(s)b \\ $R_{25}$, kpc (LEDA) & 22.5 \\ $B_T^0$ (LEDA) & 9.27 \\ $M_B$ (LEDA) & --21.41 \\ $V_r(radio) $ (LEDA) & 821 $km\cdot s^{-1}$ \\ Distance, Mpc (Hughes et al. 1998) & 15.1 \\ Inclination (LEDA) & $70^o$ \\ {\it PA}$_{phot}$ (LEDA) & $171^\circ$ \\ \hline \end{tabular} \end{flushleft} \end{table} Another question which is an object of discussion for several years: is there a "dead quasar", or supermassive black hole, in the center of NGC~7331? We have begun this discussion ten years ago. Afanasiev et al. (\cite{asz89}) examining the major-axis profile of line-of-sight velocities of the ionized gas have found that central $\pm 2\arcsec$ (the scale which is comparable to our spatial resolution) are kinematically decoupled by fast solid-body rotation. We have checked the axisymmetric character of the gas rotation and have concluded that this fast rotation is caused by a compact mass concentration of order of $5 \cdot 10^8 \, M_\odot$. To see if a "dead quasar" may be in the center of NGC~7331, Bower et al. (\cite{betal}) have studied major-axis and minor-axis profiles of stellar velocities. Though the angular rotation velocity of stars in the center of NGC~7331 has been found to be as high as that of the ionized gas, namely, about of 500 km/s/kpc, the very central part of the stellar rotation curve has not appeared to be so distinct. Moreover, the stellar velocity dispersion has not a sharp maximum in the center, as it is the case, for example, in M~31 where the presence of the supermassive black hole in the center is proved (\cite{dr88,k88}). On the contrary, in NGC~7331 $\sigma (r)$ along the minor axis seems to demonstrate a local minimum at $r\approx 0\arcsec$. So Bower et al. (\cite{betal}) have concluded that the mass of the central black hole, if exists, must be less than $5 \cdot 10^8 \, M_\odot$. However, the problem of the black hole presence in the nucleus of NGC~7331 has not been closed: Cowan et al. (\cite{crb}) have reported a detection of unresolved nuclear radio source in this galaxy which has appeared to be more luminous by a factor of 3--4 than the famous Sgr~A in the center of our Galaxy, and recently Stockdale et al. (\cite{src}) have claimed an existence of the nuclear X-ray source in NGC~7331 -- they treat it as a massive black hole. If it really exists, its dynamical implications must be re-looked for more carefully. When Afanasiev et al. (\cite{asz89}) have detected a compact mass concentration in the center of NGC~7331, we have not made a definitive choice between the supermassive black hole and a compact dense stellar subsystem, such as a central star supercluster or circumnuclear stellar disk. Moreover, later we have undertaken a two-dimensional spectrophotometry of the central region of NGC~7331 and have found that its unresolved nucleus is chemically distinct: its magnesium index is much higher than that of the nearest bulge (\cite{sil92}). As the kinematically decoupled nucleus in NGC~7331 has been found to be also chemically distinct, we would like to think it to be a separate stellar subsystem. So a more careful investigation of stellar population properties in the nucleus and in the circumnuclear region of NGC~7331 must help to clarify the structure and evolution of its center and of the entire galaxy. We report our observations and other data which we use in Section~2. The brief description of the ionized-gas morphology in the center of NGC~7331 is given in Section~3. Radial variations of the stellar population age and metal abundances are analysed in Section~4, and the kinematics of ionized gas and stars in the region under consideration is discussed in Section~5. Section~6 presents our conclusions and a brief discussion of our results. \section{Observations} The observations of NGC~7331 which results are presented in this paper have been undertaken with the Multi-Pupil Field Spectrograph (MPFS, \cite{afetal90}) of the 6m telescope of the Special Astrophysical Observatory of the Russian Academy of Sciences in 1996. The red spectral range, 6300--6900~\AA, has been exposed on August 19, 1996, during 55 minutes, under the seeing $FWHM \approx 2\farcs 4$. The MPFS was equipped with CCD $520 \times 580$; we registered simultaneously 95 spectra from an area of $19\arcsec \times 12\farcs 6$, each spectrum corresponding to a spatial element of $1\farcs 58 \times 1\farcs 58$. The strongest emission line in the circumnuclear region, [NII]$\lambda$6583, has been measured to construct two-dimensional velocity field of the ionized gas; we fitted its profile by a single Gaussian. The blue-green spectral range, 4100--5600~\AA, has been exposed on October 9, 1996, during 40 minutes, under the seeing $FWHM=1\farcs 6$. The sky has been exposed separately during 20 minutes, properly normalized and subtracted from the galaxy frame. The MPFS was equipped with CCD $1040 \times 1160$; we registered simultaneously 128 spectra from an area of $11\arcsec \times 21\arcsec$, each spectrum corresponding to a spatial element of $1\farcs 33 \times 1\farcs 33$. This spectral range includes a lot of strong absorption lines, so we have used this exposure to calculate absorption-line indices H$\beta$, Mgb, Fe5270, and Fe5335 in the well-known Lick system (\cite{woretal}). We have checked our consistency with the Lick measurements by observing stars from their list (\cite{woretal}) and by calculating the absorption-line indices for the stars in the same manner as for the galaxies. The indices calculated for 9 stars are coincident with the data tabulated in (\cite{woretal}) in average within 0.05~\AA. The exposure for the galaxy has been taken long enough to provide a signal-to-noise ratios of $\sim$ 60 in the nucleus and $\sim$ 10--15 near the edge of the frame; the corresponding random error estimations made in the manner of Cardiel et al. (\cite{cgcg}) range from 0.2~\AA\ in the center to 0.6--0.7~\AA\ for the individual spatial elements at the outermost points. To keep a constant level of accuracy along the radius, we summed the spectra for the galaxies in concentric rings centered onto the nuclei and studied the radial dependencies of the luminosity-weighted properties of stars by comparing the observational index values to those from the synthetic models of old stellar populations of Worthey (\cite{worth94}) and Tantalo et al. (\cite{tantalo}). We estimate the mean accuracy of our azimuthally-averaged indices as 0.1--0.2~\AA. Besides, we have cross-correlated each elementary spectrum with the spectrum of the late-type star ADS~15470 (the brighter component) and have obtained two-dimensional velocity field for the stellar component of the central region of NGC~7331. The reciprocal dispersion during these observations was 1.6~\AA\ per pix, and the spectral resolution varied slightly over the field of view from 3.5~\AA\ to 5~\AA. As a result, we estimate our accuracy of the elementary line-of-sight velocities as 20--25 km/s. However, our spectral resolution allowed to obtain only luminosity-weighted mean line-of-sight velocitites; we were not able to separate kinematical components similar to the counterrotating bulge reported by Prada et al. (\cite{espls}). To refine kinematical analysis, we have involved high-resolution long-slit data from the La Palma Archive. The galaxy has been observed on July 19, 1996, with the ISIS (red arm) equipped with CCD TEK $1024 \times 1024$ at the William Herschel Telescope. The near-infrared spectral range, 8360--8750~\AA, containing the strong CaII absorption-line triplet, has been exposed: 60 min in the $P.A.=172\arcdeg$ (major axis) and 30 min in the $P.A.=262\arcdeg$ (minor axis). The slit width was $0\farcs95$; the reciprocal dispersion of 0.39~\AA\ per pix provided spectral resolution of 1.0~\AA. A star HR~8656 (K0III) was observed during the same night; we have cross-correlated the galaxy spectra row-by-row with the spectrum of this star after binning by three pixels (with the final spatial step of $1\arcsec$), sky subtracting and transforming into velocity scale. The subsequent Gauss analysis of obviously multi-component LOSVDs (line-of-sight velocity distributions) has allowed to extract several kinematical components along the slit aligned with the major axis of the galaxy. The basic reduction steps -- bias subtraction, flatfielding, cosmic ray hit removing, extraction of one-dimensional spectra, wavelength calibration, construction of surface brightness maps -- have been fulfiled by using the software developed in the Special Astrophysical Observatory (\cite{vlas}). To calculate the absorption-line indices and to map them we have used our own programs as well as the FORTRAN program of Dr.~Vazdekis. \section{Morphology of the Ionized Gas Distribution in the Center of NGC~7331} Fig.~1 shows raw observational data -- so called "data cube" -- in the red spectral range. One can see immediately that a characteristic ratio of emission lines, $H\alpha$-to-[NII]$\lambda$6583, indicates a LINER-like excitation inside the central region, roughly $16\arcsec \times 10\arcsec$, but a rather strong present star formation already in $11\arcsec$ to the east (along the minor axis) from the nucleus. This fact contradicts to the claims of Keel (\cite{keel}) who observed the spectra in $10\arcsec$ to the north, south, and east (!) from the nucleus and everywhere had found the dominance of [NII]$\lambda$6583 over H$\alpha$. But his aperture, $8\farcs 1$, was perhaps too large to distinguish between the central shock-excited gaseous disk and nearby star-forming sites which are concentrated in a ring with a deprojected radius of $45\arcsec$ (or $15\arcsec$ on the sky plane along the minor axis). This ring is presented best of all in the recent paper of Smith (\cite{smith}). It looks quite identical at different wavelengths: on the H$\alpha$+[NII] map from Pogge (\cite{pogge}), on the 15$\mu$m map from Smith (\cite{smith}), on the Nobeyama CO (1-0) map from Tosaki and Shioya (\cite{jap}), and on the 20\,cm radio continuum map from Cowan et al. (\cite{crb}). Obviously, this multiple coincidence together with H$\alpha$/[NII]$\lambda 6583 > 2$ in our Fig.~1 proves that this ring contains sites of intense present star formation, including formation of massive stars. If we accept the surface-brightness profile decomposition from Boroson (\cite{bor}) or from Baggett et al. (\cite{bag}), this star-forming ring is located well inside the bulge-dominated area. The nature of the very central region is not so obvious. Fig.~2 presents isophotes of the [NII] emission-line surface brightness distribution obtained with the MPFS. They are elongated and repeat approximately the shape of continuum isophotes. The consistent results are obtained by Mediavilla et al. (\cite{esptig}) (see their Fig.~9); even an asymmetry "north-south" is the same. Keel (\cite{keel}) also mentioned a gaseous disk with the radius of $2\arcsec$ having the diffuse extension up to $10\arcsec$ aligned in the $P.A.=165\arcdeg$, close to the line-of-nodes orientation. Interestingly, the similar elongated structure, with the major-axis diameter of $7\arcsec - 9\arcsec$, is present on the 15$\mu$m map of Smith (\cite{smith}), but the sense of the north-south asymmetry is opposite with respect to the [NII] map. Probably, the explanation of this asymmetry by the dust concentration to the south from the nucleus given by Mediavilla et al. (\cite{esptig}) is correct. There are no either HI (\cite{begeman}) nor CO (Tosaki \& Shioya \cite{jap}) inside this circumnuclear structure; also it is undetected in radio (Cowan et al. \cite{crb}). Therefore, there is no detectable present star formation there; but the gas is probably shock-excited, and the dust is warm. As the structure is aligned with the line of nodes, we would like to treat it preliminarily as a circumnuclear gaseous disk. \section{Stellar Population Properties in the Circumnuclear Region of NGC~7331} Figure~3 presents isolines of the surface distributions in the central $9\arcsec \times 10\arcsec$ area for the green (5100~\AA) continuum and three absorption-line indices, Mgb, $<\mbox{Fe}>\equiv$ (Fe5270+Fe5335)/2, and H$\beta$. The continuum isophotes show a strange asymmetry: within $3\arcsec$ from the center eastern halves of them seem to be more tightly packed than their western halves. It looks like a some extracomponent of light to the west from the nucleus. Interestingly, on larger scales the sense of isophote crowding is opposite due to global dust concentration on the western side of the galaxy (Boroson \cite{bor}, Smith \cite{smith}). The magnesium-index isolines are rather roundish; the distribution is strongly peaked on the nucleus of the galaxy (NGC~7331 has a chemically distinct nucleus, as we noted earlier, \cite{sil92}). But the surface distribution of another metal-line index, $<\mbox{Fe}>$, is quite different from that of Mgb and can partly explain an asymmetry of the continuum isophotes. In Fig.~3c we clearly see a compact Fe-rich disk shifted to the west from the nucleus (or do we see only its western half?). A strong north-south asymmetry is present too. If we assume that we see only south-western quarter of the disk, its radius may be as large as $3\farcs5$. This size agrees with the size of the central mid-infrared structure (Smith \cite{smith}). But what is the most interesting thing, it is the similar surface distribution of the hydrogen-line index H$\beta$ (Fig.~3d): the circumnuclear Fe-rich disk is also distinguished by the very prominent Balmer absorption line. Since iron and hydrogen absorption lines in integrated spectra are unmatched being produced by different groups of stars, the similarity of the Fig.~3c and Fig.~3d may signify that the circumnuclear Fe-rich stellar disk is rather young. Let us look at what state-of-art diagnostics of the stellar population properties can tell us. To compare our measurements to the stellar population models based on summation (with some weights) of spectra of stars, we must made corrections for the stellar velocity dispersion in galaxies which broadens absorption lines and "degrades" a spectral resolution in such way. We have calculated the correction by smoothing the spectra of K0-K3 III giants from the list of (\cite{woretal}) which we have observed and by measuring the absorption-line indices of the smoothed spectra. We have found that the index H$\beta$ is quite insensitive to the velocity dispersion when $\sigma_v$ remains to be less than 230 km/s; as for the metal-line indices, we have found the correction to be 0.1~\AA\ for $\sigma_v$=130 km/s which is typical for the central part of NGC~7331 (Bower et al. \cite{betal}). Figures~4 and 5 contains the corrected indices. The most popular models of Worthey (\cite{worth94}) allow to disentangle age and metallicity of old stellar populations by confronting some metal-line indices (e. g. Mgb, $<\mbox{Fe}>$, or [MgFe]$\equiv (\mbox{Mgb} \, <\mbox{Fe}>)^{1/2}$) with the Balmer-line index H$\beta$; but these models have been calculated for the solar magnesium-to-iron ratio. To use the Worthey's (\cite{worth94}) models, we must be sure that the stellar population has solar magnesium-to-iron ratio. Figure~4 presents the diagram (Fe5270, Mgb) where we compare azimuthally-averaged (in circular rings centered onto the nucleus) index measurements in NGC~7331 with the models of Worthey (\cite{worth94}) for the solar magnesium-to-iron ratio. Since the circumnuclear Fe-rich disk seen in Figs.~3c and 3d is located asymmetrically with respect to the nucleus and contributes only partially to the ring-integrated estimates at $r=1\farcs 3$ and $r=2\farcs 6$, we have also plotted individual-element measurements related to this disk though they are less accurate than azimuthally-averaged ones. One can see from Fig.~4 that the nucleus of NGC~7331 is surely magnesium-overabundant. The Mg/Fe ratio for the azimuthally-averaged measurements at $r=1\farcs 3$ and $r=2\farcs 6$ is obviously lower than that for the nucleus, but is it solar or not, depends on the stellar population age. Farther from the nucleus, at $r \geq 4$, the moderate magnesium overabundance can also be seen. Interestingly, the proper measurements of the circumnuclear Fe-rich disk lie much higher that the azimuthally-averaged points: the disk has the solar Mg-to-Fe ratio if it is as young as 2 billion years old and is iron-overabundant if it is older. Taking in mind all said above, let us try to determine mean ages of stars in the nucleus and at different distances from the center. Figures~5a, 5b, and 5c present various diagrams which may be useful for this purpose. As the nucleus is magnesium-overabundant, the models of Worthey (\cite{worth94}) are inapplicable for it; in Fig.~5a we have plotted the calculations of Tantalo et al. (\cite{tantalo}) for [Mg/Fe]=+0.3. From comparison with these models on the diagram (H$\beta$, $<\mbox{Fe}>$) one can see that the mean age of the nuclear stellar population in NGC~7331 is 5 billion years, and its global metal content, $Z$, is close to the solar value. Fig.~5b shows also the models from Tantalo et al. (\cite{tantalo}), but for the solar magnesium-to-iron ratio; the locations of the azimuthally-averaged points at the $r=1\farcs 3$ and $r=2\farcs 6$ imply similarly a rather young age for the nearest neighborhood of the nucleus. The elementary measurements related to the circumnuclear Fe-rich disk, though well scattered, nevertheless all lie above the model sequence with the age of 5 billion years. The next Fig.~5c presenting the diagram (H$\beta$, [Mg\,Fe]) with the models of Worthey (\cite{worth94}) calculated under [Mg/Fe]=0 confirms that four individual points for the circumnuclear Fe-rich disk agree with the age estimate of 2 billion years and the overall metallicity at least twice the solar one. If we return now to Fig.~4, we should conclude that the circumnuclear "Fe-rich" disk has indeed [Mg/Fe]=0 under the assumption of $T=2$ billion years. The work of Tantalo et al. (\cite{tantalo}) proposes also a possibility to quantify differences of stellar population properties basing on the index differences. A set of three linear equations, connecting $\Delta$[Mg/Fe], $\Delta \log Z$, and $\Delta \log T$ to the $\Delta \mbox{Mg}_2$, $\Delta <\mbox{Fe}>$, and $\Delta \mbox{H}\beta$, is written. We apply these equations to the differences between the nucleus and the bulge; the bulge is safely taken at the following values of radius, at $r=4\arcsec$, $5\farcs 3$, $6\farcs 7$, and $8\arcsec$, namely, outside the circumnuclear "Fe-rich" disk but well inside the star-forming ring. Solely, one must take in mind that the index measurements at $r > 6\arcsec$ are twice less precise than the more inner ones so they can be used mostly as a check. Having performed the set of calculations, we have obtained the parameter differences listed in Table~2. They mean that the bulge is twice older and more metal-poor by a factor of 2.5--4 than the nucleus. Surprisingly, the Mg/Fe ratios are almost equal in the nucleus and in the bulge. When we compare the absolute values of nuclear indices, $\mbox{Mg}_2=0.229 \pm 0.005$, $<\mbox{Fe}>=2.66 \pm 0.22$, and H$\beta = 1.83 \pm 0.19$, to the direct model calculations of Tantalo et al. (\cite{tantalo}), we see that the model with $Z=0.02$ (solar value), [Mg/Fe]=+0.3, and $T=5$ billion years has consistent index values. Then the bulge stellar population parameters are $Z=0.005-0.008$, [Mg/Fe]=+0.2, and $T=9-12$ billion years. The latter age estimate agrees also with the positions of bulge points in Fig.~5a. \begin{table} \caption[ ] {Stellar population parameter differences "bulge-nucleus"} \begin{flushleft} \begin{tabular}{lccc} \tableline Radius, \arcsec & $\Delta$[Mg/Fe] & $\Delta \log Z$ & $\Delta \log T$ \\ \tableline 4 & -0.09 & -0.39 & +0.26\\ 5.3 & -0.17 & -0.41 & +0.29 \\ 6.7 & -0.16 & -0.62 & +0.46 \\ 8 & -0.07 & -0.58 & +0.27 \\ \tableline \end{tabular} \end{flushleft} \end{table} The analysis undertaken in this Section has allowed to identify three quite different stellar structures within $8\arcsec$ ($\sim$ 600 pc) from the center. The unresolved star-like nucleus is rather young, $5 \cdot 10^9$ years old, strongly magnesium-overabundant and has solar global metallicity. Farther out, the circumnuclear disk with a radius of $\sim 3\arcsec$ is even younger, $\sim 2$ billion years old, has the solar Mg-to-Fe ratio, and the global metallicity higher than the solar one. The surrounding bulge is older than the nucleus and the circumnuclear disk, namely, is 9--12 billion years old, magnesium-overabundant and moderately metal-poor. \section{Kinematics of Gas and Stars} Two-dimensional line-of-sight velocity fields for stars and ionized gas obtained with the MPFS are presented in Fig.~6. Both fields look rather regular and show clear signs of rotation. However there are some differences between velocity fields of stars and gas. As we noted earlier (Afanasiev et al. \cite{asz89}), the gas rotation curve has a sharp local maximum at $R\approx 2\arcsec$, and we see a consequent feature to the north-west from the center in Fig.~6b (it is somewhat farther than it must be due to a seeing quality worse with respect to that of our long-slit observations). Meantime the rotation velocity of stars rises continuously with radius -- rapidly up to $R\approx 5\arcsec$ and more slowly farther from the center (Bower et al. \cite{betal}, Prada et al. \cite{espls}), and consistently, Fig.~6a has no closed isovelocities. Two-dimensional velocity fields can allow to check an axisymmetric character of rotation. If we have an axisymmetric mass distribution and rotation on circular orbits, the direction of maximum central line-of-sight velocity gradient (we shall call it "dynamical major axis") should coincide with the line of nodes as well as the photometric major axis; whereas in a case of triaxial potential the isovelocities align with the principal axis of the ellipsoid, and generally the dynamical and photometrical major axes diverge showing turns with respect to the line of nodes in opposite senses (e. g. \cite{mbe92}). In a simple case of cylindric (disk-like) rotation we have a convenient analytical expression for the azimuthal dependence of central line-of-sight velocity gradient within the area of solid-body rotation:\\ \noindent $dv_r/dr = \omega$ sin $i$ cos $(P.A. - P.A._0)$, \\ \noindent where $\omega$ is a deprojected central angular rotation velocity, $i$ is an inclination of the rotation plane, and $P.A._0$ is an orientation of the line of nodes. We have fitted the data presented in Fig.~6a by this formula and have obtained for the gradients taken within $r \leq 2\arcsec$:\\ \noindent $dv_r/dr$ = [33 cos $(P.A. - 183^\circ) - 2.5$] $\mbox{km} \ \mbox{s}^{-1} \ \mbox{arcsec}^{-1}$ . \\ \noindent The amplitude of the cosine curve, $\omega \, \sin i = 33 \, \mbox{km} \ \mbox{s}^{-1} \ \mbox{arcsec}^{-1}$, agrees rather well with the central slopes of the major axis long-slit cross-sections obtained for the stellar component of NGC~7331 by Bower et al. (\cite{betal}) and by Prada et al. (\cite{espls}). Moreover, it agrees rather well with the rotation velocity of ionized gas within the kinematically decoupled region: in Afanasiev et al. (\cite{asz89}) we reported an azimuthal dependence of the central line-of-sight velocity gradients based on the long-slit [NII]$\lambda$6583 emission line measurements in four different position angles, the best-fitted cosine curve formula for which was:\\ \noindent $dv_r/dr$ = [38.4 cos $(P.A. - 175^\circ) + 1.3$] $\mbox{km} \ \mbox{s}^{-1} \ \mbox{arcsec}^{-1}$ . \\ \noindent The phases of these two cosine curves are close too; but together they evidence for a marginal turn of the dynamical major axis with respect to the line of nodes which has $P.A.=166\arcdeg - 167\arcdeg$ (Prieto et al. \cite{pretal}, von Linden et al. \cite{germ}). It would be a signature of a triaxial potential, if the photometric major axis turns in opposite sense, to lesser $P.A.$. But indeed it turns in the same sense! Numerous photometric studies detected a turn of the photometric major axis in NGC~7331 to $P.A.\approx 175\arcdeg - 183\arcdeg$ inside $R=6\arcsec - 8\arcsec$ (somewhat different in different works). It is also known that this turn is stronger at shorter wavelengths. For example, Prieto et al. (\cite{pretal}) have measured the following orientations of the isophote major axis at $R\approx 4\arcsec$: $P.A._0 \approx 182\arcdeg$ in the $B$-band, $P.A._0 \approx 179\arcdeg$ in the $V$-band, and only $P.A._0 \approx 176\arcdeg$ in the $I$-band; the asymptotic $P.A._0$ value at larger radii given by them is $166\arcdeg$. Up to now the common point of view is that it is a dust effect which must be wavelength-dependent. But in the center of NGC~7331 the dust is still visible in the $I$-band, however, the measurements of $P.A._0$ in the $I$- and in the $K$-bands agree well (Prada et al. \cite{espls}). There may be another explanation: if there is a blue (young) misaligned stellar disk inside the red (older) bulge, the measurements of combined isophotes should show a stronger turn through the bluer filter. The coincidence of the dynamical and photometrical major axes at $P.A.\approx 180\arcdeg$ reveals a presence of inclined circumnuclear stellar disk, rather rapidly rotating. The long-slit observations made under higher spectral resolution may help to clarify a dynamical structure of the central region of NGC~7331. The similar observational data taken along the major axis of the galaxy in 1992 August have allowed Prada et al. (\cite{espls}) to claim a presence of counterrotating bulge in NGC~7331. Here we present the later results for the major-axis and minor-axis cross-sections. Fig.~7 shows a direct view of the LOSVD calculated along the major axis, and Figs.~8a and 8b -- the results of multi-component Gauss analysis of the LOSVDs. One can immediately see a difference between our Fig.~7 and the analogous Fig.~3 of Prada et al. (\cite{espls}). The description of the latter included only two kinematical components: fast-rotating disk and retrograde bulge rotating more slowly and seen only in the radius range of $5\arcsec - 15\arcsec$. Prada et al. (\cite{espls}) claimed an absence of prograde bulge. Meantime even a single glance at our Fig.~7 reveals two quite noticeable prograde structures: a fast-rotating disk and a slower rotating prograde bulge seen up to $\sim 35\arcsec$ from the center. The retrograde component is seen too, but it is not so prominent as it seemed to be in Fig.~3 of Prada et al. (\cite{espls}). Since our long-slit data were obtained with the same equipment -- the ISIS, red arm, of the WHT, -- and the template star is of the same spectral type as that of Prada et al. (\cite{espls}), we can only refer to our better spectral resolution -- 35 km/s (their spectral resolution was reported as 52 km/s). Fig.~8a displays the results of Gauss analysis of the LOSVD along the major axis $P.A.=172\arcdeg$ together with stellar velocity profiles derived without component separation, namely, together with the recent data of Heraudeau \& Simien (\cite{hr}) obtained under seeing of $2\farcs 5$ and with the simulated one-dimensional profile calculated from our MPFS two-dimensional velocity field shown in Fig.~6a. By analysing the long-slit data, we have extracted three kinematical components. The first, fast-rotating component dominates over the full range of radii. Since according to Prada et al. (\cite{espls}) at $R \geq 5\arcsec$ its stellar velocity dispersion does not exceed 70 km/s, we would treat it as a disk. The second component, corotating with the first one and seen at $R \geq 5\arcsec$, rotates much slower than the first one, so we would conclude that it is an ordinary bulge; but it dominates nowhere, and that is very strange for the early-type spiral galaxy with extended photometric bulge. Interestingly, its nearly solid-body slow rotation up to $R \approx 45\arcsec$ matches perfectly the rotation of the ionized gas at $R > 2\arcsec$ (see Fig.~1 in our work, Afanasiev et al. \cite{asz89}, or the lower resolution H$\alpha$ data in \cite{fr}) while the rotation curve of the dominant stellar component, disk, diverges strongly with the gas rotation curve. Finally, the third component counter-rotates with respect to the main stellar and gaseous rotation; it is just the same component which was reported by Prada et al. (\cite{espls}) as a "retrograde bulge". But there is some ambiguity with this diagnosis: in their Conclusions Prada et al. (\cite{espls}) stated that "the inner parts of the galaxy consists of a boxy component, dominating the inner $5\arcsec$. It shows position angle twisting, rotates retrograde to the rest of the galaxy, and is rounder"; meantime the retrograde component is seen only in the narrow radius range $5\arcsec < R < 20\arcsec$ according both to the data of Prada et al. (\cite{espls}) and ours! Besides, its velocity dispersion is as low as that of the disk (Prada et al. \cite{espls}), and it rotates faster than the prograde bulge though slower than the disk. So it does not resemble a bulge; it must be a rather flat structure. The central $5\arcsec$ contains only one inseparable component; but at $R \approx 5\arcsec$ it meets perfectly the disk component. As it looks like a straight solid-body rotation curve, we would identify it with a central part of the disk rotation profile. An increased velocity dispersion inside $R\approx 5\arcsec$ which has been reported by Prada et al. (\cite{espls}) is probably a result of adding slower rotating, weaker components. However, we can conclude that from a dynamical point of view the disk dominates over the whole central region of NGC~7331. \section{Discussion} We can summarize our conclusions as follows. In the Sb galaxy NGC~7331 the stellar disk is a kinematical component dominating over the full radius range. The bulge is less prominent though it can be traced up to $R\approx 45\arcsec$ (3.3 kpc); its slow solid-body rotation is very similar to the main rotation of the ionized gas. We confirm also an existence of the counter-rotating stellar component in the radius range of $5\arcsec - 20\arcsec$ (400 -- 1500 pc). The central part of the disk inside $\sim 3\arcsec$ (200 pc) -- or a separate circumnuclear stellar-gaseous disk as it is distinguished by decoupled fast rotation of the ionized gas -- is very metal-rich, rather young, $\sim 2$ billion years old, and its solar magnesium-to-iron ratio evidences for a very long duration of the last episode of star formation there. However the gas excitation mechanism in this disk now is shock-like. The star-like nucleus had probably experienced a secondary star formation burst too: its age is 5 billion years, much younger than the age of the circumnuclear bulge. But [Mg/Fe]=+0.3 and only solar global metallicity imply that the nuclear star formation burst has been much shorter than that in the circumnuclear disk. Up to now two-dimensional mapping of the absorption-line equivalent widths (or indices) is rarely used to investigate stellar population properties in galaxies. Practically unique examples of such approach are a long-slit combined study of the center of M~31 by Davidge (\cite{dav}) and a detailed TIGER investigation of NGC~4594 by Emsellem et al. (\cite{ems96}). The latter case appears to be something similar to the case of NGC~7331: the magnesium-index map of the center of NGC~4594 demonstrates a point-like peak in the nucleus of the galaxy, and the iron indices are roughly constant along the major axis up to the border of the area investigated, $R\approx 5\arcsec$ (Emsellem et al. \cite{ems96}). So the isolines of the iron indices in NGC~4594 present something like the Fe-rich circumnuclear stellar disk found by us in NGC~7331. Though the measurements of Mgb index in NGC~4594 are complicated by a presence of the rather strong emission line [NI]$\lambda$5199 (Emsellem et al. \cite{ems96}) and though two examples are not a statistics yet, perhaps, there exists some evolutionary sense in different morphologies of magnesium- and iron-index surface distributions. An interesting problem is an origin of the counter-rotating component. Usually counter-rotating stellar substructures are considered as a result of merger. Some merger event could also provoke a secondary star formation burst in the center of the galaxy in this way producing the chemically distinct nucleus and, if the merger was dissipative, the inclined circumnuclear disk. However, the circumnuclear Fe-rich disk in NGC~7331 rotates in the same sense as the rest of the galaxy and so cannot be genetically related to the counter-rotating component. Recently we have found a counter-rotating stellar component in the nearby spiral galaxy NGC~2841 (\cite{we99}). NGC~2841 and NGC~7331 look almost twins: the same morphological type, Sb, the same size and luminosity, the same inclination. They were the first galaxies where a global ring-like distribution of CO has been detected (\cite{ys}). Analysing the major-axis long-slit cross-section of NGC~2841, we have found two kinematical components in the bulge: strong prograde one which dominates up to $R\approx 25\arcsec$ and weak retrograde one (\cite{we99}). A set of other phenomena allow us to suggest an existence of extended triaxial bulge almost aligned with the line of nodes of the global disk in this galaxy, and we have thought the counter-rotating stellar component to be an intrinsic property of a slightly tumbling triaxial potential. Perhaps the similar situation takes place in NGC~7331; the only difference is that in NGC~7331 the disk dominates over the full radius range whereas in NGC~2841 the bulge is more prominent. A hypothesis of the bar presence in the center of NGC~7331 has already been proposed by von Linden et al. (\cite{germ}) to explain a central depletion of molecular gas in this galaxy. Besides, the French team (\cite{fr}) have constructed a global two-dimensional velocity field of the ionized gas in NGC~7331 from their observations of H$\alpha$ emission line with a scanning Fabry-Perot interferometer and have found a strong large-scale turn of the isovelocities which is usually treated as a signature of bar presence. Now we can add another argument in favor of a triaxial potential aligned with the line of nodes: inside $R\approx 40\arcsec$ the gas in NGC~7331 rotates more slowly than the stellar disk, and it is valid both for the ionized gas (\cite{retal65}, Afanasiev et al. \cite{asz89}, \cite{fr}) and for the molecular gas (von Linden et al. \cite{germ}). Since the emission lines are narrow (below the spectral resolution) and the gas seems to be well-settled to the global plane of the galaxy (Bosma \cite{bosma}, von Linden et al. \cite{germ}), the only explanation can be non-circular rotation of the gas caused by the triaxial potential of the bulge. The stronger response of the gaseous disk to the triaxial potential of the bulge when compared to the response of the stellar disk may be explained by a viscous nature of the gas and by a significant self-gravitation of the massive stellar disk of NGC~7331 which is a dominant dynamical component in this galaxy. Interestingly, the solid-body part of the gas rotation curve, or a zone of non-circular motions as we think, ends at $R=40\arcsec - 45\arcsec$ -- exactly at the ring of molecular gas, warm dust, and intense star formation which we have discussed in Section~3. The configuration looks like an HII ring around a bar which is often observed in classic barred galaxies. Probably, the low-contrast triaxial bulge extends up to $40\arcsec \div 50\arcsec$ from the center in NGC~7331. The last interesting question concerns a possible presence of the black hole in the center of NGC~7331. In Afanasiev et al. (\cite{asz89}) we argued that the fast decoupled axisymmetric rotation of the ionized gas inside $R=2\arcsec$ proved strong mass concentration in the nucleus; this mass concentration might be a black hole. But Bower et al. (\cite{betal}) have noticed no decoupled rotation by studying a stellar component; and what is the most important, they have not found a peak of stellar velocity dispersion in the nucleus. How can we agree gas and star behaviours? In Afanasiev et al. (\cite{asz89}), Zasov and Sil'chenko (\cite{zs96}) we noted that a misaligned minibar in the center of a galaxy may mimic a presence of kinematically decoupled nucleus by increasing a visible central slope of the major-axis velocity profile due to non-circular motions. But in NGC~7331 the situation is more complex: the rotation of the gas and stars in the very center is circular, and the bar effect (non-circular gas motions) is felt only outside the zone of decoupled rotation. However, the result is the same: the nucleus looks kinematically decoupled though no point-like central mass concentration is required for this. By summarizing, since Cowan et al. (\cite{crb}) and Stockdale et al. (\cite{src}) have found the unresolved radio- and X-ray source in the center of NGC~7331, it may be a black hole, but it cannot be a {\it supermassive} black hole, like those in NGC~4594, NGC~3115, and M~87, because its dynamical influence is nowadays indeterminate. As supermassive black holes are now detected in some three dozens galaxies, a correlation is found between black hole mass and galactic spheroid luminosity. By using this relation taken e. g. from Cattaneo et al. (\cite{bhmass}) one can try to estimate a possible black hole mass in the center of NGC~7331. Unfortunately, as we have noted in the Introduction, the bulge-disk decomposition in NGC~7331 is ambiguous: if the bulge is large as Boroson (\cite{bor}) or Baggett et al. (\cite{bag}) reported, the black hole mass may be $10^9\,M_\odot$, if it is small like that reported by Prada et al. (\cite{espls}), the mean relation $M_{BH}$ vs. $L_{B,bul}$ implies $M_{BH}=10^8\,M_\odot$ for NGC~7331. If we take also into account the large scatter of this relation (Cattaneo et al., \cite{bhmass}, give $\sigma = 0.74$) and the impression that the $M_{BH}$ estimates for spiral galaxies lie all below the mean relation defined mostly by ellipticals, the mass of the black hole in the center of NGC~7331 may be as small as $3 \cdot 10^7\,M_\odot$. This value contributes only several percents into the total mass of the circumnuclear disk ($5 \cdot 10^8\,M_\odot$, Afanasiev et al., \cite{asz89}), so it may be undetectable in kinematical studies of moderate spatial resolution like ours. \acknowledgements I thank the astronomers of the Special Astrophysical Observatory Drs. V. L. Afanasiev, S. N. Dodonov, V. V. Vlasyuk, and Mr. S. V. Drabek for supporting the observations at the 6m telescope. I am also grateful to the graduate student of the Moscow University A. V. Moiseev for the help in preparing the manuscript. The 6m telescope is operated under the financial support of Science Department of Russia (registration number 01-43). During the data analysis I have used the Lyon-Meudon Extragalactic Database (LEDA) supplied by the LEDA team at the CRAL-Observatoire de Lyon (France) and the NASA/IPAC Extragalactic Database (NED) which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. This research has made use of the La Palma Archive. The telescope WHT is operated on the island of La Palma by the Royal Greenwich Observatory in the Spanish Observatorio del Roque de los Muchachos of the Instituto de Astrofisica de Canarias. The work was supported by the grant of the Russian Foundation for Basic Researches 98-02-16196, by the grant of the President of Russian Federation for young Russian doctors of sciences 98-15-96029 and by the Russian State Scientific-Technical Program "Astronomy. Basic Space Researches" (the section "Astronomy").
\section{Introduction} On general grounds (2+1)-dimensional spacetime was long considered unlikely to support black holes, before such solutions were discovered \cite{BTZ}. Black holes were commonly conceived as places where the effects of gravity are large, surrounded by a region where these effects are asymptotically negligible. Another possible reason is the idea that black holes are ``frozen gravitational waves" and therefore exist only in a context where the gravitational field can have independent degrees of freedom. In 2+1 dimensional Einstein theory --- that is, Einstein's equations in a 3-dimensional space-time of signature $(- + +)$ --- the pure, sourceless gravitational field has no local degrees of freedom, because in three dimensions the Riemann tensor is given algebraically by the Einstein tensor, which in turn is algebraically determined by the Einstein field equations. If there is no matter source and no cosmological constant, the Riemann tensor vanishes and space-time is flat; if there is no matter but a cosmological constant $\Lambda$, the Riemann tensor is that of a space of constant curvature $\Lambda/3$. Thus gravity does not vary from place to place and it does not have any wave degrees of freedom. These were some of the reasons why the possibility of black holes was discounted, and the discovery of black hole solutions in 2+1 D spacetimes with a negative $\Lambda$ came as such a surprise. The existence of (2+1)-dimensional black holes of course does not alter the absence of gravitational waves in (2+1)-dimensional Einstein spaces, nor the lack of variation of their curvature. The curvature of spacetimes satisfying the sourceless Einstein equations with negative $\Lambda$ is constant negative, and the local geometry in the asymptotic region does not differ from that near the black hole. Indeed, black hole solutions can be obtained from the standard, simply connected spacetime of constant negative curvature (anti-de Sitter space, AdS space for short) by forming its quotient space with a suitable group of isometries.\footnote{It appears that all locally AdS spacetimes can be obtained in this way \cite{mess}. This is not so for positive curvature \cite{MJW}.} One of the criteria on the isometries is that the quotient space should not have any objectionable singularities. For example, if the group contains isometries of the rotation type, with a timelike set of fixed points, then the quotient space will have singularities of the conical kind. Such singularities can represent ``point" particles, and the corresponding spacetime can be interpreted as an interesting and physically meaningful description of the dynamics of such particles \cite{Hoo}. However, we confine attention to solutions of the sourceless Einstein equations with negative cosmological constant --- whether black holes or not --- that are at least initially nonsingular. Therefore we exclude such particle-like solutions. (Likewise, we will not consider the interesting developments in lower-dimensional dilaton gravity, nor other matter fields \cite{LD}.) On the other hand, the group used to construct our quotient space may have isometries that are locally Lorentz boosts, with spacelike sets of fixed points. The corresponding singularities are of the non-Hausdorff ``Misner" type \cite{Mis}. If such a singularity does not occur on an initial spacelike surface, and is hidden behind an event horizon, then the spacetime can be acceptable as a representation of a black hole. Finally, the isometry may not have any fixed points but still lead to regions in the quotient space that are to be considered singular for physical reasons, and such regions may again be surrounded by an event horizon, yielding other types of black holes. Thus the proper criterion characterizing a black hole in this context is not a region of large curvature or an infinite red shift (in typical representations of AdS space itself, where there is no black hole, there is an infinite red shift between the interior and the region near infinity), but existence of an event horizon. This in turn requires the existence of a suitable {\curs I}, whose neighborhood is a region in which ``distant observers" can survive for an arbitrarily long time without hitting a singularity. That is, there have to be causal curves (the worldlines of these observers) that can be continued to infinite proper time. For example, Misner space itself --- the quotient of Minkowski space by a Lorentz boost --- does not satisfy this criterion in any dimension, because all timelike curves intersect the non-Hausdorff singularity in a finite proper time. Thus the case $\Lambda = 0$ does not yield any black holes. The same is true, for similar reason, in the case $\Lambda > 0$. However, for $\Lambda < 0$ there are worldlines along which asymptotic observers can survive forever even when spacelike singularities are present. Our black holes will then not be asymptotically flat \cite{Heu}, but asymptotically AdS. We will see (in section 3) that the usual definition of black holes can be applied to these spacetimes, and even before we have come to this we will speak of them as black holes. We can understand the difference between the cases $\Lambda \geq 0$ and $\Lambda < 0$ as as a consequence of the positive ``relative acceleration" of spacelike geodesics in spaces of negative curvature. Spacelike geodesics reaching the asymptotically AdS region will increase their separation without limit. The fixed points of the identification that generates a black hole --- that is, the ``singularities" --- lie along a spacelike geodesic. Consider a set of observers located initially further and further towards the asymptotic region and along another spacelike geodesic, which does not intersect the geodesic of fixed points. The timelike distance of an observer from the singularity will then eventually increase without limit, so a sufficiently far-out observer can survive for an arbitrarily long time. We note in passing that {\em timelike} geodesics in spacetimes of constant negative curvature have the opposite property: they accelerate toward each other. Thus $\Lambda < 0$ corresponds to a universal ``attractive" gravity, and a black hole in such a spacetime exerts this same attraction on test particles, as a black hole should. The quotient of the AdS universe with the group generated by a single finite isometry that is without fixed points, at least on some initial spacelike surface, yields a single black hole, called a BTZ spacetime (for its discoverers, Ba\~nados, Teitelboim and Zanelli \cite{BTZ}). As we will see, one can make further identifications in a BTZ spacetime, obtaining more complicated black holes, and this process can be repeated an arbitrary number of times. Although the isometries used for the identification cannot be entirely arbitrary, the variety of possibilities and of the resulting spacetimes is quite large. These spacetimes cannot be described by their metric in one or in a few simple coordinate systems, because many coordinate patches would be needed to cover their possibly complicated topology. In principle such a spacetime is of course defined, and all its physical properties are computable, once we know the structure of the AdS isometries that generate it. But such a presentation does not give an accessible and easily visualizable picture of the spacetime. Therefore we prefer to describe the spacetimes combinatorially, by ``gluing together" pieces of AdS space. This view allows one to gain many important geometrical insights directly, without much algebra or analysis (even if a few of these geometrical constructions may resemble a tour de force). In section 2 we consider the simplest, time-symmetric case. Because the extrinsic curvature of the surface of time-symmetry vanishes, this surface is itself a smooth two-dimensional Riemannian space of constant negative curvature. This class of spaces has been studied in considerable detail \cite{BP}. In particular, almost all two-dimensional spacelike topologies occur already within this class. Section 3 considers the time development of these spaces; we find that all the non-compact initial states develop into black holes. The horizon can be found explicitly, although its behavior can be quite complicated. Section 4 concerns spacetimes that are not time-symmetric but have angular momentum. An important reason for studying the classical behavior of these spacetimes is their relative simplicity while still preserving many of the features of more realistic black hole spacetimes. They are therefore interesting models for testing the formalism of quantum gravity. We do not go into these developments but refer the reader to the recent book by S. Carlip \cite{Carl}. \section{Time-symmetric geometries} Three-dimensional AdS space has many totally geodesic (``time-symmetric") spacelike surfaces. Because the extrinsic curvature of such surfaces vanishes, they have constant negative curvature $\Lambda$. Each such surface remains invariant under a ``little group" of AdS isometries, which are therefore isometries of the spacelike surface, and conversely each isometry of the spacelike surface can be extended to be an isometry of the whole AdS spacetime.\footnote{Since AdS spacetime is an analytic continuation (both in signature and curvature) of the familiar spherical geometry, such properties can be considered extensions of the corresponding statements about spheres, mutatis mutandis for the difference in group structure, SO(4) vs SO(2,2). Analogous statements are true about surfaces of constant extrinsic curvature.} Therefore any identification obtained by isometries on the spacelike surface can likewise be extended to the whole spacetime. (AdS space identified by this extension coincides with the usual time development of the initial data via Einstein's equations where the latter is defined, but it even goes beyond any Cauchy horizon). Thus to identify the possible time-symmetric geometries it suffices to discuss the possible initial spacelike geometries --- although this leaves the time development still to be made explicit. \subsection{Coordinates} Although most physically and mathematically interesting facts about constant negative curvature spaces can be phrased without reference to coordinates, and even usefully so, it is convenient for the elucidation and proof of these facts to have coordinates available. Because of the large number of symmetries of AdS spacetime, its geometry takes a simple form in a large number of coordinate systems, which do not usually cover all of the spacetime, but which exhibit explicitly one or several of these symmetries. The simplest coordinates are the redundant set of four $X^\mu$, $\mu = 1,\dots 4$ in terms of which AdS space is usually defined, namely as an embedding in four-dimensional flat space with signature $(-,-,+,+)$ and metric \begin{eqnarray} ds^2=-dU^2-dV^2+dX^2+dY^ \label{emme} \end{eqnarray} by the surface \begin{eqnarray} -U^2-V^2+X^2+Y^2=-\ell^2 . \label{ads} \end{eqnarray} This spacetime is periodic in the timelike direction with the topology $S^1\times R^2$; for example, for $X^2+Y^2<\ell^2$ the curves ($X,\,Y$) = const, $ U^2+V^2 = \ell^2 - X^2-Y^2$ are closed timelike circles. In the following we assume that this periodicity has been removed by passing to the universal covering space with topology $R^3$, which we will call AdS space. If it is necessary to distinguish it from the space of Eq (\ref{ads}) we will call the latter ``periodic AdS space." Either spacetime is a solution of the vacuum Einstein equations with a negative cosmological constant $\Lambda=-1/\ell^2$. Eq (\ref{ads}) shows that AdS space is a surface of constant distance from the origin in the metric (\ref{emme}). It therefore inherits from the embedding space all the isometries that leave the origin fixed, which form the SO(2,2) group. AdS space can be described by coordinates analogous to the usual spherical polar coordinates as in Eq (\ref{ts}), but of greater interest are coordinates related to isometries that leave a plane fixed, and whose orbits lie in the orthogonal plane. These have the nature of rotations if the plane is spacelike (or double-timelike, such as the ($U,\,V$) plane), and of Lorentz transformations if the plane is timelike. Isometries corresponding to orthogonal planes commute, and we can find coordinates that exhibit such pairs of isometries explicitly. If the isometries are rotations, the coordinates cover all of AdS space; if they are Lorentz transformations the corresponding coordinates are analogous to Rindler coordinates of flat space, and need to be analytically extended in the usual fashion to cover all of the spacetime. For example, if we choose rotations by an angle $\theta$ in the ($X,\,Y$) plane and by an angle $t/\ell$ in the ($U,\,V$) plane, and specify the respective orbits on the AdS surface by $$U^2+V^2=\ell^2\cosh^2\chi \qquad {\rm and} \qquad X^2+Y^2 = \ell^2\sinh^2\chi$$ (so that, for example, $U=-\ell\cosh\chi\cos{t\over\ell},\; V=\ell\cosh\chi\sin{t\over\ell}$) we obtain the metric \begin{eqnarray} ds^2=-\cosh^2\chi dt^2+\ell^2\left( d\chi^2+\sinh^2\chi d\theta^2\right) . \label{statmet} \end{eqnarray} In order to describe the universal covering space we have to allow $t$ to range $-\infty < t < \infty$, whereas $\theta$ has its usual range, $0\leq\theta<2\pi$, and similarly $0\leq\chi<\infty$. Except for the usual polar coordinate singularity at $\chi = 0$, these coordinates cover all of AdS space by a sequence of identical (``static") two-dimensional spacelike surfaces $t =$ const having a standard metric of spaces of constant negative curvature $-1/\ell^2$. Because $U=0=V$ does not occur on (\ref{ads}), shifts in the $t$ coordinate are true translations, without fixed points. These coordinates define timelike sections ($\theta=$ const) and spacelike sections ($t=$ const) of AdS space. Each of these can be represented in a conformal diagram, shown in Fig.~1. \begin{figure} \unitlength 0.8mm \linethickness{0.4pt} \begin{picture}(140.20,86.00)(10,5) \bezier{184}(99.80,50.00)(99.80,61.28)(109.90,67.68) \bezier{180}(109.90,67.68)(120.00,72.73)(130.10,67.68) \bezier{184}(130.10,67.68)(140.20,61.28)(140.20,50.00) \bezier{184}(99.80,50.00)(99.80,38.72)(109.90,32.32) \bezier{180}(109.90,32.32)(120.00,27.27)(130.10,32.32) \bezier{184}(130.10,32.32)(140.20,38.72)(140.20,50.00) \put(20.00,15.00){\line(0,1){70.00}} \put(40.00,15.00){\line(0,1){70.00}} \put(60.00,15.00){\line(0,1){70.00}} \put(80.00,50.00){\makebox(0,0)[cc]{$t=0$}} \put(85.00,50.00){\vector(1,0){11.00}} \put(75.00,50.00){\vector(-1,0){11.00}} \put(20.00,50.00){\line(1,0){40.00}} \put(20.00,60.00){\line(1,0){40.00}} \put(20.00,70.00){\line(1,0){40.00}} \put(20.00,80.00){\line(1,0){40.00}} \put(20.00,40.00){\line(1,0){40.00}} \put(20.00,30.00){\line(1,0){40.00}} \put(20.00,20.00){\line(1,0){40.00}} \put(48.00,82.00){\line(0,-1){64.00}} \put(54.00,82.00){\line(0,-1){64.00}} \put(57.00,82.00){\line(0,-1){64.00}} \put(32.00,82.00){\line(0,-1){64.00}} \put(26.00,82.00){\line(0,-1){64.00}} \put(23.00,82.00){\line(0,-1){64.00}} \bezier{184}(103.38,50.00)(103.38,59.28)(111.69,64.54) \bezier{180}(111.69,64.54)(120.00,68.70)(128.31,64.54) \bezier{184}(128.31,64.54)(136.62,59.28)(136.62,50.00) \bezier{184}(103.38,50.00)(103.38,40.72)(111.69,35.46) \bezier{180}(111.69,35.46)(120.00,31.30)(128.31,35.46) \bezier{184}(128.31,35.46)(136.62,40.72)(136.62,50.00) \bezier{184}(106.33,50.00)(106.33,57.63)(113.16,61.97) \bezier{180}(113.16,61.97)(120.00,65.38)(126.84,61.97) \bezier{184}(126.84,61.97)(133.67,57.63)(133.67,50.00) \bezier{184}(106.33,50.00)(106.33,42.37)(113.16,38.03) \bezier{180}(113.16,38.03)(120.00,34.62)(126.84,38.03) \bezier{184}(126.84,38.03)(133.67,42.37)(133.67,50.00) \bezier{184}(112.01,50.00)(112.01,54.46)(116.00,57.00) \bezier{180}(116.00,57.00)(120.00,58.99)(124.00,57.00) \bezier{184}(124.00,57.00)(127.99,54.46)(127.99,50.00) \bezier{184}(112.01,50.00)(112.01,45.54)(116.00,43.00) \bezier{180}(116.00,43.00)(120.00,41.01)(124.00,43.00) \bezier{184}(124.00,43.00)(127.99,45.54)(127.99,50.00) \put(100.00,50.00){\line(1,0){40.00}} \put(120.00,30.00){\line(0,1){40.00}} \put(138.00,61.67){\makebox(0,0)[lb]{$\theta$}} \put(119.33,49.67){\vector(3,1){0.2}} \bezier{100}(99.33,37.00)(107.33,46.00)(119.33,49.67) \put(99.00,36.33){\makebox(0,0)[ct]{$\chi = 0$}} \put(105.67,35.67){\line(1,1){29.00}} \put(134.33,35.67){\line(-1,1){29.00}} \put(140.00,42.00){\makebox(0,0)[lt]{$\chi=\infty$}} \put(40.00,86.00){\makebox(0,0)[cb]{$\chi=0$}} \put(60.00,86.00){\makebox(0,0)[cb]{$\chi=\infty$}} \put(61.00,70.00){\makebox(0,0)[lc]{$t=$ const}} \put(40.00,10.00){\makebox(0,0)[cc]{({\bf a})}} \put(120.00,10.00){\makebox(0,0)[cc]{({\bf b})}} \thicklines \put(135.67,62.67){\vector(-1,1){0.20}} \end{picture} \caption{Conformal diagrams of the static (or sausage) coordinates of Eq (\ref{statmet}) in sections of AdS space. ({\bf a}) The $\chi,\,t$ section, both sides of the origin. The right half is, for example, $\theta = 0$, and the left half, $\theta=\pi$. ({\bf b}) The section $t=$ const is the 2D space of constant negative curvature, conformally represented as a Poincar\'e disk (see section 2.2). The conformal factors are different in the two sections, so they do not represent sections of one three-dimensional conformal diagram. (For the latter see Fig.~4b)} \label{fig1} \end{figure} We can define a ``radial" coordinate (which really measures the circumference of circles) by $$ r = \ell\sinh\chi\,.$$ The metric (\ref{statmet}) then takes the form \begin{equation} ds^2=-\left({r^2\over\ell^2}+1\right) dt^2+ \left({r^2\over\ell^2}+1\right)^{-1} dr^2+r^2 d\theta^2\,. \label{m0} \end{equation} By choosing a different radial coordinate, namely $$\rho = \ell\tanh{1\over 2}\chi$$ to replace the $\chi$ of Eq (\ref{statmet}), we can make the conformally flat nature of the spacelike section explicit and keep the metric static: \begin{equation} ds^2=-\left({1+(\rho/\ell)^2\over 1-(\rho/\ell)^2}\right)^2dt^2 + {4\over\left(1-(\rho/\ell)^2\right)^2}\left(d\rho^2 + \rho^2d\theta^2\right). \label{sausge} \end{equation} A picture like Fig.~1, with parts ({\bf a}) and ({\bf b}) put together into a 3-dimensional cylinder, can be considered a plot of AdS space in the cylindrical coordinates of Eq (\ref{sausge}). Because of the cylindrical shape of this diagram these coordinates are sometimes called {\it sausage coordinates} \cite{Ingemar}. Like the static coordinates of (\ref{statmet}), these cover all of AdS space. If we follow an analogous construction but use the timelike ($X,\, U$) and ($Y,\, V$) planes with orbits (in terms of a new coordinate $\chi$) $$-V^2+X^2=-\ell^2\cosh^2\chi \qquad {\rm and} \qquad -U^2+Y^2 = \ell^2\sinh^2\chi ,$$ and new hyperbolic coordinates $\phi$ and $t/\ell$, we obtain the metric \begin{equation} ds^2 = -\sinh^2\chi dt^2 +\ell^2\left(d\chi^2+\cosh^2\chi d\phi^2\right). \end{equation} By defining $$r = \ell\cosh\chi$$ we can change this to the Schwarzschild-coordinate form \begin{eqnarray} ds^2=-\left({r^2\over\ell^2}-1\right) dt^2+ \left({r^2\over\ell^2}-1\right)^{-1} dr^2+ r^2 d\phi^2, \label{schwmet} \end{eqnarray} which is usually derived from the ``rotationally" symmetric ansatz --- however, in this description of AdS space, $\phi$ has to be given the full range, $-\infty < \phi < \infty$ of a hyperbolic angle. The range of $r$ for which the metric (\ref{schwmet}) is regular, $\ell<r<\infty$, describes only a part of AdS space, as can be seen from the explicit expression for the embedding in terms of these coordinates, \begin{eqnarray} \left\{\begin{array}{l} U=\left(r^2-\ell^2\right)^{1/2}\sinh{t\over\ell}\\ V=r\cosh\phi\\ X=r\sinh\phi\\ Y=\left(r^2-\ell^2\right)^{1/2}\cosh{t\over\ell}\\ \end{array} \right. \label{schwc} \end{eqnarray} This regular region can be patched together in the usual way with the region $0<r<\ell$ (Fig.~2), to describe a larger part of AdS space. But if it is desired (for whatever bizarre reason) to describe all of AdS space by analytic extensions of the coordinates (\ref{schwc}), one needs also analytic extensions beyond the null surfaces $\phi=\pm\infty$ (or $r=0$), which are quite analogous to the usual Schwarzschild-type ``horizon" null surfaces $t=\pm\infty$ (or $r=\ell$). One then finds two disjoint regions of a third type (not shown in the Figure because they extend perpendicular to the plane of Fig.~2a) in which $r^2$ is negative and $\phi$ is the timelike coordinate.\footnote{Like all statements derived from embedding equations such as (\ref{schwc}) this really applies to periodic AdS space, and should be repeated an infinite number of times for the covering AdS space itself. For example, there are an infinite number of regions of the three types in AdS space.} \begin{figure} \unitlength 0.80mm \linethickness{0.4pt} \begin{picture}(131.00,90.00) \bezier{184}(89.80,50.32)(89.80,61.60)(99.90,68.00) \bezier{180}(99.90,68.00)(110.00,73.05)(120.10,68.00) \bezier{184}(120.10,68.00)(130.20,61.60)(130.20,50.32) \bezier{184}(89.80,50.32)(89.80,39.05)(99.90,32.64) \bezier{180}(99.90,32.64)(110.00,27.59)(120.10,32.64) \bezier{184}(120.10,32.64)(130.20,39.05)(130.20,50.32) \put(10.00,10.00){\line(0,1){80.00}} \put(50.00,10.00){\line(0,1){80.00}} \put(10.00,50.00){\line(1,0){40.00}} \put(10.00,30.00){\line(1,1){40.00}} \put(50.00,30.00){\line(-1,1){40.00}} \bezier{88}(10.00,42.00)(18.00,42.00)(30.00,50.00) \bezier{88}(50.00,58.00)(42.00,58.00)(30.00,50.00) \bezier{88}(10.00,58.00)(18.00,58.00)(30.00,50.00) \bezier{88}(50.00,42.00)(42.00,42.00)(30.00,50.00) \put(10.00,70.00){\line(1,0){40.00}} \bezier{204}(50.00,70.00)(34.00,50.00)(50.00,30.00) \bezier{204}(10.00,70.00)(26.00,50.00)(10.00,30.00) \bezier{204}(50.00,70.00)(30.00,54.00)(10.00,70.00) \put(30.00,5.00){\makebox(0,0)[cc]{({\bf a})}} \put(90.00,50.00){\line(1,0){40.00}} \bezier{152}(95.00,64.00)(110.00,52.00)(125.00,64.00) \bezier{40}(102.00,69.00)(104.00,64.00)(110.00,64.00) \bezier{40}(118.00,69.00)(116.00,64.00)(110.00,64.00) \bezier{152}(95.00,37.00)(110.00,49.00)(125.00,37.00) \bezier{40}(102.00,32.00)(104.00,37.00)(110.00,37.00) \bezier{40}(118.00,32.00)(116.00,37.00)(110.00,37.00) \put(10.00,30.00){\line(1,0){40.00}} \bezier{204}(50.00,30.00)(30.00,46.00)(10.00,30.00) \bezier{216}(110.00,30.00)(92.00,50.00)(110.00,70.50) \bezier{120}(110.00,30.00)(95.00,36.00)(95.00,50.00) \bezier{120}(110.00,70.50)(95.00,64.00)(95.00,50.00) \bezier{216}(110.00,30.00)(128.00,50.00)(110.00,70.50) \bezier{120}(110.00,30.00)(125.00,36.00)(125.00,50.00) \bezier{120}(110.00,70.50)(125.00,64.00)(125.00,50.00) \put(70.00,50.00){\makebox(0,0)[cc]{$t=0$}} \put(75.00,50.00){\vector(1,0){12.00}} \put(65.00,50.00){\vector(-1,0){12.00}} \put(51.00,58.00){\makebox(0,0)[lc]{$t=$ const}} \put(30.00,70.33){\makebox(0,0)[cb]{$r=0$}} \put(56.00,75.00){\makebox(0,0)[lc]{$r=$ const}} \put(39.00,64.00){\vector(-1,-2){0.2}} \bezier{92}(55.00,75.00)(43.00,74.33)(39.00,64.00) \put(44.00,60.00){\vector(-3,-1){0.2}} \bezier{80}(56.00,73.00)(54.00,63.33)(44.00,60.00) \put(110.00,55.00){\makebox(0,0)[cc]{$r$=$\ell$}} \put(131.00,50.00){\makebox(0,0)[lc]{$\phi=0$}} \put(129.00,62.00){\makebox(0,0)[lc]{$\phi$}} \put(110.00,5.00){\makebox(0,0)[cc]{({\bf b})}} \put(110.00,30.00){\line(0,1){23.00}} \put(110.00,56.00){\line(0,1){14.50}} \put(110.00,71.00){\makebox(0,0)[cb]{$\phi=\infty$}} \put(110.00,29.00){\makebox(0,0)[cc]{$\phi=-\infty$}} \thicklines \put(127.00,62.00){\vector(-1,2){0.20}} \bezier{10}(10.00,50.00)(20.00,60.00)(30.00,70.00) \bezier{10}(30.00,70.00)(40.00,60.00)(50.00,50.00) \bezier{10}(50.00,50.00)(40.00,40.00)(30.00,30.00) \bezier{10}(30.00,30.00)(20.00,40.00)(10.00,50.00) \bezier{20}(10.00,50.00)(30.00,65.00)(50.00,50.00) \bezier{20}(50.00,50.00)(30.00,35.00)(10.00,50.00) \end{picture} \caption{Conformal diagrams of the ``Schwarzschild" coordinates of Eq (\ref{schwc}) in sections of AdS space. ({\bf a})~An $r,\,t$ section, continued across the $r=\ell$ coordinate singularity. The outer vertical lines correspond to $r=\infty$. The dotted curves show a few of the surfaces $\tau=$ const for the coordinates of Eq (\ref{cmc}), with limits at $\tau = \pm\pi\ell/2$. ({\bf b})~An $r,\,\phi$ section ($r>\ell$) is a two dimensional space of constant negative curvature, conformally represented as a Poincar\'e disk (see below). The approximately vertical curves are lines of constant $r$; they are equidistant in the hyperbolic metric. The approximately horizontal curves are lines of constant $\phi$; they are geodesics in the hyperbolic metric. The outer circle corresponds to $r=\infty$} \end{figure} Another interesting coordinate system is closely related to ordinary polar coordinates on the three-sphere: \begin{eqnarray} \left\{\begin{array}{l} U=\ell\sin{\tau\over\ell}\\ V=\left(r^2-\ell^2\right)^{1/2}\cos{\tau\over\ell}\\ X=r\cos{\tau\over\ell} \cos\phi \\ Y=r\cos{\tau\over\ell} \sin\phi\\ \end{array} \right. \label{ts} \end{eqnarray} with the metric \begin{equation} ds^2=-d\tau^2 + \cos^2\left({\tau\over\ell}\right) \left[\left({r^2\over\ell^2}-1\right)^{-1}dr^2+ r^2 d\phi^2\right]. \label{cmc} \end{equation} This is a time development of the same initial data as in (\ref{schwmet}) (at $t=0$ resp.\ $\tau=0$) but with unit lapse function $N=1$. The surfaces $\tau=$ const have constant extrinsic curvature, and they just cover the domain of dependence of those initial values. Finally one can introduce coordinates that correspond to the flat sections of de Sitter space: \begin{eqnarray} \left\{\begin{array}{l} U+Y=r\\ U-Y=r\left(\phi^2-t^2\right)+{1\over r}\\ X=r\phi\\ V=rt\\ \end{array} \right. \label{horo} \end{eqnarray} The metric then takes the form \begin{equation} ds^2=-r^2dt^2+{dr^2\over r^2}+r^2d\phi^2. \label{horom} \end{equation} Here the $r=$ const sections are manifestly flat.\footnote{These subspaces are the analog in the case of Lorentzian metrics of {\it horospheres} of hyperbolic spaces (see, for example, \cite{BP}).} Fig.~3 shows the conformal picture of these coordinates. \begin{figure} \unitlength 0.80mm \linethickness{0.4pt} \begin{picture}(131.00,90.00)(5,0) \bezier{184}(89.80,50.32)(89.80,61.60)(99.90,68.00) \bezier{180}(99.90,68.00)(110.00,73.05)(120.10,68.00) \bezier{184}(120.10,68.00)(130.20,61.60)(130.20,50.32) \bezier{184}(89.80,50.32)(89.80,39.05)(99.90,32.64) \bezier{180}(99.90,32.64)(110.00,27.59)(120.10,32.64) \bezier{184}(120.10,32.64)(130.20,39.05)(130.20,50.32) \put(10.00,10.00){\line(0,1){80.00}} \put(50.00,10.00){\line(0,1){80.00}} \put(30.00,5.00){\makebox(0,0)[cc]{({\bf a})}} \put(70.00,50.00){\makebox(0,0)[cc]{$t=0$}} \put(75.00,50.00){\vector(1,0){12.00}} \put(65.00,50.00){\vector(-1,0){12.00}} \put(131.00,50.00){\makebox(0,0)[lc]{$\phi=0$}} \put(127.00,61.50){\vector(-2,3){1.00}} \put(127.67,63.00){\makebox(0,0)[lc]{$\phi$}} \put(110.00,5.00){\makebox(0,0)[cc]{({\bf b})}} \put(10.00,50.00){\line(1,1){40.00}} \put(10.00,50.00){\line(1,-1){40.00}} \put(10.00,50.00){\line(1,0){40.00}} \put(90.00,50.00){\line(1,0){40.00}} \bezier{144}(90.00,50.00)(108.00,52.00)(110.00,70.00) \bezier{144}(90.00,50.00)(108.00,48.00)(110.00,30.00) \bezier{160}(90.00,50.00)(107.00,49.00)(124.00,64.00) \bezier{160}(90.00,50.00)(107.00,51.00)(124.00,36.00) \bezier{120}(89.66,50.22)(89.66,57.85)(96.49,62.18) \bezier{120}(96.49,62.18)(103.33,65.60)(110.17,62.18) \bezier{120}(110.17,62.18)(117.00,57.85)(117.00,50.22) \bezier{120}(89.66,50.22)(89.66,42.59)(96.49,38.25) \bezier{120}(96.49,38.25)(103.33,34.83)(110.17,38.25) \bezier{120}(110.17,38.25)(117.00,42.59)(117.00,50.22) \put(96.00,50.00){\circle{12.00}} \bezier{380}(50.00,90.00)(24.00,50.00)(50.00,10.00) \bezier{204}(22.00,50.00)(22.00,58.00)(50.00,90.00) \bezier{204}(22.00,50.00)(22.00,42.00)(50.00,10.00) \bezier{180}(10.00,50.00)(36.00,66.00)(50.00,66.00) \bezier{180}(10.00,50.00)(36.00,34.00)(50.00,34.00) \put(110.00,71.00){\makebox(0,0)[lb]{$\phi=$const}} \put(70.00,29.00){\makebox(0,0)[cc]{$r=$ const}} \put(50.50,65.00){\makebox(0,0)[lb]{$t=$ const}} \put(82.00,29.00){\vector(2,1){16.00}} \put(59.00,28.00){\vector(-3,-1){15.00}} \put(22.00,62.00){\makebox(0,0)[cc]{$r$=0}} \end{picture} \caption{Conformal diagram of the ``extremal" Schwarzschild coordinates of Eq (\ref{horo}) in sections of AdS space. ({\bf a}) An $r,\,t$ section. ({\bf b}) An $r,\,\phi$ section. The lines $r=$ const are {\it horocycles} of the Poincar\'e disk.} \label{fig3} \end{figure} The spacelike surfaces $t=$ const are conformally flat as are all two-dimensional surfaces, and as is manifest in Eq (\ref{sausge}). Less trivially, the three-dimensional AdS spacetime also has this property, so neighborhoods of AdS space can be conformally mapped to flat space (one of the few cases where a three-dimensional conformal diagram exists). Such a map is the ``stereographic" projection, a projection by straight lines in the embedding space from a point in the surface of Eq (\ref{ads}) onto a plane tangent to that surface at the antipodal point, analogous to the familiar stereographic projection of a sphere (Fig.~4a). By projection from the point $(U,\,V,\,X,\,Y) = (-\ell,\,0,\,0,\,0)$ to the plane $U=\ell$ we obtain the coordinates (provided $U > -\ell$) \begin{equation} x^\mu = {2\ell X^\mu\over U+\ell} \qquad X^\mu \neq U \label{stc} \end{equation} with the metric (where $X^0=V,\,x^0=t$) \begin{equation} ds^2=\left({1\over 1-r_c^2}\right)^2\left(-dt^2+dx^2+dy^2\right)\quad {\rm where} \quad r_c^2={-t^2+x^2+y^2\over 4\ell^2}\,. \label{stm} \end{equation} This metric is time-symmetric about $t=0$ but not static. It remains invariant under the Lorentz group of the flat 2+1-dimensional Minkowski space ($t,\,x,\,y$). In addition the origin may be shifted and the projection ``centered" about any point in AdS space (by projecting from the corresponding antipodal point). \begin{figure} \unitlength 0.90mm \linethickness{0.4pt} \begin{picture}(127.20,65.00)(12,0) \bezier{184}(63.00,55.20)(67.47,55.20)(70.00,45.10) \bezier{180}(70.00,45.10)(71.99,35.00)(70.00,24.90) \bezier{184}(70.00,24.90)(67.47,14.80)(63.00,14.80) \bezier{30}(63.00,55.20)(58.53,55.20)(56.00,45.10) \bezier{30}(56.00,45.10)(54.01,35.00)(56.00,24.90) \bezier{30}(56.00,24.90)(58.53,14.80)(63.00,14.80) \bezier{184}(30.00,55.20)(33.67,55.20)(35.76,45.10) \bezier{180}(35.76,45.10)(37.40,35.00)(35.76,24.90) \bezier{184}(35.76,24.90)(33.67,14.80)(30.00,14.80) \bezier{184}(30.00,55.20)(26.33,55.20)(24.24,45.10) \bezier{180}(24.24,45.10)(22.60,35.00)(24.24,24.90) \bezier{184}(24.24,24.90)(26.33,14.80)(30.00,14.80) \bezier{160}(31.00,55.00)(47.00,43.00)(63.00,55.00) \bezier{24}(63.00,55.20)(61.51,55.20)(59.00,53.00) \put(33.00,17.00){\line(-1,1){9.00}} \put(70.00,45.00){\line(-1,0){34.50}} \put(35.50,45.00){\line(-1,1){8.20}} \put(59.00,53.00){\line(-1,0){24.50}} \bezier{28}(70.00,45.00)(63.67,49.07)(59.00,53.00) \bezier{14}(34.00,53.00)(31.63,52.96)(27.00,53.00) \put(33.00,17.00){\line(1,0){34.00}} \bezier{28}(67.00,17.00)(60.70,22.59)(56.00,26.00) \bezier{64}(56.00,26.00)(39.96,25.93)(24.00,26.00) \put(47.00,21.00){\circle*{0.80}} \bezier{80}(61.00,15.20)(47.00,26.85)(31.00,15.00) \put(47.00,21.00){\line(1,2){16.00}} \put(63.30,53.00){\makebox(0,0)[lc]{A}} \put(63.00,53.00){\circle*{0.80}} \put(60.50,48.00){\circle*{0.80}} \put(61.00,48.00){\makebox(0,0)[lc]{B}} \put(47.00,20.93){\makebox(0,0)[rb]{P}} \put(28.00,23.00){\makebox(0,0)[lb]{2}} \put(30.20,51.20){\makebox(0,0)[lc]{1}} \bezier{184}(126.80,20.90)(129.00,14.93)(117.10,12.00) \bezier{180}(117.10,12.00)(107.00,10.01)(96.90,12.00) \bezier{184}(96.90,12.00)(86.80,14.53)(86.80,19.00) \bezier{30}(127.20,19.00)(127.20,23.47)(117.10,26.00) \bezier{30}(117.10,26.00)(107.00,27.99)(96.90,26.00) \bezier{30}(96.90,26.00)(86.80,23.47)(86.80,19.00) \bezier{184}(127.20,52.00)(127.20,48.33)(117.10,46.24) \bezier{184}(127.20,52.00)(127.20,48.33)(117.10,46.24) \bezier{180}(117.10,46.24)(107.00,44.60)(96.90,46.24) \bezier{184}(96.90,46.24)(86.80,48.33)(86.80,52.00) \bezier{184}(127.20,52.00)(127.20,55.67)(117.10,57.76) \bezier{180}(117.10,57.76)(107.00,59.40)(96.90,57.76) \bezier{184}(96.90,57.76)(86.80,55.67)(86.80,52.00) \bezier{160}(127.00,51.00)(115.00,35.00)(126.80,21.00) \bezier{180}(87.20,21.00)(98.85,35.00)(87.00,51.00) \put(48.00,5.00){\makebox(0,0)[cc]{({\bf a})}} \put(107.00,5.00){\makebox(0,0)[cc]{({\bf b})}} \put(107.00,35.00){\vector(0,1){30.00}} \put(107.00,35.00){\vector(4,-1){19.00}} \put(107.00,35.00){\vector(3,2){10.00}} \put(117.00,42.00){\makebox(0,0)[lb]{$y$}} \put(126.00,30.00){\makebox(0,0)[lc]{$x$}} \put(107.00,65.20){\makebox(0,0)[cb]{$t$}} \end{picture} \caption{AdS space in stereographic projection. ({\bf a}) The hyperboloid is 2-dimensional AdS space embedded in 3-dimensional flat space as in Eq (\ref{ads}), restricted to $Y=0$. It is projected from point P onto the plane 1 ($U=\ell$). The image of point A in the hyperboloid is point B in the plane. The part of the hyperboloid that lies below plane 2 is not covered by the stereographic coordinates. ({\bf b}) When plotted in the stereographic coordinates (\ref{stc}), AdS space is the interior of a hyperboloid. The boundary of the hyperboloid is (part of) the conformal boundary of AdS space.} \label{fig4} \end{figure} Because of the condition $U>-\ell$ the stereographic projection fails to cover a part of AdS space, even in the periodically identified version (Fig.~4a). The 3-dimensional conformal diagram is the interior of the hyperboloid $r_c=1$, where the conformal factor of the metric (\ref{stm}) is finite (Fig.~4b). On the surface of time-symmetry, $t=0$, the stereographic metric agrees with the sausage metric (\ref{sausge}). Many similar coordinate systems, illustrating various symmetries of AdS space, are possible; for examples see \cite{Li}. \subsection{Isometries and Geodesics} To discuss the identifications that lead to time-symmetric black holes and other globally non-trivial 2+1-dimensional solutions we need a convenient representation of isometries and other geometrical relations in a spacelike initial surface of time-symmetry. Such a representation is the conformal map of Figs 1 and 2, in which this spacelike surface is shown as a disk, known as the {\it Poincar\'e disk}. This representation has been extensively studied (see, for example, \cite{BP}), and we only mention the features that are most important for our task. \begin{figure} \unitlength 0.63mm \linethickness{0.4pt} \begin{picture}(144.00,109.00) \bezier{436}(43.00,95.00)(80.00,55.00)(117.00,95.00) \bezier{160}(43.00,96.00)(42.67,100.04)(60.00,102.83) \bezier{188}(60.00,102.83)(80.00,105.48)(100.00,102.83) \bezier{160}(100.00,102.83)(115.67,100.92)(117.00,96.00) \bezier{160}(43.00,96.00)(42.67,91.96)(60.00,89.17) \bezier{190}(60.00,89.17)(80.00,86.52)(100.00,89.17) \bezier{160}(100.00,89.17)(115.67,91.08)(117.00,96.00) \bezier{100}(43.00,15.00)(80.00,55.00)(117.00,15.00) \bezier{160}(41.15,75.00)(40.80,79.24)(59.00,82.17) \bezier{188}(59.00,82.17)(80.00,84.95)(101.00,82.17) \bezier{160}(101.00,82.17)(117.45,80.17)(118.85,75.00) \bezier{160}(41.15,75.00)(40.80,70.76)(59.00,67.83) \bezier{188}(59.00,67.83)(80.00,65.05)(101.00,67.83) \bezier{160}(101.00,67.83)(117.45,69.83)(118.85,75.00) \bezier{70}(40.00,95.00)(60.00,75.00)(80.00,55.00) \bezier{70}(80.00,55.00)(98.00,73.00)(120.00,95.00) \bezier{40}(43.00,14.00)(42.67,18.04)(60.00,20.83) \bezier{45}(60.00,20.83)(80.00,23.48)(100.00,20.83) \bezier{40}(100.00,20.83)(115.67,18.92)(117.00,14.00) \bezier{40}(43.00,14.00)(42.67,9.96)(60.00,7.17) \bezier{45}(60.00,7.17)(80.00,4.52)(100.00,7.17) \bezier{40}(100.00,7.17)(115.67,9.08)(117.00,14.00) \put(122.00,109.00){\vector(-1,-1){9.00}} \put(124.00,109.00){\makebox(0,0)[lc]{hyperboloid}} \put(142.00,75.00){\vector(-1,0){10.00}} \put(144.00,75.00){\makebox(0,0)[lc]{plane}} \put(28.00,69.00){\makebox(0,0)[rc]{limit circle}} \put(125.00,7.00){\vector(-4,3){8.00}} \put(126.00,8.00){\makebox(0,0)[lc]{other sheet of hyperboloid}} \put(126.00,5.50){\makebox(0,0)[lt]{(not used in construction)}} \put(29.00,69.00){\vector(1,0){23.00}} \put(123.00,85.00){\line(5,-6){16.67}} \thicklines \put(140.00,65.00){\line(-1,0){105.00}} \put(35.00,65.00){\line(-3,5){12.00}} \thinlines \put(23.00,85.00){\line(1,0){100.00}} \put(80.00,55.00){\vector(0,1){40.00}} \put(80.00,96.00){\makebox(0,0)[cb]{$U$}} \put(80.00,55.00){\vector(4,-1){30.00}} \put(111.00,47.00){\makebox(0,0)[lc]{$X$}} \put(80.00,55.00){\vector(3,2){12.00}} \put(93.00,63.00){\makebox(0,0)[lc]{$Y$}} \put(80.00,45.00){\makebox(0,0)[cc]{$\ell$}} \put(80.00,48.00){\vector(0,1){7.00}} \put(80.00,43.00){\vector(0,-1){8.00}} \bezier{50}(118.00,73.00)(99.00,54.00)(80.00,35.00) \bezier{50}(80.00,35.00)(60.67,54.33)(43.00,72.00) \put(80.00,35.00){\line(-1,2){27.00}} \put(53.00,89.00){\circle*{1.0}} \put(61.00,73.00){\circle*{1.0}} \put(80.00,35.00){\circle*{1.0}} \put(54.00,89.00){\makebox(0,0)[lc]{A}} \put(62.00,73.00){\makebox(0,0)[lc]{B}} \put(80.00,34.00){\makebox(0,0)[ct]{P}} \end{picture} \caption{The two-dimensional space $H^2$ of constant curvature $1/\ell^2$ is embedded in flat Minkowski space as one sheet of the hyperboloid of Eq (\ref{emb}). Under a stereographic projection from point P to the plane, point A on the hyperboloid is mapped to point B in the plane. Thus the hyperboloid ($H^2$) is mapped onto the Poincar\'e disk, the interior of the curve marked ``limit circle" \end{figure} All totally geodesic, time-symmetric surfaces $H^2$ in AdS space are isometric to the typical hyperboloid (Fig.~5) obtained by restricting Eq (\ref{ads}) to $V=0$, \begin{equation} X^2+Y^2-U^2=-\ell^2 \label{emb} \end{equation} This surface has zero extrinsic curvature and therefore constant negative Gaussian curvature $-1/\ell^2$. The Poincar\'e disk can be obtained as a map of $H^2$ by the stereographic projection of Fig.~5, which illustrates Eq (\ref{stc}) when restricted to $V=0$ similar to the way Fig.~4 illustrates it when restricted to $X=0$. In this way all of $H^2$ is mapped into the interior of a disk of radius $2\ell$, whose boundary, called the limit circle, represents points at (projective or conformal) infinity. Because the map is conformal, angles are faithfully represented. Other geometrical objects in $H^2$ appear distorted in the Euclidean geometry of the disk, but by assigning new roles to these ``distorted" objects and manipulating those according to Euclidean geometry one can perform constructions equivalent to those in the $H^2$-geometry directly on the Poincar\'e disk. For example, on the surface $H^2$ as described by Eq (\ref{emb}), all geodesics are intersections of planes through the origin with the surface; that is, they satisfy a linear relation between $X,\,Y,\,U$. From Eq (\ref{stc}) it follows directly that Eq (\ref{emb}) becomes such a linear relation if $x,\,y$ satisfy the equation of a circle that has radius $(a^2-4\ell^2)$ if it is centered at $(x,\,y)=(a_x,\,a_y)$, hence meets the limit circle at right angles. Because two such circles intersect in at most one point in the interior of the Poincar\'e disk, it follows that two geodesics in $H^2$ meet at most in one point (as in Euclidean space). An important difference occurs if two geodesics do {\it not} meet: in Euclidean space they are then equidistant; whereas in the Poincar\'e disc the geodesic between points on two disjoint geodesics (Euclidean circles perpendicular to the limit circle) approaches a complete geodesic as the points approach the limit circle. Since the conformal factor in the metric of Eq (\ref{stm}), restricted to $t=0$, \begin{equation} ds^2 = \left(1-{x^2+y^2\over 4\ell^2}\right)^{-2}\left(dx^2+dy^2\right) \label{cm} \end{equation} increases without limit as $x^2+y^2\rightarrow 4\ell^2$ on $H^2$ the geodesic distance between two given disjoint geodesics typically increases without bound as we go along the given geodesics in either direction. However, the geodesic distance between points on two given disjoint geodesics of course has a lower bound. If this is nonzero there is a unique geodesic segment of minimal length joining the two given geodesics at right angles to either. On the other hand, if we have a family of equidistant curves, at most one of them can be a geodesic, and then the representation of the others on the Poincar\'e disk are arcs of circles, not perpendicular to the limit circle, but meeting the geodesic asymptotically at the limit circle. The curves $r=$ const of Fig.~2b are examples, with $r=\ell$ the geodesic of the family. These equidistant curves have constant acceleration (with respect to their arclength parameter), and they also illustrate how the conformal factor in (\ref{cm}) distorts the apparent (Euclidean) distances of the disk into the true distances of $H^2$. Because the surface (\ref{emb}) in Minkowski space has constant extrinsic curvature, any isometry of the surface geometry can be extended to an isometry of the embedding space. But we know all those isometries: they form the homogeneous isochronous Lorentz group. Thus any Lorentz transformation implies, by the projection of Fig.~5, a corresponding transformation of the Poincar\'e disk that represents an isometry of $H^2$, and all $H^2$ isometries can be obtained in this way. In the Euclidean metric of the disk such transformations must be conformal transformations leaving the limit circle fixed, since they are isometries of the conformal metric (\ref{cm}). Knowing this we can now classify\footnote{We confine attention to orientation-preserving transformations; they can be combined with a reflection about a geodesic (with an infinite number of fixed points) to obtain the rest.} the isometries of $H^2$. Proper Lorentz transformations in 3D Minkowski space have an axis of fixed points that may be a spacelike, null, or timelike straight line. If the axis is timelike, it intersects the hyperboloid (\ref{emb}). If the axis is null, it intersects the hyperboloid asymptotically. If the axis is spacelike, it does not intersect the hyperboloid, but there are two fixed null directions perpendicular to the axis. Correspondingly on the Poincar\'e disk there is either one fixed point within the disk (``elliptic"), or one fixed point on the limit circle (``parabolic"), or two fixed points on the limit circle (``hyperbolic") for these transformations. Fig.~1b illustrates by the transformation $\theta \rightarrow \theta+$const the case with one finite fixed point (the origin). Figs.~2b and 3b illustrate by the transformation $\phi \rightarrow \phi+$const the case with two fixed points and one fixed point, respectively, on the limit circle ($\phi =\pm\infty$). In the case of two fixed points there is a unique geodesic ($r=\ell$ in Fig.~2b) left fixed by the isometry, and conversely the isometry, which we will call ``along" the geodesic, is uniquely defined by the invariant geodesic and the distance by which a point moves along that geodesic. Except for the rotation about the center of the disk as in Fig.~1b these are not isometries of the disk's flat, Euclidean metric, but they are of course conformal isometries of this metric. Such conformal transformations, mapping the limit circle into itself, are conveniently described as {\em M\"obius transformations} of the complex coordinate \begin{equation} z = {x+iy\over\ell} \quad {\rm by} \quad z\rightarrow z'={az+b\over \bar b z+\bar a}\,, \label{moeb} \end{equation} where $a,\,b$ are complex numbers with $|a|^2-|b|^2=1$. When we consider an isometry or identification abstractly, it can always be implemented concretely by such a M\"obius transformation. In particular, hyperbolic isometries are described by M\"obius transformations with real $a$. As the examples of Figs.~1-3 show, each of these isometries is part of a family depending on a continuous parameter (the constant in $\phi\rightarrow\phi +$const, for example). There is therefore an ``infinitesimal" version of each isometry, described by a Killing vector ($\partial/\partial\phi$ in the example). Conversely an (orientation-preserving) isometry can be described as the exponential of its Killing vector. \subsection{Identifications} The hyperbolic transformations, which have no fixed points in $H^2$, are suitable for forming nonsingular quotient spaces that have the same local geometry as AdS space, and hence satisfy the same Einstein equations. In the context of Fig.~2b and Eq (\ref{schwmet}) the transformation that comes to mind is described by $\phi\rightarrow\phi+2\pi$. The quotient space is the space in which points connected by this transformation are regarded as identical, which is the same as the space in which $\phi$ is a periodic coordinate with the usual period. Eq (\ref{schwmet}) with this periodicity in $\phi$ already gives us the simplest BTZ metric for a single, non-rotating 2+1-dimensional black hole. It is asymptotically AdS, as shown by comparing Eqs (\ref{schwmet}) and (\ref{m0}). The minimum distance between the two identified geodesics occurs at $r=\ell$ and is $2\pi\ell$. This is the minimum distance around the black hole, and plays the role of the horizon ``area". If we identify $\phi$ with a different period $2\pi a$, we get a metric with a different horizon size. We can then redefine the coordinates so that $\phi$ has its usual period, $$\phi\rightarrow a\phi,\qquad r\rightarrow r/a\qquad t\rightarrow at$$ and the metric takes this standard form, called the BTZ metric \cite{BTZ}: \begin{equation} ds^2=-\left({r^2\over\ell^2}-m\right) dt^2+\left({r^2\over\ell^2}-m\right)^{-1} dr^2+ r^2 d\phi^2, \label{BTZ} \end{equation} where $m=1/a^2$. Here the dimensionless quantity $m$ is called the mass parameter. Although it can be measured in the asymptotic region, it is more directly related to the horizon size, the length of the minimal geodesic at the horizon, $2\pi\ell\sqrt{m}$. The metric (\ref{BTZ}) is a solution also for $m=0$, as shown by Eq (\ref{horom}), but that is not the AdS metric itself. The latter is also described by Eq (\ref{BTZ}), but with $m=-1$, as shown by Eq (\ref{m0}). By contrast, the $m=0$ initial state is obtained by identifying the geodesics $\phi=0$ and $\phi=2\pi$ in Fig.~3b. To describe the identification more explicitly, we may say that we have cut a strip from $\phi=0$ to $\phi=2\pi$ out of Fig.~2b, and glued the edges together. This strip is a ``fundamental domain" for our identification, a region that contains images of its own points under the group only on its boundary, and that together with all its images covers the full AdS space. To obtain a fundamental domain for the BTZ black hole we might have used as the boundaries some other curve on the Poincar\'e disk and its image under the transformation, provided only that the curve and its image do not intersect. But since it is always possible to avoid apparent asymmetries by choosing boundaries composed of geodesics that meet at right angles, we will generally do so. We can think of the identification in yet another way, by a process that has been called ``doubling": cut a strip from $\phi=0$ to $\phi=\pi$ from Fig.~2b, and cut another identical strip. Put one on top of the other and glue the two edges together, obtaining again the black hole initial state. The gluing makes the two strips reflections of each other with respect to either of the original edges. Back on the Poincar\'e disk the composition of the two reflections is a translation in $\phi$ by $2\pi$, that is, the isometry of the identification. Any (orientation-preserving) isometry of a hyperbolic space can be decomposed into two reflections \cite{Bach}; hence any quotient space can be considered the double of a suitable region (possibly in another quotient space), and a fundamental domain is obtained from the region and one of its reflections. The process of gluing together a constant negative curvature space from a fundamental domain of the Poincar\'e disk can be reversed: we cut the space by geodesics into its fundamental domain, make many copies of the domain, and put these down on the disk so that boundaries coming from the same cut touch, until the entire disk is covered. The resulting pattern is called a ``tiling" of the disk (although the ``tiles" corresponding to the $t=0$ section of the BTZ black hole look more like strip flooring). Thus we have two equivalent ways of describing our identified space: by giving a fundamental domain and rules of gluing the boundaries, or by a tiling together with rules relating each tile to its neighbors. \begin{figure} \unitlength 0.70mm \linethickness{0.4pt} \begin{picture}(150.00,82.00)(-5,0) \thicklines \bezier{40}(5.00,17.00)(5.00,47.00)(5.00,77.00) \bezier{40}(5.00,77.00)(35.00,77.00)(65.00,77.00) \bezier{40}(35.00,17.00)(35.00,47.00)(35.00,77.00) \bezier{40}(65.00,17.00)(65.00,47.00)(65.00,77.00) \bezier{40}(5.00,47.00)(35.00,47.00)(65.00,47.00) \bezier{40}(5.00,17.00)(35.00,17.00)(65.00,17.00) \put(35.00,47.00){\line(5,3){9.00}} \put(65.00,47.00){\line(-3,5){21.00}} \put(44.00,82.00){\line(-5,-3){22.00}} \put(65.00,47.00){\line(-5,-3){13.00}} \put(52.00,39.20){\line(-3,5){7.68}} \put(65.00,47.00){\circle*{2.00}} \put(35.00,47.00){\circle*{2.00}} \put(57.00,60.00){\circle*{2.00}} \put(35.00,47.00){\line(-3,5){13.20}} \thinlines \put(22.00,39.00){\line(5,3){36.00}} \put(27.20,60.00){\line(-5,-3){22.20}} \put(27.20,30.00){\line(-5,-3){22.20}} \put(57.20,30.00){\line(-5,-3){22.20}} \put(65.00,17.00){\line(-5,-3){13.00}} \put(52.00,9.20){\line(-3,5){7.68}} \put(65.00,77.00){\line(-5,-3){13.00}} \put(35.00,47.00){\line(-3,5){21.00}} \put(14.00,82.00){\line(-5,-3){9.00}} \put(35.00,17.00){\line(-3,5){21.00}} \put(65.00,17.00){\line(-3,5){13.80}} \put(35.00,17.00){\line(-5,-3){13.00}} \put(22.00,9.20){\line(-3,5){7.68}} \thicklines \put(108.00,40.00){\line(0,1){12.00}} \put(108.00,52.00){\line(1,0){12.00}} \put(120.00,52.00){\line(0,1){18.00}} \put(108.00,40.00){\line(-1,0){18.00}} \put(90.00,40.00){\line(0,1){30.00}} \put(90.00,70.00){\line(1,0){30.00}} \bezier{23}(90.00,40.00)(105.00,46.00)(120.00,52.00) \bezier{23}(90.00,40.00)(96.00,55.00)(102.00,70.00) \bezier{23}(108.00,22.00)(123.00,28.00)(138.00,34.00) \bezier{23}(108.00,22.00)(114.00,37.00)(120.00,52.00) \bezier{23}(120.00,52.00)(135.00,58.00)(150.00,64.00) \bezier{23}(120.00,52.00)(126.00,67.00)(132.00,82.00) \bezier{23}(138.00,34.00)(144.00,49.00)(150.00,64.00) \bezier{23}(78.00,10.00)(84.00,25.00)(90.00,40.00) \bezier{23}(102.00,70.00)(117.00,76.00)(132.00,82.00) \bezier{23}(78.00,10.00)(93.00,16.00)(108.00,22.00) \thinlines \put(138.00,34.00){\line(0,1){18.00}} \put(126.00,22.00){\line(-1,0){18.00}} \put(120.00,52.00){\line(1,0){18.00}} \put(108.00,40.00){\line(0,-1){18.00}} \put(108.00,40.00){\line(1,0){12.00}} \put(120.00,40.00){\line(0,1){12.00}} \put(120.00,52.00){\circle*{2.00}} \put(120.00,40.00){\circle*{2.00}} \put(90.00,40.00){\circle*{2.00}} \put(96.00,10.00){\line(0,1){12.00}} \put(96.00,22.00){\line(1,0){12.00}} \put(96.00,10.00){\line(-1,0){18.00}} \put(78.00,10.00){\line(0,1){30.00}} \put(78.00,40.00){\line(1,0){12.00}} \put(120.00,70.00){\line(0,1){12.00}} \put(138.00,52.00){\line(0,1){12.00}} \put(138.00,64.00){\line(1,0){12.00}} \put(150.00,64.00){\line(0,1){18.00}} \put(120.00,82.00){\line(1,0){30.00}} \put(138.00,34.00){\line(1,0){12.00}} \put(126.00,22.00){\line(0,-1){12.00}} \put(102.00,70.00){\line(0,1){12.00}} \put(90.00,58.00){\line(-1,0){12.00}} \put(126.00,22.00){\line(0,1){12.00}} \put(126.00,34.00){\line(1,0){12.00}} \put(144.00,10.00){\line(0,1){6.00}} \put(144.00,16.00){\line(1,0){6.00}} \put(35.00,2.00){\makebox(0,0)[cc]{({\bf a})}} \put(114.00,2.00){\makebox(0,0)[cc]{({\bf b})}} \put(40.00,64.00){\makebox(0,0)[cc]{$a^2$}} \put(54.00,50.00){\makebox(0,0)[cc]{$b^2$}} \put(105.00,55.00){\makebox(0,0)[cc]{$a^2$}} \put(114.00,46.00){\makebox(0,0)[cc]{$b^2$}} \end{picture} \caption{Two different ways of tiling the plane prove the theorem of Pythagoras in ({\bf a}) Euclidean space and ({\bf b}) Minkowski space} \end{figure} \subsubsection{Tiling and Pythagoras} To fix ideas, consider an application of tiling found among the numerous proofs of the theorem of Pythgoras (a local boy who contributed to the early fame of Samos). This proof is based on the fact that all fundamental domains of a given group of isometries have equal area. In the Euclidean plane we consider the group generated by two translations specified in direction and amounts by two adjacent sides of the square above the hypotenuse of a right triangle, whose vertices are the three larger dots in Fig.~6a. This square is a fundamental domain of the group, and part of the tiling by this square is shown by the horizontal and vertical dotted lines. The region drawn in heavy outline is an alternative fundamental domain of the same group of isometries, and that domain is made from the squares above the sides of the same triangle. Part of its tiling of the plane is shown by the lightly drawn lines of Fig.~6a. Either fundamental domain can be glued together to form the same quotient space, a ``square" torus, so the areas are equal, $c^2 = a^2+b^2 =$ area of torus. In special relativity the theorem of Pythagoras is valid with a different sign, $c^2=a^2-b^2$ if we choose the hypotenuse and one of the sides to be spacelike, and of course the right angles of the triangle and of squares are to be drawn in accordance with the Minkowski metric. Fig.~6b shows the proof by the tiling that derives from a Minkowski torus of area $c^2$. (Here we use, at least implicitly, the fact that the area of a two-dimensional figure is the same in Euclidean and Minkowski spaces if their metrics differ only by a sign.) \begin{figure} \unitlength 0.70mm \linethickness{0.4pt} \begin{picture}(155.08,136.96) \bezier{50}(4.92,20.00)(4.92,28.41)(12.46,33.19) \bezier{50}(12.46,33.19)(20.00,36.96)(27.54,33.19) \bezier{50}(27.54,33.19)(35.08,28.41)(35.08,20.00) \bezier{50}(4.92,20.00)(4.92,11.59)(12.46,6.81) \bezier{50}(12.46,6.81)(20.00,3.04)(27.54,6.81) \bezier{50}(27.54,6.81)(35.08,11.59)(35.08,20.00) \bezier{50}(44.92,20.00)(44.92,28.41)(52.46,33.19) \bezier{50}(84.92,20.00)(84.92,28.41)(92.46,33.19) \bezier{50}(124.92,20.00)(124.92,28.41)(132.46,33.19) \bezier{50}(52.46,33.19)(60.00,36.96)(67.54,33.19) \bezier{50}(92.46,33.19)(100.00,36.96)(107.54,33.19) \bezier{50}(132.46,33.19)(140.00,36.96)(147.54,33.19) \bezier{50}(67.54,33.19)(75.08,28.41)(75.08,20.00) \bezier{50}(107.54,33.19)(115.08,28.41)(115.08,20.00) \bezier{50}(147.54,33.19)(155.08,28.41)(155.08,20.00) \bezier{50}(44.92,20.00)(44.92,11.59)(52.46,6.81) \bezier{50}(84.92,20.00)(84.92,11.59)(92.46,6.81) \bezier{50}(124.92,20.00)(124.92,11.59)(132.46,6.81) \bezier{50}(52.46,6.81)(60.00,3.04)(67.54,6.81) \bezier{50}(92.46,6.81)(100.00,3.04)(107.54,6.81) \bezier{50}(132.46,6.81)(140.00,3.04)(147.54,6.81) \bezier{50}(67.54,6.81)(75.08,11.59)(75.08,20.00) \bezier{50}(107.54,6.81)(115.08,11.59)(115.08,20.00) \bezier{50}(147.54,6.81)(155.08,11.59)(155.08,20.00) \put(89.00,9.00){\line(1,1){11.00}} \put(100.00,20.00){\line(1,-1){11.00}} \bezier{120}(49.00,9.00)(60.00,18.00)(60.00,35.00) \bezier{120}(71.00,9.00)(60.00,18.00)(60.00,35.00) \bezier{120}(31.00,9.00)(21.00,20.00)(31.00,31.00) \bezier{120}(9.00,9.00)(19.00,20.00)(9.00,31.00) \bezier{50}(4.92,120.00)(4.92,128.41)(12.46,133.19) \bezier{50}(12.46,133.19)(20.00,136.96)(27.54,133.19) \bezier{50}(27.54,133.19)(35.08,128.41)(35.08,120.00) \bezier{50}(4.92,120.00)(4.92,111.59)(12.46,106.81) \bezier{50}(12.46,106.81)(20.00,103.04)(27.54,106.81) \bezier{50}(27.54,106.81)(35.08,111.59)(35.08,120.00) \bezier{50}(44.92,120.00)(44.92,128.41)(52.46,133.19) \bezier{50}(84.92,120.00)(84.92,128.41)(92.46,133.19) \bezier{50}(124.92,120.00)(124.92,128.41)(132.46,133.19) \bezier{50}(52.46,133.19)(60.00,136.96)(67.54,133.19) \bezier{50}(92.46,133.19)(100.00,136.96)(107.54,133.19) \bezier{50}(132.46,133.19)(140.00,136.96)(147.54,133.19) \bezier{50}(67.54,133.19)(75.08,128.41)(75.08,120.00) \bezier{50}(107.54,133.19)(115.08,128.41)(115.08,120.00) \bezier{50}(147.54,133.19)(155.08,128.41)(155.08,120.00) \bezier{50}(44.92,120.00)(44.92,111.59)(52.46,106.81) \bezier{50}(84.92,120.00)(84.92,111.59)(92.46,106.81) \bezier{50}(124.92,120.00)(124.92,111.59)(132.46,106.81) \bezier{50}(52.46,106.81)(60.00,103.04)(67.54,106.81) \bezier{50}(92.46,106.81)(100.00,103.04)(107.54,106.81) \bezier{50}(132.46,106.81)(140.00,103.04)(147.54,106.81) \bezier{50}(67.54,106.81)(75.08,111.59)(75.08,120.00) \bezier{50}(107.54,106.81)(115.08,111.59)(115.08,120.00) \bezier{50}(147.54,106.81)(155.08,111.59)(155.08,120.00) \thicklines \put(89.00,109.00){\line(1,1){11.00}} \put(100.00,120.00){\line(1,-1){11.00}} \bezier{120}(49.00,109.00)(60.00,118.00)(60.00,135.00) \bezier{120}(71.00,109.00)(60.00,118.00)(60.00,135.00) \bezier{120}(31.00,109.00)(21.00,120.00)(31.00,131.00) \bezier{120}(9.00,109.00)(19.00,120.00)(9.00,131.00) \thinlines \bezier{10}(9.79,9.00)(9.79,10.32)(14.90,11.07) \bezier{10}(14.90,11.07)(20.00,11.66)(25.10,11.07) \bezier{10}(25.10,11.07)(30.21,10.32)(30.21,9.00) \bezier{50}(9.79,9.00)(9.79,7.68)(14.90,6.93) \bezier{50}(14.90,6.93)(20.00,6.34)(25.10,6.93) \bezier{50}(25.10,6.93)(30.21,7.68)(30.21,9.00) \bezier{50}(9.79,31.00)(9.79,32.32)(14.90,33.07) \bezier{50}(14.90,33.07)(20.00,33.66)(25.10,33.07) \bezier{50}(25.10,33.07)(30.21,32.32)(30.21,31.00) \bezier{50}(9.79,31.00)(9.79,29.68)(14.90,28.93) \bezier{50}(14.90,28.93)(20.00,28.34)(25.10,28.93) \bezier{50}(25.10,28.93)(30.21,29.68)(30.21,31.00) \bezier{10}(49.79,9.00)(49.79,10.32)(54.90,11.07) \bezier{10}(54.90,11.07)(60.00,11.66)(65.10,11.07) \bezier{10}(65.10,11.07)(70.21,10.32)(70.21,9.00) \bezier{50}(49.79,9.00)(49.79,7.68)(54.90,6.93) \bezier{50}(54.90,6.93)(60.00,6.34)(65.10,6.93) \bezier{50}(65.10,6.93)(70.21,7.68)(70.21,9.00) \bezier{10}(89.79,9.00)(89.79,10.32)(94.90,11.07) \bezier{10}(94.90,11.07)(100.00,11.66)(105.10,11.07) \bezier{10}(105.10,11.07)(110.21,10.32)(110.21,9.00) \bezier{50}(89.79,9.00)(89.79,7.68)(94.90,6.93) \bezier{50}(94.90,6.93)(100.00,6.34)(105.10,6.93) \bezier{50}(105.10,6.93)(110.21,7.68)(110.21,9.00) \bezier{10}(129.79,9.00)(129.79,10.32)(134.90,11.07) \bezier{10}(134.90,11.07)(140.00,11.66)(145.10,11.07) \bezier{10}(145.10,11.07)(150.21,10.32)(150.21,9.00) \bezier{50}(129.79,9.00)(129.79,7.68)(134.90,6.93) \bezier{50}(134.90,6.93)(140.00,6.34)(145.10,6.93) \bezier{50}(145.10,6.93)(150.21,7.68)(150.21,9.00) \bezier{112}(130.00,9.00)(140.00,19.00)(150.00,9.00) \bezier{10}(9.79,59.00)(9.79,60.32)(14.90,61.07) \bezier{10}(14.90,61.07)(20.00,61.66)(25.10,61.07) \bezier{10}(25.10,61.07)(30.21,60.32)(30.21,59.00) \bezier{50}(9.79,59.00)(9.79,57.68)(14.90,56.93) \bezier{50}(14.90,56.93)(20.00,56.34)(25.10,56.93) \bezier{50}(25.10,56.93)(30.21,57.68)(30.21,59.00) \bezier{50}(9.79,81.00)(9.79,82.32)(14.90,83.07) \bezier{50}(14.90,83.07)(20.00,83.66)(25.10,83.07) \bezier{50}(25.10,83.07)(30.21,82.32)(30.21,81.00) \bezier{50}(9.79,81.00)(9.79,79.68)(14.90,78.93) \bezier{50}(14.90,78.93)(20.00,78.34)(25.10,78.93) \bezier{50}(25.10,78.93)(30.21,79.68)(30.21,81.00) \bezier{10}(49.79,59.00)(49.79,60.32)(54.90,61.07) \bezier{10}(89.79,59.00)(89.79,60.32)(94.90,61.07) \bezier{10}(54.90,61.07)(60.00,61.66)(65.10,61.07) \bezier{10}(94.90,61.07)(100.00,61.66)(105.10,61.07) \bezier{10}(65.10,61.07)(70.21,60.32)(70.21,59.00) \bezier{10}(105.10,61.07)(110.21,60.32)(110.21,59.00) \bezier{50}(49.79,59.00)(49.79,57.68)(54.90,56.93) \bezier{50}(89.79,59.00)(89.79,57.68)(94.90,56.93) \bezier{50}(54.90,56.93)(60.00,56.34)(65.10,56.93) \bezier{50}(94.90,56.93)(100.00,56.34)(105.10,56.93) \bezier{50}(65.10,56.93)(70.21,57.68)(70.21,59.00) \bezier{50}(105.10,56.93)(110.21,57.68)(110.21,59.00) \bezier{100}(50.00,59.70)(58.00,59.70)(59.50,95.00) \bezier{88}(90.00,59.70)(97.00,59.70)(100.00,75.00) \bezier{52}(11.00,80.00)(16.00,79.00)(16.00,70.00) \bezier{64}(16.00,70.00)(16.00,61.00)(10.00,59.70) \put(15.00,134.00){\line(1,0){10.00}} \put(14.00,120.00){\line(1,0){12.00}} \put(14.00,122.00){\line(1,0){12.00}} \put(26.00,124.00){\line(-1,0){12.00}} \put(13.00,126.00){\line(1,0){14.00}} \put(28.00,128.00){\line(-1,0){16.00}} \put(10.00,130.00){\line(1,0){20.00}} \put(29.00,132.00){\line(-1,0){18.00}} \bezier{50}(12.46,106.81)(20.00,103.04)(27.54,106.81) \put(15.00,106.00){\line(1,0){10.00}} \put(14.00,118.00){\line(1,0){12.00}} \put(26.00,116.00){\line(-1,0){12.00}} \put(13.00,114.00){\line(1,0){14.00}} \put(28.00,112.00){\line(-1,0){16.00}} \put(10.00,110.00){\line(1,0){20.00}} \put(29.00,108.00){\line(-1,0){18.00}} \put(55.00,106.00){\line(1,0){10.00}} \put(70.00,110.00){\line(-1,0){20.00}} \put(52.00,112.00){\line(1,0){16.00}} \put(66.00,114.00){\line(-1,0){12.00}} \put(56.00,116.00){\line(1,0){8.00}} \put(57.00,118.00){\line(1,0){6.00}} \put(62.00,120.00){\line(-1,0){4.00}} \put(59.00,122.00){\line(1,0){2.00}} \put(59.00,124.00){\line(1,0){2.00}} \put(59.67,126.00){\line(1,0){0.67}} \put(95.00,106.00){\line(1,0){11.00}} \put(109.00,108.00){\line(-1,0){17.00}} \put(90.00,110.00){\line(1,0){20.00}} \put(108.00,112.00){\line(-1,0){16.00}} \put(94.00,114.00){\line(1,0){12.00}} \put(104.00,116.00){\line(-1,0){8.00}} \put(98.00,118.00){\line(1,0){4.00}} \put(135.00,134.00){\line(1,0){10.00}} \put(131.00,132.00){\line(1,0){18.00}} \put(129.00,130.00){\line(1,0){22.00}} \put(128.00,128.00){\line(1,0){24.00}} \put(154.00,126.00){\line(-1,0){28.00}} \put(126.00,124.00){\line(1,0){28.00}} \put(126.00,122.00){\line(1,0){28.00}} \put(155.00,120.00){\line(-1,0){30.00}} \put(129.00,110.00){\line(1,0){22.00}} \put(128.00,112.00){\line(1,0){24.00}} \put(154.00,114.00){\line(-1,0){28.00}} \put(126.00,116.00){\line(1,0){28.00}} \put(126.00,118.00){\line(1,0){28.00}} \put(135.00,106.00){\line(1,0){10.00}} \put(131.00,108.00){\line(1,0){18.00}} \put(20.00,45.00){\makebox(0,0)[cc]{$m>0$}} \put(60.00,45.00){\makebox(0,0)[cc]{$m=0$}} \put(100.00,45.00){\makebox(0,0)[cc]{$-1<m<0$}} \put(140.00,45.00){\makebox(0,0)[cc]{$m=-1$}} \bezier{52}(29.00,80.00)(24.00,79.00)(24.00,70.00) \bezier{64}(24.00,70.00)(24.00,61.00)(30.00,59.70) \bezier{100}(70.00,59.70)(62.00,59.70)(60.50,95.00) \bezier{88}(110.00,59.70)(103.00,59.70)(100.00,75.00) \put(51.00,108.00){\line(1,0){18.00}} \end{picture} \caption{Three representations of the geometry of the $t=0$ geometry of metric (\ref{BTZ}) for different ranges of the mass parameter $m$: The BTZ black hole for $m>0$; the extremal BTZ black hole for $m=0$; the point particle (conical singularity) for $m<0$; and AdS (``vacuum") space itself for $m=-1$. {\bf Top row:} shaded regions of the Poincar\'e disk, to be identified in each figure along the left and right boundaries, drawn in thicker lines. {\bf Second row:} an embedding of the central part ($r\leq\ell\sqrt{1-m}$) of these spaces as surfaces in three-dimensional flat Euclidean space. The embedding cannot be continued beyond the outer edges of each figure. {\bf Bottom row:} the entire surface can be embedded in a 3D space of constant negative curvature, shown as a Poincar\'e ball. (The figure is schematic only; for example, the angle at the conical tips ought to be the same in the second and last row, to represent the same surface)}\label{fig6} \end{figure} \subsubsection{Embeddings} To visualize the geometry of our glued-together surface --- the $t=0$ surface of a static BTZ black hole --- it helps to embed this surface in a three-dimensional space in which the gluing can be actually carried out. This is analogous to the embedding of the $t=0$ surface of the Schwarzschild black hole, with one angle suppressed, in three-dimensional flat space (the surface of rotation of the {\it Flamm parabola} \cite{MTW}). For the BTZ initial surface only a finite part can be so embedded. The embedding stops where the rate of increase of circumference of the circle $r=$ const with respect to the true distance in the radial direction exceeds that rate in flat space. (The remainder of the surface could then be embedded in Minkowski space, but the switch between embeddings is an artifact and corresponds to no local intrinsic property.) However, the entire surface can be embedded in $H^3$, the {\it Riemannian} (positive definite metric) space of constant negative curvature. By the obvious generalization of the Poincar\'e disk this space can be conformally represented as a ball in three-dimensional flat space. Fig.~7 shows this embedding, where the surface for $m>0$ is seen to have two asymptotic sheets, similar to the corresponding Schwarzschild surface. \subsection{Multiple Black Holes} We saw that a single hyperbolic isometry (call it $a$) used as an identification to obtain an AdS initial state always yields a (single) BTZ black hole state, with horizon size and location depending on $a$. For other types of initial states we therefore need to use more than one such isometry, for example $a$ and $b$. Assuring that there are no fixed points (which would lead to singularities of the quotient space) would then seem to be much more difficult: If we know that $a$ has no fixed point, then the whole group consisting of powers $a^n$ has no fixed points (except the identity, $n=0$); but for the group generated by two isometries $a,\,b$ we have to check that no ``word" formed from these and their inverses, such as $ab^{-2}a^3b$ has fixed points. Although this may seem complicated, it is easy if we have a fundamental domain such that the isometry $a$ maps one of a pair of boundaries into the other, and the isometry $b$ does the same for a different pair of boundaries. Now tile the Poincar\'e disk with copies of this fundamental domain (see Fig.~10 for an example). Once we fix the original tile (associated with the identity isometry), there is a one-to-one correspondence between tiles and words. Therefore every non-trivial word moves all points in the original tile to some different tile, and there can be no fixed points in the open disk. How to obtain such a fundamental domain? A simple way is by doubling a region bounded by any number $k$ of non-intersecting geodesics \cite{DB}. Fig.~8a shows this for the case $k=3$. In Fig.~8b we see the fundamental domain. Half of it is the original (heavily outlined) region, shifted to the right so that the center of the Poincar\'e disk lies on the geodesic boundary of the region rather than at its center. The other half is the reflection of this original region across that geodesic boundary. Thus $2k-2 = 4$ boundaries remain to be identified in pairs, as indicated in the figure for the top pair. To construct the isometry that moves one member of such a pair into the other we find the unique common normal geodesic H$_2$ (shown for the bottom pair), and its intersection with the limit circle; these intersections are the fixed points of one of the hyperbolic isometries that have this fundamental domain. For example, in Fig.~8b the isometry associated with H$_2$ moves one of the bottom boundaries into the other. Similarly we find $k-2$ other isometries, each of them associated with a common normal. After the identification are made these common normals are smooth closed geodesics that separate an asymptotically AdS region from the rest of the manifold. We call such curves horizons. In addition to the $k-1$ horizons found this way there is another one, so there is a total of $k$ horizons. The additional one can be found in the above way from a different fundamental domain, obtained by reflecting the original region about a different geodesic boundary, but it is more easily found from the doubling picture, as shown by the H$_3$ in Fig.~8a. \begin{figure} \unitlength 0.750mm \linethickness{0.4pt} \begin{picture}(140.50,64.00) \thicklines \bezier{144}(26.91,35.01)(40.88,29.00)(26.91,22.99) \bezier{140}(40.69,38.50)(45.65,30.28)(64.67,32.29) \bezier{140}(40.69,19.54)(45.65,27.72)(64.67,25.63) \thinlines \bezier{64}(26.91,35.01)(31.87,37.90)(40.69,38.50) \bezier{64}(64.67,32.45)(67.52,29.00)(64.67,25.63) \bezier{64}(26.91,22.99)(31.87,20.10)(40.69,19.54) \bezier{40}(40.69,38.50)(58.52,39.02)(64.67,32.45) \bezier{40}(26.91,35.01)(16.44,29.00)(26.91,22.99) \bezier{40}(40.69,19.54)(58.52,18.98)(64.67,25.63) \bezier{15}(58.55,36.37)(54.58,34.71)(52.52,32.21) \bezier{64}(52.52,32.21)(50.47,29.00)(52.68,25.79) \bezier{15}(52.68,25.79)(54.44,23.23)(58.55,21.63) \put(26.00,41.00){\line(-2,-3){16.00}} \thicklines \put(10.00,17.00){\line(1,0){68.00}} \thinlines \put(78.00,17.00){\line(-2,3){16.00}} \put(62.00,41.00){\line(-1,0){36.00}} \thicklines \bezier{144}(26.91,58.71)(40.88,54.00)(26.91,49.29) \bezier{140}(40.69,61.44)(45.65,55.01)(64.67,56.57) \bezier{140}(40.69,46.59)(45.65,52.99)(64.67,51.36) \thinlines \bezier{64}(26.91,58.71)(31.87,60.97)(40.69,61.44) \bezier{64}(64.67,56.70)(67.52,54.00)(64.67,51.36) \bezier{64}(26.91,49.29)(31.87,47.03)(40.69,46.59) \bezier{40}(40.69,61.44)(58.52,61.85)(64.67,56.70) \bezier{40}(26.91,58.71)(16.44,54.00)(26.91,49.29) \bezier{40}(40.69,46.59)(58.52,46.15)(64.67,51.36) \bezier{15}(58.55,59.78)(54.58,58.47)(52.52,56.51) \bezier{64}(52.52,56.51)(50.47,54.00)(52.68,51.49) \bezier{15}(52.68,51.49)(54.44,49.48)(58.55,48.22) \multiput(26.00,63.40)(-0.12,-0.14){134}{\line(0,-1){0.14}} \thicklines \put(10.00,44.60){\line(1,0){68.00}} \thinlines \multiput(78.00,44.60)(-0.12,0.14){134}{\line(0,1){0.14}} \put(62.00,63.40){\line(-1,0){36.00}} \put(13.00,39.00){\vector(2,-1){21.00}} \put(13.00,43.00){\vector(2,1){21.00}} \put(13.00,41.00){\makebox(0,0)[cc]{identify}} \put(46.00,6.00){\makebox(0,0)[cc]{({\bf a})}} \thicklines \put(120.00,20.00){\line(0,1){40.00}} \thinlines \put(120.00,6.00){\makebox(0,0)[cc]{({\bf b})}} \bezier{36}(111.17,57.83)(115.17,60.00)(120.00,60.00) \thicklines \bezier{36}(120.00,60.00)(124.83,60.00)(129.00,58.00) \thinlines \bezier{20}(129.00,58.00)(137.17,54.17)(139.50,44.17) \thicklines \bezier{36}(139.50,44.17)(140.50,40.00)(139.50,35.67) \bezier{104}(139.50,44.33)(126.83,44.50)(129.00,57.83) \bezier{36}(120.00,20.00)(124.83,20.00)(129.00,22.00) \thinlines \bezier{36}(111.17,22.17)(115.17,20.00)(120.00,20.00) \bezier{20}(129.00,22.00)(137.17,25.83)(139.50,35.83) \thicklines \bezier{104}(139.50,35.67)(126.83,35.50)(129.00,22.17) \thinlines \bezier{20}(111.00,58.00)(102.83,54.17)(100.50,44.17) \bezier{36}(100.50,44.17)(99.50,40.00)(100.50,35.67) \bezier{104}(100.50,44.33)(113.17,44.50)(111.00,57.83) \bezier{20}(111.00,22.00)(102.83,25.83)(100.50,35.83) \bezier{104}(100.50,35.67)(113.17,35.50)(111.00,22.17) \put(71.00,41.00){\vector(-3,-1){12.00}} \put(72.00,41.00){\makebox(0,0)[lc]{H$_3$}} \put(110.00,51.00){\vector(3,-4){0.2}} \bezier{72}(110.00,64.00)(104.00,58.33)(110.00,51.00) \put(120.00,64.00){\makebox(0,0)[cc]{identify}} \put(130.00,51.00){\vector(-3,-4){0.2}} \bezier{72}(130.00,64.00)(136.00,58.33)(130.00,51.00) \bezier{156}(104.33,27.33)(120.00,39.33)(135.67,27.33) \put(145.00,20.50){\vector(-4,3){8.00}} \put(145.00,20.50){\makebox(0,0)[lc]{H$_2$}} \end{picture} \caption{Initial state for an AdS spacetime containing three black holes. ({\bf a}) Representation by doubling a region on the Poincar\'e disk. The top and bottom surfaces are to be glued together along pairs of heavily drawn curves, such as the pair labeled ``identify". The resulting topology is that of a pair of pants, with the waist and the legs flaring out to infinity at the limit circle. The heavier part of the curve H$_3$ becomes a closed geodesic at the narrowest part of a leg. ({\bf b}) The fundamental region, obtained from one of the regions of part ({\bf a}) by adding its reflection about the geodesic labeled ``identify" in ({\bf a}). In ({\bf b}) only two boundaries remain to be identified; the top pair are so labeled. For the bottom pair the minimal connecting curve H$_2$ is shown} \end{figure} The topology of the resulting space may be easiest to see in the doubling picture: there are $k$ asymptotically AdS regions, which can be regarded as $k$ punctures (``pants' legs") on a 2-sphere. With each asymptotic region there is associated a horizon, namely the geodesic normal to the corresponding adjacent boundaries of the original region (because it is normal it will become a smooth, circular, minimal geodesic after the doubling). On the outside of each horizon the geometry is the same as that obtained from the isometry corresponding to that horizon alone, so it is exactly the exterior of a BTZ black hole geometry. Therefore the whole space contains $k$ black holes, joined together inside each hole's horizon. \subsubsection{Parameters} The time-symmetric (zero angular momentum) BTZ black hole in AdS space of a given cosmological constant is described by a single parameter, the mass $m$. For an initial state of several black holes we have analogously the several masses, and in addition the relative positions of the black holes. These are however not all independent. Consider a $k$ black hole initial state obtained by doubling a simply-connected region bounded by $k$ non-intersecting geodesics. Find the $k$ minimal geodesic segments $\sigma_i$ between adjacent geodesics.\footnote{Two geodesics of a set are adjacent if each has an end point (at infinity) such that between those end points there is no end point of any other geodesic of the set.} The parts $s_i$ of the original geodesics between the endpoints of those segments, together with the segments $\sigma_i$ themselves, form a geodesic $2k$-gon with right-angle corners. Clearly the $\sigma_i$ are half the horizon size and hence a measure of the masses, and the $s_i$ may be considered a measure of the distances between the black holes. If $2k-3$ of the sides of a $2k$-gon are given, then the geodesics that will form the $2k-2$ side (orthogonal at the end of the $2k-3$ side) and the $2k$ side (orthogonal at the end of the first side) are well-defined. They have a unique common normal geodesic that forms the $2k-1$ side, hence the whole polygon is uniquely defined. Thus only $2k-3$ of the $2k$ numbers measuring the masses and the distances of this type of multi-black-hole are independent. In the case $k=3$ (corresponding to a geodesic hexagon) one can show that alternating sides (either the three masses or the three distances) can be {\it arbitrarily} chosen. Higher $2k$-gons can be divided by geodesics into hexagons, so at least all the masses (or all the distances) can be chosen arbitrarily. (The remaining $k-3$ parameters may have to satisfy inequalities.) Composing the $2k$-gon out of geodesic hexagons means, for the doubled surface, that the multi-black-hole geometry is made out of $k-2$ three-black-hole geometries with $2k-6$ of the asymptotic AdS regions removed and the horizons glued together pairwise. In the five-black-hole example of Fig.~9 the three-black-hole parts are labeled 1, 2, and 3. One asymptotic AdS regions was removed from 1 and 3, and two such regions are missing from 2. The geometries obtained by doubling this are however not the most general time-symmetric five-black-hole configuration. For example, in Fig.~9 the curve separating regions 2 and 3 is a closed geodesic. If we cut and re-glue after a hyperbolic isometry along this geodesic the geometry is still smooth; the operation amounts to rotating the top and bottom part of Fig.~9b with respect to each other, as indicated by the arrows. (In general we can make $k-3$ such re-identifications.) That the result is in general different after this rotation is shown, for example, by the change in angle between the boundary geodesic and another closed geodesic which, before the rotation, is indicated by the dotted line in Fig.~9a. \begin{figure} \unitlength 0.89mm \linethickness{0.4pt} \begin{picture}(141.55,62.66)(10,0) \bezier{30}(10.44,35.00)(10.44,48.71)(22.73,56.49) \bezier{30}(22.73,56.49)(35.00,62.63)(47.27,56.49) \bezier{30}(47.27,56.49)(59.56,48.71)(59.56,35.00) \bezier{30}(10.44,35.00)(10.44,21.29)(22.73,13.51) \bezier{30}(22.73,13.51)(35.00,7.37)(47.27,13.51) \bezier{30}(47.27,13.51)(59.56,21.29)(59.56,35.00) \put(10.70,32.50){\oval(3.90,3.40)[r]} \put(10.70,37.50){\oval(3.90,3.40)[r]} \bezier{74}(11.17,28.67)(18.33,30.33)(24.33,24.50) \bezier{70}(24.33,24.50)(29.00,18.67)(27.67,11.50) \bezier{74}(58.83,28.67)(51.67,30.33)(45.67,24.50) \bezier{70}(45.67,24.50)(41.00,18.67)(42.33,11.50) \bezier{74}(58.83,41.33)(51.67,39.67)(45.67,45.50) \bezier{70}(45.67,45.50)(41.00,51.33)(42.33,58.50) \bezier{74}(11.17,41.33)(18.33,39.67)(24.33,45.50) \bezier{70}(24.33,45.50)(29.00,51.33)(27.67,58.50) \bezier{88}(104.00,60.00)(111.67,56.00)(101.00,48.00) \bezier{80}(101.00,48.00)(91.00,45.00)(93.00,54.00) \bezier{32}(89.00,51.00)(92.67,51.00)(90.00,47.00) \bezier{32}(90.00,47.00)(86.67,43.33)(85.00,46.00) \bezier{60}(85.00,41.00)(87.00,44.00)(93.00,43.00) \bezier{80}(93.00,43.00)(99.00,40.33)(97.00,25.00) \bezier{68}(97.00,25.00)(96.00,15.00)(91.00,14.00) \bezier{150}(136.00,39.00)(128.67,35.00)(118.00,47.00) \bezier{120}(118.00,47.00)(112.00,56.33)(118.00,60.00) \bezier{68}(107.00,61.77)(101.00,60.00)(107.00,58.23) \bezier{44}(107.00,58.23)(110.67,57.44)(116.00,58.52) \bezier{52}(116.00,58.52)(120.00,60.00)(116.00,61.48) \bezier{52}(116.00,61.48)(111.00,62.66)(106.00,61.48) \thicklines \bezier{68}(97.00,37.00)(100.00,33.00)(111.00,37.00) \bezier{72}(111.00,37.00)(121.67,42.00)(118.00,47.00) \bezier{92}(116.00,25.00)(129.00,32.33)(135.50,25.00) \bezier{80}(109.00,11.00)(105.67,18.00)(116.00,25.00) \bezier{68}(134.98,38.30)(131.95,32.00)(134.98,25.70) \bezier{44}(134.98,25.70)(136.83,22.90)(139.53,26.75) \bezier{52}(139.53,26.75)(141.55,32.00)(139.53,37.25) \bezier{52}(139.53,37.25)(137.00,41.45)(134.47,37.25) \bezier{64}(91.00,14.00)(89.67,9.22)(98.00,7.55) \bezier{72}(98.00,7.55)(107.00,6.62)(109.00,11.00) \bezier{24}(95.00,47.30)(96.00,45.00)(93.00,43.00) \thinlines \put(92.00,46.00){\makebox(0,0)[cc]{1}} \put(107.00,44.00){\makebox(0,0)[cc]{2}} \put(111.00,30.00){\makebox(0,0)[cc]{3}} \put(24.50,24.50){\line(1,1){21.00}} \bezier{94}(20.00,42.00)(24.00,35.00)(20.00,28.00) \put(16.00,35.00){\makebox(0,0)[cc]{1}} \put(32.00,42.00){\makebox(0,0)[cc]{2}} \put(41.00,30.00){\makebox(0,0)[cc]{3}} \bezier{25}(83.17,41.39)(82.43,43.50)(83.17,45.61) \bezier{25}(83.17,45.61)(84.29,47.02)(85.41,45.61) \bezier{25}(85.41,45.61)(86.28,43.50)(85.29,41.15) \bezier{20}(85.29,41.15)(84.17,39.98)(83.17,41.39) \bezier{30}(88.33,52.00)(88.17,50.33)(90.83,51.83) \bezier{40}(90.83,51.83)(94.67,54.83)(91.83,54.33) \bezier{30}(91.83,54.33)(89.83,53.83)(88.33,52.00) \put(113.33,40.00){\vector(2,1){0.2}} \bezier{36}(105.00,37.00)(109.67,38.00)(113.33,40.00) \put(103.33,33.67){\vector(-4,-1){0.2}} \bezier{44}(113.33,36.67)(107.67,34.00)(103.33,33.67) \put(35.00,3.00){\makebox(0,0)[cc]{({\bf a})}} \put(110.00,3.00){\makebox(0,0)[cc]{({\bf b})}} \thicklines \bezier{15}(25.00,45.00)(35.00,35.00)(45.00,25.00) \end{picture} \caption{A five-black-hole time-symmetric initial state is obtained by doubling the region on the Poincar\'e disk in ({\bf a}). Part ({\bf b}) shows a somewhat fanciful picture of the result of the doubling, cut off at the flare-outs, which should extend to infinity.} \end{figure} The $2k-3$ distance parameters and the $k-3$ rotation angles describe a $3k-6$-dimensional space of $k$-black-hole geometries. Equivalently we may say that a $k$-black-hole initial state is given by a fundamental domain bounded by $2k-2$ geodesics to be identified in pairs by $k-1$ M\"obius transformations. Since each M\"obius transformation depends on 3 parameters, and the whole fundamental domain can be moved by another M\"obius transformation, the number of free parameters is $3k-6$. Such a space of geometries is known as a Teichm\"uller space, and the length and twist parameters are known as Fenchel-Nielsen coordinates on this space \cite{BP}. Instead of cutting and re-gluing along closed geodesics as in Fig.~9 one can do this operation on the identification geodesics used in the doubling procedure. For example, in Fig.~8a on the pair of geodesics marked ``identify" one can identify each point on the bottom geodesic with one that is moved by a constant distance along the top geodesic. For the fundamental region this means the following: so far, whenever two identification geodesics on the boundary of the fundamental domain were to be identified, it was done by the unique hyperbolic transformation along the minimal normal geodesic between the identification geodesics. If we follow this transformation by a hyperbolic isometry along one of the identification geodesics, the two geodesics will still fit together, and the identified surface will be smooth but with a difference in global structure (like that produced by the re-gluing in Fig.~9). Of course the two transformations combine into one, and conversely any isometry that maps one identification geodesic into another can be decomposed into a ``move" along the normal geodesic, and a ``shift" along a identification geodesic. Since each hyperbolic transformation is a Lorentz transformation in the embedding picture (Fig.~5) the combination is again hyperbolic, so no finite fixed points (singularities) occur in this more general identification process. If we identify with a non-zero shift, there is of course still a minimal geodesic between the two identified geodesics, but it is no longer orthogonal to those geodesics. Nevertheless the identified geometry is that of a black hole. To make the correspondence to the $\phi\rightarrow\phi+2\pi$ identification of Eq (\ref{schwmet}) one would have to change the identification geodesics to be normal to the minimal one (which can complicate the fundamental domain). \subsubsection{Fixed points} It is useful to understand the fixed points at infinity (the limiting circle of the Poincar\'e disk) of the identifications that glue a black hole geometry out of a fundamental domain of AdS space. The fixed points are directly related to the minimal geodesics associated with the identification, and they can indicate whether we have a black hole or not: there must be open sets free of fixed points if the initial data is to be asymptotically AdS. We know that the identifications can have some fixed points at infinity, but if the fixed points cover all of infinity, there is no place left for an asymptotically AdS region, and the space is not a black hole space. Thus even in the relatively simple time-symmetric case it is useful to understand the tiling and the fixed points of the M\"obius transformation associated with the identifications. As an example, consider again the three-black-hole case. Let $a$ and $b$ be the identifications of the top and the bottom pair of geodesics of a figure like 8b. Then the free group generated by these, that is, any ``word" formed from $a,\,b$ and their inverses $A,\,B$ is also an identification. Since the identified geometry is everywhere smooth, none of these can have a fixed point in the finite part of the disk, so all fixed points must lie on the limit circle. The pattern of fixed points is characteristic of the identifications and constitutes a kind of hologram \cite{Suss} of the multi-black-hole spacetime. \begin{figure} \unitlength 1.00mm \linethickness{0.4pt} \begin{picture}(69.85,68.58)(20,15 \bezier{49}(20.15,50.00)(20.15,66.66)(35.08,76.12) \bezier{48}(35.08,76.12)(50.00,83.58)(64.92,76.12) \bezier{49}(64.92,76.12)(79.85,66.66)(79.85,50.00) \bezier{49}(20.15,50.00)(20.15,33.34)(35.08,23.88) \bezier{48}(35.08,23.88)(50.00,16.42)(64.92,23.88) \bezier{49}(64.92,23.88)(79.85,33.34)(79.85,50.00) \bezier{88}(55.19,20.56)(53.33,30.93)(61.11,39.07) \bezier{84}(61.11,39.07)(68.70,46.48)(79.44,44.81) \bezier{32}(59.25,21.67)(58.17,24.83)(61.50,27.33) \bezier{32}(61.50,27.33)(65.58,29.00)(68.00,26.00) \bezier{12}(69.75,27.42)(69.00,28.83)(70.00,29.92) \bezier{32}(78.33,40.75)(75.17,41.83)(72.67,38.50) \bezier{32}(72.67,38.50)(71.00,34.42)(74.00,32.00) \bezier{12}(72.58,30.25)(71.17,31.00)(70.08,30.00) \put(65.67,24.33){\circle{1.67}} \put(70.67,28.33){\circle{1.00}} \put(62.00,22.67){\circle*{1.67}} \bezier{8}(63.50,23.00)(63.50,24.17)(64.33,23.33) \put(67.83,38.83){\makebox(0,0)[cc]{$a$}} \put(63.00,25.80){\makebox(0,0)[cc]{$a^2$}} \put(74.67,37.00){\makebox(0,0)[cc]{$ab$}} \bezier{88}(79.44,55.19)(69.07,53.33)(60.93,61.11) \bezier{84}(60.93,61.11)(53.52,68.70)(55.19,79.44) \bezier{32}(78.33,59.25)(75.17,58.17)(72.67,61.50) \bezier{32}(72.67,61.50)(71.00,65.58)(74.00,68.00) \bezier{12}(72.58,69.75)(71.17,69.00)(70.08,70.00) \bezier{32}(59.25,78.33)(58.17,75.17)(61.50,72.67) \bezier{32}(61.50,72.67)(65.58,71.00)(68.00,74.00) \bezier{12}(69.75,72.58)(69.00,71.17)(70.00,70.08) \bezier{88}(44.81,79.44)(46.67,69.07)(38.89,60.93) \bezier{84}(38.89,60.93)(31.30,53.52)(20.56,55.19) \bezier{32}(40.75,78.33)(41.83,75.17)(38.50,72.67) \bezier{32}(38.50,72.67)(34.42,71.00)(32.00,74.00) \bezier{12}(30.25,72.58)(31.00,71.17)(30.00,70.08) \bezier{32}(21.67,59.25)(24.83,58.17)(27.33,61.50) \bezier{32}(27.33,61.50)(29.00,65.58)(26.00,68.00) \bezier{12}(27.42,69.75)(28.83,69.00)(29.92,70.00) \bezier{88}(20.56,44.81)(30.93,46.67)(39.07,38.89) \bezier{84}(39.07,38.89)(46.48,31.30)(44.81,20.56) \bezier{32}(21.67,40.75)(24.83,41.83)(27.33,38.50) \bezier{32}(27.33,38.50)(29.00,34.42)(26.00,32.00) \bezier{12}(27.42,30.25)(28.83,31.00)(29.92,30.00) \bezier{32}(40.75,21.67)(41.83,24.83)(38.50,27.33) \bezier{32}(38.50,27.33)(34.42,29.00)(32.00,26.00) \bezier{12}(30.25,27.42)(31.00,28.83)(30.00,29.92) \put(65.00,65.00){\makebox(0,0)[cc]{$B$}} \put(35.00,65.00){\makebox(0,0)[cc]{$b$}} \put(35.00,35.00){\makebox(0,0)[cc]{$A$}} \put(75.00,64.00){\makebox(0,0)[cc]{$B\!A$}} \put(73.00,72.00){\makebox(0,0)[cc]{$Ba$}} \put(63.30,75.00){\makebox(0,0)[cc]{$B^2$}} \put(37.00,75.00){\makebox(0,0)[cc]{$b^2$}} \put(30.00,71.00){\makebox(0,0)[rb]{$bA$}} \put(25.50,63.50){\makebox(0,0)[cc]{$ba$}} \put(25.00,37.00){\makebox(0,0)[cc]{$A\!B$}} \put(29.50,28.00){\makebox(0,0)[rc]{$Ab$}} \put(37.00,25.00){\makebox(0,0)[cc]{$A^2$}} \put(38.00,22.67){\circle*{1.67}} \put(72.00,29.33){\circle*{1.00}} \put(75.67,34.33){\circle*{1.67}} \put(77.33,38.00){\circle{1.67}} \put(77.33,62.00){\circle{1.67}} \thicklines \bezier{64}(38.00,23.00)(42.04,30.00)(50.00,30.00) \bezier{64}(50.00,30.00)(58.15,30.00)(62.00,23.00) \bezier{44}(65.17,25.17)(65.33,30.67)(70.17,28.67) \bezier{44}(74.83,34.83)(69.33,34.67)(71.33,29.83) \bezier{64}(77.00,62.00)(70.00,57.96)(70.00,50.00) \bezier{64}(70.00,50.00)(70.00,41.85)(77.00,38.00) \bezier{64}(23.00,38.00)(30.00,42.04)(30.00,50.00) \bezier{64}(30.00,50.00)(30.00,58.15)(23.00,62.00) \bezier{64}(62.00,77.00)(57.96,70.00)(50.00,70.00) \bezier{64}(50.00,70.00)(41.85,70.00)(38.00,77.00) \put(50.00,30.00){\vector(1,0){1.00}} \put(70.00,50.00){\vector(0,1){1.00}} \put(50.00,70.00){\vector(-1,0){1.00}} \put(30.00,50.00){\vector(0,-1){1.00}} \thinlines \put(51.00,31.00){\makebox(0,0)[cc]{$a$}} \put(49.00,69.00){\makebox(0,0)[cc]{$b$}} \put(69.00,50.00){\makebox(0,0)[rc]{$B\!A$}} \put(31.00,50.00){\makebox(0,0)[lc]{$A\!B$}} \put(66.70,33.30){\vector(1,-1){4.00}} \put(67.00,33.00){\makebox(0,0)[rb]{$aB$}} \put(62.00,17.00){\vector(0,1){5.00}} \put(62.40,17.00){\makebox(0,0)[ct]{$a^n$}} \put(64.00,20.00){\vector(0,1){3.00}} \put(64.00,20.80){\makebox(0,0)[lt]{$a^2B$}} \put(71.00,24.00){\vector(-1,0){4.00}} \put(71.00,24.00){\makebox(0,0)[lc]{$a(ab)^n$}} \put(76.00,28.00){\vector(-1,0){5.00}} \put(76.00,28.00){\makebox(0,0)[lb]{$a(B\!A)^n$}} \put(79.00,38.00){\makebox(0,0)[lc]{$(ab)^n$}} \put(79.00,62.00){\makebox(0,0)[lc]{$(B\!A)^n$}} \put(38.00,21.00){\makebox(0,0)[ct]{$A^n$}} \put(77.00,34.00){\makebox(0,0)[lc]{$ab^n$}} \put(50.00,50.00){\makebox(0,0)[cc]{\bf 1}} \end{picture} \caption{Part of the tiling of the Poincar\'e disk obtained by ``unwrapping" a three-black-hole initial geometry as in Fig.~8. A fundamental domain {\bf 1} is imaged by combinations of identification maps $a$ and $b$ and their inverses $A=a^{-1},\,B=b^{-1}$. Repeating $n$ times a map such as $ab$ leads to a point $(ab)^n$ on the limit circle, in the limit $n\rightarrow\infty$. Some geodesics (``horizons") connecting such a limit point and its inverse limit (such as $a(ab)^n$ and $a(ab)^{-n}=a(BA)^n$) are shown as heavy curves} \label{fig10} \end{figure} In Fig.~10 the initial fundamental domain is denoted by {\bf 1}. The identifications are given by hyperbolic M\"obius transformations $a,\,b$, with inverses $A,\,B$ that connect the top and the bottom boundaries, respectively. Any ``word" made up of these four letters is, first, also an identification. Secondly each word can be used to label a tile, because each tile is some image, $a${\bf 1}, $A${\bf 1}, $aB${\bf 1},\dots of the initial domain {\bf 1}, shown simply as $a,\,A,\,aB,$ \dots in the Figure. Finally, there is a closed minimal geodesic associated with each pair of identified boundaries, hence each word also corresponds to a geodesic.\footnote{In this connection we regard a word, its inverse, and the permuted word as equal, in order to have a unique correspondence to geodesics; see \cite{sor}.} (For example, $Ba$ connects $(Ab)^n$ to $(Ba)^n$.) Horizons are special geodesics that bound asymptotically AdS regions. Some of these are shown by the heavy curves. The ones that cut through the basic domain are labeled by the isometries that leave them invariant, $a,\,b$, and $AB=BA$. The words for the other horizons are obtained from these by conjugation, for example the horizon connecting the points labeled $a(ab)^n$ and $a(BA)^n$ is ``called" by the word $a(BA)A$. Every words is a hyperbolic isometry, hence has two fixed points on the Poincar\'e limit circle. We can find the fixed points by applying the word (or its inverse) many times to any finite region, because in the limit the images will converge to a point on the limit circle (see, for example, the equidistant curves in Fig.~2). Some of these fixed points are shown by open and by filled circles in the Figure, and labeled by an $n$th power, where the limit $n\rightarrow\infty$ is understood. The two fixed points of a hyperbolic transformation define a geodesic that ends at them, and that is the minimal geodesic along which the transformation acts. Because the infinity side of a horizon is isometric to the asymptotic region of a single black hole, there are no fixed points on that side of the horizon. (Cf.~Fig.~2, where the only fixed points of the horizon isometry $\phi\rightarrow\phi+$ const are on the horizon $r=\ell$.) Between two different horizons (between open and filled circles of the Figure) there will however be further horizons, with fixed points at their ends. Thus the set of fixed points for multi-black-holes has the fractal structure of a Cantor set. By contrast, for some identifications the fixed points are everywhere dense on the limit circle. This happens, for example, if we try to build, by analogy to the multi-black-hole construction, a geometry containing three $m=0$ black holes. The tiles, analogs of those of Fig.~10, would be ``ideal" quadrilaterals, that is, each tile is a geodesic polygon whose four corners lie on the limit circle. This space is smooth and contains three ends of the type shown in the second column of Fig.~7 (instead of the ``legs" in such pictures as Fig.~9); but since there is then no fixed-point-free region on the limit circle, this space is not asymptotically AdS and hence does not contain BTZ-type black holes. \subsection{Other Topologies} It is well known that time-symmetric AdS initial states, that is, spaces of constant negative curvature, admit a large variety of topologies. In the context of (orientable) black hole spaces one can construct all of these out of pieces of the three-black-hole space as in Fig.~7. These pieces are: three BTZ-exteriors, that is, the regions outside each of the three horizons; and one region interior to the horizons. The interior piece is sometimes called the ``convex core" or ``trousers."\footnote{Previously we have used the image of flared pants' legs for the asymptotically AdS regions, which need to be cut off to obtain the core, so it would be more consistent to call the latter ``cut-offs" or ``shorts," but we will use ``trousers."} Fig.~11 shows how other topologies can be constructed out of these pieces. \begin{figure} \unitlength 1.00mm \linethickness{0.4pt} \begin{picture}(117.00,70.00)(10,0) \put(30.00,30.00){\circle{10.00}} \put(27.00,46.00){\line(-1,0){8.00}} \put(19.00,46.00){\line(-1,1){6.00}} \thicklines \put(13.00,52.00){\line(1,0){17.00}} \thinlines \put(33.00,46.00){\line(1,0){8.00}} \put(41.00,46.00){\line(1,1){6.00}} \thicklines \put(47.00,52.00){\line(-1,0){17.00}} \put(30.00,49.00){\makebox(0,0)[cc]{$\cal E$}} \bezier{32}(19.78,30.00)(19.78,34.14)(22.71,37.07) \bezier{32}(39.78,30.00)(39.78,34.14)(36.85,37.07) \bezier{32}(19.78,30.00)(19.78,25.86)(22.71,22.93) \bezier{32}(22.71,22.93)(25.64,20.00)(29.78,20.00) \bezier{32}(39.78,30.00)(39.78,25.86)(36.85,22.93) \bezier{32}(36.85,22.93)(33.92,20.00)(29.78,20.00) \bezier{8}(27.13,44.00)(27.13,44.53)(27.95,44.91) \bezier{8}(27.95,44.91)(28.77,45.29)(29.94,45.29) \bezier{8}(32.75,44.00)(32.75,44.53)(31.93,44.91) \bezier{8}(31.93,44.91)(31.10,45.29)(29.94,45.29) \thinlines \bezier{8}(27.13,44.00)(27.13,43.47)(27.95,43.09) \bezier{8}(27.95,43.09)(28.77,42.71)(29.94,42.71) \bezier{8}(32.75,44.00)(32.75,43.47)(31.93,43.09) \bezier{8}(31.93,43.09)(31.10,42.71)(29.94,42.71) \bezier{24}(32.75,44.00)(32.75,47.00)(37.00,48.40) \bezier{24}(27.21,44.00)(27.21,47.00)(22.96,48.40) \thicklines \bezier{8}(27.13,40.40)(27.13,40.93)(27.95,41.31) \bezier{8}(27.95,41.31)(28.77,41.69)(29.94,41.69) \bezier{8}(32.75,40.40)(32.75,40.93)(31.93,41.31) \bezier{8}(31.93,41.31)(31.10,41.69)(29.94,41.69) \bezier{16}(36.84,37.07)(35.64,38.40)(33.38,39.60) \bezier{4}(33.38,39.60)(32.98,39.87)(32.71,40.53) \bezier{16}(22.84,37.07)(24.04,38.40)(26.31,39.60) \bezier{4}(26.31,39.60)(26.71,39.87)(26.98,40.53) \thinlines \bezier{16}(30.18,24.93)(31.24,25.07)(31.38,22.40) \bezier{12}(31.38,22.40)(30.98,20.53)(30.04,20.13) \put(97.00,64.00){\line(-1,0){8.00}} \put(89.00,64.00){\line(-1,1){6.00}} \thicklines \put(83.00,70.00){\line(1,0){17.00}} \put(117.00,70.00){\line(-1,0){17.00}} \thinlines \put(103.00,64.00){\line(1,0){8.00}} \put(111.00,64.00){\line(1,1){6.00}} \put(100.00,67.00){\makebox(0,0)[cc]{$\cal E$}} \thicklines \bezier{8}(97.13,62.00)(97.13,62.53)(97.95,62.91) \bezier{8}(97.95,62.91)(98.77,63.29)(99.94,63.29) \bezier{8}(102.75,62.00)(102.75,62.53)(101.93,62.91) \bezier{8}(101.93,62.91)(101.10,63.29)(99.94,63.29) \thinlines \bezier{8}(97.13,62.00)(97.13,61.47)(97.95,61.09) \bezier{8}(97.95,61.09)(98.77,60.71)(99.94,60.71) \bezier{8}(102.75,62.00)(102.75,61.47)(101.93,61.09) \bezier{8}(101.93,61.09)(101.10,60.71)(99.94,60.71) \bezier{24}(102.75,62.00)(102.75,65.00)(107.00,66.40) \bezier{24}(97.21,62.00)(97.21,65.00)(92.96,66.40) \put(100.00,51.00){\circle{8.00}} \thicklines \bezier{8}(97.13,59.00)(97.13,59.53)(97.95,59.91) \bezier{8}(97.95,59.91)(98.77,60.29)(99.94,60.29) \bezier{8}(102.75,59.00)(102.75,59.53)(101.93,59.91) \bezier{8}(101.93,59.91)(101.10,60.29)(99.94,60.29) \bezier{8}(96.84,42.05)(96.84,42.64)(97.74,43.05) \bezier{8}(97.74,43.05)(98.64,43.47)(99.93,43.47) \bezier{8}(103.03,42.05)(103.03,42.64)(102.13,43.05) \bezier{8}(102.13,43.05)(101.21,43.47)(99.93,43.47) \thinlines \bezier{8}(96.84,42.05)(96.84,41.47)(97.74,41.05) \bezier{8}(97.74,41.05)(98.64,40.63)(99.93,40.63) \bezier{8}(103.03,42.05)(103.03,41.47)(102.13,41.05) \bezier{8}(102.13,41.05)(101.21,40.63)(99.93,40.63) \thicklines \bezier{12}(102.81,58.97)(102.81,57.90)(104.09,57.44) \bezier{36}(104.09,57.44)(107.94,55.38)(107.94,51.03) \bezier{36}(104.09,44.40)(107.94,46.67)(107.94,51.03) \bezier{20}(104.09,44.40)(103.00,44.00)(103.00,42.26) \bezier{12}(97.19,58.97)(97.19,57.90)(95.91,57.44) \bezier{36}(95.91,57.44)(92.06,55.38)(92.06,51.03) \bezier{36}(95.91,44.40)(92.06,46.67)(92.06,51.03) \bezier{20}(95.91,44.40)(97.00,44.00)(97.00,42.26) \thinlines \put(100.00,28.00){\circle{10.00}} \thicklines \bezier{32}(89.78,28.00)(89.78,32.14)(92.71,35.07) \bezier{32}(109.78,28.00)(109.78,32.14)(106.85,35.07) \bezier{32}(89.78,28.00)(89.78,23.86)(92.71,20.93) \bezier{32}(92.71,20.93)(95.64,18.00)(99.78,18.00) \bezier{32}(109.78,28.00)(109.78,23.86)(106.85,20.93) \bezier{32}(106.85,20.93)(103.92,18.00)(99.78,18.00) \bezier{8}(97.13,38.40)(97.13,38.93)(97.95,39.31) \bezier{8}(97.95,39.31)(98.77,39.69)(99.94,39.69) \bezier{8}(102.75,38.40)(102.75,38.93)(101.93,39.31) \bezier{8}(101.93,39.31)(101.10,39.69)(99.94,39.69) \bezier{16}(106.84,35.07)(105.64,36.40)(103.38,37.60) \bezier{4}(103.38,37.60)(102.98,37.87)(102.71,38.53) \bezier{16}(92.84,35.07)(94.04,36.40)(96.31,37.60) \bezier{4}(96.31,37.60)(96.71,37.87)(96.98,38.53) \thinlines \bezier{16}(100.18,22.93)(101.24,23.07)(101.38,20.40) \bezier{12}(101.38,20.40)(100.98,18.53)(100.04,18.13) \bezier{20}(92.14,50.97)(94.02,52.0)(95.91,50.97) \bezier{20}(107.79,50.97)(105.91,52.0)(104.02,50.97) \put(100.00,57.00){\makebox(0,0)[cc]{$\cal T$}} \put(100.00,44.00){\makebox(0,0)[cb]{$\cal T$}} \put(100.00,36.00){\makebox(0,0)[cc]{$\cal T$}} \put(30.00,38.00){\makebox(0,0)[cc]{$\cal T$}} \put(51.00,42.00){\vector(-1,0){16.00}} \put(52.00,42.00){\makebox(0,0)[lc]{Minimal circles}} \put(76.00,43.00){\vector(2,1){15.00}} \put(75.74,41.49){\vector(1,0){19.00}} \put(32.00,19.00){\vector(-1,1){0.2}} \bezier{172}(52.00,40.00)(40.00,10.51)(32.00,19.00) \put(30.00,10.00){\makebox(0,0)[cc]{({\bf a})}} \put(100.00,10.00){\makebox(0,0)[cc]{({\bf b})}} \end{picture} \caption{Examples of the class of spaces considered here, constructed by sewing together one or several trousers and one asymptotic region. The latter looks asymptotically flat in this topological picture, but metrically it has constant negative curvature everywhere, just like the trousers} \end{figure} The resulting geometry is smooth if we choose the freely specifiable mass parameters of each exterior or core to match those of its neighbors at the connection horizons\footnote{The horizons along which the legs were cut off from the cores may no longer be horizons of the space-time if the cores are re-assembled differently. Nonetheless, in the present section we will still call them by that name.}: the intrinsic geometries then match, and the extrinsic geometries match because the horizons are geodesics. Conversely we can decompose a given $k$-black-hole initial geometry of genus $g$ into asymptotic regions and trousers by cutting it along minimal circles of different and non-trivial homotopy types. We can choose $3g+2k-3$ such circles that divide the space into $k$ exteriors $\cal E$ and $2g+k-2$ trousers $\cal T$. Fig.~11 illustrates the construction for $k=1$ and $g=1$ (left) resp.\ $g=2$ (right). A fundamental region on the Poincar\'e disk, and hence the M\"obius transformations that implement the identifications, can be constructed for these spaces in a similar way, by putting together geodesic, right-angle octagons representing trousers and analogous asymptotic regions. For example, the $k=1,\, g=1$ geometry can be represented by identifying two of the horizon geodesics of a trousers octagon and adding an exterior to the third. The resulting fundamental domain is bounded by geodesics, but it is not unique. We can cut it into pieces and re-assemble it in a different way \cite{DBS}, or we can cut the original space along some geodesics (not necessarily those of the trousers decomposition) only until it becomes one simply connected piece. If we can lay these geodesics so that they start and end at infinity and therefore do not cross we obtain a simple fundamental region bounded only by complete geodesics. Figure 12a shows the two geodesic cuts necessary for the case of our example of Fig.~11a, and the fundamental domain so obtained is seen in Fig.~12b. The pattern of tiling for this case is identical to that of Fig.~10, but the labeling is different. For example, rather than three horizon words there is only one, $abAB$, corresponding to the existence of only one horizon in the identified manifold. \begin{figure} \unitlength 0.80mm \linethickness{0.4pt} \begin{picture}(121.27,66.23) \bezier{44}(76.73,33.00)(76.73,45.43)(87.87,52.49) \bezier{40}(87.87,52.49)(99.00,58.06)(110.13,52.49) \bezier{44}(110.13,52.49)(121.27,45.43)(121.27,33.00) \bezier{44}(76.73,33.00)(76.73,20.57)(87.87,13.51) \bezier{40}(87.87,13.51)(99.00,7.94)(110.13,13.51) \bezier{44}(110.13,13.51)(121.27,20.57)(121.27,33.00) \thicklines \bezier{88}(117.28,45.81)(110.69,40.84)(110.69,33.00) \bezier{88}(117.28,20.19)(110.69,25.16)(110.69,33.00) \bezier{88}(80.72,45.81)(87.31,40.84)(87.31,33.00) \bezier{88}(80.72,20.19)(87.31,25.16)(87.31,33.00) \bezier{64}(90.66,53.65)(93.28,47.92)(99.00,47.92) \bezier{64}(107.34,53.65)(104.72,47.92)(99.00,47.92) \bezier{64}(90.66,12.35)(93.28,18.08)(99.00,18.08) \bezier{64}(107.34,12.35)(104.72,18.08)(99.00,18.08) \thinlines \put(99.00,18.08){\vector(0,1){29.85}} \put(87.43,33.00){\vector(1,0){23.13}} \put(40.00,28.67){\circle{12.00}} \thicklines \bezier{64}(21.67,30.00)(21.67,40.23)(30.84,46.04) \bezier{64}(49.16,46.04)(58.33,40.23)(58.33,30.00) \bezier{64}(21.67,30.00)(21.67,19.77)(30.84,13.96) \bezier{60}(30.84,13.96)(40.00,9.38)(49.16,13.96) \bezier{64}(49.16,13.96)(58.33,19.77)(58.33,30.00) \bezier{48}(30.84,46.04)(36.67,50.00)(37.00,54.00) \bezier{68}(37.00,54.00)(37.67,63.00)(28.00,63.00) \bezier{48}(49.16,46.04)(43.33,50.00)(43.00,54.00) \bezier{68}(43.00,54.00)(42.33,63.00)(52.00,63.00) \thinlines \bezier{64}(27.59,63.00)(27.59,61.40)(33.80,60.49) \bezier{60}(33.80,60.49)(40.00,59.77)(46.20,60.49) \bezier{64}(46.20,60.49)(52.41,61.40)(52.41,63.00) \bezier{64}(27.59,63.00)(27.59,64.60)(33.80,65.51) \bezier{60}(33.80,65.51)(40.00,66.23)(46.20,65.51) \bezier{64}(46.20,65.51)(52.41,64.60)(52.41,63.00) \thicklines \bezier{68}(40.00,35.00)(45.00,34.00)(42.00,45.00) \bezier{64}(42.00,45.00)(40.00,55.33)(42.00,60.00) \bezier{14}(40.00,35.00)(34.67,35.00)(37.50,49.17) \bezier{10}(37.50,49.17)(38.83,56.33)(38.17,60.17) \bezier{32}(38.17,60.17)(36.83,64.83)(33.67,65.50) \thinlines \bezier{14}(40.24,22.39)(38.86,22.39)(38.07,19.69) \bezier{10}(38.07,19.69)(37.45,16.98)(38.07,14.28) \bezier{14}(38.07,14.28)(38.86,11.56)(40.24,11.56) \bezier{34}(40.24,22.39)(41.63,22.39)(42.42,19.69) \bezier{30}(42.42,19.69)(43.03,16.98)(42.42,14.28) \bezier{34}(42.42,14.28)(41.63,11.56)(40.24,11.56) \put(95.67,33.33){\makebox(0,0)[cb]{$a$}} \put(99.33,39.67){\makebox(0,0)[lc]{$b$}} \put(42.87,17.00){\makebox(0,0)[lc]{$a$}} \put(46.33,28.33){\makebox(0,0)[lc]{$b$}} \put(44.33,33.00){\vector(-1,1){0.17}} \put(41.76,21.44){\vector(-1,3){0.2}} \put(40.00,3.00){\makebox(0,0)[cc]{({\bf a})}} \put(100.00,3.00){\makebox(0,0)[cc]{({\bf b})}} \end{picture} \caption{Construction of a BTZ exterior with toroidal interior. Rather than cutting the geometry shown in part ({\bf a}) by minimal geodesics as in Fig.~11, the cuts, shown by the heavy lines, are chosen to reach infinity and divide this into four regions. Thus one obtains the fundamental domain shown in part ({\bf b}). The identifications $a$ and $b$ are shown as arrows. The lines of these arrows are also the minimal geodesics between the lines that are to be identified. In the identified manifold these are closed geodesics, as shown in part ({\bf a}). The possible geometries are characterized by the lengths of these two closed geodesics, and the angle between them (shown here as $90^\circ$)} \end{figure} Out of an even number of trousers only one can construct locally AdS initial data that contain no asymptotically AdS region at all. Such compact spaces can be interpreted as closed universes, and for lack of other physical content they can be considered to contain several black holes, associated with the horizons that were glued together in the construction. (Of course these horizons and black holes are only analogies, for example there are no observers that see them as black, i.e.\ of infinite redshift.) Our reasoning about the number of parameters that specify a $k$-black-hole geometry can be generalized to the case that the internal geometry has genus $g$. If we cut off two asymptotic AdS regions and identify the two horizons that go with them, we decrease $k$ by two and increase $g$ by one. The number of parameters does not change: we lose one mass parameter, since the masses of the two horizons that will be identified have to be equal, but we gain one rotation parameter which specifies with what shift the horizons are to be identified. Thus from the formula in section 2.4 we find that the (orientable) time-symmetric initial states, of genus $g$ and $k$ asymptotic AdS regions, form a $(6g+3k-6)$-dimensional Teichm\"uller space. If this number is non-positive, no state of that type is possible. (However, the formula cannot be applied to the time-symmetric BTZ initial state itself: it has one free parameter, the mass $m$, but no integral value of $k$ makes the formula valid; the BTZ state is not a multi-black-hole geometry in the sense of this section.) For example, if we want a single exterior region ($k=1$) we need a genus of at least $g=1$ (Fig.~12). Here the number of parameters is $6g+3k-6=3$, for example the minimal distances (lengths of closed geodesics $a$ and $b$) for each of the two identifications, and the angle between these geodesics. It is clear from the figure that these distances must be large enough, and the angle close enough to a right angle, that an asymptotic region remains in Fig.~12b. (If the geodesics crossed and formed a quadrilateral, there would be an angle deficit at the crossing point, which could be interpreted as a toroidal universe that is not empty, but contains one point particle.) The formula for the number of free parameters tells us that there is no time-symmetric torus ($k=0,\,g=1$) initial state. However, all topologies of higher genus or with at least one asymptotic AdS region do occur; and the spatial torus topology does occur among all locally AdS spacetimes, for example as Eq~(\ref{schwmet}) for $r^2<\ell^2$ with $\phi$ and $t$ periodically identified --- the analog of a closed Kantowski-Sachs universe. \section{Time Development} The identifications used on a time-symmetric surface of AdS space to generate black hole and other initial values have a unique extension to all of AdS space, and thus define a unique time development (even beyond any Cauchy horizon). A fundamental domain in 3-dimensional AdS space can be generated by extending normal timelike geodesics from the geodesic boundaries of the two-dimensional fundamental domain on the initial surface. Due to the negative curvature of AdS space such {\it timelike} geodesics accelerate towards each other and will eventually cross. Such crossing of fundamental domain boundaries is the space-time analog of a conical singularity. A prototype of this is the ``non-Hausdorff singularity" of Misner space \cite{Mis}. Although not a curvature singularity, these points are considered not to be part of the space-time. This in turn provides an end of {\curs I}\ and hence the possibility of a black hole horizon. A metric for the time development of the finite part of any multi-black-hole or multiply-connected time-symmetric initial geometry is provided by Eq (\ref{cmc}) when we replace the expression in the bracket by the initial multi-black-hole metric. The result is a metric adapted to free-fall observers, and it shows that they all reach the singularity after the same proper time, $\tau=\pi\ell/2$, when the $\cos^2$ factor vanishes. (This can be seen geometrically from Fig.~4a, where geodesics are intersections with planes through the origin, and the collapse time is one quarter of the period around the hyperboloid.) But these coordinates do not cover the time development of conformal infinity (cf.\ the dotted curves in Fig.~2). A more complete picture emerges from the continuation of the identification group to AdS spacetime, for example via the embedding of AdS space according to Eq (\ref{ads}). In the embedding of the initial surface in the 3-dimensional Minkowski space $V=0$, each identification corresponds to a Lorentz ``rotation" about some (spacelike) axis $A$. This is uniquely extended to an SO(2,2) ``rotation" of the four-dimensional embedding space by requiring that the $V$-axis also remain invariant; that is, we rotate by the same hyperbolic angle about the $A,V$ plane. This plane intersects the AdS space (\ref{ads}) in a spacelike geodesic of fixed points. All such geodesics from all the identifications are to be considered singularities after the identifications are made, so they are not points in the identified spacetime. Three-dimensional pictures that include conformal infinity and all of the singularities can be had in sausage and in stereographic coordinates, Eqs (\ref{sausge}) and (\ref{stm}). Because all timelike geodesics starting normally on a time-symmetric initial surface collapse together to a point C, all the totally geodesic boundaries of the fundamental domain also meet at C, forming a tent-like structure with a tip at C. Their intersections may be timelike or spacelike. If an intersection is timelike, the sides typically intersect there at a right angle ``corner," and the intersection passes through the initial surface. If the initial geometry is smooth, such intersections are innocuous.\footnote{We have not encountered such corners in our pictures, but they must appear in spaces composed only of trousers, for example in the time development of a $k=0,\,g=2$ surface that can be represented by a right-angled octagon on the Poincar\'e disk, as in Fig.~3b of \cite{DBS}.} Spacelike intersections are called ``folds" of the tent, and they are the geodesics of fixed points, which likewise meet at C.\footnote{The reason that corners can be regular and but folds are not is that four corners can be put together to make a line without angle deficit, but no finite number of folds can eliminate the Misner-space singularity.} \begin{figure} \unitlength 1.00mm \linethickness{0.4pt} \begin{picture}(84.04,55.00)(-7,15) \bezier{30}(20.15,32.00)(20.15,38.59)(35.08,42.34) \bezier{30}(64.92,42.34)(79.85,38.59)(79.85,32.00) \bezier{30}(20.15,32.00)(20.15,25.41)(35.08,21.66) \bezier{40}(35.08,21.66)(50.00,18.71)(64.92,21.66) \bezier{30}(64.92,21.66)(79.85,25.41)(79.85,32.00) \bezier{14}(34.17,60.00)(34.17,63.17)(42.09,64.97) \bezier{20}(42.09,64.97)(50.00,66.39)(57.91,64.97) \bezier{14}(57.91,64.97)(65.83,63.17)(65.83,60.00) \bezier{14}(34.17,60.00)(34.17,56.83)(42.09,55.03) \bezier{20}(42.09,55.03)(50.00,53.61)(57.91,55.03) \bezier{14}(57.91,55.03)(65.83,56.83)(65.83,60.00) \bezier{264}(24.00,35.00)(50.00,55.00)(76.00,35.00) \bezier{56}(39.30,55.79)(46.32,39.30)(56.32,20.53) \bezier{56}(60.70,55.79)(53.68,39.30)(43.68,20.53) \put(50.00,31.00){\circle*{0.00}} \put(49.00,30.00){\circle*{0.00}} \put(50.00,29.00){\circle*{0.00}} \put(49.00,28.00){\circle*{0.00}} \put(50.00,27.00){\circle*{0.00}} \put(48.00,27.00){\circle*{0.00}} \put(49.00,26.00){\circle*{0.00}} \put(47.00,26.00){\circle*{0.00}} \put(48.00,25.00){\circle*{0.00}} \put(50.00,25.00){\circle*{0.00}} \put(49.00,24.00){\circle*{0.00}} \put(47.00,24.00){\circle*{0.00}} \put(46.00,23.00){\circle*{0.00}} \put(48.00,23.00){\circle*{0.00}} \put(50.00,23.00){\circle*{0.00}} \put(49.00,22.00){\circle*{0.00}} \put(47.00,22.00){\circle*{0.00}} \put(45.00,22.00){\circle*{0.00}} \put(46.00,21.00){\circle*{0.00}} \put(48.00,21.00){\circle*{0.00}} \put(50.00,21.00){\circle*{0.00}} \put(44.00,21.00){\circle*{0.00}} \put(47.02,20.18){\circle*{0.00}} \put(48.95,20.18){\circle*{0.00}} \put(51.00,30.00){\circle*{0.00}} \put(52.00,29.00){\circle*{0.00}} \put(51.00,28.00){\circle*{0.00}} \put(52.00,27.00){\circle*{0.00}} \put(51.00,26.00){\circle*{0.00}} \put(53.00,26.00){\circle*{0.00}} \put(52.00,25.00){\circle*{0.00}} \put(54.00,25.00){\circle*{0.00}} \put(51.00,24.00){\circle*{0.00}} \put(53.00,24.00){\circle*{0.00}} \put(54.00,23.00){\circle*{0.00}} \put(52.00,23.00){\circle*{0.00}} \put(51.00,22.00){\circle*{0.00}} \put(53.00,22.00){\circle*{0.00}} \put(55.00,22.00){\circle*{0.00}} \put(54.00,21.00){\circle*{0.00}} \put(52.00,21.00){\circle*{0.00}} \put(56.00,21.00){\circle*{0.00}} \put(52.98,20.18){\circle*{0.00}} \put(51.05,20.18){\circle*{0.00}} \put(54.00,68.00){\vector(-1,-2){3.50}} \put(55.00,68.00){\makebox(0,0)[lc]{C (Center of projection)}} \put(29.00,65.00){\vector(1,-1){4.50}} \put(28.50,65.00){\makebox(0,0)[rc]{Conformal infinity}} \put(20.00,40.00){\vector(1,0){5.0}} \put(19.50,40.00){\makebox(0,0)[rc]{P}} \bezier{36}(23.86,35.09)(21.75,31.75)(24.74,28.25) \bezier{92}(38.25,22.63)(49.82,20.88)(61.40,22.63) \bezier{60}(24.91,28.07)(28.77,24.39)(38.42,22.63) \bezier{36}(76.14,35.09)(78.25,31.75)(75.26,28.25) \bezier{60}(75.09,28.07)(71.23,24.39)(61.58,22.63) \put(50.00,32.63){\line(0,-1){14.56}} \put(15.96,31.75){\line(5,-2){4.91}} \put(24.74,28.25){\line(5,-2){13.68}} \put(44.04,20.53){\line(5,-2){5.96}} \put(15.96,31.75){\line(5,2){4.56}} \bezier{30}(25.61,39.65)(33.16,47.54)(39.30,55.80) \bezier{26}(34.04,60.00)(34.04,50.35)(25.09,39.30) \put(25.79,39.82){\line(-6,-5){9.82}} \put(71.93,47.37){\vector(-3,-2){9.65}} \put(72.28,47.37){\makebox(0,0)[lc]{Initial surface}} \put(84.04,31.75){\line(-5,-2){4.91}} \put(75.26,28.25){\line(-5,-2){13.68}} \put(84.04,31.75){\line(-5,2){4.56}} \bezier{30}(74.39,39.65)(66.84,47.54)(60.70,55.80) \bezier{26}(65.96,60.00)(65.96,50.35)(74.91,39.30) \put(74.21,39.82){\line(6,-5){9.82}} \put(55.96,20.53){\line(-5,-2){5.96}} \thicklines \put(20.88,29.79){\line(5,-2){3.86}} \put(38.42,22.77){\line(5,-2){5.61}} \bezier{16}(20.53,33.68)(19.47,31.75)(20.53,29.82) \bezier{36}(44.00,20.50)(50.00,19.65)(56.00,20.50) \put(44.04,20.70){\line(1,2){5.88}} \put(49.91,32.46){\line(0,1){27.54}} \bezier{16}(20.53,33.51)(22.11,34.39)(23.68,35.09) \put(20.53,33.51){\line(5,6){5.12}} \put(50.00,60.00){\line(-6,-5){24.21}} \put(20.53,29.87){\line(1,2){4.87}} \put(79.12,29.79){\line(-5,-2){3.86}} \put(61.58,22.77){\line(-5,-2){5.61}} \bezier{16}(79.47,33.68)(80.53,31.75)(79.47,29.82) \put(55.96,20.70){\line(-1,2){5.88}} \bezier{16}(79.47,33.51)(77.89,34.39)(76.32,35.09) \put(79.47,33.51){\line(-5,6){5.12}} \put(50.00,60.00){\line(6,-5){24.21}} \put(79.47,29.87){\line(-1,2){4.87}} \bezier{80}(75.00,28.00)(65.67,32.33)(62.00,23.00) \bezier{80}(25.00,28.00)(34.33,32.33)(38.00,23.00) \end{picture} \caption{The identification surfaces near the collapse point C in stereographic coordinates. AdS spacetime is the interior of the lightly dotted hyperboloid. The hyperboloid itself represents conformal infinity. The initial surface is a Minkowski hyperboloid (like that of Fig.~5) and in that sense is shown in its true metric. The triangular regions on the infinity hyperboloid, one of which is dotted, are the part of {\curs I}\ that can be shown in this coordinate neighborhood} \end{figure} The tent has a simple, pyramid shape in a stereographic mapping centered at C. Since all geodesics through the center of the map are represented by straight lines in such a map, the sides of the tent are timelike planes (that is, linear spaces in stereographic coordinates), and the folds are spacelike straight lines. Figure~13 shows a tent with no corners but four folds. This can be the spacetime fundamental domain for the $k=3,\,g=0$ three-black-hole of Fig.~10 or for the $k=1,\,g=1$ toroidal black hole of Fig.~12, depending on the identification rule. In the three-black-hole case two of the folds, on opposite sides, are fixed points of the identifications $a$ and $b$ of Fig.~10 that generate the group. The other two folds are fixed points of $ab$ and of $ba$. For the toroidal black hole the fixed points of $a$ and of $b$ of Fig.~12 are not folds, they would be horizontal lines through the tip of the tent. Instead the folds are fixed points of $aba^{-1}b^{-1}$ and its three cyclic permutations. In each case the folds are fixed points of transformations associated with a horizon. All the other fixed points lie outside of the fundamental domain. Because the stereographic picture is centered at a particular time, it can be misleading in that it does not exhibit the time symmetry about the initial surface, nor the early history before the time-symmetric moment. The time-independent sausage coordinates are more suitable for the global view of a black hole spacetime. Since the BTZ black hole (Fig.~14a) involves only one identification, its fundamental domain has only one geodesic of fixed points to the future of the initial surface, the $r=0$ line in Fig.~2a. The sides of the tent are the surfaces $\phi=0$ and $\phi=2\pi$. Fig.~14b is the sausage coordinate version of Fig.~13. \begin{figure} \unitlength 0.90mm \linethickness{0.4pt} \begin{picture}(121.81,77.01)(0,0) \bezier{80}(42.21,16.71)(42.21,13.47)(35.11,11.84) \bezier{80}(35.11,11.84)(28.00,10.22)(20.89,11.84) \bezier{80}(20.89,11.84)(13.79,13.47)(13.79,16.71) \bezier{80}(42.21,44.56)(42.21,41.88)(35.11,40.55) \bezier{80}(35.11,40.55)(28.00,39.22)(20.89,40.55) \bezier{80}(20.89,40.55)(13.79,41.88)(13.79,44.56) \bezier{32}(32.26,40.16)(31.40,40.97)(33.69,41.78) \bezier{44}(33.69,41.78)(36.61,42.47)(39.37,41.78) \bezier{32}(23.74,48.95)(24.60,48.14)(22.31,47.34) \bezier{44}(22.31,47.34)(19.39,46.64)(16.63,47.34) \bezier{80}(42.21,72.40)(42.21,74.71)(35.11,75.86) \bezier{80}(35.11,75.86)(28.00,77.01)(20.89,75.86) \bezier{80}(20.89,75.86)(13.79,74.71)(13.79,72.40) \bezier{80}(42.21,72.40)(42.21,70.09)(35.11,68.94) \bezier{80}(35.11,68.94)(28.00,67.78)(20.89,68.94) \bezier{80}(20.89,68.94)(13.79,70.09)(13.79,72.40) \put(13.79,16.71){\line(0,1){55.68}} \put(42.21,16.71){\line(0,1){55.68}} \put(28.00,5.58){\makebox(0,0)[cc]{({\bf a})}} \thicklines \multiput(16.98,69.89)(0.46,0.12){46}{\line(1,0){0.46}} \multiput(16.63,13.37)(0.38,0.12){58}{\line(1,0){0.38}} \bezier{536}(16.98,69.89)(47.90,38.66)(16.63,13.37) \bezier{148}(37.95,75.18)(42.92,67.01)(42.21,55.69) \bezier{128}(40.79,33.98)(37.95,41.96)(40.79,50.13) \bezier{80}(40.79,50.13)(42.21,55.69)(42.21,55.69) \bezier{88}(40.79,33.98)(42.21,29.52)(42.21,27.85) \bezier{44}(42.21,27.85)(42.21,23.40)(38.66,20.33) \bezier{34}(37.95,75.18)(29.90,67.57)(25.87,59.03) \bezier{228}(25.87,59.03)(19.47,45.30)(30.13,30.64) \bezier{23}(30.13,30.64)(32.97,25.81)(38.66,20.33) \bezier{148}(16.63,13.37)(13.08,22.10)(13.79,33.42) \bezier{128}(15.21,56.25)(18.05,47.16)(15.21,38.99) \bezier{80}(15.21,38.99)(13.79,33.42)(13.79,33.42) \bezier{88}(15.21,56.25)(13.79,59.59)(13.79,62.38) \bezier{44}(13.79,62.93)(13.79,65.72)(16.98,69.89) \thinlines \put(15.92,69.43){\vector(4,1){0.2}} \multiput(8.10,67.39)(0.43,0.11){18}{\line(1,0){0.43}} \put(7.39,67.94){\makebox(0,0)[cc]{P}} \put(42.92,44.56){\makebox(0,0)[lc]{$t=0$}} \bezier{40}(42.21,16.71)(42.21,18.66)(38.66,20.33) \bezier{19}(38.66,20.33)(31.55,22.84)(25.87,22.28) \bezier{17}(25.87,22.28)(18.41,21.73)(14.85,18.94) \bezier{20}(14.85,18.94)(13.79,18.00)(13.79,16.71) \bezier{32}(13.79,44.56)(13.50,46.12)(16.63,47.38) \bezier{10}(16.63,47.38)(19.85,48.64)(23.45,49.01) \bezier{40}(23.45,49.01)(27.05,49.38)(30.75,49.16) \bezier{14}(30.75,49.16)(36.24,48.72)(39.94,47.08) \bezier{24}(39.94,47.08)(41.84,46.41)(42.21,44.56) \bezier{80}(109.73,16.71)(109.73,19.96)(102.62,21.59) \bezier{28}(88.41,21.59)(90.90,21.95)(93.03,22.00) \bezier{28}(102.62,21.59)(100.13,21.95)(98.00,22.00) \bezier{80}(88.41,21.59)(81.30,19.96)(81.30,16.71) \bezier{80}(109.73,16.71)(109.73,13.47)(102.62,11.84) \bezier{80}(102.62,11.84)(95.51,10.22)(88.41,11.84) \bezier{80}(88.41,11.84)(81.30,13.47)(81.30,16.71) \bezier{20}(109.73,44.56)(109.73,47.23)(102.62,48.57) \bezier{20}(102.62,48.57)(95.51,49.90)(88.41,48.57) \bezier{20}(88.41,48.57)(81.30,47.23)(81.30,44.56) \bezier{80}(109.73,44.56)(109.73,41.88)(102.62,40.55) \bezier{80}(102.62,40.55)(95.51,39.22)(88.41,40.55) \bezier{80}(88.41,40.55)(81.30,41.88)(81.30,44.56) \bezier{32}(82.01,43.17)(84.50,43.45)(86.99,42.71) \bezier{44}(86.99,42.71)(89.12,41.76)(86.99,40.85) \bezier{32}(99.78,40.16)(99.78,41.49)(101.20,41.78) \bezier{44}(101.20,41.78)(104.12,42.47)(106.89,41.78) \bezier{36}(91.20,49.01)(92.35,47.68)(88.27,47.08) \bezier{6}(88.27,47.08)(86.47,46.78)(84.00,47.38) \bezier{12}(109.02,45.95)(108.26,45.84)(107.31,45.87) \bezier{15}(107.31,45.87)(100.58,46.63)(104.19,48.27) \bezier{80}(109.73,72.40)(109.73,74.71)(102.62,75.86) \bezier{80}(102.62,75.86)(95.51,77.01)(88.41,75.86) \bezier{80}(88.41,75.86)(81.30,74.71)(81.30,72.40) \bezier{80}(109.73,72.40)(109.73,70.09)(102.62,68.94) \bezier{80}(102.62,68.94)(95.51,67.78)(88.41,68.94) \bezier{80}(88.41,68.94)(81.30,70.09)(81.30,72.40) \put(81.30,16.71){\line(0,1){55.68}} \put(109.73,16.71){\line(0,1){55.68}} \thicklines \bezier{252}(93.38,25.62)(80.59,42.51)(93.38,54.02) \bezier{248}(93.38,54.02)(106.17,40.29)(93.38,25.62) \bezier{208}(109.02,29.52)(104.54,40.84)(109.02,57.92) \bezier{36}(109.02,57.92)(109.73,54.86)(109.73,52.35) \bezier{116}(109.73,52.35)(108.31,46.51)(109.73,36.20) \bezier{52}(109.73,36.20)(109.73,32.86)(109.02,29.52) \bezier{96}(81.30,51.80)(82.72,42.89)(81.30,38.43) \bezier{52}(82.01,59.03)(81.30,54.86)(81.30,51.80) \bezier{48}(82.01,31.75)(81.30,36.20)(81.30,38.43) \thinlines \bezier{50}(82.01,59.03)(86.77,46.97)(82.01,31.75) \bezier{16}(97.65,35.65)(99.54,36.94)(99.78,37.87) \bezier{54}(99.78,37.87)(109.49,46.97)(97.65,62.38) \bezier{108}(97.65,35.65)(90.94,42.39)(91.25,49.01) \bezier{28}(91.25,49.01)(90.94,49.07)(91.96,52.91) \bezier{20}(91.96,52.91)(93.30,56.75)(97.65,62.93) \bezier{44}(97.65,35.65)(96.94,34.26)(96.23,29.52) \bezier{23}(96.23,29.52)(95.87,23.68)(95.51,16.71) \thicklines \bezier{152}(82.01,31.75)(91.25,29.15)(95.51,16.71) \bezier{156}(95.51,16.71)(99.07,29.52)(109.02,29.52) \bezier{84}(93.38,25.62)(95.28,26.74)(95.51,16.71) \bezier{148}(82.01,59.03)(92.67,61.26)(95.51,72.40) \bezier{148}(95.51,72.40)(99.78,60.71)(109.02,57.92) \bezier{132}(93.38,54.02)(95.28,60.33)(95.51,72.40) \thinlines \bezier{19}(95.51,72.40)(96.94,62.56)(97.65,62.93) \put(92.67,54.16){\vector(4,-1){0.2}} \multiput(70.64,58.48)(0.61,-0.12){36}{\line(1,0){0.61}} \put(70.64,59.03){\vector(1,0){11.02}} \put(69.93,59.03){\makebox(0,0)[rc]{P}} \put(121.10,57.92){\vector(-1,0){11.73}} \bezier{76}(121.10,59.03)(115.41,62.38)(109.73,63.49) \put(97.65,63.49){\vector(-4,-1){0.2}} \bezier{68}(109.73,63.49)(104.52,64.60)(97.65,63.49) \put(121.81,58.48){\makebox(0,0)[lc]{P}} \put(95.51,5.58){\makebox(0,0)[cc]{({\bf b})}} \bezier{130}(88.53,47.62)(87.34,39.27)(89.24,32.40) \end{picture} \caption{Fundamental regions and their boundary ``tents" in sausage coordinates, for ({\bf a}) the BTZ black hole and ({\bf b}) a three-black-hole or toroidal black hole configuration} \end{figure} Since folds are spacelike they extend to infinity, and therefore the initial fundamental domain must also have asymptotic regions. Conversely, the tent of an initial state without asymptotic regions has only corners and a tip but no folds: any closed time-symmetric AdS universe always collapses to a point in the finite time $\pi\ell/2$. The holes in the tent are important for the black hole interpretation, for they are the regions at infinity, {\curs I}. The edges of the holes of course disappear once the identifications are made, and the only remaining boundaries of {\curs I}\ appear as points such as those marked P in the figure. The backwards lightcone from P is the boundary of the past of {\curs I}, i.e.\ the horizon. It surrounds the singularity whose end is P. It is now clear that all the initial configurations that have a horizon in the sense of section 2.4 do have spacetime horizons and hence are black holes: a horizon word extended to spacetime is an identification that has fixed points along some fold of the tent-shaped boundary of the spacetime fundamental domain. The intersection of the fold with conformal infinity is an endpoint of a {\curs I}, and the backwards lightcone of that endpoint is the spacetime horizon. For a given fold we can consider a region in the fundamental domain sufficiently near infinity (spatially) and the fold (temporally) so that the only relevant identification is the one that has fixed points on that fold (because the other identifications would move points out of the region). In that region the spacetime is then indistinguishable from that of a BTZ black hole, and the spacetime horizon behaves in the same way as a BTZ black hole horizon. For example, the backward lightcone from P does intersect the initial surface in the minimal horizon geodesic. As we follow the horizon further backward in time it changes from the BTZ behavior only when it encounters other horizons or another part of itself, coming from another copy of the point P in the fundamental domain. For example, in the toroidal black hole interpretation of Fig.~14, all four openings of the tent are parts of one {\curs I}, and there is a single spacetime horizon consisting of the four ``quarter" backwards lightcones from the four copies of the point P. As we go backwards in time below the initial surface these lightcones eventually touch and merge and shown in Fig.~15. \begin{figure} \unitlength 1.05mm \linethickness{0.4pt} \begin{picture}(120.00,30.00)(9,0) \bezier{15}(10.00,20.00)(10.00,24.14)(12.93,27.07) \bezier{15}(12.93,27.07)(15.86,30.00)(20.00,30.00) \bezier{15}(30.00,20.00)(30.00,24.14)(27.07,27.07) \bezier{15}(27.07,27.07)(24.14,30.00)(20.00,30.00) \bezier{15}(10.00,20.00)(10.00,15.86)(12.93,12.93) \bezier{15}(12.93,12.93)(15.86,10.00)(20.00,10.00) \bezier{15}(30.00,20.00)(30.00,15.86)(27.07,12.93) \bezier{15}(27.07,12.93)(24.14,10.00)(20.00,10.00) \bezier{15}(40.00,20.00)(40.00,24.14)(42.93,27.07) \bezier{15}(70.00,20.00)(70.00,24.14)(72.93,27.07) \bezier{15}(100.00,20.00)(100.00,24.14)(102.93,27.07) \bezier{15}(42.93,27.07)(45.86,30.00)(50.00,30.00) \bezier{15}(72.93,27.07)(75.86,30.00)(80.00,30.00) \bezier{15}(102.93,27.07)(105.86,30.00)(110.00,30.00) \bezier{15}(60.00,20.00)(60.00,24.14)(57.07,27.07) \bezier{15}(90.00,20.00)(90.00,24.14)(87.07,27.07) \bezier{15}(120.00,20.00)(120.00,24.14)(117.07,27.07) \bezier{15}(57.07,27.07)(54.14,30.00)(50.00,30.00) \bezier{15}(87.07,27.07)(84.14,30.00)(80.00,30.00) \bezier{15}(117.07,27.07)(114.14,30.00)(110.00,30.00) \bezier{15}(40.00,20.00)(40.00,15.86)(42.93,12.93) \bezier{15}(70.00,20.00)(70.00,15.86)(72.93,12.93) \bezier{15}(100.00,20.00)(100.00,15.86)(102.93,12.93) \bezier{15}(42.93,12.93)(45.86,10.00)(50.00,10.00) \bezier{15}(72.93,12.93)(75.86,10.00)(80.00,10.00) \bezier{15}(102.93,12.93)(105.86,10.00)(110.00,10.00) \bezier{15}(60.00,20.00)(60.00,15.86)(57.07,12.93) \bezier{15}(90.00,20.00)(90.00,15.86)(87.07,12.93) \bezier{15}(120.00,20.00)(120.00,15.86)(117.07,12.93) \bezier{15}(57.07,12.93)(54.14,10.00)(50.00,10.00) \bezier{15}(87.07,12.93)(84.14,10.00)(80.00,10.00) \bezier{15}(117.07,12.93)(114.14,10.00)(110.00,10.00) \bezier{30}(46.38,29.13)(47.25,27.25)(50.00,26.96) \bezier{30}(53.62,29.13)(52.75,27.25)(50.00,26.96) \bezier{30}(40.87,23.62)(42.75,22.75)(43.04,20.00) \bezier{30}(40.87,16.38)(42.75,17.25)(43.04,20.00) \bezier{50}(12.78,12.93)(15.71,15.86)(15.71,20.00) \bezier{50}(12.93,27.03)(15.86,24.10)(20.00,24.10) \bezier{50}(27.07,27.03)(24.14,24.10)(20.00,24.10) \bezier{50}(27.22,12.93)(24.29,15.86)(24.29,20.00) \bezier{50}(27.22,27.07)(24.29,24.14)(24.29,20.00) \bezier{50}(27.07,12.97)(24.14,15.90)(20.00,15.90) \bezier{50}(12.93,12.97)(15.86,15.90)(20.00,15.90) \bezier{50}(12.78,27.07)(15.71,24.14)(15.71,20.00) \thicklines \bezier{12}(42.90,21.16)(47.83,22.17)(48.84,27.10) \bezier{12}(42.90,18.84)(47.83,17.83)(48.84,12.90) \bezier{12}(57.10,21.16)(52.17,22.17)(51.16,27.10) \bezier{12}(57.10,18.84)(52.17,17.83)(51.16,12.90) \thinlines \bezier{30}(46.38,10.87)(47.25,12.75)(50.00,13.04) \bezier{30}(53.62,10.87)(52.75,12.75)(50.00,13.04) \bezier{30}(59.13,23.62)(57.25,22.75)(56.96,20.00) \bezier{30}(59.13,16.38)(57.25,17.25)(56.96,20.00) \bezier{40}(74.44,28.33)(76.33,25.33)(80.00,25.33) \bezier{40}(85.56,28.33)(83.67,25.33)(80.00,25.33) \bezier{40}(88.33,14.44)(85.33,16.33)(85.33,20.00) \bezier{40}(88.33,25.56)(85.33,23.67)(85.33,20.00) \bezier{40}(71.67,25.56)(74.67,23.67)(74.67,20.00) \bezier{40}(71.67,14.44)(74.67,16.33)(74.67,20.00) \bezier{40}(85.56,11.67)(83.67,14.67)(80.00,14.67) \bezier{40}(74.44,11.67)(76.33,14.67)(80.00,14.67) \thicklines \bezier{8}(80.00,25.33)(82.00,22.00)(85.33,20.00) \bezier{8}(80.00,25.33)(78.00,22.00)(74.67,20.00) \bezier{8}(80.00,14.67)(82.00,18.00)(85.33,20.00) \bezier{8}(80.00,14.67)(78.00,18.00)(74.67,20.00) \thinlines \bezier{30}(106.22,20.00)(106.22,22.50)(105.22,24.78) \bezier{10}(105.22,24.78)(104.56,26.20)(103.56,27.30) \bezier{30}(106.22,20.00)(106.22,17.50)(105.22,15.22) \bezier{10}(105.22,15.22)(104.56,13.80)(103.56,12.70) \bezier{30}(110.00,16.22)(107.50,16.22)(105.22,15.22) \bezier{10}(105.22,15.22)(103.80,14.56)(102.70,13.56) \bezier{30}(110.00,16.22)(112.50,16.22)(114.78,15.22) \bezier{10}(114.78,15.22)(116.20,14.56)(117.30,13.56) \bezier{30}(113.78,20.00)(113.78,17.50)(114.78,15.22) \bezier{30}(113.78,20.00)(113.78,22.50)(114.78,24.78) \bezier{30}(110.00,23.78)(112.50,23.78)(114.78,24.78) \bezier{30}(110.00,23.78)(107.50,23.78)(105.22,24.78) \bezier{10}(114.78,15.22)(115.44,13.80)(116.44,12.70) \bezier{10}(114.78,24.78)(115.44,26.20)(116.44,27.30) \bezier{10}(114.78,24.78)(116.20,25.44)(117.30,26.44) \bezier{10}(105.22,24.78)(103.80,25.44)(102.70,26.44) \thicklines \bezier{3}(110.00,21.50)(109.50,20.50)(108.50,20.00) \bezier{3}(110.00,21.50)(110.50,20.50)(111.50,20.00) \bezier{3}(110.00,18.50)(109.50,19.50)(108.50,20.00) \bezier{3}(110.00,18.50)(110.50,19.50)(111.50,20.00) \thinlines \put(27.50,27.50){\makebox(0,0)[lb]{P}} \put(12.50,27.50){\makebox(0,0)[rb]{P}} \put(12.50,12.50){\makebox(0,0)[rt]{P}} \put(27.50,12.50){\makebox(0,0)[lt]{P}} \put(110.00,3.30){\makebox(0,0)[cc]{d}} \put(20.00,3.00){\makebox(0,0)[cc]{a}} \put(50.00,3.30){\makebox(0,0)[cc]{b}} \put(80.00,3.00){\makebox(0,0)[cc]{c}} \end{picture} \caption{Slices of the sausage in Fig.~14 to show the time development of the horizon. Part a is the latest and part d the earliest sausage time. The geometry of each time slice is the constant curvature space represented by a Poincar\'e disk. The geodesics shown by solid lines are to be identified as before for the toroidal black hole. Where these geodesics intersect we have a fixed point of some identification, a physical singularity. Slice a is at the sausage time of the end point P of {\curs I}. As we go backwards in time, the horizon (dotted arcs of circles) spreads out from those points at infinity. Slice b is the moment of time-symmetry. The horizon remains smooth until slice c, when its different parts meet each other at the identification surfaces. Prior to that time the event horizon has four kinks} \end{figure} \subsection{Fixed Points at Infinity} As the above examples of multi-black-hole time developments show, any time-symmetric initial state with an asymptotic region ending at a horizon is isometric to a corresponding region of a BTZ black hole, so that each such region will look like a black hole from infinity for at least a finite time. It is maybe not so clear whether this is also true for the unlimited time necessary for a true black hole, for example because other singularities (fixed points) might intervene. By an interesting method due to {\AA}minneborg, Bengtsson and Holst \cite{ABH} one can directly find all of the universal covering space of {\curs I}\ from a knowledge of spatial infinity on an initial surface. (The universal covering space gives information about horizons and is natural in many contexts, for example topological censorship questions reduce to existence of certain geodesics in AdS space \cite{DBS}.) Since our black holes are quotient spaces of AdS space, the covering space of their {\curs I}\ will be a subset of conformal infinity of AdS space. To describe this conformal infinity in a finite way we follow the usual Penrose procedure and multiply the AdS metric by a factor so that the resulting metric is finite in the asymptotic region. An obviously suitable conformal factor in Eq (\ref{sausge}) is $(1-(\rho/\ell)^2)^2$, giving the metric at infinity, $\rho=\ell$, $$ds^2_\infty = 4 (dt^2 + \ell^2 d\theta^2)$$ This is the flat metric of a cylinder of radius $\ell$. Consider first the covering space of {\curs I}\ of a single black hole in this description, and recall that the identification is a ``Lorentz boost" in the embedding space (\ref{ads}). As we apply this transformation $n$ times to get the $n$th tile of the covering space, we are boosting the fundamental domain to the limiting velocity, and since the identification boundaries are timelike in our description, they become two null surfaces in the limit $n\rightarrow\pm\infty$. These null surfaces (called ``singularity surfaces" in \cite{HP}) are then of course invariant under the identification transformation. Hence, if ${\bf K}=\partial/\partial\phi$ is the Killing vector corresponding to the identification, these surfaces are described by ${\bf K}^2=0$. They intersect where the vector {\bf K} itself vanishes, that is at the fixed points at infinity on the initial surface and at the singularity inside the black hole. The intersection of these surfaces with conformal infinity of AdS space is the boundary ($n\rightarrow\infty$) of the covering space of {\curs I}. To find it we only need to draw null lines from the endpoints of the horizon at $t=0$ toward each other (Fig.~16a). Their future intersection is the nearest future fixed point to this initial surface, it is the end of {\curs I}, and the covering space of {\curs I}\ is the diamond-shaped region between the future and past null lines. Furthermore the future null lines are also the intersection of the covering space of the horizon with conformal infinity, since the horizon is the backward null cone from the end of {\curs I}. Thus a knowledge of the initial fixed points gives us the ``holographic" information about the exterior and the horizon of the black hole. \begin{figure} \unitlength 1.20mm \linethickness{0.4pt} \begin{picture}(106.00,32.00)(10,0) \put(10.00,20.00){\line(1,-1){10.00}} \put(20.00,10.00){\line(1,1){20.00}} \put(40.00,30.00){\line(1,-1){10.00}} \put(50.00,20.00){\line(-1,-1){10.00}} \put(40.00,10.00){\line(-1,1){20.00}} \put(20.00,30.00){\line(-1,-1){10.00}} \put(10.00,8.00){\line(0,1){24.00}} \put(50.00,32.00){\line(0,-1){24.00}} \thicklines \bezier{104}(40.00,30.00)(48.00,20.00)(40.00,10.00) \thinlines \bezier{72}(40.00,30.00)(47.00,22.33)(47.00,20.00) \bezier{76}(40.00,30.00)(48.67,21.00)(48.67,20.00) \bezier{72}(40.00,10.00)(47.00,17.67)(47.00,20.00) \bezier{76}(40.00,10.00)(48.67,19.00)(48.67,20.00) \thicklines \bezier{104}(20.00,30.00)(12.00,20.00)(20.00,10.00) \thinlines \bezier{72}(20.00,30.00)(13.00,22.33)(13.00,20.00) \bezier{76}(20.00,30.00)(11.33,21.00)(11.33,20.00) \bezier{72}(20.00,10.00)(13.00,17.67)(13.00,20.00) \bezier{76}(20.00,10.00)(11.33,19.00)(11.33,20.00) \thicklines \put(40.00,10.00){\vector(0,1){20.00}} \put(20.00,30.00){\line(0,-1){20.00}} \thinlines \put(30.00,20.00){\vector(1,0){20.00}} \put(30.00,20.00){\vector(-1,0){20.00}} \put(51.00,20.00){\makebox(0,0)[lc]{$\phi$}} \put(65.00,20.00){\line(1,-1){10.00}} \put(75.00,10.00){\line(1,1){20.00}} \put(95.00,30.00){\line(1,-1){10.00}} \put(105.00,20.00){\line(-1,-1){10.00}} \put(95.00,10.00){\line(-1,1){20.00}} \put(75.00,30.00){\line(-1,-1){10.00}} \put(65.00,8.00){\line(0,1){24.00}} \put(105.00,32.00){\line(0,-1){24.00}} \bezier{84}(85.00,20.00)(81.79,17.95)(75.00,20.00) \bezier{84}(85.00,20.00)(88.21,22.05)(95.00,20.00) \put(65.00,20.00){\vector(-2,-1){0.2}} \bezier{84}(75.00,20.00)(68.21,22.05)(65.00,20.00) \put(105.00,20.00){\vector(2,1){0.2}} \bezier{84}(95.00,20.00)(102.05,17.95)(105.00,20.00) \put(106.00,20.00){\makebox(0,0)[lc]{$\varphi$}} \thicklines \bezier{64}(95.00,10.00)(92.95,13.21)(95.00,20.00) \put(95.00,30.00){\vector(-1,2){0.2}} \bezier{84}(95.00,20.00)(96.92,26.67)(95.00,30.00) \thinlines \put(95.00,30.51){\makebox(0,0)[cb]{$\tau$}} \put(40.00,30.51){\makebox(0,0)[cb]{$t$}} \thicklines \bezier{84}(75.00,10.00)(72.95,13.21)(75.00,20.00) \bezier{84}(75.00,20.00)(76.92,26.67)(75.00,30.00) \bezier{60}(95.00,10.00)(99.00,17.00)(98.83,19.00) \bezier{68}(98.83,19.00)(98.00,25.17)(95.00,30.00) \thinlines \bezier{88}(95.00,10.00)(101.50,17.50)(101.50,19.00) \bezier{86}(101.50,19.00)(101.50,20.67)(95.00,30.17) \bezier{86}(95.00,10.00)(103.50,18.83)(102.83,19.00) \bezier{84}(102.83,19.00)(104.50,19.50)(95.00,30.00) \thicklines \bezier{60}(75.00,30.00)(71.00,23.00)(71.17,21.00) \bezier{68}(71.17,21.00)(72.00,14.83)(75.00,10.00) \thinlines \bezier{88}(75.00,30.00)(68.50,22.50)(68.50,21.00) \bezier{86}(68.50,21.00)(68.50,19.33)(75.00,9.83) \bezier{86}(75.00,30.00)(66.50,21.17)(67.17,21.00) \bezier{84}(67.17,21.00)(65.50,20.50)(75.00,10.00) \put(85.00,3.00){\makebox(0,0)[cc]{({\bf b})}} \put(30.00,3.00){\makebox(0,0)[cc]{({\bf a})}} \end{picture} \caption{Universal covering spaces of conformal infinity {\curs I}\ for ({\bf a}) non-rotating, ({\bf b})~rotating black holes on the conformal infinity cylinder of AdS space. To show the cylinder in this flat picture it has been cut along the vertical lines, which are to be identified with each other in each part of the figure. The angle $\phi$ resp.\ $\varphi$ runs from $-\infty$ to $+\infty$, in the direction of the arrow, in each of the diamond-shaped regions. The intersection of the fundamental domain with conformal infinity of AdS space is shown by the heavier boundaries. These boundaries are at the values $0$ and $2\pi$ of $\phi$ resp.\ $\varphi$. A few of the tiles obtained by the isometries that change these angles by $2\pi n$ are shown for positive multiples $n$. In the limit $n\rightarrow\pm\infty$ the tiles converge to the null boundaries of the two diamond-shaped regions. These null boundaries intersect on the initial surface ($t=0$ resp.\ $\tau=0$) at the $n\rightarrow\infty$ limit at conformal infinity of the initial surface, and they end in the future at the end of {\curs I}} \end{figure} The situation for time-symmetric multi-black-holes is similar, except that the construction yields an infinite number of copies of the covering space of {\curs I}. We saw in Fig.~10 that each horizon word has two fixed points at spatial infinity. Each such pair yields a diamond-shaped region for which the transformation of its horizon word looks like that of Fig.~16a, and which is free of fixed points. \section{Angular Momentum} In the metric (\ref{BTZ}) for the static BTZ black hole, introduce new coordinates $T,\,\varphi,\,R$, \begin{eqnarray} t = T + \left({J\over 2m}\right)\varphi\nonumber \\ \phi = \varphi + \left({J\over 2m\ell^2}\right)T \\ R^2=r^2\left(1-{J^2\over 4m^2\ell^2}\right) + {J^2\over 4m}\nonumber \label{coord} \end{eqnarray} where $J<2m\ell$ is a constant with dimension of length, and define another new constant \begin{equation} M = m + {J^2\over 4m\ell^2}. \label{mass} \end{equation} In terms of these new quantities the metric (\ref{BTZ}) may be written as \begin{equation} ds^2 = -N^2 dT^2 + N^{-2}dR^2 + R^2\left(d\varphi + {J\over 2R^2} dT\right)^2 \label{BTZJ} \end{equation} where $$N^2 = \left({R\over\ell}\right)^2 - M + \left({J\over 2R}\right)^2.$$ The metric (\ref{BTZJ}) now looks like a (2+1)-dimensional analog of the metric for a black hole that carries angular momentum. Although metric (\ref{BTZJ}) was obtained by a coordinate transformation from (\ref{BTZ}) and is therefore locally isometric to the latter (as all of our spaces are locally isometric to AdS space), it differs in its global structure: we have silently assumed that the new metric (\ref{BTZJ}) is periodic with period $2\pi$ in the {\it new} angular variable $\varphi$, rather than in the old variable $\phi$. This means, in coordinate-independent language, that we have changed the identification group that creates this new spacetime from AdS space. As for the non-rotating BTZ black hole, the new group for this ``single" rotating black hole is still generated by all the powers of a single isometry of AdS space, but this isometry does not leave invariant a totally geodesic spacelike surface of time symmetry. The surface $T =$ const that is obviously left invariant by a displacement of $\varphi$ is twisted, as measured by its extrinsic curvature, and this is one indication of the global difference from the static metric. When only the one new coordinate $\varphi$ changes by $2\pi$, the old coordinates of (\ref{BTZ}) change by \begin{equation} t\rightarrow t+{\pi J\over m} \qquad \phi \rightarrow \phi + 2\pi. \label{twist} \end{equation} A change in either $t$ or $\phi$ is of course an isometry of the metric (\ref{BTZ}), and because $t$ and $\phi$ are coordinates, the two changes commute. The identification for a rotating black hole involves the two isometries applied simultaneously. Either one is a ``boost" about an axis of fixed points; the change in $\phi$ has fixed points in the future, at $r=0$, and the change in $t$ has fixed points at $r=\ell\sqrt{m}$, the horizon of (\ref{BTZ}).\footnote{Since these isometries are also isometries of the periodically identified embedding (\ref{ads}), each axis of fixed points is really repeated an infinite number of times in AdS space itself.} The combination of the two does not have any fixed points at all (either one moves points on the fixed axis of the other in the direction of the axis): the length $R^2$ of the corresponding Killing vector $\partial/\partial\varphi$ vanishes where the vector is null but not zero, since its scalar product with the finite $\partial/\partial T$ is the finite constant $J$. Earlier we argued that a spacelike set of fixed points of the identification isometry becomes a kind of singularity after the identification, and its removal from the spacetime gave us the end of {\curs I}\ and associated horizon. What happens when we do not have this singularity? \subsection{Is it a Black Hole?} The geometry of metric (\ref{BTZJ}) --- more properly speaking, the geometry of its analytic extension, or of AdS space identified according to the $\varphi\rightarrow\varphi+2\pi$ isometry exhibited by this metric --- satisfies the definition of a black hole if we are somewhat creative about the definition of ``singularity." We expect the singularity to occur at $R=0$, but because there are no fixed points, the identified spacetime is regular there, and can be continued to negative $R^2$. But then the closed $\varphi$-direction becomes timelike, hence the spacetime has a region of closed timelike lines. We shall follow the usual practice to regard these as sufficiently unphysical that they should be eliminated from the spacetime, like a singularity. So we confine attention to $R>0$. Our spacetime then ends at the singularity surfaces where the square of the Killing vector $\partial/\partial\varphi$ vanishes, $R^2=0$. The corresponding $r^2$ of Eq (\ref{coord}) is negative. We recall from Sect.~2.1 that this occurs on two timelike surfaces in a region where $\phi$ is timelike, unlike the non-rotating black hole whose singularity occurs on the spacelike line $r=0$. Since there is a singularity-free region between the two singularity surfaces, not all timelike lines that ``fall into the black hole" (cross the horizon) end at the singularity; they can escape through the hole left open by the singularity surfaces, as is the case in a three-dimensional Kerr black hole. However, at conformal infinity the difference between $R$ and $r$ disappears, the two singularity surfaces come together at the point where the spacelike line $r=0$ meets conformal infinity.\ Thus the covering space of {\curs I}\ for the rotating black hole looks the same as that of the non-rotating one that corresponds to it via Eq (\ref{mass}), only the identification is different, as shown in Fig.~16b. We see that {\curs I}\ has an endpoint, there is a horizon, so the identified spacetime is a black hole. We can recognize a (rotating) black hole in a spacetime by the presence of a closed, non-contractible spacelike geodesic $\gamma$. If we have such a $\gamma$ we consider all spacelike geodesics that start normal to $\gamma$. We assume that these can be divided into two types, which we might call right-starting and left-starting (with respect to an arbitrarily chosen direction of $\gamma$). If all geodesics of one type reach infinity, then they cover the outside of a black hole. In this region the totally geodesic timelike surfaces normal to $\gamma$ are surfaces of constant $\phi$. Within these surfaces one can introduce coordinates so that the metric takes the form (\ref{BTZJ}). (If the normals to those surfaces, not at $\gamma$, also integrate to closed curves after one circuit of $\gamma$, we have $J=0$.) \subsection{Does it rotate?} The asymptotically measurable properties of (2+1)-dimensional black holes can be defined in various way, for example: from the ADM form of the Einstein action; as the conserved quantities that go with the Killing vectors $\partial/\partial t$ and $\partial/\partial\varphi$; as the Noether charges associated with $t$- and $\varphi$-displacements; and so on \cite{mass,BTZ}. All of these yield $M$ as the mass and $J$ as the angular momentum. $J$ can also be measured ``quasi-locally" in the neighborhood of the horizon. We find an extremal closed spacelike geodesic (corresponding to $r=\ell$) and parallel transport an orthogonal vector around this geodesic. According to Eq (\ref{twist}) the hyperbolic ``holonomy" angle between the original and rotated vectors is $\pi J/m\ell$. \subsection{Multiple Black Holes with Angular Momentum} It is fairly straightforward to extend the methods of section 2.4 to obtain metrics with several asymptotic regions, or with non-standard topologies, that have angular momentum as measured in these asymptotic regions; the main difference is that we will deal with spacetimes rather than initial values. Our aim is only to show that rotating multi-black-holes are possible, and to indicate what the free parameters are. We begin with a three-black-hole spacetime, whose time development can be described by the geometry of Fig.~13. We suppose that the front left and right surfaces are identified, and similarly the back left and right surfaces. The corresponding fixed points are the front and back edges of the pyramid. As we have seen, there is then a third black hole associated with the left and right edges (which are identified with each other). We cut this figure into two halves by the plane $S$ (a totally geodesic surface) spanned by these left and right edges. This surface cuts the third black hole into two equal parts, which we can think of, for example, as $\phi = 0$ to $\phi = \pi$ and $\phi = \pi$ to $\phi = 2\pi$, respectively. Now we re-identify the two halves with a ``boost" between them, that is an isometry with fixed points along the normal to the plane $S$ at the center of the initial surface, as illustrated in Fig.~17. \begin{figure} \unitlength 1.0mm \linethickness{0.4pt} \begin{picture}(86.00,68.00)(-7,10) \bezier{30}(20.15,32.00)(20.15,38.59)(35.08,42.34) \bezier{30}(64.92,42.34)(79.85,38.59)(79.85,32.00) \bezier{30}(20.15,32.00)(20.15,25.41)(35.08,21.66) \bezier{40}(35.08,21.66)(50.00,18.71)(64.92,21.66) \bezier{30}(64.92,21.66)(79.85,25.41)(79.85,32.00) \bezier{14}(34.17,60.00)(34.17,63.17)(42.09,64.97) \bezier{20}(42.09,64.97)(50.00,66.39)(57.91,64.97) \bezier{14}(57.91,64.97)(65.83,63.17)(65.83,60.00) \bezier{14}(34.17,60.00)(34.17,56.83)(42.09,55.03) \bezier{20}(42.09,55.03)(50.00,53.61)(57.91,55.03) \bezier{14}(57.91,55.03)(65.83,56.83)(65.83,60.00) \bezier{264}(24.00,35.00)(50.00,55.00)(76.00,35.00) \put(54.00,68.00){\vector(-1,-2){3.50}} \put(55.00,68.00){\makebox(0,0)[lc]{C (Center of projection)}} \put(29.00,65.00){\vector(1,-1){4.50}} \put(28.50,65.00){\makebox(0,0)[rc]{Conformal infinity}} \bezier{36}(23.86,35.09)(21.75,31.75)(24.74,28.25) \bezier{92}(38.25,22.63)(49.82,20.88)(61.40,22.63) \bezier{60}(24.91,28.07)(28.77,24.39)(38.42,22.63) \bezier{36}(76.14,35.09)(78.25,31.75)(75.26,28.25) \bezier{60}(75.09,28.07)(71.23,24.39)(61.58,22.63) \bezier{26}(34.04,60.00)(34.04,50.35)(25.09,39.30) \put(71.93,47.37){\vector(-3,-2){9.65}} \put(72.28,47.37){\makebox(0,0)[lc]{Initial surface}} \bezier{26}(65.96,60.00)(65.96,50.35)(74.91,39.30) \thicklines \put(44.00,56.00){\line(6,1){12.00}} \put(44.00,56.00){\line(1,-6){6.33}} \bezier{20}(43.00,62.00)(42.30,61.20)(36.00,54.00) \bezier{14}(42.90,62.00)(43.05,61.20)(43.85,56.00) \bezier{18}(43.00,62.00)(43.70,61.50)(50.00,57.00) \bezier{18}(57.00,62.00)(56.30,61.50)(50.00,57.00) \bezier{20}(57.00,62.00)(57.70,61.30)(64.00,55.00) \put(50.00,60.00){\circle*{0.00}} \put(56.00,58.00){\line(6,-5){30.00}} \put(85.9,32.58){\line(-5,-2){36.00}} \put(44.00,56.00){\line(-6,-5){28.50}} \put(15.25,32.1){\line(5,-2){34.50}} \thinlines \bezier{10}(57.00,62.00)(56.90,61.60)(56.00,58.00) \put(56.00,58.00){\line(-1,-4){3.25}} \put(86.00,33.00){\line(-5,2){16.00}} \put(15.50,32.20){\line(5,2){10.00}} \end{picture} \begin{caption} {A three-black-hole geometry obtained by cutting Fig.~13 into two tetrahedra by the plane $S$ of the paper (passing through C), and re-gluing after an isometry with axis normal to that plane. The isometry moves the top of the front tetrahedron from C to the left (and up), and the top of the back tetrahedron from C to the right (and up). The dotted outlines show these two tops. The solid figure approximates the convex region bounded by the four planes that are identified pairwise (but it is {\em not} the fundamental domain). The edges where the planes intersect are drawn only to identify these planes; they are simplified as straight lines (but ought to be hyperbolic arcs, representing geodesics). Unlike in Fig.~13 the edges are not to be considered as singularities, except for the front and back edges, which are fixed points of the two basic identifications that generate this spacetime. The other ``singularities" are the boundaries of the regions of closed timelike lines, not drawn (and not easily identified) in this figure.} \end{caption} \end{figure} The four planes stick out of the conformal infinity surface at the four bottom corners, uncovering four parts of conformal infinity. As in Fig.~13, the left and right infinity parts combine into one continuous region due to the identifications. So this spacetime has three conformal infinities with ends, and therefore represents three black holes. The two black holes associated with the front and back edges, as seen from their respective asymptotic regions, are unchanged by this re-identification: by a ``boost" isometry either of these edges and associated planes (but not both together) can be moved back to their old position. Since the two planes alone determine the asymptotic behavior of the black hole, either of these holes has the same mass, and vanishing angular momentum, as before. But the third black hole changes, because the left and right edges no longer lie in the same plane. As we go once around this third black hole, we cross the surface $S$ twice, and its effects add (as a right-handed screw is right-handed from either end). The black hole therefore acquires angular momentum. Unfortunately this is not directly described by Eqs (19-22), because the ``boost" in Eq (\ref{twist}) has fixed points at the horizon of the non-rotating black hole, whereas the fixed points of the boost of Fig.~17 lie along a geodesic connecting the asymptotic regions of the two other holes. However, for the third black hole this difference is asymptotically negligible: as seen from its own infinity it does have angular momentum. (Its standard form (\ref{BTZJ}) would correspond to identification surfaces different from any of those drawn in Fig.~17.) By a similar re-identification any one of a $k$-black-hole time-symmetric spacetime can be given angular momentum; further momentum parameters will be needed to describe how the asymptotic regions fit to an interior. Generally we expect one momentum parameter for each configuration parameter of the corresponding time-symmetric spacetime. For example, the toroidal black hole constructed as in Fig.~12 should allow three independent momenta. Of these the state in which there is angular momentum of the black hole as seen from infinity has been constructed \cite{ABH}. (Another state with momentum can be obtained from Fig.~17 by identifying opposite rather than adjacent planes.) \section{Conclusions} We have seen that a considerable variety of black hole and multiply-connected spacetimes can be constructed by cutting a region out of anti-de Sitter space and identifying the cuts in various ways. Many of the properties, such as horizon structure and topological features of the time-symmetric spacetimes, have been investigated in detail. Comparatively little beyond existence is known about the spacetimes with angular momentum (but see \cite{ABH}). \section*{Acknowledgment} This research was supported in part by the National Science Foundation under Grant No.~PHY94-07194.
\section{Introduction}\label{sec:intro} The galactic black hole candidate (BHC) GX~339$-$4 is unique among persistent sources in that it shows a wide variety of spectral states and transitions among these states. In presumed order of increasing bolometric luminosity, GX~339$-$4 exhibits states with hard, power-law spectra (`off state', \def\astroncite##1##2{##1\ ##2}\@internalcite{ilovaisky:86a}; `low state', \def\astroncite##1##2{##1\ ##2}\@internalcite{grebenev:91a}); a soft state with \emph{no} evidence of a power law tail (`high state'; \def\astroncite##1##2{##1\ ##2}\@internalcite{grebenev:91a}); and a very bright, soft state with extended power-law tail (`very high state'; \def\astroncite##1##2{##1\ ##2}\@internalcite{miyamoto:91a}). There also are apparently times when the flux is high, but the spectrum is not as soft as the `high' or `very high state'. \def\astroncite##1##2{##1\ (##2)}\@internalcite{mendez:97a} refer to this as the `intermediate state'. We also note that there is some evidence of overlap between the states. The broad-band (GRANAT/SIGMA) hard state data presented by \def\astroncite##1##2{##1\ (##2)}\@internalcite{grebenev:91a} apparently represents a more luminous state than does the broad-band soft state data taken with the same instrument. [\def\astroncite##1##2{##1\ (##2)}\@internalcite{miyamoto:95a} has suggested the possibility of hysteresis in galactic BHC state transitions.] Similarly diverse sets of states have been observed in X-ray transients such as Nova Muscae (\def\astroncite##1##2{##1\ ##2}\@internalcite{kitamoto:92a,miyamoto:94a}); however, GX~339$-$4 is closer to being a persistent source. Although there have been a number of observations of GX~339$-$4 in the near-infrared and optical (\def\astroncite##1##2{##1\ ##2}\@internalcite{doxsey:79a,motch:83a,motch:85a,steiman:90a,imamura:90a,cowley:91b}), including detection of a 14.8\,hr periodicity in the optical (\def\astroncite##1##2{##1\ ##2}\@internalcite{callanan:92a}), there is no convincing mass function for the system. In the optical, the system is faint, variable ($M_{\rm V} \approx 16$--$20$), and reddened ($A_{\rm V} = 3.5$). The physical source of the optical emission is unknown. It has been hypothesized that it is \emph{entirely} dominated by the accretion disk, as the optical flux is apparently anticorrelated with the soft X-ray emission (\def\astroncite##1##2{##1\ ##2}\@internalcite{steiman:90a,imamura:90a}). These properties of the emission have made it difficult to obtain a good distance measurement, with estimates ranging from 1.3\,kpc (\def\astroncite##1##2{##1\ ##2}\@internalcite{predhel:91a}) to 8\,kpc (\def\astroncite##1##2{##1\ ##2}\@internalcite{grindlay:79a}), with many authors choosing 4\,kpc (\def\astroncite##1##2{##1\ ##2}\@internalcite{doxsey:79a,cowley:87a}). A careful study of these distance estimates is presented by \def\astroncite##1##2{##1\ (##2)}\@internalcite{zdziarski:98a} who argue for a distance of $\approx 4$\,kpc. GX~339$-$4 also has been detected in the radio (\def\astroncite##1##2{##1\ ##2}\@internalcite{sood:94a}), and possibly even has exhibited extended emission (\def\astroncite##1##2{##1\ ##2}\@internalcite{fender:97a}). Within the hard state, the radio spectrum is flat/inverted with a spectral index of $\alpha=0.1$--0.2 (\def\astroncite##1##2{##1\ ##2}\@internalcite{fender:97a,corbel:98a}), where the radio flux density $S_\nu \propto \nu^{\alpha}$. Furthermore, in this state the radio flux is correlated with the X-ray and gamma-ray fluxes (\def\astroncite##1##2{##1\ ##2}\@internalcite{hanni:98a}), but the radio flux disappears as GX~339$-$4 transits to a higher X-ray flux/softer state (Fender 1998, priv.\ comm.), which is comparable to the behavior of Cyg~X-1 (\def\astroncite##1##2{##1\ ##2}\@internalcite{pooley:98a}). During the Rossi X-ray Timing Explorer (RXTE) Cycle~2 observing phase (1996~December -- 1998~February), we performed a series of eight RXTE observations of GX~339$-$4. The first three observations were spaced a week apart from one another from 1997 February~4 to 1997 February~18. These three observations were scheduled to be simultaneous with 8.3--9.1\,GHz radio observations that were conducted at the Australian Telescope Compact Array (ATCA). The results of the radio observations have been reported by \def\astroncite##1##2{##1\ (##2)}\@internalcite{corbel:98a}. Additionally, three 843\,MHz observations performed at the Molongolo Observatory Synthesis Telescope (MOST) and reported by \def\astroncite##1##2{##1\ (##2)}\@internalcite{hanni:98a} are also simultaneous with these RXTE observations. This paper is structured as follows. We discuss the spectral analysis of archival Advanced Satellite for Cosmology and Astrophysics (ASCA) data in section~\ref{sec:asca}. We look for evidence of Fe lines in the data and we characterize the soft ($\aproxlt 1$\,keV) X-ray data. In section~\ref{sec:xte} we present the RXTE data. We first discuss the All Sky Monitor (ASM) data, and then we discuss the pointed observations. We perform spectral analysis much akin to that which we considered for Cyg~X-1 (\def\astroncite##1##2{##1\ ##2}\@internalcite{dove:98a}). Here, however, we consider Advection Dominated Accretion Flow (ADAF) models as well by using the models described by \def\astroncite##1##2{##1\ (##2)}\@internalcite{dimatteo:98a}. We discuss the implications of the simultaneous radio data in section~\ref{sec:radio}. In section~\ref{sec:discuss} we discuss the implications of the X-ray observations for theoretical models. We summarize our results in section~\ref{sec:summ}. We present timing analysis of the RXTE data in a companion paper (\def\astroncite##1##2{##1\ ##2}\@internalcite{nowak:98c}, henceforth paper~II). \section{Archival ASCA Observations}\label{sec:asca} \begin{table*} \caption{\small Log of the ASCA observations. \label{tab:ascalog}} \smallskip \center{ \begin{tabular}{lllll} \hline \hline \noalign{\vspace*{1mm}} Obs. & Date & Integration time & SIS0 & 3--9\,keV Flux \\ & & (ksec) & (cts s$^{-1}$)& ($10^{-9}~{\rm ergs~cm^{-2}~s^{-1}}$)\\ \noalign{\vspace*{1mm}} \hline \noalign{\vspace*{1mm}} 1 & 1994 August 24 & 15 & 3.6 & 0.11 \\ \noalign{\vspace*{1mm}} 2 & 1994 September 12 & 17 & 6.4 & 0.19 \\ \noalign{\vspace*{1mm}} 3 & 1995 September 8 & 30 & 17.5 & 0.63 \\ \noalign{\vspace*{1mm}} \hline \end{tabular}} \tablecomments{All observations were taken in Bright 1-CCD mode. SIS0: filtered SIS0 count rate.} \end{table*} The ASCA archives contain four observations of the GX~339$-$4 region. A log of the observations is given in in Table~\ref{tab:ascalog}. In appendix~\ref{sec:ascaapp}, we describe the methods that we used to extract, filter, and analyze these ASCA observations. To the best of our knowledge, an analysis of these observations has not been published previously, except for a power spectrum for one of the observations (date not given, \def\astroncite##1##2{##1\ ##2}\@internalcite{dobrin:97a}). The first of the observations (1993 September 16) did not detect the source, with the upper limit to the 3--9\,keV flux being $\approx 10^{-12}~{\rm ergs~s^{-1}~cm^{2}}$. As we will discuss further below, the inferred 3--9\,keV fluxes for the remaining three observations (Table~\ref{tab:ascalog}) are lower by a factor of two to ten than the fluxes of the RXTE observations discussed in \S\ref{sec:xte}. We chose to fit the ASCA data with a phenomenological model consisting of a multicolor blackbody spectrum plus a broken power law, considered with and without a narrow Gaussian line at $\approx 6.4$\,keV. These fits are similar to those performed for ASCA observations of the hard state of Cyg~X-1 (\def\astroncite##1##2{##1\ ##2}\@internalcite{ebisawa:96b}), which shows evidence for a weak and narrow Fe line with equivalent width $\approx 40$\,eV, as well as for a soft excess well-modeled as a multicolor blackbody with peak temperature $\approx 150$\,eV. \begin{table*} \caption{\small Parameters for a multicolor blackbody plus broken power law plus Gaussian line fits to ASCA data. \label{tab:ascafita}} \smallskip {\small \center{ \begin{tabular}{llllllllllll} \hline \hline \noalign{\vspace*{1mm}} Date & $T_{\rm in}$ & $A_{\rm dbb}$ & $\Gamma_1$ & $E_{\rm b}$ & $\Gamma_2$ & $A_{\rm bpl}$ & $E_{\rm l}$ & $A_{\rm l}$ & EW & $\chi^2$/dof & $\chi^2_{\rm red}$ \\ & (keV) & ($\times 10^4)$ & & (keV) & & ($\times 10^{-2}$) & (keV) & ($\times 10^{-4}$) & (eV) \\ \noalign{\vspace*{1mm}} \hline \noalign{\vspace*{1mm}} 1994 Aug. 24 & \errtwo{0.14}{0.01}{0.02} & \errtwo{2.2}{0.9}{1.6} & \errtwo{1.78}{0.03}{0.03} & \errtwo{3.4}{0.4}{0.5} & \errtwo{1.62}{0.04}{0.03} & \errtwo{4.2}{0.1}{0.1} & \nodata & \nodata & \nodata & 1500/1439 & 1.04 \\ \noalign{\vspace*{1mm}} 1994 Aug. 24 & \errtwo{0.14}{0.01}{0.02} & \errtwo{2.2}{0.7}{1.6} & \errtwo{1.78}{0.03}{0.03} & \errtwo{3.3}{0.5}{0.6} & \errtwo{1.64}{0.03}{0.03} & \errtwo{4.2}{0.1}{0.1} & {\it 6.4} & \errtwo{0.6}{0.4}{0.3} & \errtwo{34}{25}{19} & 1494/1438 & 1.04 \\ \noalign{\vspace*{1mm}} 1994 Sept. 12 & \errtwo{0.15}{0.01}{0.01} & \errtwo{2.5}{0.6}{1.0} & \errtwo{1.81}{0.02}{0.01} & \errtwo{3.8}{0.2}{0.2} & \errtwo{1.56}{0.03}{0.03} & \errtwo{7.2}{0.1}{0.1} & \nodata & \nodata & \nodata & 1603/1621 & 0.99 \\ \noalign{\vspace*{1mm}} 1994 Sept. 12 & \errtwo{0.15}{0.00}{0.00} & \errtwo{2.5}{0.1}{0.4} & \errtwo{1.81}{0.01}{0.01} & \errtwo{3.8}{0.1}{0.1} & \errtwo{1.59}{0.01}{0.02} & \errtwo{7.2}{0.0}{0.1} & \errtwo{6.36}{0.08}{0.09} & \errtwo{1.6}{0.7}{0.2} & \errtwo{56}{26}{7} & 1580/1619 & 0.98 \\ \noalign{\vspace*{1mm}} 1995 Sept. 08 & \errtwo{0.19}{0.00}{0.01} & \errtwo{2.4}{0.1}{0.3} & \errtwo{1.93}{0.02}{0.01} & \errtwo{3.7}{0.1}{0.1} & \errtwo{1.60}{0.01}{0.02} & \errtwo{25.6}{0.3}{0.3} & \nodata & \nodata & \nodata & 2597/1838 & 1.41 \\ \noalign{\vspace*{1mm}} 1995 Sept. 08 & \errtwo{0.19}{0.00}{0.00} & \errtwo{2.4}{0.0}{0.1} & \errtwo{1.93}{0.01}{0.00} & \errtwo{3.7}{0.1}{0.0} & \errtwo{1.63}{0.01}{0.01} & \errtwo{25.6}{0.0}{0.1} & \errtwo{6.51}{0.07}{0.06} & \errtwo{3.3}{0.6}{0.5} & \errtwo{40}{7}{6} & 2523/1836 & 1.37 \\ \noalign{\vspace*{1mm}} \hline \end{tabular} } } \tablecomments{$T_{\rm in}$: peak multicolor blackbody temperature. $A_{\rm dbb}$: multicolor blackbody normalization. $\Gamma_1$, $\Gamma_2$: broken power law photon indices. $E_{\rm b}$: break energy. $A_{\rm bpl}$: Power law normalization (photons\,keV$^{-1}$\,cm$^{-2}$\,s$^{-1}$ at 1\,keV). $E_{\rm l}$: line energy. $A_{\rm l}$: Line normalization (photons\,cm$^{-2}$\,s$^{-1}$ in the line). EW: line equivalent width. Uncertainties are at the 90\% confidence level for one interesting parameter ($\Delta \chi^2 = 2.71$). } \end{table*} \begin{figure*} \centerline{ \psfig{figure=Fig1.ps,width=0.55\textwidth,angle=270} } \caption{\small GX~339$-$4 ASCA observation of 1994 August~24 with energy bins rebinned by a factor of 25 for clarity. Model and associated residuals (data/model) are for the best fit multicolor blackbody plus broken power law \emph{without} a Gaussian line component. For clarity, only the SIS0 (circles) and GIS2 (triangles) data are shown.} \label{fig:ascaphem} \end{figure*} The fits with the phenomenological models yield $\chi^2_{\rm red}$ ranging from $0.98$ to $1.4$. The brightest data set showed the greatest evidence for structure beyond this simple model. A sample fit is shown in Figure~\ref{fig:ascaphem}. Note that the neutral hydrogen column was fixed to $6 \times 10^{21}~{\rm cm}^{-2}$. Allowed to freely vary, the neutral hydrogen column tended to float between $4$ and $8 \times 10^{21}~{\rm cm}^{-2}$, depending upon what combination of phenomenological models was chosen, with minimal changes in the $\chi^2$ of the fits. Associated with these changes in best fit neutral hydrogen column were $\aproxgt \pm 30\%$ changes of the best fit peak temperature of the multicolor blackbody and even larger changes (factors of $\approx 3$) in the best fit normalization of the multicolor blackbody component. We should thus associate systematic error bars with these two parameters that are somewhat larger than the statistical error bars presented in Table~\ref{tab:ascafita}. All fits improved with the addition of a narrow Gaussian line. In all fits we fixed the line width to $0.1$\,keV (see \def\astroncite##1##2{##1\ ##2}\@internalcite{ebisawa:96b}; who always found $\sigma < 0.1$\,keV in fits to ASCA data of Cyg~X-1), and for the lowest flux data set we also fix the line energy to 6.4\,keV. For the lowest flux data set the $\Delta \chi^2 = 5.5$ for one additional parameter. By the F-test (\def\astroncite##1##2{##1\ ##2}\@internalcite{bevington}), this is an improvement to the fit for one additional parameter at the 98\% confidence level. The other two data sets show even more significant improvements to the $\chi^2$. The best fit line equivalent widths ranged from $\approx 30$ to 60\,eV. There is no compelling evidence for a strong flux dependence to the equivalent width of the line. The transition to the bright, soft state typically occurs at 3--9\,keV luminosities $\aproxgt 10^{-9}~{\rm ergs~cm^{-2}~s^{-1}}$; i.e., a factor of two to ten brighter than these ASCA observations. Thus these observations offer useful tests of ADAF models, which are hypothesized to be most relevant to low-luminosity, hard state systems (\def\astroncite##1##2{##1\ ##2}\@internalcite{narayan:96e,esin:97c}). ADAF models predict a detectable correlation between the temperature of the soft excess, the strength of the Fe line, and the source luminosity. They hypothesize that the luminosity decay of BH transients is due, in part, to an increase of the radius at which the accretion flow transits from cold, geometrically thin, and radiatively efficient to hot, geometrically thick, and advective (\def\astroncite##1##2{##1\ ##2}\@internalcite{esin:97c}). In some models, the transition radius can grow to as large as ${\cal O}(10^4~R_{\rm G})$, where $R_{\rm G} \equiv GM/c^2$. (Such large transition radii are \emph{not} a strict requirement of ADAF models; in \S\ref{sec:adaf} we show that somewhat smaller transition radii, $\approx 200$--$400~R_{\rm G}$, are preferred for ADAF models of the RXTE data.) As discussed by \def\astroncite##1##2{##1\ (##2)}\@internalcite{esin:97c}, one then expects the peak temperature of the soft excess to decrease below $150$\,eV and the equivalent width of any Fe line to decrease to values less than $\approx 30$\,eV. The best fit equivalent widths found for GX~339$-$4 are greater than can be accommodated in ADAF models with a large transition radius, and they are also slightly larger than predicted by the `sphere+disk' corona models described in \S\ref{sec:corona} (see also \def\astroncite##1##2{##1\ ##2}\@internalcite{dove:97b,dove:98a}). These latter models have a similar geometry to the ADAF models, and they often posit a coronal radius $\aproxlt 100~R_{\rm G}$. Likewise, we do not detect any large decreases in the best-fit disk temperatures with decreasing luminosity. Although it is dangerous to make a one-to-one correspondence between a phenomenological fit component and a \emph{physical} component, these best-fit values are suggestive of, but not definitive proof of, temperatures hotter than can be accommodated in models where cold, soft X-ray emitting material exists at very large radii. \section{RXTE Observations}\label{sec:xte} \subsection{The Monitoring Campaign}\label{sec:asm} To study the long-term behavior of GX~339$-$4, and to place our pointed observations within the context of the overall behavior of the source, we used data from the All Sky Monitor (ASM) on RXTE. The ASM provides lightcurves in three energy bands, 1.3--3.0\,keV, 3.0--5.0\,keV, and 5.0--12.2\,keV, typically consisting of several 90\,s measurements per day (see \def\astroncite##1##2{##1\ ##2}\@internalcite{levine:96a,remillard:97a,lochner:97a}). In Figure~\ref{fig:asm} we present the ASM data of GX~339$-$4 up until Truncated Julian Date (TJD) $\approx 1000$ (1998 July 6). We also indicate the dates of our RXTE observations, as well as the dates of ATCA and MOST radio observations (\def\astroncite##1##2{##1\ ##2}\@internalcite{fender:97a,corbel:98a,hanni:98a}). We discuss the long timescale variability of this lightcurve in paper~II. \begin{figure*} \centerline{ \psfig{figure=Fig2a.eps,width=0.45\textwidth} \psfig{figure=Fig2b.eps,width=0.45\textwidth} } \caption{\small \emph{Left:} RXTE All Sky Monitor data for GX~339$-$4 (5 day averages in the 1.3--12.2\,keV band) vs. Truncated Julian Date (TJD) $\equiv$ Julian Date (JD) $-2450000.5$. Dashes indicate dates of our RXTE pointed observations, diamonds indicate dates of MOST radio observations (Hannikainnen et al. 1998), and triangles indicate dates of ATCA radio observations (Fender et al. 1997, Corbel et al. 1998). \emph{Right:} ASM colors for GX~339$-$4 vs TJD. Colors shown are: (1.3--3\,keV)/(3-5\,keV), (3--5\,keV)/(5--12\,keV)+5, and (1.3--3\,keV)/(5--12.2\,keV)+10 (the latter two ratios have been offset for clarity).} \label{fig:asm} \end{figure*} \begin{table*} \caption{\small Approximate expected ASM colors for the different states of GX~339$-$4. \label{tab:asmcol}} \smallskip \center{ \begin{tabular}{llll} \hline \hline \noalign{\vspace*{1mm}} State & count rate & 1.3--3\,keV/ & 1.3--3\,keV/ \\ & cts\,s$^{-1}$ & 3--5\,keV & 5--12.2\,keV \\ \noalign{\vspace*{1mm}} \hline \noalign{\vspace*{1mm}} low & 7 & $2$ & $2$ \\ \noalign{\vspace*{1mm}} high & 15 & $4$ & $30$\\ \noalign{\vspace*{1mm}} very high & 60 & $4$ & $15$\\ \noalign{\vspace*{1mm}} \hline \end{tabular}} \end{table*} Based upon model fits to the observations of \def\astroncite##1##2{##1\ (##2)}\@internalcite{grebenev:91a} (low and high state), and \def\astroncite##1##2{##1\ (##2)}\@internalcite{miyamoto:91a} (very high state), we expect the different states of GX~339$-$4 to have ASM count rates as indicated in Table~\ref{tab:asmcol}. The ASCA and RXTE observations discussed here are most characteristic of weak to average luminosity hard states. Confirmation that the eight RXTE observations taken between TJD 481 and 749 do indeed represent a typical low/hard state comes from the broad band spectral analysis presented in \S\ref{sec:xte}, as well as from the timing analysis presented in paper II. The X-ray variability of these observations show root mean square variability of ${\cal O}(30\%)$ and show a power spectrum (PSD) that, roughly, is flat below 0.1\,Hz, $\propto f^{-1}$ between 0.1--3\,Hz, and $\propto f^{-2}$ above 3\,Hz. Time lags and coherence functions were also comparable to previously observed hard states of Cygnus X-1 (see paper II, and references therein). Further discussion and analyses of the timing data can be found in paper II. The transition to a higher flux level that occurs at TJD $\approx 800$ appears to have a characteristic flux of a high state, but is not as soft in the 2--10\,keV bands as expected from the above cited high and very high states. This might be an example of what \def\astroncite##1##2{##1\ (##2)}\@internalcite{mendez:97a} refer to as an `intermediate state' between hard and soft. No pointed RXTE observations were taken during the transition, and four pointed observations, which were not part of our monitoring campaign, occured shortly after the transition. Detailed confirmation of the spectral state suggested by the ASM data awaits analysis of these pointed observations. Note that the radio flux became quenched over the course of this transition to a higher ASM flux level (\def\astroncite##1##2{##1\ ##2}\@internalcite{hanni:98a}). The variations observed in both the ASM lightcurve (prior to TJD $\approx 800$) and the pointed RXTE observations discussed below represent more than a factor five variation in observed flux. Comparable variations have been observed in the radio, and furthermore the radio lightcurves show evidence of a correlation with both the ASM and Burst and Transient Survey Explorer (BATSE) lightcurves (\def\astroncite##1##2{##1\ ##2}\@internalcite{hanni:98a}). \subsection{PCA and HEXTE Observations} In this section we present the results from our analysis of the data from the two pointed instruments on RXTE: the Proportional Counter Array (PCA), and the High Energy X-ray Timing Experiment (HEXTE). See appendix~\ref{sec:rxteapp} for a description of the instruments and of the details of the data extraction and processing. A log of the RXTE pointed observations and the simultaneous radio observations is given in Table~\ref{tab:log}. \begin{table*} \caption{\small Log of RXTE and radio observations. \label{tab:log}} \smallskip \centerline{\small \begin{tabular}{lllcccccccc} \hline \hline \noalign{\vspace*{1mm}} Obs & Date & TJD & $T$ & Rate & 3--9\,keV & 9--30\,keV & 30--100\,keV & ATCA & MOST & $\alpha$ \\ & & & (ksec) & (cps) & & (${\rm 10^{-9}~erg~cm^{-2}~s^{-1}}$) & & (mJy) & (mJy) \\ \noalign{\vspace*{1mm}} \hline \noalign{\vspace*{1mm}} 01 & 1997 Feb. 02 & 481 & 11 & 830 & 1.07 & 1.68 & 2.65 & $9.1\pm0.2$ & $7.0\pm0.7$ & 0.11 \\ \noalign{\vspace*{1mm}} 02 & 1997 Feb. 10 & 489 & 10 & 730 & 0.94 & 1.50 & 2.41 & $8.2\pm0.2$ & $6.3\pm0.7$ & 0.11 \\ \noalign{\vspace*{1mm}} 03 & 1997 Feb. 17 & 496 & 8 & 700 & 0.90 & 1.43 & 2.35 & $8.7\pm0.2$ & $6.1\pm0.7$ & 0.15 \\ \noalign{\vspace*{1mm}} 04 & 1997 Apr. 29 & 567 & 10 & 470 & 0.60 & 0.97 & 1.55 \\ \noalign{\vspace*{1mm}} 05 & 1997 Jul. 07 & 636 & 10 & 200 & 0.25 & 0.43 & 0.75 \\ \noalign{\vspace*{1mm}} 06 & 1997 Aug. 23 & 683 & 11 & 650 & 0.74 & 1.18 & 1.98 \\ \noalign{\vspace*{1mm}} 07 & 1997 Sep. 19 & 710 & 10 & 730 & 0.96 & 1.48 & 2.36 \\ \noalign{\vspace*{1mm}} 08 & 1997 Oct. 28 & 749 & 10 & 480 & 0.63 & 1.01 & 1.68 \\ \noalign{\vspace*{1mm}} \hline \end{tabular}} \tablecomments{We list: $T$, the duration of the RXTE observations; the average PCA count rate; the average (3--9\,keV), (9--30\,keV), and (30--100\,keV) energy fluxes (all normalized to the PCA calibration); the flux density of the 8.3--9.1\,GHz ATCA observations; the flux density of the 843\,MHz MOST observations; and $\alpha = \Delta \log \nu / \Delta \log S_\nu$, the spectral index of the radio observations (\protect\def\astroncite##1##2{##1\ ##2}\@internalcite{fender:97a,corbel:98a,hanni:98a}).} \end{table*} As we show in appendix~\ref{sec:rxteapp}, there is a difference in the power-law slopes obtained from an analysis of spectra of the Crab with both instruments upon RXTE. In order to minimize the impact of this difference in the instrumental calibration onto the data analysis, we primarily analyze the data from both instruments individually and use the difference in model parameters between instruments as a gauge of the systematic uncertainties. We do perform some joint analysis of PCA and HEXTE data using various reflection models. In the following sections we discuss in detail the implications of the calibration uncertainty for our analysis. For our analysis of the RXTE broad band spectrum, we used several different spectral models consistent with the range of parameterizations currently used in the literature to describe the spectra of BHC. As in the ASCA analysis (\S\ref{sec:asca}), we fixed $N_{\rm H} = 6 \times 10^{21}~{\rm cm^2}$. We first used the purely phenomenological exponentially cutoff power law and broken power law models as a broad characterization of the data. The results of this modeling are given in \S\ref{sec:phenom} and in Tables~\ref{tab:jwstud} and \ref{tab:jwstudb}. We then applied the three more physically motivated models that are currently discussed in the literature: reflection of a power law off an (ionized) accretion disk (\S\ref{sec:reflect} and Table~\ref{tab:jwstud2}), `sphere and disk' corona Comptonization models (\S\ref{sec:corona} and Table~\ref{tab:jwstud3}), and ADAF models (\S\ref{sec:adaf}). The ADAF models are only applied to the unfolded data. Residuals for each of the best fits to the data from Observations~5 and~7 are shown in Figures~\ref{fig:pcaalone},~\ref{fig:hextealone}, and \ref{fig:jointfit} (except for the ADAF models, where we present Observations 1 and 5). We chose to present these former two observations because not only are they at extremes in terms of observed luminosity (Observation~5 is the faintest, and Observation~7 is the second brightest, Table~\ref{tab:log}), but also because they show detectable differences in their timing properties (paper~II). \begin{figure*} \centerline{ \psfig{figure=Fig3a.eps,width=0.45\textwidth} \psfig{figure=Fig3b.eps,width=0.45\textwidth}} \caption{\small Spectral modeling of the PCA data from Observation~5 (left) and~7 (right). Residues are shown as the contribution to $\chi$. a) Count rate spectrum and the best-fit broken power-law with Gaussian line, b) Contribution to $\chi$ from the broken power-law with Gaussian line, c) Contribution to $\chi$ from the ionized reflection model (pexriv) with Gaussian line, d) Contribution to $\chi$ from the sphere+disk model with additional Gaussian line.} \label{fig:pcaalone} \end{figure*} \begin{figure*} \centerline{ \psfig{figure=Fig4a.eps,width=0.45\textwidth} \psfig{figure=Fig4b.eps,width=0.45\textwidth}} \caption{\small Spectral modeling of the HEXTE data from observation~5 (left) and~7 (right). Residues are shown as the contribution to $\chi$. a) Count rate spectrum and the best-fit power-law with exponential cutoff, b) Contribution to $\chi$ from the power-law with exponential cutoff, c) Contribution to $\chi$ from the sphere+disk model.} \label{fig:hextealone} \end{figure*} \begin{figure*} \centerline{ \psfig{figure=Fig5a.eps,width=0.45\textwidth} \psfig{figure=Fig5b.eps,width=0.45\textwidth}} \caption{\small Spectral modeling of the joint PCA and HEXTE data for Observation~5 and~7. Residues are shown as the contribution to $\chi$. a) Count rate spectrum and the best fit broken power-law with Gaussian line. b) Contribution to $\chi$ from the best fit broken power-law with a Gaussian line (parameters not given in text). c) Contribution to $\chi$ from the best fit ionized reflector model with a Gaussian line.} \label{fig:jointfit} \end{figure*} \subsubsection{Phenomenological Models}\label{sec:phenom} Results from the purely phenomenological fits to the data, i.e., the broken power law and exponentially cutoff power law, are presented in Tables~\ref{tab:jwstud} and \ref{tab:jwstudb}. We see that a broken power law plus a Gaussian models the PCA data very well. The low $\chi^2_{\rm red}$ (0.15--0.32) indicates that to some extent we may be fitting systematic features in the PCA response. The same may be true for the best fit parameters of the Fe line. The line widths ($\sigma \approx 0.6$\,keV) and equivalent widths ($\approx 130$\,eV) are larger than for the ASCA observations, with the exception of Observation~5 which has line parameters comparable to the ASCA observations. As Observation~5 has the lowest count rate, it is more dominated by statistical errors and less dominated by systematic errors than the other observations. Even ignoring the possible systematic effects, however, we see that any observed line is narrower and weaker than is commonly observed in AGN. \begin{table*} \caption{\small Parameters for Gaussian line plus broken power law models (PCA only).} {\small \begin{center} {Model: $\exp(-N_{\rm H}\sigma_a)\left( {\sf gauss}(E_{\rm line},\sigma,A_{\rm line}) + {\sf bknpower}(\Gamma_1,\Gamma_2,E_{\rm break},A_{\rm bkn}) \right)$}\\[3mm] \begin{tabular}{ccccccccccc} \hline \hline \noalign{\vspace*{0.7mm}} Obs & $E_{\rm line}$ & $\sigma $ & $A_{\rm line}$ & EW & $\Gamma_1 $ & $E_{\rm break}$ & $\Gamma_2$ & $A_{\rm bkn} $ & $\chi^2/{\rm dof}$ & $\chi^2_{\rm red}$ \\ & $\,({\rm keV})$ & $\,({\rm keV})$ & ($\times 10^{-3}$) & $\,({\rm eV})$ & & $\,({\rm keV})$ & & \\ \noalign{\vspace*{0.7mm}} \hline \noalign{\vspace*{0.7mm}} 01 & $\errtwa{ 6.48}{ 0.14}{ 0.15}$ & $\errtwa { 0.6}{ 0.2}{ 0.2}$ & $\errtwa{ 1.97}{ 0.42}{ 0.36}$ & $\errtwa{ 130}{24}{22}$ & $\errtwa{ 1.80}{ 0.01}{ 0.01}$ & $\errtwa{ 11.2}{ 0.3} { 0.4}$ & $\errtwa{ 1.53}{ 0.02}{ 0.02}$ & $\errtwa{ 0.44} { 0.01}{ 0.01}$ & 13.6/ 52 & 0.26 \\ \noalign{\vspace*{0.7mm}} \noalign{\vspace*{0.7mm}} 02 & $\errtwa{ 6.47}{ 0.15}{ 0.16}$ & $\errtwa { 0.6}{ 0.2}{ 0.2}$ & $\errtwa{ 1.73}{ 0.40}{ 0.35}$ & $\errtwa{129}{ 26}{24}$ & $\errtwa{ 1.80}{ 0.01}{ 0.01}$ & $\errtwa{ 10.9}{ 0.4} { 0.4}$ & $\errtwa{ 1.53}{ 0.02}{ 0.02}$ & $\errtwa{ 0.38} { 0.01}{ 0.01}$ & 16.6/52 & 0.32 \\ \noalign{\vspace*{0.7mm}} \noalign{\vspace*{0.7mm}} 03 & $\errtwa{ 6.47}{ 0.14}{ 0.15}$ & $\errtwa { 0.5}{ 0.2}{ 0.2}$ & $\errtwa{ 1.55}{ 0.29}{ 0.31}$ & $\errtwa{121}{24}{23}$ & $\errtwa{ 1.79}{ 0.01}{ 0.01}$ & $\errtwa{ 10.9}{ 0.4} { 0.4}$ & $\errtwa{ 1.53}{ 0.02}{ 0.03}$ & $\errtwa{ 0.37} { 0.01}{ 0.01}$ & 13.5/ 52 & 0.26 \\ \noalign{\vspace*{0.7mm}} \noalign{\vspace*{0.7mm}} 04 & $\errtwa{ 6.45}{ 0.08}{ 0.15}$ & $\errtwa { 0.5}{ 0.2}{ 0.2}$ & $\errtwa{ 1.05}{ 0.25}{ 0.22}$ & $\errtwa{121}{28}{23}$ & $\errtwa{ 1.78}{ 0.01}{ 0.01}$ & $\errtwa{ 10.8}{ 0.4} { 0.5}$ & $\errtwa{ 1.54}{ 0.03}{ 0.03}$ & $\errtwa{ 0.24} { 0.00}{ 0.00}$ & 20.7/ 52 & 0.38 \\ \noalign{\vspace*{0.7mm}} \noalign{\vspace*{0.7mm}} 05 & $\errtwa{ 6.43}{ 0.15}{ 0.17}$ & $\errtwa { 0.2}{ 0.3}{ 0.2}$ & $\errtwa{ 0.31}{ 0.09}{ 0.08}$ & $\errtwa{84}{25}{21}$ & $\errtwa{ 1.72}{ 0.01}{ 0.01}$ & $\errtwa{ 10.8}{ 0.8} { 0.8}$ & $\errtwa{ 1.49}{ 0.04}{ 0.05}$ & $\errtwa{ 0.09} { 0.00}{ 0.00}$ & 22.9/ 52 & 0.44 \\ \noalign{\vspace*{0.7mm}} \noalign{\vspace*{0.7mm}} 06 & $\errtwa{ 6.40}{ 0.14}{ 0.15}$ & $\errtwa { 0.5}{ 0.2}{ 0.2}$ & $\errtwa{ 1.32}{ 0.28}{ 0.27}$ & $\errtwa{123}{23}{23}$ & $\errtwa{ 1.80}{ 0.01}{ 0.01}$ & $\errtwa{ 10.9}{ 0.4} { 0.4}$ & $\errtwa{ 1.51}{ 0.03}{ 0.03}$ & $\errtwa{ 0.30} { 0.01}{ 0.01}$ & 25.2/ 52 & 0.48 \\ \noalign{\vspace*{0.7mm}} \noalign{\vspace*{0.7mm}} 07 & $\errtwa{ 6.45}{ 0.13}{ 0.13}$ & $\errtwa { 0.6}{ 0.2}{ 0.2}$ & $\errtwa{ 1.92}{ 0.38}{ 0.34}$ & $\errtwa{140}{23}{23}$ & $\errtwa{ 1.83}{ 0.01}{ 0.01}$ & $\errtwa{ 10.9}{ 0.3} { 0.4}$ & $\errtwa{ 1.55}{ 0.02}{ 0.02}$ & $\errtwa{ 0.42} { 0.01}{ 0.01}$ & 27.3/ 52 & 0.53 \\ \noalign{\vspace*{0.7mm}} \noalign{\vspace*{0.7mm}} 08 & $\errtwa{ 6.40}{ 0.14}{ 0.15}$ & $\errtwa { 0.5}{ 0.2}{ 0.2}$ & $\errtwa{ 1.20}{ 0.27}{ 0.23}$ & $\errtwa{130}{26}{23}$ & $\errtwa{ 1.79}{ 0.01}{ 0.01}$ & $\errtwa{ 10.8}{ 0.5} { 0.5}$ & $\errtwa{ 1.54}{ 0.03}{ 0.03}$ & $\errtwa{ 0.26} { 0.01}{ 0.01}$ & 28.3/ 52 & 0.54 \\ \noalign{\vspace*{0.7mm}} \hline \end{tabular} \end{center} " } \label{tab:jwstud} \end{table*} \begin{table*} \caption{\small Parameters for exponentially cutoff power law models (HEXTE only).} {\small \begin{center} {Model: ${\sf const.}\left( A_{\rm PL} E^{-\Gamma} \exp\left(-E/E_{\rm cut}\right) \right)$}\\[3mm] \begin{tabular}{cccccccc} \hline \hline \noalign{\vspace*{0.7mm}} Obs & ${\rm const.}$ & $\Gamma $ & $E_{\rm cut} $ & $A_{\rm PL} $ & ${\rm const.}$& $\chi^2/{\rm dof}$ & $\chi^2_{\rm red}$ \\ & & & $\,({\rm keV})$ & & & & \\ \noalign{\vspace*{0.7mm}} \hline \noalign{\vspace*{0.7mm}} 01 & ${\it 1.00}$ & $\errtwa{ 1.25}{ 0.06}{ 0.06}$ & $\errtwa{ 101} { 18}{ 14}$ & $\errtwa{ 0.082}{ 0.014}{ 0.012}$ & $\errtwa { 0.99}{ 0.01}{ 0.01}$ & 69.2/ 80 & 0.87 \\ \noalign{\vspace*{0.7mm}} 02 & ${\it 1.00}$ & $\errtwa{ 1.12}{ 0.08}{ 0.08}$ & $\errtwa{ 79} { 13}{ 10}$ & $\errtwa{ 0.052}{ 0.011}{ 0.009}$ & $\errtwa { 0.92}{ 0.02}{ 0.02}$ & 69.0/ 80 & 0.86 \\ \noalign{\vspace*{0.7mm}} 03 & ${\it 1.00}$ & $\errtwa{ 1.16}{ 0.17}{ 0.18}$ & $\errtwa{ 94} { 63}{ 28}$ & $\errtwa{ 0.052}{ 0.029}{ 0.019}$ & $\errtwa { 1.08}{ 0.04}{ 0.04}$ & 71.8/ 80 & 0.90 \\ \noalign{\vspace*{0.7mm}} 04 & ${\it 1.00}$ & $\errtwa{ 1.15}{ 0.11}{ 0.11}$ & $\errtwa{ 85} { 25}{ 16}$ & $\errtwa{ 0.035}{ 0.011}{ 0.009}$ & $\errtwa { 0.98}{ 0.02}{ 0.02}$ & 76.2/ 80 & 0.95 \\ \noalign{\vspace*{0.7mm}} 05 & ${\it 1.00}$ & $\errtwa{ 1.18}{ 0.19}{ 0.23}$ & $\errtwa{ 115} { 85}{ 46}$ & $\errtwa{ 0.016}{ 0.011}{ 0.007}$ & $\errtwa { 0.99}{ 0.05}{ 0.05}$ & 71.3/ 80 & 0.89 \\ \noalign{\vspace*{0.7mm}} 06 & ${\it 1.00}$ & $\errtwa{ 1.19}{ 0.09}{ 0.09}$ & $\errtwa{ 103} { 29}{ 19}$ & $\errtwa{ 0.049}{ 0.012}{ 0.010}$ & $\errtwa { 0.93}{ 0.02}{ 0.02}$ & 87.9/ 80 & 1.10 \\ \noalign{\vspace*{0.7mm}} 07 & ${\it 1.00}$ & $\errtwa{ 1.21}{ 0.07}{ 0.08}$ & $\errtwa{ 95} { 20}{ 14}$ & $\errtwa{ 0.066}{ 0.014}{ 0.012}$ & $\errtwa { 0.96}{ 0.02}{ 0.02}$ & 101.5/ 80 & 1.27 \\ \noalign{\vspace*{0.7mm}} 08 & ${\it 1.00}$ & $\errtwa{ 1.08}{ 0.10}{ 0.10}$ & $\errtwa{ 81} { 19}{ 13}$ & $\errtwa{ 0.031}{ 0.009}{ 0.007}$ & $\errtwa { 0.98}{ 0.02}{ 0.02}$ & 106.6/ 80 & 1.33 \\ \noalign{\vspace*{0.7mm}} \hline \end{tabular} \end{center} } \label{tab:jwstudb} \end{table*} The $\approx 3$--$10$\,keV spectral power-law slope is close to the `canonical value' of $\Gamma = 1.7$; however, the PCA shows evidence for a hardening of this spectral slope above $\approx 10$\,keV. HEXTE data alone also show the high energy spectrum to be harder than the 3--10\,keV spectrum (Fig.~\ref{fig:hextealone} and Table~\ref{tab:jwstudb}). Note that the difference between the PCA and HEXTE photon indices is \emph{greater} than the discrepancy between the PCA and HEXTE best fit Crab photon indices (appendix~\ref{sec:rxteapp}), and therefore it is unlikely to be a systematic effect. Such a hard HEXTE spectrum is consistent with previous observations by the Oriented Scintillation Spectrometer Experiment (OSSE) on board the Compton Gamma-ray Observatory (CGRO) (\def\astroncite##1##2{##1\ ##2}\@internalcite{grabelsky:95a}; see also \def\astroncite##1##2{##1\ ##2}\@internalcite{zdziarski:98a}). \def\astroncite##1##2{##1\ (##2)}\@internalcite{grabelsky:95a} found a slightly harder photon index of $\Gamma = 0.88$ and an exponential cutoff of $E_{\rm cut} = 68$\,keV, somewhat lower than observed here. Note, however, that the OSSE observations extended to $\approx 500$\,keV as opposed to the $\sim 110$\,keV of our HEXTE observation. Therefore, the HEXTE data for GX~339$-$4 do not strongly constrain the exponential rollover, and slightly harder power laws with lower exponential cutoffs are permitted. \subsubsection{Reflection Models}\label{sec:reflect} A spectral hardening above $\approx 7$\,keV is the expected signature of reflection of a hard continuum off of cold material (\def\astroncite##1##2{##1\ ##2}\@internalcite{magdziarz:95a}). \def\astroncite##1##2{##1\ (##2)}\@internalcite{ueda:94a} applied reflection models to \textsl{Ginga} data of GX~339$-$4 and found strong evidence of reflection, whereas \def\astroncite##1##2{##1\ (##2)}\@internalcite{grabelsky:95a} found no evidence of reflection in OSSE data. \def\astroncite##1##2{##1\ (##2)}\@internalcite{zdziarski:98a} jointly fit these simultaneously observed data sets and find that reflection models, albeit with a large Fe abundance, provide a very good description of the data. We have applied the models of \def\astroncite##1##2{##1\ (##2)}\@internalcite{magdziarz:95a}, as implemented in XSPEC (\emph{pexrav}, \emph{pexriv}), to the GX~339$-$4 data. These models consider an exponentially cutoff power law reflected off of neutral (\emph{pexrav}) or partially ionized (\emph{pexriv}) cold material. \begin{table*} \caption{\small Parameters for Gaussian line plus multicolor disk plus ionized reflection models after Magdziarz \& Zdziarski (1995). \label{tab:jwstud2}} {\small \begin{sideways} \begin{minipage}{0.95\textheight} \begin{center} {Model: ${\sf const.} \cdot \exp(-N_{\rm H}\sigma_{a}) \left ( {\sf diskbb}(A_{\rm dbb},kT_{\rm dbb}) + {\sf pexriv}(E_{\rm fold},\Gamma,f,z,A_{\rm X}, A_{\rm Fe},\cos i, A_{\rm pex}) + {\sf gauss}(E_{\rm line},\sigma, A_{\rm line}) \right ) $} \nl \begin{tabular}{ccccccccccccccc} \hline \hline \noalign{\vspace*{0.7mm}} Obs & $A_{\rm dbb}$ & $\Gamma$ & $\Gamma_{\rm HEXTE}$ & $f$ & $A_{\rm Fe}$ & $\xi$ & $A_{\rm pex}$ & $E_{\rm line}$ & $A_{\rm line}$ & EW & const. & const. & $\chi^2/$dof & $\chi^2_{\rm red}$ \nl & $(\times 10^{5})$ & & & & & $({\rm erg~cm~s^{-1}})$ & & (keV) & ($\times 10^{-4}$) & (eV) \nl \noalign{\vspace*{0.7mm}} \hline \noalign{\vspace*{0.7mm}} 01& $\errtwa{ 1.64}{ 0.82}{ 0.83}$ & $\errtwa{ 1.81}{ 0.02}{ 0.02}$ & --- & $ \errtwa{ 0.41}{ 0.06}{ 0.05}$ & $ \errtwa{ 1.54}{ 0.67}{ 0.40}$ & $ \errtwa{ 78.2}{ 28.7}{ 23.7}$ & $ \errtwa{ 0.43}{ 0.01}{ 0.01}$ & $ \errtwa{ 6.22}{ 0.22}{ 0.23}$ & $ \errtwa{ 7.00}{ 2.95}{ 3.10}$ & $ \errtwa{ 41}{ 18}{ 18}$ & --- & --- & $ 25./51$ & 0.49 \nl \noalign{\vspace*{0.7mm}} 01& $ \errtwa{ 1.54}{ 0.81}{ 0.82}$ & $\errtwa{ 1.81}{ 0.02}{ 0.02}$ & $\errtwa{ 1.76}{ 0.04}{ 0.03}$ & $ \errtwa{ 0.47}{ 0.07}{ 0.06}$ & $ \errtwa{ 2.26}{ 1.01}{ 0.65}$ & $ \errtwa{ 60.5}{ 20.8}{ 21.8}$ & $ \errtwa{ 5.85}{ 2.75}{ 2.98}$ & {\it 6.4} & $ \errtwa{ 5.85}{ 2.75}{ 2.98}$ & $ \errtwa{ 35}{ 18}{ 17}$ & $ \errtwa{ 0.57}{ 0.09}{ 0.07}$ & $ \errtwa{ 0.57}{ 0.09}{ 0.07}$ & $ 113./132$ & 0.86 \nl \noalign{\vspace*{0.7mm}} \noalign{\vspace*{0.7mm}} 02& $\errtwa{ 1.68}{ 0.73}{ 0.73}$ & $\errtwa{ 1.81}{ 0.02}{ 0.02}$ & --- & $ \errtwa{ 0.41}{ 0.06}{ 0.05}$ & $ \errtwa{ 1.36}{ 0.58}{ 0.36}$ & $ \errtwa{ 82.2}{ 31.6}{ 23.9}$ & $ \errtwa{ 0.37}{ 0.01}{ 0.01}$ & $ \errtwa{ 6.16}{ 0.23}{ 0.25}$ & $ \errtwa{ 5.96}{ 2.60}{ 2.82}$ & $ \errtwa{ 40}{ 17}{ 19}$ & --- & --- & $ 21./51$ & 0.41 \nl \noalign{\vspace*{0.7mm}} 02& $ \errtwa{ 1.49}{ 0.73}{ 0.73}$ & $\errtwa{ 1.80}{ 0.02}{ 0.02}$ & $\errtwa{ 1.75}{ 0.05}{ 0.04}$ & $ \errtwa{ 0.52}{ 0.09}{ 0.07}$ & $ \errtwa{ 2.46}{ 1.50}{ 0.80}$ & $ \errtwa{ 54.0}{ 21.3}{ 22.5}$ & $ \errtwa{ 4.08}{ 2.45}{ 2.68}$ & {\it 6.4} & $ \errtwa{ 4.08}{ 2.45}{ 2.68}$ & $ \errtwa{ 18}{ 28}{ 10}$ & $ \errtwa{ 0.57}{ 0.12}{ 0.08}$ & $ \errtwa{ 0.52}{ 0.11}{ 0.07}$ & $ 124./132$ & 0.94 \nl \noalign{\vspace*{0.7mm}} \noalign{\vspace*{0.7mm}} 03& $\errtwa{ 1.34}{ 0.70}{ 0.69}$ & $\errtwa{ 1.81}{ 0.02}{ 0.02}$ & --- & $ \errtwa{ 0.42}{ 0.06}{ 0.05}$ & $ \errtwa{ 1.49}{ 0.62}{ 0.42}$ & $ \errtwa{ 59.8}{ 24.1}{ 24.8}$ & $ \errtwa{ 0.36}{ 0.01}{ 0.01}$ & $ \errtwa{ 6.22}{ 0.24}{ 0.27}$ & $ \errtwa{ 5.56}{ 2.53}{ 2.66}$ & $ \errtwa{ 39}{ 16}{ 18}$ & --- & --- & $ 14./51$ & 0.28 \nl \noalign{\vspace*{0.7mm}} 03& $ \errtwa{ 1.23}{ 0.70}{ 0.70}$ & $\errtwa{ 1.81}{ 0.02}{ 0.02}$ & $\errtwa{ 1.68}{ 0.05}{ 0.05}$ & $ \errtwa{ 0.44}{ 0.07}{ 0.06}$ & $ \errtwa{ 1.61}{ 0.76}{ 0.46}$ & $ \errtwa{ 50.9}{ 22.6}{ 22.9}$ & $ \errtwa{ 4.96}{ 2.46}{ 2.56}$ & {\it 6.4} & $ \errtwa{ 4.96}{ 2.46}{ 2.56}$ & $ \errtwa{ 36}{ 19}{ 18}$ & $ \errtwa{ 0.42}{ 0.09}{ 0.07}$ & $ \errtwa{ 0.45}{ 0.09}{ 0.07}$ & $ 89./132$ & 0.68 \nl \noalign{\vspace*{0.7mm}} \noalign{\vspace*{0.7mm}} 04& $\errtwa{ 1.13}{ 0.47}{ 0.48}$ & $\errtwa{ 1.79}{ 0.02}{ 0.02}$ & --- & $ \errtwa{ 0.37}{ 0.07}{ 0.06}$ & $ \errtwa{ 1.33}{ 0.77}{ 0.43}$ & $ \errtwa{ 70.2}{ 31.3}{ 27.3}$ & $ \errtwa{ 0.23}{ 0.00}{ 0.01}$ & $ \errtwa{ 6.17}{ 0.23}{ 0.23}$ & $ \errtwa{ 4.56}{ 1.80}{ 1.90}$ & $ \errtwa{ 47}{ 18}{ 19}$ & --- & --- & $ 15./51$ & 0.30 \nl \noalign{\vspace*{0.7mm}} 04& $ \errtwa{ 1.03}{ 0.47}{ 0.47}$ & $\errtwa{ 1.79}{ 0.02}{ 0.02}$ & $\errtwa{ 1.73}{ 0.07}{ 0.05}$ & $ \errtwa{ 0.46}{ 0.12}{ 0.08}$ & $ \errtwa{ 2.26}{ 1.78}{ 0.86}$ & $ \errtwa{ 44.2}{ 24.1}{ 23.7}$ & $ \errtwa{ 3.61}{ 1.60}{ 1.88}$ & {\it 6.4} & $ \errtwa{ 3.61}{ 1.60}{ 1.88}$ & $ \errtwa{ 39}{ 18}{ 20}$ & $ \errtwa{ 0.54}{ 0.14}{ 0.09}$ & $ \errtwa{ 0.52}{ 0.14}{ 0.08}$ & $ 100./132$ & 0.76 \nl \noalign{\vspace*{0.7mm}} \noalign{\vspace*{0.7mm}} 05& $\errtwa{ 0.39}{ 0.23}{ 0.24}$ & $\errtwa{ 1.73}{ 0.03}{ 0.02}$ & --- & $ \errtwa{ 0.43}{ 0.49}{ 0.13}$ & $ \errtwa{ 2.43}{11.04}{ 1.35}$ & $ \errtwa{ 10.1}{ 35.8}{ 10.1}$ & $ \errtwa{ 0.09}{ 0.00}{ 0.00}$ & $ \errtwa{ 6.32}{ 0.27}{ 0.27}$ & $ \errtwa{ 1.84}{ 1.06}{ 0.94}$ & $ \errtwa{ 48}{ 29}{ 25}$ & --- & --- & $ 18./51$ & 0.34 \nl \noalign{\vspace*{0.7mm}} 05& $ \errtwa{ 0.36}{ 0.22}{ 0.23}$ & $\errtwa{ 1.73}{ 0.03}{ 0.02}$ & $\errtwa{ 1.65}{ 0.12}{ 0.08}$ & $ \errtwa{ 0.44}{ 0.24}{ 0.12}$ & $ \errtwa{ 2.41}{ 3.49}{ 1.22}$ & $ \errtwa{ 6.1}{ 25.2}{ 6.1}$ & $ \errtwa{ 1.87}{ 1.09}{ 0.96}$ & {\it 6.4} & $ \errtwa{ 1.87}{ 1.09}{ 0.96}$ & $ \errtwa{ 50}{ 30}{ 48}$ & $ \errtwa{ 0.49}{ 0.24}{ 0.12}$ & $ \errtwa{ 0.49}{ 0.23}{ 0.12}$ & $ 88./132$ & 0.67 \nl \noalign{\vspace*{0.7mm}} \noalign{\vspace*{0.7mm}} 06& $\errtwa{ 1.61}{ 0.57}{ 0.57}$ & $\errtwa{ 1.80}{ 0.02}{ 0.02}$ & --- & $ \errtwa{ 0.47}{ 0.10}{ 0.07}$ & $ \errtwa{ 1.87}{ 1.05}{ 0.56}$ & $ \errtwa{ 55.0}{ 24.0}{ 24.2}$ & $ \errtwa{ 0.29}{ 0.00}{ 0.01}$ & $ \errtwa{ 6.08}{ 0.22}{ 0.23}$ & $ \errtwa{ 5.45}{ 2.16}{ 2.30}$ & $ \errtwa{ 45}{ 18}{ 19}$ & --- & --- & $ 17./51$ & 0.33 \nl \noalign{\vspace*{0.7mm}} 06& $ \errtwa{ 1.46}{ 0.55}{ 0.55}$ & $\errtwa{ 1.81}{ 0.02}{ 0.02}$ & $\errtwa{ 1.71}{ 0.05}{ 0.04}$ & $ \errtwa{ 0.52}{ 0.10}{ 0.07}$ & $ \errtwa{ 2.38}{ 1.26}{ 0.74}$ & $ \errtwa{ 36.7}{ 22.0}{ 19.2}$ & $ \errtwa{ 4.06}{ 2.08}{ 2.05}$ & {\it 6.4} & $ \errtwa{ 4.06}{ 2.08}{ 2.05}$ & $ \errtwa{ 26}{ 30}{ 8}$ & $ \errtwa{ 0.50}{ 0.10}{ 0.07}$ & $ \errtwa{ 0.47}{ 0.09}{ 0.07}$ & $ 113./132$ & 0.86 \nl \noalign{\vspace*{0.7mm}} \noalign{\vspace*{0.7mm}} 07& $\errtwa{ 2.29}{ 0.73}{ 0.73}$ & $\errtwa{ 1.83}{ 0.02}{ 0.02}$ & --- & $ \errtwa{ 0.42}{ 0.06}{ 0.05}$ & $ \errtwa{ 1.37}{ 0.58}{ 0.37}$ & $ \errtwa{ 84.8}{ 32.5}{ 23.9}$ & $ \errtwa{ 0.40}{ 0.01}{ 0.01}$ & $ \errtwa{ 6.16}{ 0.18}{ 0.19}$ & $ \errtwa{ 8.36}{ 2.67}{ 2.96}$ & $ \errtwa{ 54}{ 18}{ 19}$ & --- & --- & $ 25./51$ & 0.48 \nl \noalign{\vspace*{0.7mm}} 07& $ \errtwa{ 2.10}{ 0.72}{ 0.72}$ & $\errtwa{ 1.83}{ 0.02}{ 0.02}$ & $\errtwa{ 1.74}{ 0.04}{ 0.03}$ & $ \errtwa{ 0.48}{ 0.07}{ 0.06}$ & $ \errtwa{ 1.97}{ 0.93}{ 0.57}$ & $ \errtwa{ 61.5}{ 21.9}{ 22.7}$ & $ \errtwa{ 6.23}{ 2.76}{ 2.48}$ & {\it 6.4} & $ \errtwa{ 6.23}{ 2.76}{ 2.48}$ & $ \errtwa{ 43}{ 20}{ 17}$ & $ \errtwa{ 0.50}{ 0.08}{ 0.06}$ & $ \errtwa{ 0.48}{ 0.08}{ 0.06}$ & $ 144./132$ & 1.09 \nl \noalign{\vspace*{0.7mm}} \noalign{\vspace*{0.7mm}} 08& $\errtwa{ 1.25}{ 0.47}{ 0.47}$ & $\errtwa{ 1.80}{ 0.02}{ 0.02}$ & --- & $ \errtwa{ 0.38}{ 0.06}{ 0.05}$ & $ \errtwa{ 1.36}{ 0.67}{ 0.42}$ & $ \errtwa{ 72.6}{ 30.7}{ 26.6}$ & $ \errtwa{ 0.25}{ 0.01}{ 0.01}$ & $ \errtwa{ 6.15}{ 0.19}{ 0.20}$ & $ \errtwa{ 5.55}{ 1.95}{ 1.87}$ & $ \errtwa{ 55}{ 19}{ 18}$ & --- & --- & $ 18./51$ & 0.35 \nl \noalign{\vspace*{0.7mm}} 08& $ \errtwa{ 1.14}{ 0.47}{ 0.47}$ & $\errtwa{ 1.80}{ 0.02}{ 0.02}$ & $\errtwa{ 1.71}{ 0.07}{ 0.05}$ & $ \errtwa{ 0.50}{ 0.15}{ 0.09}$ & $ \errtwa{ 2.81}{ 2.23}{ 1.11}$ & $ \errtwa{ 39.0}{ 24.8}{ 21.7}$ & $ \errtwa{ 4.14}{ 1.89}{ 1.72}$ & {\it 6.4} & $ \errtwa{ 4.14}{ 1.89}{ 1.72}$ & $ \errtwa{ 44}{ 20}{ 19}$ & $ \errtwa{ 0.51}{ 0.14}{ 0.09}$ & $ \errtwa{ 0.50}{ 0.14}{ 0.09}$ & $ 141./132$ & 1.07 \nl \noalign{\vspace*{0.7mm}} \noalign{\vspace*{0.7mm}} \hline \end{tabular} \end{center} \end{minipage} \end{sideways} } \vskip -0.5 true in \tablecomments{Models have been fit to: PCA data only (51~dof), and PCA and HEXTE data where the photon index of the reflected power law has been allowed to differ between the PCA and the HEXTE data (132~dof). Fit parameters are: the normalization of the multicolor disk blackbody, $A_{\rm dbb}$ (the disk temperature has been fixed to 250\,eV); the power law slope, $\Gamma$; cutoff energy, $E_{\rm fold}$; relative reflection fraction, $f\equiv \Delta \Omega/2 \pi$; Fe abundance relative to solar, $A_{\rm Fe}$; the disk ionization parameter, $\xi\equiv$ luminosity/(density $\times$ radius$^2$); and a normalization, $A_{\rm pex}$. The ionized reflection models also have parameters for the temperature of the disk, $T_{\rm disk}$ (fixed at $10^6$\,K); abundances of elements heavier than He relative to solar, $A_{\rm X}$ (fixed at unity); and disk inclination angle, $i$ (fixed at $45^\circ$). Parameters typeset in italics have been held constant for that particular fit. The iron line width was fixed at 0.1\,keV. } \end{table*} In Table~\ref{tab:jwstud2} we show the fit results for reflection off of partially ionized material similar to the models presented by \def\astroncite##1##2{##1\ (##2)}\@internalcite{zdziarski:98a}. Just as in \def\astroncite##1##2{##1\ (##2)}\@internalcite{zdziarski:98a}, we include a disk component where we fix the inner disk temperature to 250\,eV. As PCA does not usefully constrain models below 3\,keV, the disk component is not strongly constrained; typically the $\chi^2$ values were higher by 5--20 without this component. We also fix the reflector inclination angle at $45^\circ$, fix the disk temperature at $T_{\rm disk} = 10^6$\,K, freeze the abundances at solar, but let the Fe abundance be a free parameter. In all our fits we found that the Gaussian line width, $\sigma$, would tend to drift toward $0$, so we fixed $\sigma = 0.1$\,keV. For the combined PCA and HEXTE data, we also fixed the Gaussian line energy to 6.4\,keV. For fits to PCA data alone and joint PCA/HEXTE data, the exponential cutoff energy, $E_{\rm fold}$, would drift towards very large energy ($\gg 1000$\,keV). We therefore only considered pure power laws without cutoffs. \def\astroncite##1##2{##1\ (##2)}\@internalcite{zdziarski:98a} have argued that the high energy cutoff is sharper than exponential, which one would not expect to be strongly constrained by the combined PCA/HEXTE data. As for the \textsl{Ginga} data of GX~339$-$4 (\def\astroncite##1##2{##1\ ##2}\@internalcite{ueda:94a}), the PCA data alone were extremely well described by reflection models. Again, however, the extremely low $\chi^2_{\rm red}$ (as low as 0.28) makes us caution that these models might partly be fitting systematics in the PCA response. PCA and \textsl{Ginga} are also very similar instruments in terms of design, and so to some extent they should exhibit similar systematic effects (as discussed in appendix~\ref{sec:rxteapp}, the internal consistency of the PCA calibration is now as good as or better than that for the \textsl{Ginga} calibration.). Note that the best fit Fe line equivalent widths here are significantly smaller than those found with the purely phenomenological models discussed in \S\ref{sec:phenom}. The fits for the HEXTE data alone (not shown) were similar to the OSSE results of \def\astroncite##1##2{##1\ (##2)}\@internalcite{grabelsky:95a}. Namely, if one allows for an exponential cutoff (typically $\approx 100$\,keV) to the power law, the best fit reflection fraction becomes $f \aproxlt 0.01$. Such a small reflection fraction is not surprising considering how well a pure exponentially cutoff power law fits the HEXTE data (Table~\ref{tab:jwstudb}). If one does not allow an exponential cutoff, the reflection fraction becomes $f \aproxgt 3$. Such a fit is trying to mimic a hard power law with a high energy cutoff. A joint analysis of the PCA and the HEXTE data should be similar to a joint analysis of the Ginga and OSSE data. Indeed, such an analysis yields results comparable to those presented by \def\astroncite##1##2{##1\ (##2)}\@internalcite{zdziarski:98a} if we constrain the photon index of the incident power law to be the same for both the PCA and HEXTE data. Notable for the results of such fits (not presented) is the fairly large overabundance of Fe ($A_{\rm Fe} = 3.2$--$5.2$). Similarly, \def\astroncite##1##2{##1\ (##2)}\@internalcite{zdziarski:98a} find a large $A_{\rm Fe}=2.5$--$3.0$ \emph{except} for a short data set, more likely dominated by statistical errors rather than systematic errors, where they find $A_{\rm Fe}=1.6$--$2.0$. For our joint PCA/HEXTE data, the best-fit reflection fractions were approximately 20\% larger than the best-fit reflection fraction for PCA data alone. Such an increase in reflection fraction in general will reproduce the spectral hardening seen in the HEXTE energy bands. Increasing the average best fit Fe abundance from $\langle A_{\rm Fe} \rangle = 1.6$ (PCA data alone) to $\langle A_{\rm Fe} \rangle = 4.0$ (joint PCA/HEXTE data) also leads to an increased spectral hardening above $\approx 7$\,keV, while leaving the spectrum below $\approx 7$\,keV relatively unchanged. That is, such a fit helps to reproduce the spectral break at $\approx 10$\,keV. For the joint PCA/HEXTE analysis, there is clearly a worry that these results are influenced by the systematic differences between the PCA and the HEXTE responses. We therefore performed reflection model fits where we constrained all fit parameters to be the same for the PCA and the HEXTE data \emph{except} for the incident power law photon index, which we allowed to vary between the two instruments\footnote{The photon index was constrained to be the same for HEXTE Cluster~A and~B. The necessary different normalizations between the PCA and the HEXTE models were subsumed into the constants multiplying the HEXTE models. As HEXTE requires a harder power law, these constants were now $0.42$--$0.57$, as opposed to $\approx 0.7$. Furthermore, the constants showed larger uncertainties, as the uncertainty of the HEXTE photon index now couples strongly to the value of the constants.}. Such models provided reasonably good fits to the data, with $\chi^2_{\rm red}$ ranging from 0.67 to 1.09. The difference between the PCA and the HEXTE best fit photon indices ranged from 0.05 to 0.13, with an average value of 0.08. This is consistent with the systematic difference between the best-fit photon indeces for the Crab spectrum. For these models we find $\langle A_{\rm Fe} \rangle = 2.3$, which is more consistent with the results for fits to the PCA data only, and is slightly smaller than the results found by \def\astroncite##1##2{##1\ (##2)}\@internalcite{zdziarski:98a}. Note that we also find smaller values of the ionization parameter, $\xi$, than were found by \def\astroncite##1##2{##1\ (##2)}\@internalcite{zdziarski:98a}. \subsubsection{Corona Models}\label{sec:corona} We considered `sphere+disk' Comptonization models (\def\astroncite##1##2{##1\ ##2}\@internalcite{dove:97b}) of the GX~339$-$4 observations. We have previously applied these models to an RXTE observation of Cygnus~X--1 (\def\astroncite##1##2{##1\ ##2}\@internalcite{dove:98a}). The models consist of a central, spherical corona surrounded by a geometrically thin, flat disk. Seed photons for Comptonization come from the disk, which has a radial temperature distribution $kT_{\rm disk}(R) \propto R^{-3/4}$ and a temperature of 150\,eV at the inner edge of the disk. Hard flux from the corona further leads to reflection features from the disk or to soft photons due to thermalization of the hard radiation. The (non-uniform) temperature and pair balance within the corona are self-consistently calculated from the radiation field (\def\astroncite##1##2{##1\ ##2}\@internalcite{dove:97b}). As described by \def\astroncite##1##2{##1\ (##2)}\@internalcite{dove:97b}, we parameterize our models by the coronal compactness \begin{equation} \ell_{\rm c} \equiv \frac{\sigma_{\rm T}}{m_{\rm e} c^3}\frac{L_{\rm C}}{R_{\rm C}} ~~, \label{eq:compact} \end{equation} where $\sigma_{\rm T}$ is the Thomson cross section, $m_{\rm e}$ is the electron mass, $L_{\rm C}$ is the luminosity of the corona, and $R_{\rm C}$ is the radius of the corona. Likewise, we define a disk compactness, $\ell_{\rm d} \equiv (1-f_{\rm c}) (\sigma_{\rm T}/m_{\rm e} c^3) P_{\rm G}/R_{\rm C}$, where $P_{\rm G}$ is the \emph{total} rate of gravitational energy dissipated in the system, and $f_{\rm c}$ is the fraction dissipated in the corona. In calculating the numerical models, we set $\ell_{\rm d} = 1$. Models with other values of $\ell_{\rm d}$ yield the same ranges of self-consistent coronal temperatures and opacities. In general $f_{\rm c} = \ell_{\rm c} / (\ell_{\rm d} + \ell_{\rm c})$ (\def\astroncite##1##2{##1\ ##2}\@internalcite{dove:97b}). Based upon the `sphere+disk' geometry, a fraction $f \approx 0.32$ of the coronal flux is absorbed by the disk (\def\astroncite##1##2{##1\ ##2}\@internalcite{dove:97b}). The models are further parameterized by an initial electron coronal optical depth, $\tau_{\rm c}$ (approximately equal to the total optical depth, as pair production is negligible for the parameters of interest to us), and a normalization constant $A_{\rm kot}$. From the best-fit compactness and optical depth, the average coronal temperature can be calculated \emph{a posteriori}. \begin{table*} \caption{\small Models of `sphere+disk' Comptonization plus a Gaussian line fit to PCA data only and HEXTE data only. \label{tab:jwstud3}} \smallskip {\small \begin{tabular}{cccccccccccc} \hline \hline \noalign{\vspace*{0.7mm}} Obs & $E_{\rm Line}$ & $\sigma$ & $A_{\rm Line}$ & EW & $l_{\rm c}$ & $\tau_{\rm c}$ & $A_{\rm kot}$ & ${\rm const.}$ & $kT_{\rm c}$ & $\chi^2/{\rm dof}$ & $\chi^2_{\rm red}$ \\ & (keV) & (keV) & ($\times 10^{-3}$) & (eV) & & & & & (keV) \\ \noalign{\vspace*{0.7mm}} \hline \noalign{\vspace*{0.7mm}} 01 & $\errtwa{ 6.39}{ 0.18}{ 0.17}$ & $\errtwa { 0.8}{ 0.2}{ 0.1}$ & $\errtwa{ 3.09}{ 0.60}{ 0.53}$ & $\errtwa{197}{30}{30}$ & $\errtwa{ 0.62}{ 0.05}{ 0.04}$ & $\errtwa{ 3.3} { 0.1}{ 0.1}$ & $\errtwa{ 2.38}{ 0.06}{ 0.08}$ & \nodata & $\errtwa{28.6}{0.4}{0.4}$ & 19.2/53 & 0.36 \\ \noalign{\vspace*{0.7mm}} 01 & \nodata & \nodata & \nodata & \nodata & $\errtwa{1.83}{0.17}{0.13}$ & $\errtwa{2.9}{0.5}{0.3}$ & $\errtwa{0.91}{0.09}{0.09}$ & $\errtwa{0.99}{0.01}{0.01}$ & $\errtwa{43.9}{8}{8}$ & 71.2/80 & 0.89 \\ \noalign{\vspace*{0.7mm}} 02 & $\errtwa{ 6.36}{ 0.19}{ 0.19}$ & $\errtwa { 0.8}{ 0.2}{ 0.1}$ & $\errtwa{ 2.69}{ 0.57}{ 0.50}$ & $\errtwa{194}{41}{36}$ & $\errtwa{ 0.63}{ 0.06}{ 0.05}$ & $\errtwa{ 3.3} { 0.1}{ 0.1}$ & $\errtwa{ 2.07}{ 0.06}{ 0.07}$ & \nodata & $\errtwa{27.9}{0.4}{0.4}$ & 15.7/53 & 0.30 \\ \noalign{\vspace*{0.7mm}} 02 & \nodata & \nodata & \nodata & \nodata & $\errtwa{1.96}{0.12}{0.12}$ & $\errtwa{3.6}{0.3}{0.3}$ & $\errtwa{0.73}{0.04}{0.03}$ & $\errtwa{0.92}{0.02}{0.02}$ & $\errtwa{34.0}{3}{2}$ & 67.6/80 & 0.84 \\ \noalign{\vspace*{0.7mm}} 03 & $\errtwa{ 6.36}{ 0.21}{ 0.16}$ & $\errtwa { 0.8}{ 0.2}{ 0.2}$ & $\errtwa{ 2.42}{ 0.44}{ 0.49}$ & $\errtwa{182}{33}{37}$ & $\errtwa{ 0.66}{ 0.06}{ 0.06}$ & $\errtwa{ 3.4} { 0.1}{ 0.1}$ & $\errtwa{ 1.95}{ 0.06}{ 0.06}$ & \nodata & $\errtwa{27.9}{0.4}{0.4}$ & 16.7/53 & 0.31 \\ \noalign{\vspace*{0.7mm}} 03 & \nodata & \nodata & \nodata & \nodata & $\errtwa{2.21}{0.48}{0.34}$ & $\errtwa{3.4}{0.7}{1.1}$ & $\errtwa{0.61}{0.17}{0.08}$ & $\errtwa{1.08}{0.04}{0.04}$ & $\errtwa{37.4}{26}{7}$ & 72.4/80 & 0.90 \\ \noalign{\vspace*{0.7mm}} 04 & $\errtwa{ 6.40}{ 0.22}{ 0.17}$ & $\errtwa { 0.8}{ 0.2}{ 0.2}$ & $\errtwa{ 1.53}{ 0.32}{ 0.34}$ & $\errtwa{174}{36}{39}$ & $\errtwa{ 0.70}{ 0.08}{ 0.07}$ & $\errtwa{ 3.6} { 0.1}{ 0.1}$ & $\errtwa{ 1.27}{ 0.05}{ 0.05}$ & \nodata & $\errtwa{26.6}{0.4}{0.4}$ & 13.5/53 & 0.25 \\ \noalign{\vspace*{0.7mm}} 04 & \nodata & \nodata & \nodata & \nodata & $\errtwa{2.09}{0.19}{0.22}$ & $\errtwa{3.4}{0.6}{0.5}$ & $\errtwa{0.46}{0.04}{0.04}$ & $\errtwa{0.98}{0.02}{0.02}$ & $\errtwa{36.9}{9}{6}$ & 78.1/80 & 0.98 \\ \noalign{\vspace*{0.7mm}} 05 & $\errtwa{ 6.28}{ 0.31}{ 0.97}$ & $\errtwa { 0.9}{ 0.3}{ 0.3}$ & $\errtwa{ 0.52}{ 0.73}{ 0.17}$ & $\errtwa{138}{100}{100}$ & $\errtwa{ 0.68}{ 0.15}{ 0.08}$ & $\errtwa{ 4.4} { 0.3}{ 0.1}$ & $\errtwa{ 0.53}{ 0.02}{ 0.02}$ & \nodata & $\errtwa{20.6}{0.4}{0.4}$ & 20.5/53 & 0.39 \\ \noalign{\vspace*{0.7mm}} 05 & \nodata & \nodata & \nodata & \nodata & $\errtwa{2.89}{0.78}{0.65}$ & $\errtwa{3.1}{1.1}{2.0}$ & $\errtwa{0.16}{0.03}{0.03}$ & $\errtwa{0.99}{0.05}{0.05}$ & $\errtwa{44.9}{103}{13}$ & 71.8/80 & 0.90 \\ \noalign{\vspace*{0.7mm}} 06 & $\errtwa{ 6.26}{ 0.24}{ 0.22}$ & $\errtwa { 0.9}{ 0.2}{ 0.2}$ & $\errtwa{ 2.06}{ 0.50}{ 0.46}$ & $\errtwa{186}{45}{42}$ & $\errtwa{ 0.66}{ 0.07}{ 0.05}$ & $\errtwa{ 3.6} { 0.1}{ 0.1}$ & $\errtwa{ 1.57}{ 0.04}{ 0.06}$ & \nodata & $\errtwa{26.2}{0.4}{0.4}$ & 16.5/53 & 0.31 \\ \noalign{\vspace*{0.7mm}} 06 & \nodata & \nodata & \nodata & \nodata & $\errtwa{2.33}{0.21}{0.18}$ & $\errtwa{3.1}{0.6}{0.4}$ & $\errtwa{0.54}{0.08}{0.04}$ & $\errtwa{0.93}{0.02}{0.02}$ & $\errtwa{42.3}{11}{8}$ & 88.7/80 & 1.11 \\ \noalign{\vspace*{0.7mm}} 07 & $\errtwa{ 6.40}{ 0.19}{ 0.18}$ & $\errtwa { 0.8}{ 0.2}{ 0.2}$ & $\errtwa{ 2.67}{ 0.56}{ 0.49}$ & $\errtwa{192}{41}{35}$ & $\errtwa{ 0.65}{ 0.06}{ 0.05}$ & $\errtwa{ 3.3} { 0.1}{ 0.1}$ & $\errtwa{ 2.06}{ 0.05}{ 0.07}$ & \nodata & $\errtwa{28.4}{0.4}{0.4}$ & 15.6/53 & 0.30 \\ \noalign{\vspace*{0.7mm}} 07 & \nodata & \nodata & \nodata & \nodata & $\errtwa{1.99}{0.13}{0.16}$ & $\errtwa{3.1}{0.4}{0.4}$ & $\errtwa{0.75}{0.09}{0.04}$ & $\errtwa{0.96}{0.02}{0.02}$ & $\errtwa{40.7}{10}{6}$ & 102.9/80 & 1.29 \\ \noalign{\vspace*{0.7mm}} 08 & $\errtwa{ 6.34}{ 0.23}{ 0.21}$ & $\errtwa { 0.9}{ 0.2}{ 0.2}$ & $\errtwa{ 1.74}{ 0.40}{ 0.40}$ & $\errtwa{189}{43}{43}$ & $\errtwa{ 0.71}{ 0.08}{ 0.07}$ & $\errtwa{ 3.7} { 0.1}{ 0.1}$ & $\errtwa{ 1.30}{ 0.05}{ 0.05}$ & \nodata & $\errtwa{25.9}{0.4}{0.4}$ & 15.5/53 & 0.29 \\ \noalign{\vspace*{0.7mm}} 08 & \nodata & \nodata & \nodata & \nodata & $\errtwa{2.37}{0.29}{0.23}$ & $\errtwa{3.7}{0.4}{0.5}$ & $\errtwa{0.43}{0.03}{0.03}$ & $\errtwa{0.98}{0.02}{0.02}$ & $\errtwa{34.6}{7}{5}$ & 109.0/80 & 1.36 \\ \noalign{\vspace*{0.7mm}} \hline \end{tabular} } \tablecomments{The three fit parameters of the Comptonization model are the compactness of the corona, $\ell_{\rm c}$, the coronal optical depth, $\tau_{\rm c}$, and a normalization constant, $A_{\rm kot}$. The Gaussian line is parameterized as in the previous tables. From the best fit parameters, the equivalent width of the line, EW, and the density averaged coronal temperature, $kT_{\rm c}$, are derived.} \end{table*} Attempts to fit these models to the joint PCA/HEXTE data failed. Typical $\chi^2_{\rm red}$ values, even allowing for the inclusion of an extra Gaussian line component, were $\aproxgt 1.3$. These fits showed a clear tendency for a hardening in the HEXTE band, and therefore we considered them to be influenced by the cross-calibration uncertainties between the PCA and the HEXTE instruments (Note that our previous fits of Cyg~X-1 used an earlier version of the PCA response where we applied 1.5\% systematic uncertainties across the \emph{entire} PCA band; these fits yielded $\chi^2_{\rm red} \approx 1.6$, \emph{without} considering an additional Gaussian component). We therefore considered `sphere+disk' models fit to the PCA and the HEXTE data separately. In Table~\ref{tab:jwstud3}, we present the best-fit parameters for these models applied to our GX~339$-$4 data. Although our numerical `sphere+disk' models do include reflection and a fluorescent Fe line (typical equivalent width $\approx 25$\,eV) from the disk, the PCA data showed residuals in the 5--7\,keV band, similar as in in our fits to RXTE data of Cyg~X--1 (\def\astroncite##1##2{##1\ ##2}\@internalcite{dove:98a}). We included an additional Gaussian component to our fits. The equivalent widths of the additional lines were $\approx 150$\,eV, and they appeared to be broad ($\sigma \approx 0.8$\,keV). This additional line may be attributable partly to uncertainties in the PCA response. For these fits, as well as for the reflection model fits, lines with energies significantly less than 6.4\,keV can be fit, and this is likely a systematic effect. Part of the discrepancy between the data and the model, however, is significant. As we have discussed for our fits to the RXTE Cyg~X--1 data, there are several possible physical interpretations for the additional required equivalent width: There may be an overlap between the disk and the sphere (our models invoke a sharp transition), the disk may be flared (we model a flat disk), the disk may have non-solar abundances, or alternatively one might invoke a `patchy disk' embedded in the corona (\def\astroncite##1##2{##1\ ##2}\@internalcite{zdziarski:98a}). The best-fit reflection fractions of $f \approx 0.4$--$0.5$ found above are further indication that our models may require an additional source of reflected flux. Allowing an additional Gaussian line component, the fits to the PCA data yield extremely low $\chi^2_{\rm red}$, which could be indicating that we are partly fitting systematic features in the PCA response. Note also that the PCA data fits yielded consistently larger optical depths and consistently lower compactness parameters than the HEXTE data fits. The latter was more significant, and is again indicative of the HEXTE response being harder than the PCA response. Both instruments yielded optical depths $\tau_{\rm c} \approx 3$--4; however, due to the discrepancy in the best-fit spectral slopes between the PCA and the HEXTE bands, the best-fit average coronal temperatures range from 21--30\,keV (PCA) to 34--45\,keV (HEXTE). \subsubsection{ADAF Models}\label{sec:adaf} The basic picture of mass accretion via an ADAF in the context of galactic BHC was introduced by \def\astroncite##1##2{##1\ (##2)}\@internalcite{ichimaru:77a} and has been elaborated upon in a series of papers by Narayan and collaborators (\def\astroncite##1##2{##1\ ##2}\@internalcite{narayan:97a,esin:97c}). The accretion flow is divided into two distinct zones: the inner part is modeled as a hot, optically thin ADAF similar in some respects to the spherical corona discussed above, while the outer part consists of a standard optically thick, geometrically thin disk. The transition radius between the two zones, $r_{\rm tr}=R_{tr}/R_{\rm G}$, is one of the model parameters. We compute the ADAF spectrum according to the procedure described by \def\astroncite##1##2{##1\ (##2)}\@internalcite{dimatteo:98a}. The electrons in an ADAF cool via three processes: bremsstrahlung, synchrotron radiation, and inverse Compton scattering. In addition we add the emission from a thin disk--- calculated as a standard multicolor blackbody--- and include the Compton reflection component due to the scattering of high energy photons incident on the disk. In the ADAF models discussed here, we fix the black hole mass to be $m \equiv M/M_\odot = 6$, assume the magnetic field to be in equipartition with thermal pressure ($\beta =0.5$), and set the standard Shakura-Sunyaev viscosity parameter (\def\astroncite##1##2{##1\ ##2}\@internalcite{shakura:73a}) to be $\alpha_{\rm SS} =0.3$. We normalize the accretion rate to $\dot m \equiv \dot M c^2 / L_{\rm Edd}$, where $L_{\rm Edd}$ is the Eddington luminosity of the source. The hard state corresponds to mass accretion rates $\dot{m} \le \dot{m}_{\rm crit}=10^{-2}$, where $\dot{m}_{\rm crit}$ is the critical value above which an ADAF no longer exists. As $\dot{m}$ increases towards $\dot{m}_{\rm crit}$, the scattering optical depth of the ADAF goes up which causes the spectrum to become harder and smoother. Most of the flux from the ADAF plus disk configuration is emitted around 100\,keV and the spectrum falls off at higher energies. The model spectrum changes mainly as a function of $r_{\rm tr}$ and $\dot{m}$. The various spectral states correspond to different values of these parameters. For example, \def\astroncite##1##2{##1\ (##2)}\@internalcite{esin:97c} attempt to explain the initial transition from soft to hard seen in the decay of Nova Muscae by a large change in $r_{\rm tr}$ (from $r_{\rm tr} \approx 10$ to $r_{\rm tr} \approx 10^4$), followed by an exponential decay in $\dot{m}$ for the subsequent evolution of this transient system. The ASCA data of GX~339$-$4 discussed in \S\ref{sec:asca} imply that comparably large changes in $r_{\rm tr}$ are not relevant to those observations. Here, however, unfolded RXTE data from Observation~1 and Observation~5, the brightest and faintest observation respectively, can be described by ADAF models with $r_{\rm tr} = 200$, $\dot m = 0.08$ and $r_{\rm tr} = 400$, $\dot m = 0.05$, respectively. These model spectra and RXTE spectra for Observations 1 and 5 unfolded with a cutoff broken power law plus Gaussian line are shown in Fig.~\ref{fig:tiz}. In these ADAF models, the observed spectral and luminosity changes of GX~339$-$4 are predominantly driven by changes in the transition radius; the implied accretion rate change is substantially smaller than the factor of 5 for the observed luminosity change. \begin{figure*} \centerline{ \psfig{figure=Fig6.ps,width=0.45\textwidth} } \caption{\small Advection Dominated Accretion Flow models for the unfolded RXTE data from Observation 1 (solid line) and Observation 5 (dash-dot line). Parameters consistent with the unfolded data are described in the text. A source distance of 4 kpc was assumed.} \label{fig:tiz} \end{figure*} For $r_{\rm tr} \sim 200$, which provides a rough description of Observation~1, the disk blackbody emission peaks in the far UV/soft X-rays and dominates over the synchrotron emission, which peaks in the optical/UV (see also \def\astroncite##1##2{##1\ ##2}\@internalcite{zdziarski:98a}). The peak synchrotron emission frequency scales as $\propto m^{-1/2}\dot{m}^{1/2}r_{\rm tr}^{5/4}$ and peaks in the range $\nu = 10^{11}$--$10^{12}$\,Hz for supermassive black holes and $\nu = 10^{15}$--$10^{16}$\,Hz for galactic black holes. The spectrum below the peak is approximately $S_\nu \propto \nu^{2}$. The synchrotron emission can contribute significantly to the radio emission of super-massive black holes (although see \def\astroncite##1##2{##1\ ##2}\@internalcite{dimatteo:98a}); however, the predicted radio flux of GX~339$-$4 is ten orders of magnitude below the observed 7 mJy flux at 843 MHz. Thus, there must be an extended source of radio emission, which we further discuss in the next section. \section{Simultaneous Radio Observations}\label{sec:radio} The first three of our RXTE observations were simultaneous with 843\,MHz observations taken with the Molongolo Observatory Synthesis Telescope (MOST), and with 8.3--9.1\,GHz observations taken at the Australian Telescope Compact Array (ATCA). Extensive discussion of the MOST and ATCA observations can be found in \def\astroncite##1##2{##1\ (##2)}\@internalcite{hanni:98a} and \def\astroncite##1##2{##1\ (##2)}\@internalcite{corbel:98a}, respectively (see also Table~\ref{tab:log}). An estimate of the minimum size of the radio emitting region can be obtained by noting that observationally the brightness temperatures of radio sources usually are not larger than $10^{12}$\,K, else the electrons will suffer catastrophic inverse Compton losses. The brightness temperature of a uniformly bright spherical source is given by $(c D/d \nu)^2 S_\nu/2\pi k$, where $d$ is the diameter of the source, $D$ is its distance, $\nu$ is the observed radio frequency, $S_\nu$ is the observed radio flux density, $c$ is the speed of light, and $k$ is the Boltzmann constant. Taking the 7\,mJy observed at 843\,Mz by MOST and the fact that GX~339$-$4 is unresolved, we derive \begin{equation} d \aproxgt 4 \times 10^{12}\,{\rm cm} \left ( {{D}\over{4\,{\rm kpc}}} \right ) \approx 3 \times 10^6\,R_{\rm G} \left ( {{D}\over{4\,{\rm kpc}}} \right ) \left ( {{M}\over{10\,M_\odot}} \right )^{-1} , \label{eq:bright} \end{equation} which is orders of magnitude larger than the inferred size of the X-ray emitting region, even for models that posit extremely extended coronae (e.g., \def\astroncite##1##2{##1\ ##2}\@internalcite{esin:97c,kazanas:97a}). This size scale is $\propto \nu^{-1}$, so that emission at 8.6 GHz could arise in a region an order of magnitude smaller than that responsible for the emission at 843 MHz. Indeed, the flat spectrum emission is likely to arise in a conical jet with a radially decreasing optical depth (e.g., \def\astroncite##1##2{##1\ ##2}\@internalcite{hjellming:88a}). Thus the outflow likely has an extent of ${\cal O}(10^7~GM/c^2)$ or greater. Similar estimates for source size have been made for the other persistent black hole candidate and Z-source neutron star X-ray binaries by \def\astroncite##1##2{##1\ (##2)}\@internalcite{fender:99a}. Assuming a radio spectral index of $\alpha = 0.1$ and a sharp cutoff at $10\mu {\rm m}$ (a reasonable upper limit for where the radio flux becomes optically thin, and typical of where ADAF models become optically thin in the radio), the radio flux is approximately 0.1\% of the 3--100\,keV X-ray flux. The correlation between the X-ray and radio fluxes found by \def\astroncite##1##2{##1\ (##2)}\@internalcite{hanni:98a}, comparable to the X-ray/radio correlation observed in Cyg~X-1 (\def\astroncite##1##2{##1\ ##2}\@internalcite{pooley:98a}), suggests that there is a coupling between the inner accretion disk and the extended outflow on timescales of 7 days or less. Matter leaving the corona at the escape velocity (0.25$c$ at $30 R_{\rm G}$) and thereafter decelerating under the influence of gravity would take roughly 7~days to travel a distance of $10^7 R_{\rm G}$. As 7 days is the upper bound to the correlation timescale, the radio emitting outflow must leave at slightly greater than escape velocity, or there must be at least some amount of acceleration of the outflow. Although the radio observations are strictly simultaneous with our first three RXTE observations, GX~339$-$4 exhibits less than 1\% rms variability over the shortest timescales for which a reasonable radio flux estimate can be made ($\aproxgt 10$ minutes). Thus there are no strong features to correlate between the radio and X-ray bands. \section{Discussion}\label{sec:discuss} \smallskip \noindent \emph{Coronal Size and Luminosity Variation:} The relationship between the inferred size of the corona and the magnitude of the observed flux depends upon which spectral model we are considering. As discussed in \S\ref{sec:adaf}, for ADAF models one can associate lower fluxes with \emph{increased} coronal radii. A larger coronal radius implies a lower efficiency and hence a decreased observed flux, even for constant accretion rates. Paper~II shows that the characteristic power spectral density (PSD) timescale for GX~339$-$4 decreases for the lowest observed flux (Observation 5). If one associates the PSD timescale with characteristic disk timescales, this could be in agreement with an increased coronal radius. However, in paper~II we also show that the time lags between hard and soft X-ray variability \emph{decreases} with decreasing flux, which seems counter to a positive correlation between flux and coronal size. The `sphere+disk' coronal models make no assumptions about the radiative efficiency of the accretion. The flux can be either positively or negatively correlated with coronal radius, depending upon the variations of the coronal compactness, $\ell_{\rm c}$, and the temperature, $T_{\rm d}$, at the inner edge of the accretion disk that surrounds the corona (\def\astroncite##1##2{##1\ ##2}\@internalcite{dove:97b}). Note that the `sphere+disk' models used in this work, contrary to many ADAF models, do not consider synchrotron photons as a source of seed photons for Comptonization. Using the definitions of $\ell_{\rm c}$, $\ell_{\rm d}$, and $f$ given in \S\ref{sec:corona}, energy balance in the `sphere+disk' system determines the coronal radius, to within factors of order unity, to be given by \begin{eqnarray} R_{\rm C} &\approx& 160 \left ( \frac{\ell_{\rm d}+f \ell_{\rm c}}{\ell_{\rm d} + \ell_{\rm c}} \right )^{1/2} \left ( \frac{k T_{\rm d}}{150~{\rm eV}} \right )^{-2} \left ( \frac{6~M_\odot}{M} \right ) \cr &\times& \left ( \frac{D}{4~{\rm kpc}} \right ) \left ( \frac{F_{\rm tot}}{10^{-8}{\rm ergs~cm^{-2}~s^{-1}}} \right )^{1/2} R_{\rm G} ~~, \label{eq:coronarad} \end{eqnarray} where $M$ is the mass of the compact object, $D$ is the distance to the source, and $F_{\rm tot}$ is the bolometric flux of the source. If $T_{\rm d}$, $f$, $\ell_{\rm d}$, and $\ell_{\rm c}$ were held fixed, then the coronal radius would be positively correlated with flux. Whereas this might pose some problems for understanding the flux dependence of the characteristic timescales observed in the PSD, this would agree with the flux dependence of the X-ray variability time lags (paper~II). However, as the RXTE bandpass does not usefully extend below $\approx 3$\,keV, we do not have a good handle on the flux dependence of $T_{\rm d}$. If $T_{\rm d} \propto F_{\rm tot}^\beta$ with $\beta > 1/4$, then increasing flux could imply decreasing coronal radius. \smallskip \noindent \emph{Correlations Among Spectral Parameters:} \begin{figure*} \centerline{ \psfig{figure=Fig7a.eps,width=0.33\textwidth} \psfig{figure=Fig7b.eps,width=0.33\textwidth} \psfig{figure=Fig7c.eps,width=0.33\textwidth} } \caption{\small {\it Left:} Reflection fraction vs. photon index, $\Gamma$, for models fit to PCA data only (squares), and models fit to PCA plus HEXTE data that allowed the PCA and HEXTE photon indeces and normalizations to be different (diamonds; HEXTE photon index shown). {\it Middle:} Photon index, $\Gamma$, vs. observed 3--9\,keV flux for the same reflection models as on the left. Also shown, without error bars, is the best fit compactness, $\ell_c$, for `sphere+disk' coronal models fit to HEXTE data (small triangles). {\it Right:} Disk ionization parameter, $\xi$, in units of ${\rm ergs~cm~s^{-1}}$ for the same reflection models as on the left. \label{fig:reflect}} \end{figure*} \def\astroncite##1##2{##1\ (##2)}\@internalcite{ueda:94a} claimed that reflection models of GX~339$-$4 exhibited a correlation between photon index, $\Gamma$, and reflection fraction, $f$, with softer spectra implying greater reflection. \def\astroncite##1##2{##1\ (##2)}\@internalcite{zdziarski:98b} has claimed that this correlation extends to reflection models of Seyfert~1 galaxies as well. Such a correlation is not unreasonable to expect. For example, if we allow the corona and disk to overlap to some extent in the `sphere+disk' model (\def\astroncite##1##2{##1\ ##2}\@internalcite{poutanen:97b}), then we expect the increase in the flux of seed photons to cool the corona and lead to a softer spectral index. Likewise, the covering fraction of the disk would be increased, in agreement with the suggested correlation. In Figure~\ref{fig:reflect} we plot $f$ vs. $\Gamma$ for our reflection model fits to GX~339$-$4. Contrary to the claims of \def\astroncite##1##2{##1\ (##2)}\@internalcite{ueda:94a} and \def\astroncite##1##2{##1\ (##2)}\@internalcite{zdziarski:98b}, however, there is no strong evidence for a correlation. Fitting the reflection fraction with a function linear in $\Gamma$, as opposed to fitting with the mean value of $f$, improves the $\chi^2$ of the fits by $0.2$, which is not significant. Fitting with the mean gives $\chi^2_{\rm red} = 0.2$. We do note two possible trends from the reflection model fits. First, as has been noted for other hard state galactic black hole candidates (\def\astroncite##1##2{##1\ ##2}\@internalcite{tanaka:95a} and references therein), there may be a correlation between flux and photon index with lower flux implying a harder source. (The significance of the correlation is driven by Observation 5, the faintest and hardest of the observations. However, a similar correlation is also present in color-intensity diagrams.) Such a correlation is consistent with the expectations of ADAF models where the radius increases with decreasing flux (Figure~\ref{fig:tiz}). Again, the `sphere+disk' corona models do not predict a clear trend without knowing the flux dependences of other parameters such as $T_{\rm d}$. Second, the ionization parameter, $\xi$, is positively correlated with flux. Such a correlation was noted by \def\astroncite##1##2{##1\ (##2)}\@internalcite{zycki:98a} for Ginga observations of Nova Muscae. It is not unreasonable to expect the disk to become increasingly ionized with increasing flux. We again caution that it is dangerous to make one-to-one correlations between a model fit parameter and a true physical parameter. Furthermore, the significance of the correlation is again almost entirely determined by Observation 5, the faintest observation, which has $\xi \approx 0$. However, if we take the flux dependence of $\xi$ as being real and interpret it physically, it provides some constraints on the flux dependence of the coronal radius. The ionization parameter is $\propto F_{\rm tot} / (\rho R^2)$, where $\rho$ is the density of the disk. For a gas pressure-dominated Shakura-Sunyaev $\alpha$-disk, $\rho \propto R^{-1.65}$ (\def\astroncite##1##2{##1\ ##2}\@internalcite{shakura:73a}). In order for $\xi$ to be roughly linear in flux (the actual dependence is not strongly constrained by the data), we require $F_{\rm tot} \proptwid R^{1.35}$. (For the `sphere+disk' models this would further require $T_{\rm d} \proptwid R^{-1/3}$, depending upon the flux dependences of $\ell_{\rm d}$, $\ell_{\rm c}$, $g$, etc.) Taken physically and in the context of a gas pressure dominated Shakura-Sunyaev $\alpha$-model, the flux dependence of $\xi$ implies that the coronal radius increases with increasing flux. \section{Summary}\label{sec:summ} We have presented a series of observations of the black hole candidate GX~339$-$4 in low luminosity, spectrally hard states. These observations consisted of three separate archival ASCA and eight separate RXTE data sets. All of these observations exhibited (3--9\,keV) flux $\aproxlt 10^{-9}~{\rm ergs~s^{-1}~cm^{-2}}$, and the observed fluxes spanned roughly a factor of 5 in range for both the ASCA and RXTE data sets. Subject to uncertainties in the cross calibration between ASCA and RXTE, the faintest ASCA observation was approximately a factor of two fainter than the faintest RXTE observation. All of these observations showed evidence for an $\approx 6.4$\,keV Fe line with equivalent widths in the range of $\approx 20$--$140$\,eV. The ASCA observations further showed evidence for a soft excess that was well-modeled by a power law plus a multicolor blackbody spectrum with peak temperatures in the range $\approx 150-200$\,eV. Both of these factors considered together argue against `sphere+disk' or ADAF type-geometry coronae with extremely large coronal radii of ${\cal O}(10^4~R_{\rm G})$ (e.g., \def\astroncite##1##2{##1\ ##2}\@internalcite{esin:97c}). The RXTE data sets were well-fit by `sphere+disk' Comptonization models with coronal temperatures in the range $20$--$50$\,keV and optical depths in the range of $\tau \approx 3$. These fits were similar to our previous fits to RXTE data of Cyg~X-1. Advection Dominated Accretion Flow models, which posit a similar geometry, also provided reasonable descriptions of the unfolded RXTE data. The `sphere+disk' and ADAF models were not able, however, to also model the observed radio fluxes. Thus, a static corona seems to be ruled out by the observations. The ADAF models can imply that the coronal radius increases with decreasing flux. The `sphere+disk' corona models do not make a specific prediction for the dependence of the coronal radius on the flux; however, they can be consistent with a positive correlation between coronal radius and flux. As described in paper~II, a positive correlation between flux and coronal radius is consistent with the observed flux dependence of the time lags between hard and soft X-ray variability. We also considered `reflection models' of the RXTE data. These models showed evidence of a hardening of the RXTE spectra with decreasing X-ray flux. They further showed evidence of a positive correlation between the best-fit ionization parameter, $\xi$, and the observed flux. Especially the latter of these correlations, however, was dominated by the model fits of the faintest observation. The reflection models did not exhibit any evidence of a correlation between the photon index of the incident power law flux and the solid angle subtended by the reflector. Three of the RXTE observations were strictly simultaneous with 843\,MHz and 8.3--9.1\,GHz radio observations. The most likely source of the radio flux is synchrotron emission from an extended outflow with a size of ${\cal O}(10^7~GM/c^2)$. The correlation between radio and X-ray emission on timescales of 7 days or less (\def\astroncite##1##2{##1\ ##2}\@internalcite{hanni:98a}) implies a strong coupling of the inner disk accretion flow with this spatially extended outflow as is expected by recent theoretical arguments (\def\astroncite##1##2{##1\ ##2}\@internalcite{blandford:98a}). Further simultaneous radio/X-ray observations, preferably with the addition of IR/optical monitoring to constrain the location of the synchrotron break and with the addition of soft X-ray monitoring to constrain the accretion disk parameters, are required to test such models in detail. \acknowledgements We would like to thank Dr. Christopher Reynolds for keeping a stiff upper lip while explaining ASCA data analysis to us. We would also like to thank K.~Mukai of the ASCA GOF for useful advice. W.A.~Heindl and D.~Gruber kindly provided assistance with the HEXTE data extraction, and S.~Corbel provided assistance with the radio data. We are grateful to B.~Stern for writing the original version of kotelp, and, more importantly, for finally telling us what the name means (`cauldron'). We would also like to acknowledge useful conversations with M.~Begelman, J. Chiang, B.A.~Harmon, K.~Pottschmidt, R.~Staubert, C.~Thompson, and A.~Zdziarski. This work has been financed by NASA Grants {NAG5-4731} and {NAG5-3225} (MAN, JBD). MN was supported in part by the National Science Foundation under Grant No. Phy94-07194. JW was supported by a travel grant from the Deutscher Akademischer Austauschdienst, RPF was supported by an EC Marie Curie Fellowship (ERBFMBICT 972436), and TDM thanks Trinity College and PPARC for financial support. This research has made use of data obtained through the High Energy Astrophysics Science Archive Research Center Online Service, provided by the NASA/Goddard Space Flight Center.
\section{Introduction} In the decade that has passed since the pioneering analytical study of G\'alfi and R\'acz~\cite{galfi}, there has emerged a substantial body of research devoted to experimental~\cite{expt,leger}, computational~\cite{jiang,cornell,cornell97,araujo,havlin,kozatait,tait2} and analytical~\cite{schenkel,vanbaalen,koza1,koza2,peletier96,peletier99} studies of reaction-diffusion systems with two initially separated, diffusing species A and B reacting to produce an inert product C according to the chemical formula \begin{equation} m^\prime\mbox{A} + n^\prime\mbox{B} \rightarrow \mbox{C (inert)} , \end{equation} where $m^\prime$ and $n^\prime$, the stoichiometric coefficients, are positive integers. Theoretical studies have focused almost exclusively on the \lq\lq one-dimensional'' case of an infinite, flat reaction front between two regions of homogeneous composition of either A or B (see Fig.~\ref{fig:cartoon}). This idealized situation is believed to capture much of the essential physics of reaction fronts commonly observed in various chemical~\cite{grindrod,alberty} and biological~\cite{murray77,murray93} systems. The standard continuum model for such a one-dimensional reaction front involves a pair of nonlinear partial differential equations~\cite{galfi,cornell,schenkel,vanbaalen,koza1} \begin{subeqnarray} \frac{\partial \rho_A}{\partial T} & = & D_A \frac{\partial^2 \rho_A}{\partial X^2} - m^\prime R(\rho_A,\rho_B) \slabel{eq:coup1} \\ \frac{\partial \rho_B}{\partial T} & = & D_B \frac{\partial^2 \rho_B}{\partial X^2} - n^\prime R(\rho_A,\rho_B) \slabel{eq:coup2} , \label{eq:coup} \end{subeqnarray} subject to the boundary conditions \begin{equation} \rho_A(-\infty,T) = 0, \ \ \rho_A(\infty,T) = \rho_A^o, \ \ \rho_B(-\infty,T) = \rho_B^o, \ \ \rho_B(\infty,T) = 0 \label{eq:rhoBC} \end{equation} and the initial conditions \begin{equation} \rho_A(X,0) = \rho_A^o H(X) , \ \ \rho_B(X,0) = \rho_B^o H(-X) \label{eq:rhoIC} \end{equation} where $\rho_A(X,T)$ and $\rho_B(X,T)$ are the concentrations, $D_A$ and $D_B$ the diffusion coefficients of A and B, respectively, $\rho_A^o > 0$ and $\rho_B^o > 0$ are constants, $H(X)$ is the Heaviside unit step function and $R(\rho_A,\rho_B)$ is the reaction rate density for production of species C. (Note that upper-case letters denote quantities with dimensions, e.g. $X$ and $T$ for space and time, respectively. Lower-case letters for the corresponding dimensionless quantities are introduced in section~\ref{sec:dim}.) The reactants are completely separated at first according to (\ref{eq:rhoIC}), but for $T > 0$ they diffuse together and react, which decreases the concentrations wherever $\rho_A(X,T)\rho_B(X,T) > 0$. Diffusion acts to replenish any depleted regions. As a result the system develops a localized, moving region, the \lq\lq reaction front,'' where the reaction rate $R(\rho_A,\rho_B)$ is greatest and which is fed by diffusion from the distant particle reservoirs described by the boundary conditions. The dynamics of this reaction front are described by the long-time asymptotics of the nonlinear initial-boundary-value problem (\ref{eq:coup})--(\ref{eq:rhoIC}). The nonlinear reaction term $R(\rho_A,\rho_B)$ is usually assumed to have the form of a power law \begin{equation} R(\rho_A,\rho_B) = k \rho_A^m \rho_B^n , \label{eq:Rpower} \end{equation} where $k$ is a rate constant, and $m$ and $n$ are respectively the \lq\lq kinetic orders'' of A and B in the reaction~\cite{alberty}. For a one-step reaction with sufficient mixing (see below) $m=m^\prime$ and $n=n^\prime$, but for more complex, multi-step reactions $m$ and $n$ are determined by the stoichiometric coefficients of the (often unknown) rate-limiting step. Although $m$ and $n$ are usually taken to be positive integers, non-integer values of $m$ and $n$ can arise in certain situations~\cite{alberty}. We will see that a well-defined reaction front exists for any real numbers $m, n \geq 1$, but not for $m<1$ or $n<1$. Technically, by assuming in (\ref{eq:coup}) that the reaction rate $R$ depends only on the average local concentrations (and not on any fluctuations or many-body effects) we have made the \lq\lq mean-field approximation''~\cite{cardy}. In low-dimensional systems, such as ion channels ($d=1$) or catalytic surfaces ($d=2$), the mean-field approximation can break down because the reacting particles cannot mix efficiently enough, but as the dimension of the system is increased above a certain \lq\lq upper critical dimension'' $d_c$, such statistical anomalies disappear. For two diffusing reactants with a simple one-step reaction it is known~\cite{cornell,havlin,cardy} that $d_c = 2/(m+n-1)$. Since $d_c \leq 2$ for $m,n \geq 1$ the mean-field approximation should be perfectly valid in the usual case $d=3$, which is consistent with experimental findings~\cite{expt}. In contrast to the case of two diffusing reactants described above, relatively little is known~\cite{jiang,havlin,koza2,peletier96,peletier99} about the case of one diffusing reactant ($D_A > 0$) and one static reactant ($D_B = 0$). This situation, depicted schematically in Fig.~\ref{fig:cartoon}, describes the corrosion of a porous solid B saturated with a fluid solvent and exposed to an initially separated colloidal reactant A, as shown in recent electrochemical experiments (described below) ~\cite{leger}. Jiang and Ebner~\cite{jiang} first pointed out (for $m=n=1$) that setting $D_B=0$ in (\ref{eq:coup}) is a non-trivial, i.e. singular, limit leading to different long-time behavior than in the case of $D_B > 0$ (no matter how small), which they explained with simple scaling arguments supported by Monte Carlo computer simulations. For an analytical description of such one-dimensional diffusion with one static reactant, we adopt the power-law form of the reaction term and study the coupled equations \begin{subeqnarray} \frac{\partial \rho_A}{\partial T} & = & D_A \frac{\partial^2 \rho_A}{\partial X^2} - m^\prime k \rho_{A}^m\rho_{B}^n \slabel{eq:rhoA} \\ \frac{\partial \rho_{B}}{\partial T} & = & - n^\prime k \rho_{A}^m\rho_{B}^n . \slabel{eq:rhoB} \label{eq:rho} \end{subeqnarray} In the simplest case $m=n=1$, the initial-boundary-value problem (\ref{eq:rhoBC})--(\ref{eq:rho}) has been solved numerically by Havlin et al.~\cite{havlin} and analyzed in the limit of ``long times'' $T \rightarrow \infty$ by Koza~\cite{koza2}, using various asymptotic approximations introduced by G\'alfi and R\'acz~\cite{galfi}. Rigorous analysis has been reported in the analogous limit of ``fast reactions'' $k\rightarrow\infty$ by Hilhorst et al.~\cite{peletier96}, but these authors only address the behavior at the diffusive length scale $X \propto \sqrt{T}$ (see section~\ref{sec:diff} below) and do not consider the structure of the reaction front studied by Koza~\cite{koza2}, which is of primary interest here. Hilhorst et al. have also recently considered the effect of a more general reaction term at the diffusive scale~\cite{peletier99}, but the present work appears to be the first to analyze the nontrivial effect of changing reaction orders at the reactive length scale (see below) in the general case ($m,n\geq 1$) with one static reactant. The relevance of (\ref{eq:rho}) for a given porous-solid corrosion system rests on several key assumptions that are less obviously satisfied {\it a priori} than in the case of two diffusing reactants. First, the solid matrix containing the static reactant B must be sufficiently porous that the moving reactant A can diffuse freely to the exposed surfaces with an effective diffusion constant (averaged over many pores) comparable to that in the bulk solvent. The concentration of A must also be dilute enough that $D_A$ is constant. Another reason that the concentrations of A and B must be dilute is that the inert product C must be created in small enough quantities that its presence does not affect the reaction dynamics (e.g. by inhibiting diffusion or initiating convection). Finally, one might worry about the breakdown of the mean-field approximation since the (possibly fractal) pore structure may influence statistical averaging. For example, it is known that fluctuations alter the reaction-front dynamics when the diffusion is confined to a percolating cluster in two dimensions~\cite{havlin}. In spite of these concerns, however, the one-dimensional mean-field model (\ref{eq:rho}) can in fact describe certain corrosion systems. An important motivation for the present analytical study is afforded by the recent experiments of L\'eger {\it et al.}~\cite{leger}, which are the first to examine in detail the case of one static and one diffusing reactant. These experiments involve the corrosion of ramified copper electrodeposits exposed to a cupric chloride electrolyte to produce cuprous chloride crystallites via the reaction \begin{equation} \mbox{CuCl$_2$ (aq) \ + \ Cu (solid) \ $\rightarrow$ \ 2 CuCl (solid)} \label{eq:reaction} \end{equation} immediately following electrodeposition. It is found that the long-time behavior of (\ref{eq:rho}) with $m=1$ matches the experimentally observed front speed and concentration profile of diffusing reactant (CuCl$_2$) rather well, in spite of the complex fractal geometry of the electrodeposits and the presence of the inert product (CuCl)~\cite{leger}. Since the reaction rate and the concentration of the static species (Cu) are not directly measured, however, the interpretation of these kinds of corrosion experiments can be aided by the analysis presented here of the mean-field model with $m, n \geq 1$. There is an extensive mathematical literature~\cite{grindrod,murray77,murray93,gmira,barenblatt96} on the subject of single reaction-diffusion equations of the general form \begin{equation} \frac{\partial \rho}{\partial T} = D \frac{\partial^2 \rho}{\partial X^2} - f(\rho) \label{eq:single} \end{equation} which arise in many applications (e.g. chemical reactions, combustion and population dynamics). A common theme in these studies is the appearance of two distinct (time-dependent) length scales in the intermediate asymptotic regime ($t \rightarrow \infty$) which correspond to either ``weakly nonlinear behavior'', where it has been established in many cases that the reaction term is negligible and the dynamics are purely diffusive, or ``strongly nonlinear behavior'', where the reaction and diffusion terms balance (in the nomenclature of Gmira and Veron~\cite{gmira}). This separation of scales also arises in coupled systems of reaction-diffusion equations like (\ref{eq:coup}), but owing to their greater complexity, much less rigorous analysis has been reported. In the case of two diffusing reactants, G\'alfi and R\'acz~\cite{galfi} pointed out that if the diffusion constants are the same, $D_A = D_B$, then the difference in concentrations $\rho_A - \rho_B$ obeys a pure diffusion equation which can be easily integrated, thereby reducing the coupled system (\ref{eq:coup}) to a single equation with the form of (\ref{eq:single}). Another simplification occurs if also $\rho_A^o = \rho_B^o$ in which case the reaction front is perfectly symmetric and does not move. In this simplified case with $m=n=1$, Schenkel et al.~\cite{schenkel} were able to prove that the asymptotic solution of G\'alfi and R\'acz~\cite{galfi}, which combines different approximations at the diffusive and reactive scales, is approached uniformly as $T \rightarrow \infty$ starting from the initial conditions of (\ref{eq:rhoIC}), and they also reported rigorous bounds on the transient decay to the asymptotic solution. Recently, van Baalen et al.~\cite{vanbaalen} have extended this analysis to the case of symmetric, high-order reactions $m=n>3$, where the reaction-front scaling is altered. The analyses of Refs.~\cite{schenkel}--\cite{vanbaalen} represent an important contribution because, at least in the case $D_A=D_B$, $\rho_A^o=\rho_B^o$ and $m=n$, they provide a rigorous mathematical justification for various {\it ad hoc} assumptions introduced by G\'alfi and R\'acz~\cite{galfi,koza1} to describe the local structure of the reaction front which have otherwise been validated only by numerical simulations. Unfortunately, however, since the analysis in Refs.~\cite{schenkel,vanbaalen} relies on a comparison principle for single parabolic equations~\cite{aronson,gmira} of the form (\ref{eq:single}), it does not (as the authors indicate) appear to be applicable when $D_A \neq D_B$ (which also leads to a moving reaction front). Van Baalen et al.~\cite{vanbaalen} also remark that their analysis is not easily extended to certain intermediate reaction orders ($1 < m=n \leq 3$). These difficulties are reflected in Koza's recent studies of the general cases $D_A > D_B > 0$ ~\cite{koza1} and $D_B = 0$ ~\cite{koza2}, in which several {\it ad hoc} (but reasonable) approximations are made and transients are ignored. In the present article, the long-time asymptotics of the initial-boundary-value problem (\ref{eq:rhoBC})--(\ref{eq:rho}) are studied. This special case of (\ref{eq:coup}) is more tractable analytically than the general case because (\ref{eq:rhoB}) can be integrated exactly in time, thereby reducing the coupled system to a single integro-partial differential equation. This useful simplification is presented in section~\ref{sec:prelim} where the problem is recast in a dimensionless form. It is also noted that similarity solutions are expected to exist because there is no natural length or time scale in the problem~\cite{barenblatt96,barenblatt87,dresner}. Although it may be possible to prove that the system actually approaches such a self-similar solution starting from the prescribed initial conditions, we instead pursue the more modest goal of proving that if an asymptotic similarity solution exists, it must have a certain unique form, {\it i.e.} we explore the consequences of the ``quasi-stationary approximation''~\cite{cornell,koza1,koza2}. In section~\ref{sec:similarity}, the similarity solution is systematically derived, and it is shown that a \lq\lq diffusion layer'' (where the reaction term is dominated by the diffusion term) with different scaling properties than the \lq\lq reaction front'' (where the reaction and diffusion terms balance) must exist to satisfy the boundary conditions. In section~\ref{sec:transients}, the transient decay of the reaction rate in the diffusion layer is analyzed, thereby proving {\it a posteriori} that the reaction term can indeed be neglected in the dominant balance. In section~\ref{sec:uniform}, a uniformly-valid, asymptotic approximation is constructed by matching the self-similar forms in the two different regions. Finally, in section~\ref{sec:disc} some general physical conclusions are drawn from the analysis, and in the Epilogue certain similarities are discussed between this work and the literature on combustion waves. (Note that section \ref{sec:transients} is more technical and may be skipped in a first reading.) \section{Preliminaries} \label{sec:prelim} \subsection{Dimensionless Formulation} \label{sec:dim} With the definitions, \begin{subeqnarray} t \equiv m^\prime k (\rho_A^o)^{m-1}(\rho_B^o)^n T, \ & \ & \ x \equiv X \sqrt{m^\prime k (\rho_A^o)^{m-1}(\rho_B^o)^n/D_A} \slabel{eq:units} \\ a(x,t) \equiv \rho_A(X,T)/\rho_A^o, \ & \ & \ b(x,t) \equiv \rho_B(X,T)/\rho_B^o , \end{subeqnarray} the initial-boundary-value problem (\ref{eq:rhoBC})--(\ref{eq:rho}) may be expressed in a dimensionless form \begin{subeqnarray} \frac{\partial a}{\partial t} & = & \frac{\partial^2 a}{\partial x^2} - a^mb^n \slabel{eq:a} \\ \frac{\partial b}{\partial t} & = & - q a^mb^n . \slabel{eq:b} \\ a(\infty,t) = 1, \ b(\infty,t) = 0, & \ & a(-\infty,t) = 0, \ b(-\infty,t) = 1 , \slabel{eq:bc} \\ a(x,0) = H(x), & \ & b(x,0) = H(-x) \slabel{eq:ic} \label{eq:eqs} \end{subeqnarray} which involves only one dimensionless parameter: \begin{equation} q \equiv \frac{n^\prime\rho_A^o}{m^\prime\rho_B^o} . \end{equation} Note that the dimensionless problem (\ref{eq:eqs}) depends only upon the initial concentrations $\rho_A^o$ and $rho_B^o$ and the stochiometric coefficients $m^\prime$ and $n^\prime$ through the parameter $q$; the reaction rate $k$ are the diffusion constant $D_A$ simply set the natural scales for length and time. From (\ref{eq:units}), we see that the limit of ``fast reactions'' $k \rightarrow \infty$ (with $X$ and $T$ fixed) corrseponds to the limit of long (dimensionless) times $t\rightarrow\infty$ at the diffusive scale $x \propto \sqrt{t}$. (See Ref.~\cite{peletier96} for another discussion of this correspondence of limits.) \subsection{The Governing Integro-Partial Differential Equation} The statement of the problem (\ref{eq:eqs}) will be used in deriving the asymptotic similarity solution below, but for the transient analysis described in section ~\ref{sec:transients} it will be convenient to first integrate (\ref{eq:b}) exactly in time. Note that (\ref{eq:b}) and (\ref{eq:ic}) imply that $b(x,t) = 0$ for $x>0$ at all times $t \geq 0$, which reflects the fact that species B cannot diffuse out of its initial region. For $x < 0$, we integrate (\ref{eq:b}) using the initial condition (\ref{eq:ic}) to express $b(x,t)$ as \begin{equation} b(x,t) = \left\{ \begin{array} {ll} e^{-q \phi_m(x,t)} & \ \mbox{if} \ n=1 \\ \left[ 1 + q(n-1)\phi_m(x,t) \right]^{-1/(n-1)} & \ \mbox{if} \ n \neq 1 \end{array} \right. , \ \ \ x < 0 \label{eq:bint} \end{equation} which involves the time-integral of $a(x,t)^m$: \begin{equation} \phi_m(x,t) \equiv \int_0^t a(x,\tau)^m d\tau . \label{eq:phidef} \end{equation} (Note that a partial differential equation satisfied by $\phi_1(x,t)$ is given in Ref.~\cite{peletier96}.) Substituting for $b(x,t)$ in (\ref{eq:a}), we obtain a single, nonlinear integro-partial differential equation for $a(x,t)$, either \begin{equation} \frac{\partial a(x,t)}{\partial t} = \frac{\partial^2 a(x,t)}{\partial x^2} - H(x) a(x,t)^m e^{-q\int_0^t a(x,\tau)^m d\tau} \label{eq:ide1} \end{equation} if $n=1$ or \begin{equation} \frac{\partial a(x,t)}{\partial t} = \frac{\partial^2 a(x,t)}{\partial x^2} - \frac{H(x) a(x,t)^m}{\left[ 1 + q(n-1)\int_0^t a(x,\tau)^m d\tau \right]^{n/(n-1)}} \label{eq:ide2} \end{equation} if $n \neq 1$. Although these equations involve only one unknown function $a(x,t)$, they are somewhat unwieldy, so we will first seek long-time ($t \rightarrow \infty$) asymptotic solutions to the coupled system (\ref{eq:eqs}) in section~\ref{sec:similarity}. The time-dependent properties of (\ref{eq:ide1}) and (\ref{eq:ide2}) will be studied in section~\ref{sec:transients}. Before proceeding, however, we digress to show that physically meaningful solutions exist only if $n \geq 1$. Later in the analysis we will see that $m \geq 1$ is required as well. \subsection{A Reaction Front Does Not Exist if $n < 1$} \label{sec:exist_n} Consider any point $x_o < 0$. Since species A diffuses to $x_o$ from a reservoir of constant concentration ($a(\infty,t) = 1$) while species B is removed by reactions without ever being replenished ($\partial b(x_o,t)/\partial t < 0$ for all $t > 0$), it is clear that after long times $a(x_o,t)$ must eventually differ from zero. Therefore, there exists some $a_\ast(x_o) > 0$ and $t_o > 0$ such that $a(x_o,t) > a_\ast(x_o)$ for all $t > t_o$. This implies $\phi_m(x_o,t) > a_\ast(x_o)^m \cdot(t-t_o)$ from (\ref{eq:phidef}) and thus $b(x_o,t) \rightarrow 0$ from (\ref{eq:bint}) since $q>0$, but a singularity arises if $n < 1$: The concentration of static reactant $b(x_0,t)$ vanishes at some finite time $t_1$ given by $q(1-n)\phi_m(x_o,t_1)=1$ (which exists because $\phi(x_o,t)$ is continuous, $\phi_m(x_o,0)=0$ and $\phi_m(x_o,\infty)=\infty$). For $t > t_1$, Eq.~(\ref{eq:bint}) predicts imaginary, negative, or diverging solutions for $n < 1$, none of which are physically meaningful. Therefore, when $n<1$, the solutions to the model equations break down physically in a finite time, and in that sense there does not exist a stable, moving reaction front. Nevertheless, it should be noted that Hilhorst et al.~\cite{peletier96,peletier99} have shown that well-defined solutions with free boundaries (at the diffusive scale $x \propto \sqrt{t}$) can exist when $0 < n < 1$ . \section{Derivation of the Asymptotic Similarity Solution} \label{sec:similarity} \subsection{Scaling of the Reaction Front} \label{sec:Ac} The initial-boundary-value problem (\ref{eq:eqs}) possesses no natural length or time scale, {\it i.e.} it is invariant under power-law \lq\lq stretching transformations''~\cite{dresner}, and consequently in the limit $t \rightarrow \infty$ the system is expected to approach an asymptotic similarity solution in which distance and time are coupled by power-law scalings~\cite{barenblatt96,barenblatt87}. Since reactant A diffuses while reactant B does not, the (presumably unique) point of maximal reaction rate $r(a,b) = a^mb^n$ moves in the $-x$ direction toward the reservoir of reactant B. Therefore, an asymptotic similarity solution, if one exists, must involve a moving frame of reference centered on some point $x_f(t)$ identifying the position of the reaction front at or near the point of maximal reaction rate (with $dx_f/dt < 0$). Let $x_f(t) = -2 \nu t^\sigma$, where $\nu(q) > 0$ is a constant (akin to the \lq\lq speed'' of the front) to be determined self-consistently during the analysis, and consider an arbitrary coordinate stretching transformation in the moving reference frame, \begin{equation} \eta \equiv \frac{x + 2\nu t^\sigma}{t^\alpha} . \end{equation} where $w(t) = t^\alpha$ is the width of the reaction front indicated in Fig.~\ref{fig:cartoon}. (The factor of two is included only for algebraic convenience.) In the neighborhood of $x_f(t)$, we also allow the magnitude of $a(x,t)$ to vary with a power-law scaling, \begin{equation} \tilde{{\cal A}}(\eta,t) \equiv t^{\gamma} a(x,t) . \label{eq:Acdef} \end{equation} If $\gamma \neq 0$, then another similarity solution far away from the reaction front (in the \lq\lq diffusion layer'' shown in Fig.~\ref{fig:cartoon} and defined below) will be needed to satisfy the boundary condition $a(\infty,t)=1$. This possibility that {\it two regions with different asymptotically self-similar dynamics for $a(x,t)$} could arise is suggested by the fact that there are two driving terms, representing diffusion and reaction, on the right-hand side of (\ref{eq:a}) with different behaviors under stretching transformations. On the other hand, there is only the reaction term on the right-hand side of (\ref{eq:b}), so {\it $b(x,t)$ can exhibit only one type of asymptotic scale invariance}. This is the main mathematical consequence of the physical fact that reactant B does not diffuse. Since $b(-\infty,1)=1$, we consider the transformation \begin{equation} \tilde{{\cal B}}(\eta,t) \equiv b(x,t) . \end{equation} Note that the reaction term $r(a,b) = a^mb^n$ has the scaling, $r = t^{-\beta}\tilde{{\cal A}}^m\tilde{{\cal B}}^n$, where $\beta = m\gamma$ in the notation of G\'alfi and R\'acz~\cite{galfi}. These transformations leave the governing equations in the form: \begin{subeqnarray} t^{(m-1)\gamma} \frac{\partial \tilde{{\cal A}}}{\partial t} - t^{(m-1)\gamma-1} \left(\gamma\tilde{{\cal A}} + \alpha\eta\frac{\partial \tilde{{\cal A}}}{\partial \eta}\right) & & \nonumber \\ + t^{(m-1)\gamma-1-\alpha+\sigma} 2\sigma\nu \frac{\partial \tilde{{\cal A}}}{\partial \eta} & = & t^{(m-1)\gamma-2\alpha} \frac{\partial^2 \tilde{{\cal A}}}{\partial \eta^2} - \tilde{{\cal A}}^m \tilde{{\cal B}}^n , \slabel{eq:Aclong} \\ t^{m\gamma} \frac{\partial \tilde{{\cal B}}}{\partial t} - t^{m\gamma-1} \alpha \eta \frac{\partial \tilde{{\cal B}}}{\partial \eta} + t^{m\gamma-1-\alpha+\sigma} 2\sigma\nu \frac{\partial \tilde{{\cal B}}}{\partial \eta} & = & - q \tilde{{\cal A}}^m \tilde{{\cal B}}^n . \slabel{eq:Bclong} \label{eq:AcBclong} \end{subeqnarray} We now look for asymptotically invariant solutions \begin{subeqnarray} & & \tilde{{\cal A}}(\eta,t) \rightarrow {\cal A}(\eta), \ \ \ \frac{\partial \tilde{{\cal A}}}{\partial \eta}(\eta,t) \rightarrow {\cal A}^\prime(\eta), \ \ \ \frac{\partial^2 \tilde{{\cal A}}}{\partial \eta^2}(\eta,t) \rightarrow {\cal A}^{\prime\prime}(\eta) \slabel{Ac_converge} \\ & & \tilde{{\cal B}}(\eta,t) \rightarrow {\cal B}(\eta), \ \ \ \frac{\partial \tilde{{\cal B}}}{\partial \eta}(\eta,t) \rightarrow {\cal B}^\prime(\eta) \label{eq:AcBc_converge} \end{subeqnarray} as $t \rightarrow \infty$ with $|\eta| < \infty$ fixed. For consistency with our definition of the reaction front, we require that there is in each equation a dominant balance between the reaction term ${\cal A}^m {\cal B}^n$ and at least one other non-vanishing term. In order for time invariance to be attained, we assume that time-dependent terms in the transformed coordinates are negligible compared to the reaction term, {\it i.e.} \begin{equation} \lim_{t\rightarrow\infty}t^{(m-1)\gamma} \frac{\partial \tilde{{\cal A}}}{\partial t} = \lim_{t\rightarrow\infty}t^{m\gamma} \frac{\partial \tilde{{\cal B}}}{\partial t} = 0 , \label{eq:rfstat} \end{equation} which is a precise statement of the assumption of \lq\lq quasi-stationarity''~\cite{koza2}. A dominant balance with the reaction term in (\ref{eq:Aclong}) implies that at least one of the following three cases must be true: \begin{description} \item[Case A1:] $\ \ (m-1)\gamma-2\alpha = 0, \ \ (m-1)\gamma-1-\alpha+\sigma\leq 0, \ \ (m-1)\gamma-1\leq 0,$ \item[Case A2:] $\ \ (m-1)\gamma-2\alpha \leq 0, \ \ (m-1)\gamma-1-\alpha+\sigma= 0, \ \ (m-1)\gamma-1\leq 0,$ \item[Case A3:] $\ \ (m-1)\gamma-2\alpha\leq 0, \ \ (m-1)\gamma-1-\alpha+\sigma\leq 0, \ \ (m-1)\gamma-1= 0.$ \end{description} Likewise a dominant balance in (\ref{eq:Bclong}) requires that one of the following two cases must hold: \begin{description} \item[Case B1:] $\ \ m\gamma-1 = 0, \ \ m\gamma-1-\alpha+\sigma\leq 0,$ \item[Case B2:] $\ \ m\gamma-1 \leq 0, \ \ m\gamma-1-\alpha+\sigma = 0.$ \end{description} There are only two combinations of these cases that are logically consistent: \begin{description} \item[Traveling Wave Case:] (A2,\ B2) $\ \ \alpha \geq 0,\ \ \gamma = 0, \ \ \sigma = 1 + \alpha$, and \item[Diffusing Front Case:] (A1,\ B2) $\ \ \alpha = (m-1)\gamma/2, \ \ \sigma = 1 - (m+1)\gamma/2, \ \ 0 < \gamma \leq 1/m$. \end{description} In the first case, we have $\sigma \geq 1 $, which implies that the reaction front advances at least linearly, {\it e.g.} as a traveling wave $x_f \sim t$, but in the second case, the front advances sublinearly, {\it e.g.} as a diffusing front $x_f \sim t^{1/2}$. In both cases, note that the reaction order $n$ of the static species B plays no role in the scaling behavior. The same conclusion is also true of the reaction order $m$ of the diffusing species A in the Traveling Wave Case, but $m$ does affect the scaling exponents in the Diffusing Front Case. Consider the possibility $\gamma=0$, which is only consistent with the Traveling Wave Case. In this case, a single asymptotic scale invariance is attained everywhere, and the equations for ${\cal A}(\eta)$ and ${\cal B}(\eta)$ are \begin{subeqnarray} 2\sigma\nu {\cal A}^\prime & = & \delta_{\alpha,0} {\cal A}^{\prime\prime} - {\cal A}^m{\cal B}^n \\ 2\sigma\nu {\cal B}^\prime & = & -q {\cal A}^m{\cal B}^n \end{subeqnarray} where $\delta_{x,y}$ is the Kronecker delta. By combining these equations and integrating once using the boundary conditions behind the front, {\it i.e.} ${\cal A}(\infty) = 1$, ${\cal A}^\prime(\infty)=0$ and ${\cal B}(\infty)=0$, we obtain \begin{equation} 2\sigma\nu({\cal B} + q) = q (2\sigma\nu {\cal A} - \delta_{\alpha,0} {\cal A}^{\prime}) . \end{equation} Applying the boundary conditions ahead of the front, ${\cal A}(-\infty)={\cal A}^{\prime}(-\infty)=0$ and ${\cal B}(-\infty)=1$, to this equation then implies $\sigma\nu(1+q) = 0$, which is a contradiction since $\sigma > 0$ and $\nu > 0$ are needed for the reaction front to move at all (and $q > 0$). In this way, we are forced to consider at least {\it two regions with different scale invariance} if there is to be any hope of an asymptotic similarity solution. Since the second type of scale invariance is associated with the dominance of the diffusion term versus the reaction term in (\ref{eq:a}), it must occur only on the back ($+x$) side of the reaction front due to the reservoir of reactant A at infinity, $a(\infty,t) = 1$ (see Fig.~\ref{fig:cartoon}). To describe the scale invariance of the diffusion layer, we postulate another power law $W(t) = t^\delta$ for the asymptotic width of the diffusion layer. \subsection{Scaling of the Diffusion Layer} Since $\delta \neq \alpha$, there are two possibilities, each involving a singular perturbation $w/W = t^{\alpha-\delta}$: \begin{description} \item[Infinitely Thin Reaction Front Case:] \ $\delta > \alpha$, \ $w = o(W)$, \item[Infinitely Thin Diffusion Layer Case:] \ $\delta < \alpha$, \ $W = o(w)$. \end{description} Since chemical reactions are typically much faster than diffusion, the former case seems more reasonable on physical grounds, but we do not rule out the latter case {\it a priori}. In the Infinitely Thin Reaction Front Case, the reaction front is defined by $x-x_f = O(w)$ and the diffusion layer by $W = O(x-x_f)$, $x>x_f$, whereas in the Infinitely Thin Diffusion Layer Case, the reaction front is defined by $w = O(x-x_f)$, $x<x_f$ and the diffusion layer by $x-x_f = O(W)$. In both cases, we view the reaction front as representing the \lq\lq inner problem'' (with similarity variable $|\eta|<\infty$) and the diffusion layer as representing the \lq\lq outer problem'' (with similarity variable $\zeta > 0$ defined below). The two regions are connected by asymptotic matching of the limits $\eta\rightarrow \infty$ and $\zeta \rightarrow 0^+$ (described in the next section) ~\cite{hinch,bender}. In order to treat the outer problem, we transform the original equations using a new reduced coordinate with power-law scalings, \begin{equation} \zeta \equiv \frac{x + 2\nu t^\sigma}{2 t^\delta} , \ \ \ \ \tilde{A}(\zeta,t) \equiv a(x,t), \ \ \ \ \tilde{B}(\zeta,t) \equiv b(x,t) . \label{eq:zeta} \end{equation} (Another factor of two is included in $\eta= 2t^{\delta-\alpha}\zeta$ for algebraic convenience. Note the use of $A$ and $B$ for the diffusion layer versus ${\cal A}$ and ${\cal B}$ for the reaction front.) Under this transformation, the equations take the form, \begin{subeqnarray} t^{2\delta}\frac{\partial \tilde{A}}{\partial t} - t^{2\delta-1} \delta\zeta \frac{\partial \tilde{A}}{\partial \zeta} + t^{\delta+\sigma-1} \sigma\nu \frac{\partial \tilde{A}}{\partial \zeta} & = & \frac{1}{4}\frac{\partial^2 \tilde{A}}{\partial \zeta^2} - t^{2\delta} \tilde{A}^m \tilde{B}^n , \slabel{eq:Along} \\ t^{2\delta}\frac{\partial \tilde{B}}{\partial t} - t^{2\delta-1} \delta\zeta \frac{\partial \tilde{B}}{\partial \zeta} + t^{\delta+\sigma-1} \sigma\nu \frac{\partial \tilde{B}}{\partial \zeta} & = & - t^{2\delta} q \tilde{A}^m \tilde{B}^n . \slabel{eq:Blong} \label{eq:ABlong} \end{subeqnarray} Seeking an asymptotic similarity solution, we assume that invariance is achieved in the transformed equations \begin{subeqnarray} & & \tilde{A}(\zeta,t) \rightarrow A(\zeta), \ \ \ \frac{\partial \tilde{A}}{\partial \zeta}(\zeta,t) \rightarrow A^\prime(\zeta), \ \ \ \frac{\partial^2 \tilde{A}}{\partial \zeta^2}(\zeta,t) \rightarrow A^{\prime\prime}(\zeta) \slabel{eq:A_converge} \\ & & \tilde{B}(\zeta,t) \rightarrow B(\zeta), \ \ \ \frac{\partial \tilde{B}}{\partial \zeta}(\zeta,t) \rightarrow B^\prime(\zeta) \slabel{eq:B_converge} \label{eq:AB_converge} \end{subeqnarray} assuming time-variations in (\ref{eq:A_converge}) are small relative to the diffusion term \begin{equation} \lim_{t \rightarrow\infty} t^{2\delta}\frac{\partial \tilde{A}}{\partial t}(\zeta,t) = 0,\ \ \mbox{for}\ \zeta>0. \label{eq:dlstat} \end{equation} In order to obtain a different scaling from the reaction front, the reaction term must also not enter into the dominant balance \begin{equation} \label{eq:Rzero} \lim_{t\rightarrow\infty} t^{2\delta} \tilde{A}(\zeta,t)^m \tilde{B}(\zeta,t)^n = 0 \ \ \mbox{for $\zeta > 0$,} \end{equation} a condition that we will check {\it a posteriori} for consistency in section~\ref{sec:transients}. Note that this limit vanishes trivially for $\zeta > \nu$ since we have already noted that $b(x,t)=0$ for all $x>0$. From (\ref{eq:Blong}), this condition would imply $\partial \tilde{B}/\partial \zeta = 0$, which together with the boundary condition $\tilde{B}(\infty) = 0$ would imply $\tilde{B}(\zeta) = 0$. With the reaction term gone, one of the terms on the left-hand side of (\ref{eq:Along}) must balance the $\partial^2 \tilde{A}/\partial \zeta^2$ term on the right side; if not, we would have $\partial^2 \tilde{A}/\partial \zeta^2=0$, which cannot satisfy all of the boundary conditions. There are only two possible dominant balances: \begin{description} \item[Case D1:] $\delta+\sigma-1 =0 \ \ \mbox{and} \ \ 2\delta-1 \leq 0$, \item[Case D2:] $\delta+\sigma-1 \leq 0 \ \ \mbox{and} \ \ 2\delta-1 = 0$. \end{description} The former case implies $\sigma > 1/2$ and hence contains the Traveling Wave Case (and not the Diffusing Front Case). With the scaling relations of case D1, Eq.~(\ref{eq:Along}) has the asymptotic form \begin{equation} \sigma\nu A^\prime = \frac{1}{4} A^{\prime\prime}. \end{equation} The solutions to this equation exhibit exponential growth as $\zeta \rightarrow \infty$, which is incompatible with the boundary condition $A(\infty) = 1$. Therefore, we conclude that Case D1, and hence also the Traveling Wave Case, is not consistent with the boundary conditions. At this point, we are left with Case D2 together with the Diffusing Front Case (A1,B2), which imply $W \sim \sqrt{t}$, thus justifying the term \lq\lq diffusion layer" for the region $\zeta > 0$. \subsection{Asymptotic Matching of the Reaction Front and Diffusion Layer} One more condition is needed to uniquely determine the scaling exponents, and it comes from asymptotic matching: The \lq\lq outer limit'' $\eta \rightarrow \infty$ of the inner approximation must match with the \lq\lq inner limit'' $\zeta \rightarrow 0^+$ of the outer approximation (because both are asymptotic representations of the same function). Unlike the more familiar case of boundary layers of ordinary differential equations~\cite{hinch,bender}, however, our system of partial differential equations will require extra care for matching because the limit $t \rightarrow \infty$ (with either $\zeta$ or $\eta$ fixed) must be taken before the inner and outer limits. Since $B(\zeta) = 0$ for all $\zeta > 0$, the only matching condition for $b(x,t)$ is trivial, ${\cal B}(\infty) = 0$, but the matching conditions for $a(x,t)$ are more subtle. Since $\gamma > 0$, the concentration of species $A$ approaches $0$ in the reaction front: $a(x,t) = O(t^{-\gamma})$ as $t \rightarrow \infty$ with $|\eta|<\infty$ fixed. Therefore, a boundary condition on the outer problem is $A(\zeta) = 0$, but unfortunately this does not provide a boundary condition on the inner problem. Instead, we must consider matching at the next (linear) order of Taylor expansion in the intermediate region: \begin{equation} \frac{\partial a}{\partial x} = \left\{ \begin{array}{ll} \frac{\partial \tilde{A}}{\partial \zeta} \frac{\partial \zeta}{\partial x} \sim \frac{A^\prime(\zeta)}{2t^\delta} & \ \ \mbox{as} \ t \rightarrow \infty \ \mbox{with} \ 0 < \zeta < \infty \ \mbox{fixed} \\ \frac{1}{t^\gamma} \frac{\partial \tilde{{\cal A}}}{\partial \eta} \frac{\partial \eta}{\partial x} \sim \frac{{\cal A}^\prime(\eta)}{t^{\alpha+\gamma}} & \ \ \mbox{as} \ t \rightarrow \infty \ \mbox{with} \ |\eta| < \infty \ \mbox{fixed} \end{array} \right. . \end{equation} Now requiring that the two intermediate limits match yields the final scaling relation: \begin{equation} \alpha + \gamma = \delta \label{eq:agd} \end{equation} as well as the missing boundary condition on the inner problem: \begin{equation} {\cal A}^{\prime}(\infty) = A^\prime(0)/2 . \label{eq:fluxmatch} \end{equation} (Note that $A(\zeta)$ is already fully determined by matching at zeroth order.) This scaling relation (\ref{eq:agd}) can be understood physically as expressing conservation of mass between the diffusion layer and reaction front \cite{jiang}. Similarly, the matching condition (\ref{eq:fluxmatch}) simply means that the diffusive flux entering the reaction front equals the flux leaving the diffusion layer. By examining all possible similarity solutions with power-law couplings of distance and time, we finally arrive at a {\it unique} set of scaling exponents from cases A1, B2 and D2 and (\ref{eq:agd}): \begin{equation} \alpha = \frac{m-1}{2(m+1)},\ \ \beta = \frac{m}{m+1}, \ \ \gamma = \frac{1}{m+1}, \ \ \sigma = \delta = \frac{1}{2} . \label{eq:exponents} \end{equation} Therefore, after long times the reaction front itself \lq\lq diffuses" according to $x_f(t) = -2\nu\sqrt{t}$, where $\nu(q)^2$ is now interpreted as an effective diffusion constant for the front. Although for $m = 1$ the reaction zone settles down to a constant width ($\alpha = 0$), for $m>1$ the front width grows in time ($\alpha > 0$). In all cases the reaction front is \lq\lq infinitely thin" compared to the diffusion layer ($\alpha < \delta$). Note that as $m$ increases, $\gamma$ tends to zero, meaning that the concentration in the reaction front does not decrease as quickly for higher-order reactions as it does for first-order reactions. The exponents $\alpha=0$ and $\gamma = 1/2$ for the case $m=1$ were first obtained by Jiang and Ebner~\cite{jiang} based on physical arguments supported by Monte Carlo simulations and later discussed in an analytical context by Koza~\cite{koza2}, but to our knowledge prior to this work neither have the general expressions for $m \neq 1$ been given nor have the scaling exponents been proven to be unique. With the scaling exponents and matching boundary conditions now determined, we proceed to solve the inner and outer boundary-value problems in the following sections. \subsection{Concentration Profiles in the Diffusion Layer} \label{sec:diff} In the diffusion layer from (\ref{eq:ABlong})--(\ref{eq:Rzero}) we have \begin{equation} -2(\zeta - \nu)A^{\prime} = A^{\prime\prime}, \ \ \ A(0)=0, \ A(\infty)=1 . \label{eq:Aeq} \end{equation} The exact solution to this boundary value problem can be expressed in terms of error functions~\cite{abra} \begin{equation} A(\zeta) = \frac{\mbox{erf}(\zeta-\nu) + \mbox{erf}(\nu)}{1 + \mbox{erf}(\nu)} \label{eq:diffA} \end{equation} and is depicted in Fig.~\ref{fig:A}. Note that the dimensionless flux entering the reaction front \begin{equation} {\cal A}_1(\nu) \equiv \frac{A^\prime(0)}{2} = \frac{e^{-\nu^2}}{\sqrt{\pi}(1 + \mbox{erf}(\nu))} \label{eq:A1} \end{equation} is needed for asymptotic matching in (\ref{eq:fluxmatch}). The effect of varying $q = n^\prime \rho_A^o/(m^\prime\rho_B^o)$ is easily understood in terms of the mathematical model. As $q$ is decreased, the reaction front slows down since reactions in the front region remove species A much faster than diffusion can replenish it. In the limit $q\rightarrow 0$, {\it i.e.} $\rho_A^o\rightarrow 0$, the front comes to a complete stop, $\nu(0) = 0$. For very small, but finite $q > 0$, the concentration of diffusing reactant approximately obeys \begin{equation} \frac{\partial a}{\partial t} = \frac{\partial^2 a}{\partial x^2}, \ \ \ a(x,0) = H(x), \ \ \ a(0,t) = 0, \ \ a(\infty,t) = 1 , \ \ x \geq 0 \end{equation} (at least for $t \ll \nu^{-2}$ since the reaction front is stationary only for short times). This classical diffusion problem has the exact similarity solution \begin{equation} a(x,t) = \mbox{erf}\left(\frac{x}{2\sqrt{t}}\right) , \label{eq:Anuzero} \end{equation} which is precisely the $\nu=0$ curve in Fig.~\ref{fig:A}. From (\ref{eq:diffA}) note that even when the front has moved significantly ($t \gg \nu^{-2}$) the concentration still has the same shape, $A(\zeta) \approx \mbox{erf}(\zeta)$, in the (very slowly) moving reference frame, as long as $\nu(q) \ll 1$. On the other hand, as $q$ is increased, the reaction term becomes progressively less important compared to the diffusion term in (\ref{eq:a}). In the limit $q \rightarrow \infty$, {\it i.e.} $\rho_B^o \rightarrow 0$, we recover another classical diffusion problem (after sufficiently long times $t \gg \nu^{-2}$) \begin{equation} \frac{\partial a}{\partial t} = \frac{\partial^2 a}{\partial x^2}, \ \ \ a(x,0) = H(x),\ \ \ a(-\infty,t) = 0, \ \ a(\infty,t) = 1 , \end{equation} which has the exact similarity solution \begin{equation} a(x,t) = \frac{1}{2} \left[ 1 + \mbox{erf}\left(\frac{x}{2\sqrt{t}}\right)\right] \label{eq:Anuinfty} \end{equation} Note that in the limit $q \rightarrow \infty$ the reaction front instantly speeds off to $-\infty$ having consumed only a negligible amount of reactant A, resulting in a pure diffusion problem for $a(x,t)$. Indeed, we will see below that $\nu(\infty) = \infty$. It remains, of course, to relate $\nu$ and $q$. In Fig.~\ref{fig:A}(b), we see how the true asymptotic similarity solution $A(\zeta)$ interpolates between the limiting forms (\ref{eq:Anuzero}) and (\ref{eq:Anuinfty}) as $\nu$ goes from $0$ to $\infty$, respectively. It turns out that for $\nu \geq 2$ the asymptotic behavior of the original reaction-diffusion system is almost indistinguishable from (\ref{eq:Anuinfty}) for $t \gg 1/4$. \subsection{Diffusion Constant of the Reaction Front} In the reaction front, we have a third-order system of nonlinear ordinary differential equations \begin{subeqnarray} 0 & = & {\cal A}^{\prime\prime} - {\cal A}^m {\cal B}^n \slabel{eq:Ac} \\ \nu {\cal B}^\prime & = & - q {\cal A}^m {\cal B}^n , \slabel{eq:Bc} \label{eq:AcBc} \end{subeqnarray} with four boundary conditions \begin{equation} {\cal A}(-\infty) = 0,\ \, {\cal B}(-\infty) = 1, \ \ {\cal B}(\infty) = 0, \ \ \mbox{and} \ \ {\cal A}^\prime(\infty) = {\cal A}_1 , \end{equation} where ${\cal A}_1(\nu)$ is known via (\ref{eq:A1}). Although this boundary-value problem appears to be overdetermined, the fourth boundary condition is actually necessary to determine the unknown diffusion constant of the reaction front $\nu(q)$. By comparing (\ref{eq:AcBclong}) and (\ref{eq:AcBc}), some physical insight into the dynamics of the reaction front is gained. The concentration of diffusing reactant A is determined by a local balance of reactions and \lq\lq steady state" diffusion and the concentration of static reactant B by a local balance of reactions and fictitious advection due to the translating reference frame. The latter balance reflects the special character of $D_B = 0$: Since reactant B cannot diffuse to the front, instead the front must diffuse to it. This is no longer true if $D_B>0$ (no matter how small), which explains why different scaling exponents arise in that case~\cite{galfi,jiang}. These physical properties are manifested in the mathematical model by the fact that since it multiplies the highest derivative in the equations $D_B > 0$ is a singular perturbation. One integration of (\ref{eq:AcBc}) is easy to perform and fortunately suffices to derive an exact expression for $\nu(q)$. Substituting (\ref{eq:Bc}) into (\ref{eq:Ac}), integrating and applying the boundary conditions at $\eta = \infty$, we obtain \begin{equation} \nu {\cal B} = q ({\cal A}_1 - {\cal A}^\prime) . \end{equation} Likewise enforcing the boundary conditions at $\eta = -\infty$, we find $\nu = q {\cal A}_1$. Substituting ${\cal A}_1$ from (\ref{eq:A1}), we have \begin{equation} \nu(q) = F^{-1}(q) , \label{eq:qnu} \end{equation} where \begin{equation} F(x) \equiv \sqrt{\pi} x e^{x^2}\left[ 1 + \mbox{erf}(x)\right] , \end{equation} which was first derived by Koza~\cite{koza2}. The function $\nu(q)$ is plotted in Fig.~\ref{fig:nuq}. The transcendental function $F(x)$ cannot be inverted analytically, but limiting formulae can be derived. The Maclaurin series of $F(x)$ is \begin{equation} F(x) = \sqrt{\pi} x + 2 x^2 + \sqrt{\pi} x^3 + \frac{4}{3}x^4 + \frac{\sqrt{\pi}}{2}x^5 + \frac{8}{15}x^6 + \frac{\sqrt{\pi}}{6}x^7 + \ldots , \end{equation} which can be inverted term by term to generate the Maclaurin series of $\nu(q)$, valid for small $q$, \begin{equation} \nu(q) = \frac{1}{\pi^{1/2}} q - \frac{2}{\pi^{3/2}}q^2 + \frac{8 - \pi}{\pi^{5/2}}q^3 - \ldots . \end{equation} For large $q$, approximations such as, \begin{equation} \nu(q) \sim \sqrt{\log\left(\frac{q}{\sqrt{\pi}}\right) - \log 2 - \frac{1}{2}\log\log\left(\frac{q}{\sqrt{\pi}}\right) } , \end{equation} can be generated by iteration. \subsection{Existence and Uniqueness of the Reaction-Front Scaling Functions} \label{sec:exist} With the results of the previous section, the inner boundary-value problem is reduced to a nonlinear, second-order equation for ${\cal A}(\eta)$: \begin{equation} {\cal A}^{\prime\prime} = {\cal A}^m(1 - {\cal A}^\prime/{\cal A}_1)^n, \ \ \ {\cal A}(-\infty)=0, \ \ {\cal A}^\prime(\infty)={\cal A}_1 . \label{eq:AcODE} \end{equation} Once this system is solved, ${\cal B}(\eta)$ is recovered from ${\cal B}(\eta) = 1 - {\cal A}^\prime(\eta)/{\cal A}_1$. Note that (\ref{eq:AcODE}) is invariant under translation $\eta \mapsto \eta - \eta_o$, where the arbitrary constant $\eta_o$ sets the precise location of the reaction front. Since $\eta_o$ depends on the exact initial conditions, however, it cannot be determined by considering only the long-time asymptotic limit as we have done here. Since the second-order equation (\ref{eq:AcODE}) is autonomous ({\it i.e.} $\eta$ does not appear), it is useful to consider the \lq\lq Lie diagram''~\cite{dresner} or \lq\lq phase plane''~\cite{bender,davis} of trajectories in the $({\cal A},{\cal B})$ plane parameterized by $\eta$, as shown in Fig.~\ref{fig:Lie}. By studying properties of the phase plane, it is straightforward to prove the existence and uniqueness of solutions if and only if $m,n \geq 1$ and ${\cal A}_1 > 0$. With the change of variables \begin{equation} s \equiv {\cal A}_1^{(m-1)/(m+1)} \eta, \ \ \ u(s) \equiv {\cal B}(\eta), \ \ \ v(s) \equiv {\cal A}_1^{-2/(m+1)} {\cal A}(\eta), \label{eq:uvdef} \end{equation} we begin by transforming (\ref{eq:AcODE}) into a system of first-order equations \begin{subeqnarray} u^\prime & = & - v^m u^n \slabel{eq:uODE} \\ v^\prime & = & 1 - u \slabel{eq:vODE} \label{eq:uv} \end{subeqnarray} with boundary conditions $v(-\infty)=0$ and $u(\infty) = 0$, or equivalently $u(-\infty)=1$ and $v^\prime(\infty)=1$. There is a unique fixed point at $(u,v) = (1,0)$ corresponding to the region ahead of the reaction front which contains only the static reactant B. This is the starting point ($s=-\infty$) of any trajectories that satisfy the boundary condition $v(-\infty)=0$, so our task is to identify and follow any unstable manifolds leaving $(1,0)$ to see if they satisfy the other boundary condition $u(\infty)=0$. If $m=1$, then the equations can be linearized about the fixed point, and $(1,0)$ is a hyperbolic saddle point with an unstable manifold in the $(1,-1)$ direction and a stable manifold in the $(1,1)$ direction, as shown in Fig.~\ref{fig:Lie}(a). If $m\neq 1$, then the stable and unstable manifolds are degenerate at linear order and form a cusp oriented in the $(0,1)$ direction, as shown in Figs.~\ref{fig:Lie}(b) and (c). The nonlinear stability of the fixed point can be determined by noting that $v^{\prime\prime} \sim v^m$ as $u \rightarrow 1$. This equation has solutions satisfying the boundary condition $v(-\infty) = 0$ if and only if $m \geq 1$. A stable reaction front does not exist if $m<1$ because the concentration of diffusing reactant A would become negative ahead of the front. Therefore, since $m \geq 1$ implies $\alpha = (m-1)/2(m+1) \geq 0$, the front width $w(t) \sim t^\alpha$ either stays the same (for $m=1$) or increases (for $m > 1$) but cannot decrease in time. For $m \geq 1$, we integrate $v^{\prime\prime} \sim v^m$ once and substitute into (\ref{eq:vODE}) to obtain the separatrices in the upper half-plane ($v>0$): \begin{equation} u \sim 1 \pm \sqrt{\frac{2}{m+1}} v^{(m+1)/2}, \ \mbox{as} \ (u,v) \rightarrow (1,0^+) , \label{eq:separ} \end{equation} where the upper sign corresponds to the stable manifold ($\eta \rightarrow \infty$) and the lower sign to the unstable manifold ($\eta \rightarrow -\infty$). Let us briefly consider trajectories in the lower half-plane ($v <0$) in the neighborhood of the fixed point ($u \approx 1$). Of course, such trajectories are not physically allowed, but it is satisfying to prove that the model equations exclude such possibilities. If $m$ is either irrational or a rational number of the irreducible form $k_1/k_2$ where $k_2$ is even, then such trajectories do not exist because in that case $v^m$ (and hence $u^\prime$) would not be a real number. If $m = k_1/k_2$ where $k_1$ is even and $k_2$ is odd, then the direction field $(u^\prime,v^\prime)$ is an even function of $v$, which in light of (\ref{eq:separ}) implies that there are no other separatrices in the lower half-plane and that trajectories merely circle the fixed point, as shown in Fig.~\ref{fig:Lie}(b). Finally, if $m=k_1/k_2$ where both $k_1$ and $k_2$ are odd, then $u^\prime$ is an odd function of $v$ (while $v^\prime$ is even), and $(1,0)$ is a saddle point. In this case, Eq.~(\ref{eq:separ}) also describes separatrices in the lower half-plane with the lower sign corresponding to the stable manifold and the upper to the unstable manifold, as shown in Fig.~\ref{fig:Lie}(c). As it leaves the fixed point, this branch of the unstable manifold enters the region $(u>1,v<0)$, throughout which $u^\prime>0$ and $v^\prime<0$, and thus it heads off to $u=v=-\infty$ and cannot satisfy the other boundary condition, as shown in Figs.~\ref{fig:Lie}(a) and (c). Therefore, any solutions must lie entirely in the first quadrant of the phase plane ($u>0, v>0$). In this way we are left with only one possible solution, which leaves the fixed point along the unstable manifold of (\ref{eq:separ}) and enters the region defined by $0<u<1$ and $v>0$, throughout which $u^\prime<0$ and $v^\prime>0$. Since the $v$-axis ($u=0$) is itself a trajectory, which cannot be crossed, this candidate solution must reach an asymptote $u(\infty)=u_o$, for some constant $0\leq u_o < 1$. However, it is clear from (\ref{eq:uODE}) that $u_o=0$ is the only possible asymptote, which implies that the trajectory (if it exists) must satisfy the other boundary condition $u(\infty)=0$. To check the existence of this solution in the limit $s \rightarrow\infty$, note that $v^\prime \sim 1$ which implies $u^\prime \sim -(s - s_o)^m u^n$ for some constant $s_o$ with $m>1$. Solutions to this equation satisfying $u(\infty)=0$ exist if and only if $n \geq 1$. A stable reaction front does not exist if $n < 1$ because the concentration of static reactant B would be negative behind of the front. \subsection{Concentration Profiles in the Reaction Front} Although solutions to (\ref{eq:AcODE}) exist for $m,n \geq 1$, they are not easily expressed in terms of elementary functions. The exact trajectories in the phase plane, however, can be obtained. The ratio of (\ref{eq:vODE}) and (\ref{eq:uODE}) yields a separable, first-order equation for $v(u)$ \begin{equation} \frac{dv}{du} = \frac{u-1}{v^m u^n} \label{eq:vu} \end{equation} which can be integrated to obtain the one-parameter family of trajectories (for $u>0$) plotted in Fig.~\ref{fig:Lie} \begin{equation} (m+1)^{-1} v^{m+1} = \left\{ \begin{array}{ll} c_1 + u - \log u & \ \mbox{if} \ n=1 \\ c_2 +\log u + u^{-1} & \ \mbox{if} \ n=2 \\ c_n - (n-2) u^{2-n} + (n-1)^{-1} u^{1-n} & \ \mbox{if} \ n>1, n\neq 2 \end{array} \right. \end{equation} indexed by the real number $c_n$. Applying the boundary conditions $v(u=1)=0$ (which determines $c_n$) and $v(0)=\infty$ (which selects the positive branch when $v(u)$ is multivalued), we arrive at the exact phase-plane trajectories (in the region $0<u<1$, $v>0$) of the solution to the inner problem: \begin{equation} v = \left\{ \begin{array}{ll} \left[ (m+1) \left(u - 1 - \log u\right) \right]^{1/(m+1)} & \ \mbox{if} \ n=1 \\ \left[ (m+1) \left(u^{-1} - 1 + \log u\right)\right]^{1/(m+1)} & \ \mbox{if} \ n=2 \\ \left[\frac{(m+1)\left( 1 - (n-1)u^{2-n} + (n-2)u^{1-n}\right)}{(n-1)(n-2)}\right]^{1/(m+1)} & \ \mbox{if} \ n>1, n\neq 2 \end{array} \right. \label{eq:Lie} \end{equation} which is an algebraic equation $v = g_{m,n}(u)$ relating ${\cal A}(\eta)$ and ${\cal B}(\eta)$ via (\ref{eq:uvdef}). This equation is transcendental, but in some cases it is easily solved for $u = g_{m,n}^{-1}(v)$, {\it e.g.} for $n=3$ we have \begin{equation} u = g_{m,3}^{-1}(v) = \frac{\sqrt{2(m+1)^{-1}v^{m+1}} - 1} {2(m+1)^{-1}v^{m+1} - 1} . \end{equation} Note that $g_{m,n}^{-1}(0) = 1$ and $g_{m,n}^{-1}(\infty) = 0$. By substituting (\ref{eq:Lie}) into (\ref{eq:vODE}) we arrive at a first-order equation for $v(\eta)$ \begin{equation} v^\prime = 1 - g_{m,n}^{-1}(v) \label{eq:vg} \end{equation} without any boundary conditions (because the conditions $v(-\infty)=0$ and $v^\prime(\infty)=1$ are automatically satisfied). Since (\ref{eq:vg}) is separable, the solution to the inner problem can be expressed in the form: \begin{subeqnarray} {\cal A}(\eta) & = & {\cal A}_1^{2/(m+1)} h_{m,n}^{-1}\left( {\cal A}_1^{(m-1)/(m+1)}(\eta-\eta_o)\right) \slabel{eq:Ac_soln} \\ {\cal B}(\eta) & = & g_{m,n}^{-1}\left[h_{m,n}^{-1}\left( {\cal A}_1^{(m-1)/(m+1)}(\eta-\eta_o)\right) \right] \slabel{eq:Bc_soln} \label{eq:AcBc_soln} \end{subeqnarray} where \begin{equation} h_{m,n}(v) \equiv \int_{v_o}^v \frac{ds}{1-g_{m,n}^{-1}(s)} , \label{eq:hmn} \end{equation} The precise location of the reaction front is set by choosing $v(s_o) = v_o > 0$ for some constant $s_o = {\cal A}_1^{(m-1)/(m+1)}\eta_o$. With these results, the inner problem is reduced to the solution of two algebraic equations for $g_{m,n}^{-1}(v)$ and $h_{m,n}^{-1}(\eta)$ and one quadrature (\ref{eq:hmn}). In practice, however, it is simpler to integrate (\ref{eq:AcODE}) directly. Numerical solutions of the rescaled inner problem \begin{equation} v^{\prime\prime} = v^m(1-v^\prime)^n, \ \ \ v(-\infty)=0, \ \ \ v^{\prime}(\infty) = 1 \label{eq:veq} \end{equation} are obtained by a shooting method, with the results shown in Fig.~\ref{fig:Ac}(a) for $m=n=1$ and $m=n=2$. The position of the front is chosen such that $v(5)=5$. The static reactant concentration $u(s)$ and the reaction rate density $v^m u^n$ are shown in Figs.~\ref{fig:Ac}(b) and (c), respectively. Note that the concentration fields decay to their asymptotic values as $|\eta| \rightarrow \infty$ more slowly as $m$ and $n$ are increased above unity, a phenomenon that we explore analytically in next section. \subsection{Localization of the Reaction Front} The width of the reaction front varies in time according to $w(t) \sim t^\alpha$ where $\alpha = (m-1)/2(m+1)$. On the scale of the diffusion layer width $W(t) \sim t^{1/2}$, the reaction front is \lq\lq localized" after long times because $\alpha < 1/2$. Note that the overall localization of the front $w(t)/W(t)$ is controlled by $m$ (the reaction order of the diffusing species A), but we now show that both $m$ and $n$ (the reaction order of the static species B) affect localization on the scale of $w(t)$. Specifically, we derive the spatial decay of the scaled reaction rate ${\cal R}(\eta) = {\cal A}(\eta)^m {\cal B}(\eta)^n$ in terms of the inner similarity variable $\eta \rightarrow \pm \infty$ (see Fig.~\ref{fig:Ac}(c)). The actual reaction rate decays uniformly to zero in time, $r \sim t^{-\beta}{\cal R}(\eta)$ with $\beta = m/(m+1)$, but here we are only concerned with the shape of ${\cal R}(\eta)$. Ahead of the reaction front in the limit $\eta \rightarrow -\infty$, we have ${\cal A}^{\prime\prime} \sim {\cal A}^m$ from (\ref{eq:AcODE}), which can be integrated to obtain the decay of concentration fields: \begin{subeqnarray} {\cal A}(\eta) & \sim & \left\{ \begin{array}{ll} {\cal A}_2 e^{-|\eta|} & \ \mbox{if} \ m=1 \\ {\cal A}_2 |\eta|^{-2/(m-1)} & \ \mbox{if} \ m>1 \end{array} \right. \slabel{eq:Ac_ahead} \\ 1- {\cal B}(\eta) & \sim & \left\{ \begin{array}{ll} {\cal A}_3 e^{-|\eta|} & \ \mbox{if} \ m=1 \\ {\cal A}_3 |\eta|^{-(m+1)/(m-1)} & \ \mbox{if} \ m>1 \end{array} \right. \slabel{eq:Bc_ahead} \\ {\cal R}(\eta) & \sim & \left\{ \begin{array}{ll} {\cal A}_2^m e^{-m|\eta|} & \ \mbox{if} \ m=1 \\ {\cal A}_2^m |\eta|^{-2m/(m-1)} & \ \mbox{if} \ m>1 \end{array} \right. \end{subeqnarray} where ${\cal A}_2$ and ${\cal A}_3$ are constants. Note that the localization of ${\cal A}(\eta)$ and ${\cal R}(\eta)$ ahead of the front is entirely controlled by the reaction order $m$ of the depleted reactant (which is the diffusing species $A$). There is a transition from an exponential decay for $m=1$ to a slower power-law decay for $m>1$. Next we consider localization of ${\cal R}(\eta)$ behind the reaction front in the limit $\eta \rightarrow \infty$. From asymptotic matching with the diffusion layer we have already derived ${\cal A}(\eta) \sim {\cal A}_1 \eta$. The asymptotic decay of ${\cal B}(\eta)$ and ${\cal R}(\eta)$ is obtained by integrating ${\cal A}_1 {\cal B}^\prime \sim - ({\cal A}_1 \eta)^m {\cal B}^n$: \begin{subeqnarray} \slabel{eq:Bctail} {\cal B}(\eta) & \sim & \left\{ \begin{array}{ll} {\cal B}_1 \exp\left[-{\cal A}_1^{m-1}\eta^{m+1}/(m+1)\right] & \ \mbox{if} \ n=1 \\ {\cal B}_1 \eta^{-(m+1)/(n-1)} & \ \mbox{if} \ n>1 \end{array} \right. \\ \slabel{eq:Rctail} {\cal R}(\eta) & \sim & \left\{ \begin{array}{ll} {\cal A}_1^m {\cal B}_1^n \eta^m \exp\left[-{\cal A}_1^{m-1}\eta^{m+1}/(m+1)\right] & \ \mbox{if} \ n=1 \\ {\cal A}_1^m {\cal B}_1^n \eta^{-(m+n)/(n-1)} & \ \mbox{if} \ n>1 \end{array} \right. \end{subeqnarray} where ${\cal B}_1$ is a constant. Once again, a higher reaction order $n$ for the depleted species (which is the static species B) broadens the front: There is another transition from exponential decay for $n=1$ to a power-law decay for $n>1$. Note, however, that increasing the reaction order $m$ of the diffusing species A contracts the back side of the front. The fact that ${\cal R}(\eta)$ has a fairly broad, power-law decay for $n > 1$ toward the diffusion layer should cause concern since we have previously assumed in (\ref{eq:Rzero}) that the reaction term is negligible in the diffusion layer. Fortunately, however, for all $m,n\geq 1$ the decay of ${\cal B}(\eta)$ is just fast enough to satisfy (\ref{eq:Rzero}) in the intermediate region where $B(0<\zeta\ll 1) \approx {\cal B}(\eta \gg 1)$. From (\ref{eq:Bctail}) along with $\tilde{A}(\zeta,t) \sim A(\zeta)$, $\tilde{{\cal B}}(\eta,t) \sim {\cal B}(\eta)$, $\eta = 2\zeta t^{1/(m+1)}$ and $\delta = 1/2$, we have \begin{equation} t^{2\delta} \tilde{A}(\zeta,t)^m \tilde{B}(\zeta,t)^n = \left\{ \begin{array}{ll} O\left( t \exp\left[-{\cal A}_1^{m-1}(2\zeta)^{m+1}t/(m+1)\right] \right) & \ \mbox{if} \ n = 1 \\ O\left( \zeta^{-n(m+1)/(n-1)} t^{-n/(n-1)} \right) & \ \mbox{if} \ n > 1 \end{array} \right. \end{equation} as $t \rightarrow \infty$ with $0<\zeta\ll 1$ fixed, which verifies (\ref{eq:Rzero}) in the intermediate region for any $m, n \geq 1$. However, we now prove that (\ref{eq:Rzero}) actually holds throughout the diffusion layer for all $\zeta > 0$. \section{Transient Decay in the Diffusion Layer} \label{sec:transients} \subsection{Assumption of Quasi-Stationarity} The analysis of long-time asymptotics in the previous section rests on two basic assumptions: $(i)$ \lq\lq scale separation", given by \begin{subeqnarray} 0 < \lim_{t\rightarrow \infty} \tilde{{\cal A}}(\eta,t)^m \tilde{B}(\zeta,t)^n & < & \infty \ \ \ \mbox{for} \ |\eta| < \infty \\ \lim_{t\rightarrow \infty} t \cdot \tilde{A}(\zeta,t)^m \tilde{B}(\zeta,t)^n & = & 0 \ \ \ \mbox{for} \ \zeta > 0 , \slabel{eq:SS} \label{eq:SSboth} \end{subeqnarray} and $(ii)$ \lq\lq quasi-stationarity'', given by (\ref{eq:AcBc_converge}) and (\ref{eq:AB_converge}) along with \begin{subeqnarray} \lim_{t\rightarrow \infty} t^{(m-1)/(m+1)}\frac{\partial \tilde{{\cal A}}}{\partial t}(\eta,t) & = & 0 \ \ \ \mbox{for} \ |\eta|<\infty \slabel{eq:QSAc} \\ \lim_{t\rightarrow \infty} t^{m/(m+1)}\frac{\partial \tilde{{\cal B}}}{\partial t}(\eta,t) & = & 0 \ \ \ \mbox{for} \ |\eta|<\infty \slabel{eq:QSBc} \\ \lim_{t\rightarrow \infty} t \cdot \frac{\partial \tilde{A}}{\partial t}(\zeta,t) & = & 0 \ \ \ \mbox{for} \ \zeta > 0 \slabel{eq:QSA} . \label{eq:QS} \end{subeqnarray} Assumption $(i)$ states that two spatial regions with disparate length scales, the reaction front and the diffusion layer, arise where the reaction term on the right-hand side of the governing partial integro-differential equation (\ref{eq:ide1}) or (\ref{eq:ide2}) is, respectively, either comparable to or dominated by the diffusion term. Assumption $(ii)$ states that, when viewed on scales appropriate for each region, the solution to the initial-boundary-value problem (\ref{eq:eqs}) approaches an asymptotically self-similar form, which is suggested by the fact that there is no natural length scale in the problem. These ubiquitous assumptions~\cite{galfi,koza1,koza2} have been rigorously justified~\cite{schenkel,vanbaalen} in the special case of a perfectly symmetric ($D_A=D_B$, $\rho_A^o=\rho_B^o$), and thus stationary ($\nu=0$), reaction front involving two diffusing reactants with certain kinetic orders ($m=n=1$ and $m=n>3$). To our knowledge, a similar mathematical validation of these assumptions has not been given for the general situation of a moving reaction front with arbitrary kinetic orders for either one or two diffusing reactants. For one static reactant, however, convergence at the diffusive scale has been rigorously established by Hilhorst et al.~\cite{peletier96}, even with a very general reaction term~\cite{peletier99}. In this section, we prove the more modest result that quasi-stationarity implies scale separation, {\it i.e.} (\ref{eq:QS}) implies (\ref{eq:SSboth}). (It suffices to show (\ref{eq:SS}) since (\ref{eq:SSboth}) follows from the definition of $\eta$ in section~\ref{sec:Ac}.) Although this analysis justifies {\it a posteriori} the assumption of scale separation in our fairly general situation ($D_B=0$, $\nu > 0$, $m,n \geq 1$), it more importantly reveals the transient decay to the asymptotic similarity solution in the diffusion layer. Specifically, we derive exact formulae for the asymptotic decay of the reaction-rate density and static-reactant concentration in the diffusion layer. We begin by precisely stating our assumptions related to (\ref{eq:QS}). From (\ref{eq:QSAc}) and its consequence (\ref{eq:Ac_ahead}), we conclude that the diffusing reactant concentration vanishes on the scale $W(t) = \sqrt{t}$ ahead of the front (where there is no diffusion layer) since $\tilde{A}(\zeta,t) \sim {\cal A}(2\zeta t^{1/(m+1)}) \rightarrow 0$ as $t \rightarrow \infty$ with $\zeta < 0$ fixed. This result can be combined with (\ref{eq:A_converge}) to obtain a statement of quasi-stationarity on the scale $W(t)$: \begin{equation} \tilde{A}(\zeta,t) \rightarrow A(\zeta)H(\zeta) \ \ \mbox{and} \ \ \frac{\partial \tilde{A}}{\partial \zeta}(\zeta,t) \rightarrow A^\prime(\zeta)H(\zeta) \label{eq:AH} \end{equation} as $t \rightarrow \infty$ with $\zeta \neq 0$ fixed. We have already derived the exact form of the similarity function $A(\zeta)$ in (\ref{eq:diffA}) as a consequence of neglecting reactions (\ref{eq:SS}). To avoid a circular argument, however, we must now establish (\ref{eq:SS}) without using (\ref{eq:diffA}), thus giving {\it a posteriori} justification for the latter equation. Throughout section~\ref{sec:transients} our only assumptions about $A(\zeta)$ are $A(0)=0$ and $A^\prime(0)>0$. These properties follow from matching with the reaction front, where $a \rightarrow 0$ ({\it i.e.} $\gamma > 0$) follows from quasi-stationarity, as shown in section~\ref{sec:Ac}. \subsection{Direction of the Diffusing-Reactant Flux} Let us prove that $A(\zeta)$ is strictly increasing in the diffusion layer, $A^\prime(\zeta) > 0$ for all $\zeta \geq 0$, as a consequence of (\ref{eq:SSboth}). Combining (\ref{eq:QSA}) with (\ref{eq:Along}), we have \begin{equation} -2(\zeta-\nu)\frac{\partial \tilde{A}}{\partial \zeta} \sim \frac{\partial^2 \tilde{A}}{\partial \zeta^2} - t \tilde{A}^m \tilde{B}^n \ \ \ \mbox{as} \ t\rightarrow \infty \ \mbox{with} \ \zeta > 0 \ \mbox{fixed} \end{equation} which is easily integrated once using an integrating factor, \begin{equation} \frac{\partial \tilde{A}}{\partial \zeta}(\zeta,t) \sim e^{-(\zeta-\nu)^2} \left[ e^{\nu^2} \frac{\partial \tilde{A}}{\partial \zeta}(0,t) + t \int_0^\zeta \tilde{A}(\xi,t)^m \tilde{B}(\xi,t)^n e^{(\xi-\nu)^2}d\xi \right] . \label{eq:dAt} \end{equation} Since the second term on the right-hand side is non-negative for all $t>0$, we can pass to the limit $t\rightarrow \infty$ for any fixed $\zeta> 0$ to obtain the desired bound \begin{equation} A^\prime(\zeta) \geq A^\prime(0)e^{\nu^2-(\zeta-\nu)^2} > 0 , \label{eq:dAbound} \end{equation} which expresses the physical fact that everywhere in the diffusion layer a nonzero flux of the diffusing species is directed toward the reaction front (at sufficiently large times). \subsection{Decay of the Static-Reactant Concentration} We now prove that $b(x,t)$ vanishes asymptotically in the diffusion layer as a consequence of quasi-stationarity, which implies \begin{equation} \tilde{B}(\zeta,t) \rightarrow B(\zeta) = H(-\zeta)\ \ \ \mbox{as} \ t \rightarrow \infty \ \mbox{with} \ \zeta \neq 0 \ \mbox{fixed} . \label{eq:Bzeta} \end{equation} For $\zeta < 0$, this follows from (\ref{eq:Bc_ahead}) since there is no diffusion layer ahead of the front, and therefore $\tilde{B}(\zeta,t) \sim {\cal B}(2\zeta t^{1/(m+1)}) \rightarrow 1 = B(\zeta)$ as $t \rightarrow \infty$ with $\zeta < 0$ fixed. Likewise, in section \ref{sec:dim}, we have already established (\ref{eq:Bzeta}) for $\zeta > \nu$ since $\tilde{B}(\zeta,t)$ vanishes there identically for all times. Therefore, it only remains to prove that $B(\zeta) = 0$ for $0 < \zeta < \nu$. In light of the expression for $b(x,t)$ in (\ref{eq:bint}), the definition of $\phi_m(x,t)$ in (\ref{eq:phidef}) and the restriction $n \geq 1$, it suffices to show that $\Phi_m(\zeta,t) \equiv \phi_m(x,t) \rightarrow \infty$ as $t \rightarrow \infty$ for $0 < \zeta < \nu$ fixed. Using $\zeta = \nu + x/2\sqrt{t}$ (since $\delta=\sigma=\frac{1}{2}$), we transform $\phi_m(x,t)$ into the diffusion-layer coordinates $(x,t) \mapsto (\zeta,t)$ \begin{equation} \Phi_m(\zeta,t) = \int_0^t a\left(2\sqrt{t}(\zeta-\nu),\tau\right)^m d\tau , \end{equation} and express this in terms of the diffusion-layer scaling function $a(x,\tau) = \tilde{A}(\nu + x/2\sqrt{\tau},\tau)$ \begin{equation} \Phi_m(\zeta,t) = \int_0^t \tilde{A}\left(\sqrt{\frac{t}{\tau}} (\zeta-\nu)+\nu, \tau\right)^m d\tau . \end{equation} It is convenient to work with the partial time-derivative of $\Phi_m(\zeta,t)$ given by the Leibniz rule: \begin{equation} \frac{\partial \Phi_m}{\partial t} = \tilde{A}(\zeta,t)^m + \int_0^t \frac{\partial \tilde{A}^m}{\partial \zeta} \left(\sqrt{\frac{t}{\tau}}(\zeta-\nu)+\nu, \tau\right) \frac{1}{2t}\sqrt{\frac{t}{\tau}}(\zeta-\nu)d\tau . \end{equation} Focusing on the region $0 < \zeta < \nu$, we make the transformation $\xi = \sqrt{t/\tau}(\zeta - \nu) + \nu$, \begin{equation} \frac{\partial \Phi_m}{\partial t} = \tilde{A}(\zeta,t)^m - \int_{-\infty}^\zeta \frac{\partial \tilde{A}^m}{\partial\zeta} \left(\xi,\left(\frac{\zeta-\nu}{\xi-\nu}\right)^2t\right) \left(\frac{\zeta-\nu}{\xi-\nu}\right)^2 d\xi , \label{eq:phitrans} \end{equation} and pass the limit $t\rightarrow \infty$ inside the integral to obtain \begin{equation} \lim_{t\rightarrow \infty}\frac{\partial \Phi_m}{\partial t} = A(\zeta)^m - \int_0^\zeta \frac{dA^m}{d\zeta}(\xi) \left(\frac{\zeta-\nu}{\xi-\nu}\right)^2 d\xi , \label{eq:dphi0} \end{equation} where the lower limit of integration follows from (\ref{eq:AH}) since $A(\zeta) = A^\prime(\zeta) = 0$ for $\zeta < 0$. This step is justified by the Dominated Convergence Theorem~\cite{evans} because, by virtue of (\ref{eq:AH}), there exist constants $M, t_o > 0$ such that the integrand in (\ref{eq:phitrans}) is bounded for all $t > t_o$ by $M/(\xi - \nu)^2$, which is integrable on $(-\infty,\zeta)$, if $\zeta < \nu$. Since $A(0) = 0$ is required by matching between the two regions of quasi-stationarity, Eq.~(\ref{eq:dphi0}) can be written in the form $\partial \Phi_m/\partial t \sim f_m(\zeta)$, where \begin{equation} f_m(\zeta) \equiv \int_0^\zeta \frac{dA^m}{d\zeta}(\xi) \left[ 1 - \left(\frac{\zeta-\nu}{\xi-\nu}\right)^2\right] d\xi . \label{eq:fm} \end{equation} Note that $f_m(\zeta)>0$ for $0 < \zeta < \nu$ since $A^\prime(\zeta)>0$ in this region, as shown in (\ref{eq:dAbound}). Therefore, with an integration of (\ref{eq:fm}), we arrive at the desired result \begin{equation} \Phi_m(\zeta) \sim f_m(\zeta)t \ \ \mbox{as} \ \ t \rightarrow \infty \ \ \mbox{with} \ \ 0 < \zeta < \nu \ \ \mbox{fixed,} \label{eq:Phit} \end{equation} thus completing the proof that $B(\zeta) = 0$ for $\zeta> 0$. By substituting (\ref{eq:Phit}) into (\ref{eq:bint}), we obtain the transient decay of $\tilde{B}(\zeta,t)$ in part of the diffusion layer where the reaction front has already passed ($0 < \zeta \leq \nu$): \begin{equation} \tilde{B}(\zeta,t) \sim \left\{ \begin{array}{ll} e^{-qf_m(\zeta)t} & \ \mbox{if} \ n=1 \\ \left[ q(n-1)f_m(\zeta) t \right]^{-1/(n-1)} & \ \mbox{if} \ n >1 \end{array} \right. \label{eq:Blimit} \end{equation} Note that $\tilde{B}(\zeta,t)$ vanishes with exponential decay if $n=1$ and with a power-law decay if $n > 1$. Therefore, by measuring the asymptotic decay (either exponential or power-law) of the static-reactant concentration in the diffusion layer, the reaction order $n$ could in principle be inferred from experimental data (although such measurements are difficult in practice~\cite{leger}). \subsection{Decay of the Reaction-Rate Density} From (\ref{eq:Blimit}) we easily obtain the asymptotic decay on the reaction rate density in the diffusion layer as $t \rightarrow \infty$ with $\zeta>0$ fixed: \begin{equation} t \tilde{A}(\zeta,t)^m \tilde{B}(\zeta,t)^n \sim \left\{ \begin{array}{ll} A(\zeta)^m t e^{-qf_m(\zeta)t} & \ \mbox{if} \ n=1 \\ A(\zeta)^m \left[ q(n-1)f_m(\zeta)\right]^{-\frac{n}{n-1}} t^{-\frac{1}{n-1}} & \ \mbox{if} \ n >1 \end{array} \right. , \label{eq:Rtransient} \end{equation} which establishes (\ref{eq:SS}). The reaction term in the diffusion layer has previously been neglected based only on physical intuition~\cite{koza2}, but here we have given a mathematical justification. \subsection{The Decay Time When $n=1$} Since the reaction term vanishes sufficiently fast in the diffusion layer to justify {\it a posteriori} the analysis in section~\ref{sec:similarity}, the exact expression for $A(\zeta)$ from (\ref{eq:diffA}) may be substituted into (\ref{eq:fm}) to evaluate the function $f_m(\zeta)$. If $n>1$, then $f_m(\zeta)$ affects the transient decay in (\ref{eq:Blimit}) and (\ref{eq:Rtransient}) only as a multiplicative prefactor in a power law, which would be difficult to measure in a real experiment. If $n=1$, however, then $f_m(\zeta)$ sets the characteristic time $\tau_m(\zeta)^{-1} \equiv q f_m(\zeta)$ of an exponential decay, which is easier to measure experimentally. Therefore, we now derive an exact expression for the decay time $\tau_1(\zeta)$ ($0<\zeta\leq \nu$) in the case $m=1$: \begin{eqnarray} \tau_1(\zeta)^{-1} & = & q f_1(\zeta) \nonumber \\ & = & \int_0^\zeta 2\nu e^{\nu^2-(\nu-\xi)^2} \left[ 1 - \left(\frac{\zeta-\nu}{\xi-\nu}\right)^2\right] d\xi \nonumber \\ & = & 2(\nu-\zeta)\left[ (\nu-\zeta) - \nu e^{\nu^2-(\nu-\zeta)^2}\right] \nonumber \\ & & + \sqrt{\pi} \nu e^{\nu^2} \left[ 1 + 2(\nu-\zeta)^2 \right]\cdot \left[ \mbox{erf}(\nu) - \mbox{erf}(\nu-\zeta) \right] . \label{eq:tau} \end{eqnarray} Note that $\tau_1(0) = \infty$ in the vicinity of the reaction front ($\zeta = 0$) because (\ref{eq:Rtransient}) and (\ref{eq:Blimit}) no longer hold. Within the diffusion layer, the decay time is a decreasing function of distance $\zeta$ away from the reaction front, as shown in Fig.~\ref{fig:tau}. These results may be used to infer reaction orders and perhaps even kinetic constants in diffusion-limited corrosion experiments from transient decay measurements of the reaction-rate density in the diffusion layer~\cite{leger}. \section{Uniformly Valid Asymptotic Approximations} \label{sec:uniform} In the previous two sections we have argued for the existence of a unique asymptotic similarity solution (up to an unknown constant $\eta_o$) contingent upon certain \lq\lq quasi-stationarity" conditions, which are likely to be satisfied for the specified initial conditions (see below). This solution, valid after long times, consists of two different asymptotic approximations for $a(x,t)$, the concentration of the diffusing reactant A, which reflect the different couplings of length and time in the reaction front and the diffusion layer. A single asymptotic approximation for $a(x,t)$ that is uniformly valid across all space is obtained by adding the two contributions from the reaction front (the inner region) and the diffusion layer (the outer region) and subtracting the overlap (from the intermediate region)~\cite{hinch,bender}: \begin{equation} a(x,t) \sim \left[{\cal A}(\eta-\eta_o)- {\cal A}_1 \cdot (\eta-\eta_o) H(\eta-\eta_o)\right]t^{-1/(m+1)} + A(\zeta) H(\zeta), \ \mbox{as} \ t \rightarrow\infty \ \mbox{for\ all\ } x . \label{eq:a_unif} \end{equation} where the reaction-front and diffusion-layer similarity variables are \begin{subeqnarray} \eta(x,t) & = & \frac{x + 2\nu t^{1/2}}{t^{(m-1)/2(m+1)}} \\ \mbox{and} \ \ \zeta(x,t) & = & \frac{x + 2\nu t^{1/2}}{2t^{1/2}} , \end{subeqnarray} $\nu(q)^2$ is the diffusion constant of the reaction front (see (\ref{eq:qnu}) and Fig.~\ref{fig:nuq}), ${\cal A}_1 = \nu(q)/q$ is a constant proportional to the diffusive flux entering the front, ${\cal A}(\eta)$ is the reaction-front similarity function (see (\ref{eq:Ac_soln}) and Fig.~\ref{fig:Ac}), $A(\zeta)$ is the diffusion-layer similarity function (see (\ref{eq:diffA}) and Fig.~\ref{fig:A}) and $\eta_o$ is an undetermined constant depending upon the initial conditions as well as the precise definition of the reaction-front location. The uniform approximation has been determined analytically up to the solution of two algebraic equations (\ref{eq:AcBc_soln}) and one quadrature (\ref{eq:hmn}). A subtle point in the construction of this uniformly valid approximation is that shifting the position of the front by $\eta \mapsto \eta - \eta_o$ does not affect matching with the diffusion layer because in that case $\zeta \mapsto (\eta-\eta_o)t^{-1/(m+1)}/2 \sim \eta t^{-1/(m+1)}/2 = \zeta$. In other words, because the reaction front is \lq\lq infinitely thin" compared to the diffusion layer, translating its similarity variable by a constant $\eta_o$, or any other function of time that is $o(t^{1/2})$, does not require that the diffusion-layer similarity variable $\zeta$ be shifted as well. The situation for $b(x,t)$, the concentration of the static reactant B, is much simpler. By comparing the asymptotic bound on $b(x,t)$ in the diffusion layer given by (\ref{eq:Blimit}) with the tail of the reaction-front approximation given by (\ref{eq:Bctail}) with $\eta = 2\zeta t^{1/(m+1)}$, we see that that the asymptotic behavior of $b(x,t)$ is identical in the two regions. Therefore, \begin{equation} b(x,t) \sim {\cal B}(\eta-\eta_o), \ \mbox{as} \ t \rightarrow\infty \ \mbox{for\ all\ } x \label{eq:b_unif} \end{equation} is a uniformly valid approximation, where ${\cal B}(\eta)$ is the reaction-front similarity function given by (\ref{eq:Bc_soln}). At this point the initial conditions have not yet entered the analysis except in (\ref{eq:bint}), which only influences the prefactors of the transient-decay formulae in section~\ref{sec:transients}. Therefore, the asymptotic similarity solution is universal up to a constant shift of the reaction front by $\eta_o$ for some broad set of initial conditions which presumably contains (\ref{eq:ic}). In general, this \lq\lq universality class'' of initial conditions leading to the same asymptotic similarity solution (up to different values of $\eta_o$) is expected to be attained whenever the initial reaction-rate distribution $r(x,0) = a(x,0)^m b(x,0)^n$ is sufficiently well localized and the reactants are sufficiently well separated. This class surely contains all initial conditions for which $r(x,0)$ has compact support, {\it e.g.} $r(x,0) = 0$ for $x\neq 0$ in (\ref{eq:ic}), or exponential decay, {\it e.g.} $r(x,0) < Me^{-|x|/x_o}$ for some $M,x_o>0$, but perhaps not slower power-law decay. \section{Discussion} \label{sec:disc} In this article we have studied the long-time asymptotics of solutions to the initial-boundary-value problem of (\ref{eq:rhoBC})--(\ref{eq:rho}), which is a generic mean-field model for the corrosion of a porous solid by a diffusing chemical. We have derived a uniformly valid asymptotic approximation (\ref{eq:a_unif})--(\ref{eq:b_unif}) consisting of matched similarity solutions in two distinct regions, the reaction front and diffusion layer, each possessing different power-law scaling behavior. The existence and uniqueness of the similarity functions and the scaling exponents have been established if and only if $m,n \geq 1$, and through an analysis of transients in the diffusion layer the asymptotic scale separation has been shown to follow from the assumption of quasi-stationarity. Since quasi-stationarity has been observed in recent experiments on the corrosion of ramified electrodeposits~\cite{leger}, the present analysis therefore suffices to establish the theoretical predictions of the mean-field equations for at least one particular corrosion system. Although the case considered here ($m,n\geq 1$, $q \neq 1$) is more complicated, it would be useful to perform a rigorous transient analysis along the lines of Schenkel {\it et al.}~\cite{schenkel} (who considered only the case $D_A=D_B\neq 0$, $q=1$ and $m=n=1$). Nevertheless, we have at least provided a firm mathematical justification for the scale separation between the diffusion layer and reaction front. In this work we have paid special attention to the effect of higher-order reactions ($m,n > 1$). First of all, the scaling exponents vary with the reaction order $m$ of the diffusing reactant in precisely the same way as they do on the sum $m+n$ in the case of two diffusing reactants~\cite{cornell}, as shown in Table ~\ref{table:exponents}. Moreover, the spatial localization of the reaction rate $r(a,b)$ on each side of the front depends primarily on the reaction order of the depleted reactant: As the appropriate reaction order is increased from unity, the spatial dependence of the reaction rate away from the front changes from an exponential decay to a progressively broader power-law decay. Similarly, the temporal decay of the depleted (static) reactant concentration in the diffusion layer depends sensitively on its reaction order, undergoing a transition from exponential to power-law decay (in time) as $n$ is increased from unity. These properties may have general relevance for more complicated multi-component reaction-diffusion systems. Other qualitiative features of our analysis that might have more general applicability are the dominant balances in the reaction front, where the concentration of a diffusing reactant is determined by a balance between reactions and \lq\lq steady state" diffusion, {\it i.e.} a mobile species diffuses slowly to the front where it immediately reacts. On the other hand, the concentration of a static reactant is determined by a balance between reactions and fictitous advection due to the moving reference frame, {\it i.e.} since the static species cannot diffuse to the front, the front must diffuse to it. These guiding principles might help simplify more complicated reaction-diffusion equations for which asymptotically self-similar solutions do not exist. \section*{Epilogue} After the writing of this article, a referee pointed out some interesting similarities between our analysis of chemical reaction fronts and various existing studies in combustion theory~\cite{ZFK,zeldovich}. Indeed, the equations for combustion waves introduced by Zeldovich and Frank-Kamenetzki~\cite{ZFK}, which have since been generalized by many authors (e.g. Matkowsky and Sivashinsky~\cite{MS}), bear some resemblance to (\ref{eq:coup}) since they describe the diffusion of a fuel substance coupled to the diffusion of heat. In combustion theory, however, the usual reaction term is quite different from (\ref{eq:Rpower}) because it involves exponential Arrhenius temperature dependence, and the initial and boundary conditions also differ from those considered here. As a result, combustion waves tend to exhibit qualitatively different behavior from reaction-diffusion fronts. (For example, simple flame fronts have constant width and constant velocity.) Nevertheless, combustion waves exhibit multiple scales analogous to the reaction-front and diffusion layers described here, which have also been analyzed using matched asymptotic expansions~\cite{schult} (although not in the dynamical setting of this work). The idea of matching derivatives between the inner and outer regions actually appears to have its origin in the pioneering paper of Zeldovich and Frank-Kamenetzki~\cite{ZFK}, in which the velocity of a simple flame front is determined by analyzing a single-component equation like (\ref{eq:single}). In hindsight, it is somewhat surprising that the recent parallel literatures on two-species reaction-diffusion fronts and combustion waves have developed quite independently of each other, without any cross-references (at least, none to our knowledge). It is hoped, therefore, that this paper will initiate the ``diffusion'' of ideas between these two mature but related disciplines. \section*{Acknowledgments} The authors thank C. L\'eger and R. R. Rosales for useful discussions. This work was supported by an NSF infrastructure grant (MZB) and grants from the Harvard MRSEC DMR-980-9363 and the Army Research Office DAAG-55-97-1-0114 (HAS).
\section{Introduction} The nonrelativistic quark model (NQM) explains qualitatively many of the strong, electromagnetic and weak interaction properties of the nucleon and other octet and decuplet baryons in terms of three valence quarks whose dynamics is motivated by quantum chromodynamics (QCD), the gauge field theory of the strong interaction. Due to the spontaneous chiral symmetry breakdown ($\chi $SB) of QCD, the effective degrees of freedom at the scale $\Lambda _{QCD}$ are expected to be quarks and Goldstone bosons. Here we point out that a naive use of a constituent quark mass for all observables, and in axialvector quark-Goldstone boson couplings in particular, leads to disagreement with the proton spin fractions, because the non-spinflip contributions dominate over spinflip at low momentum. Understanding the internal proton structure is one of the goals of particle physics and, over the past dozen years or more, has led to extensive studies of the spin and flavor contents of the nucleon in terms of measured deep inelastic structure functions (DIS)~\cite{SMC,E143}. Since spin fractions ultimately must derive from and be consistent with polarized DIS proton structure functions that are integrated over Bjorken $x$, we have used such a general DIS formalism~\cite{HJW} to analyze recent chiral models that have succeded in reproducing the spin fractions of the proton. \par The effects of chiral dynamics on the spin fractions are discussed in Sects. II and III in the framework of chiral field theory applied to deep inelastic scattering. \section{Quark Spin Fractions from Chiral Dynamics} Chiral field theory involves the effective strong interactions commonly used in chiral perturbation theory ($\chi$PT~\cite{CPT}) and applies at scales from $\Lambda _{QCD}$ up to the chiral symmetry restoration scale $\sim\Lambda _{\chi }=4\pi f_{\pi }\sim 1.17$ GeV, where $f_{\pi }=0.093$ GeV is the pion decay constant. \par If the chiral symmetry breakdown is based on $SU(3)_L \times SU(3)_R$, then the effective interaction between quarks and the octet of Goldstone boson (GB) fields $\Phi_i$ involves the axial vector coupling \begin{eqnarray} {\cal L}_{int}=-{g_A\over 2f_\pi }\sum_{i=1}^{8}\bar q \partial_\mu \gamma ^\mu \gamma _5 \lambda _i \Phi_i q \label{lint} \end{eqnarray} that is well known from soft-pion physics. In Eq.~\ref{lint}, $g_A$ is the dimensionless axial vector-quark coupling constant that is taken to be 1 here. As a consequence, the polarization of quarks flips in chiral fluctuations, $q_{\uparrow,\downarrow,}\rightarrow q_{\downarrow,\uparrow}+GB$, into pseudoscalar mesons of the SU(3) flavor octet of Goldstone bosons, but for massive quarks the non-spinflip transitions from $\gamma_{\pm}\gamma_5 k^{\pm}$ that depend on the quark masses are not negligible. Let us also emphasize that, despite the nonperturbative nature of the chiral symmetry breakdown, the interaction between quarks and Goldstone bosons is small enough for a perturbative expansion to apply. \par Chiral field theory dissolves a dynamical or constituent quark into a current quark and a cloud of virtual Goldstone bosons. In this context, it was first shown in~\cite{EHQ} that chiral dynamics leads to a reduction of the proton spin fractions carried by the valence quarks and also to a reduction of the axial vector coupling constant $g_A^{(3)}$, based on one overall chiral strength parameter, $a$, that contains the scale (cf. Table 1). It is well known that relativistic effects reduce the axial charge further, and this causes problems for the spin fractions~\cite{WB}. In addition, the violation of the Gottfried sum rule~\cite{Go}, which signals an isospin asymmetric quark sea in the proton, i.e. $\bar u<\bar d$, became plausible. SU(3) symmetric chiral spin fraction models explain spin and sea quark observables of the proton, except for the ratio of axial charges $\Delta_3/\Delta_8=$5/3 and the weak axial vector coupling constant of the nucleon, $g_A^{(3)}={\cal F}+{\cal D}$. In~\cite{SMW,WSK} the effects of SU(3) breaking (adding $(1+\lambda _8\epsilon )$ in ${\cal L}_{int}$) were more systematically built into these chiral models and shown to lead to a remarkable further improvement of the spin and quark sea fractions in comparison with the data. \par It was first shown in ref.~\cite{WSK}, and subsequently confirmed~\cite{XS,OS} that the $\eta '$ meson, proposed in~\cite{CLi} mainly to decrease the antiquark fraction $\bar u/\bar d$ from the SU(3) symmetric value $3/4$ to $\sim 0.5$, gives an almost negligible contribution to the spin fractions of the nucleon, not only because of its large mass but also due to the small singlet chiral coupling constant. It is therefore often ignored, and this is consistent with the understanding that, due to the axial anomaly, the $\eta '$ meson is not a genuine Goldstone boson. Pions and kaons are well established Goldstone bosons. For a discussion of the controversial role of the $\eta$ meson as the hypercharge or octet Goldstone boson, see~\cite{KW}. \par Chiral fluctuations occur with probability densities $f(u_{\uparrow} \rightarrow \pi^+ + d_{\downarrow})$,... which, from Eq.~\ref{lint}, may be written as coefficients in the following chiral reactions: \begin{eqnarray}\nonumber u_{\uparrow} &\rightarrow& f_{u\rightarrow \pi ^{+}d}(x_{\pi }, \vec{k}_{\perp}) (\pi^+ + d_{\downarrow}) +f_{u\rightarrow \eta u} {1\over 6}(\eta +u_{\downarrow}) +f_{u\rightarrow \pi ^0 u} {1\over 2}(\pi^0 + u_{\downarrow}) \\\nonumber &&+f_{u\rightarrow K^{+}s} (K^+ + s_{\downarrow}),\\\nonumber d_{\uparrow} &\rightarrow& f_{d\rightarrow \pi ^{-} u} (\pi^- + u_{\downarrow}) +f_{d\rightarrow \eta d} {1\over 6}(\eta +d_{\downarrow}) +f_{d\rightarrow \pi ^0 d} {1\over 2}(\pi^0 + d_{\downarrow})\\\nonumber &&+f_{d\rightarrow K^0 s} (K^0 + s_{\downarrow}),\\ s_{\uparrow} &\rightarrow &f_{s\rightarrow \eta s} {2\over 3}(\eta +s_{\downarrow})+f_{s\rightarrow K^- u} (K^- + u_{\downarrow}) +f_{s\rightarrow \bar K^0 d} (\bar K^0 + d_{\downarrow}), \label{fluc} \end{eqnarray} and corresponding ones for the other initial quark helicity. The factors $1/2,1/6,1,...$ in Eq.~\ref{fluc} for $u\rightarrow \pi ^0 u, u\rightarrow \eta u, u\rightarrow \pi ^{+}d,...$, respectively, originate from the flavor dependence in Eq.~\ref{lint} and are denoted as $p_m$ for the Goldstone boson $m$ for brevity. After integrating over transverse momentum in the infinite momentum frame, the coefficients in Eq.~\ref{fluc} become the polarized ($-$ sign) and unpolarized ($+$ sign) chiral splitting functions, \begin{eqnarray} P^{\pm}_{GB/q}(x)=\int d^2 k_{\perp}f^{\pm}_{q\rightarrow q'GB} (x,\vec{k}_{\perp}). \label{splf} \end{eqnarray} The unpolarized splitting function $P^+$ determines the (spinflip plus non-spinflip) probability for finding a Goldstone boson of mass $m_{GB}$ carrying the longitudinal momentum fraction $x_{GB}$ of the parent quark's momentum and a recoil quark $q'$ with momentum fraction $1-x_{GB}$ for each fluctuation in Eq.~\ref{fluc}. \par In (non-renormalizable) chiral field theory with cutoff $\Lambda _{\chi} $ of ref.~\cite{EHQ}, the unpolarized chiral splitting function takes the form \begin{eqnarray} P^{+}_{q\rightarrow q'+GB}(x_{GB})={g^{2}_{A}\over f^{2}_{\pi }} {x_{GB}\over 32\pi ^2}(m_q+m_{q'})^2\int^{t_{min}}_{-\Lambda _{\chi }^2}dt {(m_q-m_{q'})^2-t\over (t-m^{2}_{GB})^2}\ , \label{EHQ} \end{eqnarray} where $t=k^2=-[(k_{\perp})^2+x_{GB}[(m'_{q})^{2}-(1-x_{GB})(m_{q})^2]]/ (1-x_{GB})$ is the square of the Goldstone boson four-momentum. The polarized splitting function is obtained using that it contains the difference of non-flip and helicity-flip probabilities. Since quarks are on their mass shell in the light front dynamics used here, the axialvector quark-Goldstone boson interaction is equivalent to the simpler $\gamma _5$ coupling. Except for an overall factor, the relevant unpolarized chiral transition probability is proportional to \begin{eqnarray} -{1\over 2}tr[(\gamma \cdot p+m_q)\gamma _5(\gamma \cdot p+m_q)\gamma _5] =2(pp'-m_q m'_q)=(m_q-m'_q)^2-k^2,\ \label{unpol} \end{eqnarray} where $2pp'=m'^{2}_{q}+m^{2}_{q}-k^2.$ Eq.~\ref{unpol} is the numerator in Eq.~\ref{EHQ} which can also be written as \begin{eqnarray} {1\over 1-x_{GB}}[(k_{\perp})^2+[m'_{q}-(1-x_{GB})m_q]^2], \label{nu} \end{eqnarray} and has the following physical interpretation. Recall that the axialvector quark-Goldstone boson coupling ${\gamma_{\mu }\gamma_5 k^{\mu}}$ in Eq.~\ref{lint} involves the spin raising and lowering operators $\sigma_1\pm i\sigma_2$ in a scalar product with the transverse momentum $\vec{k}_{\perp}$ of the recoil quark, which suggests that the $k^{2}_{\perp}$ term in $P^+$ of Eq.~\ref{nu} represents the helicity-flip probability of the chiral fluctuation, while the longitudinal and time components, ${\gamma_{\pm}\gamma_5 k^{\pm}}$, induce the non-spinflip probability, which depends on the quark masses. This can be seen from the helicity non-flip probability \begin{eqnarray} |\bar u'_{\uparrow}\gamma _5 u_{\uparrow}|^2=|\bar u'_{\downarrow}\gamma _5 u_{\downarrow}|^2\sim (m'_q-x' m_q)^2,\ x'=1-x_{GB}, \label{nofli} \end{eqnarray} using light cone spinors and suppressing the spinor normalizations. Thus Eq.~\ref{nofli} identifies the mass term in Eq.~\ref{nu} as the helicity non-flip chiral transition. Similarly, the helicity-flip probability is obtained from \begin{eqnarray} |\bar u'{\downarrow}\gamma _5 u_{\uparrow}|^2\sim (p'_{\perp})^2+x'^2 (p_{\perp})^2-x'p'_{\perp}\cdot p_{\perp} \label{fli} \end{eqnarray} which, in frames where $\vec{p}_{\perp}=0,$ reduces to $(\vec{k}_{\perp})^2,$ i.e. the net helicity flip probability generated by the chiral splitting process. In the nonrelativistic limit, where $|\vec{p}'_{\perp}|\ll m'_{q},$ $|\vec{p}_{\perp}|\ll m_q,$ clearly non-spinflip dominates over spinflip, while spinflip dominates at high momentum (it is not clear how ref.~\cite{EHQ} reached the opposite conclusion which, obviously, flies in the face of an extensive low-energy nuclear physics lore). \par The polarized splitting function $P^-$ therefore has the same quark mass dependence as $P^+$, but involves the helicity flip probability with the opposite sign, i.e. has \begin{eqnarray} {1\over 1-x_{GB}}[-(\vec{k}_{\perp})^2+\left(m'_{q}-(1-x_{GB})m_q\right)^2] \label{polnu} \end{eqnarray} in its numerator, which agrees with ref.\cite{SW}. Only for massless quarks there are no non-spinflip chiral transitions, so that $P^-=-P^+$ holds which is characteristic of pure spinflip chiral transitions. \par The splitting of quarks into a Goldstone boson and a recoil quark corresponds to a factorization of DIS structure functions that leads to a convolution of quark distributions with splitting functions. Thus, chiral fluctuations in lowest order of perturbation theory contribute convolution integrals \begin{eqnarray}\nonumber \sum_{q,m} p_m P^{+}_{u_{\uparrow}\rightarrow q_{\downarrow} m}\otimes u^{0}_{\uparrow}+\sum_{q,m} p_m P^{+}_{d_{\uparrow} \rightarrow q_{\downarrow} m} \otimes d^0_{\uparrow},\\ \sum_{q,m} p_m P^{+}_{u_{\downarrow}\rightarrow +q_{\uparrow} m} \otimes u^0_{\downarrow}\ +\sum_{q,m} p_m P^{+}_{d_{\downarrow}\rightarrow q_{\uparrow} m} \otimes d^0_{\downarrow} \label{rqp} \end{eqnarray} to the quark distributions $q_{\downarrow,\uparrow}$, respectively. Chiral fluctuations also cause a reduction of the valence quark probabilities \begin{eqnarray} (1-P_q)q^{0}_{\uparrow,\downarrow}, \label{valq} \end{eqnarray} where $P_q$ are the total fluctuation probabilities. \section{Spin Distribution Results} In chiral dynamics, antiquarks originate only from the Goldstone bosons via their standard quark-antiquark composition. Therefore, {\em antiquarks are unpolarized}. Small antiquark polarizations are consistent with the most recent SMC data~\cite{SMC}, so that we expect only small corrections if we use $\bar u_\uparrow = \bar u_\downarrow$ in the spin fractions $\Delta u = u_\uparrow -u_\downarrow +\bar u_\uparrow -\bar u_\downarrow$, etc., i.e. $\Delta s=\Delta s_{sea}$, $\Delta \bar u=\Delta \bar d=\Delta \bar s =0$. \par Let us now return to the polarized quark distributions and their lowest moments, the spin fractions. \par Upon generalizing the chiral spin fraction formalism of~\cite{WSK,XS,CLi} to the polarized quark distributions, the probabilities displayed in Eq.~\ref{fluc},\ref{rqp},\ref{valq} yield \begin{eqnarray} q_{\uparrow}(x)=(1-P_{q})q^{0}_{\uparrow}(x)+\sum_{m,q'} p_m P^{+}_{q'\rightarrow q m}\otimes q'^{0}_{\downarrow}+... \label{oldq} \end{eqnarray} which obviously are based on pure spinflip chiral transitions. The corresponding result holds for the other quark helicity. The ellipses in Eq.~\ref{oldq} denote double convolution terms with $q^{0}_{\uparrow}$ from a Goldstone boson $m$ that cancel in $q_{\uparrow}-q_{\downarrow}$. The opposite quark helicity on the rhs of Eq.~\ref{oldq} implies the {\em negative} sign of all chiral contributions to the spin distributions \begin{eqnarray} \Delta q(x) = (1-P_{q})\Delta q^0(x) -\sum_{m,q'} p_m P^{+}_{q'\rightarrow q m} \otimes \Delta q'^0. \label{delp} \end{eqnarray} This result is common to all recent successful chiral models of spin fractions. Let us emphasize that the general reduction of the valence quark spin fractions $\Delta q^{0}$ by chiral fluctuations in lowest order in Eq.~\ref{delp} is the crucial property responsible for the remarkable success of chiral field theory for the proton spin fractions. Eq.~\ref{delp} can be compared to the corresponding one from DIS involving the polarized splitting functions \begin{equation} P^{-}_{q\rightarrow q' GB}(y)=\int d^2 k_{\perp} f^{-}_{q\rightarrow q' GB}(y, \vec{k}_{\perp}), \label{polspl} \end{equation} which has the same form as Eq.~\ref{delp}, \begin{eqnarray} \Delta q(x) = (1-P_q)\Delta q^0(x)+\sum_{m,q'}p_m P^{-}_{q'\rightarrow q m} \otimes \Delta q'^0 \label{delq} \end{eqnarray} except for the replacement of $-P^+$ by the corresponding polarized splitting function $P^-$. A comparison with Eqs.~\ref{nu},\ref{polnu} shows that this approximation, $P^-\approx -P^+$, corresponds to neglecting the non-spinflip probability, which is valid only if the quark mass term in the numerator (i.e. Eq.~\ref{polnu} of the splitting functions is negligible compared to the tranverse momentum scale. This is not the case for constituent quarks~\cite{HJW,SW}. Thus, when these $\Delta q(x)$ of Eq.~\ref{delq} are integrated over Bjorken $x$, the lowest moments reproduce precisely the structure of the results for the spin fractions~\cite{WB,WSK} (cf. Table 1). \begin{table}[tbh] \caption{\label{tab:1} Quark Spin Observables of the Proton (from ref.~\protect\cite{WSK}), $a=$chiral strength, $\zeta =$relative singlet to octet coupling and $\epsilon =SU_3$ breaking parameter in $\lambda _8$ direction.} \begin{tabular}{|c|c|c|c|c|} \hline & Data & NQM & $a=$0.12 & $a=$0.12\\ & E143\cite{E143} & & no $\eta '$ & with $\eta '$ \\ & at 3 GeV$^2$ & & & $\zeta=$-0.3 \\ & SMC\cite{SMC} & & $\epsilon =$ 0.2 & $\epsilon =$ 0.2 \\ & at 5 GeV$^2$ & & & \\ \hline $\Delta u$ & 0.84$\pm$ 0.05 & 4/3 & 0.83 & 0.81 \\ & 0.82$\pm$0.02 & & & \\ $\Delta d$ & -0.43$\pm$0.05 & -1/3& -0.39& -0.39 \\ & -0.43$\pm$0.02 & & & \\ $\Delta s$ & -0.08$\pm$0.05 & 0 & -0.07 & -0.07\\ & -0.10$\pm$0.02 & & & \\ $\Delta \Sigma $ & 0.30$\pm$0.06 & 1 & 0.36 & 0.35 \\ & 0.29$\pm$0.06 & & & \\ $\Delta_3/\Delta_8$ & 2.09$\pm$0.13 & 5/3 & 2.12 & 2.13 \\ $g^{(3)}_A$ & 1.2573$\pm$0.0028\cite{PDG}& 5/3 & 1.22 & 1.21 \\ ${\cal F}/{\cal D}$ & 0.575$\pm$0.016 & 2/3 & 0.58 & 0.58 \\ $I_G$ & 0.235$\pm$0.026 & 1/3 & 0.27 & 0.25 \\ \hline \end{tabular} \end{table} \section{Conclusions} \par The detailed comparison in Sect. IV shows unambiguously that the success of several recent chiral models~\cite{EHQ,WB,SMW,WSK,XS,OS,CLi} for the spin fractions $\Delta q$ can be attributed to pure spinflip quark-Goldstone boson couplings that do not account for quark mass terms from the non-spinflip chiral transitions in the lowest moments of the splitting functions in standard chiral field theory. Whenever {\em constituent quark masses} are used in Eqs.~\ref{nu},\ref{polnu}, then the positive contributions in $P^-$ corresponding to the non-spinflip probability represented by the quark mass term substantially reduce the chiral subtractions so that {\em no agreement with the proton spin data can be achieved}~\cite{HJW,SW}. These results were obtained with quite different initial valence quark distributions and both show that the proton spin observable $\Delta \Sigma $ stays above the value 1/2. This disagreement can be interpreted so that the constituent quark model in the framework of standard chiral field theory, which is often called chiral quark model~\cite{GM}, is ruled out by the proton spin data. \par The success of the chiral spin fraction models suggests that axialvector quark-Goldstone boson couplings are consistent with current quarks, but not constituent quarks. This usage conforms with the chiral field theory practiced in chiral perturbation theory~\cite{CPT}. It is interesting to note that proton spin data are successfully described by models based on an instanton fluid in the vacuum, where the instanton-quark interaction is also pure spinflip~\cite{DK}. The proton spin data does not challenge the vast nuclear theory based on pion exchange directly -- as the pion exchange potential has been successfully tested through its pion exchange current predictions -- but it suggests understanding better its derivation from QCD concepts such as current quark masses and quark condensates, and not merely from constituent quark models.
\section{Introduction} Clusters of galaxies are excellent cosmological probes. Their size suggests that their constituents provide a fair sample of the universe. Their structure and hydrodynamic state provide information on their formation and evolution, and thus upon models of structure formation. Measurements of the abundance of clusters of a given mass allows constraint of the amplitude of mass fluctuations in the universe; measurements of abundance evolution can be used to constrain the mass density $\Omega_{\rm m}$. Many of these approaches depend upon some knowledge of the mass, or mass distribution, of the cluster. Most techniques for measuring cluster masses are based upon some equilibrium assumption which relates the cluster mass to an observable such as the temperature of the intracluster plasma or the velocity dispersion of cluster galaxies. Recently, however, it has become feasible to measure the surface density distribution of a cluster through observations of weak gravitational lensing of the background galaxy field by the cluster. An attractive quality of this technique is that no assumptions about the dynamical or thermodynamical state of the cluster components need be made. In other words, weak lensing analyses probe the mass distribution directly. However, analyses of the mass distribution of a cluster drawn from weak lensing observations are not without problems (for a recent review, see Mellier~\cite{Mellier}). Many of these relate to details of the procedure adopted to go from the observed ellipticity distribution to the mass, or from instrumental effects. We will not discuss these in this {\em Letter}. We are interested here in the degree to which weak lensing mass estimates of clusters are affected by lensing from material outside but nearby the cluster. This is a source of systematic error which is not well understood (though it has been alluded to in earlier work, e.g.~Miralda-Escude~\cite{ME}; Wambsganss, Cen \& Ostriker~\cite{WamCenOst}). Since clusters form in overdense regions, the volume surrounding a cluster is likely to contain infalling overdense matter (Gunn \& Gott~\cite{GunGot}). This infalling matter could add to an observed lensing signal and result in an overestimate of the cluster mass. It is possible that this bias could be quite severe. In modern hierarchical models of structure formation, such as the Cold Dark Matter (CDM) model, numerical simulations imply that clusters form primarily at the intersections of filaments in a web of cosmic structure, accreting additional diffuse mass and smaller collapsed objects that drain along these filaments. It is thus reasonable to expect a beaded filamentary structure surrounding most clusters of galaxies. Tentative observational evidence of filamentary structure near clusters has been reported recently (Kull \& Boehringer~\cite{KulBoe}, Kaiser et al.~\cite{MS0302}). A filament lying near the line-of-sight will also lens the background galaxies, and therefore contribute spuriously to the lensing signal. If the observed lensing signal were attributed solely to the cluster, the inferred cluster mass could be much larger than its actual mass. In this {\em Letter}, we use mock clusters from numerical simulations to explore the significance of the systematic mass overestimation induced by the additional lensing signal of both diffuse and filamentary material near the cluster. We find that this effect can be quite significant and must be considered when evaluating the results of lensing mass reconstruction techniques. In the next section, we describe the numerical data and our procedure for evaluating the errors introduced into cluster mass estimates by nearby material. Section 3 describes the results of our analyses. We discuss these results and outline plans for future study at the end. \section{Method} \begin{table} \begin{center} \begin{tabular}{lcccc} & & \multicolumn{3}{c}{Number} \\ & & 0 & 2 & 4 \\ $R_{\rm sphere}$ & $(h^{-1}{\rm Mpc})$ & 12.9 & 15.1 & 14.9 \\ $r_{200}$ & $(h^{-1}{\rm Mpc})$ & 3.14 & 2.76 & 2.60 \\ $M_{200}$ & $(h^{-1} 10^{15} M_{\sun})$ & 2.16 & 1.47 & 1.23 \\ $M_{>70}$ & $(h^{-1} 10^{15} M_{\sun})$ & 2.41 & 3.11 & 2.35 \\ $M_{>10}$ & $(h^{-1} 10^{15} M_{\sun})$ & 3.24 & 4.24 & 3.37 \\ $M_{\rm tot}$ & $(h^{-1} 10^{15} M_{\sun})$ & 3.89 & 5.09 & 4.15 \\ \end{tabular} \end{center} \caption{\tfa Parameters for the 3 simulated clusters discussed in this {\em Letter}. $R_{\rm sphere}$ is the size of the sphere, centered on the cluster, which we consider in this work, $r_{200}$ is the 3D radius defined in Eq.~\protect\ref{eq:r200def} and $M_{200}$ the mass enclosed. $M_{>70}$ and $M_{>10}$ are masses cutting out particles above thresholds of 70 and 10 times local density respectively (see text). $M_{\rm tot}$ is the total mass in the sphere.} \label{tab:numbers} \end{table} Weak lensing mass reconstruction techniques produce a map of shear or convergence. These are integrals of the mass along the line-of-sight times a projection kernel. This kernel is quite wide in the redshift direction, scaling as $D_LD_{LS}/D_S$ where $D_L$ is the distance from the observer to the lens, $D_S$ from the observer to the source and $D_{LS}$ from the lens to the source\footnote{For a distribution of source distances, one takes an appropriate integral of this expression.} (Mellier~\cite{Mellier}). Under the assumption that the cluster is the most massive object along the line-of-sight and is well localized in space (the thin-lens approximation), the convergence map is proportional to the projected surface density map of the lensing cluster itself. Any additional mass located near the cluster and along the line-of-sight will also contribute to the lensing signal. Since the kernel is such a slowly varying function of distance, material even large distances from the cluster will contribute within the thin lens approximation. For a source at $z=1$ and a cluster at $z\sim0.5$ the kernel changes by only 1\% within $\pm40\,h^{-1}\,$Mpc of the cluster in a universe with $\Omega_{\rm m}=0.3=1-\Omega_\Lambda$, with similar results in other cosmologies. As a result, weak lensing observations will probe the projected density of a cluster {\em plus\/} all of the material in its vicinity. Note that this ``nearby'' material is essentially ``at'' the redshift of the cluster for the purposes of lensing, and so cannot be distinguished by using extra information such as source redshifts. To study the effect of the surrounding mass upon the projected mass inferred from lensing observations of the simulated clusters, we have examined the mass distribution around several clusters of galaxies extracted from a large cosmological simulation. The simulated clusters were taken from the X-Ray Cluster Data Archive of the Laboratory for Computational Astrophysics of the National Center for Supercomputing Applications (NCSA), and the Missouri Astrophysics Research Group of the University of Missouri. To produce these clusters, a particle-mesh N-body simulation incorporating adaptive mesh refinement was performed in a volume $256\,h^{-1}$Mpc on a side. Regions where clusters formed were identified; for each cluster, the simulation was then re-run (including a baryonic fluid) with finer resolution grids centered upon the cluster of interest. In the adaptive mesh refinement technique, the mesh resolution dynamically improves as needed in high-density regions. The ``final'' mesh scale at the highest resolution was $15.6\,h^{-1}$ kpc, allowing good resolution of the filamentary structure around the cluster. Inside the cluster, the characteristic separation between the smallest--mass particles, given by $d\,=\,\left(4\pi\ r_{200}^3/3N\right)^{1/3}$ with $N$ the number of particles inside the region, was approximately 86 kpc for all three clusters examined here. The code is described in detail in Norman \& Bryan (\cite{TheCode}). The clusters used here were extracted at $z=0$ from simulations of a $\Lambda$CDM model, with parameters $\Omega_{\rm m}=0.3$, $\Omega_{\rm B}=0.026$, $\Omega_{\Lambda}=0.7$, $h=0.7$, and $\sigma_8=0.928$. In this {\em Letter\/} we observe these clusters as if they were at $z=0.5$. In future work we plan to investigate the dependence of these results on cosmology and on cluster redshift. Since the number density of rich clusters is approximately $\phi_{*}\sim 10^{-5}$ Mpc$^{-3}$, the typical separation between them is $\phi_{*}^{-1/3}\sim 40\,h^{-1}$Mpc. This is a characteristic scale for filaments: volumes containing a cluster and with one-dimensional extent $\sim 40\,h^{-1}$Mpc should also contain much of the nearby filamentary structure. Three such volumes, containing a rich cluster (Clusters 0, 2 and 4) as well as satellites and filaments, were extracted from the archive. We selected clusters that did not appear to be mergers or have a large secondary mass concentration nearby. Such systems might be excluded observationally from studying, for instance, the galaxy line-of-sight velocity distribution. For each volume, the ``center'' of our cluster was determined using a maximum-density algorithm. As the extracted volumes were not spherical, it was possible that some lines of sight could contain more mass than others simply by geometry. To avoid such biases we restrict our analysis to particles that lay within the largest sphere, centered on the cluster, which was contained entirely within the extracted volume. The radii of these spheres, $R_{\rm sphere}$, are listed in Table~\ref{tab:numbers}, along with other properties of the clusters. Note that these radii are large compared to the projected values of $r_{200}$ obtained for each cluster; thus no significant radial surface density gradient is introduced by a decreasing chord length through the sphere with radius. It is also important to note that since our volumes are by necessity limited, our results should be interpreted as a {\em lower limit\/} to the size of the effect; the magnitude of the lensing kernel is still significant at the edge of our spherical volume. Each of the clusters we examined was surrounded by a large amount of mass. Most of this material appeared by eye to be collapsed into ``beads'' along a filamentary structure, although a small number of clumps could be found outside the filaments. We show a projection of a fraction of the points from the simulation of cluster 4 in Fig.~\ref{fig:cl004}. The filamentary structure and satellites are easily evident. Note that this filamentary structure extends well beyond our radius $R_{\rm sphere}$. No single projection can show the full 3D nature of the structure, in which the filamentarity is even more apparent. Since much of this mass is at low density it is unlikely it would be a site for galaxy formation or otherwise emit light. Thus this structure would not be easy to constrain by observations of redshifts near the cluster. \begin{figure}[t] \centerline{\epsfxsize=3.3in \epsfbox{f1.eps}} \caption{\tfa Cluster 4, with a fraction of the particles projected onto the $x-z$ plane. Note the filamentary structure, with clumps beading up in the filaments. The dashed circle marks the sphere of radius $R_{\rm sphere}$ to within which we have restricted our analysis (see text).} \label{fig:cl004} \end{figure} We observed each of the three selected clusters from $10,000$ randomly chosen viewing angles. For each cluster and viewing angle, the projected surface density map was constructed and used to estimate $r_{200}$, the radius within which the mean interior density contrast is 200. In three dimensions, this radius is defined in terms of the enclosed mass by \begin{equation} M\left(<r_{200}\right)=200\times \left({4\pi\over3}\right) \Omega_{\rm m}\, \rho_{\rm crit}\, r_{200}^3. \label{eq:r200def} \end{equation} The projected estimate of $r_{200}$ was extracted from the surface density map by considering the radius of the circle, centered on the cluster, which contained the amount of mass given by Eq.~(\ref{eq:r200def}) above, i.e. \begin{equation} \int_0^{2\pi}{\rm d}\theta \int_0^{r_{200}} R\, {\rm d}R \,\ \Sigma\left(R,\theta\right) = M\left(<r_{200}\right) \end{equation} with $\Sigma\left(R,\theta\right)$ the surface density on the map in terms of a two-dimensional radius $R$. This radius was compared to the cluster's true $r_{200}$, extracted from the three--dimensional mass distribution. The ratio of the projected mass to true mass is given simply by the cube of the ratio of the estimated value of $r_{200}$ to the true value. For each cluster, a value of this ratio was obtained for each viewing angle. \section{Results} Our main result is displayed in Fig.~\ref{fig:mass}, where we show the distribution of projected vs.~``true'' cluster mass in each of the three simulated clusters. We have checked that the features in the histogram do not come from shot noise due to discrete particles in the simulation. However, the ``spikiness'' {\em is\/} due to a discrete number of objects in the neighborhood of the cluster. A small lump of matter near the cluster will project entirely within $r_{200}$ for a fraction of the lines of sight. For any such line of sight, the effect on the projected value of $r_{200}$ is identical. \begin{figure}[t] \centerline{\epsfysize=9cm \epsfbox{f2.eps}} \caption{\tfa The ratio of projected to actual mass within $r_{200}$ for our 3 simulated clusters. In all cases nearby mass (primarily in filaments) has biased the projected mass distribution to higher values. The typical bias is a few tens of percent (b-c), with less of an effect on massive clusters (a). The vertical dashed line marks the mean of the distribution.} \label{fig:mass} \end{figure} We expect the ratio $M_{200}/M_{\rm true}$ to be greater than unity since only additional mass can be included in the projection. The size of the smallest offset from unity for clusters 0 and 4 is approximately twice what would be expected for material uniformly distributed at the mean density. This suggests that matter near the cluster is itself clustered and at higher than mean density. The width of the histogram in Fig.~\ref{fig:mass}, as a fraction of the true mass, depends on the true mass of the cluster. Though we have only a few clusters in this study, it appears that the mass in nearby material is not proportional to the cluster mass, thus the relative effect of this material is smaller the larger the cluster. The total mass in the sphere, $M_{\rm tot}$, is also listed in Table~\ref{tab:numbers} for reference. The signal shown in Fig.~\ref{fig:mass} comes from (primarily filamentary) material outside the cluster and is {\em not\/} the well known projection effect arising from cluster asphericity. To verify this, we repeated the procedure described above for a subset of particles aimed at selecting the cluster alone. This was done by first selecting out particles with a local density contrast of greater than 70 (chosen because density profiles near $r^{-2}$ reach a local density contrast near 70 at a mean interior density contrast of 200); a small sphere containing the cluster but little nearby material was then cut out of this subset. The histogram produced by viewing the clearly prolate cluster at a large number of randomly chosen viewing angles produced a much narrower distribution, with a maximum offset of less than $10\%$ in the mass ratio and a mean offset of approximately half that value. While it is beyond the scope of this {\em Letter\/} to perform a detailed modelling of any observational weak lensing strategy, we show in Fig.~\ref{fig:zeta} two sample profiles obtained from aperture densitometry on our noiseless projected mass maps. Specifically, for Cluster 4, we show the profile along the lines of sight giving the largest and smallest $r_{200}$, for comparison. The $\zeta$ statistic is the mean value of the convergence, $\kappa$, within a disk of radius $r_1$ minus the mean value within an annulus $r_1\le r\le r_2$ (Fahlman et al.~\cite{Fahlman}, Kaiser~\cite{K95}). Here we calculate $\zeta$ directly from the projected mass, though observationally it would be computed from the tangential shear. We have taken $r_2=800''$. Such a large radius is not (currently) achievable observationally, but it minimizes the impact of objects nearby the cluster and provides a lower limit on the size of the projection effect. We have explicitly checked that reducing the radius to half this value does not change our result. \begin{figure}[t] \centerline{\epsfxsize=3.3in \epsfbox{f3.eps}} \caption{\tfa The profile $\zeta(r_1,r_2)$ (see text) for cluster 4 along two lines of sight with extremal values of $r_{200}$. We have used $r_2=800''$ in making this figure. Such a large radius is not currently achievable observationally, but it minimizes the impact of objects nearby the cluster and provides a lower limit on the effect. Reducing this radius by a factor of two does not change the result. Note that both profiles appear smooth and well behaved, even though they differ in mass by a factor of 1.7.} \label{fig:zeta} \end{figure} In calculating the convergence $\kappa$, the cluster was again assumed to be at a redshift of $0.5$, with the lensed sources at a redshift of $1.0$. In any real observation, of course, the lensed sources will span a range of redshifts. For material very close to the cluster, such as here, this will not affect our conclusions, and the error introduced by incorrectly modelling the redshift distribution of the background sources is not the subject of this work. In addition to being easy to estimate, the $\zeta$ statistic is robust and minimizes contamination by foreground mass (Mellier~\cite{Mellier}) because it is insensitive to the sheet mass degeneracy. This does not, however, remove the sensitivity to {\em clustered\/} material, as can be seen in Fig.~\ref{fig:zeta}. The mass which would be inferred from Fig.~\ref{fig:zeta} along two lines of sight differ by a factor of $1.7$. While the distribution of the projection effect varies from cluster to cluster, it seems clear that positive biases in projected mass of $20\%$ are typical, with biases above $30\%$ not uncommon. Furthermore, we emphasize again that these estimates are in fact {\em lower limits\/}; some lines of sight through the untruncated volume produced overestimates as large as $80\%$ or more. While appropriate modeling of a mean density profile outside $r_{200}$ (drawn perhaps from simulations such as these) can be used to produce an unbiased estimator, the width of these histograms implies a large amount of scatter around such an estimator of the true (unprojected) mass. It is clear that this effect can be quite significant and must be taken into account when attempting to understand the results of mass reconstruction analyses. \section{Conclusions} Clusters of galaxies are part of the large-scale structure of the universe and observations of them should be considered within this context. The filamentary structure near a cluster can contain a reasonable fraction of the mass of a cluster in tenuous material. Should a filament lie close to the line-of-sight to a cluster it will contribute to the weak lensing signal and positively bias the projected mass. We have shown that such a bias could easily be 30\% (see Fig.~\ref{fig:mass}). Weak lensing remains one of the best methods for determining the mass of clusters of galaxies. However methods which obtain the mass from estimates of the projected surface density must consider the effect outlined in this {\em Letter}. This is clearly of particular significance for attempts to determine the mass function of clusters directly through surveys of weak lensing-determined masses. We have not attempted to address the detailed question of how much this filamentary signal would affect a {\em particular\/} reconstruction algorithm; the answer is no doubt algorithm dependent. We hope to return to this issue in future work, as well as to consider the effect of cosmological model and evolution with cluster redshift. The authors would like to thank Greg Bryan and Greg Daues for assistance in understanding the archive data, and Gordon Squires and Albert Stebbins for useful conversations on lensing. This research was supported by the NSF.
\section{Introduction} The charge coupled device (CCD) is a standard X-ray detector due to its high X-ray detection efficiency, moderate X-ray resolving power, and high spatial resolving power. The Solid-state Imaging Spectrometer, SIS, onboard ASCA is the first CCD camera used as a photon counting detector and equipped on board the satellite (\cite{Tanaka}). Following SIS, many satellites such as HETE (\cite{Ricker}), Chandra (\cite{ACIS}), XMM (\cite{XMM}), and Astro-E (\cite{Hayashida}) carry a X-ray CCD camera on their focal planes. \section{MAXI} The International Space Station (ISS) will be placed in a nearly circular, high inclination (51.6$^\circ$), low Earth orbit having a 96 minute orbital period with respect to a point in the sky. ISS will rotate synchronously with its orbit so that one side will always point toward the center of the Earth and the opposite side will permanently view the sky. A payload is attached to the main structure of the JEM which rotates and has unpredictable disturbances. Therefore, pointed observations are very difficult on the JEM. On the other hand, synchronous rotation with orbital revolution provides access to the entire sky in one orbit without a moving mechanism. Considering these characteristics, we conclude that a monitoring mission or survey of a large field of the sky is suitable, and can produce significant scientific results. The schematic view of MAXI is shown in figure~\ref{maxi}. MAXI can scan almost the entire sky with a precision of 1$^\circ$. MAXI carries two kinds of X-ray detectors: a one-dimensional position-sensitive proportional counter (GSC) and an X-ray CCD camera (SSC). Combining these two cameras, MAXI can monitor X-ray sources with an energy band of $0.5-30$ keV. The total weight of MAXI is about 500 kg. Simulations of the data expected from MAXI have been performed in (\cite{Rubin}). A detailed description of MAXI can be found in (\cite{Matsuoka}, \cite{Matsuoka2}). \section{SSC} The SSC is an X-ray CCD camera system. The SSC consists of two X-ray CCD cameras, each comprising 16 CCD chips. The block diagram of the SSC camera is shown in figure~\ref{block-diagram}. The SSC consists of three parts: two CCD cameras, analogue electronics (SSCE), and a digital processing unit (DP). Detailed specifications of the SSC are shown in table~\ref{table:spec}. The CCD is fabricated by Hamamatsu Photonics K.K. (HPK). The CCD chip is three-side buttable with full-frame transfer and has 1024 $\times$ 1024 pixels of 24$\mu$m$\times$24$\mu$m size with two phase gate structures. The CCD chip is covered by $\sim$2000\AA\ Al to block optical light. The CCD is operated at $-60^\circ$C, which is achieved by using a passive cooling system and a Peltier cooling system (TEC). TEC is supported with glasses to hold out the launch shock. The SSCE is developed by Meisei Electronics. There are several CCD signal processing techniques (\cite{CCD} and references therein). To measure the voltage of each charge packet, we need a reference voltage between the floating level and the signal level. The correlated double sampling technique is widely used for this purpose. In practice, it is advantageous to integrate or take the sum of the signals rather than merely spot sample floating and signal levels. Thus, a delay-line circuit is used in SIS/ASCA and an integrated circuit is introduced for SXC/HETE, ACIS/Chandra and XIS/Astro-E. We plan to develop all these circuits for the SSC and will select the one that possesses the lowest readout noise. Since the data rate of CCD is fairly high, an onboard data reduction system is important. We developed an efficient reduction system based on our experiences with SIS/ASCA and XIS/Astro-E. There are three parts in DP: the control unit, the event handling unit (EHU), and the telemetry unit. Two CPU boards (RAD 6000) on the VME bus will be used for EHU and another CPU board will be used for the control unit, the telemetry unit, and GSC data processing. There are two interfaces between MAXI and JEM Exposed Facility (EF): medium-speed interface (10Base-T ethernet) and low-speed interface (MIL1553B). All CCD data will be downlinked through the ethernet whereas part of health and status (HS) data will be transferred through MIL1553B. Based on the SIS/ASCA, we have learnt much about radiation damage on the CCD (\cite{ayamashi}). One serious problem is the increase in dark current and its non-uniformity. To minimize the effects of radiation damage on the CCD, we allocate a dark level buffer for every pixel. The dark level for each pixel is updated for every frame based on the pulse height of pixel of interest. For the recovery of the radiation damage, we use an annealing process. However, we think that the radiation damage would be small because the lifetime of MAXI is two years (might be extended) and the orbit is lower than other missions ($\simeq$ 400km). Since the SSC is a one-dimensional X-ray camera, we use the spatial resolving power of the CCD only for the horizontal axis. Thus, we operate CCD in the parallel summing mode (same as the fast mode for SIS/ASCA). The vertical axis of the CCD corresponds to the time. The binning number can be changed as $2^{\rm n}$ (n=2$\sim$8). 16 CCD chips in one camera are read cyclically using a multiplexer. \section{SSC Engineering Model} The engineering model (EM) of the CCD chip has been completed and tested at the Osaka University X-ray CCD laboratory. EM of CCD is shown in figure~\ref{ccd-chip} where CCD is fixed on the Al plate. Two cables connected to the CCD are used for the Peltier cooler. There are three types of CCD produced for EM: a standard chip (standard), a deep depletion type I (deep-I), and a deep depletion type II (deep-II). There is a difference both in the depletion layer and in the dark current among these three types of CCDs. The details of these three chips can be referred to in Miyaguchi et al. (1999, \cite{Miyaguchi}). EM of the SSCE has been fabricated by MEISEI on the VME board. The function test of the EM SSCE is underway. \section{X-ray Responsivity} \subsection{Experimental Setup and Analysis} We evaluated the X-ray responsivity of deep-I EM CCD. We cooled the CCD chip down to $-100^\circ$ with a He cryogenic system in the vacuum chamber. We used the C4880 CCD camera system, which is the X-ray CCD data acquisition system manufactured by HPK. Exposure time was set at 5 seconds. CCD frame files were transfered to a workstation through the ethernet with FITS format after they were acquired by C4880. HK information was collected with a workstation and stored in a hard disk. Dark current image was constructed with several CCD frame files using the same algorithm as that of XIS/Astro-E (\cite{Hayashida}). Before the X-ray event extraction, the dark current image was subtracted from each frame. \subsection{Results} Figure~\ref{fe-spec} shows the energy spectrum of X-rays from $^{55}$Fe for single-pixel events. The split threshold is $\simeq$ 70 eV. Mn K$\alpha$ and K$\beta$ lines are clearly separated. The energy resolution of Mn K$\alpha$ has a full-width at half maximum of $\simeq$ 182 eV. Readout noise is $\simeq$ 11 e$^-$rms. Since the energy resolution of HPK CCD is $\sim$40\% less than that obtained by CCDs fabricated by the MIT Lincoln Laboratory (e.g. \cite{Hayashida}). HPK plans to improve the CCD to achieve performance comparable to those of other X-ray CCD devices. \section{Conclusion} MAXI is an X-ray all-sky monitor on the International Space Station and is due for flight in 2003. It is designed to scan almost the entire sky with a precision of 1$^\circ$ and with an X-ray energy range of 0.5$-$30~keV in the course of one station orbit. We have developed the engineering model of the analogue electronics and the CCD chips for the X-ray CCD camera, SSC. We evaluated the X-ray responsivity of the EM CCD chip. The energy resolution of Mn K$\alpha$ X-rays has a full-width at a half maximum of 182 eV. Based on the EM results, we will improve the performance of CCD and its electronics.
\section{Introduction} While supersymmetry (SUSY) is generally agreed to be integral to any theory incorporating both gravity and gauge forces, techniques for investigating nonperturbative effects such as chiral symmetry breaking are still in their early stages. A small number of authors have employed Dyson Schwinger Equations (DSEs) to analyse various SUSY theories \cite{Pski,clark,ster,sham} and small inroads into numerical solutions \cite{me} of the SUSY DSE (SDSE) in Supersymmetric Quantum Electrodynamics (SQED) in 2+1 dimensions (SQED$_3$) have been made. Analyses of SUSY theories generally use the rainbow approximation to truncate the DSEs at a manageable level. One exception is Clark and Love who use the superfield formalism and derive a differential $U(1)$ gauge Ward Identity for the superfields. They find that the effective mass contains a prefactor which vanishes in Feynman gauge and conclude that there can be no spontaneous mass generation in SQED, even beyond the rainbow approximation. However the superfield approach suffers the disadvantage that each DSE contains an infinite number of terms. This is dealt with by truncating diagrams containing seagull and higher order $n$-point vertices. The work of Clark and Love has been criticized by Kaiser and Selipsky on two grounds \cite{kaiser}. Firstly they argue that the truncation of seagull diagrams is too severe as it ignores contributions even at the one-loop level. Secondly they point out that infinities arising from infrared divergences which plague the superfield formalism can counter the vanishing prefactor and allow spontaneous mass generation. These criticisms highlight some of the dangers of attempting to extract phenomenological consequences of supersymmetric DSEs by working solely with the superfield formalism. In fact, analyses in the literature \cite{sham,me} have generally found the component formalism to be the most efficient way to proceed. Koopmans and Steringa \cite{ster}, using the component formalism, also sought to be consistent with the differential $U(1)$ gauge WI in their analysis of SQED3 with two-component fermions. To this end they multiplied the bare vertices by $A(q^2)$ where the electron propagator is given by $S^{-1}(q)=i(\gamma \cdot q A(q^2) + B(q^2))$. This approach is questionable as it implicitly approximates the functions $A(p^2)$ and $B(p^2)$ as being flat. While this approximation is reasonable over most of the momentum range, it is not valid in the low momentum limit where the dynamics are largely determined. Attempts to go beyond the rainbow approximation in non-SUSY theories began with the Ball and Chiu \cite{ball} vertex ans\"atze for QED and QCD. These are the minimal vertices which ``solve'' the Ward Takahashi Identities (WTIs) while avoiding kinematic singularities. Ball and Chiu also gave the general form of the possible ``transverse'' pieces which may be added. Since then several authors have sought to construct ans\"atze which improve on the minimal Ball-Chiu vertex \cite{CP91,paulus}. That analogous progress has not been made in SQED using the component formalism is not suprising. Not only must the gauge particle vertices be dressed but the gaugino vertices also. Indeed substituting the minimal Ball and Chiu vertex for photon interactions in SQED$_3$ while leaving the other vertices bare exacerbates the SDSE's gauge violating properties \cite{me}. The problem of going beyond the rainbow approximation in SUSY theories is the problem of finding the gaugino vertices corresponding to the improved photon vertex. Gaugino vertices are not constrained by the WTI since the gaugino is invariant to gauge transformations. However they are related to the gauge particle vertices by SUSY Ward Identities (SWIs). It is the purpose of this paper to derive and solve the SWIs for SQED and obtain the most general form of the three-point vertex functions consistent with both SUSY and $U(1)$ gauge Ward identities. Sec.\ref{twopoint} gives the SWIs between the various two-point functions of SQED and their solution which is unique once the electron propagator is known. Sec.\ref{auxiliary} shows how to treat proper functions of auxiliary fields. Sec.\ref{SWI} gives the SWIs constraining the three-point proper functions and finds that the rainbow approximation violates SUSY. The most general form of the vertices consistent with these identities is presented in Sec.\ref{threepoint} and proven to be so in Appendix \ref{proof}. \setcounter{equation}{0} \section{$U(1)$ and Supersymmetric Ward Identities} \label{twopoint} The conventions used in this paper are that $g^{\mu \nu} =$ diag$(1,-1,-1,-1)$, $\{\gamma^\mu,\gamma^\nu\}=2g^{\mu \nu}$, and $\gamma_5 = i\gamma^0 \gamma^1 \gamma^2 \gamma^3$. The Lagrangian of SQED, \begin{eqnarray} \label{lagrang} L &=& |f|^2 + |g|^2+ |\partial_\mu a|^2 + |\partial_\mu b|^2 - \bar{\psi} \not \! \partial \psi \nonumber \\ \nonumber \\ &&-m(a^*f + af^* + b^*g + bg^* + i\bar{\psi} \psi)\nonumber \\ \nonumber \\ &&- ieA^\mu(a^\ast \stackrel{\leftrightarrow}{\partial}_\mu a + b^\ast \stackrel{\leftrightarrow}{\partial}_\mu b + \bar{\psi} \gamma_\mu \psi) \nonumber \\ \\ &&- e[\bar{\lambda}(a^\ast + i\gamma_5 b^\ast)\psi - \bar{\psi}(a + i\gamma_5 b) \lambda] \nonumber \\ \nonumber \\ &&+ ieD(a^\ast b - a b^\ast) +e^2 A_\mu A^\mu (|a|^2 + |b|^2) \nonumber \\ \nonumber \\ &&-\frac{1}{4}F^{\mu \nu}F_{\mu \nu} - \frac{1}{2}\bar{\lambda} \not \! \partial \lambda + \frac{1}{2} D^2, \nonumber \end{eqnarray} is, by construction, invariant with respect to both $U(1)$ gauge transformations and SUSY transformations where the SUSY transformations are given by \cite{wz} \begin{eqnarray} \label{super} &\delta_S a = -i\bar{\zeta} \psi,& \nonumber \\ &\delta_S b = \bar{\zeta} \gamma_5 \psi,& \nonumber \\ &\delta_S \psi = [f + i\gamma_5 g + i\gamma \cdot \partial(a + i\gamma_5 b) -e \gamma \cdot A (a-i\gamma_5 b)]\zeta,& \\ &\delta_S f = \bar{\zeta} [\gamma \cdot \partial \psi + e[-a\lambda -i b \gamma_5 \lambda + i\gamma \cdot A \psi ]],& \nonumber \\ &\delta_S g = i\bar{\zeta} [\gamma_5 \gamma \cdot \partial \psi + e[-\gamma_5 \lambda -i b \lambda -i \gamma \cdot A \gamma_5 \psi]],& \nonumber \end{eqnarray} for the chiral multiplet and \begin{eqnarray} &\delta_S A_\mu = \bar{\zeta} \gamma_\mu \lambda,& \nonumber \\ &\delta_S \lambda = \sigma^{\nu \mu} \partial_\mu A_\nu \zeta +i\gamma_5 D \zeta,& \nonumber \\ &\delta_S D = i\bar{\zeta} \gamma_5 \gamma \cdot \partial \lambda,& \nonumber \end{eqnarray} for the vector multiplet. It is important to note that the transformations in Eqn.(\ref{super}) are not true SUSY transformations but SUSY transformations plus a gauge transformation. This is a manifestation of the Wess-Zumino (WZ) gauge which is used to make the Lagrangian polynomial \cite{wz}. A true SUSY transformation spoils the WZ gauge and must be followed by a gauge transformation which restores it for the Lagrangian to be invariant. It is from this invariance that the SWIs arise. The SWIs completely specify the selectron propagators in terms of the electron propagator \cite{iz}. The SWIs relating the scalar propagators to the electron propagator \cite{iz} are \begin{equation} \label{SUSY_WI} \langle \psi \bar{\psi} \rangle = i\langle a^* f\rangle - i\gamma \cdot p \langle a^* a\rangle = i\langle b^* g\rangle - i\gamma \cdot p \langle b^* b\rangle, \end{equation} and \begin{equation} \gamma \cdot p \langle \psi \bar{\psi} \rangle = -i\langle f^* f\rangle + i\gamma \cdot p \langle f^* a\rangle = -i\langle g^* g\rangle + i\gamma \cdot p \langle g^* b\rangle. \end{equation} Substituting in the fermion propagator \begin{equation} \label{fermiprop} S(p) \equiv \langle \psi \bar{\psi} \rangle = \frac{-i}{\gamma \cdot p A(p^2) + B(p^2)}, \end{equation} gives the scalar propagators \begin{equation} \label{boseprop} D(p^2) \equiv \langle a^* a\rangle = \langle b^* b\rangle = \frac{A(p^2)}{p^2 A^2 (p^2) - B^2 (p^2)}, \end{equation} \begin{equation} \label{atofprop} \langle a^* f\rangle = \langle b^* g\rangle = \frac{B(p^2)}{p^2 A^2 (p^2) - B^2 (p^2)} = \frac{B(p^2)}{A(p^2)} D(p^2), \end{equation} and \begin{equation} \label{auxprop} \langle f^* f\rangle = \langle g^* g\rangle = \frac{p^2 A(p^2)}{p^2 A^2 (p^2) - B^2 (p^2)}. \end{equation} SWIs hold between proper vertices too of course. Taking $\Gamma$ to be the effective action we define $\Gamma_{X..Z} \equiv \frac{\delta^n \Gamma} {\delta X ... \delta Z}$. The two-point proper vertices are constrained by \begin{eqnarray} \Gamma_{\bar{\psi} \psi} & \equiv & \langle \psi \bar{\psi} \rangle^{-1} \nonumber \\ &=& -i\Gamma_{f^* a}(p) + i\gamma \cdot p \Gamma_{f^* f}(p) = -i\Gamma_{g^* b}(p) + i\gamma \cdot p \Gamma_{g^* g}(p), \end{eqnarray} \begin{equation} \gamma \cdot p \Gamma_{\bar{\psi} \psi}(p) = i\Gamma_{a^* a}(p) - i\gamma \cdot p \Gamma_{a^* f}(p) = i\Gamma_{b^* b}(p) - i\gamma \cdot p \Gamma_{b^* g}(p), \end{equation} to be \begin{eqnarray} \label{twoeleven} \Gamma_{a^* a}(p) = \Gamma_{b^* b}(p) = p^2 A(p^2), \\ \Gamma_{a^* f}(p) = \Gamma_{f^* a}(p) = \Gamma_{b^* g}(p) = \Gamma_{g^* b}(p) = - B(p^2), \label{twotwelve} \\ \Gamma_{f^* f}(p) = \Gamma_{g^* g}(p) = A(p^2). \label{twothirteen} \end{eqnarray} It is interesting that $\Gamma_{a^* a}(p) = \Gamma_{b^* b}(p) \neq D(p^2)^{-1}$. This can be attributed to the presence of the auxiliary fields $f$ and $g$. The treatment of proper functions involving selectrons is discussed in the next section. \setcounter{equation}{0} \section{Handling the Proper Functions of Auxiliary Fields} \label{auxiliary} One of the difficulties of the component notation in SQED is that of dealing with the auxiliary fields $f,g$ and $D$. The first two are particularly difficult as they contribute off-diagonal quadratic terms which give the scalar propagators an unfamiliar form. To make the free field theory manifestly Gaussian we define, \begin{eqnarray} \mbox{$[a]$} &\equiv& \left( \begin{array}{c} a \\ f \end{array} \right), \\ \mbox{$[b]$} &\equiv& \left( \begin{array}{c} b \\ g \end{array} \right), \\ \mbox{$[a]^\dagger$} &\equiv& \left( \begin{array}{cc} a^* & f^* \end{array} \right), \\ \mbox{$[b]^\dagger$} &\equiv& \left( \begin{array}{cc} b^* & g^* \end{array} \right). \end{eqnarray} The Lagrangian becomes \begin{eqnarray} \label{newlagrang} L &=& [a]^\dagger \left[ \begin{array}{cc} -\partial^2 & -m \\ -m & 1 \end{array} \right] [a] + [b]^\dagger \left[ \begin{array}{cc} -\partial^2 & -m \\ -m & 1 \end{array} \right] [b] - \bar{\psi} (\not \! \partial + im)\psi \nonumber \\ \nonumber \\ &&- ieA^\mu(\, [a]^\dagger \left[ \begin{array}{cc} \stackrel{\leftrightarrow}{\partial}_\mu & 0 \\ 0 & 0 \end{array} \right] [a] + [b]^\dagger \left[ \begin{array}{cc} \stackrel{\leftrightarrow}{\partial}_\mu & 0 \\ 0 & 0 \end{array} \right] [b] + \bar{\psi} \gamma_\mu \psi) \nonumber \\ \nonumber \\ &&- e[\bar{\lambda}([a]^\dagger + i\gamma_5 [b]^\dagger) \left[ \begin{array}{c} 1 \\ 0 \end{array} \right] \psi - \bar{\psi} [ \begin{array}{cc} 1 & 0 \end{array} ] ([a] + i\gamma_5 [b]) \lambda] \nonumber \\ \\ &&+ ieD([a]^\dagger \left[ \begin{array}{cc} 1 & 0 \\ 0 & 0 \end{array} \right] [b] - [b]^\dagger \left[ \begin{array}{cc} 1 & 0 \\ 0 & 0 \end{array} \right] [a]) \nonumber \\ \nonumber \\ && +e^2 A_\mu A^\mu ([a]^\dagger \left[ \begin{array}{cc} 1 & 0 \\ 0 & 0 \end{array} \right] [a] + [b]^\dagger \left[ \begin{array}{cc} 1 & 0 \\ 0 & 0 \end{array} \right] [b]) \nonumber \\ \nonumber \\ &&-\frac{1}{4}F^{\mu \nu}F_{\mu \nu} - \frac{1}{2}\bar{\lambda} \not \! \partial \lambda + \frac{1}{2} D^2, \nonumber \end{eqnarray} and the problem of ``interpreting'' auxiliary fields is therefore side-stepped. We shall denote the propagators or proper vertices involving $[a]$ or $[b]$ by enclosing them in square brackets to distinguish them from the propagators or vertices of the single component fields $a,b,f$ and $g$. Thus the $[a]$ and $[b]$ propagators are \begin{equation} [D(p^2)] \equiv \left[ \begin{array}{cc} \langle a^* a\rangle & \langle a^* f\rangle \\ \langle f^* a\rangle & \langle f^* f\rangle \end{array} \right] = \left[ \begin{array}{cc} \langle b^* b\rangle & \langle b^* g\rangle \\ \langle g^* b\rangle & \langle g^* g\rangle \end{array} \right]; \end{equation} their photon interaction is \begin{equation} [\Gamma_{(a,b)^* A_\mu (a,b)}](p,q) \equiv \left[ \begin{array}{cc} \Gamma_{(a,b)^* A_\mu (a,b)}(p,q) & \Gamma_{(a,b)^* A_\mu (f,g)}(p,q) \\ \Gamma_{(f,g)^* A_\mu (a,b)}(p,q) & \Gamma_{(f,g)^* A_\mu (f,g)}(p,q) \end{array} \right]; \end{equation} the photino interactions are \begin{equation} [\Gamma_{\bar{\lambda} (a,b)^* \psi}](p,q) \equiv [ \begin{array}{cc} \Gamma_{\bar{\lambda} (a,b)^* \psi}(p,q) & \Gamma_{\bar{\lambda} (f,g)^* \psi}(p,q) \end{array} ], \end{equation} and \begin{equation} [\Gamma_{\bar{\psi} (a,b) \lambda}](p,q) \equiv \left[ \begin{array}{c} \Gamma_{\bar{\psi} (a,b) \lambda}(p,q) \\ \Gamma_{\bar{\psi} (f,g) \lambda}(p,q) \end{array} \right]; \end{equation} and their $D$ interactions are \begin{equation} [\Gamma_{(a,b)^* D (b,a)}](p,q) \equiv \left[ \begin{array}{cc} \Gamma_{(a,b)^* D (b,a)}(p,q) & \Gamma_{(a,b)^* A_\mu (g,f)}(p,q) \\ \Gamma_{(f,g)^* A_\mu (b,a)}(p,q) & \Gamma_{(f,g)^* A_\mu (g,f)}(p,q) \end{array} \right]. \end{equation} One readily checks that Eqs.(\ref{twoeleven}) to (\ref{twothirteen}) are consistent with \begin{equation} [\Gamma_{(a,b)^* (a,b)}](p) \equiv \left[ \begin{array}{cc} \Gamma_{(a,b)^* (a,b)}(p) & \Gamma_{(a,b)^* (f,g)}(p) \\ \Gamma_{(f,g)^* (a,b)}(p) & \Gamma_{(f,g)^* (f,g)}(p) \end{array} \right] = [D(p^2)]^{-1}. \end{equation} With the Lagrangian in its familiar form and our notation established it is a simple matter to write down the DSE for the electron in SQED, namely \begin{eqnarray} \lefteqn{S^{-1}(p) - S_{bare}^{-1}(p)} \\ &=&-\int \frac{d^4 p}{(2\pi)^4}\{ D_{\mu \nu}(p-q) \gamma^\mu S(q) \Gamma_{\bar{\psi} A_\mu \psi}^\nu (q,p) + D_\lambda (p-q) [ \begin{array}{cc} 1 & 0 \end{array} ] [D(q)] [\Gamma_{\bar{\lambda} a^* \psi}](q,p)\}, \nonumber \end{eqnarray} where $D_{\mu \nu}$ is the photon propagator, and $D_\lambda$ the photino propagator. \setcounter{equation}{0} \section{Supersymmetric Vertex Ward Identities} \label{SWI} Before we can find the vertices to substitute into the SDSE, we need the SWIs which constrain them. These are found by taking functional \begin{table}[h] \caption{Each SWI is derived from a functional derivative of $\delta_S \Gamma =0$. The functional derivative leading to each SWI (indicated by its equation number) is given in this table.\label{fnctl}} \vspace{0.2cm} \begin{center} \footnotesize \begin{tabular}{|c|c||c|c|} \hline \raisebox{0pt}[13pt][7pt]{Functional Derivative of $\delta_S \Gamma =0$} & \raisebox{0pt}[13pt][7pt]{SWI} & \raisebox{0pt}[13pt][7pt]{Functional Derivative of $\delta_S \Gamma =0$} & \raisebox{0pt}[13pt][7pt]{SWI}\\ \hline \raisebox{0pt}[13pt][7pt]{$\delta^3 \Gamma/ (\delta a(y) \delta a^* (x) \delta \bar{\lambda} (z))$}& \raisebox{0pt}[13pt][7pt]{\ref{ephotinoa}} & \raisebox{0pt}[13pt][7pt]{$\delta^3 \Gamma/ (\delta \psi(y) \delta D(z) \delta a^*(x))$} & \raisebox{0pt}[13pt][7pt]{\ref{aDpsi}}\\ \hline \raisebox{0pt}[13pt][7pt]{$\delta^3 \Gamma/ (\delta b(y) \delta b^* (x) \delta \bar{\lambda} (z))$} & \raisebox{0pt}[13pt][7pt]{\ref{ephotinob}} & \raisebox{0pt}[13pt][7pt]{$\delta^3 \Gamma/ (\delta \psi(y) \delta D(z) \delta b^*(x))$} & \raisebox{0pt}[13pt][7pt]{\ref{bDpsi}}\\ \hline \raisebox{0pt}[13pt][7pt]{$\delta^3 \Gamma/ (\delta f(y) \delta a^*(x) \delta \bar{\lambda}(z))$} & \raisebox{0pt}[13pt][7pt]{\ref{ephotinof}} & \raisebox{0pt}[13pt][7pt]{$\delta^3 \Gamma/ (\delta \psi(y) \delta D(z) \delta f^*(x))$} & \raisebox{0pt}[13pt][7pt]{\ref{fDpsi}}\\ \hline \raisebox{0pt}[13pt][7pt]{$\delta^3 \Gamma/ (\delta g(y) \delta b^*(x) \delta \bar{\lambda}(z))$} & \raisebox{0pt}[13pt][7pt]{\ref{ephotinog}} & \raisebox{0pt}[13pt][7pt]{$\delta^3 \Gamma/ (\delta \psi(y) \delta D(z) \delta g^*(x))$} & \raisebox{0pt}[13pt][7pt]{\ref{gDpsi}}\\ \hline \raisebox{0pt}[13pt][7pt]{$\delta^3 \Gamma/ (\delta f(y) \delta f^*(x) \delta \bar{\lambda}(z))$} & \raisebox{0pt}[13pt][7pt]{\ref{efpsi}} & \raisebox{0pt}[13pt][7pt]{$\delta^3 \Gamma/ (\delta b(y) \delta D(z) \delta a^*(x))$} & \raisebox{0pt}[13pt][7pt]{\ref{alambdab}}\\ \hline \raisebox{0pt}[13pt][7pt]{$\delta^3 \Gamma/ (\delta g(y) \delta g^*(x) \delta \bar{\lambda}(z))$} & \raisebox{0pt}[13pt][7pt]{\ref{egpsi}} & \raisebox{0pt}[13pt][7pt]{$\delta^3 \Gamma/ (\delta a(y) \delta \lambda (z) \delta b^*(x))$} & \raisebox{0pt}[13pt][7pt]{\ref{blambdaa}}\\ \hline \raisebox{0pt}[13pt][7pt]{$\delta^3 \Gamma/ (\delta \psi(y) \delta A_\mu(z) \delta f^*(x))$} & \raisebox{0pt}[13pt][7pt]{\ref{ephotone}} & \raisebox{0pt}[13pt][7pt]{$\delta^3 \Gamma/ (\delta g(y) \delta \lambda (z) \delta a^*(x))$} & \raisebox{0pt}[13pt][7pt]{\ref{alambdag}}\\ \hline \raisebox{0pt}[13pt][7pt]{$\delta^3 \Gamma/ (\delta \psi(y) \delta A_\mu(z) \delta g^*(x))$} & \raisebox{0pt}[13pt][7pt]{\ref{ephotone2}} & \raisebox{0pt}[13pt][7pt]{$\delta^3 \Gamma/ (\delta f(y) \delta \lambda (z) \delta b^*(x))$} & \raisebox{0pt}[13pt][7pt]{\ref{blambdaf}}\\ \hline \raisebox{0pt}[13pt][7pt]{$\delta^3 \Gamma/ (\delta \psi(y) \delta A_\mu(z) \delta a^*(x))$} & \raisebox{0pt}[13pt][7pt]{\ref{apsi}} & \raisebox{0pt}[13pt][7pt]{$\delta^3 \Gamma (\delta a(y) \delta \lambda (z) \delta g^*(x))$} & \raisebox{0pt}[13pt][7pt]{\ref{glambdaa}}\\ \hline \raisebox{0pt}[13pt][7pt]{$\delta^3 \Gamma/ (\delta \psi(y) \delta A_\mu(z) \delta b^*(x))$} & \raisebox{0pt}[13pt][7pt]{\ref{bpsi}} & \raisebox{0pt}[13pt][7pt]{$\delta^3 \Gamma/ (\delta b(y) \delta \lambda (z) \delta f^*(x))$} & \raisebox{0pt}[13pt][7pt]{\ref{flambdab}}\\ \hline \raisebox{0pt}[13pt][7pt]{$\delta^3 \Gamma/ (\delta \psi_\alpha (y) \delta \bar{\psi}^\beta (x) \delta \lambda_\kappa (z))$} & \raisebox{0pt}[13pt][7pt]{\ref{indices}} & \raisebox{0pt}[13pt][7pt]{$\delta^3 \Gamma/ (\delta g(y) \delta \lambda (z) \delta f^*(x))$} & \raisebox{0pt}[13pt][7pt]{\ref{flambdag}}\\ \hline \multicolumn{2}{c||}{}& \raisebox{0pt}[13pt][7pt]{$\delta^3 \Gamma/ (\delta f(y) \delta \lambda (z) \delta g^*(x))$} & \raisebox{0pt}[13pt][7pt]{\ref{glambdaf}}\\ \cline{3-4} \end{tabular} \end{center} \end{table} derivatives of $\delta_S \Gamma =0$ where $\Gamma$ is the effective action and $\delta_S$ is defined in Eqn.(\ref{super}). The functional derivatives of $\delta_S \Gamma =0$ corresponding to the following SWIs are given in table \ref{fnctl}: \begin{eqnarray} \label{ephotinoa} \lefteqn{\gamma_\mu \Gamma^\mu_{a^* A_\mu a}(p,q)} \\ &=& \Gamma_{\bar{\lambda} a^* \psi}(p,q) \gamma \cdot q + e(B(p^2) - B(q^2)) + \Gamma_{\bar{\lambda} a^* \psi}(-q,-p) \gamma \cdot p, \nonumber \\ \label{ephotinob} \lefteqn{\gamma_\mu \Gamma^\mu_{b^* A_\mu b}(p,q)} \\ &=& -i\Gamma_{\bar{\lambda} b^* \psi}(p,q)\gamma_5 \gamma \cdot q - e(B(p^2) - B(q^2)) - i\Gamma_{\bar{\lambda} b^* \psi}(-q,-p)\gamma_5 \gamma \cdot p, \nonumber \end{eqnarray} \begin{eqnarray} \label{ephotinof} \gamma_\mu \Gamma^\mu_{f^* A_\mu a}(p,q) + e A(p^2) &=& \Gamma_{\bar{\lambda} a^* \psi}(-q,-p) + \Gamma_{\bar{\lambda} f^* \psi}(p,q)\gamma \cdot q, \\ \label{ephotinog} \gamma_\mu \Gamma^\mu_{g^* A_\mu b}(p,q) -e A(p^2) &=& i\Gamma_{\bar{\lambda} b^* \psi}(-q,-p) \gamma_5 + i\Gamma_{\bar{\lambda} g^* \psi}(p,q) \gamma \cdot q \gamma_5, \end{eqnarray} \begin{eqnarray} \label{efpsi} \gamma_\mu \Gamma^\mu_{f^* A_\mu f}(p,q) &=& \Gamma_{\bar{\lambda} f^* \psi}(-q,-p) - \Gamma_{\bar{\lambda} f^* \psi}(p,q), \\ \label{egpsi} \gamma_\mu \Gamma^\mu_{g^* A_\mu g}(p,q) &=& i\Gamma_{\bar{\lambda} g^* \psi}(-q,-p)\gamma_5 - i\Gamma_{\bar{\lambda} g^* \psi}(p,q) \gamma_5, \end{eqnarray} \begin{eqnarray} \label{ephotone} \lefteqn{i\sigma^{\mu \nu}(p-q)_\nu \Gamma_{\bar{\lambda} f^* \psi}(p,q)} \\ & = & \Gamma^\mu_{\bar{\psi} A_\mu \psi}(p,q) -i\gamma \cdot q \Gamma^\mu_{f^* A_\mu f}(p,q) +i\Gamma^\mu_{f^* A_\mu a}(p,q) - ie\gamma^\mu A(p^2), \nonumber \\ \label{ephotone2} \lefteqn{i\sigma^{\mu \nu}(p-q)_\nu \Gamma_{\bar{\lambda} g^* \psi}(p,q)} \\ & = & i\gamma_5 \Gamma^\mu_{\bar{\psi} A_\mu \psi}(p,q) +\gamma_5 \gamma \cdot q \Gamma^\mu_{g^* A_\mu g}(p,q) -\gamma_5 \Gamma^\mu_{g^* A_\mu b}(p,q) + e\gamma_5 \gamma^\mu A(p^2), \nonumber \end{eqnarray} \begin{eqnarray} \label{apsi} \lefteqn{i\sigma^{\mu \nu}(p-q)_\nu \Gamma_{\bar{\lambda} a^* \psi}(p,q)} \\ & = & i\Gamma^\mu_{a^* A_\mu a}(p,q) - i\gamma \cdot q \Gamma^\mu_{a^* A_\mu f}(p,q) -e\gamma^\mu S^{-1}(p) \nonumber -\gamma \cdot p \Gamma^\mu_{\bar{\psi} A_\mu \psi}(p,q) \\ && + ie\gamma^\mu B(p^2), \nonumber \\ \lefteqn{i\sigma^{\mu \nu}(p-q)_\nu \Gamma_{\bar{\lambda} b^* \psi}(p,q)} \label{bpsi} \\ & = & -\gamma_5 \Gamma^\mu_{b^* A_\mu b}(p,q) + \gamma_5 \gamma \cdot q \Gamma^\mu_{b^* A_\mu g}(p,q) \nonumber -i\gamma_5 e\gamma^\mu S^{-1}(p) \\ && - i\gamma_5 \gamma \cdot p \Gamma^\mu_{\bar{\psi} A_\mu \psi}(p,q) - e\gamma_5 \gamma^\mu B(p^2). \nonumber \end{eqnarray} It follows from both (\ref{apsi}) and (\ref{bpsi}) that the rainbow approximation, that is, dressed vertices replaced by bare vertices, violates SUSY in the same way that it violates $U(1)$ gauge invariance. From \begin{eqnarray} \label{indices} 0 &=& -i(\gamma \cdot q)_\sigma^{\; \; \alpha} (\Gamma_{\bar{\psi} f \lambda}(p,q))_\beta^{\; \; \kappa} + (\gamma_5 \gamma \cdot q)_\sigma^{\; \; \alpha} (\Gamma_{\bar{\psi} g \lambda}(p,q))_\beta^{\; \; \kappa} \\ && - i(\gamma \cdot p C)_{\beta \sigma} \frac{\delta^2}{\delta \psi_\alpha(q) \delta f^*(p)} (\frac{\delta \Gamma}{\delta \lambda(p-q)} C^{-1})^\kappa \nonumber \\ && -(\gamma_5\gamma \cdot p C)_{\beta \sigma} \frac{\delta^2}{\delta \psi_\alpha(q) \delta g^*(p)} (\frac{\delta \Gamma}{\delta \lambda(p-q)} C^{-1})^\kappa \nonumber \\ && - (\gamma_5 (\gamma \cdot p - \gamma \cdot q))_\sigma^{\; \; \kappa} (\Gamma_{\bar{\psi} D \psi}(p,q))_\beta^{\; \; \alpha}, \nonumber \end{eqnarray} where $C$ is the charge conjugation matrix, we obtain \begin{eqnarray} 0 & = & (\gamma \cdot p - \gamma \cdot q)\gamma_5 \mbox{Tr}(\Gamma_{\bar{\psi} D \psi}(p,q)) + \gamma_\mu \mbox{Tr}(\Gamma^\mu_{\bar{\psi} A_\mu \psi}(p,q)) + i\Gamma_{\bar{\psi} a \lambda}(p,q) \nonumber \\ &&- \gamma_5 \Gamma_{\bar{\psi} b \lambda}(p,q) - i\Gamma_{\bar{\psi} a \lambda\psi}(-q,-p) + \gamma_5 \Gamma_{\bar{\psi} b \lambda}(-q,-p) \\ && - i\gamma \cdot q \Gamma_{\bar{\psi} f \lambda}(p,q) + \gamma_5 \gamma \cdot q \Gamma_{\bar{\psi} g \lambda}(p,q) - i\gamma \cdot p \Gamma_{\bar{\psi} f \lambda}(-q,-p) \nonumber \\ && + \gamma_5 \gamma \cdot p \Gamma_{\bar{\psi} g \lambda}(-q,-p), \nonumber \end{eqnarray} by setting $\beta = \alpha$ and summing, and \begin{eqnarray} \label{eDpsi} 0 & = & i\mbox{Tr}(\Gamma_{\bar{\psi} a \lambda} (p,q)) - \gamma_5 \mbox{Tr}(\Gamma_{\bar{\psi} b \lambda} (p,q)) - i\gamma \cdot q \mbox{Tr}(\Gamma_{\bar{\psi} f \lambda} (p,q)) \\ && + \gamma_5 \gamma \cdot q \mbox{Tr}(\Gamma_{\bar{\psi} g \lambda} (p,q)) - i\Gamma_{\bar{\lambda} a^* \psi}(p,q) + \gamma_5 \Gamma_{\bar{\lambda} b^* \psi}(p,q) - i\gamma \cdot p \Gamma_{\bar{\lambda} f^* \psi}(p,q) \nonumber \\ && - \gamma \cdot p \gamma_5 \Gamma_{\bar{\lambda} g^* \psi}(p,q)) + \gamma_\mu \Gamma^\mu_{\bar{\psi} A_\mu \psi} (p,q) - \gamma_5 (\gamma \cdot p - \gamma \cdot q) \Gamma_{\bar{\psi} D \psi} (p,q), \nonumber \end{eqnarray} by setting $\beta = \kappa$ and summing. Finally there are the SWIs governing the vertices of the $D$ particle; \begin{eqnarray} \label{aDpsi} \lefteqn{i\gamma_5 \Gamma_{\bar{\lambda} a^* \psi}(p,q)} \\ & = & \gamma \cdot p \Gamma_{\bar{\psi} D \psi} (p,q) + \gamma_5 \Gamma_{a^* D b}(p,q) - \gamma_5 \gamma \cdot q \Gamma_{a^* D g}(p,q), \nonumber \\ \label{bDpsi} \lefteqn{i\gamma_5 \Gamma_{\bar{\lambda} b^* \psi}(p,q)} \\ & = & i\gamma_5 \gamma \cdot p \Gamma_{\bar{\psi} D \psi} (p,q) - i\Gamma_{b^* D a}(p,q) + i\gamma \cdot q \Gamma_{b^* D f}(p,q), \nonumber \end{eqnarray} \begin{eqnarray} \label{fDpsi} \lefteqn{\gamma_5 \Gamma_{f^* D b}(p,q)} \\ & = & i\gamma_5 \Gamma_{\bar{\lambda} f^* \psi}(p,q) + \gamma_5 \gamma \cdot q \Gamma_{f^* D g}(p,q) + \Gamma_{\bar{\psi} D \psi} (p,q), \nonumber \\ \lefteqn{\gamma_5 \Gamma_{g^* D a}(p,q)} \label{gDpsi} \\ & = & - \Gamma_{\bar{\lambda} g^* \psi}(p,q) + \gamma_5 \gamma \cdot q \Gamma_{g^* D f}(p,q) - \Gamma_{\bar{\psi} D \psi} (p,q), \nonumber \end{eqnarray} \begin{eqnarray} \label{alambdab} \lefteqn{\gamma_5(\gamma \cdot p - \gamma \cdot q) \Gamma_{a^* D b}(p,q)} \\ & = & \Gamma_{\bar{\lambda} b^* \psi}(-q,-p) \gamma \cdot p + i\Gamma_{\bar{\lambda} a^* \psi}(p,q) \gamma \cdot q \gamma_5 + ie\gamma_5 (B(p^2) - B(q^2)), \nonumber \\ \lefteqn{\gamma_5(\gamma \cdot p - \gamma \cdot q) \Gamma_{b^* D a}(p,q)} \label{blambdaa} \\ & = & i\Gamma_{\bar{\lambda} a^* \psi}(-q,-p) \gamma \cdot p \gamma_5 + \Gamma_{\bar{\lambda} b^* \psi}(p,q) \gamma \cdot q + ie\gamma_5 (B(p^2) - B(q^2)), \nonumber \end{eqnarray} \begin{eqnarray} \label{alambdag} \lefteqn{\gamma_5(\gamma \cdot p - \gamma \cdot q) \Gamma_{a^* D g}(p,q)} \\ & = &\Gamma_{\bar{\lambda} g^* \psi}(-q,-p) \gamma \cdot p - i\Gamma_{\bar{\lambda} a^* \psi}(p,q) \gamma_5 + ie\gamma_5 A(q^2), \nonumber \\ \lefteqn{\gamma_5(\gamma \cdot p - \gamma \cdot q) \Gamma_{b^* D f}(p,q)} \label{blambdaf} \\ & = &i\Gamma_{\bar{\lambda} f^* \psi}(-q,-p) \gamma \cdot p \gamma_5 - \Gamma_{\bar{\lambda} b^* \psi}(p,q) + ie\gamma_5 A(q^2), \nonumber \\ \lefteqn{\gamma_5(\gamma \cdot p - \gamma \cdot q) \Gamma_{g^* D a}(p,q)} \label{glambdaa} \\ & = &\Gamma_{\bar{\lambda} g^* \psi}(p,q) \gamma \cdot q + i\Gamma_{\bar{\lambda} a^* \psi}(-q,-p) \gamma_5 - ie\gamma_5 A(p^2), \nonumber \\ \lefteqn{\gamma_5(\gamma \cdot p - \gamma \cdot q) \Gamma_{f^* D b}(p,q)} \label{flambdab} \\ & = &i\Gamma_{\bar{\lambda} f^* \psi}(p,q) \gamma \cdot q \gamma_5 + \Gamma_{\bar{\lambda} b^* \psi}(-q,-p) - ie\gamma_5 A(p^2), \nonumber \end{eqnarray} \begin{eqnarray} \gamma_5(\gamma \cdot p - \gamma \cdot q) \Gamma_{f^* D g}(p,q) & = & \Gamma_{\bar{\lambda} g^* \psi}(-q,-p) \label{flambdag} - i\Gamma_{\bar{\lambda} f^* \psi}(p,q) \gamma_5, \\ \gamma_5(\gamma \cdot p - \gamma \cdot q) \Gamma_{g^* D f}(p,q) & = & i\Gamma_{\bar{\lambda} f^* \psi}(-q,-p) \gamma_5 - \Gamma_{\bar{\lambda} g^* \psi}(p,q) \label{glambdaf}. \end{eqnarray} These make up the entire set of SWIs containing only three-or-fewer point proper functions, modulo charge conjugation. A suitable vertex ansatz must also be consistent with the WTIs; \begin{eqnarray} \label{WTI} (p-q)_\mu [\Gamma_{(a.b)^* A_\mu (a,b)}]^\mu (p,q) &=& e[\Gamma_{(a,b)^* (a,b)}] (p) - e[\Gamma_{(a,b)^* (a,b)}] (q), \\ (p-q)_\mu \Gamma^\mu_{\bar{\psi} A_\mu \psi}(p,q) &=& eS^{-1}(p) - eS^{-1}(q). \end{eqnarray} We also have from charge conjugation invariance that \begin{equation} \begin{array}{l} \label{conjugate} \mbox{[$\Gamma_{\bar{\psi} (a,b) \lambda}$]}(p,q) = - C [\Gamma_{\bar{\lambda} (a^*,b^*) \psi}](-q,-p)^T C^{-1}, \\ \mbox{[$\Gamma_{(a^*,b^*) D (b,a)}$]}(p,q) = - [\Gamma_{(b^*,a^*) D (a,b)}](-q,-p). \end{array} \end{equation} \setcounter{equation}{0} \section{Solution to SWIs and WTIs in SQED} \label{threepoint} Below is a solution for the SWIs and WTIs. It is the most general set of vertices consistent with both the WTIs and the SWIs and free of kinematic singularities if one assumes charge conjugation invariance and \begin{equation} \label{condition} [\Gamma_{a^* A_\mu a}]^\mu (p,q) = [\Gamma_{b^* A_\mu b}]^\mu (p,q). \end{equation} Proof of this is presented in Appendix \ref{proof}. The assumption of Eqn.(\ref{condition}) is true to all orders in perturbation theory, and any nonperturbative violations of this assumption are restricted by the WTIs to lie completely within their transverse components. Our general solution is as follows: \newline The scalar-photon vertices are \begin{eqnarray} \lefteqn{\Gamma^\mu_{a^* A_\mu a}(p,q) = \Gamma^\mu_{b^* A_\mu b}(p,q)} \label{onea} \\ &=& \frac{e}{p^2 - q^2}(p^2 A(p^2) - q^2 A(q^2))(p+q)^\mu + [p^\mu (q^2 - p\cdot q) + q^\mu (p^2 - p\cdot q)] T_{aa}(p^2,q^2),p\cdot q), \nonumber \\ \lefteqn{\Gamma^\mu_{a^* A_\mu f}(p,q) = \Gamma^\mu_{b^* A_\mu g}(p,q) = \Gamma^\mu_{f^* A_\mu a}(p,q) = \Gamma^\mu_{g^* A_\mu b}(p,q)} \label{oneb} \\ &=& \frac{-e}{p^2 - q^2}(B(p^2) - B(q^2))(p+q)^\mu + [p^\mu (q^2 - p\cdot q) + q^\mu (p^2 - p\cdot q)] T_{af}(p^2,q^2,p\cdot q) \nonumber, \\ \lefteqn{\Gamma^\mu_{f^* A_\mu f}(p,q) = \Gamma^\mu_{g^* A_\mu g}(p,q)} \label{onec} \\ &=& \frac{e}{p^2 - q^2}(A(p^2) - A(q^2))(p+q)^\mu + [p^\mu (q^2 - p\cdot q) + q^\mu (p^2 - p\cdot q)] T_{ff}(p^2,q^2,p\cdot q), \nonumber \end{eqnarray} where the three functions $T_{aa}(p^2,q^2,p\cdot q),T_{af}(p^2,q^2,p\cdot q)$ and $T_{ff}(p^2,q^2,p\cdot q)$, each satisfying $T(p^2,q^2,p\cdot q) = T(q^2,p^2,p\cdot q)$, are free of kinematic singularities and represent the only degrees of freedom inherent in the solution. The forms (\ref{onea}) to (\ref{onec}) are equivalent to that given by Ball and Chiu \cite{ball} in the context of non SUSY scalar QED. The photino vertices are \begin{eqnarray} \label{thirtysix} \Gamma_{\bar{\lambda} a^* \psi}(p,q) &=& \frac{e}{p^2 - q^2}(p^2 A(p^2) - q^2 A(q^2)) + \frac{e}{p^2 - q^2}(B(p^2) - B(q^2))\gamma \cdot q \nonumber \\ && + \frac{1}{2} e (p^2 - \gamma \cdot q \gamma \cdot p)T_{aa}(p^2,q^2,p\cdot q) \\ && + \frac{1}{2} e p^2(q^2 - \gamma \cdot p \gamma \cdot q)T_{ff}(p^2,q^2,p\cdot q) \nonumber \\ &&+ \frac{1}{2} e [\gamma \cdot p (p^2 - q^2) - 2\gamma \cdot q (p^2 - p\cdot q)] T_{af}(p^2,q^2,p\cdot q), \nonumber \end{eqnarray} and \begin{eqnarray} \label{thirtyfive} \Gamma_{\bar{\lambda} f^* \psi}(p,q) &=& \frac{-e}{p^2 - q^2}(A(p^2) - A(q^2))\gamma \cdot q -\frac{e}{p^2 - q^2}(B(p^2) - B(q^2)) \nonumber \\ && +\frac{1}{2} e(\gamma \cdot p - \gamma \cdot q)T_{aa}(p^2,q^2,p\cdot q) \\ && + \frac{1}{2} e(p-q)^2 T_{af}(p^2,q^2,p\cdot q) \nonumber \\ && - \frac{1}{2} e\gamma \cdot q(p^2 - \gamma \cdot p \gamma \cdot q)T_{ff}(p^2,q^2,p\cdot q). \nonumber \end{eqnarray} The electron-photon vertex must be restricted at least to the form given by Ball and Chiu\cite{ball} for non SUSY QED. For the SUSY case we find \begin{eqnarray} \Gamma^\mu_{\bar{\psi} A_\mu \psi}(p,q) &=& \Gamma^\mu_{BC}(p,q) + \frac{ie}{p^2 - q^2}(A(p^2) - A(q^2)) [\frac{1}{2} T_3^\mu - T_8^\mu] \nonumber \\ && - \frac{ie}{p^2 - q^2}(B(p^2) - B(q^2))T_5^\mu + \frac{1}{2} ie T_{aa}(p^2,q^2,p\cdot q) T_3^\mu \nonumber \\ && +ie T_{af}(p^2,q^2,p\cdot q) [\frac{1}{2} (p-q)^2 T_5^\mu - T_1^\mu] \\ && + \frac{1}{2} ie T_{ff}(p^2,q^2,p\cdot q) [T_2^\mu - p\cdot q T_3^\mu - (p-q)^2 T_8^\mu], \nonumber \end{eqnarray} where \begin{eqnarray} \Gamma^\mu_{BC}(p,q) &=& \frac{1}{2} \frac{ie}{p^2 - q^2}(\gamma \cdot p + \gamma \cdot q) (A(p^2) - A(q^2))(p+q)^\mu \\ && + ie\frac{1}{2} (A(p^2) + A(q^2)) \gamma^\mu + \frac{ie}{p^2 - q^2}(B(p^2) - B(q^2))(p+q)^\mu, \nonumber \end{eqnarray} \begin{eqnarray} \label{transverse} T_1^\mu &=& p^\mu(q^2 - p\cdot q) + q^\mu (p^2 - p\cdot q), \\ T_2^\mu &=& (\gamma \cdot p + \gamma \cdot q)T_1^\mu, \\ T_3^\mu &=& \gamma^\mu(p-q)^2 - (\gamma \cdot p - \gamma \cdot q)(p-q)^\mu], \\ T_5^\mu &=& \sigma^{\mu \nu}(p-q)_\nu, \\ T_8^\mu &=& \frac{1}{2}(\gamma \cdot p \gamma \cdot q \gamma^\mu - \gamma^\mu \gamma \cdot q \gamma \cdot p). \end{eqnarray} Finally there are the vertices for the $D$-boson, namely, \begin{eqnarray} \lefteqn{\Gamma_{a^* D b} (p,q) = -\Gamma_{b^* D a} (p,q)} \\ & = & \frac{ie}{p^2 - q^2}(p^2 A(p^2) - q^2 A(q^2)) - iep\cdot q T_{a^* a}(p^2,q^2,p\cdot q) \nonumber \\ && + \frac{1}{2} ie p^2 q^2 T_{ff}(p^2,q^2,p\cdot q), \nonumber \\ \nonumber \\ \lefteqn{\Gamma_{f^* D g} (p,q) = -\Gamma_{g^* D f} (p,q)} \\ &=&\frac{ie}{p^2 - q^2}(A(p^2) - A(q^2)) + ie T_{a^* a}(p^2,q^2,p\cdot q) \nonumber \\ &&- ie p\cdot q T_{f^* f}(p^2,q^2,p\cdot q), \nonumber \end{eqnarray} \begin{eqnarray} \Gamma_{g^* D a} (p,q)&=& \frac{ie}{p^2 - q^2}(B(p^2) - B(q^2)) \\ &&-ie (q^2 - p\cdot q) T_{af}(p^2,q^2,p\cdot q), \nonumber \\ \Gamma_{a^* D g} (p,q) &=& \frac{-ie}{p^2 - q^2}(B(p^2) - B(q^2)) \\ &&+ie (p^2 - p\cdot q) T_{af}(p^2,q^2,p\cdot q), \nonumber \\ \Gamma_{f^* D b} (p,q) &=& \frac{-ie}{p^2 - q^2}(B(p^2) - B(q^2)) \\ &&+ie (q^2 - p\cdot q) T_{af}(p^2,q^2,p\cdot q), \nonumber \\ \Gamma_{b^* D f} (p,q) &=& \frac{ie}{p^2 - q^2}(B(p^2) - B(q^2)) \\ &&-ie (p^2 - p\cdot q) T_{af}(p^2,q^2,p\cdot q), \nonumber \end{eqnarray} and \begin{eqnarray} \Gamma_{\bar{\psi} D \psi}(p,q) &=& \frac{1}{2} ie\gamma_5 [(p^2 - q^2)T_{af}(p^2,q^2,p\cdot q) \\ && +(\gamma \cdot p + \gamma \cdot q)T_{a^* a}(p^2,q^2,p\cdot q) \nonumber \\ && -(\gamma \cdot q p^2 + \gamma \cdot p q^2)T_{ff}(p^2,q^2,p\cdot q)]. \nonumber \end{eqnarray} \setcounter{equation}{0} \section{Conclusion} We have derived the three-point SWIs for SQED and found a solution, given in sections \ref{twopoint} and \ref{threepoint}, which, under the reasonable assumptions of charge conjugation invariance and symmetry between $[a]$ and $[b]$ with respect to their photon interaction, comprises the most general set of vertices consistent with both the SWIs and WTIs and free of kinematic singularities. They are, in fact, the SUSY equivalent of the Ball-Chiu vertex. These SUSY Ball-Chiu vertices have only three degrees of freedom between them once the electron propagator is known, compared with non SUSY QED which has eight. The loss of degrees of freedom occurs entirely within the electron-photon vertex. The scalar-photon vertices remain unchanged from non SUSY scalar QED (with auxiliary fields). We have given the form of the electron DSE. There is no need to consider also the DSE for scalar partners since SWIs ensure that the propagators of all chiral multiplet fields can be written in terms of the same two scalar functions $A(p^2)$ and $B(p^2)$ (See Sec.(\ref{twopoint})). Solving the DSE for any chiral multiplet field can therefore be accomplished by projecting from the electron DSE a pair of coupled integral equations for $A(p^2)$ and $B(p^2)$. Numerical solutions of the analogous calculation in non SUSY QED \cite{HW95,HWR96,HSW97,RW94} and QED$_3$ \cite{BR91,CPW92} using the minimal Ball-Chiu and Curtis-Pennington \cite{CP91} vertex ans\"atze exist in the literature. The same task in SUSY is conceptually similar and the presence of extra terms in the DSE is not expected to reduce its feasibility. Indeed such numerical work has been done already in the rainbow approximation in SQED$_3$ \cite{me}. The way now lies open to transcend the rainbow approximation in the analysis of SQED and SQED$_3$ in the nonperturbative limit.
\section{Introduction} A solution to the gauge hierarchy, based on extra spacial dimensions, was recently proposed by Arkani-Hamed, Dimopoulos and Dvali \cite{theory}. They assumed the space-time is $4+n$ dimensional, with the standard model (SM) particles living on a brane. While the electromagnetic, strong, and weak forces are confined to this brane, gravity can propagate in the extra dimensions. To solve the gauge hierarchy problem they proposed the ``new'' Planck scale $M_S$ is of the order of TeV with very large extra dimensions. The size $R$ of these extra dimensions can be as large as 1 mm, which corresponds to a compactification scale $R^{-1}$ as low as $10^{-4}$ eV. The usual Planck scale $M_G=1/\sqrt{G_N} \sim 1.22 \times 10^{19}$ GeV is related to this effective Planck scale $M_S$ using the Gauss's law: \begin{equation} R^n \, M_S^{n+2} \sim M_G^2 \;, \end{equation} where $n$ is the number of extra dimensions. For $n=1$ it gives a large value for $R$, which is already ruled out by gravitational experiments. On the other hand, $n=2$ gives $R \mathrel{\raisebox{-.6ex}{$\stackrel{\textstyle<}{\sim}$}} 1$ mm, which is in the margin beyond the reach of present gravitational experiments. The graviton including its excitations in the extra dimensions can couple to the SM particles on the brane with an effective strength of $1/M_S$ (instead of $1/M_G$) after summing the effect of all excitations collectively, and thus the gravitation interaction becomes comparable in strength to weak interaction at TeV scale. Hence, it can give rise to a number of phenomenological activities testable at existing and future colliders \cite{collider}. So far, studies show that there are two categories of signals: direct and indirect. The indirect signal refers to exchanges of gravitons in the intermediate states, while direct refers to production or associated production of gravitons in the final state. Indirect signals include fermion pair, gauge boson pair production, correction to precision variables, etc. There are also other astrophysical and cosmological signatures and constraints \cite{others}. Among the constraints the cooling of supernovae by radiating gravitons places the strongest limit on the effective Planck scale $M_S$ of order 50 TeV for $n=2$, which renders collider signatures for $n=2$ uninteresting. Thus, we concentrate on $n\ge 3$. In this work, we perform a global analysis of the lepton-quark neutral current data on the low scale gravity model. The global data include HERA neutral current deep-inelastic scattering, Drell-yan production at the Tevatron, and total hadronic, $b\bar b$ and $c\bar c$ pair cross sections at LEPII. In addition, we also include the leptonic cross sections $e^+ e^- \to \mu^+\mu^-, \tau^+ \tau^-$ at LEPII in our analysis. The $\nu$-N scattering data from CCFR and NuTeV have been shown by Rizzo \cite{collider} to be very insignificant in constraining the low scale gravity, and so we shall not include this data set in our analysis. We shall see that the Drell-yan production, due to the large invariant mass data, provides the strongest constraint among the global data. By combining all data, the effective Planck scale $M_S$ must be larger than about 1.12 TeV for $n=3$ and 0.94 TeV for $n=4$ at 95\%CL. This is our main result. The organization of the paper is as follows. In the next section, we shall describe each set of data used in our global analysis and derive the effect of the low scale gravity. In Sec. III, we give the numerical results and interpretations for our fits, {}from which we can draw the limits on the effective Planck scale. \section{Global Data} Before we come to each data set, let us first give a general expression for the square of amplitude for $e^- e^+ \to q \bar q$: \begin{eqnarray} \sum \left| {\cal M} \right |^2 &=& 4 u^2 \left( |M_{LL}|^2 + |M_{RR}|^2 \right) + 4 t^2 \left( |M_{LR}|^2 + |M_{RL}|^2 \right) \nonumber \\ &+& 2\pi^2 \; \left( \frac{\cal F}{M_S^4} \right )^2 \; ( t^4 - 6 t^3 u + 18 t^2 u^2 - 6 t u^3 + u^4 ) \nonumber \\ &+& 8 \pi e^2 Q_e Q_q \; \frac{\cal F}{M_S^4} \; \frac{(u-t)^3}{s} \nonumber \\ &+& \frac{8 \pi e^2}{\sin^2\theta_{\rm w} \cos^2 \theta_{\rm w} } \; \frac{\cal F}{M_S^4} \; \frac{1}{s - M_Z^2} \; \biggr[ g_a^e g_a^q ( t^3 - 3t^2 u - 3t u^2 + u^3 ) + g_v^e g_v^q ( u-t)^3 \biggr ] \;, \end{eqnarray} where $s,t,u$ are the usual Mandelstam variables, $e = \sqrt{4\pi \alpha}$ is the electromagnetic coupling constant, $Q_f$ is the electric charge of the fermion $f$, $\theta_{\rm w}$ is the Weinberg mixing angle, $g_a^f$ and $g_v^f$ are, respectively, the axial-vector and the vector $Z$ couplings to the fermion $f$. The reduced amplitudes $M_{\alpha\beta}$ ($\alpha,\beta= L,R$) are \begin{equation} \label{reduce} M_{\alpha\beta} = \frac{e^2 Q_e Q_q}{s} + \frac{e^2}{\sin^2\theta_{\rm w} \cos^2 \theta_{\rm w} } \; \frac{g_\alpha^e g_\beta^q}{s - M_Z^2} \end{equation} where \begin{eqnarray} g_L^f = T_{3f} - Q_f \sin^2\theta_{\rm w}\;, && g_R^f = - Q_f \sin^2\theta_{\rm w} \nonumber \\ g_v^f = (g_L^f + g_R^f)/2\;, && g_a^f = (g_L^f - g_R^f)/2 \nonumber \;. \end{eqnarray} In the derivation, we have followed the notation in Han {\it et al}. \cite{collider} and we have taken $M_S^2 \gg s, |t|, |u|$ and in this case the propagator factor $D(s)=D(|t|)=D(|u|)$, which is given by \begin{equation} \kappa^2 |D(s)| = \frac{16 \pi}{M_S^4}\; \times {\cal F} \;, \end{equation} where the factor ${\cal F}$ is given by \begin{equation} \label{F} {\cal F} = \Biggr \{ \begin{array}{l} \log \left( \frac{M_S^2}{s} \right ) \;\; {\rm for}\;\; n=2 \;, \\ \frac{2}{n-2} \;\;\;\;\;\;\;\;\;\;\; {\rm for}\;\; n>2 \;. \end{array} \end{equation} For $n\ge 3$ ${\cal F}$ is always positive. The amplitude square for the crossed channels $e^\pm \stackrel{(-)}{q} \to e^\pm \stackrel{(-)}{q}$ can be obtained using the crossing symmetry. \subsection{HERA neutral-current data} Both H1 and ZEUS have measured the neutral-current deep-inelastic scattering cross sections at high-$Q^2$ region. Despite the excess in cross section reported in early 1997, the 1997 data alone agreed well with the SM. The double differential cross section for $e^+ p \to e^+ X$ is given by \begin{eqnarray} \frac{d^2\sigma}{dy dQ^2} &=& \sum_q \frac{f_q(x)}{16\pi}\, \frac{Q^2}{s y^2} \; \Biggr \{ (1-y)^2 ( |M_{LL}|^2 + |M_{RR}|^2) + |M_{LR}|^2 + |M_{RL}|^2 \nonumber \\ &&+ \frac{\pi^2}{2}\, \left( \frac{Q^2}{y} \right )^2 \, \left( \frac{\cal F}{M_S^4} \right )^2 \; (32 - 64 y +42 y^2 -10y^3 + y^4 ) \nonumber \\ &&+ 2 \pi e^2 Q_e Q_q \;\left( \frac{\cal F}{M_S^4} \right )\; \frac{(2-y)^3}{y} \nonumber \\ &&+ \frac{2\pi e^2 }{\sin^2\theta_{\rm w} \cos^2\theta_{\rm w}}\; \left( \frac{\cal F}{M_S^4} \right )\; \left( \frac{Q^2}{y} \right )\; \frac{1}{-Q^2 - M_Z^2} \; \biggr[ g_a^e g_a^q (6y -6y^2 +y^3 ) + g_v^e g_v^q (y-2)^3 \biggr ] \Biggr \} \nonumber \\ &&+ \frac{\pi}{2}\, f_{g}(x) \, \frac{Q^2}{s y^2} \; \left( \frac{\cal F}{M_S^4} \right )^2 \; \left( \frac{Q^2}{y} \right)^2 \; (1-y)(y^2 -2y + 2) \;, \end{eqnarray} where $Q^2 =s x y$ is the square of the momentum-transfer, $\sqrt{s}=300$ GeV, and $f_{q/g}(x)$ are the parton distribution functions. The reduced amplitudes $M_{\alpha\beta}$ are given in Eq. (\ref{reduce}) with $s$ replaced by $-Q^2$. The last term in the above equation comes from the additional channel $e^+ g \to e^+ g$ allowed by the low scale gravity interactions. The channels $\sum_{\bar q} e^+ \bar q \to e^+ \bar q$ are also included. The $Q^2$ data of ZEUS \cite{h1-zeus}, including systematics, are recently published while the H1 data \cite{h1-zeus} used in our analysis are shown in Table \ref{hera}. \subsection{Drell-yan at the Tevatron} CDF measured the double differential cross section $d^2\sigma/dM_{\ell\ell} dy\;(\ell=e,\mu)$ for the Drell-yan production, where $M_{\ell\ell}$ and $y$ are, respectively, the invariant mass and the rapidity of the lepton pair. Essentially, CDF measured the cross section in the range $-1<y<1$ and then average over $y$ to obtain the double differential cross section. The double differential cross section for $p \bar p \to \ell^+ \ell^-$ is given by \begin{eqnarray} \frac{d^2\sigma}{dM_{\ell\ell} dy} &=& K \Biggr\{ \sum_q \; \biggr [ \frac{M_{\ell\ell}^3}{72\pi s} \; f_q(x_1) f_{\bar q}(x_2)\; \left( |M_{LL}|^2 + |M_{LR}|^2 + |M_{RL}|^2 + |M_{RR}|^2 \right ) \nonumber \\ &+& \frac{\pi M_{\ell\ell}^7}{120 s}\; f_q(x_1) f_{\bar q}(x_2) \; \left( \frac{\cal F}{M_S^4} \right )^2 \; \biggr ] + \frac{\pi M_{\ell\ell}^7}{80 s}\; f_g(x_1) f_g (x_2) \; \left( \frac{\cal F}{M_S^4} \right )^2 \Biggr \} \;, \end{eqnarray} where $x_{1,2}=M_{\ell\ell} e^{\pm y}/\sqrt{s}$, $\sqrt{s}=1800$ GeV, $f_{q,g}(x_i)$ are the parton distribution functions, and the sum over all possible $q\bar q$ pairs is understood. Note that after integrating the scattering angle $\cos\theta^*$ over the whole range in the center-of-mass frame, the interference terms proportional to $({\cal F}/M_S^4)$ vanish. The reduced amplitudes $M_{\alpha\beta}\;(\alpha,\beta=L,R)$ are given in Eq. (\ref{reduce}) with $s$ replaced by $\hat s=M_{\ell\ell}^2$. The QCD $K$ factor is given by $K=1+ \frac{\alpha_s(\hat s)}{2\pi}\frac{4}{3} ( 1+ \frac{4\pi^2}{3})$. The second term of the above equation comes from $q\bar q \to G \to \ell^+ \ell^-$ and the third one comes from $gg \to G \to \ell^+ \ell^-$. This $K$ factor is, in principle, not valid for the $gg$ channel, but it will not affect our calculation because we normalize our SM calculation to the expected number of events in each bin given by CDF when we deal with the CDF data (similarly, we normalize to the first bin of the D0 data when we deal with D0 data.) On the other hand, D0 measured the differential cross section $d\sigma/dM_{ee}$ with the rapidity range integrated. The Drell-yan data \cite{dy-cdf-d0} we used in our analysis are shown in Table \ref{dy}. \subsection{LEPII fermion pair Cross sections} LEPII has measured the leptonic cross sections, hadronic cross sections, and heavy flavor ($b$ and $c$) production. Since the low scale gravity interactions are naturally flavor-blind, we shall include the effects on all flavors here. For a massless $q$ the expression for $\sigma(e^+ e^- \to q \bar q)$ is given by \begin{equation} \sigma(e^+ e^- \to q \bar q ) = K \Biggr\{ \frac{s}{16\pi} \left( |M_{LL}|^2 + |M_{LR}|^2 + |M_{RL}|^2 + |M_{RR}|^2 \right ) + \frac{3\pi s^3}{80} \left( \frac{\cal F}{M_S^4} \right )^2 \Biggr \} \;, \end{equation} where the reduced amplitudes $M_{\alpha\beta}\;(\alpha,\beta=L,R)$ are given in Eq. (\ref{reduce}), and the QCD $K$ factor is $K=1+ \alpha_s/\pi + 1.409 (\alpha_s/\pi)^2 -12.77 (\alpha_s/\pi)^3$. The leptonic cross sections $\sigma(e^+ e^- \to \mu^+\mu^-, \tau^+ \tau^-)$ do not have this $K$ factor. The LEPII data \cite{lep} used in our analysis are given in Tables \ref{lep-h} and \ref{lep-l}. \subsection{$\nu$-N Scattering} The recent measurements by CCFR and NuTeV \cite{ccfr} are the most precise ones on the neutrino-quark couplings. Since the gravitons also couple to neutrinos, the measurements should, in principle, place a constraint on the scale $M_S$. However, the analysis by Rizzo in Ref. \cite{collider} showed that the constraint coming from these low energy $\nu$-N scattering is very weak. This is because the center-of-mass energies of these collisions are only of order of tens of GeV's, even though the neutrino beam energy is as high as a few hundred GeV. In addition, the effective operators induced by the low scale gravity interactions are at least of dimension eight. This is in contrast to the traditional four-fermion contact interactions, which are only of dimension six. Therefore, at such low center-of-mass energies in these collisions, the effect of low scale gravity is extremely minimal. We are not going to include these data in our global fit. \section{Fits} For the fitting we follow the procedures in Refs. \cite{contact}, where the four-fermion contact interactions are analyzed with a similar but larger global data set. The difference is that Refs. \cite{contact} include also the data sets from the low energy $e$-N scattering and the atomic parity violation, which constrain parity violating interactions. But the gravity is certainly parity conserving and thus it receives no restriction from these parity-violating experiments. The interference effect between the SM amplitude and the low scale gravity scales as ${\cal F}/M_S^4$, while the pure low scale gravity scales as $({\cal F}/M_S^4)^2$. To linearize the fitting we use the parameter $\eta={\cal F}/M_S^4$. We use MINUIT to minimize the $\chi^2$. The best estimates of $\eta$ for each individual data set and the combined set are shown in Table \ref{fit}. The SM fit gives a $\chi^2=94.10$ for 118 d.o.f., while the fit with non-zero $\eta$ gives a $\chi^2=94.05$ for 117 d.o.f. Since the fits do not show any evidence for new physics, we can then place limits on the $\eta={\cal F}/M_S^4$. The physical region of $\eta$ is $\eta \ge 0$, and so we define the 95\%CL upper limit for $\eta$ as \begin{equation} 0.95 = \frac{\int_0^{\eta_+} {\cal P}(\eta) \; d\eta} {\int_0^\infty {\cal P}(\eta) \; d\eta}\;, \end{equation} where \begin{equation} {\cal P}(\eta) = \frac{1}{\sigma \sqrt{2\pi}}\; \exp \left( -\, \frac{\chi^2(\eta) - \chi^2_{\rm min} }{2} \right ) \;. \end{equation} {}From $\eta_+$ we find the limits as $M_S/{\cal F}^{1/4} = \eta_+^{-1/4}$. The limits on $M_S/{\cal F}^{1/4}$ for each data set and the combined are also shown in Table \ref{fit}. The combined limit is $M_S/{\cal F}^{1/4} > 939$ GeV, or \begin{equation} \label{result} M_S > \Biggr\{ \begin{array}{cc} 1120 \; {\rm GeV}\;\; &{\rm for}\;\; n=3 \\ 939 \; {\rm GeV}\;\; & {\rm for}\;\; n=4 \\ \end{array} \end{equation} at 95\%CL. It is the Drell-yan production at the Tevatron that gives the strongest constraint among all data. This is easy to understand because the $\hat s$ of the Drell-yan process is the largest of all. The new HERA data are now consistent with the SM and only cause a very slight pull of the fit, about one third of a $\sigma$ from zero: see Table \ref{fit}. Both the ``ALL w/o HERA'' and the ``ALL'' fits agree well with the SM. We show in Fig. \ref{fig1} the Drell-yan production for the SM and for the low scale gravity model with $M_S$ and $n$ given in Eq. (\ref{result}), together with the D0 data. To conclude, we have used the global lepton-quark neutral current data plus the leptonic cross sections at LEPII to constrain the low scale gravity interactions arising from the extra dimensions. The limit that we place on the effective Planck scale $M_S$ is about 1.12 TeV for $n=3$ and 0.94 TeV for $n=4$ at 95\%CL. \section*{\bf Acknowledgments} This research was supported in part by the U.S.~Department of Energy under Grants No. DE-FG03-91ER40674 and by the Davis Institute for High Energy Physics.
\section{Introduction} \label{intro} There is no longer any controversy about the physical relevance of noise effects in spatially extended, nonlinear systems \cite{Bishop,GSbook}: Indeed, the pervasive, joint role of nonlinearity and (static or dynamic) disorder has already been recognized in biophysics, electronics, optics, fluids, condensed matter, computational physics, etc. In most of these fields, nonlinear phenomena involve nonlinear coherent excitations, such as solitons or solitary waves, which play a key part in the corresponding system dynamics. It is because of this nowadays well established fact that much effort has been devoted to understand how stochastic perturbations affect solitons, mostly during the decade of the 80's (see \cite{Bass,% yo1,KVbook} for reviews). In fact, early numerical simulations \cite{early} already revealed that $\phi^4$ solitary waves underwent Brownian-like motion in the presence of additive white noise, i.e., of thermal fluctuations. Subsequent works focused on the study of soliton diffusion, since it may be crucial in a number of problems, such as photoexcitation dynamics, photoconductivity of conducting polymers, or transport by phase solitons in charge-density-wave systems, to name a few \cite{aplic}. Among the different soliton-bearing nonlinear models which have been studied in the above context, one which has received a great deal of attention is the sG equation. The interest in this model is both theoretical, as it displays the main features of more realistic and complicated cases while remaining analytically tractable, and applied, as it very approximately describes the dynamics of many physically relevant systems, such as one-dimensional (1D) magnets \cite{mikeska} or long Josephson junctions \cite{barone}, for instance. Soliton diffusion governed by the sG (and other nonlinear Klein-Gordon equations) has been studied along two main, different lines which are discussed and compared, e.g., in Ref. \cite{ivanov}. The first one consists of considering extended excitations of the system (phonons) in equilibrium with both a single sG soliton and a heat bath at temperature $T$. This approach leads to two distinct diffusion regimes: anomalous diffusion, characterized by a diffusion constant proportional to $T^2$, and viscous diffusion, when the appearance of a dynamical damping coefficient yields a diffusion constant proportional to $T^{-1}$. We will not follow this approach here; the interested reader is referred to the detailed review by Y.\ Wada \cite{wada}. The second manner is {\em \`a la} Langevin, i.e., introducing the influence of an external thermal bath by means of local fluctuations of the string and a local damping force related to that by an appropriate fluctuation-dissipation relationship. The corresponding equation of motion is then \begin{equation} \phi_{tt} - \phi_{xx} + \sin(\phi) = -\alpha \phi_{t} + \eta(x,t), \label{ecua1} \end{equation} with \begin{mathletters} \begin{eqnarray} \langle\eta(x,t)\rangle &=& 0, \\ \langle\eta(x,t)\eta(x',t')\rangle &=& D \delta(x-x') \delta(t-t'), \end{eqnarray} \noindent \label{ecua3} \end{mathletters} \noindent where the diffusion coefficient $D=2 \alpha k_{b} T$, $k_{b}$ being the Boltzmann constant, and $-\alpha \phi_{t}$ being the damping term with a dissipation coefficient $\alpha$. This equation has been considered a number of times in the literature (see, e.g., \cite{Bass} and references therein; see also \cite{cast} for related experimental work). In this work, we focus on the Langevin version of the problem, with the aim of improving the analytical results obtained in the aforementioned works as well as of verifying them by numerical simulations specifically planned to that end. Furthermore, we concern ourselves with the overdamped limit of the sG equation, which reads \begin{equation} \alpha \phi_{t} - \phi_{xx} + \sin(\phi) = \epsilon\eta(x,t,\phi_,...). \label{ecua2} \end{equation} Note that we have introduced a factor $\epsilon$ in front of the noise term for convenience in the analytical calculations in section II. This equation (without noise, $\epsilon=0$) was already considered by Eilenberger in \cite{bre}, as the limit of the sG equation (\ref{ecua1}) in the case when the dissipation effect is strong enough in Eq.\ (\ref{ecua1}) and there is an input of energy into the system (see, e.g., \cite{km,bennet} and references therein). On the other hand, Eq.\ (\ref{ecua2}), with additive noise as in (\ref{ecua3}), is interesting in itself: For example, it has been proposed as a model for crystal growth (see \cite{ks,ancai,rangel} and references therein). Equation (\ref{ecua2}) has also been studied to analyze the kink contribution to transport properties when the system is driven and thermally activated \cite{ks,butland1,butland2,kaup}. In particular, the work of Kaup \cite{kaup} is the most closely related to the present one, as it presents a singular perturbation theory to compute the first-order (in $T$) correction to the kink mobility as well as the change of its shape. However, to our knowledge the free diffusion problem for the overdamped sG equation has not been adressed in the literature to date and, therefore, we believe that our results will be interesting by themselves. On the other hand, we also hope that what we learn in this case can be used towards obtaining a more complete, accurate picture of the full sG problem; we will discuss this question in the conclusions. The outline of the paper is as follows: In section II, using a general perturbative method \cite{bre} which we recall in detail, we calculate the correlation functions of the position and the velocity of the kink center up to second-order in $k_{b} T$, as well as the diffusion coefficient and the mean value $\langle\phi(x,t)\rangle$ for fixed $t$. In section III we numerically integrate the stochastic partial differential equation (\ref{ecua2}), with noise given by Eq.\ (\ref{ecua3}), using the Heun scheme \cite{maxra} and compute the time correlation function of the position of the kink center and the diffusion coefficient. We compare these results with the theoretical ones obtained in section II and find an excellent agreement. Finally, in section IV we discuss our results, summarize our main conclusions, and sketch lines for future research. \section{A general perturbative approach} Following the {\em Ansatz} proposed in \cite{bre,raj}, we assume that the solution of Eq.\ (\ref{ecua2}) can be expanded as \begin{equation} \phi(x,t)=\phi_{0}[x-X(t)] + \int_{-\infty}^{+\infty}dk\,A_{k}(t) f_{k}[x-X(t)], \label{ecua4} \end{equation} where $f_{k}[x-X(t)]$ are the eigenfunctions of the linearized version of Eq.\ (\ref{ecua2}) [with $\epsilon=0$], which along with $\displaystyle f_{T} [x-X(t)] = \frac{\partial{\phi_{0}}}{\partial{x}}[x-X(t)]$, form a complete set of orthogonal eigenfunctions (see Appendix I). The first term in the expansion (\ref{ecua4}) represents the translational mode related to the position of the kink center $X(t)$, whereas the second one characterizes the phonon modes (linear excitations around a kink) of the system. We will focus on the kink center motion as described by $X(t)$, as it is well established that such a particle-like picture is very generally enough to describe the behavior of the kink as a whole ($X$ playing the r\^ole of a collective coordinate; see, e.g., \cite{siam} for a review). In order to calculate the dynamics of the kink center, we begin by inserting (\ref{ecua4}) in (\ref{ecua2}), and projecting on the orthogonal basis [see Appendix I, relationships (\ref{ap6})] we obtain a system of differential equations for the unknown functions $X(t)$ and $A_{k}(t)$: \begin{eqnarray} \dot{X}(t) & = & - \frac{1}{8} \dot{X}(t) \int_{-\infty}^{+\infty}dk A_{k}(t) I_{1}(k) - \frac{1}{16 \alpha} \int_{-\infty}^{+\infty} dk \int_{-\infty}^{+\infty}dk' A_{k}(t) A_{k'}(t) R_{3}(k,k') + \nonumber \\ & + & \frac{\sqrt{D}}{8 \alpha} \int_{-\infty}^{+\infty} f_{T} [x - X(t)] \,\ \eta(x,t) \,\ dx - \nonumber \\ & -& \frac{1}{48 \alpha} \int_{-\infty}^{+\infty} dk \int_{-\infty}^{+\infty} dk_{1} \int_{-\infty}^{+\infty} dk_{2} A_{k}(t) A_{k_{1}}(t) A_{k_{2}}(t) R_{6}(k,k_{1},k_{2}),\label{ecua5}\\ \frac{\partial{A_{k}}}{\partial{t}} + \frac{\omega_{k}^{2}}{\alpha} A_{k}(t) & = & \dot{X}(t) \int_{-\infty}^{+\infty} dk A_{k}(t) I_{3}(k,k') + \frac{1}{2 \alpha} \int_{-\infty}^{+\infty} dk \int_{-\infty}^{+\infty} dk' A_{k}(t) A_{k'}(t) R_{4}(k,k') - \nonumber \\ & - & \frac{\sqrt{D}}{\alpha} \int_{-\infty}^{+\infty} f_{k'}^{*} [x - X(t)] \,\ \eta(x,t) \,\ dx + \nonumber \\ & + & \frac{1}{6 \alpha} \int_{-\infty}^{+\infty} dk \int_{-\infty}^{+\infty} dk_{1} \int_{-\infty}^{+\infty} dk_{2} A_{k}(t) A_{k_{1}}(t) A_{k_{2}}(t) R_{7}(k,k_{1},k_{2}), \label{ecua6} \end{eqnarray} where \begin{eqnarray} I_{1}(k) & = & \int_{-\infty}^{+\infty} \frac{\partial f_{k}}{\partial \theta} f_{T}(\theta) d\theta = \frac{i \pi \omega_{k}}{\sqrt{2 \pi} \,\ \cosh\Big(\displaystyle{\frac{\pi k}{2}}\Big)}, \nonumber \\ R_{3}(k,k') & = & \int_{-\infty}^{+\infty} f_{T}(\theta) \frac{\partial f_{T}}{\partial \theta} f_{k}(\theta) f_{k'}^{*}(\theta) d\theta = -\frac{i (\omega_{k}^{2} - \omega_{k'}^{2})^{2}}{4 \omega_{k} \omega_{k'} \,\ \sinh\Big(\displaystyle{\frac{\pi \Delta k}{2}}\Big)},\,\ \Delta k=k'-k, \nonumber \\ I_{3}(k,k') & = & \int_{-\infty}^{+\infty} \frac{\partial f_{k}}{\partial \theta} f_{k'}^{*}(\theta) d\theta, \nonumber \\ R_{4}(k,k') & = & \int_{-\infty}^{+\infty} [f_{k'}^{*}(\theta)]^{2} \frac{\partial f_{T}}{\partial \theta} f_{k}(\theta) d\theta, \,\ R_{4}(k,k) = \frac{3 i \omega_{k}}{8 \sqrt{2 \pi} \,\ \cosh\Big(\displaystyle{\frac{\pi k}{2}}\Big)}, \nonumber \\ R_{6}(k,k_{1},k_{2}) & = & \int_{-\infty}^{+\infty} \frac{\partial^{2} f_{T}}{\partial \theta^{2}} f_{k}(\theta) f_{k_1}^{*}(\theta) f_{k_2}(\theta) d\theta, \,\ \nonumber \\ R_{7}(k,k_{1},k_{2}) & = & \int_{-\infty}^{+\infty} \cos(\phi_{0}) f_{k'}^{*}(\theta) f_{k}(\theta) f_{k_1}^{*}(\theta) f_{k_2}(\theta) d\theta. \label{ecua7} \end{eqnarray} We now recall that, if we set $\epsilon = 0$ in (\ref{ecua2}), the static kink is an exact solution; hence, in what follows we will consider $\epsilon$ as a small perturbative parameter, and expand $A_{k}(t)$ and $X(t)$ in powers of $\epsilon$. By substituting the series $A_{k}(t) = \sum_{n=1}^{\infty} \epsilon^{n} A_{k}^{n}(t)$ and $X(t) = \sum_{n=1}^{\infty} \epsilon^{n} X_{n}(t)$ in (\ref{ecua5}) and (\ref{ecua6}) we find a set of linear equations for the coefficients of these series. We only write down here the systems of equations up to order $\epsilon^{3}$, leaving out the cumbersome (albeit straightforward) equation for $A_{k}^{3}(t)$: \underline{$O(\epsilon)$} \begin{mathletters} \begin{eqnarray} \dot{X}_{1}(t) & = & \epsilon_{1}(t), \,\ \langle \epsilon_{1}(t)\rangle =0, \,\ \langle \epsilon_1(t) \epsilon_1(t')\rangle = \frac{D}{8 \alpha^{2}} \delta(t-t'), \label{ecua8}\\ \frac{\partial {A_{k}^{1}}}{\partial t}(t) + \frac{\omega_{k}^{2}}{\alpha} A_{k}^{1}(t) & = & \frac{\epsilon_{k}(t)}{\alpha}, \,\ \langle\epsilon_{k}(t)\rangle =0, \,\ \langle\epsilon_{k}(t) \epsilon_{k'}(t')\rangle = \frac{D}{\alpha^{2}} \delta(t-t') \delta(k-k'); \label{ecua9} \end{eqnarray} \end{mathletters} \underline{$O(\epsilon^2)$} \begin{mathletters} \begin{eqnarray} \dot{X}_{2}(t) & = & -\frac{\dot{X}_{1}(t)}{8} \int _{-\infty}^{+\infty} dk A_{k}^{1}(t) I_{1}(k) - \nonumber \\ & - & \frac{1}{16 \alpha} \int_{-\infty}^{+\infty} dk \int_{-\infty}^{+\infty} dk' A_{k}^{1}(t) A_{k'}^{1}(t) R_{3}(k,k'), \label{ecua10}\\ \frac{\partial {A_{k}^{2}}}{\partial t}(t) + \frac{\omega_{k}^{2}}{\alpha} A_{k}^{2}(t) & = & \dot{X}_{1}(t) \int_{-\infty}^{+\infty} dk A_{k}^{1}(t) I_{3}(k,k) + \nonumber \\ & + & \frac{1}{2 \alpha} \int_{-\infty}^{+\infty} dk \int_{-\infty}^{+\infty} dk' A_{k}^{1}(t) A_{k'}^{1}(t) R_{4}(k,k'); \label{ecua11} \end{eqnarray} \end{mathletters} \underline{$O(\epsilon^{3})$} \begin{eqnarray} \dot{X}_{3}(t) & = & -\frac{\dot{X}_{1}(t)}{8} \int_{-\infty}^{+\infty} dk A_{k}^{2}(t) I_{1}(k) - \frac{\dot{X}_{2}(t)}{8} \int_{-\infty}^{+\infty} dk A_{k}^{1}(t) I_{1}(k) - \nonumber \\ & - & \frac{1}{16 \alpha} \int_{-\infty}^{+\infty} dk \int_{-\infty}^{+\infty} dk' A_{k}^{2}(t) A_{k'}^{1}(t) R_{3}(k,k') - \nonumber \\ & - & \frac{1}{16 \alpha} \int_{-\infty}^{+\infty} dk \int_{-\infty}^{+\infty} dk' A_{k}^{1}(t) A_{k'}^{2}(t) R_{3}(k,k') - \nonumber \\ & - &\frac{1}{48 \alpha} \int_{-\infty}^{+\infty} dk \int_{-\infty}^{+\infty} dk_{1} \int_{-\infty}^{+\infty} dk_{2} A_{k}^{1}(t) A_{k_{1}}^{1}(t) A_{k_{2}}^{1}(t) R_{6}(k,k_{1},k_{2}). \label{ecua12} \end{eqnarray} We now proceed with the first-order equations. The solutions of (\ref{ecua8}) and (\ref{ecua9}) can be written as \begin{eqnarray} X_{1}(t) & = & \int_{0}^{t} \epsilon_{1}(\tau) d \tau, \quad A_{k}^{1}(t) = \exp(-\frac{\omega_{k}^{2} \,\ t}{\alpha}) \int_{0}^{t} \exp(\frac{\omega_{k}^{2} \tau}{\alpha}) \epsilon_{k}(\tau) d \tau, \label{ecua13} \end{eqnarray} respectively. From these relations we can immediately compute averages over the quantities of interest, such as \begin{eqnarray} \langle X_{1}(t)\rangle & = & 0, \,\ \langle X_{1}(t) X_{1}(t')\rangle = \frac{D}{8 \alpha^{2}} M, \label{ecua14}\\ \langle \dot{X}_{1}(t)\rangle & = & 0, \,\ \langle \dot{X}_{1}(t) \dot{X}_{1}(t')\rangle = \frac{D}{8 \alpha^{2}} \delta(t-t'),\label{ecua15}\\ \langle A_{k}^{1}(t)\rangle & = & 0, \quad \langle A_{k}^{1}(t) A_{k}^{1}(t')\rangle = \frac{D}{2 \alpha \omega_{k}^{2}} \Big[\exp\Big(-\frac{\omega_{k}^{2} |t'-t|}{\alpha}\Big) - \exp\Big(-\frac{\omega_{k}^{2} (t+t')}{\alpha}\Big)\Big], \label{ecua16} \end{eqnarray} where $M=min(t,t')$. For the next orders, the calculations are more involved but not difficult. After some tedious algebra, from Eqs.\ (\ref{ecua10})-(\ref{ecua12}) we find the average values of the position and velocity of the kink center \begin{eqnarray} \langle X_{2}(t)\rangle & = & 0, \,\langle \dot{X}_{2}(t)\rangle = 0, \label{ecua17}\\ \langle X_{3}(t)\rangle & = & 0, \,\langle \dot{X}_{3}(t)\rangle = 0; \label{ecua18} \end{eqnarray} whereas it can be shown that, for large enough times, \begin{mathletters} \begin{eqnarray} \langle |A_{k}^{2}(t)|\rangle &\sim & \frac{3 \sigma k_{b} T}{16 \sqrt{2 \pi} \omega_{k}^{2}},\\ \sigma &=& \int_{-\infty}^{+\infty} \frac{dk}{\omega_{k} \cosh\Big(\displaystyle{\frac{\pi k}{2}}\Big)} \approx 1.62386. \end{eqnarray} \end{mathletters} The corresponding correlation functions for $X_{2}(t)$ and $\dot{X}_{2}(t)$ are \begin{eqnarray} \langle X_{2}(t) X_{2}(t')\rangle & = & \frac{D^{2} M}{512 \alpha^{3}} + \frac{D^{2} \pi}{4096 \alpha^{2}} \int_{-\infty}^{+\infty} \frac{\Big[\exp\Big(-2 \omega_{k}^{2} M/\alpha \Big)-1 \Big] dk}{\omega_{k}^{2} \cosh^{2}\displaystyle{\Big(\frac{\pi k}{2}\Big)}}, \label{ecua19}\\ \langle \dot{X}_{2}(t) \dot{X}_{2}(t')\rangle & = & \langle \dot{X}_{1}(t) \dot{X}_{1}(t')\rangle \times \nonumber \\ & \times & \frac{D \pi}{256 \alpha} \int_{-\infty}^{+\infty} \frac{\exp\Big(- \omega_{k}^{2} |t'-t|/\alpha \Big) - \exp\Big(- \omega_{k}^{2} (t'+t)/\alpha \Big) dk} {\cosh^{2}\displaystyle{\Big(\frac{\pi k}{2}\Big)}}. \label{ecua20} \end{eqnarray} Notice that the cross correlation function of $X_{1}(t)$ and $X_{3}(t)$ is of the same order as $\langle X_{2}(t) X_{2}(t')\rangle $, and also that $\langle X_{1}(t) X_{2}(t')\rangle =0$. So, from Eqs.\ (\ref{ecua8}) and (\ref{ecua12}) we have \begin{equation} \begin{array}{l} \displaystyle {\langle X_{3}(t) X_{1}(t')\rangle = \langle X_{2}(t) X_{2}(t')\rangle - \frac{D^{2}}{256 \alpha^{3}} \int_{-\infty}^{+\infty} dk \,\ I_{1}(k)} \times \\ \\ \displaystyle {\times \Big\{ \Big( \int_{-\infty}^{+\infty} dm \frac{R_{4}(m,m)}{\omega_{m}^{2}} \Big) \,\ \Big[\frac{M}{\omega_{k}^{2}} + \frac{\alpha (\exp(-\omega_{k}^{2} M/\alpha)-1)}{\omega_{k}^{4}} \Big]} - \\ \\ \displaystyle {- \int_{-\infty}^{+\infty} dn \frac{R_{4}(n,n)}{\omega_{n}^{2}} \,\ \frac{\alpha}{2 \omega_{n}^{2}-\omega_{k}^{2}} \Big[\frac{(\exp(-2 \omega_{n}^{2} M/\alpha)-1)}{2 \omega_{n}^{2}} - \frac{(\exp(-\omega_{k}^{2} M/\alpha)-1)}{\omega_{k}^{2}}\Big] \Big\}}, \label{ecua21} \end{array} \end{equation} \begin{equation} \begin{array}{l} \displaystyle {\langle \dot{X}_{3}(t) \dot{X}_{1}(t')\rangle = - \frac{1}{8} \langle \dot{X}_{1}(t) \dot{X}_{1}(t')\rangle \times } \,\ \nonumber \\ \displaystyle{\times \int_{-\infty}^{+\infty} dk\Big \{ \langle A_{k}^{2}(t)\rangle I_{1}(k) - \frac{1}{8} \langle [A_{k}^{1}(t)]^{2}\rangle |I_{1}(k)|^{2} \Big \}}. \label{ecua22} \end{array} \end{equation} Finally, from (\ref{ecua15}), (\ref{ecua20}) and (\ref{ecua22}) we obtain the final result, namely that for large $t$ [i.e., taking the limit $t\to\infty$ in Eqs.\ (\ref{ecua20}) and (\ref{ecua22}) in all terms except those related to $X_1(t)$] the correlation function $\langle \dot{X}(t) \dot{X}(t')\rangle $ is given up to order $\epsilon^{4}$ by \begin{eqnarray} \langle \dot{X}(t) \dot{X}(t')\rangle & = & \epsilon^{2} \langle \dot{X}_{1}(t) \dot{X}_{1}(t')\rangle + \nonumber \\ & + & \epsilon^{4} (\langle \dot{X}_{2}(t) \dot{X}_{2}(t')\rangle + \langle \dot{X}_{1}(t) \dot{X}_{3}(t')\rangle + \langle \dot{X}_{3}(t) \dot{X}_{1}(t')\rangle )+... \nonumber \\ & = & \frac{\epsilon^{2}}{8} \langle \dot{X}_{1}(t) \dot{X}_{1}(t')\rangle \Big\{ 1+ \epsilon^{2} \Big(\frac{3}{32} + \frac{3}{128} \sigma^{2}\Big) k_{b} T \Big \} + o(\epsilon^{4}). \label{ecua23} \end{eqnarray} We now return to our original equation notation: We set $\epsilon$ equal to one and consider $\sqrt{k_{b} T}$ as the small parameter. When $t$ goes to infinity and imposing $\epsilon=1$, from Eqs.\ (\ref{ecua14}), (\ref{ecua19}) and (\ref{ecua21}) we find \begin{eqnarray} \langle [X(t)]^{2}\rangle & = & \frac{k_{b} T}{4 \alpha} t \Big\{ 1+ \Big(\frac{3}{32} + \frac{3}{128} \sigma^{2}\Big) k_{b} T \Big \}. \label{corrx} \end{eqnarray} Note that the slope of this function is the kink diffusion coefficient, so if one takes into account the second-order correction one obtains that the diffusion coefficient is a quadratic function of the temperature. We postpone our comments to Sec. IV, where a comparison with the previously available results will be made. To complete this work, we can calculate in a very simple way the average value of the wave function $\phi(x,t)$ in first order: {}From Eq.\ (4) we have that \begin{equation} \langle \phi(x,t)\rangle =\langle \phi_{0}[x-\epsilon X_1(t)]\rangle + O(\epsilon^{2}). \label{avphi} \end{equation} In this last relation we have taken into account that $\langle A_{k}(t)\rangle =\epsilon \langle A_{k}^{1}(t)\rangle + O(\epsilon^{2})$ and $\langle A_{k}^{1}(t)\rangle =0$ [see Eq.\ (\ref{ecua16})]. If we solve the corresponding Fokker-Planck equation for $X_{1}$ [see Eq.\ (\ref{ecua8})], we obtain that the probability distribution function for $X_{1}$ is a Gaussian function given by \begin{equation} p(X_{1})=\sqrt{\frac{4 \alpha^{2}}{\pi t D}} \exp\Big(-\frac{4 \alpha^{2} X_{1}^{2} }{D t} \Big). \label{pdf} \end{equation} So, one can define the average value $\langle \phi(x,t)\rangle $ as \begin{equation} \langle \phi(x,t)\rangle = \int_{-\infty}^{+\infty} dX_{1} p(X_{1}) \phi_{0}[x-\epsilon X_{1}(t)]. \label{inte} \end{equation} Unfortunately we have not found the analytical expression for this integral. But we have calculated it numerically, and below we will compare it to the simulations for the full partial differential equation. \section{Numerical simulations} For our numerical simulations of the partial differential equation (\ref{ecua2}), we have used the Heun method \cite{maxra}, whose stochastic properties are well known and suitable for comparison to our theoretical predictions. We numerically integrate Eqs.\ (\ref{ecua2}), with white noise (\ref{ecua3}), starting from an unperturbed kink at rest and taking the average values over 1000 realizations. The other parameters are $\alpha=1$, $\Delta x = 0.05$, and $\Delta t = 0.001$. In the evaluation of the simulations, we have defined the center of the kink as follows: We first find all the lattice points $i$ such that $\phi_{i} \le \pi$ and $\phi_{i+1} \ge \pi$ or vice versa. We then interpolate to obtain the points $x_{i}$ where the field $\phi$ crosses $\pi$. In case that, due to the noise-induced deformation of the kink, there is more than one such $x_{i}$, we average them to finally obtain the numerical kink center position, $x_{c}$. As discussed below, this introduces some error, but other alternatives we tested (such as the center of mass, for instance) gave results which did not really represent the kink location, and moreover its calculation from numerics is much less accurate. Once the center is obtained, we computed also its variance $\langle [X(t)]^{2}\rangle $. \begin{figure} \begin{center} \hspace{-2.4cm} \epsfig{file=fig1a.eps, width=2.7in, angle=-90} \epsfig{file=fig1b.eps, width=2.7in, angle=-90}\\ \hspace{-2.4cm} \epsfig{file=fig1c.eps, width=2.7in, angle=-90} \epsfig{file=fig18.eps, width=2.7in, angle=-90}\\ \ \\[15mm] \caption[]{Simulations with initial condition given by a static kink initially located at $X(0)=0$, and subject to a thermal bath. As a continuous (but wiggly) line, we have plotted $\langle [X(t)]^{2}\rangle $ as obtained by numerical integration of Eq.\ (\ref{ecua2}) for (a) $k_{b} T=0.2$, (b) $k_{b} T=0.4$, (c) $k_{b} T=0.6$ and (d) $k_{b} T=0.8$. Overimposed to these lines the linear regression of the numerical results for $t\ge30$ is shown (long-dashed line). The solid line is the theoretical prediction $\langle [X(t)]^{2}\rangle $ from Eq.\ (\ref{corrx}); this line practically overlaps with the linear regression in Fig.\ a, b, and c. The first-order result $\langle [X(t)]^{2}\rangle $ from Eq.\ (\ref{ecua14}) is shown as a dot-dash line.} \label{graph1} \end{center} \end{figure} Figures \ref{graph1}(a)-(d) show a comparison of our numerical results with the analytical predictions, Eqs.\ (\ref{ecua14}) and (\ref{corrx}), for different values of $k_{b} T$. We see that there is an excellent agreement between theory and numerics except for the highest value of $k_bT$ [Fig.\ \ref{graph1}(d)]. We have checked that this disagreement arises from the way we compute the kink center: For such large values of the noise, points where $\phi(x,t)=\pi$ are found all over the system, irrespective of their distance to the kink center (we note, however, that the temperature was not high enough to create new kink-antikink pairs). Those points contribute to the center position through our averaging procedure, and in fact their contribution can be shown to be additive, i.e., it amounts to move the whole curve $\langle [X(t)]^{2}\rangle $ upwards. This is indeed what occurs in Fig.\ \ref{graph1}(d), and as we will see below the slope is very close to the predicted one. The same behavior is found for higher temperatures in so far no new kinks are created (not shown). Interestingly, a first conclusion that can be drawn from these figures is that already for not so high temperatures, $k_bT\ge 0.4$, as time passes the kink behavior becomes more and more different from the first-order prediction, showing clearly the necessity for the second-order correction. \begin{figure} \begin{center} \hspace{-2.4cm} \epsfig{file=figu2.eps, width=5.0in, angle=-90} \ \\[15mm] \caption{Lower solid line: the function $D_{1}$; upper solid line: $D_{2}$, which represent the first- and second-order results for the kink diffusion coefficient [see Eqs.\ (\ref{ecua14}) and (\ref{corrx})]. Diamonds represent the numerical values of the kink diffusion coefficient, obtained by numerical integration of Eq.\ (\ref{ecua2}) with final time $t_{f}=200$ (as in Fig.\ \ref{graph1}) and different values of $k_{b} T$. A quadratic regression of these numerical values is also plotted (long-dashed line).} \label{graph2} \end{center} \end{figure} We have calculated the numerical values of the diffusion coefficient for several temperatures by taking the slope of $\langle [X(t)]^{2}\rangle $, which we obtain from a linear fit of the data for not so early times ($t \ge 30$) to avoid transient effects coming from the adjustment of the kink to the heat bath. Note also that our prediction for the second-order contribution was obtained in the large-time limit, so we should not try to fit the whole evolution. The figures also show those linear regressions. Subsequently, in Fig.\ \ref{graph2} we have compared the computed slopes with the first and second-order coefficients $\displaystyle{D_{1}=\frac{k_{b} T}{4 \alpha}}$, and $\displaystyle{D_{2}=\frac{k_{b} T}{4 \alpha} \Big\{ 1+ \Big(\frac{3}{32} + \frac{3}{128} \sigma^{2}\Big) k_{b} T \Big \}}$ [see. Eqs.\ (\ref{ecua14}) and (\ref{corrx})]. The comparison is once again very good, and points out very clearly that for values of $k_bT$ as low as 0.3, the first-order prediction begins to deviate from the diffusion constant measured in the simulations. In addition, the quadratic fit to the simulation results, shown as a long-dashed line in Fig.\ \ref{graph2}, practically coincides with the second-order prediction in the whole studied range. As a final verification of our results, in Fig.\ \ref{graph3} we have plotted the mean value $\langle \phi(x,t) \rangle$ of the wave function at three different times along its evolution, both as obtained from the numerical simulation of the partial differential equation and {}from the numerical evaluation of Eq.\ (\ref{inte}). The perfect agreement between these expressions provides us with a hint to derive an approximate analytical estimate of the evolution of $\langle \phi \rangle$ {}from the integral (\ref{inte}). {}From Fig.\ \ref{graph3} one immediately concludes that the width of $\langle \phi \rangle$ increases with temperature and time. Let us define the width of $\langle \phi \rangle$ by \begin{figure} \begin{center} \hspace{-2.4cm} \epsfig{file=k400m.eps, width=2.7in, angle=-90} \epsfig{file=k800m.eps, width=2.7in, angle=-90} \ \\[15mm] \caption{Solid lines: Snapshots of the evolution of $\langle \phi(x,t)\rangle $, obtained from numerical simulations of the partial differential equation, for fixed times 40, 120 and 200, respectively. The initial kink (unperturbed, at rest, is also included for comparison). The width of $\langle \phi \rangle $ increases as time progresses. The overimposed points have been computed numerically from the integral (\ref{inte}). Plots correspond to $k_{b} T=0.4$ (a), and 0.8 (b); the width of $\langle \phi \rangle $ is seen to increase also with temperature.} \label{graph3} \end{center} \end{figure} \begin{equation} \displaystyle { L(t)=\sqrt{\frac{\int_{-\infty}^{+\infty} x^{2} \langle [\phi_x(x,t)]^{2}\rangle dx} {\int_{-\infty}^{+\infty} \langle [\phi_x(x,t)]^{2}\rangle dx}}. } \label{anwidth} \end{equation} With this definition, we can now calculate $\langle [\phi_x(x,t)]^2\rangle $ by using the distribution function of $X_{1}(t)$; this procedure yields \begin{equation} \displaystyle { L(t) \approx \sqrt{L_{0}^{2}+\langle [X_{1}(t)]^{2}\rangle }}, \label{eqL} \end{equation} where ${\displaystyle L_{0}^{2}=\frac{\int_{-\infty}^{+\infty} dx \,\ [x^{2}/\rm cosh^{2}(x)]} {\int_{-\infty}^{+\infty} dx \,\ [1/\rm cosh^{2}(x)]}=0.8225}$. It is important to note that, of course, we could define $L(t)$ using $\langle \phi_x(x,t)\rangle $ instead of $\langle [\phi_x(x,t)]^{2}\rangle $ in the above expression, or equivalently another quantity which is localized around the kink center. However, as all possible (and sensible) definitions of $L(t)$ give more or less the same results, the difference between them becomes a constant factor when $\langle [X_{1}(t)]^{2}\rangle $ increases above $L_{0}^{2}$ (for example, for large enough $t$). So, we expect that the ratio \begin{equation} \label{ratio} \frac{L(t)}{L(t_{fix})} \to \sqrt{\frac{t}{t_{fix}}}, \end{equation} for large enough $t$ and $t_{fix}$. Figure \ref{graph4} shows a comparison of this prediction with the numerical evaluation of the width of $\langle \phi \rangle $ from Eq.\ (\ref{inte}). {}From these plots we see that the broadening of $\langle \phi \rangle $ behaves indeed as $\sqrt{t}$: We can compare the analytical slope equal to $0.5$ with the numerical ones equal to $0.4276$ and $0.4517$ for plots (a) and (b) respectively. The slope in (b) is closer to the analytical value due to the fact that $t_{fix}$ is larger than in case (a), which agrees with the above considerations. \begin{figure} \begin{center}\hspace{-2.4cm} \epsfig{file=figu4.eps, width=2.7in, angle=-90} \epsfig{file=figu4a.eps, width=2.7in, angle=-90} \ \\[15mm] \caption{Solid lines: Analytical values of $\rm ln[L(t)/L(t_{fix})]$ for $t_{fix}=40$ (a) and $t_{fix}=80$ (b). In both cases $\alpha=1$ and $k_{b} T=0.6$. Long-dashed lines: numerical values, calculated from (\ref{inte}). The solid lines over the long-dashed ones correspond to the linear regression of the numerical points.} \label{graph4} \end{center} \end{figure} \section{Discussion and conclusions} As we have seen in the previous section, our second-order theoretical predictions constitute a very accurate description of the kink dynamics for a wide range of temperatures, up to a value of $k_bT\simeq 1$. In fact, the range of validity of the analytical results might be somewhat higher, provided a better way to estimate the kink center from the numerical simulations could be devised. In any event, the occurrence of $\pi$-crossings far away from the kink center for values around $k_bT\simeq 1$ indicates that further increments of the temperature would undoubtedly produce kink-antikink pairs, thus invalidating our collective coordinate approach which necessarily relies on the identification of the individual kink propagation. We note that this value is a little over 10\% of the kink rest mass ($M_0=8$ in our units); in this respect, a similar result was obtained in \cite{jacek} for the overdamped $\phi^4$ model by means of a similar perturbative approach (with the caveat that the numerical data presented in \cite{jacek} only allow one to guess what is the range of validity of their results). It is interesting to pursue further the comparison of the results for the sG and $\phi^4$ cases. In our calculations for sG, we have found that the second-order correction is clearly smaller (albeit relevant) than the first-order one. The structure of the perturbative calculation allows to identify the origin of that correction: It comes from the interaction of the phonons (described by the functions $A_k$) with the kink. Now, in the $\phi^4$ case, the situation is quite different: Indeed, the second-order correction is much {\em larger} than the one we find here, and the reason is the so-called internal mode, present for $\phi^4$ kinks and absent in the sG case. The coupling between this internal mode (which has been shown to act as a reservoir of energy available for exchange with the kink translation mode \cite{Campbell}) and the kink motion can be shown, by a careful examination of the calculation in \cite{jacek}, to be responsible for most of the second-order correction, while the phonons produce a second-order term comparable to the one we have found. We thus see that, while the range of validity of the analytical approach is in principle the same in both cases, the physics is certainly different, and in fact the question arises as to the validity of this kind of perturbative calculation for the $\phi^4$ problem in view of the large contribution of the internal mode. This is an interesting question that deserves further analytical and numerical work. Coming back to our results for the sG kink, the fact that the second-order correction is smaller than the first-order term makes us confident that our expansion is likely to be free of problems coming from secular terms. This belief is reinforced by the result that, up to the validity range discussed above and limited by kink-antikink creation phenomena, the second-order result describes very accurately the kink behavior, which deviates very little from the predicted diffusive motion. It is then reasonable to expect higher-order contributions (whose calculation is extremely cumbersome, but feasible in principle) to be negligible, thus yielding our theoretical result as the final one for the kink diffusion in the overdamped sG problem. In this context, it is also important to realize that Eqs.\ (\ref{ecua8})-(\ref{ecua9}), which are only first-order, can also be obtained following the McLaughlin and Scott procedure \cite{McL} (see also \cite{siam}). However, the advantages of the perturbative scheme we have used are, on one hand, that we were able to obtain the next order in the expansion, and on the other hand, we demonstrated that the second order originates in the interaction between phonon and translational modes of the sG kink. A final remark on our results relates to the mean value of the wave function $\langle \phi(x,t)\rangle $ as a function of $t$, that must {\it{not}} be interpreted as the shape of the kink; in contrast to the interpretation in \cite{jacek}. We first note that the width of the kink cannot increase from its value when unperturbed at rest; the sG equation, being Lorentz invariant, implies that the kink width diminishes when in motion, and therefore an increasing of the width would be very difficult to understand on physical grounds. Indeed that is not the case. The broadening of the mean wave function comes in fact from the dispersion of the individual realizations, as is immediately seen from Fig.\ \ref{graph5}. As may be seen all individual realizations show a width comparable to the initial kink width, which agrees with our physical intuition. The observed $\sqrt{t}$ behavior, discussed at the end of the preceding section, is then evidently related to the fact that the variance of the kink position has that behavior too. The correct interpretation of the width of $\langle \phi(x,t)\rangle $ is that it represents the area in which the kink can be located as its diffusive motion progresses. A similar result has been found for multiplicative noise in \cite{armero} (see also \cite{GSbook} and references therein). \begin{figure} \begin{center} \hspace{-2.4cm} \epsfig{file=figu5.eps, width=5.0in, angle=-90} \ \\[15mm] \caption{Average of the wave function for $k_{b}T=0.4$ and $t=200$ obtained from 1000 realizations (wider solid line), compared to the average of only 5 realizations (dot-dashed line). Also represented are 3 of these individual realizations. Note the different slope and width of the average values as compared to individual realizations. } \label{graph5} \end{center} \end{figure} To conclude, we want to stress that our main result is the quadratic dependence of the diffusion constant on the temperature, stemming from the kink-phonon interactions. This has been verified numerically to a high degree of accuracy. We have carried out standard Langevin dynamics simulations following a well grounded procedure, the Heun method, as far as statistical properties are concerned \cite{maxra}. We can thus be sure that what we are dealing with is indeed the dynamics of a sG kink at finite temperature. Therefore, our analytical calculations and our numerical simulations establish firmly the quadratic dependence of the kink diffusion constant on the temperature for the first time. Now the question remains as to the behavior of {\em underdamped} sG kinks. Preliminary calculations \cite{us} seem to indicate that for underdamped sG kinks the second-order correction is of the same order as that found here, which would support the applicability of the previous calculations at least for small temperatures and not too small damping. To date, no detailed comparison with numerical simulations has ever been done to check the importance of the second-order correction. On the other hand, it would be interesting to compare the results of our approach with the theoretical analysis and experiments in \cite{cast}. Such comparison would provide much insight into the importance of second and higher-order corrections in actual physical systems. Work along these lines is in progress \cite{us} \section*{Acknowledgement} We are grateful to Esteban Moro, Grant Lythe, and Jos\'e Mar\'\i a Sancho for discussions. Work at GISC (Legan\'es) has been supported by CICyT (Spain) grant MAT95-0325 and DGES (Spain) grant PB96-0119. Travel between Bayreuth and Madrid is supported by ``Acciones Integradas Hispano-Alemanas'', a joint program of DAAD (Az.\ 314-AI) and DGES. This research is part of a project supported by NATO grant CRG 971090. \bigskip {\em Note added in proof:} After acceptance of this paper, we have implemented an improved algorithm for detecting the kink center in our code. With this new procedure, no spurious contributions (see discussion below Fig.\ \ref{graph1}) to the variance appear: Specifically, Fig.\ \ref{graph1}(d) is largely improved, and the numerical results overlap the theoretical prediction, thus confirming the interpretation we have provided of the discrepancy. A detailed report will be given in \cite{us}. \section*{Appendix I} One class of solutions of (\ref{ecua2}) [with $\epsilon=0$] is represented by a static kink \begin{equation} \phi_{0}(x,t)= 4 \,\ \mbox{arctan}[\exp(x)]. \label{ap1} \end{equation} The perturbations over this equation may be treated by assuming that the solution of (\ref{ecua2}) [with $\epsilon=0$] has the form \begin{equation} \phi(x,t)=\phi_{0}(x) + \psi(x,t), \,\ \,\ \psi(x,t) \ll \phi_{0}(x). \label{ap2} \end{equation} If we substitute Eq. (\ref{ap2}) in (\ref{ecua2}) [with $\epsilon=0$] and linearize around $\phi_{0}(x)$, we obtain the following equation for $\psi(x,t)$ \begin{equation} \alpha \psi_{t} = \psi_{xx} - \Big[1-\frac{2}{\cosh^{2}(x)}\Big] \psi. \label{ap3} \end{equation} Then, the solution of (\ref{ap3}) may be written as $\displaystyle {\psi(x,t) = f_{k}(x) \exp\Big(-\frac{\omega_{k}^{2} \,\ t}{\alpha}\Big)}$, where $f_{k}(x)$ satisfies the eigenvalue problem given by \begin{equation} -\frac{\partial^{2} f_{k}}{\partial x^{2}} + \Big[1-\frac{2}{\cosh^{2}(x)}\Big] f_{k} = \omega_{k}^{2} f_{k}. \label{ap4} \end{equation} This equation admits the following eigenfunctions with their respective eigenvalues \begin{eqnarray} f_{T}(x) = \frac{2}{\cosh(x)}, \,\ \omega_{T}^{2} & = & 0, \\ f_{k}(x) = \frac{\exp(i k x) \,\ [k + i \mbox{tanh}(x)]}{\sqrt{2 \pi} \,\ \omega_{k}}, \,\ \omega_{k}^{2} & = & 1 + k^{2}. \label{ap5} \end{eqnarray} Notice, that $f_{T}(x)$ and $f_{k}(x)$ form a complete set of functions with the orthogonality relations \begin{mathletters} \begin{eqnarray} \int_{-\infty}^{+\infty} f_{T}^{2}(x) \,\ dx & = & 8, \quad \int_{-\infty}^{+\infty} f_{T}(x) f_{k}(x) \,\ dx = 0, \\ \int_{-\infty}^{+\infty} f_{k}(x) f_{k'}^{*}(x) \,\ dx & = &\delta(k-k'). \end{eqnarray} \label{ap6} \end{mathletters}
\section{Introduction} \label{sec:intro} In many terrestrial flows of granular materials, gravity consolidates the medium to a state where sustained frictional contact between the particles is the dominant mode of momentum transfer. In this regime of high solids fraction and low deformation rate, models based on concepts in metal plasticity and soil mechanics have been traditionally used to describe the flow~\citep{jackson83}. While many gross features of granular flows can be predicted using these models, one aspect they fail to capture is the thickness of shear layers; often when granular materials are sheared, large portions of the material do not suffer sustained deformation. In the experiments of \citet{roscoe70, neddermanandlaohakul80, gudehusandtejchman91}, the velocity gradients are confined to layers approximately 5--40 particle diameters in thickness. Moreover, the thickness of the shear layers is influenced by the nature of the boundaries; when the flowing medium is confined by smooth walls, it is found that the thickness of the shear layers is less than that in the case of rough walls \citep{neddermanandlaohakul80, tejchmanandgudehus93}. Conventional models of plasticity do not predict shear layers~\citep{tejchmanandwu93, mohanetal97}. The failure of the frictional models to predict the thickness of the shear layers accurately has been attributed to the absence of a material length scale in their constitutive equations~\citep[cited in~\citealp{muhlhausandvardoulakis87}]{muhlhaus86}. To overcome this deficiency of the classical models, the particle size must be incorporated in the constitutive equations. In the absence of a comprehensive micro-mechanical model to describe friction, a continuum theory that includes a material length scale in the constitutive equations can be constructed by modelling the granular material as a Cosserat continuum~\citep{muhlhaus86}. We shall argue later that the frictional nature of particle interactions provides sufficient grounds for using this approach. We note here that models based on kinetic theory~\citep[see, for example,][]{lunetal84}, involve the particle diameter in the constitutive relations. However, these models are expected to hold only for rapid flows, where particle interactions may be approximated by instantaneous collisions. In this paper, we explore the use of a Cosserat plasticity model to describe steady, fully developed, plane flow of a granular material in a vertical channel under the action of gravity. The predictions of the model will be compared with data reported in the literature. While Cosserat plasticity models have been applied to problems in granular flow in the past \citep{muhlhausandvardoulakis87,muhlhaus89,tejchmanandwu93,% tejchmanandgudehus93,tejchmanandwu94}, these studies address unsteady flows and are posed in terms of strain increments; with this formulation, \citet{tejchmanandgudehus93} found it difficult to integrate the equations numerically for large times. To the best of our knowledge, the present work represents the first attempt to examine steady flow in this context. We indicate below how the model is developed, and then apply it to channel flow. \section{The Cosserat model} \label{sec:cosserat} The field variables of the classical continuum are the density $\rho$, the linear velocity $\Vector{v}$, and the stress tensor $\Tensor{\sigma}$. A Cosserat continuum~(\citep[p.~223]{jaunzemis67}; \citep{cowin74}) involves two additional field variables, namely, the angular velocity $\Vector{\omega}$, and the couple stress tensor $\Tensor{M}$. Considering a Cartesian coordinate system (\fig{1}), $M_{xz}$ represents the couple per unit area exerted about the $z$-axis on a plane~$x = \mathrm{constant}$, by the material to the left of this plane. A positive value of $M_{xz}$ is taken to impose an anti-clockwise rotation on this plane~(\fig{1}). For a Cosserat continuum, the mass and linear momentum balances must be supplemented by the angular momentum balance, which relates $\Vector{\omega}$, $\Tensor{M}$, and $\Tensor{\sigma}$. For steady, fully developed flow, spatial gradients of $\Tensor{M}$ cause $\Tensor{\sigma}$ to be asymmetric. This is in contrast to the classical continuum, which assumes implicitly that there are no couple stresses, body couples, and intrinsic angular momentum; hence the angular momentum balance can be satisfied identically by requiring $\Tensor{\sigma}$ to be symmetric. There is enough analytical evidence to motivate the use of a Cosserat model in the present problem. \citet{dahler59} used a statistical mechanical approach to develop expressions for the stresses in a fluid composed of diatomic molecules. For molecules interacting via central forces, which are directed along the lines joining the centers of mass of the molecules, $\Tensor{\sigma}$ is found to be symmetric. However, his model suggests that non-central forces may cause $\Tensor{\sigma}$ to be asymmetric. \citet{campbell93a} simulated the shearing of circular discs between parallel plates, assuming that the collisions between discs, and between a disc and the wall, were instantaneous. In the latter case, wall roughness was incorporated by imposing (after collision) either (i)~a zero relative velocity between the surface of the disc and the wall, or~(ii)~a zero relative velocity between the center of the disc and the wall. In both cases $\Tensor{\sigma}$ was asymmetric near the wall, and there were non-zero couple stresses. \citet{jenkins89} constructed a micro-mechanical model for an assembly of identical spheres. They found that asymmetric stresses resulted when the distribution of contact normals was anisotropic; however, they secured the symmetry of the stress tensor by suitably enforcing the rotation of particles. Dry friction, the dominant mode of momentum transfer in high-density flows, introduces non-central forces in an inherently complex fashion. Hence, we expect that a micro-mechanical model for dry friction would result in a continuum with asymmetric stresses; such materials can be modelled as Cosserat continua. A satisfactory micro-mechanical model is not yet available, and it is hoped that this issue will be addressed by future investigators. \subsection{Equations of motion} \label{sec:eqnofmotion} It is instructive to write the equations for the case of steady plane flow, and later simplify them for the case of fully developed flow. For flow parallel to the $xy$ plane~(\fig{1}), the velocity field has the following form: \begin{equation} \label{e-1} v_{x} = v_{x} (x, y); \; v_{y} = v_{y} (x, y); \; v_{z} = 0; \; \omega_{x} = \omega_{y} = 0; \; \omega_{z} = \omega_{z} (x, y), \end{equation} where $v_{x}$ and $\omega_{x}$ are the $x$ components of the linear and angular velocity, respectively. A positive value of $\omega_{z}$ is associated with an anti-clockwise rotation about the $z$-axis. The balances for mass and linear momentum are \begin{align} \label{e-5} \pder{}{x} (\nu v_{x}) + \pder{}{y} (\nu v_{y}) &= 0,\\ \label{e-6} \pder{\sigma_{xx}}{x} + \pder{\sigma_{yx}}{y} + \rho_{p}\nu\matder{}\:v_{x} &= 0,\\ \label{e-7} \pder{\sigma_{xy}}{x} + \pder{\sigma_{yy}}{y} + \rho_{p}\nu\matder{}\:v_{y} &= - \rho_{p} \nu g, \end{align} where $\nu$ is the solids fraction or the volume fraction of solids, $\sigma_{ij}$'s are the components of the Cauchy stress tensor, defined in the compressive sense, $\rho_{p}$ is the intrinsic density of the particles, assumed constant, and $g$ is the acceleration due to gravity. Following~\citet[p.~233]{jaunzemis67}, the $z$ component of the angular momentum balance is \begin{equation} \label{e-8} \pder{M_{xz}}{x} + \pder{M_{yz}}{y} - \rho_{p} \nu \zeta_{z} + \sigma_{xy} - \sigma_{yx} + \rho_{p}\nu\matder{}\:\eta_{z} = 0, \end{equation} where $M_{iz}$'s are the couple stresses, $\eta_{z}$ is the $z$ component of the intrinsic angular momentum (per unit volume), and $\zeta_{z}$ is the $z$ component of the body couple acting on the material. To close the above set of equations, constitutive relations for the $\sigma_{ij}$ and $M_{iz}$ are required. \subsection{Constitutive equations} \label{sec:consteqn} \citet{muhlhausandvardoulakis87} and~\citet{tejchmanandwu93} have developed Cosserat plasticity models for studying the development of shear bands in granular flow. In their models, the yield condition and the flow rule were modified to account for the influence of the couple stress and to provide a relation for the angular velocity. We have adapted their model to the present problem. The constitutive equations comprise of a yield condition and a flow rule, which are elaborated below. \subsubsection{Yield condition} Following~\citet[cited in~\citealp{lippmann95}]{besdo74}, \citet{deBorst93}, and~\citet{tejchmanandwu93}, we use a yield condition of the form \begin{equation} \label{e-9} F \equiv \tau - Y = 0, \end{equation} where \begin{equation} \label{e-10} \tau \equiv \mysqrt{\left(a_{1} \sigma_{ij}' \sigma_{ij}' + a_{2} \sigma_{ij}' \sigma_{ji}' + \Frac{1}{(L d_{p})^{2}} M_{ij} M_{ij} \right)}, \end{equation} $\sigma_{ij}' = \sigma_{ij} - \textfracp{1}{3}\,\sigma_{kk}\, \delta_{ij}$ is the deviatoric stress, $\delta_{ij}$ is the Kronecker delta, $a_{1}$, $a_{2}$, and~$L$ are material constants, and $d_{p}$ is the particle diameter. Here $L\,d_{p}$ is a characteristic material length scale; the value of $L$ will be chosen later. \citet{deBorst93} assumes that the yield limit $Y$ depends on the mean stress $\sigma\equiv\textfrac{\sigma_{kk}}{3}$, and a hardening parameter $h$. Here we identify $h$ with the solids fraction $\nu$. \citet[p.~112]{schofieldandwroth68} and~\citet{jackson83} discuss the use of a yield condition of the form $F(\Tensor{\sigma}, \nu) = 0$ in the classical frictional models. The yield condition~\eqref{e-9} with $Y = Y(\sigma, \nu)$ represents an attempt to include couple stresses within this framework. Only two of the three parameters $a_{1}$, $a_{2}$ and $L$ in~\eqref{e-10} are independent, because the third parameter may be absorbed in the definition of $Y(\sigma,\nu)$~(see~\eqref{e-9}). Following \citet{muhlhausandvardoulakis87} we set $a_{1} + a_{2} = \textfrac{1}{2}$, without loss of generality. \citet{tejchmanandwu93} use $\mathrm{A}\equiv\textfrac{a_{2}}{a_{1}} = \textfrac{1}{3}$ and $L = 1$. Here we retain their choice of $\mathrm{A}$ and and treat $L$ as an adjustable parameter, whose value is chosen as described later. \citet{deBorst93} found that changes in the values of $\mathrm{A}$ and $L$ affected the post-peak behaviour of a sample which was sheared between parallel plates. Unfortunately, neither experiments nor satisfactory micro-mechanical models are available to guide the choice of $\mathrm{A}$. Following~\citet{prakashandrao88}, we assume the following form for the yield limit $Y$ \begin{equation} \label{e-11} Y = \sigma_{c}(\nu) \sin \phi \left( n \alpha - (n - 1) \alpha^{\textfracp{n}{(n - 1)}} \right); \quad \alpha\equiv\Frac{\sigma}{\sigma_{c}(\nu)}. \end{equation} Here $\sigma_{c}(\nu)$ is the mean stress at a critical state, $\phi$ is a material constant called the angle of internal friction, and $n$ is a material constant. The significance of a critical state will be explained shortly. The dependence of $\sigma_{c}$ on the solids fraction $\nu$ is taken to be~\citep{johnsonandjackson87} \begin{equation} \label{eq:24} \sigma_{c}(\nu) = \begin{cases} 0 &\nu < \nu_{\min},\\ \Lambda\Frac{(\nu - \nu_{\min})^{p}}{(\nu_{\max} - \nu)^{q}} &\nu_{\min} \leq \nu \leq \nu_{\max}. \end{cases} \end{equation} Here $\Lambda$, $\nu_{\min}$, $\nu_{\max}$, $p$ and $q$ are material constants. Note that $\sigma_{c}(\nu)$ has been chosen to be zero below $\nu_{\min}$, the solids fraction at loose random packing, and to diverge as $\nu$ approaches $\nu_{\max}$, the solids fraction at dense random packing. \subsubsection{Flow rule} \citet{tejchmanandwu93} have used incremental elasto-plastic constitutive equations, which they attribute to~\citet{muhlhaus89}. Elastic effects are ignored in the present work to simplify the analysis. Because we are interested in sustained flow, the plastic strain increments used by~\citet{tejchmanandwu93} are replaced by suitable velocity gradients~\citep{muhlhaus89}. In Cartesian tensor notation, the flow rule is given by \begin{equation} \label{eq:22} E_{ij} \equiv \pder{v_{i}}{x_{j}} + \varepsilon_{ijk}\omega_{k} = \dot{\lambda} \pder{G}{\sigma_{ji}}; \quad H_{ij} \equiv \pder{\omega_{i}}{x_{j}} = \dot{\lambda}\pder{G}{M_{ji}}, \end{equation} where $G(\Tensor{\sigma}, \Tensor{M}, \nu)$ is the plastic potential, $\varepsilon_{ijk}$ is the alternating tensor, and $\dot{\lambda}$ is a scalar factor. We note here that $E_{ij}$ is the sum of the rate of deformation tensor $D_{ij}$ and an objective antisymmetric tensor representing the difference between the spin tensor and the particle spin $\varepsilon_{ijk}\omega_{k}$. $E_{ij}$ and $H_{ij}$ are conjugate to the stress $\sigma_{ji}$ and the couple stress $M_{ji}$, respectively, in the sense that the rate of working per unit volume by the contact forces and couples is given by $-(\sigma_{ji}E_{ij} + M_{ji}H_{ij})$~\citep{muhlhaus89}. In a classical continuum, $\Tensor{M}$ vanishes and $\Tensor{\sigma}$ is symmetric; hence the above expression reduces to $-\sigma_{ji}D_{ij}$, where $D_{ij} = \textfracp{1}{2}\:(\textpder{v_{i}}{x_{j}} + \textpder{v_{j}}{x_{i}})$ denotes a component of the rate of deformation tensor. Lacking detailed information on the plastic potential $G$, we adopt the most commonly used closure, namely the associated flow rule \citep[p.~43]{schofieldandwroth68}: \begin{equation} \label{e-17} G \equiv F = \tau - Y. \end{equation} This form of the flow rule accounts for density changes accompanying deformation. \section{Application to channel flow} \label{sec:problem} For the case of steady, fully developed, plane flow, the velocity field is given by \begin{equation} v_{y} = v_{y}(x); \quad \omega\equiv\omega_{z}(x),\label{e-35} \end{equation} and the other velocity components vanish. Hence the mass balance~\eqref{e-5} is identically satisfied and the balances of linear and angular momentum~\eqref{e-6}--\eqref{e-8} reduce to \begin{gather} \der{\sigma_{xx}}{x} = 0;\quad\der{\sigma_{xy}}{x} = - \rho_{p} \nu g,\label{e-20}\\ \der{m}{x} + \sigma_{xy} - \sigma_{yx} = 0, \label{e-21} \end{gather} where $m \equiv M_{xz}$. It is assumed that the yield condition is satisfied at every point in the channel, so that the factor $\dot{\lambda}$ in~\eqref{eq:22} is always non-zero. In writing~\eqref{e-21}, it is assumed that there are no body couples. \subsection{The stress field} \label{sec:Some-cons-assum} For fully developed flow, it will now be shown that all the normal stresses are equal. Equation~\eqref{eq:22} implies that \begin{equation} \label{e-27} E_{xx} = \pder{v_{x}}{x} = 0 = \Frac{\dot{\lambda}}{6\tau} (2 \sigma_{xx}' - \sigma_{yy}' - \sigma_{zz}') - \Frac{\dot{\lambda}}{3}\pder{Y}{\sigma}. \end{equation} Writing the corresponding equations for $E_{yy}$ and $E_{zz}$ and summing, we get \begin{equation} \label{eq:23} \pder{Y}{\sigma} = 0, \end{equation} or using~(\ref{e-11}) \begin{equation} \label{eq:31} \sigma = \sigma_{c}(\nu); \quad Y = \sigma_{c} \sin \phi. \end{equation} Thus the material is at a critical state or a state of isochoric deformation, because $E_{ii} = \nabla \cdot \Vector{v} = 0$. Comparison of \eqref{e-11} and \eqref{eq:31} shows that the value of $n$ is not relevant. It also follows from~\eqref{e-27}, \eqref{eq:31} and~\eqref{e-20} that \begin{equation} \label{eq:33} \sigma_{xx} = \sigma_{yy} = \sigma_{zz} = \sigma_{c}(\nu) = \mathrm{constant}. \end{equation} Hence the solids fraction does not vary across the width of the channel. Using~\eqref{eq:31} and \eqref{eq:33}, the yield condition~\eqref{e-9} reduces to \begin{equation} \label{eq:34} \mysqrtp{a_{1}(\sigma_{xy}^{2} + \sigma_{yx}^{2}) + 2 a_{2} \sigma_{xy}\sigma_{yx} + \Frac{m^{2}}{{(Ld_{p})}^{2}}} - \sigma_{c}(\nu)\sin\phi = 0. \end{equation} \subsection{Velocity field} \label{sec:Const-relat} The non-trivial equations of the flow rule~\eqref{eq:22} are \begin{eqnarray} E_{xy} & = & \omega = \Frac{\dot{\lambda}}{\tau} \left(a_{1} \sigma_{yx} + a_{2} \sigma_{xy} \right),\label{e-32}\\ E_{yx} & = & \der{v_{y}}{x} - \omega = \Frac{\dot{\lambda}}{\tau} \left(a_{1} \sigma_{xy} + a_{2} \sigma_{yx} \right),\label{e-33}\\ H_{zx} & = & \der{\omega}{x} = \Frac{\dot{\lambda}}{\tau} \Frac{m}{(L d_{p})^{2}}. \label{e-34} \end{eqnarray} On eliminating $\dot{\lambda}$ we get \begin{eqnarray} \der{v_{y}}{x}& = & \Frac{(\mathrm{A} + 1)(\sigma_{xy} + \sigma_{yx})\,\omega}{\sigma_{yx} + \mathrm{A} \sigma_{xy}},\label{e-46}\\ \der{\omega}{x} & = & \Frac{2(\mathrm{A} + 1)\,m\,\omega}{(L d_{p})^{2} (\sigma_{yx} + \mathrm{A} \sigma_{xy})}. \label{e-45} \end{eqnarray} \subsection{Boundary conditions} \label{sec:Boundary-conditions-2} Considering symmetric solutions, we have \begin{equation} \sigma_{xy}(0) = 0;\qquad \omega(0) = 0.\label{e-23} \end{equation} The angular velocity $\omega$ must vanish at the centerline of the channel, because a non-zero value implies a preferred direction of rotation. Equations~\eqref{e-32} and~\eqref{e-23} imply \begin{equation} \label{eq:11} \sigma_{yx}(0) = 0, \end{equation} provided $m(0)$ is bounded, \ie, $\nu<\nu_{\max}$ (see \eqref{e-10}, \eqref{eq:24}, and \eqref{eq:34}). Because $\sigma_{xy}$ and $\sigma_{yx}$ both vanish at the centerline, the yield condition~\eqref{eq:34} implies that the couple stress at the centerline is \begin{equation} \label{eq:12} m(x = 0) = \pm L d_{p}\sigma_{c} \sin \phi. \end{equation} While both roots in~\eqref{eq:12} are mathematical solutions, only the negative root yields a physically reasonable solution. The justification for choosing this root, and the reason for discarding the other are discussed in Appendix~A. At the right hand wall $x = W$ we use the usual friction boundary condition~(\citep{brennenandpearce78}; \citep[p.~40]{nedderman92}) \begin{equation} \label{eq:5} -\Frac{\sigma_{xy}}{\sigma_{xx}} = \tan \delta \quad \text{at} \quad x = W, \end{equation} where $\delta$ is a constant called the angle of wall friction. Using~\eqref{e-20} and~\eqref{eq:33}, \eqref{eq:5} reduces to \begin{equation} \label{eq:35} \Frac{\rho_{p}\nu g W}{\sigma_{c}(\nu)} = \tan\delta \quad \text{at} \quad x = W, \end{equation} which determines the value of $\nu$ for specified values of $W$ and $\delta$. Following~\citet{tejchmanandgudehus93}, we assume that \begin{equation} \label{e-63} v_{y} = -Kd_{p}\,\omega \quad \text{at} \quad x = W, \end{equation} where $K$ is a dimensionless constant which reflects the roughness of the wall. To get a feel for this condition, consider a single spherical particle sliding or rolling down a vertical wall. Let $v_{y}'$ and $\omega{'}$ represent the linear velocity of the center of the particle and its angular velocity about an axis through its center, respectively. If the particle slides without rolling, $\omega{'} = 0$, but $v_{y}'$ is arbitrary. Conversely, if it rolls without slipping $\Modulus{v_{y}'}=\textfracp{d_{p}}{2}\:\Modulus{\omega{'}}$. For the boundary condition~\eqref{e-63}, these limits correspond to $K \rightarrow \infty$ and $K \rightarrow \textfrac{1}{2}$, respectively. Reverting to the continuum, we expect that $K$ will decrease as the wall roughness increases. One more boundary condition is needed to permit determination of all the integration constants. Here we set \begin{equation} \label{eq:26} \omega(x = W) = \omega_{w}, \end{equation} where $\omega_{w}$ is a constant whose value is determined by adjusting either the mass flow rate or the centerline velocity to match the measured value. In experiments, the mass flow rate may be varied within limits by varying the width $2 W_{s}$ of the exit slot at the bottom of the channel. Because we are considering fully developed flow, $W_{s}$ does not occur explicitly in the governing equations, but its influence is incorporated by changing $\omega_{w}$. \section{Solution procedure} \label{sec:solution} Introducing the dimensionless variables \begin{align} \xi & = \Frac{x}{W}; & \epsilon &=\Frac{d_{p}}{W}; & u &= - \Frac{v_{y}}{\mysqrtp{g W}}; \notag\\ \overline{\omega} &= \omega\left(\Frac{W}{g}\right)^{\textfrac{1}{2}}; &\overline{\sigma}_{ij} &= \Frac{\sigma_{ij}}{\rho_{p} g W}; & \overline{m} &= \Frac{m}{\rho_{p} g W d_{p}},\notag \end{align} the balance equations~\eqref{e-20} and~\eqref{e-21} may be rewritten as \begin{align} \der{\overline{\sigma}_{xx}}{\xi} &= 0,\label{e-54}\\ \der{\overline{\sigma}_{xy}}{\xi} &= - \nu,\label{e-55}\\ \epsilon\der{\overline{m}}{\xi} + \overline{\sigma}_{xy} - \overline{\sigma}_{yx} &= 0,\label{e-61} \end{align} where $\nu$ is the constant solids fraction across the channel, and \begin{equation} \label{eq:2} \epsilon\equiv\Frac{d_{p}}{W}. \end{equation} The dimensionless form of the yield condition~\eqref{eq:34} is \begin{equation} \label{eq:8} a_{1}(\overline{\sigma}_{xy}^{2} + \overline{\sigma}_{yx}^{2}) + 2 a_{2} \overline{\sigma}_{xy} \overline{\sigma}_{yx} + \Frac{\overline{m}^{2}}{L^{2}} = (\overline{\sigma}_{c}\sin\phi)^{2}. \end{equation} Here $\overline{\sigma}_{c}(\nu) = \textfrac{\sigma_{c}}{(\rho_{p} g W)}$. The flow rule~\eqref{e-46} and~\eqref{e-45} is given by \begin{align} \der{u}{\xi} &= - \Frac{(1 + \mathrm{A})(\overline{\sigma}_{xy} + \overline{\sigma}_{yx})\,\overline{\omega}} {\overline{\sigma}_{yx} + \mathrm{A} \overline{\sigma}_{xy}},\label{e-59}\\ \der{\overline{\omega}}{\xi} &= \Frac{2(1 + \mathrm{A})\, \overline{m}\,\overline{\omega}}{\epsilon L^{2} (\overline{\sigma}_{yx} + \mathrm{A}\overline{\sigma}_{xy})}.\label{e-60} \end{align} The boundary conditions are: \eqnannounce{at the centerline ($\xi = 0$)} \begin{equation} \overline{\sigma}_{xy} = 0; \quad \overline{\omega} = 0.\label{eq:6} \end{equation} \eqnannounce{at the wall ($\xi = 1$)} \begin{equation} \Frac{\nu}{\overline{\sigma}_{c}(\nu)} = \tan \delta; \quad u = \epsilon K \overline{\omega}.\label{eq:7} \end{equation} Equations~\eqref{eq:11} and~\eqref{eq:12} may be written as \begin{equation} \label{eq:54} \overline{\sigma}_{yx} = 0, \quad \overline{m} = -NL\nu \quad \text{at} \quad \xi = 0, \end{equation} where $N\equiv\textfrac{\sin\phi}{\tan\delta}$, and~\eqref{eq:7} has been used to simplify the second of equations~\eqref{eq:54}. \subsection{Method of solution} \label{sec:Method-solution} For the special case $\mathrm{A} = -1$, an analytical solution may be obtained as discussed in Appendix~A. This shows that $\overline{m}=\mathrm{constant}$, $\overline{\sigma}_{yx}=\overline{\sigma}_{xy}$, and $\overline{\omega}$ and $u - u(0)$ display a power law dependence on $\xi$. (The case $\mathrm{A}<-1$ is discussed in Appendix~A.) We now discuss the case $\mathrm{A}>-1$. Inspection of~\eqref{e-54}--\eqref{eq:54} shows that the stress field is uncoupled from the velocity field. Hence we first integrate the equations~\eqref{e-54},~\eqref{e-55} and~\eqref{eq:6}. Equations~\eqref{e-54} and~\eqref{e-55} may be solved along with the first of boundary conditions~\eqref{eq:6} to get \begin{equation} \label{eq:19} \overline{\sigma}_{xx} = \mathrm{constant}; \quad \overline{\sigma}_{xy} = -\nu\xi. \end{equation} The yield condition~\eqref{eq:8} may be solved for $\overline{\sigma}_{yx}$ to get \begin{equation} \label{eq:56} \overline{\sigma}_{yx} = \mathrm{A}\nu\xi \mp \mysqrtp{(\mathrm{A}^{2} - 1)(\nu\xi)^{2} + 2(\mathrm{A} + 1)\left(N^{2}\nu^{2} - \Frac{\overline{m}^{2}}{L^{2}}\right)}. \end{equation} In our calculations, only the root with the negative sign before the square root term in~\eqref{eq:56} was chosen. The justification for doing so is described in Appendix~A. After substituting this expression for $\overline{\sigma}_{yx}$ in~\eqref{e-61}, we solve the equation along with the boundary conditions~\eqref{eq:54} as an initial value problem by marching from $\xi = 0$ to $\xi = 1$. Equation~\eqref{e-61} is solved numerically using the \textsc{lsoda} routine~\citep{petzold83} from \textsc{odepack} in \textsc{netlib}. It should be noted that the above package estimates $\overline{m}$ at a small distance $\xi_{1}$ from $\xi = 0$ as \begin{gather*} \overline{m}(\xi_{1}) \approx -NL\nu + \der{\overline{m}}{\xi}(0)\, \xi_{1} = -NL\nu, \end{gather*} because~\eqref{eq:54} and~\eqref{e-61} imply that $\der{\overline{m}}{\xi}(0) = 0$. This causes the term under the square root in~\eqref{eq:56} to be negative. To avoid this problem~\eqref{e-61} is integrated numerically from $\xi=\xi_{1}$ to $\xi = 1$, with the initial condition given by \begin{equation*} \overline{m}(\xi_{1}) = -NL\nu + \Frac{1}{2} \dertwo{\overline{m}}{\xi}(0)\,\xi_{1}^{2}, \end{equation*} with $\xi_{1}=10^{-5}$. The use of a smaller value of $\xi_{1}$ does not significantly affect the results. Here $\dertwo{\overline{m}}{\xi}(0)$ is calculated by differentiating~\eqref{e-61} with respect to $\xi$, and using~\eqref{eq:19} and~\eqref{eq:56}. The resulting indeterminate expression is evaluated using the L'Hospital's rule to get \begin{gather} \label{eq:20} \epsilon\dertwo{\overline{m}}{\xi}(0) = (\mathrm{A} + 1)\nu \mp \nu\mysqrtp{\mathrm{A}^{2}-1+ 2(\mathrm{A} + 1)\Frac{N}{L\nu}\dertwo{\overline{m}}{\xi}(0)}. \end{gather} This can be rearranged to get a quadratic equation for $\dertwo{\overline{m}}{\xi}(0)$, and we choose the root that satisfies \begin{gather*} \epsilon\dertwo{\overline{m}}{\xi}(0) \leq (\mathrm{A} + 1)\nu, \end{gather*} as the other root is inconsistent with~\eqref{eq:a2}. Once the stresses are obtained, the velocities are calculated by integrating the flow rule~\eqref{e-59}) and~\eqref{e-60} from $\xi = 1$ to $\xi = 0$ using the initial conditions~\eqref{eq:7}. The integration is started from $\xi = 1$ because \begin{gather} B(\xi)\equiv \Frac{2(1 + \mathrm{A})\,\overline{m}}{\epsilon L^{2}(\overline{\sigma}_{yx} + \mathrm{A}\overline{\sigma}_{yx})}\label{eq:18} \end{gather} becomes unbounded as $\xi\rightarrow0$, and hence the right hand side of~\eqref{e-60} is indeterminate at $\xi = 0$. Equation~\eqref{e-60} is therefore integrated from $\xi = 1$ to get \begin{gather} \label{eq:9} \overline{\omega}(\xi) = \overline{\omega}_{w}\exp\left(- \Int{\xi}{1}{B(\xi')}{\xi'}\right), \end{gather} where $\overline{\omega}_{w} = \omega_{w}(\textfrac{W}{g})^{\textfrac{1}{2}}$ is the dimensionless angular velocity at the wall. It is shown in Appendix~B that $\Int{\xi}{1}{B(\xi')}{\xi'}$ becomes unbounded as $\xi \rightarrow 0$. Hence $\overline{\omega}$ satisfies the boundary condition $\overline{\omega}(0) = 0$ for all finite values of $\overline{\omega}_{w}$. It is also of interest to determine the behaviour of the solutions in the limit $\epsilon\equiv\textfrac{d_{p}}{W}\rightarrow0$. The issue here is the scaling of the shear layer thickness as a function of the channel half-width, $W$. For small $\epsilon$, an asymptotic solution is constructed using a perturbation technique described in Appendix~C. The predictions of this solution are discussed in the next section along with the numerical results. \subsubsection{Parameter values} \label{sec:Parameter-values} The intrinsic density of the particles ($\rho_{p}$) was taken from the studies of \citet{neddermanandlaohakul80}, \citet{natarajanetal95} and~\citet{tuzunandnedderman85}. Glass beads were used in all the experiments. For want of data, the angle of internal friction $\phi$ was taken to be equal to the reported angle of repose. The parameters $\nu_{\min}$ and $\nu_{\max}$ were chosen to be 0.5 and 0.65, respectively. The parameters in~\eqref{eq:24} were estimated as follows. \citet{jyotsnaandrao97} used the data of~\citet{fickieetal89} to obtain an expression for the variation of the mean stress at a critical state ($\sigma_{c}$) as a function of the solids fraction $\nu$. This expression was used to generate the values of $\overline{\sigma}_{c}$ for $\nu$ in the range 0.54--0.58, and the latter were used to estimate $\textfrac{\Lambda}{(\rho_{p} g W)}$, $p$ and~$q$ by the method of nonlinear least squares, using the Marquardt-Levenberg algorithm~\citep[p.~678]{pressetal92}. The parameter values were found to be $\textfrac{\Lambda}{(\rho_{p} g W)} = 817$, $p = 2.5$, and~$q = 2.2$. In the experiments of~\citet{neddermanandlaohakul80} and \citet{natarajanetal95}, a layer of particles was stuck to the walls of the channel. This will be referred to as a fully rough wall. When we compare model predictions with their data, the angle of wall friction is chosen as $\delta = \tan^{-1} (\sin \phi)$~\citep{kaza82}. For comparing the predictions with stress measurements of~\citet{tuzunandnedderman85}, the measured angle of wall friction, $\delta = 10^{\circ}$, was used. The value of the parameter $L$, which occurs in the yield condition~\eqref{eq:8} was estimated to be $10$ by matching predicted velocity profiles with the data of \citet{neddermanandlaohakul80}~(see \fig{2}). This value was used in comparisons with all other data. The parameter $K$ was set to 0.5. \section{Results} \label{sec:results} In this section we compare the predictions of the theory with the data of~\citet{neddermanandlaohakul80}, \citet{natarajanetal95}, and~\citet{tuzunandnedderman85}. \subsection{Velocity profiles} \label{sec:velocity} With $L = 10$, there is a good match between predicted and measured linear velocity profiles of~\citet{neddermanandlaohakul80} (\fig{2}). (Predictions of the theory with $L = 2$ and $L = 20$ are also shown in \fig{2} for comparison.) The solids fraction of 0.60 predicted by the model is in close agreement with the measured average value of 0.61. While there is no sharply defined plug layer in the model (and in the experiments), there is a region of low shear rate near the center of the channel. In order to compare predictions with the data of~\citet{neddermanandlaohakul80}, the apparent thickness of the ``plug'' layer, $\xi_{p}$ is calculated from \begin{equation} \label{eq:13} \Frac{u(\xi_{p})}{u(\xi = 0)} = 0.95. \end{equation} Hence the shear layer thickness, scaled by the particle diameter is $\Delta\equiv(1 - \xi_{p})(\textfrac{W}{d_{p}})$. The model predicts a central plug layer and a shear layer adjacent to the wall whose thickness is about 10.5 particle diameters. With $L = 10$, the predicted velocity profile also agrees well with the data of~\citet{natarajanetal95} as shown in~\fig{3}. This is an encouraging result because the ratio of the channel width to the particle diameter differs by a factor of 3.5 for the two sets of data. For the profile shown in \fig{3}, the solids fraction of 0.59 lies in the range 0.55--0.67 estimated from the experiments. The open circles in figures~2 and~3 show the asymptotic velocity profiles for small $\epsilon$ --- the deviation from the numerical solution is greater in \fig{3} because $\epsilon$ is larger than that in \fig{2}. For $\epsilon=\textfrac{1}{600}$, the asymptotic solution is indistinguishable from the numerical solution, as shown in \fig{4}. The angular velocity ($\overline{\omega}$) profile, shown in \fig{5}, differs slightly from that of half the dimensionless vorticity $\textfracp{1}{2}\:\textder{u}{\xi}$. As expected, the difference is more pronounced in the shear layer. (In a classical continuum, $\textfracp{1}{2}\:\textder{u}{\xi}$ represents the local angular velocity of an infinitesimal spherical material volume.) The asymptotic solution deviates significantly from the numerical solution for $\epsilon=\textfrac{1}{30}$~(\fig{5}), but the two solutions agree well for $\epsilon=\textfrac{1}{600}$ (\fig{6}). \subsection{Influence of channel width on the thickness of the shear layer} \label{sec:Infl-chann-width} For a fixed value of the particle diameter $d_{p}$, the thickness of the shear layer $\Delta$ increases with the half-width of the channel~$W$ (solid line in~\fig{7}). This is roughly in accord with the data of~\citet{neddermanandlaohakul80}, which are represented by solid symbols in \fig{7}. For each value of the $\textfrac{W}{d_{p}}$, there are three data points; these correspond to the estimates of $\Delta$ obtained by fitting three different functional forms to the measured velocity profile. For small values of $\epsilon=\textfrac{d_{p}}{W}$, the perturbation solution~(Appendix~C) shows that \begin{equation*} \Delta \sim \left(\Frac{L^{2}}{2}\right)^{\textfrac{1}{3}} \epsilon^{\textfrac{-1}{3}}. \end{equation*} Thus the dimensional thickness of the shear layer is proportional to $(\textfrac{d_{p}}{W})^{\textfrac{-1}{3}}$ when $\textfrac{d_{p}}{W} \ll 1$, and hence does not attain a constant value in this limit. It would be interesting to conduct experiments with much larger values of $\textfrac{W}{d_{p}}$ than in the range shown in \fig{7}. This would permit a more stringent test of the model predictions. \subsection{Influence of the parameter $L$ on the thickness of the shear layer} As mentioned earlier, the length scale $L d_{p}$ was chosen to fit the model predictions to the data of~\citet{neddermanandlaohakul80}. It is important to know how the predictions vary with changes in this parameter. Figure~8 shows that the thickness of the shear layer is a weak function of~$L$. In the limit of small $\epsilon$, the shear layer thickness varies as $L^{\textfrac{2}{3}}$ (Appendix~C). \subsection{Influence of the wall-roughness factor $K$ on thickness of the shear layer} The variation of the shear layer thickness with the roughness parameter $K$ is shown in figure~9. As mentioned earlier, $K \to \infty$ corresponds to a very smooth wall; it decreases as the wall roughness increases. Figure~9 shows that there is little variation with $K$ of the shear layer thickness for small $K$, but significant variation in the range $\approx$ 1--200. For $K$ greater than 200, the velocity at the wall is greater than 95\% of the centerline velocity. Hence, by our definition \eqref{eq:13}, the thickness of the shear layer is zero. As shown in Appendix~C, the shear layer thickness is independent of $K$ in the limit of small $\epsilon$. \subsection{Stresses} \subsubsection{Stress profiles} Figure~10 shows the profiles of the shear stresses $\overline{\sigma}_{xy}$ and $\overline{\sigma}_{yx}$ for $\epsilon = \textfrac{1}{30}$ and $\textfrac{1}{600}$. It is clear that the difference between $\overline{\sigma}_{xy}$ and~$\overline{\sigma}_{yx}$ increases with $\xi$. Because $\overline{\sigma}_{yx} > \overline{\sigma}_{xy}$, the couple stress $\overline{m}$ also increases with $\xi$~(\fig{11}), in accord with~\eqref{e-61}. The open symbols in figures~10 and~11 represent the asymptotic solution for small $\epsilon$. When $\epsilon=\textfrac{1}{600}$, it is clear that the asymptotic solution is indistinguishable from the exact solution, and the difference $\overline{\sigma}_{xy} - \overline{\sigma}_{yx}$ is also very small. \subsubsection{Wall stresses} We now compare the predicted wall stresses with the data of~\citet{tuzunandnedderman85} \mytable{1}. The normal and shear stresses are over-predicted, but are of the same order of magnitude as the measured values. As noted by~\citet{mohanetal97}, the dimensions of the channel used in the experiments are such that the front and the back faces may support a significant part of the weight of the material. Hence the shear stress measured at the side wall is expected to be less than the prediction, which assumes a channel of infinite depth. It is interesting to note that the estimate of~\citet{tuzunandnedderman85} for the average solids fraction is 0.63 and the model predicts a value of 0.625. \begin{table} \begin{center} \begin{tabular}{lccccc} & \multicolumn{1}{c}{$H$} & \multicolumn{2}{c}{Normal stress} & \multicolumn{2}{c}{Shear stress}\\ &&Static&Flowing&Static&Flowing\\ &0.91&2.7--3.6&4.0--5.1&0.30--0.39&0.29--0.35\\ Experiments&1.07&2.7--2.9&3.8--4.6&0.30&0.41--0.58\\ &1.22&2.7--3.6&4.6--5.7&0.52--0.65&0.58--0.76\\ Cosserat model&---&---&7.97&---&1.12\\ Kinetic model &---&---&2.51&---&1.15 \end{tabular} \caption{Comparison of predicted wall stresses with the data of~\citet{tuzunandnedderman85} for glass beads. Here $H$ is the depth measured from the top of the channel. Units: $H$--~m, stress--~kN/$\mathrm{m}^{2}$. Parameter values: $W = 0.155~\mathrm{m}$, $d_{p} = 2.29~\mathrm{mm}$, $\rho_{p} = 1180~\mathrm{kg}/\mathrm{m}^{3}$, $\phi = 30^{\circ}$, $\delta = 8^{\circ}$.} \label{tab:wallstresses} \end{center} \end{table} \section{Comparison with other models} \subsection{The classical frictional model} The classical frictional model predicts a flat velocity profile. This is consistent with the profile predicted by the Cosserat model in the limit $d_{p}\rightarrow0$ for a fixed value of $W$. Further, the Cosserat continuum reduces to the classical continuum in this limit, because $m\rightarrow0$ and $\sigma_{xy}\rightarrow\sigma_{yx}$. \subsection{The kinetic and frictional-kinetic models} \label{sec:Comp-Coss-model} The broken curves in~\fig{7} show the results obtained by using the (high density) kinetic model and the frictional-kinetic model. For these models we have used the equations given in~\citet{mohanetal97}, except that the mean stress at critical state ($\sigma_{c}$ in their paper) is evaluated using~\eqref{eq:24}. The kinetic model is based on constitutive equations derived by using the kinetic theory of dense gases~\citep[see, for example,][]{lunetal84}. Two of the underlying assumptions of this theory, namely instantaneous binary collisions between particles and molecular chaos with respect to particle velocities, are expected to break down at high solids fractions. Therefore it is surprising, and perhaps fortuitous, that the predicted thickness of the shear layer is in fair agreement with the data~(\fig{7}) even though the solids fraction is in the range of 0.64--0.65. Based on the results shown in~\fig{7}, it is difficult to discriminate between the Cosserat and the kinetic models. It should be noted that both these models contain a material length scale in their constitutive equations. As noted by~\citet{tejchmanandwu94}, this may be a pre-requisite for a satisfactory description of shear layers. The frictional-kinetic model is constructed by assuming that the stress tensor is the sum of the kinetic stress tensor and the frictional stress tensor. This model grossly underestimates the thickness of the shear layer (see the dotted curve in \fig{7}), probably because (i)~frictional effects dominate kinetic effects in the shear layer, and~(ii)~the frictional constitutive equations do not contain a material length scale. \subsection{The model of \citet{tejchmanandgudehus93}} \label{sec:Comparison-with-work} The work of \citet{tejchmanandgudehus93} appears to be the only other study which uses a Cosserat model for channel flow. They use an elasto-plastic model to examine the batch discharge of material from a cylindrical bin. The constitutive equations involve the Jaumann stress rate and the `velocity strain' tensor. Since they do not present results for steady fully developed flow, a direct comparison of our predictions with theirs is not possible. We are currently attempting to use their model for the problem at hand, but some issues require consideration before results can be obtained. For example, \citet{dienes79} have reported that the Jaumann stress rate furnishes an unrealistic oscillatory response in simple shear for a hypoelastic model. \section{Discussion} \label{sec:Discussion} Unlike the classical frictional model~\citep{mohanetal97}, the present Cosserat model predicts velocity profiles which agree well with the data of~\citet{neddermanandlaohakul80} and~\citet{natarajanetal95}. Further, the variation of the thickness of the shear layer with the width of the channel is also captured reasonably well by the model. The predicted wall stresses are of the same order of magnitude as the measured values, but there is considerable scope for improvement. In this context, it may be desirable to account for the finite spacing between the front and back walls. Our solution of the model for the limiting case of an infinitely wide channel~($\epsilon\rightarrow 0$) with fully rough walls indicates that the shear layer thickness, scaled by the particle diameter, grows as $\epsilon^{\textfrac{-1}{3}}$, where $\epsilon$ is the ratio of the particle diameter to the channel width. It would be interesting to compare this result with experiments conducted for a wide range of $\epsilon$. In the present work, and in most applications of the frictional Cosserat models, \textit{ad hoc} values are prescribed for the parameters~$a_{1}$, $a_{2}$, and~$L$ in the yield condition~\eqref{e-10}. Either suitably designed experiments, or a micro-mechanical treatment, would be valuable in providing estimates for these parameters. Similarly, it would be desirable to have a micro-mechanical basis for the kinematic boundary condition~\eqref{e-63}. An unsatisfactory feature of our model is that the solids fraction is constant across the channel. This is in variance with qualitative observations of~\citet{natarajanetal95} that the density in the shear layer is lower than that in the plug region. Perhaps the inclusion of elastic or kinetic effects in the model would correct this feature. In any case, accurate density measurements in channel flow are lacking, and more investigations in this direction are needed.
\section{Introduction} In the last twenty years increasing attention has been paid to the study of dwarf galaxies in order to understand their crucial role in galaxy formation and evolution. In hierarchical clustering theories (White \& Frenk 1991; Kauffmann, White, \& Guiderdoni 1993) these systems can constitute the building blocks from which larger systems have been created by merging, while in monolithic collapse scenarios (Tinsley \& Gunn 1976; Tinsley 1980a) they have been suggested to represent the debris of massive galaxies unable to form stars until z$\sim$1. A population of newly star--forming dwarfs at z$<$1 has been also invoked in some evolutionary models (Broadhurst, Ellis, \& Shanks 1988; Babul \& Ferguson 1996; see also Ellis 1997 for a review on the subject) to reproduce the excess of faint blue galaxies observed in deep photometric surveys, the most famous being the Hubble Deep Field (Williams et al. 1996). Early--type dwarfs (dEs and dSphs) are gas poor and constituted by intermediate and old stellar populations, while late--type dwarfs (dIrrs) are gas rich and their light is dominated predominantly by very young stars associated with bright HII regions, indicators of an ongoing star formation process. The former kind of dwarfs show a smooth luminosity distribution with a low surface brightness (like Sextants, NGC~147, Leo~I, Fornax, Sculptor, etc.), while the latter group appears with a patchy intensity distribution (NGC~6822, NGC~1569, IC~1613, etc.) and an intermediate surface brightness which can become very high in the bright blue knots of blue compact dwarf galaxies (hereinafter BCDGs). At present it is not well understood if there is an evolutionary interconnection between dIrr and dE galaxies (see e.g. Gallagher 1998), as no consistent picture of dwarf galaxy evolution has emerged yet. It has been suggested (Davies \& Phillips 1989; and references therein) that the natural evolution of dIrrs could be the condition of dEs through the phase of BCDG. In this hypothesis, a strong starburst (intense star formation episode concentrated in a very short time) can originate a galactic superwind (Heckman 1995) with the mechanical energy supplied by stellar winds and supernova explosions generated by newly formed massive stars. This wind can blow out all the gas from a dIrr, due to its shallow potential well, and transform the galaxy into a gas--poor dE. Indeed, narrowband H$\alpha$ images and X-ray maps show evidence of the existence of these superwinds in some irregulars and BCDGs, as for example NGC~1569 (Heckman 1995), NGC~1705 (Meurer et al. 1992), IZw18 (Martin 1996; Petrosian et al. 1997), etc. On the other hand, taking into account observed chemical, photometric and kinematic properties of both dwarf irregulars and ellipticals, it seems quite hard to find an efficient mechanism to transform a late-type dIrr into a dE (Jerjen \& Binggeli 1997). The derivation of the star formation history of dwarf irregular galaxies and the corresponding identification of objects with a starbursting regime become thus of primary importance to gain an insight into the nature and evolution of dwarf galaxies in general. IZw18 (also Mrk 116 or UGCA 166) is possibly the BCDG with the most striking properties. At a recession velocity of 745$\,\pm\,$3 km/sec (Dufour, Esteban, \& Casta\~{n}eda 1996a), corresponding to a distance of 10 Mpc (H$_0\,$= 75 km\,sec$^{-1}$Mpc$^{-1}$), this system shows very blue colors. The most recent estimates on emission-line corrected broadband images give U--B=\,--\,0.88 and B--V=\,--\,0.03 (Van Zee et al. 1998). These colors are indicative of a very young population, but do not exclude an underlying older one. The total mass of IZw18 from the rotation curve at a radius of 10$^{\prime\prime}\!$--12$^{\prime\prime}$ is estimated to be $\sim 10^8$ M$_{\odot}$ (e.g. Davidson \& Kinman 1985; Petrosian et al. 1997; Van Zee et al. 1998). A large amount of neutral gas is detected all around the system, totalling $\sim 7\times10^7$ M$_{\odot}$. This corresponds to $\sim$70$\%$ of the total mass, but only $ 10^7$ M$_{\odot}$ of HI are associated with the optical part of the galaxy (e.g. Lequeux \& Viallefond 1980; Van Zee et al. 1998). When discovered by Zwicky (1966), IZw18 was described as ``two galaxies separated by 5\farcs6 and interconnected by a narrow luminous bridge'', surrounded by two ``very faint flares'' at 24$^{\prime\prime}$ northwest. More recent CCD ground--based images (Davidson, Kinman, \& Friedman 1989; Dufour \& Hester 1990; hereinafter respectively DKF89 and DH90) have revealed a more complex structure: the {\it two galaxies} are in fact two star--forming regions of the same galaxy (usually indicated as NW and SE components), while the {\it two flares} are just the most prominent of a few nebulosities surrounding IZw18. These minor systems are roughly aligned toward the northwest and were initially believed at the same distance, but now we know from spectroscopic studies that only one component (referred to as component C in DKF89) is at the same distance as IZw18 and is physically associated with the main body (Dufour et al. 1996a; Petrosian et al. 1997; Van Zee et al. 1998). The other diffuse objects have been recognized as background galaxies (see Fig.~1a of Dufour et al. 1996b, D96, for an overview of the whole IZw18 system). In the following we will refer to IZw18 and to component C respectively as main body and secondary body (or companion) of IZw18. Both systems have been resolved into single stars for the first time only with HST/WFPC2 by Hunter \& Thronson (1995, hereinafter HT95) and D96. Indeed, IZw18 and its companion irregular galaxy are currently one of the most distant systems ever resolved into stars. This apparently insignificant BCDG became famous right after its discovery, when Searle \& Sargent (1972) measured from its emission-line spectrum an oxygen abundance [O/H]= --1.14, corresponding to only 7$\%$ of the solar value and indicating a quite unprocessed gas content. Furthermore, the first studies on its color and composition (Sargent \& Searle 1970; Searle \& Sargent 1972; Searle, Sargent, \& Bagnuolo 1973) already emphasized a current star formation rate (SFR) much higher than the mean value in the past. All these observational evidences brought to the formulation of the basic question on the nature of IZw18: is it a young galaxy which is presently experiencing its first burst of star formation, or is it an old system which has already formed stars in the past in at least another episode of star formation ? Subsequent spectroscopic studies in IZw18 (Lequeux et al. 1979; French 1980; Kinman \& Davidson 1981; Davidson \& Kinman 1985; Dufour, Garnett, \& Shields 1988; Garnett 1989, 1990; Pagel et al. 1992; Skillman \& Kennicutt 1993; Kunth et al. 1994; Stasi\'nska \& Leitherer 1996; Garnett et al. 1997; Izotov \& Thuan 1998) have confirmed its extreme metal deficiency, around 1/30--1/50 of Z$_{\odot}$. Despite many efforts to detect other galaxies with very low metallicity (Terlevich, Skillman, \& Terlevich 1995), IZw18 still remains the galaxy with the lowest metal and helium content known so far. This makes the system a fundamental point in the derivation of the primordial helium abundance (Izotov, Thuan, \& Lipovetsky 1994, 1997; Olive, Steigman, \& Skillman 1997; Izotov \& Thuan 1998) and in the study of the properties of chemically unevolved galaxies. However, there are several observational indications that IZw18 is not a primordial galaxy, for instance the presence of relatively high C/O and N/O abundance ratios justified only with an earlier population of low and intermediate mass stars (Dufour et al. 1988; Garnett et al. 1997) and the photometric evidence of an underlying red stellar population, both from surface photometry of the whole galaxy in the NIR (Thuan 1983) and from photometry of single stars in the optical bands (HT95, D96). For nearby galaxies the safest determination of their SF history is obtained resolving their stellar population into single stars and inferring their SFR and initial mass function (IMF) with the synthetic color--magnitude diagram (CMD) method. This method was first developed for dwarf irregular galaxies in the Local Group observed with ground-based telescopes (Ferraro et al. 1989; Tosi et al. 1991, hereinafter TGMF; Greggio et al. 1993, GMTF; Marconi et al. 1995, MTGF) and has now been updated (Greggio et al. 1998, hereinafter G98) for an optimized application to galaxies observed with HST. A procedure for the comparison between observed and synthetic CMDs has been developed also by Tolstoy \& Saha (1996), who have introduced the concept of Bayesian inference to give the relative likelihood of different models to constitute a suitable representation of the data. Gallart et al. (1996) also follow a similar approach and have introduced in the method the concept of metallicity evolution following a given law for the chemical enrichment of the interstellar medium. Here we have applied the synthetic CMD method to IZw18 using the HST archive data from HT95 and D96. The data reduction is described in Sect.2 and the resulting CMDs and LFs in Sect.3. The method and a description of the comparison of these data with theoretical synthetic CMDs and LFs are given in Sect.4, with the resulting conclusions on the recent evolution of IZw18 and its companion. An overall discussion of these conclusions in the framework of the current common knowledge on this galaxy is finally given in Sect.5. \section{Observations and Data Reduction} IZw18 has been observed with different instruments on board of HST\footnote{Observations with the NASA/ESA Hubble Space Telescope, are obtained at the Space Telescope Science Institute, which is operated by AURA for NASA under contract NAS5-26555} by different investigators. For our purposes we have retrieved from the HST archive and re-reduced all the HST/WFPC2 images available at January 1st, 1998. \subsection{The data} Two sets of deep exposures were taken in November 1994 and a third set of shorter ones in March 1995. In the first set of data (PI: Hunter, GO-5309, November 1994) IZw18 was centered on the PC CCD, with an effective plate scale of 0\farcs045 pixel$^{-1}$ and a field of view corresponding to 36$^{\prime\prime} \times$ 36$^{\prime\prime}$. The target was observed in the three broadband filters F336W, F555W and F814W (similar to the standard ground-based broadbands U, V and I), and in the two narrowband filters F469N and F656N (sampling the nebular lines HeII $\lambda$4686 and H$\alpha$ $\lambda$6563), in order to map the ionized gas and WR stars. Results from this set of data are presented in HT95. The second set of exposures (PI: Dufour, GO-5434, November 1994) consists of deep frames of IZw18 and its companion system on the WF3 CCD, with a plate scale and a field of view of 0\farcs1 and 80$^{\prime\prime} \times$ 80$^{\prime\prime}$ respectively. The frames are available in the three broadband filters F450W, F555W and F702W (corresponding indicatively to B, V and R), and in the two narrowband filters F502N and F658N, mapping the two nebular lines [OIII] $\lambda$5007 and [NII] $\lambda$6583. Photometric results from this set of data are presented by D96. Finally, the third set of images (PI: Dufour, GO-5434, March 1995), are in the three broadband filters F439W, F555W and F675W (the F439W is a filter in the B band region narrower than the F450W; the F675W is a filter in the R band region narrower than the F702W). A complete summary of all the data available for IZw18 as observed with the HST/WFPC2 is presented in Table 1, where we have indicated the filter (column 1), the WFPC2 camera where the target was centered on (column 2), the principal investigator (column 3), the epoch of observation (column 4), the integration time in seconds for each single exposure (column 5) and the image root names (column 6). We actually used only a subset of all the data available on IZw18, as indicated in Table 1 by an asterisk near the image name. In fact, we were interested in obtaining the color-magnitude diagrams V, B--V and V, V--I for IZw18 with the highest possible resolution (i.e. from the PC camera) and the color-magnitude diagram V, B--V of its companion on WF3. The observations in the narrowband filters were used to take into account the contribution of the ionized gas in the different bands. \subsection{Photometric reduction} We reduced the data applying all the corrections required to minimize photometric uncertainties and to achieve the highest photometric accuracy. All the reductions were performed in the IRAF \footnote {IRAF is distributed by the National Optical Astronomy Observatories, which is operated by the Association of Universities for Research in Astronomy, Inc., under cooperative agreement with the National Science Foundation.} environment. For each single exposure we corrected warm pixels and flagged hot pixels in the data quality files using the IRAF/STSDAS task {\it warmpix}. To restore the correct relative count numbers between pixels at different positions on the CCD, we performed geometric distorsion and CTE corrections on each single frame by applying respectively the correcting image f1k1552bu.r9h available from the Archive (see Leitherer 1995 for more details) and the linear-ramp image with the appropriate value depending on the background as indicated in Holtzman et al. (1995a, b; hereinafter H95a and H95b). Multiple frames through each filter in each dataset were simultaneously co-added and cosmic-ray removed. We also removed where possible the contribution of ionized gas from broadband images, proceeding as follows: after having adequately smoothed narrowband images to eliminate pointlike sources from gas maps, we calculated with SYNPHOT the percentage of flux detected in each narrowband filter and subtracted it from each broadband image. The pre-reduction provided the following data useful for our purposes: for IZw18 and its companion on the WF3 camera, four deep frames in the F555W, F450W, F502N and F658N filters (total integration time of 4,600 seconds for each filter); for IZw18 on the PC camera we have two images in the F555W filter, the deepest one obtained from Hunter's frames (for a total integration time of 6,600 seconds) and the shorter one from Dufour's data (1,200 seconds in total). We preferred not to combine the two final PC F555W frames, because of a great rotational displacement of one frame with respect to the other, which would have implied a repixelization and data manipulation. Also available are: one deep image in the F814W band (6,600 seconds), a less deep frame in the F439W band (2,000 seconds) and a quite deep frame in F656N filter (total integration time of 4,200 seconds). In Figs~\ref{VimagePC} and \ref{VimageWF3} we show the deepest WFPC2 images in the F555W filter respectively for IZw18 on the PC camera and for its companion galaxy on the WF3 detector: both systems are well resolved into single stars. It is also possible to distinguish HII regions in the SE part of IZw18 as well as the NW cluster, while 2 bright star clusters are evident in the center and in the NW part of the companion system. The photometric reduction of the frames was performed using the DAOPHOT package in IRAF for PSF-fitting photometry in crowded fields on both original and gas-subtracted broadband images. First we applied the automatic star detection routine {\it daofind} to the deepest F555W frame (detection threshold at 4$\sigma$ above the local background level), and then we performed an accurate inspection by eye of each single detected object to reject any feature misinterpreted as star by the routine (namely, nuclei of faint galaxies, PSF tendrils, noise spikes, etc.). The identification of stars in the other filters was then forced assuming the final positions of the stars detected in the F555W deeper image as input coordinates for the starting centering and aperture photometry in the new frame. We also tried to force the photometry from the deepest F555W image to the rotated F439W image, taking into account the coordinate transformation, but we obtained systematically higher photometric errors, introduced by transformation uncertainties. We therefore decided to couple the F439W image with the shallower but unrotated F555W one. In spite of the performed centering, forcing the photometry in the second band leads to some mismatches. These have been identified by plotting the distance between the centers in the two images as function of the F555W instrumental magnitude. To overcome this problem, we used the two routines {\it daomatch} and {\it daomaster}, kindly made available by P. B. Stetson, to finally match the coordinate lists in the two coupled filters. In order to obtain the best PSF for our frames we experimented three different methods: a) we used some well isolated stars (3 or 4) in each frame to build the observed PSF, b) we ran the {\it Tiny Tim} software (Krist \& Hook 1996) to obtain a theoretical PSF and, c) we made use of the new tool of the WFPC2 PSF library to get empirical PSFs. The different PSFs considered give quite similar results for the PC camera; eventually we preferred to adopt the observed PSF since it takes into account all technical conditions occurring at the epoch of data acquisition (real focus value, thermal breathing, etc.). For the WF3 camera theoretical PSFs seem instead to work better, since in this case we have to deal with a more dramatic undersampling of the observed/empirical PSFs which introduces a higher photometric uncertainty. Once all the stars were measured with the PSF--fitting, those with a disturbed image (as indicated by the two image--peculiarity indices $\chi^2$ and {\it sharpness}) were identified and rejected. The index $\chi^2$ gives essentially the ratio of the observed pixel--to--pixel scatter in the fitting residuals to the expected scatter based on the values of the detector characteristics (readout noise and gain). The {\it sharpness} is related to the intrinsic angular size of the astronomical object. We removed all the objects with $\chi^2\!>\,$3 and {\it sharpness} lower than --1 or larger than +1, i.e. objects with size smaller than a star (like cosmic rays or image defects) or objects too extended (like blends or semi-resolved star clusters, HII regions and galaxies). We also checked individually all the stars in critical positions in our reference CMDs (i.e. the CMD of stars with photometric error smaller than 0.2 mag in both filters), like very bright blue and red stars (possible blends, unresolved stellar clusters, or unidentified cosmic rays), very faint objects (possible peaks of noise in the forced filters, or residuals of cosmic rays detections). Finally, we checked accurately all the red stars, which are particularly important to discriminate among different SF histories. \subsection{Calibration} The instrumental magnitudes obtained with the PSF--fitting technique were then converted into calibrated magnitudes following the prescription of H95a and H95b. Since in H95a and H95b the standard calibration is given for an aperture of 0\farcs5, we transformed the instrumental magnitudes on an aperture radius of 2 pixels into the corresponding ones on an aperture radius of 0\farcs5 (11 pixels for PC and 5 pixels for WF3) by calculating the aperture correction. This turned out to be a very delicate step, due to the small number of isolated stars, suitable for aperture photometry, in the field of IZw18. The derived aperture corrections strongly depend on the choice of the adopted stars, and the small number of good templates leads to a large statistical uncertainty. Figure~\ref{apcorr} illustrates this point for the deepest V and I frames. For all the isolated stars in each frame (5--15 objects) we measured the instrumental magnitudes based on aperture photometry with different radii. For each star, the aperture corrections (i.e. the difference between the magnitude at a certain aperture radius and the PSF-fitting magnitude) are shown as open circles for different values of the aperture radius. The full dots represent the correction averaged over all the measured stars, with a 2$\sigma$ rejection algorithm, and the vertical bars show the corresponding 2$\sigma$ ranges. To these mean aperture corrections we added the encircled energy corrections as indicated by H95b; the obtained values corresponding to the conventional radius of 0\farcs5 are indicated by crosses. As apparent from Fig.~\ref{apcorr}, the uncertainty in the calibration increases with the aperture radius, with more points deviating from the mean value, as it becomes increasingly difficult to find no defects in the outer pixels which are progressively included in the calculation. However, once averaged and corrected for the encircled energy, the aperture correction turns out fairly constant. The mean of the aperture correction values over the considered range of aperture radii is indicated in Fig.~\ref{apcorr} by the horizontal thick line, which corresponds to corrections of $-0.38$ mag and $-0.54$ mag in the $V$ and $I$ bands respectively. These values are almost identical to those given by H95b, which is encouraging. We thus used directly the H95b corrections for our deepest F555W and F814W frames. For homogeneity, we extended the use of H95b aperture corrections to the other frames of IZw18 on the PC camera, as well as to the frames of the companion galaxy on WF3, where the aperture corrections are much more difficult to determine empirically. We finally corrected our measures for the gain factor, the WF3 normalization and the contamination effect, where necessary, and applied the zero points to scale the photometry into the WFPC2 synthetic system relative to Vega (Table 9 of H95a). When the value of one of these correcting parameters was not available (as for the F450W filter) we took it from public images of calibration programs (e.g. the aperture correction for F450W on WF3) or assumed it from the corresponding value of similar filters (F439W instead of F450W). \subsection{Completeness analysis} One of the larger uncertainties affecting galaxy photometry is due to crowding. Thus we carried out an accurate completeness analysis using the DAOPHOT routine {\it addstar}. For each frame we performed a series of tests, by adding each time $\sim$10\% of the stars detected in each half-magnitude bin on the original image. We then performed a new photometric reduction of the frame using the same procedure applied to the original frame, and considering the same rejection criteria for spurious objects. We then checked how many added stars were lost either because not detected by the automatic routine or because recovered with a large mag difference $\Delta$m which makes them migrate to another magnitude bin. We eventually estimated the completeness factors by averaging the results obtained repeating the test 10--20 times on each frame. Due to the uneven distribution of stars on the images, fainter objects are more easily recovered in the outer, less crowded regions. In order to derive the average completeness factors affecting our CMDs we constrained the {\it Addstar} routine to put artificial stars in the frame regions where the galaxy is actually located. For every pair of filters used to construct the CMD, we performed the completeness analysis independently on each frame. Since we forced the stellar detection in the shallower frame, the overall completeness factor is actually that corresponding to the shallower filter: in practice, either the B or the I frame. Table 2 shows the completeness factors (percentage of recovered artificial stars) and the corresponding uncertainties as a function of magnitude for all the frames considered in our photometric analysis. The listed values take also into account the star selection with photometric error smaller than 0.2 mag described in the next section. The completeness factors are averages over the whole galaxy. Clearly, incompleteness can be quite different from one region to the other, depending on crowding. For instance in the most crowded zone in the NW cluster, incompleteness is so dramatic that we recover only $\sim$20\% of the artificial stars already at the 23 magnitude. HII regions and the rich star clusters are unresolved in our images and are therefore to be considered fully incomplete. Although it could be more appropriate to apply different completeness factors in different regions, when computing our theoretical simulations we apply only the average completeness factors in Table 2. Indeed the resolved objects are not numerous enough to allow us to simulate individual different regions. The completeness tests allowed us also to evaluate the influence of blending in our photometry and the goodness of the applied rejection criteria ($\chi^2$, sharpness, error, etc.). In the deepest V and I images of IZw18 on the PC detector the rejection criteria seem to affect more intermediate/faint magnitudes (24$\la$V$\la$27 and 23$\la$I$\la$25), while in the shallower B and V frames they remove more objects at brighter magnitudes (22$\la$V$\la$24.5 and 22$\la$B$\la$24). A large fraction of the artificial stars rejected for $\chi^2\!>\,$3 and {\it sharpness} outside the range from $-$1 to 1 (actually a few objects, usually rejected for the $\chi^2$ parameter) were recognizable as blends: we have been able to actually see the companion star for $\sim$50\% of the objects rejected in the B image, for $\sim$65\% in both the V images and $\sim$75\% in the I image. For IZw18's secondary body on WF3 there were very few objects discarded, and they were all blends in both the V and B frames. This result confirms the need of applying these criteria in order to remove from the CMDs spurious objects due to blends. We also looked at the effect of blending on the derived stellar magnitude, by selecting the artificial stars which were recovered with a magnitude brighter than $\Delta$m=0.25 mag with respect to the input value. This happened for 4.7\% and 4.0\% of the stars added in the deeper V and I frames of IZw18; for 2.9\% and 5.2\% in the shallower V and B images, and around 3.5\% in both the V and B frames for the companion system on WF3. These values give an estimate of the frequency of cases in which blending affects the photometry more than our allowed photometric error in the CMDs ($\sigma_{DAO}\!<\,$0.2 mag in both filters). We conclude that our data are affected by blending at an average level of $\sim$4\%, which we consider negligible for the interpretation of the CMD with the simulation procedure. \subsection{Photometric errors} Figures~\ref{errPC} and \ref{errWF3} show the behavior of the photometric error $\sigma_{DAO}$ (estimated by DAOPHOT) as a function of magnitude, respectively for IZw18 in each combined PC image and for the companion galaxy in each combined WF3 frame. In these plots we considered all the stars fitted in both coupled filters: respectively 568 and 321 objects for the V vs V--I and V vs B--V diagrams relative to IZw18, and 117 stars for the V vs B--V diagram of the companion. When the rejection criteria of $\chi^2$ and {\it sharpness} are applied, most of the points with high $\sigma_{DAO}$ at bright magnitudes disappear, since they are most probably small unresolved stellar associations, HII regions or blends. For the theoretical interpretation we will restrict our CMD to objects with photometric error smaller than $\sim$ 0.2 mag. We notice that $\sigma_{DAO}$ remains below 0.2 mag down to m$\sim$26 in each deeper frame of both IZw18 and the companion, and down to m$\sim$25 in the less deep ones of IZw18. However, $\sigma_{DAO}$ may underestimate by $\sim$20\% the total photometric error (Stetson \& Harris 1988) and we have therefore derived an independent estimate of the latter using the outcome of the completeness tests, and looking at the amplitude of the difference $\Delta$m between the assigned and recovered magnitudes of the artificial stars. It turned out that for the companion of IZw18, the upper envelope of the distribution of the $\Delta$m coincides with that of the observed $\sigma_{DAO}$, indicating that in this case the errors are well estimated. For the PC frames with IZw18, instead, $\sigma_{DAO}$ underestimates by more than 20\% the actual error, especially at the brightest magnitudes. The larger error estimates $\Delta$m have therefore been adopted in the CMD simulations described in Sect.4. \subsection{Comparison with previous photometric analyses} We have compared the photometric results obtained for IZw18 in the deep F555W and F814W images on the PC camera with the results of HT95. The comparison of our CMD (see Fig.~\ref{VIcmdPC}) with that published by HT95 (see their Fig.~5) shows that we reach a fainter limiting magnitude. Our diagram contains many more blue and red stars fainter than V$\sim$26. We have carefully checked the faintest objects to reject any uncertain detection with ambiguous shape or profile, but most of them remain and are localized in the outer and less crowded part of the SE star-forming region where the photometry of fainter objects is easier. A possible reason for our fainter limiting magnitude could be our forcing the photometry in the shallower filter. We re-measured all the stars listed in Table 2 of HT95, adopting their x and y coordinates, in order to directly compare our mags with theirs: despite the same photometric package (DAOPHOT) and the same parameters for the photometric conversion of the instrumental magnitudes into the calibrated ones (zero points, gain, contamination, etc.), we found a shift in the zero points of $\sim$0.22 in F555W and $\sim$0.14 in F814W, in the sense that we are brighter and bluer than HT95. This can be seen in Fig.~\ref{photcomp}. The difference may arise from the use of gas subtracted images or from the PSFs and aperture corrections. However, we have verified that there is no significant difference in the photometric reduction of images with and without gas subtraction (except for some peculiar objects), and that there is no shift in the zero points resulting from the use of PSFs obtained with different techniques (as described in \S~2.2). We thus believe that the offset is due to the difficulty in estimating the aperture corrections from the stars in the images of this very problematic field. As described in \S~2.3, there are too few isolated and reliable stars and the dependence of the photometric correction on the chosen stars, shown in Fig.~\ref{apcorr}, may well lead to offsets of some tenths of mag in the calibration. We have compared also our CMD in the F450W and F555W filters for IZw18's companion in the WF3 camera (see Fig.~\ref{BVcmdWF3}) with the corresponding diagram published by D96 in their Fig.~3, accounting for the reddening correction. The general aspect of the two distributions is quite similar, despite some differences. In our CMD there are less faint objects with unphysical color (B--V)$\la-$1, probably thanks to the task for masking hot and warm pixels (that could be mistaken as faint stars), which has become available after the photometric reduction by D96. There is also an offset in the zero points of both filters, with our distribution being $\sim$0.5 mag brighter than D96 in V and in B. D96 worked on gas-subtracted and rebinned images in order to construct an artificial PSF better sampled than the observed one, while we preferred to work on the original frames (but, again, we found no differences at all using the gas subtracted images) and to use a PSF simulated with the {\it Tiny Tim} software. As for the HT95 data, we attribute the photometric offsets to uncertainties in the estimate of the aperture corrections, enhanced in this case by the undersampling of the WF3 camera. For what concerns the CMD in the F439W and F555W filters of IZw18 on the PC camera, the archival data have not been published yet. We can therefore compare only indirectly our resulting CMD (see our Fig.~\ref{BVcmdPC}) with that derived from the F450W and F555W filters on the WF3 detector and plotted in D96's Fig.~2. Both diagrams show a fairly large sequence and are very similar to each other. When accounting for the reddening, our edge of the blue plume (at V$\,\simeq\,$22.5), the faint limiting magnitude (at V$\,\simeq\,$27), and the average color of the star distribution around (B--V)$\,\simeq\,$0 seem consistent with those in D96. \section{Observed color-magnitude diagrams and luminosity functions} \subsection{IZw18 Main Body} Figures~\ref{VIcmdPC} and \ref{BVcmdPC} show the CMDs derived from the PC images of IZw18 in the HST F555W, F814W and F439W bands, that in the following will be referred to as the V vs V--I and the V vs B--V CMDs respectively. In panel a) of each figure we plotted all the objects measured in both filters with a $\chi^2\!<\,$3 and --1$<\!sharpness\!<$1 (respectively 444 in the V vs V--I and 267 in the V vs B--V), while in panel b) we show only the stars with a photometric error $\sigma_{DAO}\!<\,$0.2 in both bands, after removing spurious detections (247 objects remain in the red CMD and 106 in the blue one). The main features of the complete diagram remain unaltered after the $\sigma_{DAO}\!<\,$0.2 selection. Thus we used the CMDs in panels b) as reference diagrams for the theoretical simulations, as they have a more reliable photometry. Since this target is located at high galactic latitude (b=+45$^{\circ}$), its CMDs do not suffer from significant contamination from foreground stars belonging to our Galaxy, and it is not necessary to consider this factor of uncertainty in our simulations or to correct for it our observed CMDs. A first analysis of the observed V vs V--I diagram of IZw18 (panel b) of Fig.~\ref{VIcmdPC}) shows that we reached a limiting magnitude of V$\,\simeq\,$26.5 which goes down to V$\,\simeq\,$27 for objects with the reddest V--I color. From a morphological viewpoint, in this CMD we can easily distinguish the typical blue plume observed in ground-based observations of Local Group irregulars: this plume is populated both by main-sequence (MS) stars and by stars at the hot edge of the post-MS evolutionary phases. For IZw18 the plume extends up to V$\,\simeq\,$22 and has a median color V--I$\,\simeq\,$0, indicative of a very low reddening. This is in agreement with the most recent values proposed for this parameter, e.g. E(B--V)=0.04 in HT95. We can also notice the presence in the CMD of several bright supergiants with a wide spread of colors from blue to red, and some faint red stars. As described in the previous sections, they all turned out to be real stars after a detailed analysis of their shape and profile. All the bright blue and red stars are concentrated in the innermost and more crowded regions, particularly in the NW component of IZw18. All the HII regions recognized by HT95 (as well as many stars in the badly resolved NW cluster), were automatically removed from our V, V--I diagram because they turned out to have $\chi^2\!>\,$3 and/or $sharpness$ outside the range [--1,1]\,. Also the bright star clusters and associations are rejected for the same reason. This implies that our CMD is not sampling extremely young stars. The percentage of flux discarded with this procedure corresponds roughly to 3.5\% (2\%) of the total flux of the whole galaxy in the V (I) filter, while it is $\sim$40\% of the total flux sampled by the resolved stars in both bands. An inspection of the V vs B--V diagram shows that also in this case we reached the limiting magnitude V$\,\simeq\,$26. We have however a shallower cut-off as the B--V color becomes redder: this is due to the shorter integration time and to the lower sensitivity of the F439W filter with respect to the F814W. Again in this CMD we can recognize the typical blue plume of the MS and post-MS stars, with a median color of B--V$\,\simeq\,$0, and an upper brightness limit of V$\,\simeq\,$22.5. Also in this case we are not retaining the HII regions of HT95 and the star clusters, the total flux of the galaxy lost with the rejection criteria being 1\% (0.5\%) in the V (B) band, again $\sim$40\% of the light in the resolved stars. To estimate the masses of the stars visible in the CMDs and the corresponding lookback times, we have converted stellar evolutionary tracks into the observational plane, and superimposed them on the observed CMDs. In Fig.~\ref{tracksPC} we show the Padova tracks with Z=0.0004 (Fagotto et al. 1994) converted to the V vs V--I (panel a) and V vs B--V (panel b) plane, having adopted a distance modulus of (m--M)$_0$= 30 and a reddening value E(B--V)=0.04. The V vs V--I diagram shows that in IZw18 we have detected MS stars with masses higher than 12 M$_{\odot}$ (corresponding to lifetimes younger than $\sim$ 20 Myr) and blue-loop stars with masses down to $\sim$3--4 M$_{\odot}$ (thus with ages up to $\sim$0.2 Gyr). The faintest clump of red objects in Fig.~\ref{tracksPC} can be populated by (red) core helium burning, asymptotic giant branch (AGB) stars and bright red giant branch (RGB) stars, whose masses can be in principle as low as $\sim\,$1 M$_{\odot}$, extending the lookback time up to several Gyr. However, given the large photometric error, it is not possible to estimate precisely the mass, and therefore the age, of these faint stars. In the following we will conservatively assume a lookback time of 1 Gyr. In the V vs B--V diagram we see objects with mass larger than 12 M$_{\odot}$ in the MS stage ($\tau\la\,$20 Myr) and larger than 7 M$_{\odot}$ in the post-MS phase, thus with ages less than $\sim$ 50 Myr. We will use the V vs V--I diagram to infer the SF history of IZw18 over the last $\sim$ 1 Gyr, while the V vs B--V diagram will be used as a further check over the last $\sim$ 50 Myr. In Fig.~\ref{PCLFs} we plot the differential luminosity functions (LFs) of all the stars with $\sigma_{DAO}\!<\,$0.2 in both filters present in the V vs V--I and the V vs B--V diagrams (panels a) and b) respectively). We can see in both cases a rather smooth trend. The LFs will be used in the simulations to check the consistency between models and observations. It is clear from Fig.~\ref{tracksPC} that the blue plume of IZw18 is populated by stars both on the MS and at the hot edge of the blue loop evolutionary phase and that no safe criterion can be found to separate the two different populations. For this reason we do not even attempt to derive a MS-LF, which would be inevitably affected by too large uncertainties to be of any use. Also the derivation of the slope of the LF may turn out too uncertain, once we consider that, as listed in Table 2, the data start to be incomplete already at the brightest mags and significantly incomplete at V=24. For mere sake of comparison with other galaxies, and warning that these values should only be taken as indicative, we have nonetheless computed the slope by means of a maximum likelihood fitting on the deeper V data. Down to V=23, where the data are almost complete, but only very few stars are present, $\Delta$log\,N/$\Delta$V = 1.28 $\pm$ 0.04; at V=24, where completeness is 85\%, the slope is 0.68 $\pm 0.02$, and at V=25 (75\% of completeness) it is 0.45 $\pm$ 0.04. The latter value is consistent with those derived by HT95 from the same data for stars in three different locations (slopes between 0.58 and 0.65), once we consider that they have corrected them for incompleteness. HT95 pointed out that these slopes are steeper than those derived for other star forming systems like R136 and NGC~604. They also appear steeper than the average $0.70\pm0.03$ derived by Freedman (1985) and Hoessel (1986) from a large sample of irregulars and than those derived by us in Local Group irregulars (TGMF, GMTF, MTGF) and in NGC~1569 (G98), since those were derived in the complete portion of the stellar sample distribution. The difference is more striking if one considers that the literature values are supposed to refer to MS stars, whereas here we have all kind of objects, and in general MS-LFs are steeper than global ones, which can include bright supergiants. As discussed by HT95, it seems unlikely that such steep LF is due to a steep IMF and we certainly endorse their opinion, since we will show in the following that the data of IZw18 are actually best reproduced by assuming a flat and not a steep IMF. We are rather inclined to attribute this unusual steepness to the particular star formation history of the galaxy which can have superimposed around V=24--25 two distinct stellar populations, making that mag bin much more populated that the brighter ones. \subsection{IZw18 Companion System} Figure~\ref{BVcmdWF3} shows the V vs B--V diagram for IZw18's companion, resolved into individual stars in the field of view of the WF3 camera. In panel a) we plotted the 109 stars measured in both the F555W and F450W filters with $\chi^2\!<\,$3 and --1$<\!sharpness\!<$1, while in panel b) we considered only the subsample of stars (58 objects) with $\sigma_{DAO}\!<\,$0.2, after an accurate check for spurious detections. It is worth to point out that the rejection criteria eliminate $\sim$1\% of the total flux of the secondary body in both B and V filters, and contrary to IZw18, this corresponds only to $\sim$15\% of all the light in the measured stars. Furthermore the error constraint implies the loss of a lot of faint stars, and saves only the brightest part of the blue plume and a few red supergiants and giants. The blue plume has a median color B--V$\,\simeq\,$0 and reaches the bright limit of V$\,\simeq\,$24 (the brightest point at V$\,\simeq\,$22 in panel a) corresponds to the star cluster in the center of the system). Given the high galactic latitude of the system, again we expect no contamination problems. The foreground reddening is assumed to be E(B--V)=0.04, as for the main body. In Fig.~\ref{WF3LF} we report the differential LF in the V band which refers to the CMD of the secondary body. The derivation of the slope of the LF is even more uncertain than for the main body; we have nonetheless computed it by means of a maximum likelihood fitting. Down to V=25, incompleteness is low, but only 10 stars are present, and the slope is $\Delta$log\,N/$\Delta$V = 0.60 $\pm$ 0.16; down to V=26, where completeness is 80\%, the slope is $0.50\pm0.08$. These values are totally consistent with those derived for other irregulars (see previous subsection). As already done for the main body, in Fig.~\ref{tracksWF3} we show the comparison between the Padova tracks with Z=0.0004 and the V vs B--V diagram of the companion galaxy. Taking into account the photometric error, the resolved stars could all be MS objects, yielding a lookback time of only a few tens of Myr at most. The two brightest red stars, which appear as evolved objects of $\sim$ 7 M$_{\odot}$, are in fact the two objects circled in Fig.~\ref{VimageWF3}, located rather far from the bulk of the system, and might therefore be foreground objects. As an alternative interpretation, the observed CMD can be populated by only a few MS stars, with most of the detected objects in the blue loop phase. In this case most of the MS would be fainter than our limiting magnitude, and the CMD would be sampling the evolved progeny of 4 to 9 M$_{\odot}$ objects, yielding a lookback time of $\sim$0.2 Gyr. From a comparison between the CMDs of IZw18 and its companion it is evident that the blue plume of the main system is $\sim\,$2 magnitudes brighter in V than that of the secondary body: at first glance this may be interpreted as an indication of a considerably younger population in the bigger system. However, the interpretation can be quite different, once we take into proper account the different contributions to the blue plume of MS and post-MS stars. \section{Comparison with synthetic diagrams} In order to derive the SF history and IMF of IZw18 and its companion, we have compared the observed CMDs and LFs with theoretical simulations based on homogeneous sets of stellar evolutionary tracks. The procedure applied here for the creation of synthetic CMDs is the same described in detail in TGMF for ground-based observations of Local Group irregulars and in G98 for HST optical data of the nearest starburst dwarf NGC~1569. In the latter paper a detailed description of the whole procedure and of the conversion of synthetic CMDs to the HST/WFPC2 Vega-mag system is also given. Literature values for the distance to IZw18 range from 9.8 to 11.2 Mpc, corresponding to true distance moduli between 30 and 30.3 mag. For our simulations we adopted (m-M)$_0$ = 30, and for the reddening E(B--V)=0.04 (HT95), which turned out to provide synthetic MSs with average color in agreement with the observed one. For an adopted IMF, SFR and set of stellar evolutionary tracks of a given metallicity, the final product of the simulation is a synthetic diagram containing the same number of objects (247 for V, V--I and 106 for V, B--V in the main body, and 58 in the companion) above the same limiting magnitude as the observed CMD and with the same properties of photometric uncertainties and incompleteness (Figs \ref{errPC} and \ref{errWF3}, and Table 2). The free parameters for the CMD simulations are: IMF slope; starting epoch, duration, and ending epoch of the star formation activity; mode of the SF (continuous or episodic, constant or exponentially decreasing with time). For any adopted set of stellar evolution models, we have first generated synthetic CMDs assuming constant SF throughout all the observable lookback time and Salpeter's (1955) IMF ($\alpha$=2.35), and then modified the assumptions on each parameter to see the resulting effect on the comparison between the predicted CMD and LF and the empirical data. Here we show only a few illustrative cases; for a larger compilation of model samples see Aloisi \footnote{The Ph.D. Thesis is available in electronic form upon request to the first author} (1998, hereinafter A98). The comparison between the simulated CMDs and the observed one is carryed out in terms of the major features of the stellar distribution in the color-magnitude plane, as for example the relative number of stars in different evolutionary phases, the color and magnitude of the brightest stars or of the blue plume, etc. A more quantitative comparison is performed on the LFs. We do not perform quantitative tests on the color distribution because of the large intrinsic uncertainties in the effective temperatures of models of massive stars in their post-MS stage, which are reflected in the color determination. To constrain the model selection as much as possible, we have simulated independently the V, V--I and the V, B--V diagrams and compared the corresponding results only a posteriori. This approach may appear timewasting, since the two diagrams correspond to the same galactic area and therefore represent the same stellar population, but it adds independent and useful constraints because stars of different temperature have quite different weight on the distribution of B--V and V--I CMDs, and because the photometric errors and incompleteness factors are different in different frames. Besides, these independent simulations provide a useful test on the self consistency of the method. The B--V CMD is in general too shallow to provide by itself reliable information on the SF history of IZw18; nonetheless it is very useful for a further selection of the models providing the better agreement for the V--I data, thanks to its higher sensitivity to the younger (i.e. bluer) population. We have found that only a few of the models selected in V--I turn out to reproduce also the B--V observed features. This has significantly constrained the overall scenario able to fit both the B--V and the V--I distributions. To evaluate the theoretical uncertainties due to different stellar codes, input physics, and metallicity effects, it is always important to generate synthetic CMDs with more than one set of homogeneous tracks, when available. Since the overall metallicity of IZw18 and its companion is estimated to be between Z$\sim$0.0006 and Z$\sim$0.0004, we have performed our simulations for both systems using the Padova tracks with Z=0.0004 (Fagotto et al. 1994) and the Geneva tracks with Z=0.001 (Schaller et al. 1992) and Z=0.0004 (kindly made available by D. Schaerer). These sets of stellar models differ from each other in several aspects, some of which have significant effects on the synthetic CMDs: \begin{itemize} \item At the same nominal metallicity (Z=0.0004) the Padova tracks have the same temperatures as the Geneva ones for MS stars (core H-burning phases) and for stars at the hot edges of the blue loops (core He-burning phases), but lower temperatures for red giants and supergiants, thus spanning a larger color range. \item The Geneva tracks with Z=0.001 assign lower temperatures to the red stars (due to the higher metallicity) and cover the maximum color interval, in spite of their slightly cooler hot edges of the blue loops. \item The lifetimes of massive stars at the blue loop edge are systematically longer in the Geneva tracks than in the Padova ones. This implies that the Geneva models tend to predict post-MS massive stars mostly at the blue edge of the loops, whereas the Padova models populate more homogeneously all the colors from the red to the blue edges. The opposite occurs for intermediate mass stars, for which the Padova models predict longer lifetimes at the blue edge than the Geneva ones. \end{itemize} \subsection{Main body: Simulations with Padova tracks.} We have performed about 300 simulations with the set of Padova models with metallicity Z=0.0004, testing various values for the slope of the IMF, the starting epoch, the mode and the duration of the SF activity, under the hypotheses of one or two SF episodes occurred during the last 1 Gyr. All the synthetic CMDs based on these tracks show a short color extension: the bluest stars are properly reproduced, but the coolest objects predicted by these models have systematically V--I$\leq$1.7, whereas the empirical ones are as red as V--I$\,\simeq\,$2. The cause of this inconsistency could be an excessively low metallicity parameter in the tracks, or an inadequacy of stellar models to reproduce the effective temperatures in the coolest phases. Besides, the color conversions are more uncertain in these phases, due to the much more difficult treatment of molecules in model atmospheres. On the other hand, the calibration of HST data may still be slightly uncertain and the data may present color equations not properly taken into account. Independently of the IMF and SFR, the most evident result which clearly emerges from the comparison of synthetic and observed CMDs is that we can safely exclude that only one single recent burst has occurred, started later than a few 10$^7$ years ago. This scenario definitely does not allow us to reproduce the observed red and blue stars fainter than V$\sim$26 populating the V vs V--I diagram, and leads in general to an overabundance of bright blue stars compared to those observed both in the V--I and in the B--V CMDs. As an example, in the left hand panels of Fig.~\ref{sim1fin} we have plotted the V--I and B--V CMDs and corresponding LFs of a SF episode started 10 Myr ago and still ongoing at a constant rate. In this case the IMF is steep ($\alpha$=3.0, much steeper than Salpeter's 2.35). With such a late start, all the objects with masses lower than $\sim$ 20 M$_{\odot}$ ~are still on the MS and there is no chance to populate the blue loops at intermediate and faint magnitudes. As a result, the synthetic stars are all bluer than V--I$\,\simeq\,$0, at variance with the observational distribution. In addition, the synthetic LF turns out underpopulated in the V$\,\ga\,$25.5 portion, and overpopulated in the range V$\,\la\,$23. Flattening the IMF clearly worsens the result. Also the V, B--V synthetic diagram is inconsistent with the data, being populated only with stars bluer than B--V$\,\simeq\,$0.2. To fit the data we need an earlier start of the SF activity: from many tests, we have found that to obtain acceptable results, the SF in IZw18 must have started at least 200 Myr ago (see A98). A continuous SF provides results consistent with the data either with a currently ongoing SF, or with one stopped not earlier than 5 Myr ago. In the central panels of Fig.~\ref{sim1fin} we present the V vs V--I, V vs B--V and relative LFs for the case of one episode started 1 Gyr ago with an exponentially decreasing SF activity (e-folding time $\tau$=500 Myr) and still ongoing. The adopted IMF is $\alpha$=1.5. Notice that in the LF the maximum deviation of the model from the observational points is around 3\,$\sigma$. Similar results are obtained with a later onset of the SF. With this type of models we reproduce fairly well the CMDs and LFs, provided that the adopted IMF is flat (1.5$\leq\alpha\leq$1.8). Steeper slopes (even when coupled with a constant SFR) lead to worse results, since they don't provide enough bright stars when the faint end of the LF is matched. For instance the best model obtained with a Salpeter's slope ($\alpha$\,=\,2.35) leads to a LF which deviates in several magnitude bins by more than 4\,$\sigma$ from the observational points (A98). Trying to better reproduce the observed color distributions of the stars, we have considered a two-episode scenario. Models in which the old episode occurs from $\sim$1--0.2 Gyr to 100--50 Myr ago, and the young one from $\sim$100--30 Myr ago on, provide acceptable results (similar to those shown in the central panels of Fig.~\ref{sim1fin} for a single episode) when the IMF is flat (1.5$\leq\alpha\leq$1.8). In these cases the predicted LF is always within 1--2\,$\sigma$ from the empirical one. From these simulations we find that the SFR in the two episodes is quite similar, and that a significant quiescent intermediate phase is not necessary (see A98). At the end of the two-episode simulations, we can thus assert that IZw18 has experienced a rather continuous star-formation activity over a large fraction of the whole lookback time sampled by our CMDs. The distribution of the yellow and red supergiants in the observed CMD is quite peculiar: there is a clump of faint red stars (at V$\sim\,$26--27), and a continuous distribution of objects at brighter magnitudes (V$\sim\,$23), with a gap in between. These features are not reproduced by the simulations, unless a very specific SF history (hereinafter the burst scenario) is adopted, as we discuss below. As already mentioned, these stars have been carefully checked and confirmed to be most likely actual single objects (not extended ones or spurious detections), members of IZw18. We have thus considered two different episodes, one of which could efficiently populate the evolved portions of the tracks with masses between 12 and 15 M$_{\odot}$ (see Fig.~\ref{tracksPC}). This corresponds to force most of the stars in this mass range to be in a post-MS phase, and is equivalent to consider a burst occurred between 20 and 15 Myr ago. An older longer episode of SF populates the fainter red giant region. As an example, in the right hand panels of Fig.~\ref{sim1fin} we show the CMDs and LFs obtained assuming a first episode of SF from 1 Gyr to 30 Myr ago and a second one, ten times stronger, from 20 to 15 Myr ago, both with $\alpha$=1.5. It can be noticed that in this case all the observational features are reproduced pretty well, with deviations of models from the data always less than 1\,$\sigma$. All the models in better agreement with the data assume quite flat IMFs. However, while the second burst can reproduce the observed features only if $\alpha\simeq\,$1.5, acceptable distributions are also obtained when a Salpeter IMF is adopted for the first episode ($\alpha\simeq\,$2.35), with a proper tuning of the SF parameters. \subsection{Main body: Simulations with Geneva tracks.} The set of Geneva tracks with Z=0.0004 (kindly provided by D. Schaerer) only has models for stars with masses M$\,\ga\,$3 M$_{\odot}$. Consequently our simulations cover only the last 0.3 Gyr. The synthetic CMDs based on this set are all characterized by a color extension even smaller than that of the Padova models with the same nominal metallicity. As a general result, these simulations tend to overpredict the number of bright blue supergiants, while underpopulating the magnitude range around $V \sim 23.5$. Varying the IMF and SFR parameters the agreement can be improved, but we did not find a satisfactory representation of the data (see A98). An illustrative case is shown in Fig.~\ref{sim3nn}, which assumes two episodes of SF: the first from 300 to 100 Myr ago, and a second one started 90 Myr ago and still ongoing. The IMF slope is $\alpha$ = 2, and the SF rates in the two episodes are similar to each other. In spite of being one of our best simulations with this set of tracks, the deficiency of stars with V$\,\simeq\,$23.5 can still be noticed in the LF, where the maximum deviation is almost 4\,$\sigma$. Different from the others, the Geneva set with Z=0.001 does span a color range as large as the one observed in the main body of IZw18. In spite of the relatively large value of the $Z$ parameter, the synthetic blue plume overlaps the observed one when the canonical E(B--V)=0.04 is adopted. As already found with the other sets of stellar tracks, models with only one and recent episode of SF activity are definitely inconsistent with the observed CMD and LF. This can be easily understood from Fig.~\ref{tracksPC1} where we show the superposition of the observed V--I CMD with the Geneva Z=0.001 tracks. In order to populate the region at V--I$\,\ga\,$1, and fainter than V$\,\simeq\,$25.5, stars of 3--5 M$_{\odot}$ ~must have had the time to evolve off the MS, indicating SF activity earlier than $\sim$ 100 Myr ago. Moreover, in the faintest portion of the blue plume we find objects of $\simeq$\,5--7 M$_{\odot}$, with ages up to 100 Myr. For these reasons, simulations with SF starting later than $\sim$ 0.1 Gyr are inconsistent with the observations, as already shown in the previous subsection for the Padova set. Figure~\ref{sim2fin} shows three of the best cases obtained with this set of tracks under different assumptions for the SF history. In all the cases the IMF slope is 1.5. In the left hand panels we plot the result of assuming a constant SF over the last 1 Gyr, but stopped 5 Myr ago. Despite the flat IMF, the synthetic LF is underpopulated around V$\,\simeq\,$24, with deviations at a 2\,$\sigma$ level, an inconsistency which worsens with steeper IMFs. Had the SF continued in more recent epochs too many bright blue stars would have appeared. The central panels show a simulation with two SF episodes, the oldest one started 500 Myr ago and stopped 100 Myr ago, while the recent one started 100 Myr ago and still active. The rate of SF in the first episode is slightly higher than in the second one. In the V vs B--V diagram we can see only the stars born during the most recent activity. Both the B--V and the V--I LFs deviate from the empirical one by at most $\sim$\,3\,$\sigma$, but the color distribution, especially for the brightest stars, is not satisfactorily reproduced. As already discussed, this particular feature is difficult to reproduce due to the short lifetimes of massive stars in the post-MS phases. The only way to overcome this problem is to force the models to populate the bright part of the diagram only with evolved stars (i.e. with no contribution from the upper MS). This can be achieved assuming that the more recent SF episode stopped fairly long ago (15 Myr ago) so that no stars brighter than V$\,\simeq\,$25 can be on the MS. The right hand panels of Fig.~\ref{sim2fin} show one of the best cases of this type, with the first SF episode from 200 to 30 Myr ago and the second one from 20 to 15 Myr ago. To obtain enough stars in the brighter portion of the CMD, we find that the SF rate in the second burst has been almost 7 times higher than in the old one. These models reproduce fairly well both the observed distribution of cool and warm supergiants and the curvature of the upper blue plume in the V--I CMD. They also reproduce quite well the observational B--V CMD. \subsection{Summary of the results for the main body} The results obtained with the three sets of stellar models are consistent with each other in suggesting the overall scenario for the recent evolution in IZw18. The different values obtained for the various parameters depending on the adopted tracks give an estimate of the theoretical uncertainty still associated with stellar evolution models. Their relatively small differences support the reliability of our conclusions. Some of the results are completely independent of the adopted stellar models, like the flatness of the IMF and the presence of stars with intermediate ages. In all the simulations we have found indications of an IMF significantly flatter than Salpeter's ($\alpha\,$=2.35). The exponents which have turned out to be mostly consistent with the observations are in the range $\alpha\,$=1.5--2.0, with some preference for the flatter extreme of this range. Steeper IMFs look inappropriate also in the case of currently ongoing SF, because they imply too few massive MS stars and too many intermediate mass stars, with a consequent overpopulation of the faint blue plume. Notice, however, that the derived slope obviously refers to the visible range of masses. At the distance of IZw18, nothing can be inferred with our method on the IMF of stars less massive than $\sim$ 2 M$_{\odot}$. Given the relatively short lookback time of the empirical CMDs of IZw18, we have considered a star formation activity distributed over one or at most two episodes with a regime constant or exponentially decreasing with time. We should recall that two is possibly the maximum number of SF bursts allowed by the extremely low metallicity of this galaxy (e.g. Kunth, Matteucci, \& Marconi 1995). In no case have we been able to reproduce the observed CMDs and LFs with one single episode of SF started more recently than 0.1 Gyr ago. This rules out, beyond any reasonable doubt, that IZw18 has started very recently to form its first stars. A single SF episode can reproduce rather well the data if extending over a sufficiently long period of time ($\ga$ 0.2 Gyr). With an IMF slope of 1.5 we derive typical SF rates of $\sim 6\times10^{-3}$ M$_{\odot}$yr$^{-1}$ ~for stars more massive than 1.8 M$_{\odot}$. If the SF episodes sampled by the resolved stars in IZw18 are two, we find a better agreement between synthetic predictions and empirical data especially when the younger episode is relatively old and 7--10 times stronger than the previous one. To reproduce the observed features, the younger episode must have occurred between $\sim$ 20 and 15 Myr, with a SFR of $\sim 3 (6) \times10^{-2}$ M$_{\odot}$yr$^{-1}$ ~for the Geneva (Padova) tracks. The older episode can have started any time between 1 and 0.2 Gyr ago, and continued until approximately 30 Myr ago. An earlier stop of the latter SF activity would lead to an underpopulated blue plume at the faint magnitudes. If we want instead the SF in IZw18 to have taken place until recently, not only in the densest unresolved regions (Kunth et al. 1995; De Mello et al. 1998; Izotov \& Thuan 1998; Van Zee et al. 1998), but also in the resolved field, the best agreement is attained if the most recent of these two episodes has occurred from 0.1 Gyr ago at a rate of 2--5 $\times10^{-3}$ M$_{\odot}$yr$^{-1}$. In this case, however, as well as in the single-episode case, the yellow/red supergiants observed in IZw18 are not reproduced by the models. For this reason we definitely prefer the burst scenario with the intense SF episode between 20 and 15 Myr ago. To evaluate the actual SFR in IZw18, the value obtained from the synthetic CMDs must be extrapolated from the lower mass limit adopted in the simulation (m$_{\rm low}$ = 1.8, 2 M$_{\odot}$ ~for the Padova and Geneva sets respectively) to the physical lower mass cutoff. Since the IMF at the low mass end is still highly uncertain (Larson 1998; Leitherer 1998) both in the slope and in the lower mass cutoff, the extrapolations have been performed exploring a few simple cases. For the lower mass cutoff, we have adopted the value of 0.1 M$_{\odot}$. If $\alpha$=1.5 over the whole mass range, the extrapolation leaves basically unaltered the SFRs quoted above. Alternatively, if a Salpeter slope is adopted below m$_{\rm low}$ the corrected SFR amounts to 1.4 times the values quoted above. Since the size of IZw18 is estimated to be 840 $\times$ 610 pc$^2$ (DH90), the rates presented above and corrected for the IMF extrapolation become on average 1--2 $\times$ 10$^{-2}$ M$_{\odot}$yr$^{-1}$kpc$^{-2}$ in the cases of one or two SF episodes. Only when the second episode stops as early as 15 Myr ago, its SFR can be as high as 6--16 $\times$ 10$^{-2}$ M$_{\odot}$yr$^{-1}$kpc$^{-2}$, depending on the adopted IMF and evolutionary tracks. \subsection{Simulations for the secondary body} The fiducial stars populating the CMD of IZw18 companion are so few (58), that the comparison with the corresponding synthetic diagrams is inevitably affected by small number statistics. Besides, only the V, B--V CMD is available for this object, thus sensibly reducing the lookback time, and in general the available constraints to discriminate between different evolutionary scenarios. Nevertheless some interesting conclusions can still be drawn, thanks to the circumstance that all the sets of stellar tracks favor the same overall scenarios for its star formation history. For this reason, in the following we show only the results for the Padova set of tracks. As illustrated in Section 3.2, the blue plume of the secondary body is 1.5--2 mags fainter than that of the main body. This is not necessarily a signature of an older stellar population, since we have seen in the previous sections that the brightest blue stars in the main body are mostly post-MS objects, much brighter than their MS progenitors. As visible in Fig.~\ref{tracksWF3}, the red portion of the blue plume fainter than V=25.5 can be populated either by stars of approximately 4--6 M$_{\odot}$ in the blue loop, or by more massive stars still on the MS. As a consequence, the observed magnitude distribution of the stars in the companion can be reproduced either with a quite young SF episode (started around 50 or less Myr ago) or with a rather old one (started around 200 Myr ago). Nonetheless the color of the blue plume is (slightly) better reproduced by the older scenario. Figure~\ref{simsec1fin} shows the synthetic diagrams obtained for a SF started 10 Myr ago in the top and third panel, and a SF started 150 Myr ago in the second panel. Their luminosity functions are shown in the bottom panel (solid, dotted and dashed lines, for the top, second and third CMD, respectively). In order to compensate for the higher number of massive young stars in the top case, its IMF slope is steeper than in the second case, 2.6 and 1.5 respectively. It can be seen that both models give a fair representation of the data. The third panel (and dashed LF) shows what happens to the top panel model if one only changes the adopted IMF slope from 2.6 to 1.8. Too many massive blue supergiants populate the top of the blue plume, making it far too bright, and, correspondingly, too few MS stars populate its faint end. Similar results are obtained with the other sets of tracks, though with somewhat different values for the parameters, reflecting the different lifetimes in the various evolutionary stages. For example, slightly earlier starts for the SF activity and steeper IMF slopes are derived with the Z=0.001 Geneva sequences. In the case of the secondary body where the observational constraints are modest, the range of acceptable values for the parameters is larger than for the main body. In addition, as shown in Fig.~\ref{simsec1fin}, it is difficult to disentangle the contribution of the IMF and of the SFR to the observed stellar distribution. From the hundreds simulations performed on the secondary body, we believe that no quantitative information can be derived on its IMF slope. The range of acceptable slopes for the IMF is large (1.5--3.0), but the slopes leading to diagrams in better agreement with the data are peaked at $\alpha$=2.2, somewhat flatter than Salpeter's ($\alpha$=2.35), and definitely steeper than the slopes required for the main body. Besides, the trend that the more recent the start of the SF episode, the steeper the IMF is confirmed by all models. With $\alpha$ flatter than 2, a SF started more recently than $\sim$30 Myr ago and still active has to be excluded, and one started as early as 0.15--0.20 Gyr is preferable. For steeper IMFs, SF activities started as late as 10 Myr ago and still ongoing can be appropriate to interpret the observed features. The rate of SF is obviously inversely proportional to the duration of the activity (since the number of generated stars still visible is given by the data). Considering the extrapolation from m$_{low}$ to the lower physical mass cutoff 0.1 M$_{\odot}$ either with a single-slope IMF with $\alpha$=2.2 or with Salpeter's slope, the derived SFRs must be corrected by a factor of 2.5 or 2.9, respectively. Thus for young SF episodes, occurring in the last 10--50 Myr, the average rate is 2--5 $\times 10^{-3}$M$_{\odot}$yr$^{-1}$, depending on the adopted stellar tracks. For a SF activity started as early as 0.2 Gyr ago, the average rate is lower, 1--2 $\times$ 10$^{-3}$ M$_{\odot}$yr$^{-1}$. In terms of rate per unit area, these values translate into 0.7--1.7 $\times$ 10$^{-2}$ and 3.4--6.7 $\times$ 10$^{-3}$ M$_{\odot}$yr$^{-1}$kpc$^{-2}$, respectively, once a size of 850 $\times$ 350 pc$^2$ is assumed for the secondary body (DH90). Therefore, the average SF rate in the secondary body has been similar or $\sim$3 times lower than in the main body, depending on the preferred scenario. \section{Discussion and Conclusions} We have studied the SF history in IZw18, with the main goal of trying to disentangle the long standing question of whether or not this system is experiencing now its first burst of star formation. Other investigators have already examined this question and provided contradicting answers. To mention just one recent example: Kunth et al. (1995) inferred from a series of chemical evolution models that IZw18 can have experienced at most two SF bursts, each of which with duration no longer than 10--20 Myr, whereas Legrand \& Kunth (1998, hereinafter LK) argue from a spectro-photometric-chemical model that the observed metal abundances and colors can be better explained in terms of a very low SFR (10$^{-4}$M$_{\odot}$yr$^{-1}$), continuous during 16 Gyr, with burst occurrence (and SFR 100 times stronger) only in the last 50 Myr. Other authors have argued in favour of a relatively recent onset of the SF activity in IZw18. Both HT95 and D96 suggest a continuous, and still ongoing SF over the last 30--50 Myr, as deduced comparing the CMDs of the resolved stellar population on WFPC2 images with isochrones. From the kinematic analysis of ionized gas, Martin (1996) found a bipolar bubble with a lobe more evident in the SE than in the NW part of the galaxy. Its dynamical evolution and photometric properties are well described by a continuous SF episode started 15--30 Myr ago at a rate of $\sim$ 0.02 M$_{\odot}$yr$^{-1}$. On the other hand, in the literature there are some clues of an older SF activity in IZw18; for instance, from the comparison of the C/O ratio with predictions of chemical evolution models, Garnett et al. (1997) suggest a SF episode as long as a few hundreds Myr. Our approach is to infer the SF history of the galaxy from the CMDs and LFs of its stars resolved by HST photometry. As already mentioned in the previous sections, this method does not examine the denser, unresolvable regions where some SF has certainly occurred at very recent epochs as demonstrated by the presence of several HII regions. The derived V, V--I and V, B--V diagrams have been interpreted in terms of SF and IMF by means of theoretical simulations. In comparison with other galaxies examined with the same method, it is more difficult to derive strict constraints on the SF history of the IZw18 system, because its larger distance makes much smaller the number of resolved stars and consequently poorer the statistical significance of the results, especially for the secondary body. Nonetheless, in spite of this problem and of the difficulties described above to fully reproduce all the observed features of the galaxy, the comparison of all the synthetic CMDs and LFs with the corresponding data, has led to quite firm indications on the overall properties of the evolution of IZw18. It is clear from the results presented in the previous sections that in no way can a single SF episode started only a few tens of Myr ago reproduce the observed features of the faint blue plume of the main body. The SF in IZw18 must have been already active at least 100 Myr (but more likely 500 Myr) ago to provide all the observed faint stars, both blue and red. This same conclusion is reached with all the available sets of stellar evolution tracks and is therefore independent of the adopted models; it can then be considered quite firm. The overall scenario for the SF history of IZw18 is thus an almost constant SF activity from 1 Gyr up to $\sim$30 Myr ago coupled with a burst almost ten times stronger around 15--20 Myr ago: the oldest stellar population is practically concentrated in the SE part of the galaxy, while the other stars are both in the NW and SE inner dense regions (see A98). The presence of relatively old stars excludes one of the two alternative scenarios proposed by Kunth et al. (1995), which allows for only one ongoing episode started a few Myr ago. The alternative case, of two separate episodes is instead compatible with our results. At first glance, our results seem also in agreement with LK's scenario of an almost continuous star formation activity. However, the average SF during the epochs covered by our analysis has turned out to be $\sim 10^{-2}$M$_{\odot}$yr$^{-1}$, two orders of magnitude higher than the {\it low} level predicted by LK's model. Our rate is instead close to what LK attribute to the current burst. On the other hand, the duration of the SF activity is much longer in our scenario than in LK's burst. We do find that a burst is likely to have occurred at roughly LK's burst rate, but in a shorter time interval (from 20 to 15 Myr ago in our scenario, from 50 Myr ago until now in LK's). Thus, our quantitative conclusions do not necessarily agree with LK's values. This of course does not exclude that a continuous SF activity has taken place throughout the galaxy lifetime, but it should have had an intensity quite lower than in their model, to compensate the longer duration of the recent interval at high rate. Our derived SF history is instead in agreement with Martin (1996) and Garnett et al. (1997) results. In particular, both the epoch and the level of the SFR in the most recent episode of our burst scenario agree with Martin's (1996) finding. Thus, from the study of the resolved stars in IZw18 we find support to the idea that this episode of SF powered the bipolar bubble and possibly a galactic outflow. For a direct comparison of the derived SFRs in IZw18 with those of other dwarfs, it is more physically meaningful to consider the rate per unit area. In these units the main body of IZw18 has a SFR 10$^{-2}$--10$^{-1}$ M$_{\odot}$ yr$^{-1}$ kpc$^{-2}$, and the secondary body a SFR 3--10 $\times 10^{-3}$ M$_{\odot}$ yr$^{-1}$ kpc$^{-2}$. The SFRs derived with the same method for Irregular Galaxies of the Local Group (e.g. Tosi 1998) are in the range 10$^{-4}$--10$^{-2}$ M$_{\odot}$ yr$^{-1}$ kpc$^{-2}$, while for the extremely active dIrr NGC~1569 (G98) the estimated recent SFR is between 4 and 20 M$_{\odot}$ yr$^{-1}$ kpc$^{-2}$ depending on the adopted IMF (2.35$\leq\alpha\leq$3.0). In the solar neighborhood the present SFR is in the range (0.2--1)$\times$10$^{-2}$ M$_{\odot}$ yr$^{-1}$ kpc$^{-2}$ (Tinsley 1980b; Timmes, Woosley, \& Weaver 1995). We can thus conclude that IZw18 shows a mean SF activity comparable to that of the region around the sun and that of the most active Local Group Irregulars. As a consequence, its SFR falls short of $\sim$ 2 orders of magnitude to make IZw18 a local counterpart of the faint blue galaxies, according to Babul \& Ferguson (1996) model. In all the approximately 500 simulations performed for the CMDs of IZw18 we have found indications of an IMF significantly flatter than Salpeter's ($\alpha$=2.35). The exponents which have turned out to be mostly consistent with the observations are in the range $\alpha\,$=1.5--2.0, with some preference for the flatter extreme of this range. This is the first galaxy in our sample showing such a significant evidence in favour of a flat IMF. All the others analyzed by us with the same approach (DDO~210, NGC~1569, NGC~3109, NGC~6822, Sex~B, WLM) turned out to have IMF slopes close to Salpeter's or slightly steeper, in agreement with the current general belief (e.g. Leitherer 1998) of a roughly universal IMF in irregular galaxies. Besides, the global LF of IZw18 seems steeper and not flatter than those of other irregulars (see Sect.3.1). For this reason we have examined with particular attention all the alternatives, to evaluate the possibility that a more standard IMF could be acceptable with other parameter combinations. However, in no case have we been able to reproduce the observed CMDs and LFs if all the stars were born following Salpeter's IMF. As mentioned in Sect. 4.1, we obtained acceptable results adopting a Salpeter slope in the older SF episode, but a flat IMF in the most recent generation seems required by the data. Could this peculiarity be due to the extremely low metallicity of IZw18, following the old suggestion (e.g. Terlevich 1985; Melnick \& Terlevich 1987) that the lower the metallicity, the flatter the IMF of massive stars ? This interesting possibility is however contradicted by the possible trend of the older metal-poorer episode more consistent with a steeper IMF than the younger metal-richer one and by the circumstance that the secondary body seems to have a more standard IMF (with $\alpha \sim 2.2$), despite the same low metallicity of IZw18. It is worth to stress here also the possible correlation existing between the SF activity in IZw18 and the stellar production in its companion system. Many papers in the literature consider the stellar population in component C older than that in the main body. Different studies on its resolved stellar population (D96), ionized gas (Izotov \& Thuan 1998), integrated colors and nebular spectrum (Van Zee et al. 1998) indicate ages spanning from 100 to 300 Myr, all consistent with our older scenario for this minor system. The interpretation of the empirical CMD and LF in the secondary body is however much less constrained than for the main body, due to the small number of observed stars. Indeed we find consistency with the data also adopting a recent and still ongoing SF activity, provided that the IMF exponent is steep enough. Deeper and more accurate data would be necessary to derive tighter conclusions. On somewhat speculative grounds, the general trend may be that of a SF propagating from the secondary body through the NW part of IZw18 to its SE component, where some HII regions and the brightest young clusters are concentrated and still visible (HT95; D96; Izotov \& Thuan 1998). Admittedly, the spatial correlation between stellar production in different regions of IZw18 and in its companion system is not strong. However, in the case of recent onset of the SF activity in the minor system (50--10 Myr ago), its starting epoch is similar to that (20--15 Myr ago) of the stronger burst in the main body. Also the SFRs are roughly comparable: $\sim$0.1 M$_{\odot}$yr$^{-1}$kpc$^{-2}$ for the intense burst in IZw18, and $\sim$0.02 M$_{\odot}$yr$^{-1}$kpc$^{-2}$ for the companion system. Besides, the stars in the main body generated during the strong burst (red and yellow supergiants) are possibly located preferentially in the NW part, near the companion: this burst in IZw18 might thus have been triggered by gravitational interactions. To conclude, from the analysis of the resolved stellar population in IZw18 our major results are the following: the age of the older stars seen in the main body reaches from a few hundreds Myrs up to $\sim$1 Gyr; therefore a single recent episode of SF is ruled out. Our preferred scenario for the SF history is the burst scenario, consisting of two episodes, the younger one having occurred between 15 and 20 Myr ago at a rate 7-10 times higher than in the previous activity. This refers to our analyzed field, and not to the denser regions where an ongoing SF activity shows up through the HII regions and unresolved star clusters. The SFRs that we derive for the main body of IZw18 are similar to those of nearby irregulars and the solar neighborhood. The IMF, instead, appears to be significantly flatter than in any of these normal galaxies. This is especially true for the second of the two episodes, the real burst. The IMF in the secondary body appears instead to be less extreme, with a likely slope of $\alpha\simeq\,$2.2. {\bf Acknowledgements} We warmly thank Mark Clampin, Antonella Nota and Marco Sirianni who have been of invaluable help. We are deeply indebted with Daniel Schaerer for having computed for us the stellar models with Z=0.0004, to Peter Stetson for providing some of his software and to Manuela Zoccali for help in using it. Paolo Montegriffo has also helped a lot with his unusual skill for photometric reduction in very crowded fields. Useful conversations with Claus Leitherer, Daniel Schaerer and Michele Bellazzini are also acknowledged. We are grateful to the anonymous referee for his/her useful comments and suggestions which contributed to improve the paper. Part of this work has been funded by the Italian Space Agency ASI. \clearpage
\section{Perspective on hidden symmetries} M-theory is defined in 11 dimensions, with one time and ten space coordinates. It has an extended global supersymmetry characterized by 32 supercharges $Q_\alpha $ and 528 {\it abelian} bosonic charges that include the momentum $P_\mu $, the two-brane charge $Z_{\mu \nu }$ and the five-brane charge $Z_{\mu _1\cdots \mu _5}$ \cite{townsend}-\cite{gunaydin} \begin{equation} \left\{ Q_\alpha ,Q_\beta \right\} =\gamma _{\alpha \beta }^\mu P_\mu +\gamma _{\alpha \beta }^{\mu \nu }Z_{\mu \nu }+\gamma _{\alpha \beta }^{\mu _1\cdots \mu _5}Z_{\mu _1\cdots \mu _5}\,. \label{M-algebra} \end{equation} The charges $P_\mu $, $Z_{\mu \nu }$, $Z_{\mu _1\cdots \mu _5}$ commute among themselves and with $Q_\alpha $. In addition, the 11-dimensional SO$% \left( 10,1\right) $ Lorentz generator $J_{\mu \nu }$ has non-trivial commutation rules with $Q_\alpha ,$ $P_\mu ,$ $Z_{\mu \nu },$ $Z_{\mu _1\cdots \mu _5}$ that correspond to the classification of these charges as spinor, 1-form, 2-form, 5-form respectively in 11-dimensions. We refer to the algebra satisfied by $J_{\mu \nu }$, $Q_\alpha $, $P_\mu $, $Z_{\mu \nu } $, $Z_{\mu _1\cdots \mu _5}$ as the M-algebra. It is well known that physical input in four dimensions, such as the absence of massless interacting particles with spin higher than two, constrains the maximum number of supersymmetries to 32 \cite{nahm}. This refers to the maximum number of $Q_\alpha $ that commute with the momentum operators $% P_\mu $, as is the case in the M-algebra. However, there is no physical restriction on the number of non-linearly realized supersymmetries that do not commute with $P^\mu $. In particular, it is possible to have supersymmetries that do not commute with the Hamiltonian or $P^\mu $, while being supersymmetries of the action. For example, this is the case with the well known special superconformal symmetry generated by the fermions $% S_\alpha $ in any superconformal theory in any dimension. In addition, it has been discovered recently that very simple familiar systems have previously unnoticed hidden symmetries of the type SO$\left( d,2\right) $, that are not symmetries of the Hamiltonian \cite{lifting} \cite{super2t} but are symmetries of the action. Such symmetries are made manifest by lifting the system to the formalism of two-time physics by the addition of gauge degrees of freedom together with new gauge symmetries \cite{lifting} \cite {super2t}. In this paper we provide arguments that the supersymmetries of the different dual versions of M-theory are all unified within OSp$\left( 1/64\right) $, with 64 supercharges. We will show that the M-algebra is a subalgebra of OSp$% \left( 1/64\right) $ without any contractions. The extra 32 supercharges do not commute with $P^\mu $ within OS$p\left( 1/64\right) $. We suggest they are symmetries of the ``action'' of M-theory. The symmetry structures that emerge in this way suggest that M-theory could admit a two-time physics formulation with a total of 13 dimensions. There has been a number of hints that M-theory may contain various two-time structures \cite{duff},\cite{ibjapan},\cite{vafa}-\cite{nishino}. In this paper we will show a new embedding of the symmetries of M-theory in a higher structure suggested by the formalism of two-time physics. We show that when the M-algebra is extended by adding the 11D conformal generator $K^\mu $, the closure requires the full OSp$\left( 1/64\right) $. Duality and 11D covariance suggest that $K^\mu \,$is a hidden symmetry in M-theory. The point is that the 10D covariant type-IIB superalgebra, as well as heterotic and type I superalgebras can be obtained from the same OSp$\left( 1/64\right) $ , and the 10D type-IIA superalgebra is just the dimensional reduction of the M-theory algebra. While $K^\mu \,$is non-linearly realized and remains ``hidden'' in the 11D version of M-theory, some of its components are linearly realized in a dual version of M-theory, so $K^\mu $ is actually present non-perturbatively. A further clue is the presence of conformal symmetry in some corners of M-theory as noted through the CFT-AdS correspondance \cite{maldacena}-\cite{gunayd}. These points will be illustrated in a specific toy M-model \cite{future} with a worldline action that includes 13D p-form degrees of freedom $a^{% \tilde{M}_1\cdots \tilde{M}_p}\left( \tau \right) $ for $p=3,6$ in addition to the zero brane degrees of freedom $X^{\tilde{M}}\left( \tau \right) $, $% P^{\tilde{M}}\left( \tau \right) $, $\Theta ^{\tilde{\alpha}}\left( \tau \right) $ that normally exist in a worldline formalism \cite{super2t}. The model introduces new concepts of local symmetries, including one that is a bosonic cousin of kappa supersymmetry. Various gauge choices can be found to yield the M-algebra in 11D, the type IIA, type-IIB, heterotic, type-I extended supersymmetries in 10D, and non-Abelian superalgebras in the AdS$% _n\times S^m$ backgrounds. Thus, the symmetries of different corners of the moduli space of M-theory emerge as different gauge choices in this model \cite{future}. \section{Subgroup chains} We first show that the M-algebra is contained in OSp$\left( 1/64\right) $. The supergroup OSp$\left( 1/64\right) $ has 64 fermionic generators $Q_{% \tilde{\alpha}}$ and $\allowbreak 2080$ bosonic generators $S_{\tilde{\alpha}% \tilde{\beta}}$ that form a 64$\times $64 symmetric matrix. The Lie superalgebra is \begin{eqnarray*} \left\{ Q_{\tilde{\alpha}},Q_{\tilde{\beta}}\right\} &=&S_{\tilde{\alpha}% \tilde{\beta}},\quad \left[ S_{\tilde{\alpha}\tilde{\beta}},Q_{\tilde{\gamma}% }\right] =C_{\tilde{\alpha}\tilde{\gamma}}Q_{\tilde{\beta}}+C_{\tilde{\beta}% \tilde{\gamma}}Q_{\tilde{\alpha}}\, \\ \left[ S_{\tilde{\alpha}\tilde{\beta}},S_{\tilde{\gamma}\tilde{\delta}% }\right] &=&C_{\tilde{\alpha}\tilde{\gamma}}S_{\tilde{\beta}\tilde{\delta}% }+C_{\tilde{\alpha}\tilde{\delta}}S_{\tilde{\beta}\tilde{\gamma}}+C_{\tilde{% \beta}\tilde{\gamma}}S_{\tilde{\alpha}\tilde{\delta}}+C_{\tilde{\beta}\tilde{% \delta}}S_{\tilde{\alpha}\tilde{\gamma}}\, \end{eqnarray*} The $S_{\tilde{\alpha}\tilde{\beta}}$ form the Lie algebra of Sp$\left( 64\right) $ and the constant antisymmetric matrix $C_{\tilde{\alpha}\tilde{% \beta}}$ is the metric of Sp$\left( 64\right) $. A matrix representation of OSp$\left( 1/64\right) $ is given by 65$\times $65 supermatrices. We are interested in re-expressing these generators in various bases, that are related to the basis above by unitary transformations, such that various spacetime interpretations can be given to those bases. For this purpose several branchings of Sp$\left( 64\right) $ will be used : \begin{eqnarray} A &:&Sp\left( 64\right) \supset SU^{*}\left( 32\right) \otimes SO\left( 1,1\right) \label{A} \\ &\supset &Sp^{*}\left( 32\right) \otimes SO\left( 1,1\right) \supset SO\left( 10,2\right) \otimes SO\left( 1,1\right) \,\supset \cdots \nonumber \\ B &:&Sp\left( 64\right) \supset SO^{*}\left( 32\right) \otimes SO\left( 2,1\right) \label{B} \\ &\supset &SU^{*}\left( 16\right) \otimes U\left( 1\right) \otimes SO\left( 2,1\right) \nonumber \\ &\supset &SO\left( 9,1\right) \otimes SO\left( 2,1\right) \otimes U\left( 1\right) \,\supset \cdots \nonumber \\ C &:&Sp\left( 64\right) \supset SO\left( 11,2\right) \supset \cdots \label{C} \end{eqnarray} The (*) indicates an appropriate analytic continuation that contains the non-compact groups listed. There is another chain of interest that involves a supergroup that will come up in our discussion (we are not listing a complete set of branchings). \begin{equation} OSp\left( 1/64\right) \supset OSp\left( 1/32\right) \otimes Sp\left( 32\right) \supset \cdots \end{equation} We list the first step of the decomposition of the 64$\oplus $2080 representations for the A,B,C branches. For the A-branch we have the $% SU^{*}\left( 32\right) \otimes SO\left( 1,1\right) $ representations \begin{eqnarray} 64 &=&32^{1/2}\oplus \overline{32}^{-1/2}, \label{A1} \\ 2080 &=&\left( 1^0\oplus 1023^0\,\right) _{32\cdot \overline{32}}\,\oplus 528_{\left( 32\cdot 32\right) _s}^{+}\oplus 528_{\left( \overline{32}\cdot \overline{32}\right) _s}^{-} \label{A2} \end{eqnarray} where the superscripts correspond to the SO$\left( 1,1\right) $ charge, and subscripts indicate the products of the supercharges that produce those representations ($s$ and $a$ stand for symmetric and antisymmetric product respectively). For the B-branch we have the $SO^{*}\left( 32\right) \otimes SO\left( 2,1\right) $ representations \begin{eqnarray} 64 &=&\left( 32,2\right) , \label{B1} \\ 2080 &=&\left( 1_{(32\cdot 32)_s},3_{\left( 2\cdot 2\right) _s}\right) \,\,\,\,\oplus \left( 527_{(32\cdot 32)_s},3_{\left( 2\cdot 2\right) _s}\right) \oplus \left( 496_{\left( 32\cdot 32\right) _a},1_{\left( 2\cdot 2\right) _a}\right) . \label{B2} \end{eqnarray} For the C-branch we have the $SO\left( 11,2\right) $ representations \begin{equation} 64=64,\quad 2080=78\,\,(\tilde{J}_2)\oplus 286\,\,(\tilde{J}_3)\oplus 1716\,\,(\tilde{J}_6) \label{C1} \end{equation} where the $\tilde{J}_{\tilde{M}\tilde{N}}$ , $\tilde{J}_{\tilde{M}_1\tilde{M}% _2\tilde{M}_3}$, $\tilde{J}_{\tilde{M}_1\cdots \tilde{M}_6}$ are p-forms in 13D. The $\tilde{J}_p$ can be represented by p-products of 13D 64$\times 64$ gamma matrices. For $p=2,3,6$ these are 64$\times $64 symmetric matrices that represent the 2080 components of $S_{\tilde{\alpha}\tilde{\beta}}$ in a 13D spinor basis. The A,B,C branches in this paper are related to the A,B,C branches discussed in S-theory \cite{stheory}. The C branch provides a SO$% \left( 11,2\right) $ covariant 13D interpretation for M-theory as we will see with an explicit toy M-model. \section{11D and 12D interpretations} If SO$\left( 11,2\right) $ of the C-branch is interpreted as the conformal group in 11-dimensions then one may re-classify all the generators as representations of the Lorentz subgroups SO$\left( 10,1\right) \times SO\left( 1,1\right) $ in 11-dimensions and 2-dimensions contained in 13D. To do so, re-label the 13 dimensions with two sets, one in 10+1 and the other in 1+1 dimensions, $\tilde{M}=\mu \oplus m$ where $\mu =0,1,\cdots ,10$ and $% m=\left( +^{\prime },-^{\prime }\right) $, with the metric in the extra two dimensions taken in a lightcone type basis $\eta ^{+^{\prime }-^{\prime }}=-1 $. This conformal basis emerges naturally as one of the gauge choices in two-time physics \cite{lifting} \cite{super2t} \cite{future} as will be discussed later in this paper. The 64-spinor may be re-labelled as $Q_{% \tilde{\alpha}}\sim Q_\alpha ^{1/2}\oplus S_\alpha ^{-1/2}$ with $\alpha $ denoting the 32-spinor in 11D and $\pm \frac 12$ denoting the two chiral spinors in 2D. Then the generators $J^{\tilde{M}\tilde{N}}$ of the conformal group are identified as $J^{\mu \nu }$-SO$\left( 10,1\right) $ Lorentz transformations, $J^{+^{\prime }\mu }\equiv P^\mu $ -translations, $% J^{-^{\prime }\mu }\equiv K^\mu $ -special conformal transformations, and $% J^{+^{\prime }-^{\prime }}\equiv D$ -dilatations. All the generators have definite dimensions under the commutation relations with $D$ which generates the SO$\left( 1,1\right) $ subgroup. \begin{eqnarray} D &=&1:\quad J^{+^{\prime }\mu }\equiv P^\mu ,\quad J^{+^{\prime }\mu \nu }\equiv Z^{\mu \nu },\quad J^{+^{\prime }\mu _1\cdots \mu _5}\equiv Z^{\mu _1\cdots \mu _5} \label{D1} \\ D &=&\frac 12:\quad Q_{\alpha \frac 12}\equiv Q_\alpha \label{D12} \\ D &=&0:\quad \left\{ \begin{array}{c} J^{\mu \nu },\quad J^{+^{\prime }-^{\prime }}\equiv D,\quad \,J^{+^{\prime }-^{\prime }\mu }\equiv J^\mu ,\,\,\quad J^{\mu _1\mu _2\mu _3},\,\, \\ J^{+^{\prime }-^{\prime }\mu _1\cdots \mu _4}\equiv J^{\mu _1\cdots \mu _4},\,\,\quad J^{\mu _1\cdots \mu _6}\equiv \varepsilon ^{\mu _1\cdots \mu _6\nu _1\cdots \nu _5}J_{\nu _1\cdots \nu _5} \end{array} \right\} \label{D0} \\ D &=&\frac{-1}2:\quad Q_{\alpha \frac{-1}2}\equiv S_\alpha \label{D-12} \\ D &=&-1:\quad J^{-^{\prime }\mu }\equiv K^\mu ,\quad J^{-^{\prime }\mu \nu }\equiv \tilde{Z}^{\mu \nu },\quad J^{-^{\prime }\mu _1\cdots \mu _5}\equiv \tilde{Z}^{\mu _1\cdots \mu _5} \label{D-1} \end{eqnarray} The dimensions $D=\pm 1,\pm \frac 12,0$ provide a 5-grading of the superalgebra such that under the commutation rules the dimensions add $% [X_{D_1},X_{D_2}\}\sim X_{D_1+D_2}$. If $D_1+D_2$ is not one of the dimensions listed, the result of the commutator is zero. Then we see that the $D\geq 0$ operators $J^{\mu \nu }$, $Q_\alpha ^{1/2}$, $P^\mu $, $Z^{\mu \nu }$, $Z^{\mu _1\cdots \mu _5}$, form the M-algebra: the anti-commutator $% \left\{ Q_\alpha ,Q_\beta \right\} $ contains only the 528 generators $P^\mu $, $Z^{\mu \nu }$, $Z^{\mu _1\cdots \mu _5}$ (D=1) which commute among themselves and with $Q_\alpha ^{1/2}$ (D=1/2), while $J^{\mu \nu }$ (D=0) generates the Lorentz transformations SO$\left( 10,1\right) $. Hence, the 11D M-algebra is contained in OSp$\left( 1/64\right) $ as a subalgebra without considering any contractions! OSp$\left( 1/64\right) $ is the smallest simple supergroup that includes the M-algebra as a sub-algebra without resorting to contractions. As argued above, if the special conformal generator $K^\mu $ is also included along with the M-algebra, then all other generators of OSp$\left( 1/64\right) $ {\it must} also be included for consistency with the 5-grading and Jacobi identities. Thus, conformal symmetry in 11D together with the M-algebra demand OSp$\left( 1/64\right) $. The 11D interpretation fits into the higher algebraic structures contained in the A-branch (\ref{A}). The spinor decomposes as 64=32$^{1/2}\oplus \overline{32}^{-1/2}$ under SU$^{*}\left( 32\right) \otimes SO\left( 1,1\right) $. Each 32-spinor corresponds to the fundamental representation of Sp$^{*}\left( 32\right) \subset $SU$^{*}\left( 32\right) $ and furthermore they are classified as real Weyl spinors of SO$\left( 10,2\right) $ of {\it same} chirality. This classification defines the A-envelop of the 5-graded 11D superconformal algebra described above such that 11D is embedded in 12D. One may then combine the 11D operators $J_\mu \oplus J_{\mu \nu }\,$ into a 12D generator $J_{MN}$, and the 5-form $J_{\mu _1\cdots \mu _5}$ may be written as a 6-form $J_{M_1\cdots M_6}$ that is self-dual in 12D. The $J_{MN}$ form the SO$\left( 10,2\right) $ subgroup listed in the A-branch, and together with the $J_{M_1\cdots M_6}$ they make up the 528-adjoint of Sp$\left( 32\right) $. Similarly, the remaining operators make up complete 12D representations, such as $P_1\oplus Z_2=Z_{MN} $ and $Z_5=Z_{M_1\cdots M_6}^{+}$ and $J_3\oplus J_4=J_{M_1\cdots M_4}$, etc. Finally all of these are put together as SU$^{*}\left( 32\right) \otimes SO\left( 1,1\right) $ representations as in the first step of the A-branch in (\ref{A1}-\ref{A2}). This shows that the operators $J_{MN}$, $% Z_{MN},$ $Z_{M_1\cdots M_6}^{+},Q_\alpha $ which form a 12D envelop for the M-algebra, as used in several applications of S-theory, also fit in the two-time formalism given in this paper. \section{Type-IIA, IIB, heterotic, type-I interpretations} The 10D type-IIA version follows from re-classifying the 11D basis of (\ref {D1}-\ref{D-1}) under SO$\left( 9,1\right) $ of 10D. The 32-spinor supercharge $Q_\alpha ^{1/2}$ in 11D becomes the two opposite chirality supercharges 16+$\overline{16}$ of type IIA. Similarly the 32-spinor of special superconformal generator $S_\alpha ^{-1/2}$ becomes the two opposite chirality special superconformal supercharges 16+$\overline{16}$ of type IIA. The 11D M-algebra is rewritten trivially in 10D in the type IIA basis. In the C/B-branch, the type-IIB version follows from re-classifying the 64-spinor of 13D as the spinor$\times $spinor of 10D$\oplus $3D, namely $Q_{% \tilde{\alpha}}=Q_{\alpha a}\oplus S_{\dot{\alpha}a}$ where $\alpha ,\dot{% \alpha}$ denote the real spinors 16,$\overline{16}$ of SO$\left( 9,1\right) $ and $a$ denotes the doublet spinor of SO$\left( 2,1\right) $. Both sets of spinors $Q_{\alpha a}$, $S_{\dot{\alpha}a}$ are real, and there is no relation between them via hermitian conjugation. The $Q_{\alpha a}$ play the role of the two 10D supersymmetry generators of type IIB, while the two $S_{% \dot{\alpha}a}$ play the role of the two 10D special superconformal generators of type IIB. Similarly the vector of SO$\left( 11,2\right) $ is decomposed by using $\tilde{M}=\bar{\mu}+\bar{m}$ with $\bar{\mu}=0,1,\cdots ,9$ and $\bar{m}=+^{\prime },-^{\prime },2^{\prime }$. The $\bar{m}=\pm ^{\prime }$ components are the same as the $\pm ^{\prime }$ components used to reduce 13D to 11D, and the $2^{\prime }$ component is a re-naming of the 11th dimension in SO$\left( 10,1\right) $. The anti-commutators in the superalgebra take the sketchy form \begin{eqnarray} \left\{ Q_{\alpha a},Q_{\beta b}\right\} &=&P_{\bar{\mu}}^{\bar{m}}\oplus Z_{% \bar{\mu}_1\cdots \bar{\mu}_5}^{\bar{m}}\oplus Z_{\bar{\mu}_1\bar{\mu}_2\bar{% \mu}_3} \\ \left\{ S_{\dot{\alpha}a},S_{\dot{\beta}b}\right\} &=&\,K_{\bar{\mu}}^{\bar{m% }}\oplus \tilde{Z}_{\bar{\mu}_1\cdots \bar{\mu}_5}^{\bar{m}}\oplus \tilde{Z}% _{\bar{\mu}_1\bar{\mu}_2\bar{\mu}_3} \\ \left\{ Q_{\alpha a},S_{\dot{\beta}b}\right\} &=&D_{\bar{m}\bar{n}}\oplus J_{% \bar{\mu}_1\bar{\mu}_2}^{\bar{m}}\oplus X_{\bar{\mu}_1\cdots \bar{\mu}_4}^{% \bar{m}}\oplus \chi \oplus J_{\bar{\mu}_1\bar{\mu}_2}\oplus X_{\bar{\mu}% _1\cdots \bar{\mu}_4} \end{eqnarray} The map between the 13D notation and the 10D+3D notation follows from $% \tilde{M}=\bar{\mu}\oplus \bar{m}$ \begin{eqnarray} Q_{\tilde{\alpha}} &\sim &Q_{\alpha a}\oplus S_{\dot{\alpha}a} \\ J_{\tilde{M}_1\tilde{M}_2} &\sim &J_{\bar{\mu}_1\bar{\mu}_2}\oplus \left( P_{% \bar{\mu}}^{\bar{m}}+K_{\bar{\mu}}^{\bar{m}}\right) \oplus D_{\bar{m}_1\bar{m% }_2} \\ J_{\tilde{M}_1\tilde{M}_2\tilde{M}_3} &\sim &\left( Z_{\bar{\mu}_1\bar{\mu}_2% \bar{\mu}_3}-\tilde{Z}_{\bar{\mu}_1\bar{\mu}_2\bar{\mu}_3}\right) \oplus J_{% \bar{\mu}_1\bar{\mu}_2}^{\bar{m}}\oplus \left( P_{\bar{\mu}}^{\bar{n}}-K_{% \bar{\mu}}^{\bar{n}}\right) \varepsilon _{\bar{m}_1\bar{m}_2\bar{n}}\oplus \chi \,\,\varepsilon _{\bar{m}_1\bar{m}_2\bar{m}_3} \\ J_{\tilde{M}_1\cdots \tilde{M}_6} &\sim &\varepsilon _{\bar{\mu}_1\cdots \bar{\mu}_6\bar{\nu}_1\cdots \bar{\nu}_4}X^{\bar{\nu}_1\cdots \bar{\nu}% _4}\oplus \left[ Z_{\bar{\mu}_1\cdots \bar{\mu}_5}^{\bar{m}}\oplus Z_{\bar{% \mu}_1\cdots \bar{\mu}_5}^{\bar{m}}\right] \\ &&\oplus X_{\bar{\mu}_1\cdots \bar{\mu}_4}^{\bar{n}}\varepsilon _{\bar{m}_1% \bar{m}_2\bar{n}}\oplus \left( Z_{\bar{\mu}_1\bar{\mu}_2\bar{\mu}_3}+\tilde{Z% }_{\bar{\mu}_1\bar{\mu}_2\bar{\mu}_3}\right) \varepsilon _{\bar{m}_1\bar{m}_2% \bar{m}_3} \end{eqnarray} where $\oplus $ is direct sum, but $\pm $ imply ordinary addition or subtraction, $\varepsilon _{\bar{m}_1\bar{m}_2\bar{m}_3}$ is the SO$\left( 2,1\right) $ invariant Levi-Civita tensor, and $Z_{\bar{\mu}_1\cdots \bar{\mu% }_5}^{\bar{m}}$, $\tilde{Z}_{\bar{\mu}_1\cdots \bar{\mu}_5}^{\bar{m}}$ are self-dual and anti self-dual respectively in 10D. The SO$\left( 9,1\right) $ generators are $J_{\bar{\mu}_1\bar{\mu}_2}$ and the SO$\left( 2,1\right) $ generators are $D_{\bar{m}\bar{n}}$. The operators labelled with $\bar{m}$ are triplets of SO$\left( 2,1\right) $ while the others are singlets. The singlet operator $\chi $ is written in terms of the 13D operators as $\chi =J^{+^{\prime }-^{\prime }2^{\prime }}$. Its commutation rules with the 64 spinors $Q_{\tilde{\alpha}}$ is $\left[ \chi ,Q_{\tilde{\alpha}}\right] \sim \left( \Gamma ^{+^{\prime }-^{\prime }2^{\prime }}Q\right) _{\tilde{\alpha}}$. Since one may write $\Gamma ^{+^{\prime }-^{\prime }2^{\prime }}$= $\Gamma ^{01\cdots 9}$ for 64$\times $% 64 gamma matrices \cite{stheory}, the operator $\chi $ acts like the chirality operator on the 10D spinors $Q_{\alpha a}$, $S_{\dot{\alpha}a}$. Therefore $\chi $ provides a 5-grading for the OSp$\left( 1/64\right) \,$% operators based on their 10D chirality \begin{eqnarray} \chi &=&1:\quad P_1^1\oplus Z_5^1\oplus Z_3 \label{c1} \\ \chi &=&\frac 12:\quad Q_{\alpha a} \label{c12} \\ \chi &=&0:\quad \chi \oplus D^2\oplus J_2\oplus J_2^1\oplus X_4\oplus X_4^1 \label{c0} \\ \chi &=&\frac{-1}2:S_{\dot{\alpha}a}\quad \label{c-12} \\ \chi &=&-1:\quad K_1^1\oplus \tilde{Z}_5^1\oplus \tilde{Z}_3 \label{c-1} \end{eqnarray} The bosonic generators $P_1^1,Z_5^1$ etc. are labelled by numbers in the subscripts and superscripts $Z_p^q$ that correspond to $p$-forms in 10D and $% q$-forms in 3D respectively. So the commutation rules for OSp$\left( 1/64\right) $ may be written in a graded chirality basis in the form $% [X_{\chi _1},X_{\chi _2}\}=X_{\chi _1+\chi _2}$. If the chirality $\chi _1+\chi _2$ does not exist the result of the commutator is zero. The $\chi \geq 0$ operators $J_2$, $D^2$, $\chi ,$ $Q_{\alpha a},$ $P_1^1,$ $Z_5^1,$ $% Z_3$ define the B-algebra. The chirality grading shows that $P_1^1,Z_5^1,Z_3$ commute with each other as well as with $Q_{\alpha a}$, while the SO$\left( 9,1\right) \otimes $SO$\left( 2,1\right) \otimes $U$\left( 1\right) $ subgroup, consisting of $J_2$, $D^2$, $\chi $, map the operators $Q_{\alpha a},$ $P_1^1,$ $Z_5^1,$ $Z_3$ into themselves. We have established that the B-algebra, that is essential for understanding M-theory in a IIB basis, is included in OSp$\left( 1/64\right) $ as a subalgebra without any contractions. Furthermore, since it fits into the C-branch it is consistent with the SO$\left( 11,2\right) $ symmetry of two-time physics and the idea that the 10D+3D basis is arrived at as a gauge choice in two-time physics. Indeed this is true in the toy M-model discussed later. These C-branch 13D$=$10D+3D results intersect the B-branch 10D+3D basis as can be seen by the following larger classification under the B-branch (\ref {B}) that include the envelops SU$^{*}\left( 16\right) \times U\left( 1\right) $ and SO$^{*}\left( 32\right) \times U\left( 1\right) $% \begin{eqnarray} \left( 496,1\right) &=&\left[ \chi \oplus \left( J_2\oplus X_4\right) \right] _{\left( 16\times \overline{16}\right) \left( 2\times 2\right) _a}\oplus \left( Z_3\right) _{\left( 16\times 16\right) _a\left( 2\times 2\right) _a}\oplus \left( \tilde{Z}_3\right) _{\left( \overline{16}\times \overline{16}\right) _a\left( 2\times 2\right) _a} \\ \left( 1,3\right) &=&\left( D^2\right) _{\left( 16\times \overline{16}% \right) _s\left( 2\times 2\right) _s} \\ \left( 527,3\right) &=&\left( P_1^1\oplus Z_2^1\oplus Z_5^1\right) _{\left( 16\times 16\right) s\left( 2\times 2\right) _s}\oplus \left( K_1^1\oplus \tilde{Z}_2^1\oplus \tilde{Z}_5^1\right) _{\left( \overline{16}\times \overline{16}\right) _s\left( 2\times 2\right) _s} \\ &&\oplus \left( J_2^1\oplus X_4^1\right) _{\left( 16\times \overline{16}% \right) \left( 2\times 2\right) _s} \nonumber \end{eqnarray} The products of SO$\left( 9,1\right) \otimes $ SO$\left( 2,1\right) $ spinor representations $Q=\left( 16,2\right) $ and $S=\left( \overline{16},2\right) $ that produce the various 10D and 3D forms are indicted. The SU$^{*}\left( 16\right) \otimes U\left( 1\right) $ subgroup of SO$^{*}\left( 32\right) $ is generated by $\left( J_2\oplus X_4\right) \oplus \chi $. The forms in independent parentheses $\left( \cdots \right) $ correspond to irreducible representations under SU$^{*}\left( 16\right) \otimes SO\left( 2,1\right) \otimes U\left( 1\right) $. Their collection in each line correspond to SO$% ^{*}\left( 32\right) \otimes SO\left( 2,1\right) $ representations as given in (\ref{B2}). The $496$ is the adjoint representation of SO$^{*}\left( 32\right) $ and the $527$ is the symmetric traceless tensor of SO$^{*}\left( 32\right) $. These 2080 generators form the algebra of Sp$\left( 64\right) $. It was shown in \cite{IIB} that the B-algebra may be written in an SL$\left( 2,Z\right) $ (U-duality) basis instead of the spacetime SO$\left( 2,1\right) =$SL$\left( 2,R\right) $ basis given above. These two bases are related to each other by a deformation that involves the IIB string coupling constant $% z=a+ie^{-\phi }$ (axion and dilaton moduli), and its SL$\left( 2,Z\right) $ properties $z^{\prime }=\left( az+b\right) /\left( cz+d\right) $. The deformed B-algebra may be used to perform certain non-perturbative computations at any value of the string coupling constant (see \cite{IIB}). The entire superalgebra OSp$\left( 1/64\right) $ may be rewritten in the SL$% \left( 2,Z\right) $ basis by following the prescription in \cite{IIB}. Therefore the observations in S-theory in a IIB basis may now be interpreted as observations in the two-time physics version of M-theory taken in a particular gauge. The 10D A and B bases are obviously related to each other since they both occur in the C-branch. The map between these two corresponds to a rearrangement of the 64 fermions that are in the spinor representation of SO$% \left( 11,2\right) $. This map is clearly related to T-duality as discussed in \cite{stheory}, and in the present context of two-time physics it is interpreted as just a gauge transformation from one fixed gauge to another fixed gauge. The heterotic and type-I superalgebras in 10D are then obtained as in \cite {stheory} from the IIB sector, either by setting one of the two 16-supercharges $Q_{\alpha a}$ to zero (heterotic) or by their identification (type I). This is possible because one can work in a OSp$% \left( 1/64\right) $ representation space labelled by the commuting operators $P_{\bar{\mu}}^{\bar{m}}\oplus Z_{\bar{\mu}_1\cdots \bar{\mu}_5}^{% \bar{m}}\oplus Z_{\bar{\mu}_1\bar{\mu}_2\bar{\mu}_3}$. The heterotic or type-I sectors may be viewed as BPS-like sectors in which some of these charges are related to each other as discussed in the second paper in \cite {stheory}. \section{AdS$_n\otimes S^m$ bases} In the C-branch, starting with SO$\left( 11,2\right) $ one can come down to the basis labelled by the subgroups SO$\left( 3,2\right) \otimes $SO$\left( 8\right) $ or to SO$\left( 6,2\right) \otimes $SO$\left( 5\right) $ , which are the isometries of the spaces AdS$_4\times $S$^7$ and AdS$_7\times $S$^4$ respectively. The reduction is obtained by rewriting the 13D label $\tilde{M}% =\hat{\mu}\oplus \hat{m},$ where $\hat{\mu}$, $\hat{m}$ are labels for the vectors of SO$\left( 3,2\right) ,$SO$\left( 8\right) $ or SO$\left( 5\right) ,$SO$\left( 6,2\right) $. Also, the 64-spinor is rewritten in 32+32 form $Q_{% \tilde{\alpha}}=\psi _{\hat{\alpha}a}^{+}\oplus \psi _{\hat{\alpha}\dot{a}% }^{-}.$ Each 32=4$\times $8 since $\hat{\alpha}$ denotes the 4-spinor for Sp$% \left( 4\right) \sim SO\left( 3,2\right) $ or SO$\left( 5\right) ,$ and $a,% \dot{a}$ denote the two spinors 8$_{\pm }$ for SO$\left( 8\right) $ or SO$% \left( 6,2\right) $. For SO$\left( 3,2\right) \otimes $SO$\left( 8\right) $ both $\psi _{\hat{\alpha}a}^{+}$ and $\psi _{\hat{\alpha}\dot{a}}^{-}$ are real since the corresponding spinors are real. Hence, they each have 32 real and independent components. For SO$\left( 5\right) \otimes $SO$\left( 6,2\right) $ they are in a complex basis since the 4 and the 8$_{\pm }$ are pseudo-real. However, the pseudo-reality condition still gives 32 real and independent components in each of the $\psi _{\alpha a}^{+},\psi _{\alpha \dot{a}}^{-}$. The anti-commutators in the superalgebra take the sketchy form \begin{eqnarray} \left\{ \psi _{\hat{\alpha}a}^{+},\psi _{\hat{\beta}b}^{+}\right\} &=&J_{% \hat{\mu}\hat{\nu}}^{+}\oplus J_{+}^{\hat{m}\hat{n}}\oplus X_{\hat{\mu}}^{+% \hat{m}\hat{n}}\oplus X_{\hat{\mu}\hat{\nu}}^{+\hat{m}_1\cdots \hat{m}_4} \label{osp132+} \\ \left\{ \psi _{\hat{\alpha}\dot{a}}^{-},\psi _{\hat{\beta}\dot{b}% }^{-}\right\} &=&J_{\hat{\mu}\hat{\nu}}^{-}\oplus J_{-}^{\hat{m}\hat{n}% }\oplus X_{\hat{\mu}}^{-\hat{m}\hat{n}}\oplus X_{\hat{\mu}\hat{\nu}}^{-\hat{m% }_1\cdots \hat{m}_4} \\ \left\{ \psi _{\hat{\alpha}a}^{+},\psi _{\hat{\beta}\dot{b}}^{-}\right\} &=&Y^{\hat{m}}\oplus Y^{\hat{m}_1\hat{m}_2\hat{m}_3}\oplus Y_{\hat{\mu}}^{% \hat{m}}\oplus Y_{\hat{\mu}}^{\hat{m}_1\hat{m}_2\hat{m}_3} \\ &&\oplus Y_{\hat{\mu}\hat{\nu}}^{\hat{m}}\oplus Y_{\hat{\mu}_1\hat{\mu}_2}^{% \hat{m}_1\hat{m}_2\hat{m}_3}. \nonumber \end{eqnarray} The map to the 13D operators $\tilde{J}_{2,3,6}$ can be easily established through $\tilde{M}=\hat{\mu}\oplus \hat{m}.$ The $X_{\hat{\mu}\hat{\nu}% }^{\pm \hat{m}_1\cdots \hat{m}_4}$ are self or anti-self dual in the 8-dimensions labelled by $m$. The operators $\psi _{\hat{\alpha}a}^{\pm }$, $% J_2^{\pm },$ $J_{\pm }^2,$ $X_1^{\pm 2},$ $X_2^{\pm 4}$ form OSp$\left( 1/32\right) _{\pm }$ sub-supergroups, but OSp$\left( 1/32\right) _{+}$ does not commute with OSp$\left( 1/32\right) _{-}$ since $\left\{ \psi _{\hat{% \alpha}a}^{+},\psi _{\hat{\beta}\dot{b}}^{-}\right\} $ is not zero. However, OSp$\left( 1/64\right) \supset $ OSp$\left( 1/32\right) _{+}\otimes $Sp$% \left( 32\right) _{-}$. The generators $J_2^{\pm }\oplus J^{\pm 2}$ form SO$% \left( 3,2\right) _{\pm }\otimes SO\left( 8\right) _{\pm }$ or SO$\left( 5\right) _{\pm }\otimes SO\left( 6,2\right) _{\pm }$ subgroups embedded in each of the Sp$\left( 32\right) _{\pm }$. How is OSp$\left( 1/64\right) $ superalgebra related to the familiar AdS$\times S$ supersymmetries OSp$% \left( 8/4\right) $ or OSp$\left( 6,2/4\right) $ ? These are not sub-supergroups of OSp$\left( 1/64\right) $. From our analysis it can be seen that OSp$\left( 8/4\right) $ or OSp$\left( 6,2/4\right) $ , which includes $\psi _{\hat{\alpha}a}^{+},J_2^{+},$ $J_{+}^2$, gets enlarged by the addition of the non-Abelian operators $X_1^{+2},X_2^{+4}$ into OSp$% \left( 1/32\right) _{+}$ (it is not possible to set these operators to zero naively since they are non-Abelian and they cannot be simultaneously diagonalized in a quantum theory). In turn, OSp$\left( 1/32\right) _{+}$ is the sector of OSp$\left( 1/64\right) $ that is a singlet under Sp$\left( 32\right) _{-}$. We speculate that if the CFT-AdS conjecture \cite{maldacena} corresponds to a corner of M-theory then the enlargement of the superalgebra probably does occur on the CFT side from the point of view of the N=8 Super Yang-Mills theory in 3D. The symmetry of this theory in perturbation theory to all orders is OSp$\left( 8/4\right) $. As shown in \cite{super2t}, by taking the SO$\left( 3,2\right) $ indices $\hat{\mu}=$ $\mu ,\pm ^{\prime }$ and $\psi _{\hat{\alpha}a}^{+}\sim Q_{\alpha a}^{1/2}\oplus S_{\alpha a}^{-1/2}$, the conformal supersymmetry of the AdS$_4\times S^7$ background can be rewritten in the compact form of Eq.(\ref{osp132+}), excluding the $X_1^{+2}$ and $% X_2^{+4}$ $\,$generators. How can one see the enlargement to OSp$\left( 1/32\right) _{+}$? One begins with non-perturbative field configurations that turn on the central extensions $\left\{ Q_{\alpha a}^{1/2},Q_{\beta b}^{1/2}\right\} \sim X_{\hat{\mu}=+^{\prime }}^{+\hat{m}\hat{n}}$; then the conformal symmetry $K_\mu $ requires all $X_{\hat{\mu}=-^{\prime }}^{+\hat{m}% \hat{n}}$, $X_{\hat{\mu}=\mu }^{+\hat{m}\hat{n}},$ and their non-Abelian nature generates $X_{\hat{\mu}\hat{\nu}}^{+\hat{m}_1\cdots \hat{m}_4}$, thus completing the OSp$\left( 1/32\right) _{+}$ superalgebra of Eq.(\ref{osp132+}% ). This argument shows that the inclusion of non-perturbative physics in the AdS$_4\times $S$^7$ background could be described by the Sp$\left( 32\right) _{-}$-singlet sector of OSp$\left( 1/64\right) $. The singlet sector can arise as a result of a contraction that would be related to the limits \cite {maldacena} one must take to establish the AdS-CFT correspondance. Similarly, in the C-branch, starting with SO$\left( 11,2\right) $ we can come down to a 12D$^{\prime }$ basis by separating the 13th spacelike dimension $\tilde{M}=M\oplus 1^{\prime }$. This gives 64=32$_L+32_R$ where 32% $_{L,R}$ are the two SO$^{\prime }\left( 10,2\right) $ spinors $\psi ^L$, $% \psi ^R$ of {\it opposite} chirality (contrast 12D$^{\prime }$ to the 12D of the A-branch which gave {\it same} chirality). Next we separate 12D$^{\prime }$=6D+6D so that SO$^{\prime }\left( 10,2\right) \rightarrow SO\left( 6,2\right) \otimes SO\left( 6\right) $ which is the isometry group of the AdS% $_5\times S^5$ space. Consider the 32$_L$ real components $\psi _{\alpha a}^L $ written in the $\left( 4,4\right) $ complex spinor basis with $\alpha $ and $a$ denoting the complex 4 spinors of SU$\left( 2,2\right) =SO\left( 4,2\right) $ and SU$\left( 4\right) =SO\left( 6\right) $ respectively. Since the basis is complex we need to consider the hermitian conjugates $\left( \psi _{\alpha a}^{L,R}\right) ^{\dagger }\equiv \bar{\psi}_{\dot{\alpha}\dot{% a}}^{L,R}$ where $\dot{\alpha}$ and $\dot{a}$ denote the complex \={4} spinors. Then the 12D$^{\prime }$ basis of OSp$\left( 1/64\right) $ is reduced to the 6D+6D basis and it takes the following sketchy form \begin{eqnarray} \left\{ \psi _{\alpha a}^L,\bar{\psi}_{\dot{\beta}\dot{b}}^L\right\} &\sim &J_2^L\oplus J_L^2\oplus J_L\oplus X_2^{L2},\quad \\ \left\{ \psi _{\alpha a}^L,\psi _{\beta b}^L\right\} &\sim &Z_1^{L1}\oplus Z_3^{L3},\,\,\quad and\,\,\,\,\,h.c. \\ \left\{ \psi _{\alpha a}^R,\bar{\psi}_{\dot{\beta}\dot{b}}^R\right\} &\sim &J_2^R\oplus J_R^2\oplus J_R\oplus X_2^{R2},\quad \\ \left\{ \psi _{\alpha a}^R,\psi _{\beta b}^R\right\} &\sim &Z_1^{R1}\oplus Z_3^{R3},\,\,\quad \,and\,\,\,\,h.c., \\ \left\{ \psi _{\alpha a}^L,\bar{\psi}_{\dot{\beta}\dot{b}}^R\right\} &\sim &Y_2\oplus Y^2\oplus Y\oplus Y_2^2,\quad and\,\,\,h.c. \\ \left\{ \psi _{\alpha a}^L,\psi _{\beta b}^R\right\} &\sim &W_1^1\oplus W_1^3\oplus W_3^1\oplus W_3^3,\,\,\quad and\,\,\,\,h.c. \end{eqnarray} where the generators $J_p^q,X_p^q,Y_p^q,Z_p^q,W_p^q$ are labelled with numbers in subscripts or superscripts that are SO$\left( 4,2\right) $ p-forms or SO$\left( 6\right) $ q-forms, and $h.c.$ stand for hermitian conjugate relations. The map to the 13D operators $\tilde{J}_{2,3,6}$ can be easily established through the reduction $\tilde{M}=\tilde{\mu}\oplus \tilde{% m}\oplus 1^{\prime }$ where $\tilde{\mu},\tilde{m}$ are labels for the vectors of SO$\left( 4,2\right) ,$SO$\left( 6\right) $. $J_2^L\oplus J_L^2$ generate SU$\left( 2,2\right) \otimes SU\left( 4\right) $ which is the isometry of AdS$_5\times S^5,$ i.e. SO$\left( 4,2\right) \otimes SO\left( 6\right) \subset SO^{\prime }\left( 10,2\right) $. The supersymmetry algebra in this background is $SU(2,2/4)$. What is the relation of $OSp(1/64)$ and $SU(2,2/4)$? As before, this is not a sub-supergroup. Again, upon the inclusion of the charges $J_L\oplus X_2^{L2}\oplus Z_1^{L1}\oplus Z_3^{L3}$ it can be seen that $SU(2,2/4)$ is enlarged into $OSp(1/32)_L$ as in the first two lines of the equations above (see also \cite{craps} \cite{ferrara}). Then we see that SU$\left( 1/64\right) \supset OSp(1/32)_L\otimes Sp\left( 32\right) _R$ and the sector that is singlet under $Sp\left( 32\right) _R$ is described by $OSp(1/32)_L$. If one starts with the superconformal N=4 super Yang-Mills theory in 4D, the perturbative symmetry is SU$\left( 2,2/4\right) $. By including central extensions that correspond to non-perturbative backgrounds such as monopoles and dyons one turns on the central charge $Z_{+^{\prime }}^{L\tilde{m}}$ which is part of $Z_1^{L1}$. Conformal symmetry requires the full $Z_1^{L1}$ and its hermitian conjugate $\left( Z_1^{L1}\right) ^{\dagger }$. Their commutators generate all the other remaining charges to complete the OSp$% (1/32)_L$ superalgebra. Thus the inclusion of non-perturbative physics in the AdS$_5\times $S$^5$ background could be described by the Sp$\left( 32\right) _R$-singlet sector of OSp$\left( 1/64\right) $. The singlet sector can arise as a result of a contraction that may be related to the limits \cite{maldacena} one must take to establish the AdS-CFT correspondance. \section{Toy M-Model, cousins of kappa symmetry} Our ideas can be dynamically illustrated with a toy M-model on the worldline with a new set of gauge symmetries. The main point is that in this model the various dual bases described above emerge naturally by making appropriate gauge choices within the same theory. Conversely, one can transform one basis to another dual basis by making gauge transformations, thus imitating duality in M-theory. Here we only outline the general structure of such a model, leaving the details to a future paper \cite{future}. Similar structures could be constructed for any group $G$ and a choice of subgroup $% H $ as outlined at the end of \cite{super2t}. Consider the supergroup $G=$OSp$\left( 1/64\right) $ with the subgroup $H=$SO% $\left( 11,2\right) $. Let $X_i^{\tilde{M}}=(X^{\tilde{M}},P^{\tilde{M}})$ represent position / momentum SO$\left( 11,2\right) $ vectors. The $0$-brane vectors $X_i^{\tilde{M}}\left( \tau \right) $ form an Sp$\left( 2\right) $ doublet. Sp$\left( 2\right) $ is gauged by including the gauge potentials $% A^{ij}\left( \tau \right) $. This gauge symmetry introduces first class constraints whose solution requires two-timelike dimensions as explained in \cite{lifting}. In addition, consider the 65$\times 65$ matrix which is a group element $g\left( \tau \right) \in $OSp$\left( 1/64\right) $. It is a singlet under Sp$\left( 2\right) $. The subgroup $H=$SO$\left( 11,2\right) $ acting simultaneously on the {\it left side} of $g$ and on the vectors $X_i^{% \tilde{M}}$ is gauged. The gauge potential is $\Omega ^{\tilde{M}\tilde{N}% }\left( \tau \right) $. With this information we can write covariant derivatives as in \cite{super2t} \begin{eqnarray} D_\tau X_i^{\tilde{M}} &=&\partial _\tau X_i^{\tilde{M}}-\varepsilon _{ik}A^{kj}X_i^{\tilde{M}}-\Omega ^{\tilde{M}\tilde{N}}X_{i\tilde{N}}\,, \\ D_\tau g &=&\partial _\tau g-\frac 14\Omega ^{\tilde{M}\tilde{N}}\left( \Gamma _{\tilde{M}\tilde{N}}g\right) \,. \end{eqnarray} Consider the part of the Cartan connection $\left( D_\tau g\right) g^{-1}$ restricted to the subgroup $H$% \begin{equation} \left( D_\tau gg^{-1}\right) _H=\frac 1{32}Str\left( \Gamma _{\tilde{M}% \tilde{N}}\,D_\tau gg^{-1}\right) =\frac 1{32}Str\left( \Gamma _{\tilde{M}% \tilde{N}}\,\partial _\tau gg^{-1}\right) -\Omega _{\tilde{M}\tilde{N}} \end{equation} A Lagrangian that is invariant under the gauge symmetry Sp$\left( 2\right) \times SO\left( 11,2\right) $ is given by \begin{equation} \pounds =\frac 12\,\varepsilon ^{ij}D_\tau X_i^{\tilde{M}}X_j^{\tilde{N}% }\eta _{\tilde{M}\tilde{N}}+\frac 12\left[ \left( D_\tau gg^{-1}\right) _H\right] ^2+\frac 12\left( \varepsilon ^{ij}X_i^{\tilde{M}}X_j^{\tilde{N}% }\right) ^2. \end{equation} This action is invariant under the {\it global} symmetry OSp$\left( 1/64\right) $ that acts {\it linearly} on the {\it right side} of $g\left( \tau \right) $. On the left side of $g\left( \tau \right) $ the evident symmetry is smaller $H$=SO$\left( 11,2\right) $ because of the presence of $% \Omega $, but as we will see there is a much bigger local symmetry that is realized non-linearly. Thus the model is realized in the C-branch OSp$\left( 1/64\right) $. As we saw above, the C-branch contains all the interesting spacetime interpretations at its intersections with the A and B branches and with the OSp$\left( 1/32\right) $ branch. The group element $g\left( \tau \right) $ can be written in the form $g=ht$ where $h\in H$ and $t\in G/H$. We can use the $H=SO\left( 11,2\right) $ gauge symmetry to choose a unitary gauge by eating away $h\left( \tau \right) $. Using the equations of motion (or doing the path integral) one may solve for $\Omega _{\tilde{M}\tilde{N}}=\tilde{\Omega}_{\tilde{M}\tilde{N% }}+L_{\tilde{M}\tilde{N}}$ , with $\tilde{\Omega}_{\tilde{M}\tilde{N}}=\frac 1{32}Str\left( \Gamma _{\tilde{M}\tilde{N}}\partial _\tau tt^{-1}\right) $ and substitute back into the action to find the version of the action given in \cite{super2t} \begin{equation} \pounds =\,\partial _\tau X_1\cdot X_2-A^{ij}X_i\cdot X_j-\frac 1{32}Str\left( \Gamma _{\tilde{M}\tilde{N}}\partial _\tau tt^{-1}\right) L^{% \tilde{M}\tilde{N}}. \end{equation} where $L_{\tilde{M}\tilde{N}}=\varepsilon ^{ij}X_i^{\tilde{M}}X_j^{\tilde{N}% }=X^{\tilde{M}}P^{\tilde{N}}-X^{\tilde{N}}P^{\tilde{M}}$ is the orbital angular momentum in 13D. A total derivative has been dropped in the first term of $\pounds $. With this we see that the canonical formalism gives $% X_1^M=X^M$ as the position and $X_2^{\tilde{M}}=P^{\tilde{M}}$ as the momentum of the 13D zero-brane. We parametrize $t\left( a\left( \tau \right) ,\Theta \left( \tau \right) \right) $ it terms of 64 fermionic $\Theta _{\tilde{\alpha}}\left( \tau \right) $ and 286+1716 bosonic $a_{\tilde{M}_1\tilde{M}_2\tilde{M}_3}\left( \tau \right) ,$ $a_{\tilde{M}_1\cdots \tilde{M}_6}\left( \tau \right) $ degrees of freedom that correspond to 13D p-forms with p=3,6. As described in \cite{super2t}, the global symmetry $g\subset $OSp$\left( 1/64\right) $ acts on the {\it right side} of $t\left( \tau \right) $ and it must be compensated by a field dependent, gauge restoring transformation $% h^{-1}\left( a,\Theta ;g\right) $ on the left side \begin{equation} t\left( a,\Theta \right) \rightarrow t\left( a^{\prime },\Theta ^{\prime }\right) =h^{-1}t\left( a,\Theta \right) g\,\,. \end{equation} $h\left( \tau \right) $ acts as a field dependent SO$\left( 11,2\right) $ gauge transformation on $\tilde{\Omega}^{\tilde{M}\tilde{N}}\left( a,\Theta \right) $ and $X_i^{\tilde{M}}$. The action remains invariant because it was built as a gauge invariant under SO$\left( 11,2\right) $ and globally invariant under OSp$\left( 1/64\right) $. The action in this gauge involves the following fields on the worldline $% X_i^{\tilde{M}}\left( \tau \right) ,$ $A^{ij}\left( \tau \right) ,$ $% a_3\left( \tau \right) ,$ $a_6\left( \tau \right) ,$ $\Theta \left( \tau \right) $. The $X_i^{\tilde{M}}\left( \tau \right) ,\Theta ^{\tilde{\alpha}% }\left( \tau \right) $ may be interpreted as the 13D superspace of two-time physics as in \cite{super2t}, however the bigger dimensions in the present case, as opposed to the 3,4,6D cases of \cite{super2t}, require the extra bosonic degrees of freedom $a_3,a_6$. From the superalgebra point of view these are closely associated with the $p$-brane charges $\tilde{J}_3,\tilde{J% }_6$ of the C-branch, hence the $a_p$ may be thought of as the $\tau $% -component of $\left( p+1\right) $-form gauge potentials $A_{p+1}$ for $% p=3,6 $. There is also additional local symmetry beyond SO$\left( 11,2\right) $ that originates with transformations on the left side of $t\left( a,\Theta \right) $ with fermionic as well as bosonic parameters $b_3(\tau ),b_6\left( \tau \right) ,\kappa ^{\tilde{\alpha}}\left( \tau \right) $ that form a generalization of kappa supersymmetry. The transformation is \begin{equation} t\left( a,\Theta \right) \rightarrow t\left( a^{^{\prime \prime }},\Theta ^{^{\prime \prime }}\right) =h^{-1}t\left( b,\kappa \right) \,t\left( a,\Theta \right) , \end{equation} where again $h\left( b,\kappa ;a,\Theta \right) $ is the induced local Lorentz transformation. For infinitesimal $b,\kappa $ we have $% t=1_{65}+b_3\Gamma ^3+b_6\Gamma ^6+\kappa _{\tilde{\alpha}}F^{\tilde{\alpha}% } $ where $\Gamma ^{3,6}$ and $F^{\tilde{\alpha}}$ are 65$\times 65$ matrices that provide representations of the generators of $G/H$. Then we find that the Lagrangian is invariant under these additional gauge symmetries provided the parameters $\kappa ,b_3,b_6$ are constrained by the following relations (where $\sim X_i\cdot X_j$ means proportional up to field dependent functions with appropriate indices) \begin{equation} L_{\tilde{M}\tilde{N}}\left( \Gamma ^{\tilde{M}\tilde{N}}\kappa \right) _{% \tilde{\alpha}}\sim L_{[\tilde{M}_1}^{\,\,\tilde{N}}b_{\tilde{M}_2\tilde{M}% _3]\tilde{N}}\sim L_{[\tilde{M}_1}^{\,\,\tilde{N}}b_{\tilde{M}_2\cdots \tilde{M}_6]\tilde{N}}\sim X_i\cdot X_j. \end{equation} The left side looks like Lorentz transformations on $\kappa ,b_3,b_6$ with a Lorentz parameter $L_{\tilde{M}\tilde{N}}$ . Then the part proportional to $% X_i\cdot X_j$ is cancelled by choosing $\delta A^{ij}$ so that \pounds\ is invariant (see \cite{super2t} for an explicit example). These gauge symmetries are more than sufficient to remove the ghosts in $\Theta ,a_3,a_6$ associated with the extra timelike dimension in 13D. We will call these gauge symmetries ``extended kappa symmetries''. We may now make various gauge choices for Sp$\left( 2\right) ,$ SO$\left( 11,2\right) $, and extended kappa symmetries, that are manifestly covariant under various spacetime interpretations (one-time) as discussed in the purely algebraic discussion in the first part of this paper. To do this we reorganize the degrees of freedom $X_i^{\tilde{M}}\left( \tau \right) ,$ $% a_3\left( \tau \right) ,$ $a_6\left( \tau \right) ,$ $\Theta _{\tilde{\alpha}% }\left( \tau \right) $ according to the representations of the various subgroups of SO$\left( 11,2\right) $ as we did for the generators of OSp$% \left( 1/64\right) $. Some of the pieces of $X_i,a_3,a_6,\Theta $ are set to zero or constants by gauge fixing. Then part of OSp$\left( 1/64\right) $ is realized linearly on the remaining degrees of freedom and the remainder is realized non-linearly. For such examples in simpler cases without supersymmetry see \cite{lifting} and with supersymmetry see \cite{super2t}. This discussion makes it evident that we can choose gauges that would realize the 11D M-algebra, the 10D type IIA, type IIB, heterotic, type-I supersymmetries, or the superalgebras describing the $AdS_n\times S^m$ backgrounds. These gauge choices provide different looking toy models that have the supersymmetries of the various corners of the moduli space of M-theory \cite{future}. Since they are all derived from the same unified 13D theory by gauge choices, they can be transformed into each other by gauge transformations that correspond to dualities in M-theory. The toy M-model given here has a much richer set of possible gauge choices along the lines of \cite{lifting} that tie together systems such as free particles, Hydrogen atom, harmonic oscillators, curved spaces, and even non-relativistic systems with more general potentials, etc. This suggests that M-theory may have similar properties. It appears that M-theory could be lifted to two-time physics with a global symmetry OSp$\left( 1/64\right) $. We are hopeful that M-theory will be better understood by studying it covariantly in 13D in the formalism of two-time physics in the presence of various gauge degrees of freedom and new gauge symmetries of the type described here.
\section{Introduction} It will be one of the primary goals of a future $e^+e^-$ linear collider (LC) or $\mu^+\mu^-$ pair collider (FMC) to measure and determine the properties of the top quark, whose existence has been confirmed at the Tevatron ($M_t=173.8\pm 5$~GeV~\cite{Vancouver1}). Although the top quark will also be object of intense studies at the Run II at the Tevatron and at the LHC, the measurements at a LC are important to fill the gaps left by the measurements in the environment of the hadron colliders. One of the most dramatic improvements attainable at a LC can be expected in the determination of the top quark mass. At the LHC, where the mass is extracted from the peak in the top invariant mass spectrum of the $W$ and $b$ originating from the top decay, a final (systematics-dominated) top mass uncertainty at the level of 2--3~GeV seems realistic. More precision will be difficult owing to unavoidable conceptual and practical problems and ambiguities in disentangling the top quark invariant mass from numerous effects in the environment of hadron colliders. At a LC the top quark mass can be determined from a measurement of the line-shape of the total cross section $\sigma(e^+e^-\to Z^*,\gamma^*\to t\bar t)$ for centre-of-mass energies around the threshold, $\sqrt{q^2}\approx 350$~GeV. The rise of the cross section with increasing centre-of-mass energy is directly correlated to the mass of the top quark. Because the total cross section describes the rate of colour singlet top-antitop events, it is theoretically and practically much better under control than the top quark invariant mass distribution. Because of the large top width ($\Gamma(t\to b\,W)= \frac{G_F}{\sqrt{2}}\frac{M_t^3}{8 \pi}\approx 1.5\,\mbox{GeV}$) the top-antitop pair cannot hadronize into toponium resonances, and the cross section represents a smooth line-shape showing only a moderate peak-like enhancement, which is the broad remnant of the $1S$ resonance. At the same time the top width effectively serves as an infrared cutoff~\cite{Fadin1} and as a natural smearing mechanism~\cite{Poggio1}, which allows us to calculate the cross section in the threshold region to high precision using perturbative QCD. It is therefore possible to reliably relate the cross section line-shape to the parameters of the Standard Model, most notably the top quark mass and the strong coupling. LC simulation studies have demonstrated that for 50--100~$fb^{-1}$ total integrated luminosity an experimental uncertainty of order 100--200~MeV can be expected in the top mass determination from a line-shape scan of the total cross section~\cite{Orange1}. Evidently, at this level of precision an adequate control over theoretical uncertainties has to be achieved. In particular, a precise definition of the top mass extracted from the experiment has to be given. For centre-of-mass energies close to the top-antitop threshold, the top quarks are produced with non-relativistic velocities $v\ll 1$. Therefore the relevant physical scales, which govern the top-antitop dynamics, the top mass $M_t$, the relative momentum $M_t v$ and the top kinetic energy $M_t v^2$, are widely separated. Because ratios of the three scales arise, the cross section close to threshold cannot be calculated using the standard multi-loop expansion in the strong coupling $\alpha_s$, but rather a double expansion in $\alpha_s$ and $v$. In the non-relativistic limit the most prominent indication of this feature is known as the ``Coulomb singularity'', which originates from the ratio $M_t/(M_t v)$. The Coulomb singularity is visible as a singular $(\alpha_s/v)^n$ behaviour in the $n$-loop QCD correction to the amplitude $\gamma\to t\bar t$ for $v\to 0$. The most economic and systematic way to tackle this problem is to employ the concept of effective theories by using the hierarchy $M_t\gg M_t v\gg M_t v^2 > \Gamma_t\gg\Lambda_{QCD}$ and by successively integrating out higher momentum effects. At leading order (LO) and next-to-leading order (NLO) in the non-relativistic expansion\footnote{ We will define what is meant by LO, NLO, NNLO, etc., in the framework of the non-relativistic expansion at the beginning of Sec.~\ref{sectionNRQCDstable}. }, the use of effective field theoretical methods seems not to be vital, because, clearly, in the first approximation the top-antitop pair can be described by a non-relativistic Schr\"odinger equation~\cite{Fadin1,Fadin2} and because, luckily, the relevant current operators do not have any anomalous dimension at the one-loop level. Beyond NLO, however, anomalous dimensions arise when relativistic effects suppressed by $v^2$ are included. This makes the use of effective field theoretical methods mandatory. In addition, the effective field theoretical approach allows for the development of a power counting scheme, which allows for a systematic identification of all effects contributing to a certain order of approximation. Those power counting rules in fact confirm that at next-to-next-to-leading order (NNLO) the top-antitop pair can be completely described by a conventional Schr\"odinger equation containing an instantaneous potential. In general, the same conclusions cannot be drawn for non-relativistic bottom-antibottom or charm-anticharm systems. A large number of theoretical studies at LO and NLO~\cite{Strassler1,Kwong1,Jezabek1,Sumino1,japaner} have been carried out in the past in order to study the feasibility of the threshold scan and other measurements at the top-antitop threshold. Recently, first NNLO QCD calculations for the total vector-current-induced cross section $\sigma(e^+e^-\to \gamma^*\to t\bar t)$ have been performed in Refs.~\cite{Hoang3,Melnikov3,Yakovlev1}. These analyses were based on non-relativistic quantum chromodynamics (NRQCD)~\cite{Caswell1,Bodwin1} and on the direct matching procedure~\cite{Hoang1,Hoang2}. In this work we extend the calculations and also determine the NNLO QCD relativistic corrections to the top quark three-momentum distribution. Interconnection effects caused by gluon exchange among top-antitop decay and production processes are not considered in this work. They are known to vanish at NLO for the total cross section~\cite{Khoze1,Khoze2,Melnikov1}, but can lead to sizeable corrections in the momentum distribution~\cite{Harlander1,Sumino2}. We also include the total cross section and the three-momentum distribution induced by the axial-vector current. Because the latter quantities are suppressed by $v^2$ with respect to the vector-current-induced ones, we only determine them at leading order in the non-relativistic expansion. It turns out that the size of the axial-vector contributions is smaller than the theoretical uncertainties contained in the dominant vector-current-induced cross section. A discussion on the size of the axial-vector-current-induced contributions can also be found in~\cite{Kuehn1}. The three-momentum distributions presented in this work represent a first step towards an exclusive treatment of the top-antitop final state at NNLO close to threshold. We analyse in detail the origin of the large NNLO corrections to the peak position and the normalization in the total cross section already observed in Refs.~\cite{Hoang3,Melnikov3,Yakovlev1}, and show that the instabilities in the peak position are a consequence of the use of the pole mass scheme. We show that the pole mass parameter is irrelevant to the peak position and define a new top quark mass, the $1S$ mass, which is more suitable to parametrize the total cross section. Whereas the $1S$ mass leads to a considerable stabilization of the peak position it does not affect the large corrections to the normalization. In this paper we also propose a NNLO generalization of the energy replacement rule ``$E\to E+i \Gamma_t$'', by Fadin and Khoze, for the implementation of the top quark width by including electroweak corrections into the matching conditions of NRQCD. In general this leads to NRQCD short-distance coefficients that have an imaginary part. The program of this paper is as follows: in Sec.~\ref{sectionNRQCDstable} we review the conceptual framework of the effective theories NRQCD and potential (P) NRQCD as far as it is relevant to the NNLO calculations carried out in this work. Section~\ref{sectionlippmannschwinger} contains a derivation of the integral equations that have to be solved and Sec.~\ref{sectionregularization} describes our cutoff regularization scheme. In Secs.~\ref{sectionNRQCDstable}, \ref{sectionlippmannschwinger} and \ref{sectionregularization} the top quark width is neglected. In Sec.~\ref{sectionwidth} we discuss the effects of the top quark width from the point of view of (P)NRQCD. Some details about our numerical methods to solve the integral equations are given in Sec.~\ref{sectionnumerics}. A first analysis of the total cross section and the three momentum distribution in the pole mass scheme is given in Sec.~\ref{sectionpolescheme}. Section~\ref{sectionuncertainties} concentrates on the origin and interpretation of the large NNLO corrections in the pole mass scheme and introduces the $1S$ mass. The relation of the $1S$ mass to other mass definitions is discussed. Section~\ref{sectionconclusion} contains our conclusions. In Appendix~\ref{appendixshortdistance}, details of the NNLO matching calculation are given in the framework of our regularization scheme. \vspace{1.5cm} \section{The Conceptual Framework - NRQCD and PNRQCD} \label{sectionNRQCDstable} In this section we review the effective field theories NRQCD and PNRQCD, which form the conceptual framework in which the NNLO corrections to the top-antitop cross section close to threshold are calculated. By N$^k$LO ($k=0,1,2,\ldots$) for the total cross section we mean a resummation of all terms proportional to $\alpha_s^m v^n$, with $m+n=1,\ldots,k+1$, in perturbation theory in $\alpha_s$ supplemented by a subsequent expansion in the top quark velocity, i.e. in the limit $\alpha_s\ll v\ll 1$. Thus at the NNLO level all terms proportional to $v\sum_{n=0}^\infty(\alpha_s/v)^n [ 1; \alpha_s, v; \alpha_s^2, \alpha_s v, v^2 ]$ have to be resummed to all orders in conventional perturbation theory in $\alpha_s$, where the dominant terms in the non-relativistic limit are determined by a non-relativistic Schr\"odinger equation with a Coulomb potential $V({\mbox{\boldmath $r$}})=- C_F \alpha_s/r$, $C_F=4/3$. In this context one has to count the strong coupling $\alpha_s$ of order $v$, as long as the renormalization scale is much larger than the typical hadronization scale $\Lambda_{QCD}$. For simplicity we postpone the effects of the top quark width until Section~\ref{sectionwidth}. NRQCD~\cite{Caswell1,Bodwin1} is an effective field theory of QCD specifically designed to handle non-relativistic heavy quark-antiquark systems. NRQCD is based on the separation of the low momentum scales $M_t v$ and $M_t v^2$, which govern the non-relativistic quark-antiquark dynamics, from the high momentum scale $M_t$, which is relevant for hard effects involved in the quark-antiquark production process and quark-antiquark and quark-gluon interactions. The NRQCD Lagrangian for the top-antitop system is obtained from QCD by integrating out all hard quark and gluon momenta of order $M_t$ or larger, and the corresponding antiparticle poles of the small components. Treating all quarks except the top as massless, the resulting non-renormalizable Lagrangian reads \begin{eqnarray} \lefteqn{ {\cal{L}}_{\mbox{\tiny NRQCD}} \, = \, - \frac{1}{2} \,\mbox{Tr} \, G^{\mu\nu} G_{\mu\nu} + \sum_{q=u,d,s,c,b} \bar q \, i \mathpalette\make@slash{D} \, q }\nonumber\\& & +\, \psi^\dagger\,\bigg[\, i D_t + c_2\,\frac{{\mbox{\boldmath $D$}}^2}{2\,M_t} + c_4\,\frac{{\mbox{\boldmath $D$}}^4}{8\,M_t^3} + \ldots \nonumber\\[2mm] & & \hspace{8mm} + \frac{c_F\,g_s}{2\,M_t}\,{\mbox{\boldmath $\sigma$}}\cdot {\mbox{\boldmath $B$}} + \, \frac{c_D\,g_s}{8\,M_t^2}\,(\,{\mbox{\boldmath $D$}}\cdot {\mbox{\boldmath $E$}}-{\mbox{\boldmath $E$}}\cdot {\mbox{\boldmath $D$}}\,) + \frac{c_S\,g_s}{8\,M_t^2}\,i\,{\mbox{\boldmath $\sigma$}}\, (\,{\mbox{\boldmath $D$}}\times {\mbox{\boldmath $E$}}-{\mbox{\boldmath $E$}}\times {\mbox{\boldmath $D$}}\,) +\ldots \,\bigg]\,\psi \nonumber\\[2mm] & & +\, \chi^\dagger\,\bigg[\, i D_t - c_2\,\frac{{\mbox{\boldmath $D$}}^2}{2\,M_t} - c_4\,\frac{{\mbox{\boldmath $D$}}^4}{8\,M_t^3} + \ldots \nonumber\\[2mm] & & \hspace{8mm} - \frac{c_F\,g_s}{2\,M_t}\,{\mbox{\boldmath $\sigma$}}\cdot {\mbox{\boldmath $B$}} + \, \frac{c_D\,g_s}{8\,M_t^2}\,(\,{\mbox{\boldmath $D$}}\cdot {\mbox{\boldmath $E$}}-{\mbox{\boldmath $E$}}\cdot {\mbox{\boldmath $D$}}\,) + \frac{c_S\,g_s}{8\,M_t^2}\,i\,{\mbox{\boldmath $\sigma$}}\, (\,{\mbox{\boldmath $D$}}\times {\mbox{\boldmath $E$}}-{\mbox{\boldmath $E$}}\times {\mbox{\boldmath $D$}}\,) +\ldots \,\bigg]\,\chi \,, \label{NRQCDLagrangian} \end{eqnarray} where only those terms are displayed explicitly which are relevant to the NNLO calculations in this work. The gluonic and light quark degrees of freedom are described by the conventional relativistic Lagrangian, where $G^{\mu\nu}$ is the gluon strength field tensor, $q$ the Dirac spinor of a massless quark. The non-relativistic top and antitop quark are described by the Pauli spinors $\psi$ and $\chi$, respectively. For convenience all colour indices are suppressed and summations over colour indices are understood. $D_t$ and {\boldmath $D$} are the time and space components of the gauge covariant derivative $D_\mu$, and $E^i = G^{0 i}$ and $B^i = \frac{1}{2}\epsilon^{i j k} G^{j k}$ the electric and magnetic components of the gluon field strength tensor. The short-distance coefficients $c_2,c_4,c_F,c_D,c_S$ are normalized to one at the Born level. The subscripts $F$, $D$ and $S$ stand for Fermi, Darwin and spin-orbit. We emphasize that the mass parameter $M_t$ used for the formulation of NRQCD is the top quark pole mass. Although it is known that this choice can lead to a bad behaviour of the perturbative coefficients at large orders, the pole mass is still the most convenient mass parameter to be used at this stage, because the formulation of NRQCD is particularly simple in this scheme. In addition to integrating out hard quark and gluon momenta and the small components in the QCD Lagrangian one also has to do the same in the vector ($j_\mu^v=\bar t\gamma_\mu t$) and the axial-vector currents ($j_\mu^a=\bar t\gamma_\mu\gamma_5 t$), which produce and annihilate the top-antitop pair close to the threshold with centre-of-mass energy $\sqrt{q^2}$. This means that we have to expand the respective QCD currents in terms of NRQCD currents carrying the proper quantum numbers. In momentum space representation the expansion of the QCD vector current in terms of ${}^3\!S_1$ NRQCD currents reads ($k=1,2,3$) \begin{eqnarray} \tilde j_k^v(q) & = & c^v_1\,\Big({\tilde \psi}^\dagger \sigma_k \tilde \chi\Big)(q) - \frac{c^v_2}{6 M_t^2}\,\Big({\tilde \psi}^\dagger \sigma_k (\mbox{$-\frac{i}{2}$} \stackrel{\leftrightarrow}{\mbox{\boldmath $D$}})^2 \tilde \chi\Big)(q) + \ldots \,, \label{vectorcurrentexpansion1} \\ \tilde j_k^v(-q) & = & c^v_1\,\Big({\tilde \chi}^\dagger \sigma_k \tilde \psi\Big)(-q) - \frac{c^v_2}{6 M_t^2}\,\Big({\tilde \chi}^\dagger \sigma_k (\mbox{$-\frac{i}{2}$} \stackrel{\leftrightarrow}{\mbox{\boldmath $D$}})^2 \tilde \psi\Big)(-q) + \ldots \,, \label{vectorcurrentexpansion2} \end{eqnarray} and the expansion of the QCD axial-vector current in terms of ${}^3\!P_1$ NRQCD currents \begin{eqnarray} \tilde j_k^a(q) & = & \frac{c^a_1}{M_t}\,\Big({\tilde \psi}^\dagger (\mbox{$-\frac{i}{2}$} \stackrel{\leftrightarrow}{\mbox{\boldmath $D$}}\!\! \times{\mbox{\boldmath $\sigma$}})_k \tilde \chi\Big)(q) + \ldots \,, \label{axialvectorcurrentexpansion1} \\ \tilde j_k^a(-q) & = & \frac{c^a_1}{M_t}\,\Big({\tilde \chi}^\dagger (\mbox{$-\frac{i}{2}$} \stackrel{\leftrightarrow}{\mbox{\boldmath $D$}}\!\! \times{\mbox{\boldmath $\sigma$}})_k \tilde \psi\Big)(-q) + \ldots \,, \label{axialvectorcurrentexpansion2} \end{eqnarray} where the constants $c^v_{1,2}$ and $c^a_1$ are the short-distance coefficients normalized to one at the Born level. The time components of the currents do not contribute because the trace over the massless lepton fields that describe the $e^+e^- $ annihilation process is proportional to $(\delta^{ij}-e^i e^j)$, $(e^1,e^2,e^3)$ being the unit-vector pointing into the centre-of-mass electron direction. In addition, the zero component of the vector current vanishes. The dominant NRQCD current in the expansion of the QCD vector current has dimension three. Thus for a NNLO description of the cross section we have to expand the QCD vector current in NRQCD currents up to dimension five. The QCD axial-vector current only needs to be expanded up to dimension four. For the NNLO calculation of the cross section only the ${\cal{O}}(\alpha_s^2)$ short-distance corrections to the coefficient $c^v_1$ have to be calculated. To formulate the total $t\bar t$ production cross sections in $e^+e^-$ annihilation in the non-relativistic region at NNLO in NRQCD, we first define the vector and axial-vector-current-induced cross sections using the optical theorem and starting from their corresponding expressions in full QCD: \begin{eqnarray} R^v(q^2) & = & \frac{4\,\pi}{q^2}\,\mbox{Im}\,\bigg[\, -i\,\int\,d^4x\,e^{i\,q.x}\, \langle\, 0\,| T\,j^v_i(x) \, j^{v\,i}(0)\, |\, 0\,\rangle\,\bigg] \nonumber\\[2mm] & \equiv & \frac{4\,\pi}{q^2}\,\mbox{Im}\,[\,-i\, \langle\, 0\,| T\, \tilde j^v_i(q) \, \tilde j^{v\,i}(-q)\, |\, 0\,\rangle\,] \,, \label{vectorcrosssectioncovariant} \\[3mm] R^a(q^2) & = & \frac{4\,\pi}{q^2}\,\mbox{Im}\,\bigg[\, -i\,\int\,d^4x\,e^{i\,q.x}\, \langle\, 0\,| T\,j^a_i(x) \, j^{a\,i}(0)\, |\, 0\,\rangle\,\bigg] \nonumber\\[2mm] & \equiv & \frac{4\,\pi}{q^2}\,\mbox{Im}\,[\,-i\, \langle\, 0\,| T\, \tilde j^a_i(q) \, \tilde j^{a\,i}(-q)\, |\, 0\,\rangle] \,. \label{axialvectorcrosssectioncovariant} \end{eqnarray} In terms of $R^v$ and $R^a$ the total cross section $\sigma_{tot}^{\gamma,Z}(e^+e^-\to \gamma^*, Z^*\to t\bar t)$ reads \begin{eqnarray} \sigma_{tot}^{\gamma,Z}(q^2) & = & \sigma_{pt}\, \bigg[\, Q_t^2 - 2\,\frac{q^2}{q^2-M_Z^2}\,v_e\,v_t\,Q_t + \bigg(\frac{q^2}{q^2-M_Z^2}\bigg)^2\, \Big[ v_e^2+a_e^2 \Big]\,v_t^2 \,\bigg]\,R^v(q^2) \nonumber \\[2mm] & & +\,\sigma_{pt}\,\bigg(\frac{q^2}{q^2-M_Z^2}\bigg)^2\, \Big[ v_e^2 + a_e^2 \Big]\,a_t^2\,R^a(q^2) \,, \label{totalcrossfullQCD} \end{eqnarray} where \begin{eqnarray} \sigma_{pt} & = & \frac{4\,\pi\,\alpha^2}{3\,q^2} \,, \\[3mm] v_f & = & \frac{T_3^f - 2\,Q_f\,\sin^2\theta_W}{2\, \sin\theta_W \,\cos\theta_W} \,, \\[2mm] a_f & = & \frac{T_3^f}{2\, \sin\theta_W\, \cos\theta_W} \,. \end{eqnarray} Here, $\alpha$ is the fine structure constant, $Q_t=2/3$ the electric charge of the top quark, $\theta_W$ the Weinberg angle, and $T_3^f$ refers to the third component of the weak isospin; $Q_t^2 R^v$ is equal to the total normalized photon-induced cross section, which is usually referred to as the $R$-ratio. To determine $R^v$ and $R^a$ at NNLO in NRQCD we insert the expansions in Eqs.~(\ref{vectorcurrentexpansion1})--(\ref{axialvectorcurrentexpansion2}) into Eqs.~(\ref{vectorcrosssectioncovariant}) and (\ref{axialvectorcrosssectioncovariant}). This leads to the expressions \begin{eqnarray} R_{\mbox{\tiny NNLO}}^{v,\mbox{\tiny thr}}(q^2) & = & \frac{4\,\pi}{q^2}\,C^v\, \mbox{Im}\Big[\, {\cal{A}}^v(q^2) \,\Big] + \ldots \,, \label{vectorcrosssectionexpanded} \\[3mm] R_{\mbox{\tiny NNLO}}^{a,\mbox{\tiny thr}}(q^2) & = & \frac{4\,\pi}{q^2}\,C^a\, \mbox{Im}\Big[\, {\cal{A}}^a(q^2) \,\Big] + \ldots \,, \label{axialvectorcrosssectionexpanded} \end{eqnarray} where \begin{eqnarray} {\cal{A}}^v & = & i\,\Big\langle \, 0 \, \Big| \, \Big({\tilde\psi}^\dagger \vec\sigma \, \tilde \chi + \frac{1}{6 M_t^2} {\tilde\psi}^\dagger \vec\sigma \, (\mbox{$-\frac{i}{2}$} \stackrel{\leftrightarrow}{\mbox{\boldmath $D$}})^2 \tilde \chi \Big)\, \, \Big({\tilde\chi}^\dagger \vec\sigma \, \tilde \psi + \frac{1}{6 M_t^2}{\tilde\chi}^\dagger \vec\sigma \, (\mbox{$-\frac{i}{2}$} \stackrel{\leftrightarrow}{\mbox{\boldmath $D$}})^2 \tilde \psi \Big)\, \Big| \, 0 \, \Big\rangle \,, \label{correlatorV} \\[2mm] {\cal{A}}^a & = & i\,\Big\langle \, 0 \, \Big| \, \Big({\tilde\psi}^\dagger (\mbox{$-\frac{i}{2}$} \stackrel{\leftrightarrow}{\mbox{\boldmath $D$}}\!\! \times{\mbox{\boldmath $\sigma$}}) \, \tilde \chi \Big)\, \, \Big({\tilde\chi}^\dagger (\mbox{$-\frac{i}{2}$} \stackrel{\leftrightarrow}{\mbox{\boldmath $D$}}\!\! \times{\mbox{\boldmath $\sigma$}}) \, \tilde \psi \Big)\, \Big| \, 0 \, \Big\rangle \,, \label{correlatorA} \end{eqnarray} and \begin{eqnarray} C^v & = & (c_1^v)^2 \,, \label{Cvdef} \\[2mm] C^a & = & 1 \,. \label{Cadef} \end{eqnarray} The expressions for $R^v$ and $R^a$ at NNLO in the non-relativistic expansion in Eqs.~(\ref{vectorcrosssectionexpanded}) and (\ref{axialvectorcrosssectionexpanded}) represent an application of the factorization formalism proposed by Bodwin, Braaten and Lepage~\cite{Bodwin1}. The cross sections are written as a sum of absorptive parts of non-relativistic current correlators, each of which is multiplied by a short-distance coefficient. The vector correlator ${\cal{A}}^v$ describes the top-antitop system produced in an S-wave spin triplet state and the axial-vector correlator ${\cal{A}}^a$ describes the system in a corresponding P-wave triplet state. The axial vector-current-induced cross section is suppressed by $v^2$ with respect to the vector-current-induced cross section. We note that the correlators are defined within a proper regularization scheme. For convenience we have not expanded ${\cal{A}}^v$ into a sum of a dimension six and a dimension eight current correlator. The short distance coefficients $C^v$ and $C^a$ encode the effects of quark and gluon momenta of order the top mass or larger in top-antitop production and annihilation vertex diagrams. As the non-relativistic correlators they are regularization-scheme-dependent. In principle the non-relativistic correlators ${\cal{A}}^v$ and ${\cal{A}}^a$ could be calculated from the Feynman rules derived from the NRQCD Lagrangian~(\ref{NRQCDLagrangian}). Such a task, however, would be quite cumbersome, because an infinite number of diagrams would still have to be resummed. Clearly, we would like to derive an equation of motion for the off-shell top quark four point Green function, which in the non-relativistic limit reduces to the well known Schr\"odinger equation, which automatically carries out the resummation of the relevant diagrams. A formal and systematic way to achieve that is to also integrate out top quark and gluon modes carrying momenta of the order the inverse Bohr radius $\sim M_t v$. The resulting effective field theory is called ``potential NRQCD'' (PNRQCD)~\cite{Pineda1}. The basic ingredient to construct PNRQCD is to identify the relevant physical momentum regions in the description of heavy quark-antiquark systems in the framework of NRQCD. Those momentum regions have been identified in Ref.~\cite{Beneke1} by constructing an asymptotic expansion of Feynman diagrams describing heavy quark-antiquark production or annihilation close to threshold in terms of the heavy quark centre-of-mass velocity. Because NRQCD is not Lorentz-covariant, the time and the spatial components of the momenta have to be treated independently. There are momentum regions where time and spatial components are of a different order in the velocity counting. The relevant momentum regions are ``soft'' ($k^0\sim M_t v$, $\vec k\sim M_t v$), ``potential'' ($k^0\sim M_t v^2$, $\vec k\sim M_t v$) and ``ultrasoft'' ($k^0\sim M_t v^2$, $\vec k\sim M_t v^2$). It can be shown that heavy quarks and gluons can have soft and potential momenta, but only gluons can have ultrasoft momenta. A momentum region with $k^0\sim M_t v$, $\vec k\sim M_t v^2$ does not exist. It is in principle not excluded that there are momentum regions scaling like $M_t v^n$ with $n > 2$, but even if such regions existed they would be irrelevant to top quark production because they would represent modes below the hadronization scale. To construct PNRQCD one integrates out ``soft'' heavy quarks and gluons ($k^0\sim M_t v$, $\vec k\sim M_t v$) and ``potential'' gluons ($k^0\sim M_t v^2$, $\vec k\sim M_t v$), supplemented by an expansion in momentum components of order $M_t v^2$. Heavy quarks carrying potential momenta and gluons with ultrasoft momenta are kept as dynamical fields. This leads to spatially non-local four (heavy) quark operators which represent a coupling of a quark-antiquark pair separated by distances of order of the Bohr radius $\sim 1/M_t v$. For a quark-antiquark pair in a colour singlet state this non-local interaction is nothing else than the instantaneous potential of a quark-antiquark separated by a distance of order the inverse Bohr radius. Generically the PNRQCD Lagrangian has the form \begin{eqnarray} {\cal{L}}_{\mbox{\tiny PNRQCD}} & = & \tilde{\cal{ L}}_{\mbox{\tiny NRQCD}} \, + \, \int d^3 \mbox{\boldmath $r$} \Big(\psi^\dagger \psi\Big) (\mbox{\boldmath $r$})\,V(\mbox{\boldmath $r$})\, \Big(\chi^\dagger \chi\Big)(0) \,, \label{PNRQCDLagrangian} \end{eqnarray} where the tilde above ${\cal{ L}}_{\mbox{\tiny NRQCD}}$ on the RHS of Eq.~(\ref{PNRQCDLagrangian}) indicates that the corresponding operators only describe potential quark and ultrasoft gluonic degrees of freedom. In addition, an expansion in momentum components $\sim M_t v^2$ is understood. $V$ is the heavy quark-antiquark potential and is given below. Using the velocity counting rules for potential heavy quarks mentioned above we see that the LO contribution to $V$, the Coulomb potential $-C_F\alpha_s/|\mbox{\boldmath $r$}|$, counts as $v^2$, i.e. it is of the same order as the kinetic energy. Thus, as is well known, in the non-relativistic limit, the Coulombic interaction between the heavy quark pair has to be treated exactly rather than perturbatively. We note that the expansion of the heavy quark currents in terms of NRQCD currents (Eqs.~(\ref{vectorcurrentexpansion1} )--(\ref{axialvectorcurrentexpansion2})) is in general also affected by going from NRQCD to PNRQCD, because also in the NRQCD currents soft and potential gluonic degrees of freedom have to be integrated out. However, at NNLO this does not affect the results displayed in Eqs.~(\ref{vectorcurrentexpansion1}) and (\ref{axialvectorcurrentexpansion2}). The only (and fortunate) practical consequence is that we can neglect the gluonic contribution in the covariant derivatives. Collecting all terms from the PNRQCD Lagrangian which contribute at NNLO, i.e. count as $v^2$, $v^3$ or $v^4$, one can derive the following equation of motion in momentum space representation for the Green function of the time-independent Schr\"odinger equation, valid at NNLO in the non-relativistic expansion: \begin{equation} \bigg[\, \frac{\mbox{\boldmath $k$}^2}{M_t} - \frac{\mbox{\boldmath $k$}^4}{4M_t^3} \,-\,\bigg(\, \frac{p_0^2}{M_t} - \frac{p_0^4}{4 M_t^3} \,\bigg) \,\bigg]\, \tilde G(\mbox{\boldmath $k$},\mbox{\boldmath $k$}^\prime;q^2) \,+\, \int\frac{d^3 \mbox{\boldmath $p$}^\prime}{(2\,\pi)^3}\, \tilde V(\mbox{\boldmath $k$},\mbox{\boldmath $p$}^\prime)\, \tilde G(\mbox{\boldmath $p$}^\prime,\mbox{\boldmath $k$}^\prime;q^2) \, = \, (2\,\pi)^3\,\delta^{(3)}(\mbox{\boldmath $k$}-\mbox{\boldmath $k$}^\prime) \,, \label{NNLOSchroedinger} \end{equation} where \begin{eqnarray} \tilde V(\mbox{\boldmath $k$},\mbox{\boldmath $k$}^\prime) & = & \tilde V_c(\mbox{\boldmath $k$}-\mbox{\boldmath $k$}^\prime) + \tilde V_{\mbox{\tiny BF}}(\mbox{\boldmath $k$}, \mbox{\boldmath $k$}^\prime) + \tilde V_{\mbox{\tiny NA}}(\mbox{\boldmath $k$}- \mbox{\boldmath $k$}^\prime) \,, \label{NNLOpotential} \end{eqnarray} and \begin{eqnarray} p_0^2 & = & \frac{q^2}{4} - M_t^2 \,. \end{eqnarray} The parameter $p_0$ is equal to the centre-of-mass three-momentum of the top quarks. We have chosen this rather unusual representation for the energy parameter since it greatly simplifies, because of its symmetric form in Eq.~(\ref{NNLOSchroedinger}), the analytic matching calculations we carry out in App.~\ref{appendixshortdistance}. The same trick has already been used in~\cite{Hoang2} and later in \cite{Hoang3,Melnikov3,Yakovlev1}. It is a non-trivial fact that the PNRQCD operators, which describe interactions of the heavy quarks with ultrasoft gluons, do not contribute at NNLO, and that a conventional two-body Schr\"odinger equation with an instantaneous potential is fully capable of resumming all terms $\propto\alpha_s^m v^n$ that belong to NNLO. We come back to this issue at the end of this section. The individual potentials in momentum space representation read ($a_s\equiv\alpha_s(\mu)$, $C_A=3$, $C_F=4/3$, $T=1/2$, $\mbox{\boldmath $Q$}\equiv\mbox{\boldmath $k$}-\mbox{\boldmath $k$}^\prime$): \begin{eqnarray} \tilde V_c(\mbox{\boldmath $k$}) & = & -\,\frac{4\,\pi\,C_F\,a_s}{\mbox{\boldmath $k$}^2}\, \bigg\{\, 1 + \Big(\frac{a_s}{4\,\pi}\Big)\,\Big[\, -\beta_0\,\ln\Big(\frac{\mbox{\boldmath $k$}^2}{\mu^2}\Big) + a_1 \,\Big] \nonumber\\[2mm] & & \quad + \Big(\frac{a_s}{4\,\pi}\Big)^2\,\Big[\, \beta_0^2\,\ln^2\Big(\frac{\mbox{\boldmath $k$}^2}{\mu^2}\Big) - \Big(2\,\beta_0\,a_1 + \beta_1\Big)\,\ln\Big(\frac{\mbox{\boldmath $k$}^2}{\mu^2}\Big) + a_2 \,\Big] \,\bigg\} \,, \label{NNLOCoulomb} \\[3mm] \tilde V_{\mbox{\tiny BF}}(\mbox{\boldmath $k$}, \mbox{\boldmath $k$}^\prime) & = & \frac{\pi\,C_F\,a_s}{M_t^2} + \frac{4\,\pi\,C_F\,a_s}{M_t^2}\, \bigg[\, \mbox{\boldmath $S_t$}\mbox{\boldmath $S_{\bar t}$} - \frac{(\mbox{\boldmath $Q$\boldmath $S_t$}) (\mbox{\boldmath $Q$}\mbox{\boldmath $S_{\bar t}$})} {\mbox{\boldmath $Q$}^2} \,\bigg] \nonumber \\[2mm] & & -\,\frac{\pi\,C_F\,a_s}{M_t^2}\, \bigg[\, \frac{(\mbox{\boldmath $k$}+\mbox{\boldmath $k$}^\prime)^2} {\mbox{\boldmath $Q$}^2} - \frac{(\mbox{\boldmath $k$}^2-{\mbox{\boldmath $k$}^\prime}^2)^2} {\mbox{\boldmath $Q$}^4} \,\bigg] \nonumber\\[2mm] & & +\,6\,i\,\frac{\pi\,C_F\,a_s}{M_t^2}\, \frac{(\mbox{\boldmath $S_t$}+\mbox{\boldmath $S_{\bar t}$}) (\mbox{\boldmath $k$}\times\mbox{\boldmath $k$}^\prime)}{\mbox{\boldmath $Q$}^2} \label{NNLOBF} \,,\\[3mm] \tilde V_{\mbox{\tiny NA}}(\mbox{\boldmath $k$}) & = & -\,\frac{\pi^2\,C_A\,C_F\,a_s^2}{M_t\,|\mbox{\boldmath $k$}|} \,, \label{NNLONA} \end{eqnarray} where $\mbox{\boldmath $S_t$}$ and $\mbox{\boldmath $S_{\bar t}$}$ are the top and antitop spin operators and ($n_l=5$) \begin{eqnarray} \beta_0 & = & \frac{11}{3}\,C_A - \frac{4}{3}\,T\,n_l \,, \nonumber\\[2mm] \beta_1 & = & \frac{34}{3}\,C_A^2 -\frac{20}{3}C_A\,T\,n_l - 4\,C_F\,T\,n_l \,, \nonumber\\[2mm] a_1 & = & \frac{31}{9}\,C_A - \frac{20}{9}\,T\,n_l \,, \nonumber\\[2mm] a_2 & = & \bigg(\,\frac{4343}{162}+4\,\pi^2-\frac{\pi^4}{4} +\frac{22}{3}\,\zeta_3\,\bigg)\,C_A^2 -\bigg(\,\frac{1798}{81}+\frac{56}{3}\,\zeta_3\,\bigg)\,C_A\,T\,n_l \nonumber\\[2mm] & & -\bigg(\,\frac{55}{3}-16\,\zeta_3\,\bigg)\,C_F\,T\,n_l +\bigg(\,\frac{20}{9}\,T\,n_l\,\bigg)^2 \,. \end{eqnarray} The constants $\beta_0$ and $\beta_1$ are the one- and two-loop coefficients of the QCD beta function and $\gamma=0.577216\ldots\,$ is the Euler constant; $V_c$ is the Coulomb (static) potential. Its ${\cal{O}}(\alpha_s)$ and ${\cal{O}}(\alpha_s^2)$ corrections, which come from loops carrying soft momenta, have been determined in~\cite{Fischler1,Billoire1} and \cite{Schroeder1,Peter1}, respectively.\footnote{ The constant $a_2$ was first calculated in Ref.~\cite{Peter1}. In Ref.~\cite{Schroeder1} an error in the coefficient of the term $\propto \pi^2 C_A^2$ was corrected. } $V_{\mbox{\tiny BF}}$ is the Breit-Fermi potential known from positronium. It describes the Darwin and the spin-orbit interactions mediated by longitudinal gluons and the hyperfine interactions mediated by transverse gluons in the potential momentum region. Then, $V_{\mbox{\tiny NA}}$ is a purely non-Abelian potential generated (in Coulomb gauge) through a non-analytic term in a soft momentum one-loop vertex correction to the Coulomb potential involving the triple gluon vertex~\cite{Gupta1,Gupta2} (see also Ref.~\cite{Kummer1}). We have already noted that it is a non-trivial fact that ultrasoft gluons do not contribute at NNLO, which means that a common two-body Schr\"odinger equation, i.e. a wave equation containing an instantaneous interaction potential, is indeed capable of resumming all the terms that we count as NNLO. Although this is a well accepted fact for positronium and the hydrogen atom in QED~\cite{Bethe1}, it is not at all trivial to understand, even in the QED case, if one has to rely on arguments that are not in the framework of (P)NRQCD. In QED the corrections caused by ultrasoft photons are known as retardation effects. The Lamb shift in hydrogen is the most famous example. Based on the identification of the various momentum regions for quarks and gluons mentioned above, however, one has transparent power counting rules at hand; these show that the non-instantaneous exchange of gluons among the top quarks does not lead to any effects at NNLO. For the validity of the argument it is important that the scale $M_t v^2$ is much larger than the typical hadronization scale $\Lambda_{QCD}$.\footnote{ Even for energies very close to threshold, the scale $M_t v^2$ cannot become smaller than $\Gamma_t$ since the dominant effect of the top width is to effectively shift the energy into the positive complex plane by an amount $\Gamma_t$ (see Section~\ref{sectionwidth}). } To see that retardation effects are suppressed by at least three powers of $v$ with respect to the non-relativistic limit (LO), we recall that the only source of non-instantaneous interactions in PNRQCD are the ultrasoft gluons, for which all momentum components are of order $M_t v^2$. In addition, only transversely polarized gluons need to be considered as ultrasoft, since we can work in the Coulomb gauge where the time component of the longitudinal gluons vanishes. Thus, an exchange of an ultrasoft gluon among the heavy quark-antiquark pair is already suppressed by $v^2$ with respect to LO from the coupling of transverse gluons to the heavy quarks. To see that an additional power of $v$ arises from the loop integration over the ultrasoft gluon momentum, we compare the $v$-counting of the integration measure and the gluon propagator for ultrasoft and potential momenta. In the ultrasoft case, the product of the integration measure $d^4 k$ and the gluon propagator $1/k^2$ counts as $v^8\times v^{-4}=v^4$, whereas in the potential case the result reads $v^5\times v^{-2}=v^3$. Because the potential momenta contribute at LO we find that the ultrasoft gluons can indeed only lead to effects beyond NNLO. Even if the gluon self coupling is taken into account, this conclusion remains true, because it also leads to additional powers of $v$ if ultrasoft momenta are involved. We note that in arbitrary gauge the conclusions are true only after all gauge cancellations have been taken into account. The relation $M_t v^2 \gg \Lambda_{QCD}$ is needed for the above argumentation: otherwise, the coupling of ultrasoft gluons to the heavy quarks or among themselves, $\alpha_s(M_t v^2)$, could be of order $1$. Therefore our conclusion that retardation effects do not contribute at NNLO would not be valid for the bottom quark\footnote{ For the case of bottom-antibottom quark sum rules the conclusions can, however, still be correct if the effective smearing range is chosen larger than $\Lambda_{QCD}$.~\cite{Hoang5} } or even the charm quark case. \vspace{1.5cm} \section{Lippmann-Schwinger Equation} \label{sectionlippmannschwinger} The non-relativistic current correlators in the NRQCD factorization formulae for the total top-antitop cross section close to threshold, Eqs.~(\ref{vectorcrosssectionexpanded}) and (\ref{axialvectorcrosssectionexpanded}), are directly related to the Green function $\tilde G({\mbox{\boldmath $k$}},{\mbox{\boldmath $k$}^\prime};q^2)$ of the Schr\"odinger equation~(\ref{NNLOSchroedinger}), which describes off-shell elastic scattering of a top-antitop pair with centre-of-mass three momentum $\pm{\mbox{\boldmath $k$}}$ into a top-antitop pair with three momentum $\pm{\mbox{\boldmath $k$}^\prime}$. We emphasize that the Green function does not describe the scattering of on-shell top quarks because the common three-dimensional formulation in the form of the Schr\"odinger equation~(\ref{NNLOSchroedinger}) already contains an implicit integration over the zero-components of the momenta in the heavy quark propagators. Because the heavy quark potential is energy-independent this integration is trivial by residues. In the first part of this section we give the relations of the current correlators to the Green function in the three-dimensional formulation. In the second part, we present the generalization to four dimensions for those results that are essential for a proper treatment of the $W^+W^-b\bar b$ phase space once the top quark width is taken into account. For this section we still assume that the top quark is stable. Taking into account the partial wave decomposition of the Green function of the Schr\"odinger equation~(\ref{NNLOSchroedinger}) \begin{eqnarray} \tilde G({\mbox{\boldmath $k$}},{\mbox{\boldmath $k$}^\prime}) & = & \sum_{l=0}^\infty \tilde G^l({\mbox{\boldmath $k$}},{\mbox{\boldmath $k$}^\prime}) \,, \label{Greenfunctiondecomposition} \end{eqnarray} where $l$ is the total angular momentum quantum number, we find the following relation between the NRQCD non-relativistic current correlators~(\ref{correlatorV}) and (\ref{correlatorA}) and the $S$ and $P$ wave contributions to $\tilde G$: \begin{eqnarray} {\cal{A}}^v(q^2) & = & N_c\,\mbox{Tr} \int\frac{d^3 \mbox{\boldmath $k$}}{(2\pi)^3} \int\frac{d^3 \mbox{\boldmath $k$}^\prime}{(2\pi)^3}\, \mbox{\boldmath $\sigma$}\, \bigg(1+\frac{\mbox{\boldmath $k$}^2}{6 M_t^2}\bigg) \tilde G^0({\mbox{\boldmath $k$}},{\mbox{\boldmath $k$}^\prime})\, \bigg(1+\frac{{\mbox{\boldmath $k$}^\prime}^2}{6 M_t^2}\bigg)\, \mbox{\boldmath $\sigma$} \nonumber\\[2mm] & = & 6\,N_c\, \int\frac{d^3 \mbox{\boldmath $k$}}{(2\pi)^3} \int\frac{d^3 \mbox{\boldmath $k$}^\prime}{(2\pi)^3}\, \bigg(1+\frac{\mbox{\boldmath $k$}^2}{6 M_t^2}\bigg) \tilde G^0({\mbox{\boldmath $k$}},{\mbox{\boldmath $k$}^\prime})\, \bigg(1+\frac{{\mbox{\boldmath $k$}^\prime}^2}{6 M_t^2}\bigg)\, \,, \nonumber \label{correlatorGreenV} \\[4mm] {\cal{A}}^a(q^2) & = & N_c\,\mbox{Tr} \int\frac{d^3 \mbox{\boldmath $k$}}{(2\pi)^3} \int\frac{d^3 \mbox{\boldmath $k$}^\prime}{(2\pi)^3}\, \frac{\mbox{\boldmath $k$}\times \mbox{\boldmath $\sigma$}}{M_t} \tilde G^1({\mbox{\boldmath $k$}},{\mbox{\boldmath $k$}^\prime}) \frac{\mbox{\boldmath $k$}^\prime\times \mbox{\boldmath $\sigma$}}{M_t} \nonumber\\[2mm] & = & 4\,N_c\, \int\frac{d^3 \mbox{\boldmath $k$}}{(2\pi)^3} \int\frac{d^3 \mbox{\boldmath $k$}^\prime}{(2\pi)^3}\, \frac{\mbox{\boldmath $k$}\mbox{\boldmath $k$}^\prime}{M_t^2}\, \tilde G^1({\mbox{\boldmath $k$}},{\mbox{\boldmath $k$}^\prime}) \,, \label{correlatorGreenA} \end{eqnarray} where a proper UV regularization is understood. For simplicity we have dropped the energy argument of the Green function $\tilde G$. In this work we use the Lippmann-Schwinger equation, the Fourier transform of Eq.~(\ref{NNLOSchroedinger}) with respect to $\mbox{\boldmath $k$}^\prime$, \begin{eqnarray} \bigg[\, \frac{\mbox{\boldmath $k$}^2}{M_t} - \frac{\mbox{\boldmath $k$}^4}{4M_t^3} \,-\,\bigg(\, \frac{p_0^2}{M_t} - \frac{p_0^4}{4 M_t^3} \,\bigg) \,\bigg]\, G(\mbox{\boldmath $k$},\mbox{\boldmath $x$}) \, = \, \exp(i\,\mbox{\boldmath $k$}\mbox{\boldmath $x$}) \,-\, \int\frac{d^3 \mbox{\boldmath $p$}^\prime}{(2\,\pi)^3}\, \tilde V(\mbox{\boldmath $k$},\mbox{\boldmath $p$}^\prime)\, G(\mbox{\boldmath $p$}^\prime,\mbox{\boldmath $x$}) \,, \label{LippmannSchwinger} \end{eqnarray} to derive integral equations for $\tilde G^0$ and $\tilde G^1$, which are then solved numerically. Using the partial wave decomposition of $\exp(i\,\mbox{\boldmath $k$}\mbox{\boldmath $x$})$ ($k\equiv|\mbox{\boldmath $k$}|$, $x\equiv|\mbox{\boldmath $x$}|$): \begin{equation} \exp(i \,\mbox{\boldmath $k$}\mbox{\boldmath $x$}) \, = \, \exp(i\, k\,x\,\cos\theta) \, = \, \sum_{l=0}^\infty i^l\,(2\,l+1)\,j_l(k\,x)\,P_l(\cos\theta) \,, \end{equation} where $j_l$ are the spherical Bessel functions $j_l(x)=(-x)^l(\frac{1}{x} \frac{d}{d x})^l [\frac{\sin x}{x}]$ and $P_l$ the Legendre polynomials, one arrives at the following equations for $G^0$ and $G^1$: \begin{eqnarray} \bigg[\, \frac{\mbox{\boldmath $k$}^2}{M_t} - \frac{\mbox{\boldmath $k$}^4}{4M_t^3} \,-\,\bigg(\, \frac{p_0^2}{M_t} - \frac{p_0^4}{4 M_t^3} \,\bigg) \,\bigg]\, G^0(\mbox{\boldmath $k$},\mbox{\boldmath $x$}) & = & \frac{\sin(k \,x)}{k \,x} \, - \, \int\frac{d^3 \mbox{\boldmath $p$}^\prime}{(2\pi)^3}\, \tilde V(\mbox{\boldmath $k$},\mbox{\boldmath $p$}^\prime)\, G^0(\mbox{\boldmath $p$}^\prime,\mbox{\boldmath $x$}) \,, \label{inteqG0} \\[4mm] \bigg[\, \frac{\mbox{\boldmath $k$}^2}{M_t} \,-\, \frac{p_0^2}{M_t} \,\bigg]\, G^1(\mbox{\boldmath $k$},\mbox{\boldmath $x$}) & = & 3\, i\,\frac{\mbox{\boldmath $k$}\mbox{\boldmath $x$}}{k^2 \,x^2}\, \bigg(\,\frac{\sin(k \,x)}{k \,x}-\cos(k \,x)\,\bigg) \nonumber \\[2mm] & & - \, \int\frac{d^3 \mbox{\boldmath $p$}^\prime}{(2\pi)^3}\, \tilde V(\mbox{\boldmath $k$},\mbox{\boldmath $p$}^\prime)\, G^1(\mbox{\boldmath $p$}^\prime,\mbox{\boldmath $x$}) \,. \label{inteqG1} \end{eqnarray} Because in Eq.~(\ref{inteqG0}) only S-wave states are considered, one can, instead of the complicated form of $\tilde V_{\mbox{\tiny BF}}$, use the angular average with respect to the angle between $\mbox{\boldmath $p^\prime$}$ and $\mbox{\boldmath $k$}$ of the $1/\mbox{\boldmath $Q$}^4$ term on the RHS of Eq.~(\ref{NNLOBF}). Evaluating also the spin matrices for the S-wave state, the Breit-Fermi potential simplifies to \begin{eqnarray} \tilde V^s_{\mbox{\tiny BF}}(\mbox{\boldmath $k$}, \mbox{\boldmath $k$}^\prime) & = & \frac{1}{4\pi}\,\int d\Omega\, \tilde V_{\mbox{\tiny BF}}(\mbox{\boldmath $k$}, \mbox{\boldmath $k$}^\prime) \nonumber \\[2mm] & = & \frac{11}{3}\,\frac{\pi\,C_F\,a_s}{M_t^2}\, -2\,\frac{\pi\,C_F\,a_s}{M_t^2}\, \frac{\mbox{\boldmath $k$}^2+{\mbox{\boldmath $k$}^\prime}^2} {\mbox{\boldmath $Q$}^2} \,. \end{eqnarray} In Eq.~(\ref{inteqG1}), on the other hand, we will just use the LO Coulomb potential because the axial-vector contribution to the total cross section is already suppressed by $v^2$. For the same reason we do not include any kinematic relativistic corrections in Eq.~(\ref{inteqG1}). Defining the S-wave and the P-wave vertex Green function as \begin{eqnarray} \label{Svertex} S(\mbox{\boldmath $k$}) & = & \int\frac{d^3 \mbox{\boldmath $p$}^\prime}{(2\,\pi)^3}\, \tilde G^0(\mbox{\boldmath $k$},\mbox{\boldmath $p$}^\prime)\, \bigg(\,1+\frac{{\mbox{\boldmath $p$}^\prime}^2}{6 M_t^2}\,\bigg) \,, \\[3mm] \label{Pvertex} P(\mbox{\boldmath $k$}) & = & \int\frac{d^3 \mbox{\boldmath $p$}^\prime}{(2\,\pi)^3}\, \frac{\mbox{\boldmath $k$}\mbox{\boldmath $p$}^\prime} {\mbox{\boldmath $k$}^2}\, \tilde G^1(\mbox{\boldmath $k$},\mbox{\boldmath $p$}^\prime) \,, \end{eqnarray} we finally arrive at the integral equations for $S(\mbox{\boldmath $k$})$ and $P(\mbox{\boldmath $k$})$, \begin{eqnarray} S(\mbox{\boldmath $k$}) & = & G^f(\mbox{\boldmath $k$})\, \bigg(\,1+\frac{{\mbox{\boldmath $k$}^2}}{6 M_t^2}\,\bigg) \,-\, G^f(\mbox{\boldmath $k$})\, \int\frac{d^3 \mbox{\boldmath $p$}^\prime}{(2\pi)^3}\, \tilde V(\mbox{\boldmath $k$},\mbox{\boldmath $p$}^\prime)\, S(\mbox{\boldmath $p$}^\prime) \,, \label{Sintegral} \\[3mm] P(\mbox{\boldmath $k$}) & = & G^f(\mbox{\boldmath $k$}) \,-\, G^f(\mbox{\boldmath $k$})\, \int\frac{d^3 \mbox{\boldmath $p$}^\prime}{(2\pi)^3}\, \frac{\mbox{\boldmath $k$} \mbox{\boldmath $p$}^\prime} {\mbox{\boldmath $k$}^2}\, \tilde V_c^{\tiny \mbox{LO}}(\mbox{\boldmath $k$}, \mbox{\boldmath $p$}^\prime)\, P(\mbox{\boldmath $p$}^\prime) \,, \label{Pintegral} \end{eqnarray} where \begin{eqnarray} G^f(\mbox{\boldmath $k$}) & = & \frac{M_t}{\mbox{\boldmath $k$}^2-p_0^2-i\epsilon}\,\bigg[\, 1 + \frac{\mbox{\boldmath $k$}^2+p_0^2}{4\,M_t^2} \,\bigg] \label{Gfree} \end{eqnarray} is the free vertex function. We note that $(2\pi)^3\,\delta^{(3)}(\mbox{\boldmath $k$}- \mbox{\boldmath $k$}^\prime)\,G^f(\mbox{\boldmath $k$})$ is the Green function of Eq.~(\ref{NNLOSchroedinger}) for $\tilde V=0$. The vertex functions $S(\mbox{\boldmath $k$})$ and $P(\mbox{\boldmath $k$})$ only depend on the spatial momentum $\mbox{\boldmath $k$}$. As mentioned at the beginning of this section, their dependence on the time component $k^0$ has been eliminated by a trivial integration by residues. For a proper integration over the phase space of the top-antitop decay products, which is carried out in Sec.~\ref{sectionwidth}, we also need the full dependence of the vertex functions on $k^0$. It can be recovered by comparing $G^f(\mbox{\boldmath $k$})$ to the product of a free PNRQCD top and antitop propagator at NNLO carrying momenta $k_{t}=(k^0+(\frac{p_0^2}{2M_t}-\frac{p_0^4}{8 M_t^3}), \mbox{\boldmath $k$})$ and $k_{\bar t}=(k^0-(\frac{p_0^2}{2M_t}-\frac{p_0^4}{8 M_t^3}), \mbox{\boldmath $k$})$, respectively\footnote{ This choice corresponds to a situation in top-antitop ladder diagrams where half of the centre-of-mass energy is flowing through the top and half through the antitop line, see Fig.~\ref{figroutingladder}. }: \begin{eqnarray} G^f(k^0,\mbox{\boldmath $k$}) & \equiv & \frac{i}{k^0+(\frac{p_0^2}{2M_t}-\frac{p_0^4}{8 M_t^3}) - (\frac{\mbox{\boldmath $k$}^2}{2M_t}- \frac{\mbox{\boldmath $k$}^4}{8 M_t^3}) + i\epsilon }\, \frac{i}{k^0-(\frac{p_0^2}{2M_t}-\frac{p_0^4}{8 M_t^3}) + (\frac{\mbox{\boldmath $k$}^2}{2M_t}- \frac{\mbox{\boldmath $k$}^4}{8 M_t^3}) - i\epsilon } \nonumber \\[2mm] & = & \frac{-1}{{k^0}^2- \Big(\frac{p_0^2}{2M_t}-\frac{\mbox{\boldmath $k$}^2}{2M_t} +i\epsilon\Big)^2} + \frac{ \Big(\frac{p_0^2}{2M_t}-\frac{\mbox{\boldmath $k$}^2}{2M_t}\Big)^2 \Big(\frac{p_0^2}{2M_t^2}+\frac{\mbox{\boldmath $k$}^2}{2M_t^2}\Big) }{\Big[{k^0}^2- \Big(\frac{p_0^2}{2M_t}-\frac{\mbox{\boldmath $k$}^2}{2M_t} +i\epsilon\Big)^2\Big]^2} \,. \label{Gfreecov} \end{eqnarray} We emphasize that an expansion of the NNLO relativistic effects is understood. The relation between $G^f(\mbox{\boldmath $k$})$ and $G^f(k^0,\mbox{\boldmath $k$})$ reads \begin{eqnarray} G^f(\mbox{\boldmath $k$}) & = & -i\,\int\limits_{-\infty}^{+\infty}\frac{d k^0}{2\pi}\, G^f(k^0,\mbox{\boldmath $k$}) \,. \label{GthreetoGfourrelation} \end{eqnarray} Recalling that the potentials as well as the production and annihilation vertex corrections do not depend on the zero-components of the momenta, the integral equations for the generalized vertex functions read \begin{eqnarray} S(k^0,\mbox{\boldmath $k$}) & = & G^f(k^0,\mbox{\boldmath $k$})\, \bigg(\,1+\frac{{\mbox{\boldmath $k$}^2}}{6 M_t^2}\,\bigg) \,+\,i\, G^f(k^0,\mbox{\boldmath $k$})\, \int\frac{d^3 \mbox{\boldmath $p$}^\prime}{(2\pi)^3}\, \int\limits_{-\infty}^{+\infty}\frac{d {p^\prime}^0}{2\pi}\, \tilde V(\mbox{\boldmath $k$},\mbox{\boldmath $p$}^\prime)\, S({p^\prime}^0,\mbox{\boldmath $p$}^\prime) \,, \label{Sintegralcov} \\[3mm] P(k^0,\mbox{\boldmath $k$}) & = & G^f(k^0,\mbox{\boldmath $k$}) \,+\,i\, G^f(k^0,\mbox{\boldmath $k$})\, \int\frac{d^3 \mbox{\boldmath $p$}^\prime}{(2\pi)^3}\, \int\limits_{-\infty}^{+\infty}\frac{d {p^\prime}^0}{2\pi}\, \frac{\mbox{\boldmath $k$} \mbox{\boldmath $p$}^\prime}{k^2}\, \tilde V_c^{\tiny \mbox{LO}}(\mbox{\boldmath $k$}, \mbox{\boldmath $p$}^\prime)\, P({p^\prime}^0,\mbox{\boldmath $p$}^\prime) \,. \label{Pintegralcov} \end{eqnarray} The three- and four-dimensional versions of the vertex functions $S$ and $P$ are related by an equation similar to Eq.~(\ref{GthreetoGfourrelation}). We note that, by construction, the relation \begin{eqnarray} \frac{X(k^0,\mbox{\boldmath $k$})}{G^f(k^0,\mbox{\boldmath $k$})} & = & \frac{X(\mbox{\boldmath $k$})}{G^f(\mbox{\boldmath $k$})} \,, \qquad (X\, = \, S, P) \,, \label{amputatedrelation} \end{eqnarray} holds for the amputated vertex functions. It is straightforward to formulate the optical theorem, which relates the imaginary part of the correlators ${\cal{A}}^v$ and ${\cal{A}}^a$ to explicit phase space integrals over the modulus squared of the vertex functions $S$ and $P$. The relations read \begin{eqnarray} \mbox{Im}\Big[\,{\cal{A}}^v(q^2)\,\Big] & = & 12\,\pi^2\,N_c\, \int\frac{d^3 \mbox{\boldmath $k$}}{(2\pi)^3}\, \int\limits_{-\infty}^{+\infty}\frac{d k^0}{2\pi}\, \bigg|\, \frac{S(k^0,\mbox{\boldmath $k$})}{G^f(k^0,\mbox{\boldmath $k$})} \,\bigg|^2\, \delta\bigg(\frac{\mbox{\boldmath $k$}^2}{2 M_t}- \frac{\mbox{\boldmath $k$}^4}{8 M_t^3}- \Big(\frac{p_0^2}{2 M_t}-\frac{p_0^4}{8 M_t^3}\Big) + k^0 \bigg) \nonumber\\ & & \hspace{2cm} \times\, \delta\bigg(\frac{\mbox{\boldmath $k$}^2}{2 M_t}- \frac{\mbox{\boldmath $k$}^4}{8 M_t^3}- \Big(\frac{p_0^2}{2 M_t}-\frac{p_0^4}{8 M_t^3}\Big) - k^0 \bigg) \nonumber\\[2mm] & = & 6\,\pi\,N_c\, \int\frac{d^3 \mbox{\boldmath $k$}}{(2\pi)^3}\, \bigg|\, \frac{S(\mbox{\boldmath $k$})}{G^f(\mbox{\boldmath $k$})} \,\bigg|^2\, \delta\bigg(\frac{\mbox{\boldmath $k$}^2}{M_t}- \frac{\mbox{\boldmath $k$}^4}{4 M_t^3}- \Big(\frac{p_0^2}{M_t}-\frac{p_0^4}{4 M_t^3}\Big) \bigg) \,, \label{Sopticalstable} \\[4mm] \mbox{Im}\Big[\,{\cal{A}}^a(q^2)\,\Big] & = & 8\,\pi^2\,N_c\, \int\frac{d^3 \mbox{\boldmath $k$}}{(2\pi)^3}\, \int\limits_{-\infty}^{+\infty}\frac{d k^0}{2\pi}\, \frac{\mbox{\boldmath $k$}^2}{M_t^2}\, \bigg|\, \frac{P(k^0,\mbox{\boldmath $k$})}{G^f(k^0,\mbox{\boldmath $k$})} \,\bigg|^2\, \delta\bigg(\frac{\mbox{\boldmath $k$}^2}{2 M_t} - \frac{p_0^2}{2 M_t} + k^0 \bigg) \nonumber\\ & & \hspace{2cm} \times\, \delta\bigg(\frac{\mbox{\boldmath $k$}^2}{2 M_t} - \frac{p_0^2}{2 M_t} - k^0 \bigg) \nonumber\\[2mm] & = & 4\,\pi\,N_c\, \int\frac{d^3 \mbox{\boldmath $k$}}{(2\pi)^3}\, \frac{\mbox{\boldmath $k$}^2}{M_t^2}\, \bigg|\, \frac{P(\mbox{\boldmath $k$})}{G^f(\mbox{\boldmath $k$})} \,\bigg|^2\, \delta\bigg(\frac{\mbox{\boldmath $k$}^2}{M_t} - \frac{p_0^2}{M_t} \bigg) \,. \label{Popticalstable} \end{eqnarray} \vspace{1.5cm} \section{Regularization Scheme and Short-Distance Coefficients} \label{sectionregularization} All equations derived previously have to be considered in the framework of a proper UV regularization scheme. In fact, UV linear and logarithmic divergences arise in Eq.~(\ref{Sintegral}) from the NNLO non-Coulombic potentials and from the kinetic energy and vertex corrections. However, we emphasize that even in the case when no UV divergences arise, all integrals have to be consistently regularized, because only in a consistent regularization scheme can the short-distance coefficients be defined properly. In principle, the regularization scheme ``of choice'' would be an analytic scheme like $\overline{\mbox{MS}}$ as it is usually used in modern perturbative QCD calculations. The preference for the $\overline{\mbox{MS}}$ scheme arises from the fact that it naturally preserves gauge invariance, Ward identities and, particularly important in the framework of effective field theories, power counting rules. Unfortunately an analytic solution of Eqs.~(\ref{NNLOSchroedinger}), (\ref{Sintegral}) and (\ref{Pintegral}) is not even known for four dimensions. Thus, the only sensible way to use the $\overline{\mbox{MS}}$ scheme is to start from the known Coulomb solution of the non-relativistic Schr\"odinger equation and include NLO and NNLO corrections via time-independent perturbation theory, and then explicitly construct the spectral representation of the Green function at the NNLO level. While this program might still be feasible for the determination of the total cross section, it is rather cumbersome for the calculations of distributions at NNLO. In this work we use a momentum cutoff regularization scheme by simply excluding momenta that have a spatial component larger than the cutoff $\Lambda$. This is in fact the most natural regularization scheme for a numerical solution of Eqs.~(\ref{Sintegral}) and (\ref{Pintegral}). However, a cutoff scheme contains a number of subtleties, which shall be briefly discussed in the following. As indicated before, a cutoff scheme leads to violations of gauge invariance and Ward identities in the (P)NRQCD calculation. These effects, however, are generated at the cutoff and are therefore cancelled by corresponding terms with a different sign in the short-distance coefficients. Thus the cross section, which contains the proper combination of non-relativistic correlators and short-distance coefficients, is gauge-invariant and satisfies all Ward identities up to terms beyond the order at which the matching calculation has been carried out. Another subtlety of a cutoff prescription is that it is only well defined if a specific routing convention for the momenta in loops is adopted. For the calculations of the top-antitop cross section at NNLO, it is straightforward and easy to find such a routing convention, because only ladder-type diagrams are involved. It is natural to choose the routing used in the integral equations~(\ref{Sintegral}) and (\ref{Pintegral}), which we have, for clarity, depicted graphically in Fig~\ref{figroutingladder}. \begin{figure}[t!] \begin{center} \leavevmode \epsfxsize=6cm \leavevmode \epsffile[240 420 440 460]{routing.ps} \vskip 2.8cm \caption{\label{figroutingladder} Routing convention for loop momenta in ladder diagrams. } \end{center} \end{figure} It is important to use exactly the same routing for the (P)NRQCD diagrams calculated in the matching procedure to obtain consistent results for the NNLO short-distance coefficients. Finally, it has to be mentioned that a cutoff scheme inevitably leads to power counting breaking effects. This means in our case that a term in the Schr\"odinger equation~(\ref{NNLOSchroedinger}), which is NNLO according to the velocity counting, can in principle lead to lower order contributions in the non-relativistic current correlator. Like the terms that violate gauge invariance and Ward identities, the power counting breaking terms are also generated at the cutoff; they are therefore cancelled in the combination of the correlators and the corresponding short-distance coefficients. However, all this happens only if the cutoff is chosen of the order of $M_t$. To illustrate this issue let us first consider the LO two-loop NRQCD vector current correlator, which contains the exchange of a Coulomb gluon as shown in Fig.~\ref{figcorrelatorCoulomb}a. \begin{figure}[t!] \begin{center} \leavevmode \epsfxsize=2.5cm \leavevmode \epsffile[240 420 440 460]{corrCoul.ps} \hspace{4cm} \leavevmode \epsfxsize=2.5cm \leavevmode \epsffile[240 420 440 460]{corrCoulb.ps} \vskip 2.2cm \caption{\label{figcorrelatorCoulomb} The two-loop NRQCD vector current correlator with the exchange of one Coulomb gluon without relativistic corrections (a) and with the kinetic energy corrections (b). } \end{center} \end{figure} It is straightforward to calculate the absorptive part of the diagram as an expansion in $p_0$ (see App.~\ref{appendixshortdistance}): \begin{eqnarray} I_1^{(1)} & = & \frac{C_F\,\alpha_s\,M_t^2}{4\, \pi^2}\,\bigg[\, \frac{\pi^2}{2}-\frac{4\,p_0}{\Lambda} + {\cal{O}}(p_0^3) \,\bigg] \,. \label{2loopCoulomb} \end{eqnarray} The first term in the brackets on the RHS of Eq.~(\ref{2loopCoulomb}) is the well known Coulomb singularity, which leads to a finite cross section at order $\alpha_s$. The second term is cutoff-dependent and leads to a short-distance correction $\propto \frac{\alpha_s}{\pi}\frac{M_t}{\Lambda}$. (In App.~\ref{appendixshortdistance} the reader can convince her/himself that the scale of $\alpha_s$ in this term is indeed of order $M_t$.) Obviously, to avoid a breakdown in the separation of long- and short-distance contributions\footnote{ In this context ``long-distance'' effects are not understood as ``non-perturbative'' effects, which come from scales of order $\Lambda_{QCD}$ , but rather as effects governed by scales of order $M_t v$ or $M_t v^2$. }, $\Lambda$ has to be chosen of order $M_t$. (Choosing $\Lambda$ of order $M_t v$ would be absurd anyway, because this would cut off a large part of the dynamics. In this respect the cutoff $\Lambda$ acts in a way completely different from the ``cutoff'' scale in the $\overline{\mbox{MS}}$ scheme, which would naturally be chosen of order $M_t v$.) Let us now consider the kinetic energy corrections in the same two-loop NRQCD diagram, as shown in Fig.~\ref{figcorrelatorCoulomb}b. Using the result obtained in App.~\ref{appendixshortdistance} the expression for the diagram including combinatorial factors reads \begin{eqnarray} I_3^{(1)} & = & \frac{C_F\,\alpha_s\,M_t^2}{4\, \pi^2}\,\bigg[\, \frac{\Lambda\,p_0}{M_t^2} + \frac{p_0^2\,\pi^2}{2\,M_t^2} + {\cal{O}}( p_0^3 ) \,\bigg] \,. \label{2loopCoulombkinetic} \end{eqnarray} As expected from power counting arguments, the kinetic energy correction leads to a contribution suppressed by $p_0^2/M_t^2$ with respect to the pure Coulomb exchange diagram in Eq.~(\ref{2loopCoulomb}). However, there is also a term proportional to $p_0/M_t$ because $\Lambda$ is of order $M_t$. This is an example for a power counting breaking term. In the short-distance coefficient $C^v$ this term leads to a NNLO contribution $\propto\alpha_s\frac{\Lambda}{M_t}$. Because this term only arises if NNLO relativistic effects are taken into account, we have to count it as NNLO. Similar terms are caused by the Breit-Fermi potential $V_{\mbox{\tiny BF}}$, the non-Abelian potential $V_{\mbox{\tiny NA}}$ and the dimension-5 NRQCD vector current. We emphasize again that, as is the case for the terms that violate gauge invariance and Ward identities, all power counting breaking terms are automatically cancelled in the combination with the short-distance coefficients up to terms beyond the order at which one carries out the matching procedure. In our case, where the matching is carried out at order $\alpha_s^2$, power counting breaking terms of order $\alpha_s^3$ remain uncancelled, but they are beyond NNLO accuracy. Taking into account the issues discussed above, it is straightforward to determine the matching coefficient $C^v$. Details of this calculation are given in App.~\ref{appendixshortdistance}. The result reads ($a_s\equiv\alpha_s(\mu)$) \begin{eqnarray} C^v & = & 1 + \bigg\{\, \frac{4\,C_F\,a_s}{\pi}\, \bigg[\, -1 + \frac{M_t}{\Lambda} \,\bigg] \,\bigg\}^{\mbox{\tiny NLO}} \nonumber \\[3mm] & & + \,\bigg\{\, \frac{4\,C_F\,a_s\,\Lambda}{3\,\pi\,M_t} + \frac{C_F^2\,a_s^2}{\pi^2}\,\bigg[\, \frac{\beta_0}{C_F}\,\bigg(\, - \frac{\Lambda^2 + 12\,M_t^2}{6\,\Lambda\,M_t} + \frac{\Lambda^2 - 12\,M_t^2} {6\,\Lambda\,M_t}\,\ln\Big(\frac{\Lambda}{\mu}\Big) + 2\,\ln\Big(\frac{M_t}{\mu}\Big) \,\bigg) \nonumber \\[2mm] & & - \frac{a_1}{C_F}\,\frac{\Lambda^2 - 12\,M_t^2}{12\,\Lambda\,M_t} + \pi^2\,\bigg(\,\frac{2}{3}+\frac{C_A}{C_F}\,\bigg)\, \ln\Big(\frac{2\,M_t}{\Lambda}\,\Big) + \frac{\pi^2\,\kappa}{C_F^2} \nonumber \\[2mm] & & + \frac{16\,\Lambda^2}{9\,M_t^2} - \frac{16\,\Lambda}{3\,M_t} - \frac{16\,M_t}{\Lambda} + \frac{M_t^2\,(20 + \pi^2)}{2\,\Lambda^2} - \frac{53\,\pi^2}{24} - \frac{C_A\,\pi^2}{C_F} + \frac{25}{6} + \frac{7}{3}\zeta_3 \,\bigg] \,\bigg\}^{\mbox{\tiny NNLO}} \,. \label{CVshortdistance} \end{eqnarray} As explained in App.~\ref{appendixshortdistance} we have displayed the NLO and NNLO contributions to $C^v$ separately. Appendix~\ref{appendixshortdistance} also contains a discussion on the convergence of the perturbative series in $C^v$ in our cutoff scheme compared to the $\overline{\mbox{MS}}$ scheme~\cite{Melnikov4,Beneke2}. \vspace{1.5cm} \section{Top Quark Width} \label{sectionwidth} Up to now we have treated the top quark as a stable particle. For top quark production close to threshold this is not appropriate because the kinetic energy of the top quark $\sim M_t v^2\sim M_t \alpha_s^2\sim$~2--3~GeV is of the same order as its decay width. It has been shown by Fadin and Khoze~\cite{Fadin1} that in the non-relativistic limit the top width can be consistently implemented by calculating the cross section for stable top quarks supplemented by the replacement $E\to E+i \Gamma_t$, where $E$ is the centre-of-mass energy measured with respect to the two-particle threshold. For the total cross section, this prescription remains valid even at NLO because, in this case, order $\alpha_s$ QCD radiative corrections in form of a gluon that connects top quark production and decay vanish~\cite{Melnikov1}. A consistent implementation of the top width at NNLO has been missing so far. In the framework of NRQCD, and if one is not interested in any differential information on the top decay products, the top width can be understood as a modification of the NRQCD matching conditions caused by electroweak corrections. Because the particles involved in these corrections can be lighter than the top quark (i.e. if the top quark can decay weakly) they can lead to non-zero imaginary parts in the matching conditions and, likewise, in the short-distance coefficients of the NRQCD Lagrangian and the NRQCD currents. This is a well known concept in quantum mechanics of inelastic processes where particle decay and absorption processes are represented by potentials and couplings carrying complex coefficients, if one is not interested in the details of the decay and absorption process. In this context the effects of the top quark decay are only a small (but nevertheless the most important) part of a whole array of electroweak corrections relevant to top-antitop quark pair production close to threshold. In this work we only consider the effects from the on-shell top decay into a $W$ boson and a bottom quark, assuming that the $W$ and the $b$ are themselves stable. Thus the results presented here are, by definition, gauge-invariant. The consistent treatment of all electroweak effects, including a proper handling of the off-shell decays of the top quark, interconnection effects and of gauge-invariance-violating contributions, is beyond the scope of this paper. Such a treatment is carried out in~\cite{BenekeHoang}. Before we determine the modifications of the NRQCD Lagrangian, Eq.~(\ref{NRQCDLagrangian}), through the top decay, some remarks about the velocity counting of the top quark width are in order. Comparing the numerical size of the top width \begin{equation} \Gamma_t \, \approx \, \frac{G_F}{\sqrt{2}}\frac{M_t^3}{8 \pi} \sim \, 1.5~\mbox{GeV} \end{equation} with the binding energy of a fictitious toponium $1S$ bound state in the pole mass scheme ($\alpha_s\sim 0.15$) \begin{equation} E_{1S} \approx \frac{M_t (C_F \alpha_s)^2}{4} \, \sim \, 1.75~\mbox{GeV} \,, \end{equation} we find that we have to count $\Gamma_t/M_t$ as order $v^2$, i.e. \begin{equation} \frac{\Gamma_t}{M_t} \, \sim \, \alpha_s^2 \, \sim \, v^2 \,. \label{gammavcounting} \end{equation} Recalling the velocity counting of the operators of the NRQCD Lagrangian, this means that at NNLO and to first order in $\Gamma_t/M_t$ we have to determine the coefficients of the NRQCD operators $\psi^\dagger i \Gamma_t \psi$, $\psi^\dagger i\frac{\Gamma_t}{M_t} D_t \psi$, $\psi^\dagger i \Gamma_t \frac{\mbox{\boldmath $D$}^2}{M_t^2} \psi$ and those operators, where the top quark Pauli spinors are replaced by the antitop ones. As we show later in this section there are also contributions proportional to $i\Gamma_t$ to the photon and Z boson wave function renormalization constants, which have to be included to account for a proper treatment of the top quark decay phase space. We note that the velocity counting~(\ref{gammavcounting}) implies that $g\sim\alpha_s\sim v$, where $g$ is the SU(2) gauge coupling. Thus in a complete calculation of all electroweak effects one also has to determine ${\cal{O}}(g^4)$ electroweak contributions and ${\cal{O}}(\alpha_s^2)$ QCD corrections to the top decay width. In the framework of the Standard Model those corrections are known~\cite{Jezabek2,Czarnecki1,Denner1}. For our purposes, however, it is sufficient to consider the top quark width as an independent parameter, which we treat only to first order and which is not subject to higher-order corrections. The matching coefficients of the bilinear top spinor operators given above can be obtained by sandwiching the absorptive part of the top quark self energy $\Sigma_t$ (in the full electroweak theory) between top quark Dirac spinors and expanding the result around the complex pole position. Keeping only the terms proportional to $\Gamma_t$, the result for the bilinear top Pauli spinor terms reads ($k=(k^0,\mbox{\boldmath $k$})$) \begin{eqnarray} \bar u(\mbox{\boldmath $k$})\,\bigg[\, \mbox{Im} \Sigma_t(k) \,\bigg]\, u(\mbox{\boldmath $k$}) & \Rightarrow & \bar u(\mbox{\boldmath $k$})\,i \frac{\Gamma_t}{2}\, u(\mbox{\boldmath $k$}) \nonumber\\ & = & \tilde \psi^\dagger\,\bigg[\, i\,\frac{\Gamma_t}{2}\,\frac{M_t}{E_k} \,\bigg]\,\tilde\psi \, = \, \tilde \psi^\dagger\,\bigg[\, i\,\frac{\Gamma_t}{2}\,\bigg(\, 1 - \frac{\mbox{\boldmath $k$}^2}{2\,M_t^2} \,\bigg)\,\bigg]\,\tilde\psi \,. \end{eqnarray} The corresponding result for the bilinear antitop Pauli spinor terms reads \begin{eqnarray} \bar v(-\mbox{\boldmath $k$})\,\bigg[\, \mbox{Im} \Sigma_t(k) \,\bigg]\, v(-\mbox{\boldmath $k$}) & \Rightarrow & \bar v(-\mbox{\boldmath $k$})\,i \frac{\Gamma_t}{2}\, v(-\mbox{\boldmath $k$}) \nonumber\\ & = & -\,\tilde \chi^\dagger\,\bigg[\, i\,\frac{\Gamma_t}{2}\,\frac{M_t}{E_k} \,\bigg]\,\tilde\chi \, = \, -\,\tilde \chi^\dagger\,\bigg[\, i\,\frac{\Gamma_t}{2}\,\bigg(\, 1 - \frac{\mbox{\boldmath $k$}^2}{2\,M_t^2} \,\bigg)\,\bigg] \,\tilde\chi \,, \end{eqnarray} where $E_k\equiv(\mbox{\boldmath $k$}^2+M_t^2)^{1/2}$ and \begin{equation} u(\mbox{\boldmath $k$}) \, = \, \sqrt{\frac{E_k+M_t}{2 E_k}}\, \left(\begin{array}{cc} \tilde\psi \\ \frac{\mbox{\boldmath $\sigma$}\mbox{\boldmath $k$}}{E_k+M_t}\, \tilde\psi \end{array} \right)\,, \hspace{1cm} v(-\mbox{\boldmath $k$}) \, = \, \sqrt{\frac{E_k+M_t}{2 E_k}}\, \left(\begin{array}{cc} -\frac{\mbox{\boldmath $\sigma$}\mbox{\boldmath $k$}}{E_k+M_t}\, \tilde\chi\\ \tilde\chi \end{array} \right)\,, \end{equation} using the usual non-relativistic normalization for Dirac spinors. Thus in the presence of top decay we have to modify the NRQCD Lagrangian by adding the terms \begin{eqnarray} \delta {\cal{L}}_{\mbox{\tiny NRQCD}} & = & \psi^\dagger\,i\,\frac{\Gamma_t}{2}\,\bigg[\,1 + \frac{\mbox{\boldmath $D$}^2}{2\,M_t^2} \,\bigg]\,\psi - \chi^\dagger\,i\,\frac{\Gamma_t}{2}\,\bigg[\,1 + \frac{\mbox{\boldmath $D$}^2}{2\,M_t^2} \,\bigg]\,\chi \,. \end{eqnarray} Physically, the dimension-5 operators multiplying the top width correspond to the time dilatation correction. This leads to the following modified versions of the Schr\"odinger equation~(\ref{NNLOSchroedinger}) in momentum space representation \begin{eqnarray} \lefteqn{ \bigg[\, \frac{\mbox{\boldmath $k$}^2}{M_t} - \frac{\mbox{\boldmath $k$}^4}{4M_t^3} \,-\,\bigg(\, \frac{p_0^2}{M_t} - \frac{p_0^4}{4 M_t^3} \,\bigg) \,-\,i\,\Gamma_t\,\bigg(\, 1 - \frac{\mbox{\boldmath $k$}^2}{2\,M_t^2} \,\bigg)\, \,\bigg]\, \tilde G(\mbox{\boldmath $k$},\mbox{\boldmath $k$}^\prime;q^2) } \nonumber \\[3mm] & & + \int\frac{d^3 \mbox{\boldmath $p$}^\prime}{(2\,\pi)^3}\, \tilde V(\mbox{\boldmath $k$},\mbox{\boldmath $p$}^\prime)\, \tilde G(\mbox{\boldmath $p$}^\prime,\mbox{\boldmath $k$}^\prime;q^2) \, = \, (2\,\pi)^3\,\delta^{(3)}(\mbox{\boldmath $k$}-\mbox{\boldmath $k$}^\prime) \,, \label{NNLOSchroedingergamma} \end{eqnarray} and the free vertex functions $G^f(\mbox{\boldmath $k$})$ and $G^f(k^0,\mbox{\boldmath $k$})$ in the integral equations~(\ref{Sintegral}), (\ref{Pintegral}), (\ref{Sintegralcov}) and (\ref{Pintegralcov}): \begin{eqnarray} G^f(\mbox{\boldmath $k$}) & = & \frac{M_t}{\mbox{\boldmath $k$}^2-p_0^2-\frac{\Gamma_t^2}{4}- i\,M_t\,\Gamma_t}\,\bigg[\, 1 + \frac{\mbox{\boldmath $k$}^2+p_0^2}{4\,M_t^2}- i\,\frac{\Gamma_t}{4\,M_t} \,\bigg] \,, \label{Gfreegamma} \\ G^f(k^0,\mbox{\boldmath $k$}) & = & \frac{-1}{{k^0}^2- \Big(\frac{1}{2M_t}(p_0^2+\frac{\Gamma_t^2}{4})- \frac{\mbox{\boldmath $k$}^2}{2M_t} +i\frac{\Gamma_t}{2}\Big)^2} + \frac{ \Big(\frac{p_0^2}{2M_t}-\frac{\mbox{\boldmath $k$}^2}{2M_t} +i\frac{\Gamma_t}{2}\Big)^2 \Big(\frac{p_0^2}{2M_t^2}+\frac{\mbox{\boldmath $k$}^2}{2M_t^2} -i\frac{\Gamma_t}{2M_t}\Big) }{\Big[{k^0}^2- \Big(\frac{1}{2M_t}(p_0^2+\frac{\Gamma_t^2}{4})- \frac{\mbox{\boldmath $k$}^2}{2M_t} +i\frac{\Gamma_t}{2}\Big)^2 \Big]^2} \,. \nonumber\\ \label{Gfreecovgamma} \end{eqnarray} We note that it is not coercive to keep the term $-\frac{\Gamma_t^2}{4}$ in the LO propagator, because it is of NNLO according to the power counting. We have adopted this convention because this choice leads to a simplification of the analytic form for the rest of the NNLO corrections in Eqs.~(\ref{Gfreegamma}) and (\ref{Gfreecovgamma}). From the modified version of the Schr\"odinger equation~(\ref{NNLOSchroedingergamma}) we can immediately see that the Fadin-Khoze replacement rule is valid at LO and NLO in the non-relativistic expansion, but it is inappropriate at NNLO. In fact it is not possible at all to consistently implement the top quark width at NNLO simply by shifting the centre-of-mass energy in a calculation for stable quarks. The resulting form of the optical theorem relations reads \begin{eqnarray} \mbox{Im}\Big[\,{\cal{A}}^v(q^2)\,\Big] & = & 3\,N_c\, \int\frac{d^3 \mbox{\boldmath $k$}}{(2\pi)^3}\, \int\limits_{-\infty}^{+\infty}\frac{d k^0}{2\pi}\, |S(k^0,\mbox{\boldmath $k$})|^2\, \Gamma_t^2\,\bigg(1 - \frac{\mbox{\boldmath $k$}^2}{M_t^2} \,\bigg) \nonumber\\[2mm] & = & 6\,N_c\, \int\frac{d^3 \mbox{\boldmath $k$}}{(2\pi)^3}\, |S(\mbox{\boldmath $k$})|^2\, \Gamma_t\,\bigg(1 - \frac{\mbox{\boldmath $k$}^2}{2\,M_t^2} \,\bigg) \,, \label{Sopticalunstable} \\[4mm] \mbox{Im}\Big[\,{\cal{A}}^a(q^2)\,\Big] & = & 2\,N_c\, \int\frac{d^3 \mbox{\boldmath $k$}}{(2\pi)^3}\, \int\limits_{-\infty}^{+\infty}\frac{d k^0}{2\pi}\, \frac{\mbox{\boldmath $k$}^2}{M_t^2}\, |P(k^0,\mbox{\boldmath $k$})|^2\,\Gamma_t^2 \nonumber\\[2mm] & = & 4\,N_c\, \int\frac{d^3 \mbox{\boldmath $k$}}{(2\pi)^3}\, \frac{\mbox{\boldmath $k$}^2}{M_t^2}\, |P(\mbox{\boldmath $k$})|^2\,\Gamma_t \,. \label{Popticalunstable} \end{eqnarray} The first equality in Eqs.~(\ref{Sopticalunstable}) and (\ref{Popticalunstable}) is an explicit integration over the phase space of the top decay products for off-shell top decay keeping only the term proportional to the top and antitop width for the top decay sub phase space. We emphasize again that in Eqs.~(\ref{Sopticalunstable}) and (\ref{Popticalunstable}) no physical phase space boundaries for the integration over the top-antitop four momentum are implemented. All integrations are defined in the framework of the regularization scheme. The reader might ask whether the inclusion of the top quark width into the top quark propagators leads to non-trivial modifications of the NRQCD short-distance coefficient $C^v$.\footnote{ One could equally well ask whether there is a need to introduce the operators $\frac{\Gamma_t}{M_t} \psi^\dagger \mbox{\boldmath $\sigma$}\chi$ and $\frac{\Gamma_t}{M_t} \chi^\dagger \mbox{\boldmath $\sigma$}\psi$ in the non-relativistic expansion of the currents, which produce and annihilate the top-antitop pair. At this point we prefer the language used in the text. } Because $\frac{\Gamma_t}{M_t}\sim v^2\sim\alpha_s^2$, it is not inconceivable that there could be terms $\propto\frac{\Gamma_t}{M_t}$ in $C^v$ at NNLO. To show that this is not the case we recall that $C^v$ is the modulus squared of the short-distance coefficient $c_1^v$ of the non-relativistic current $\psi^\dagger\mbox{\boldmath $\sigma$}\chi$ in Eq.~(\ref{vectorcurrentexpansion1}). As mentioned in App.~\ref{appendixshortdistance}, the short-distance coefficient $c_1^v$ is the sum of amputated vertex diagrams in the full theory where the three momenta in the loop integrations are larger than the cutoff $\Lambda\sim M_t$ for $\sqrt{s}=2 M_t$. Thus, there is no integration over the top-antitop pole located at three momenta $\sim p_0\sim M_t v$, and we can conclude that the short-distance coefficient does not contain non-analytic terms involving the top quark width. Therefore $c_1^v$ can be expanded in $i\Gamma_t$. For the same reason, the coefficients of an expansion in $i\Gamma_t$ do not contain a top-antitop cut and are real numbers. Because $C^v$ is the modulus squared of $c_1^v$, the first non-vanishing term of an expansion of $C^v$ is proportional to $(\Gamma_t/M_t)^2\sim\alpha_s^4$, which is indeed beyond NNLO. However, we emphasize that, in a complete calculation of all electroweak effects, $C^v$ will receive corrections $\propto g^2, {g^\prime}^2\sim G_F M_W^2$, $g^\prime$ being the U(1) gauge coupling, which are formally of order $\Gamma_t/M_t$. These corrections do not come from the width contained in the top and antitop propagators but for instance from electroweak corrections to the vertices $\gamma, Z\to t\bar t$~\cite{Guth1}. These corrections exist even in the case when the top quark is treated as a stable particle. If the top quark width is included in the NRQCD framework there is only one additional source of a linear dependence on $\Gamma_t$ which comes from the phase space integration in Eqs.~(\ref{Sopticalunstable}) and (\ref{Popticalunstable}). In contrast to the case of a stable top quark, where the phase space is restricted to those top quark four momenta allowed by the centre-of-mass energy, the physical boundaries of the phase space integration $\int\frac{d^4 k}{(2\pi)^4}$ in the case of unstable top quarks are determined by the allowed invariant masses of the top quark decay products. Assuming that the bottom quark is massless and that bottom quark and $W$ boson are stable, and taking into account our routing convention (see Fig.~\ref{figroutingladder}), the boundaries of the physical four-dimensional phase space integrations read ($k^2\equiv|\mbox{\boldmath $k$}|^2$) \begin{equation} \int\limits_{\tiny\mbox{phase space}}\frac{d^4 k}{(2\pi)^4} \, = \, \int\limits_0^{\sqrt{\frac{q^2}{4}-M_W^2}}\, \frac{d^3 \mbox{\boldmath $k$}}{(2\pi)^3} \int\limits_{-(\frac{\sqrt{q^2}}{2}-\sqrt{k^2+M_W^2})}^{ (\frac{\sqrt{q^2}}{2}-\sqrt{k^2+M_W^2})}\,\frac{d k^0}{2\pi} \,. \label{phasespacephysical} \end{equation} It is obvious that the physical limits of integration are not equivalent to the actual limits of integration on the RHS of Eqs.~(\ref{Sopticalunstable}) and (\ref{Popticalunstable}) as defined through our cutoff regularization scheme: \begin{equation} \int\limits_{\tiny\mbox{cutoff scheme}}\frac{d^4 k}{(2\pi)^4} \, = \, \int\limits_0^{|\mbox{\boldmath $k$}|<\Lambda}\, \frac{d^3 \mbox{\boldmath $k$}}{(2\pi)^3}\,\, \int\limits_{-\infty}^{+\infty}\,\frac{d k^0}{2\pi} \,. \label{phasespaceactual} \end{equation} For the total cross section on the LHS of Eqs.~(\ref{Sopticalunstable}) and (\ref{Popticalunstable}), the difference between using the phase space integrations~(\ref{phasespaceactual}) or (\ref{phasespacephysical}) can be expanded in $p_0^2$ and $\Gamma_t$, and can be accounted for by introducing additional photon and Z boson wave function renormalization constants into the NRQCD Lagrangian, which are proportional to $i\Gamma_t$. The calculation of these counter-terms is straightforward. This leads to the following form of the optical theorem relations for the vector- and axial-vector-current-induced correlators, \begin{eqnarray} \mbox{Im}\,\bigg[\, {\cal{A}}^v + i \frac{3\,N_c\,M_t\,\Gamma_t}{2\,\pi^2}\,\Big[\, 2\frac{M_t}{\Lambda}-(1+\sqrt{3}) \,\Big] \,\bigg] & = & 3\,N_c\, \int\limits_{\tiny\mbox{phase space}}\frac{d^4 k}{(2\pi)^4}\, |S(k^0,\mbox{\boldmath $k$})|^2\, \Gamma_t^2\,\bigg(1 - \frac{\mbox{\boldmath $k$}^2}{M_t^2} \,\bigg) \,, \label{Sopticalunstableproper} \\[5mm] \mbox{Im}\,\bigg[\, {\cal{A}}^a + i \frac{N_c\,M_t\,\Gamma_t}{\pi^2}\,\Big[\, -2\frac{\Lambda}{M_t}+2\,(\sqrt{3}-1) \,\Big] \,\bigg] & = & 2\,N_c\, \int\limits_{\tiny\mbox{phase space}}\frac{d^4 k}{(2\pi)^4}\, \frac{\mbox{\boldmath $k$}^2}{M_t^2}\, |P(k^0,\mbox{\boldmath $k$})|^2\,\Gamma_t^2 \,, \label{Popticalunstableproper} \end{eqnarray} where we have displayed the counter-terms for the phase space, each at the Born level, to first order in $\Gamma_t$ and first order in the non-relativistic expansion. For simplicity we have set the bottom quark and the W boson masses to zero. As far as the velocity counting of the width is concerned, we in principle should have included also the ${\cal{O}}(\alpha_s)$ contributions for the counter-term in Eq.~(\ref{Sopticalunstableproper}), because we formally have to count the vector-current-induced total cross section as order $v$. However, the counter-term contributions are only at the per cent level and not (yet) important phenomenologically, compared with the much larger uncertainties in the normalization of the total cross section from QCD, which are discussed in Sec.~\ref{subsectionnormalization}. We also note that the phase space counter-terms can be calculated entirely within the non-relativistic effective theory, and that there is no need to match to the full theory. This is because within the physical phase space boundaries the non-relativistic expansion of the phase space integration should be convergent.\footnote{ This statement is a conjecture. In this context ``convergence'' for the phase space counter-terms is not meant to be associated with an expansion in $\Gamma_t$ or $p_0$, but to the convergence in the coefficient multiplying the term linear in $\Gamma_t$. All higher-order terms in the non-relativistic expansion under the phase space integral can contribute to the term linear in $\Gamma_t$. } We also note that the calculation of the phase space counter-term for the S-P-wave interference, contributing e.g. in the top quark angular distribution, is in complete analogy to the calculation of the counter-terms in Eqs.~(\ref{Sopticalunstableproper}) and (\ref{Popticalunstableproper}). We emphasize that the non-relativistic current correlators ${\cal{A}}^v$, ${\cal{A}}^a$ and the vertex functions $S$ and $P$ in relations~(\ref{Sopticalunstableproper}) and (\ref{Popticalunstableproper}) are still calculated in our cutoff scheme presented in Sec.~\ref{sectionregularization}. The expressions for the vector-current- and axial-vector-current-induced total cross sections valid at NNLO in the non-relativistic expansion, and properly including all effects of the top width at the Born level and leading order in the non-relativistic expansion, read \begin{eqnarray} R_{\mbox{\tiny NNLO}}^{v,\mbox{\tiny thr}}(q^2) & = & \frac{4\,\pi}{q^2}\,C^v\, \mbox{Im}\Big[\, {\cal{A}}^v(q^2) \,\Big] + \frac{3\,N_c\,\Gamma_t}{2\,\pi\,M_t}\,\Big[\, 2\frac{M_t}{\Lambda}-(1+\sqrt{3}) \,\Big] \,\bigg] \,, \label{vectorcrosssectionexpandedgamma} \\[3mm] R_{\mbox{\tiny NNLO}}^{a,\mbox{\tiny thr}}(q^2) & = & \frac{4\,\pi}{q^2}\,C^a\, \mbox{Im}\Big[\, {\cal{A}}^a(q^2) \,\Big] + \frac{N_c\,\Gamma_t}{\pi\,M_t}\,\Big[\, -2\frac{\Lambda}{M_t}+2\,(\sqrt{3}-1) \,\Big] \,\bigg] \,. \label{axialvectorcrosssectionexpandedgamma} \end{eqnarray} From Eqs.~(\ref{Sopticalunstableproper}) and (\ref{Popticalunstableproper}) we can derive the centre-of-mass three-momentum distributions of the top quarks ($k\equiv|\mbox{\boldmath $k$}|$), \begin{eqnarray} \frac{d R_{\mbox{\tiny NNLO}}^{v,\mbox{\tiny thr}}(q^2)} {d |\mbox{\boldmath$k$}|} & = & C^v\, \frac{6\,N_c}{\pi\,q^2}\, \Gamma_t^2\,\bigg(1 - \frac{\mbox{\boldmath $k$}^2}{M_t^2} \,\bigg)\, \mbox{\boldmath$k$}^2 \int\limits_{-(\frac{\sqrt{q^2}}{2}-\sqrt{k^2+M_W^2})}^{ (\frac{\sqrt{q^2}}{2}-\sqrt{k^2+M_W^2})}\,\frac{d k^0}{2\pi}\, |S(k^0,\mbox{\boldmath $k$})|^2 \,, \label{Sopticalunstablepropergamma} \\[5mm] \frac{d R_{\mbox{\tiny NNLO}}^{a,\mbox{\tiny thr}}(q^2)} {d |\mbox{\boldmath$k$}|} & = & C^a\, \frac{4\,N_c}{\pi\,q^2}\, \Gamma_t^2\, \frac{\mbox{\boldmath $k$}^4}{M_t^2}\, \int\limits_{-(\frac{\sqrt{q^2}}{2}-\sqrt{k^2+M_W^2})}^{ (\frac{\sqrt{q^2}}{2}-\sqrt{k^2+M_W^2})}\,\frac{d k^0}{2\pi}\, |P(k^0,\mbox{\boldmath $k$})|^2 \,. \label{Popticalunstablepropergamma} \end{eqnarray} Unless $|\mbox{\boldmath $k$}|$ is chosen close to the endpoint $(q^2/4-M_W^2)^{1/2}$ the numerical difference obtained by replacing the physical limits of the $k^0$ integrations by $\pm\infty$ is negligible. We note that the three-momentum distributions shown in Eqs.~(\ref{Sopticalunstablepropergamma}) and (\ref{Popticalunstablepropergamma}) are not equal to the physical observable three-momentum distributions, because the exchange of gluons between the top decay and production processes leads to additional non-negligible corrections at NLO~\cite{Harlander1,Sumino2} and NNLO. These ``interconnection'' effects belong to the electroweak corrections, which are not treated in this work. The NNLO corrections to the three-momentum distribution calculated here are only a first step towards a complete NNLO treatment of the three-momentum distribution. We also want to mention that the three momentum distribution is strictly speaking an ambiguous quantity since the three (as well as the four) momentum of a coloured particle is an ambiguous concept. This is in contrast to the total cross section, which describes the rate of colour singlet top-antitop events. We conclude this section with some remarks on the inconsistencies that can arise if the Fadin-Khoze replacement rule ``$E\to E+i \Gamma_t$'' is employed at NNLO for the calculation of the top-antitop cross section close to threshold. We emphasize that there is nothing wrong, in principle, in calculating the current correlators for stable top quarks via the Schr\"odinger equation~(\ref{NNLOSchroedinger}), supplemented afterwards by the replacement $\sqrt{q^2}-2 M_t\to \sqrt{q^2}-2 M_t + i \Gamma_t$. This corresponds essentially to the modification \begin{equation} \sum_n\hspace{-5mm}\int\, \frac{|n\rangle\langle n|}{E_n-E-i \epsilon} \, \to \, \sum_n\hspace{-5mm}\int\, \frac{|n\rangle\langle n|}{E_n-E-i \Gamma_t} \end{equation} in the spectral representation of the Green function of Eq.~(\ref{NNLOSchroedinger}) and is equivalent to keeping only the terms $\psi^\dagger i\Gamma_t \psi$ and $\chi^\dagger i\Gamma_t \chi$ in the modified version of the NRQCD Lagrangian. In this approach, also the optical theorem remains valid in the form \begin{eqnarray} \mbox{Im}\Big[\,{\cal{A}}^v(q^2)\,\Big] & = & 6\,N_c\, \int\frac{d^3 \mbox{\boldmath $k$}}{(2\pi)^3}\, |S(\mbox{\boldmath $k$})|^2\, \Gamma_t \,, \label{Sopticalunstableconsistent} \end{eqnarray} for the vector current correlator, as an example. (For simplicity, we neglect the subtleties of the phase space effects, because they are irrelevant to this discussion.) However, there is a caveat, since it is possible, for the case of zero width, to simplify the form of Eq.~(\ref{NNLOSchroedinger}) in a way that for the stable top quark case the results remain correct, whereas inconsistencies arise if the results undergo the replacement rule ``$E\to E+i \Gamma_t$''. Such simplifications, based on the assumption that certain singular terms, which arise during the simplification, can be neglected, have in fact been carried out in Refs.~\cite{Hoang2,Hoang3,Melnikov3,Yakovlev1}. In Refs.~\cite{Hoang2,Melnikov3,Hoang5} it was shown that the NNLO kinetic energy corrections and the Breit-Fermi potential in Eq.~(\ref{NNLOSchroedinger}), if they are treated as a perturbation to first order, can be rewritten in terms of an energy dependent Coulomb potential $\propto C_F\alpha_s/\mbox{\boldmath$Q$}^2 \times (p_0^2/M_t^2)$, a Darwin-like constant potential, and a potential $\propto \alpha_s^2/|\mbox{\boldmath$Q$}|$. Neglecting all NNLO corrections except the energy-dependent corrections to the Coulomb potential, the simplified version of the NNLO Schr\"odinger equation has the form\footnote{ We emphasize that the neglect of the rest of the NNLO corrections does not affect the validity of the following arguments, because they are independent of the top quark width after the replacement rule ``$E\to E+i \Gamma_t$'' has been applied. } \begin{eqnarray} \bigg[\, \frac{\mbox{\boldmath $k$}^2}{M_t} - \frac{p_0^2}{M_t} \,\bigg]\, \tilde G(\mbox{\boldmath $k$},\mbox{\boldmath $k$}^\prime) \,+\, \int\frac{d^3 \mbox{\boldmath $p$}^\prime}{(2\,\pi)^3}\, \tilde V^{sim}(\mbox{\boldmath $k$}-\mbox{\boldmath $p$}^\prime)\, \tilde G(\mbox{\boldmath $p$}^\prime,\mbox{\boldmath $k$}^\prime) \, = \, (2\,\pi)^3\,\delta^{(3)}(\mbox{\boldmath $k$}-\mbox{\boldmath $k$}^\prime) \,, \label{NNLOSchroedingersimplified} \end{eqnarray} where \begin{equation} \tilde V^{sim}(\mbox{\boldmath $Q$}) \, = \, -\,\frac{C_F\,4\,\pi\,\alpha_s} {\mbox{\boldmath $Q$}^2}\, \bigg(\, 1+\frac{3\,p_0^2}{2\,M_t^2} \,\bigg) \,. \end{equation} Equation~(\ref{NNLOSchroedingersimplified}) is much easier to solve than the original Schr\"odinger equation~(\ref{NNLOSchroedinger}). For real energies, and if the corrections from the NNLO terms in Eqs.~(\ref{NNLOSchroedinger}) and (\ref{NNLOSchroedingersimplified}) are treated as a perturbation to first order only, the result obtained from Eqs.~(\ref{NNLOSchroedinger}) and (\ref{NNLOSchroedingersimplified}) are indeed equivalent, after a proper renormalization has been carried out. This was in fact the case for which the form of Eq.~(\ref{NNLOSchroedingersimplified}) has been derived in Refs.~\cite{Hoang2,Melnikov3,Hoang5}. However, Eq.~(\ref{NNLOSchroedingersimplified}) leads to inconsistencies for complex energies. This can be seen from the fact that for the total cross section calculated from Eq.~(\ref{NNLOSchroedingersimplified}), after applying the replacement rule $\frac{p_0^2}{M_t}\to \frac{p_0^2}{M_t}+i \Gamma_t$, the actual form of the optical theorem relation reads \begin{eqnarray} \mbox{Im} {\cal{A}}^v & = & 6\,N_c\,\mbox{Im}\,\bigg[\, \int\frac{d^3 \mbox{\boldmath $k$}}{(2\pi)^3} \int\frac{d^3 \mbox{\boldmath $k$}^\prime}{(2\pi)^3}\, \tilde G({\mbox{\boldmath $k$}},{\mbox{\boldmath $k$}^\prime}) \,\bigg] \nonumber \\[3mm] & = & 6\,N_c\, \int\frac{d^3 \mbox{\boldmath $k$}}{(2\pi)^3}\, |S(\mbox{\boldmath $k$})|^2\, \Gamma_t \, - \, 6\,N_c\, \int\frac{d^3 \mbox{\boldmath $k$}}{(2\pi)^3}\, \int\frac{d^3 \mbox{\boldmath $k$}^\prime}{(2\pi)^3}\, S^*(\mbox{\boldmath $k$})\, \mbox{Im}\,\bigg[\, \tilde V^{sim}(\mbox{\boldmath $k$}-\mbox{\boldmath $k$}^\prime) \,\bigg]\, S(\mbox{\boldmath $k$}^\prime) \,, \label{Sopticalunstableinconsistent} \end{eqnarray} rather than Eq.~(\ref{Sopticalunstableconsistent}). The additional term on the RHS of Eq.~(\ref{Sopticalunstableinconsistent}) originates from the energy dependent Coulomb-type potential in Eq.~(\ref{NNLOSchroedingersimplified}). However, the additional term does not correspond to any physical final state, because it corresponds to an absorption process in the potential. In other words, it is impossible to recover the total cross section from the momentum distribution, if one defines it as the integral over the physical final states represented by the first term on the RHS of Eq.~(\ref{Sopticalunstableinconsistent}). Including also the rest of the NNLO corrections not displayed in Eq.~(\ref{NNLOSchroedingersimplified}), we have checked, with the numerical methods described in the next section, that the size of the second term on the RHS of Eq.~(\ref{Sopticalunstableinconsistent}) is between about $5\%$ (for $\sqrt{q^2}-2M_t\sim 5$~GeV) and $20\%$ (for $\sqrt{q^2}-2M_t\sim -5$~GeV) for the choices of parameters employed in the analysis of Sec.~\ref{sectionpolescheme}. (Similar results can already be obtained by analysing the known analytic solutions of the non-relativistic Coulomb problem~\cite{Wichmann1,Hostler1,Schwinger1} for a Coulomb potential with a complex coupling.) Thus for the determination of the total top-antitop cross section close to threshold, there is an unacceptable discrepancy between the integrated momentum distribution over physical final states and the absorptive part of the non-relativistic current correlator. We believe that the size of the second term on the RHS of Eq.~(\ref{Sopticalunstableinconsistent}) should in principle be taken as an estimate for the inherent uncertainties of using the simplified NNLO Schr\"odinger equation~(\ref{NNLOSchroedingersimplified}) supplemented by the replacement rule of Fadin and Khoze. We have checked, however, that the LHS of Eq.~(\ref{Sopticalunstableinconsistent}) is much closer to the correct result, obtained from the original Schr\"odinger equation~(\ref{NNLOSchroedingergamma}), than the integrated momentum distribution. From this point of view the use of the simplified Schr\"odinger equation~(\ref{NNLOSchroedingersimplified}) might be justified for the total cross section, but is questionable for the momentum distribution. \vspace{1.5cm} \section{Numerical Implementation} \label{sectionnumerics} In this work we use numerical methods described in Refs.~\cite{Jezabek1,Teubnerdip,Harlanderdip} to determine the vertex functions $S$ and $P$, which are the building blocks for the calculation of the total top-antitop production cross section and the three momentum distribution. Because the three- and four-dimensional versions of the vertex functions are related through Eq.~(\ref{amputatedrelation}), it is sufficient to determine the amputated vertex functions $S/G^f$ and $P/G^f$ from the three-dimensional integral equations~(\ref{Sintegral}) and (\ref{Pintegral}). The amputated vertex function are spherically symmetric and depend only on the modulus of the three momentum $\mbox{\boldmath $k$}$. It is therefore possible to reduce Eqs.~(\ref{Sintegral}) and (\ref{Pintegral}) to one-dimensional integral equations. Obviously, when solving Eqs.~(\ref{Sintegral}) and (\ref{Pintegral}), the singular behaviour of the potentials $\tilde V(\mbox{\boldmath $k$},\mbox{\boldmath $p$}^\prime)$ and $\tilde V_c^{\tiny\mbox{LO}}(\mbox{\boldmath $k$}, \mbox{\boldmath $p$}^\prime)$ for $\mbox{\boldmath $p$}^\prime \to \mbox{\boldmath $k$}$ requires special treatment. To avoid numerical problems we rewrite the integral equations for the amputated vertex functions as \begin{eqnarray} {\cal K}^v(\mbox{\boldmath $k$}) & \equiv & \frac{S(\mbox{\boldmath $k$})}{G^f(\mbox{\boldmath $k$})} \nonumber\\[2mm] & = & 1+\frac{{\mbox{\boldmath $k$}^2}}{6 M_t^2} \,-\, \int\frac{d^3 \mbox{\boldmath $p$}^\prime}{(2\pi)^3}\, \tilde V(\mbox{\boldmath $k$},\mbox{\boldmath $p$}^\prime)\, G^f(\mbox{\boldmath $p$}^\prime)\, {\cal K}^v(\mbox{\boldmath $p$}^\prime) \nonumber \\[2mm] & = & 1+\frac{{\mbox{\boldmath $k$}^2}}{6 M_t^2} \,-\, \int\frac{d^3 \mbox{\boldmath $p$}^\prime}{(2\pi)^3}\, \tilde V(\mbox{\boldmath $k$},\mbox{\boldmath $p$}^\prime)\, G^f(\mbox{\boldmath $p$}^\prime)\, \Big({\cal K}^v(\mbox{\boldmath $p$}^\prime) - {\cal K}^v(\mbox{\boldmath $k$}) \Big) - {\cal B}^v(\mbox{\boldmath $k$}) \, {\cal K}^v(\mbox{\boldmath $k$}) \,, \label{Samputatednew} \\[4mm] {\cal K}^a(\mbox{\boldmath $k$}) & \equiv & \frac{P(\mbox{\boldmath $k$})}{G^f(\mbox{\boldmath $k$})} \nonumber\\[2mm] & = & 1\,-\, \int\frac{d^3 \mbox{\boldmath $p$}^\prime}{(2\pi)^3}\, \frac{\mbox{\boldmath $k$} \mbox{\boldmath $p$}^\prime}{k^2}\, \tilde V_c^{\tiny \mbox{LO}}(\mbox{\boldmath $k$}, \mbox{\boldmath $p$}^\prime)\, G^f(\mbox{\boldmath $p$}^\prime)\, {\cal K}^a(\mbox{\boldmath $p$}^\prime) \nonumber \\[2mm] & = & 1\,-\, \int\frac{d^3 \mbox{\boldmath $p$}^\prime}{(2\pi)^3}\, \tilde V_c^{\tiny \mbox{LO}}(\mbox{\boldmath $k$}, \mbox{\boldmath $p$}^\prime)\, G^f(\mbox{\boldmath $p$}^\prime)\, \left(\frac{\mbox{\boldmath $k$} \mbox{\boldmath $p$}^\prime}{k^2}\, {\cal K}^a(\mbox{\boldmath $p$}^\prime) - {\cal K}^a(\mbox{\boldmath $k$}) \right) - {\cal B}^a(\mbox{\boldmath $k$}) \, {\cal K}^a(\mbox{\boldmath $k$}) \,, \label{Pamputatednew} \end{eqnarray} where ${\cal B}^v$ and ${\cal B}^a$ are defined as \begin{eqnarray} {\cal B}^v(\mbox{\boldmath $k$}) & \equiv & \int\frac{d^3 \mbox{\boldmath $p$}^\prime}{(2\pi)^3}\, \tilde V(\mbox{\boldmath $k$},\mbox{\boldmath $p$}^\prime)\, G^f(\mbox{\boldmath $p$}^\prime)\,, \label{Bvintegral} \\[2mm] {\cal B}^a(\mbox{\boldmath $k$}) & \equiv & \int\frac{d^3 \mbox{\boldmath $p$}^\prime}{(2\pi)^3}\, \tilde V_c^{\tiny \mbox{LO}}(\mbox{\boldmath $k$}, \mbox{\boldmath $p$}^\prime)\, G^f(\mbox{\boldmath $p$}^\prime)\,. \label{Baintegral} \end{eqnarray} As mentioned above, both ${\cal K}^v$ and ${\cal K}^a$ depend only on the modulus of the three momentum. The angular dependence of the integrand in Eqs.~(\ref{Samputatednew}) and (\ref{Pamputatednew}) (including ${\cal B}^v$ and ${\cal B}^a$) is only coming from the potentials and the dot product $\mbox{\boldmath $k$} \mbox{\boldmath $p$}^\prime$. The angular integration can be carried out analytically. The remaining one-dimensional integral equations are then solved numerically by discretization: the integrals $\int d p^\prime$ ($p^\prime\equiv |\mbox{\boldmath $p$}^\prime|$) are transformed into sums $\sum_i$ over a fixed set of momenta ${p^\prime}^i$, and the integral equations for ${\cal K}^v(\mbox{\boldmath $k$})$ and ${\cal K}^a(\mbox{\boldmath $k$})$ are each reduced to a system of linear equations, where the same set of momenta has to be used for the $k^i$.\footnote{At this point the subtraction carried out in Eqs.~(\ref{Samputatednew}) and (\ref{Pamputatednew}) becomes crucial. Without it the singularities in the potentials for $p^\prime \to k$ would be manifest for ${p^i}^\prime = k^i$, even in the case of integrable singularities.} The resulting (complex) matrices are then inverted numerically to give the amputated vertex functions ${\cal K}^v$ and ${\cal K}^a$ for the momenta $k^i$. In practice it turned out that the use of the Gaussian quadrature formulae is very efficient for the discretization. A surprisingly small number of points (of the order of 100) already leads to a high numerical accuracy. In addition, integrals were split into two (or more) parts and a suitable transformation of integration variables was applied wherever needed. It should also be noted that the finite width of the top quark is essential for the numerical stability of the method. It makes potentially dangerous denominators in the integrands of Eqs.~(\ref{Samputatednew}) and (\ref{Pamputatednew}), which originate from the free Green function, well behaved (compare Eqs.~(\ref{Gfree}) and (\ref{Gfreegamma})). Clearly, the UV regularization by a momentum cutoff as discussed in Sec.~\ref{sectionregularization} is most naturally implemented in our numerical approach. It is sufficient to choose the momenta of the (Gauss-Legendre) grid to be limited by the value of the cutoff. Such a cutoff is, in principle, not needed in the case of pure Coulomb potentials. There the solution of the integral equations is possible without any cutoff and would correspond to a different regularization scheme with different short-distance coefficients. However, the potentials $\tilde V_{\mbox{\tiny BF}}$, $\tilde V_{\mbox{\tiny NA}}$ and the kinematic corrections introduced at NNLO require a UV regularization already for purely numerical reasons, which can be seen from naive power counting in Eq.~(\ref{Samputatednew}). \vspace{1.5cm} \section{A First Analysis in the Pole Mass Scheme} \label{sectionpolescheme} In this section we carry out a first brief analysis of the total cross sections $Q_t^2 R^v$ and $R^a$ and their three-momentum distributions in the pole mass scheme. We do this even though it is known that in the pole mass scheme there are uncomfortably large NNLO corrections in the location of the $1S$ peak position as well as in the normalization of the total vector-current-induced cross section $R^v$~\cite{Hoang3,Melnikov3,Yakovlev1}. We will show in Sec.~\ref{sectionuncertainties} that these large corrections are a consequence of the pole mass scheme, and that the pole mass definition has to be abandoned as far as a precise extraction of a top quark mass from experimental data is concerned. Nevertheless, the pole mass is a well defined quantity in the framework of perturbation theory~\cite{Kronfeld1,Tarrach1}, and, despite all its problems at larger orders of perturbation theory, remains a very convenient mass parameter to use for the formulation of (P)NRQCD and for calculations of the cross section. Thus a brief analysis in the pole mass scheme serves as a reference point with which results obtained with different approaches can be compared and from which we can visualize in which way alternative top quark mass definitions can improve the situation and in which way they cannot. \begin{figure}[t!] \begin{center} \leavevmode \epsfxsize=3.8cm \epsffile[200 400 400 530]{fig3a.ps} \hspace{4.cm} \epsfxsize=3.8cm \leavevmode \epsffile[200 400 400 530]{fig3b.ps}\\[3cm] \leavevmode \epsfxsize=3.8cm \epsffile[200 400 400 530]{fig3c.ps} \vskip 2.7cm \caption{\label{figtotpole} The total vector-current-induced cross section $Q_t^2 R^v$ for centre-of-mass energies $344\,\mbox{GeV}< \sqrt{q^2}< 352$~GeV in the pole mass scheme. The dependence on the renormalization scale $\mu$ (a), on the cutoff $\Lambda$ (b) and on $\alpha_s(M_Z)$ (c) is displayed. More details and the choice of parameters are given in the text. } \end{center} \end{figure} \begin{figure}[t!] \begin{center} \leavevmode \epsfxsize=3.8cm \epsffile[200 400 400 530]{fig4a.ps} \hspace{4.2cm} \epsfxsize=3.8cm \leavevmode \epsffile[200 400 400 530]{fig4b.ps}\\[3cm] \leavevmode \epsfxsize=3.8cm \epsffile[200 400 400 530]{fig4c.ps} \vskip 2.7cm \caption{\label{figtotpoleax} The total axial-vector-current induced cross section $R^a$ for centre-of-mass energies $344\,\mbox{GeV}< \sqrt{q^2}< 352$~GeV in the pole mass scheme. The dependence on the renormalization scale $\mu$ (a), on the cutoff $\Lambda$ (b) and on $\alpha_s(M_Z)$ (c) is displayed. More details and the choice of parameters are given in the text. } \end{center} \end{figure} \begin{figure}[t!] \begin{center} \leavevmode \epsfxsize=3.8cm \epsffile[200 400 400 530]{fig5a.ps} \hspace{4.4cm} \epsfxsize=3.8cm \leavevmode \epsffile[200 400 400 530]{fig5b.ps}\\[3cm] \leavevmode \epsfxsize=3.8cm \epsffile[200 400 400 530]{fig5c.ps} \vskip 2.7cm \caption{\label{figdistpole} The three-momentum distribution of the vector-current-induced cross section $Q_t^2 R^v$ for centre-of-mass energies $\sqrt{q^2}=M_{peak}$ and $M_{peak}+5$~GeV in the pole mass scheme. The dependence on the renormalization scale $\mu$ (a), on the cutoff $\Lambda$ (b) and on $\alpha_s(M_Z)$ (c) is displayed. More details and the choice of parameters are given in the text. } \end{center} \end{figure} \begin{figure}[t!] \begin{center} \leavevmode \epsfxsize=3.8cm \epsffile[200 400 400 530]{fig6a.ps} \hspace{4.4cm} \epsfxsize=3.8cm \leavevmode \epsffile[200 400 400 530]{fig6b.ps}\\[3cm] \leavevmode \epsfxsize=3.8cm \epsffile[200 400 400 530]{fig6c.ps} \vskip 2.7cm \caption{\label{figdistpoleax} The three-momentum distribution of the axial-vector-current-induced cross section $R^a$ for centre-of-mass energies $\sqrt{q^2}=M_{peak}$ and $M_{peak}+5$~GeV in the pole mass scheme. The dependence on the renormalization scale $\mu$ (a), on the cutoff $\Lambda$ (b) and on $\alpha_s(M_Z)$ (c) is displayed. More details and the choice of parameters are given in the text. } \end{center} \end{figure} In Figs.~\ref{figtotpole} the total vector-current-induced cross section $Q_t^2 R^v$ is displayed for $344\,\mbox{GeV}<\sqrt{q^2}<352\,\mbox{GeV}$ at LO (dotted lines), NLO (dashed lines) and NNLO (solid lines). The LO curves are determined from Eq.~(\ref{NNLOSchroedingergamma}), excluding all relativistic corrections and taking into account only the ${\cal{O}}(\alpha_s)$ contribution to the Coulomb potential and setting $C^v=1$. In addition, the NLO curves also include the NLO corrections to $C^v$ and the ${\cal{O}}(\alpha_s^2)$ contributions of the Coulomb potential. At NNLO all contributions mentioned in this paper are included. Only at NNLO have we taken into account the phase space counter-terms displayed in Eqs.~(\ref{vectorcrosssectionexpandedgamma}) and (\ref{axialvectorcrosssectionexpandedgamma}). In all figures shown in this section the top quark width is chosen as $\Gamma_t=1.43$~GeV and the top quark pole mass as $M_t=175$~GeV. Figure~\ref{figtotpole}a displays the dependence on the renormalization scale for $\mu=15$, $30$ and $60$~GeV for $\alpha_s(M_Z)=0.118$, $\Lambda=175$~GeV. For $\sqrt{q^2}\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 347$~GeV, $\mu=15$, $30$ and $60$~GeV corresponds to the upper, middle and lower curves, respectively. We note that in Refs.~\cite{Hoang3,Melnikov3,Yakovlev1} the renormalization scale governing the strong coupling of the potentials in the Schr\"odinger equation~(\ref{NNLOSchroedinger}) has been chosen between $50$ and $100$~GeV. Regarding the fact that the inverse Bohr radius $\sim M_t\alpha_s$ is the scale that governs the cross section at NNLO, the natural renormalization scale is of order $30$~GeV. This causes logarithms of the ratio $\Lambda/\mu\sim 1/\alpha_s$ in the vector current correlator (see App.~\ref{appendixshortdistance}), which, however, are connected only to the running of the strong coupling. As has already been demonstrated in previous publications on the NNLO corrections~\cite{Hoang3,Melnikov3,Yakovlev1}, the position of the $1S$ peak varies considerably at the different orders of approximation and for the different choices of the renormalization scale $\mu$. (For explicit numbers see Table~\ref{tabpeakpositionpole}.) It is also evident that the normalization of the total cross section is subject to large corrections. We see that, in general, the NNLO corrections to the normalization are of order $20\%$ and as large as the NLO ones. Further, the dependence on the renormalization scale is even larger at NNLO than at NLO. In Fig.~\ref{figtotpole}b the dependence of $Q_t^2 R^v$ on the choice of the cutoff $\Lambda$ is shown for $\alpha_s(M_Z)=0.118$, $\mu=30$~GeV and $\Lambda=90$, $175$ and $350$~GeV. For LO and NNLO $\Lambda=90$, $175$ and $350$~GeV correspond to the lower, middle and upper curves, respectively. For NLO $\Lambda=90$, $175$ and $350$~GeV correspond to the lower, upper and middle curve, respectively. Whereas the dependence of the LO cross section on the choice of $\Lambda$ is quite dramatic, because at LO there is no short-distance correction that could compensate for a variation in $\Lambda$, the variations at NLO and NNLO are significantly smaller (or order 5--10\%). However, there is again no reduction of the variation from NLO to NNLO. The variation of the cross section with respect to $\Lambda$ is small compared with the variation with respect to the renormalization scale $\mu$ for centre-of-mass energies closer and below the peak. Figure~\ref{figtotpole}c displays the dependence of $Q_t^2 R^v$ on the choice of $\alpha_s(M_Z)$ for $\alpha_s(M_z)=0.113, 0.118$ and $0.123$ and $\Lambda=175$~GeV, $\mu=30$~GeV. As for the variations in the renormalization scale we see that the position of the peak depends considerably on the choice of $\alpha_s$. We observe a strong positive correlation between the choice of $M_t$ and $\alpha_s(M_Z)$. Because the peak in the total cross section is the most pronounced feature of the total cross section, its behaviour directly reflects the quality of a top mass extraction from experimental data. Thus, if one would like to fit the pole mass to data for the cross section line-shape from a threshold scan, one finds a strong positive correlation between the pole mass and the strong coupling~\cite{Orange1} and a strong dependence of $M_t$ on the choice of the renormalization scale leading to quite large theoretical uncertainties in the pole mass measurements. In Figs.~\ref{figtotpoleax}a,b,c the total axial-vector-current-induced cross section $R^a$ is displayed for the same input parameters as in Figs.~\ref{figtotpole}. Figure~\ref{figtotpoleax}a shows the dependence on the renormalization scale for $\mu=15$ (solid line), $30$ (dashed line) and $60$~GeV (dotted line). Figure~\ref{figtotpoleax}b exhibits the dependence of the cutoff for $\Lambda=90$ (solid line), $175$ (dashed line) and $350$~GeV (dotted line). Figure~\ref{figtotpoleax}c shows $R^a$ for $\alpha_s(M_z)=0.113$ (solid line), $0.118$ (dashed line) and $0.123$ (dotted line). We observe that due to the $v$ suppression of the axial-vector currents no peak-line enhancement as in the vector current case is visible. The variations of $R^a$ with respect to the renormalization scale and the cutoff are quite large, because $R^a$ has only been determined at leading order, and the short-distance coefficient $C^a$ does not contain any corrections that could compensate for the variations. We note that the NLO corrections to $R^a$ can be implemented in the same way as for $R^v$. They are, however, beyond NNLO and have therefore not been included into our analysis. From the phenomenological point of view the next-to-leading order corrections are irrelevant if one takes into account the small normalization of $R^a$ compared to $R^v$. In Figs.~\ref{figdistpole}a,b and c the LO (dotted lines), NLO (dashed line) and NNLO (solid lines) top-antitop vector-current-induced three momentum distribution $Q_t^2 d R^v/d |\mbox{\boldmath$k$}_t|$ is shown for $0 < |\mbox{\boldmath$k$}_t| < 60$~GeV and both for centre-of-mass energies exactly on top of the visible peak, $\sqrt{q^2}=M_{peak}$, and for $\sqrt{q^2}=M_{peak}+5$~GeV. The input parameters have been chosen as in Figs.~\ref{figtotpole}. Figure~\ref{figdistpole}a shows the distributions for $\mu=15$ and $60$~GeV. At LO and NNLO $\mu=15$~GeV corresponds to the upper curves below the peak and $\mu=60$~GeV to the lower curves. At NLO $\mu=60$~GeV corresponds to the higher peak and $\mu=15$~GeV to the lower. Figure~\ref{figdistpole}b displays the dependence of the distributions on the cutoff for $\Lambda=90$ (lower curves) and $350$~GeV (upper curves), and Fig.~\ref{figdistpole}c exhibits the dependence of the distributions on the strong coupling for $\alpha_s(M_Z)=0.113$ and $0.123$. Below the peak, the larger value of $\alpha_s(M_Z)$ always corresponds to the upper curve. As for the total cross section, we observe a strong dependence of the normalization of the distributions on the renormalization scale and the cutoff. For $\sqrt{q^2}=M_{peak}+5$~GeV, the dependence of the peak position on the renormalization scale is particularly strong.\footnote{ The peak position is always located approximately at $|\mbox{\boldmath$k$}_t|\approx (p_0^4+M_t^2\Gamma_t^2)^{1/4}$, which can be regarded as the effective three momentum of the top quarks. } Finally, in Figs.~\ref{figdistpoleax}a,b and c the top-antitop axial-vector-current-induced three momentum distribution $d R^a/d |\mbox{\boldmath$k$}_t|$ is shown for $0 < |\mbox{\boldmath$k$}_t| < 60$~GeV and for centre-of-mass energies exactly on top of the visible peak, $\sqrt{q^2}=M_{peak}$, and for $\sqrt{q^2}=M_{peak}+5$~GeV. The input parameters have been chosen as before. Figure~\ref{figdistpoleax}a shows the distributions for $\mu=15$ (solid curves), $30$ (dashed curves) and $60$~GeV (dotted curves). Figure~\ref{figdistpoleax}b displays the dependence of the distributions on the cutoff for $\Lambda=90$ (solid curves) $175$ (dashed curves) and $350$~GeV (dotted curves), and Fig.~\ref{figdistpoleax}c exhibits the dependence of the distributions on the strong coupling for $\alpha_s(M_Z)=0.113$ (solid curves), $0.118$ (dashed curves) and $0.123$ (dotted curves). As expected, the momentum distribution is strongly suppressed for smaller centre-of-mass energies. The variations of the normalization of the distributions are comparable to the variations of the normalization of the total cross sections. \vspace{1.5cm} \section{Theoretical Uncertainties} \label{sectionuncertainties} In the previous section we have seen that the location of the peak position and the normalization of the vector-current-induced total cross section, as well as of its three-momentum distribution, receive large NNLO corrections in the pole mass scheme. In the following two subsections we discuss the origin of the large corrections in the vector-current-induced total cross section. For the peak position of the total cross section we propose a solution, which allows for a considerable stabilization. \vspace{1cm} \subsection{The $1S$ Top Quark Mass and the Peak Position} \label{subsectionmass} From past experience in the theoretical description of $B$-meson decays~\cite{Beneke3,Bigi1}, it is well known that the pole mass, defined as the pole of the perturbative quark propagator, although infrared-finite to all orders in perturbation theory~\cite{Kronfeld1,Tarrach1}, is a concept that is ambiguous to an amount of order $\Lambda_{QCD}$. This might be a reflection of the fact that the perturbative quark pole does not exist in reality because of confinement. Within the framework of perturbation theory this ambiguity is caused by an $n$-factorial increase of the coefficients in the perturbative relation between the pole mass and a short-distance mass such as $\overline{\mbox{MS}}$. The large corrections are caused by an increasing infrared sensitivity of the perturbative coefficients for large orders. It has also been shown that the large top width does not lead to a suppression of these large corrections~\cite{Willenbrock1}. From this point of view the unstable behaviour of the peak position in the total cross section is not unexpected and it would be quite appealing conceptually to conclude that the use of a short-distance mass instead of the pole definition would cure the problem. In fact, it has been demonstrated in Refs.~\cite{Hoang6,Beneke4} that the dominant source of infrared sensitivity in the Green function of the Schr\"odinger equation~(\ref{NNLOSchroedinger}) comes from the terms in the static energy $2 M_t + V(\mbox{\boldmath $r$})$. Whereas the rest (pole) mass energy $2 M_t$ and the potential energy $V$ (which has traditionally been calculated in the pole mass scheme~\cite{Fischler1,Billoire1,Schroeder1,Peter1}) are individually ambiguous to an amount of order $\Lambda_{QCD}$~\cite{Beneke3,Bigi1,Aglietti1}, the sum of both is not~\cite{Hoang6,Beneke4}. This shows that quantities such as spectra or the total cross section calculated from Eq.~(\ref{NNLOSchroedinger}) are much less sensitive to infrared momenta than the pole mass itself, rendering it an irrelevant mass parameter. Thus, any sensible mass definition used to parametrize the top-antitop cross section close to threshold should have no ambiguity of order $\Lambda_{QCD}$, i.e. it should be a short-distance mass. However, we emphasize that, in practice, large corrections at lower orders in perturbation theory in the pole scheme do not necessarily come from the ambiguity in the pole mass. This is because the cancellations of infrared sensitive contributions in static and rest mass energy is a phenomenon that is relevant in large orders where the corresponding series are dominated by the most infrared sensitive contributions in the loop integrals. Thus, large corrections in perturbation theory at low orders could very well come from scales which are not infrared. To get a clearer picture for the case of the peak position in the total cross section, let us have a look at the size of the individual corrections to the peak position. In Table~\ref{tabpeakpositionpole} we have displayed the LO, NLO and NNLO contributions to the peak position with respect to the pole rest mass energy \begin{eqnarray} M_{peak} & = & 2M_t \, - \, \delta M_{peak}^{LO} \, - \, \delta M_{peak}^{NLO} \, - \, \delta M_{peak}^{NNLO,\beta_0} \, - \, \delta M_{peak}^{NNLO,rest} \nonumber \\[3mm] & = & 2M_t \, - \, \delta M_{peak} \,, \end{eqnarray} for $M_t=175$~GeV, $\Gamma_t=1.43$~GeV, $\alpha_s(M_Z)=0.113, 0.118$ and $0.123$, and $\mu=15, 30$ and $60$~GeV. \begin{table}[t] \vskip 7mm \begin{center} \begin{tabular}{|c||c||c|c|c|c|c|} \hline $\mu [\mbox{GeV}]$ & $\alpha_s(M_Z)$ & $\delta M_{peak}^{LO} $ & $\delta M_{peak}^{NLO} $ & $\delta M_{peak}^{NNLO,\beta_0} $ & $\delta M_{peak}^{NNLO,rest} $ & $\delta M_{peak} $ \\ \hline\hline $15$ & $0.113$ & $1.60$ & $0.28$ & $0.20$ & $0.32$ & $2.40$ \\\hline $30$ & & $1.02$ & $0.70$ & $0.27$ & $0.24$ & $2.23$ \\\hline $60$ & & $0.35$ & $1.18$ & $0.36$ & $0.18$ & $2.07$ \\ \hline\hline $15$ & $0.118$ & $1.89$ & $0.31$ & $0.24$ & $0.37$ & $2.80$ \\\hline $30$ & & $1.26$ & $0.75$ & $0.31$ & $0.27$ & $2.58$ \\\hline $60$ & & $0.69$ & $1.09$ & $0.40$ & $0.20$ & $2.39$ \\\hline\hline $15$ & $0.123$ & $2.17$ & $0.33$ & $0.29$ & $0.43$ & $3.23$ \\\hline $30$ & & $1.49$ & $0.81$ & $0.36$ & $0.31$ & $2.96$ \\\hline $60$ & & $0.92$ & $1.12$ & $0.45$ & $0.23$ & $2.72$ \\ \hline \end{tabular} \caption{\label{tabpeakpositionpole} LO, NLO and NNLO contributions to the peak position of the total vector current induced cross section $R^v$ in GeV in the pole mass scheme for $M_t=175$~GeV, $\Gamma_t=1.43$~GeV, $\alpha_s(M_Z)=0.113, 0.118$ and $0.123$, and $\mu=15, 30$ and $60$~GeV, respectively. For the strong coupling two-loop running has been employed. The results are insensitive to the choice of the cutoff scale $\Lambda\sim M_t$. } \end{center} \vskip 3mm \end{table} At NNLO we have separated from the rest the contributions with the highest power of $\beta_0$, which represent the contributions most sensitive to infrared momenta\footnote{ These contributions are determined by using the replacement rule $n_f\to -\frac{3}{2}\beta_0$ for the terms with the highest power of $n_f$, where $n_f$ is the number of light quark species. This method is called ``naive non-Abelianization'' and accounts for the most infrared sensitive contributions in perturbative series relating the pole mass to a short-distance mass~\cite{Beneke5}. The result of this replacement is referred to as the ``large-$\beta_0$'' limit. }. All the large-$\beta_0$ terms originate from the Coulomb potential $V_c$, Eq.~(\ref{NNLOCoulomb}). From the numbers presented in Table~\ref{tabpeakpositionpole} we see that depending on the choice of the renormalization scale the large-$\beta_0$ contributions at NNLO contribute between about $30$ and $60\%$ to the total NNLO corrections to the peak position. Thus at NNLO the large shift in the peak position consists to approximately equal parts of corrections very sensitive to infrared momenta and corrections coming from subleading infrared terms and relativistic corrections. From this we can conclude that using some unspecified short-distance mass definition instead of the pole mass does not necessarily lead to smaller NNLO corrections to the peak position because the most infrared sensitive terms are not yet dominating at NNLO. For the same reason, we cannot conclude that some unspecified short-distance mass definition necessarily leads to a significantly reduced renormalization scale dependence of the NNLO peak position or to a smaller correlation between the peak position and the strong coupling. Thus, the question of which mass definition one should use is not only a conceptual issue, but also a practical one. We formulate two requirements for a proper top mass definition for the total cross section close to threshold: \begin{itemize} \item[A)] it must be a short-distance mass, and \item[B)] it must lead to a considerable stabilization of the peak location with respect to the order of approximation used and also to variations of parameters such as the strong coupling or the renormalization scale. \end{itemize} Requirement A reflects the necessity that, if a top mass determination at the linear collider with uncertainties of $200$~MeV or even better is intended, the corresponding mass parameter must be free of intrinsic ambiguities of order $\Lambda_{QCD}$. In addition, only a short-distance mass can be reliably related to the $\overline{\mbox{MS}}$ top quark mass, which is the preferred mass parameter used in calculations at high energies and for top quark corrections to electroweak precision observables. Requirement B ensures that the mass parameter can be extracted from experimental data with small systematic uncertainties. The mass that seems to be most appropriate to us to fulfil this task, because it is closely related to the peak position in the vector-current-induced cross section, is what we call the $1S$ mass, $M_{1S}$. The $1S$ mass is defined as half the perturbative mass of the fictitious toponium $1\,{}^3\!S_1$ ground state, where the top quark is assumed to be stable. Expressed in terms of the pole mass, the $1S$ mass reads ($a_s=\alpha_s(\mu)$)~\cite{Melnikov5,Pineda2}: \begin{eqnarray} M_{1S} & = & M_t \, - \, \epsilon \frac{M_t\,C_F^2\,a_s^2}{8} \nonumber \\ & & - \,\epsilon^2 \, \frac{M_t\,C_F^2\,a_s^2}{8}\, \Big(\frac{a_s}{\pi}\Big)\,\bigg[\, \beta_0\,\bigg( L + 1 \,\bigg) + \frac{a_1}{2} \,\bigg] \nonumber \\ & & - \,\epsilon^3 \frac{M_t\,C_F^2\,a_s^2}{8}\, \Big(\frac{a_s}{\pi}\Big)^2\, \bigg[\, \beta_0^2\,\bigg(\, \frac{3}{4} L^2 + L + \frac{\zeta_3}{2} + \frac{\pi^2}{24} + \frac{1}{4} \,\bigg) + \beta_0\,\frac{a_1}{2}\,\bigg(\, \frac{3}{2}\,L + 1 \,\bigg) \nonumber\\[3mm] & & \hspace{1.5cm} + \frac{\beta_1}{4}\,\bigg(\, L + 1 \,\bigg) + \frac{a_1^2}{16} + \frac{a_2}{8} + \bigg(\, C_A - \frac{C_F}{48} \,\bigg)\, C_F \pi^2 \,\bigg] \,, \label{1Sdef} \end{eqnarray} where \begin{eqnarray} L & \equiv & \ln\Big(\frac{\mu}{C_F\,a_s\,M_t}\Big) \,, \end{eqnarray} and the contributions at LO, NLO and NNLO are labelled by the powers $\epsilon$, $\epsilon^2$ and $\epsilon^3$, respectively, of the auxiliary parameter $\epsilon=1$. In general, the electroweak corrections not calculated in this work can lead to further corrections in Eq.~(\ref{1Sdef}). We note that $2 M_{1S}$ is not equal to the actual peak position visible in the vector-current-induced total cross section because the top quark width leads to an additional shift of the peak by about $+200$~MeV.\footnote{ The difference between $2 M_{1S}$ and the peak location of the vector-current-induced total cross section is proportional to $\Gamma_t$ times a function of $\Gamma_t/(M_t \alpha_s^2)$. For $\Gamma_t\ll M_t \alpha_s^2$ the difference $ M_{peak}-2 M_{1S}$ is proportional to $\Gamma_t^4/(M_t C_F^2 \alpha_s^2)^3$. The size of the difference between $2 M_{1S}$ and the peak location of about $+200$~MeV can be seen in Figs.~\ref{figtotm1S}. } In principle it would also be possible to define a mass that would be equal to half the actual visible peak position. Except for additional corrections coming from the top quark width, this would also require the inclusion of an additional shift coming from the axial-vector-induced cross section. Such a definition would, however, not necessarily be more useful, since the experimentally measurable line-shape of the total cross section at a future $e^+e^-$ or $\mu^+\mu^-$ collider will be distorted by initial state radiation and beamstrahlung. In the case of a muon collider these effects lead to an additional shift and in the case of the $e^+e^-$ linear collider even to the disappearance of the peak~\cite{Orange1}. Thus one has to consider the $1S$ mass, like the $\overline{\mbox{MS}}$ mass, as a fictitious mass parameter, which to NNLO is defined through the perturbative series given in Eq.~(\ref{1Sdef}). Nevertheless, twice the $1S$ mass is quite close to the peak location and, by construction, leads to a considerable reduction of the variation of the peak position with respect to the order of approximation, the strong coupling and the renormalization scale. To show that $M_{1S}$ is indeed a short-distance mass we recall that the static energy $2M_t + V_c(\mbox{\boldmath r})$ represents the dominant infrared sensitive contribution in the Schr\"odinger equation~(\ref{NNLOSchroedinger}). The difference between $M_{1S}$ and the pole mass is\footnote{ Strictly speaking, the simple form of Eq.~(\ref{1Spolefull}) is true only up to NNLO because of retardation effects, which set in at N$^3$LO. Thus, in general, there would also be a non-trivial integration over time components. The form of our proof also depends on the assumption that the static potential is an infrared finite quantity. That this is most probably not the case was already pointed out some time ago in Ref.~\cite{Appelquist1}, because the perturbative static potential might become sensitive to scales below the inverse Bohr radius at ${\cal{O}}(\alpha_s^4)$. Some contributions at ${\cal{O}}(\alpha_s^4)$ have recently been calculated in Ref.~\cite{Pineda3}. Up to ${\cal{O}}(\alpha_s^3)$, i.e. NNLO in the non-relativistic expansion, the perturbative potential has been proven to be finite by complete calculations~\cite{Schroeder1,Peter1}. Because $M_{1S}$ is defined as a physical quantity this would not affect the final conclusion that it is a short-distance mass, but it would change the form of the proof considerably. } \begin{eqnarray} M_{1S}-M_t & = & \frac{1}{2}\, \int\frac{d^3\mbox{\boldmath $p$}}{(2\pi)^3} \frac{d^3\mbox{\boldmath $q$}}{(2\pi)^3}\, \tilde\Phi_{1S}^*(\mbox{\boldmath $p$})\, {\cal{H}}(\mbox{\boldmath $p$},\mbox{\boldmath $q$})\, \tilde\Phi_{1S}(\mbox{\boldmath $q$}) \,, \label{1Spolefull} \end{eqnarray} where ${\cal{H}}$ is the Hamiltonian of Eq.~(\ref{NNLOSchroedinger}) and $\tilde\Phi_{1S}$ the normalized wave function of the $1S$ state in momentum space representation. The dominant infrared sensitive contribution in relation~(\ref{1Spolefull}) reads \begin{eqnarray} \Big(\,M_{1S}-M_t\,\Big)^{IR} & \sim & \frac{1}{2}\, \int\frac{d^3\mbox{\boldmath $p$}}{(2\pi)^3} \frac{d^3\mbox{\boldmath $q$}}{(2\pi)^3}\, \tilde\Phi_{1S}^*(\mbox{\boldmath $p$})\, \tilde V_c(\mbox{\boldmath $p$}-\mbox{\boldmath $q$})\, \tilde\Phi_{1S}(\mbox{\boldmath $q$}) \,. \label{1SpoleIR1} \end{eqnarray} Because the infrared region in Eq.~(\ref{1SpoleIR1}) is given by $|\mbox{\boldmath $p$}-\mbox{\boldmath $q$}|<\mu_f$, where $\mu_f$ is much smaller than the inverse Bohr radius $\sim M_t\alpha_s$, the characteristic scale governing the dynamics described by the wave function, we can simplify relation~(\ref{1SpoleIR1}), \begin{eqnarray} \Big(\,M_{1S}-M_t\,\Big)^{IR} & \sim & \frac{1}{2}\, \int\limits^{|\mbox{\boldmath $p$}-\mbox{\boldmath $q$}|<\mu_f} \frac{d^3\mbox{\boldmath $p$}}{(2\pi)^3} \frac{d^3\mbox{\boldmath $q$}}{(2\pi)^3}\, |\tilde\Phi_{1S}(\mbox{\boldmath $p$})|^2\, \tilde V_c(\mbox{\boldmath $p$}-\mbox{\boldmath $q$}) \nonumber \\ & \sim & \frac{1}{2}\,\int\limits^{|\mbox{\boldmath $q$}|<\mu_f}\, \frac{d^3\mbox{\boldmath $q$}}{(2\pi)^3}\, \tilde V_c(\mbox{\boldmath $q$}) \,. \label{1SpoleIR2} \end{eqnarray} It has been shown in Refs.~\cite{Beneke4,Bigi1} that the RHS of Eq.~(\ref{1SpoleIR2}) is equivalent to the dominant infrared contributions of the difference between the $\overline{\mbox{MS}}$ and pole mass. Therefore the relation between $M_{1S}$ and the $\overline{\mbox{MS}}$ mass $\bar m_t$ only contains subleading infrared contributions, which are suppressed by at least one power of $1/M_t$. In other words the ambiguity in the relation between $M_{1S}$ and $\bar m_t$ is parametrically of order $\Lambda_{QCD}^2/M_t$. This proves that $M_{1S}$ is a short-distance mass. We also see from Eq.~(\ref{1SpoleIR2}) that, if the pole mass is expressed in terms of the $1S$ mass and if the resulting mass difference $2(M_t-M_{1S})$ is absorbed into the potential, the rest mass and the potential energy term contained in the total static energy are individually free of ambiguities of order $\Lambda_{QCD}$. The RHS of Eq.~(\ref{1SpoleIR2}) just subtracts the low momentum (i.e. dominant infrared sensitive) contribution from the Coulomb potential $V_c(\mbox{\boldmath $x$})$. We note that in order to implement the $1S$ mass definition into our numerical codes, which solve the integral equations~(\ref{Sintegral}) and (\ref{Pintegral}), we have to invert relation~(\ref{1Sdef}). It has been shown in Refs.~\cite{Hoang7,Hoang8} that a consistent way to achieve this task is to carry out the inversion with respect to the auxiliary parameter $\epsilon$. For the reason that this modified perturbative expansion has been applied for the first time to express inclusive $B$ decays in terms of the $\Upsilon(1S)$ mass, it has been called the ``Upsilon expansion''.~\cite{Hoang7,Hoang8} If the $1S$ mass is expressed in terms of the $\overline{\mbox{MS}}$ mass, which is related to the pole mass by a series of the form $\bar m_t-M_t = M_t \sum_{n=1}^\infty a_n\alpha_s^n$, one has to consider a term $\propto \alpha_s^n$ in this relation of order $\epsilon^n$ in the Upsilon expansion. In other words, if one relates the $1S$ mass to a mass which is different from the pole mass, one must combine terms of different order in $\alpha_s$. As an example, this means that in order to relate the N$^k$LO $1S$ mass to the $\overline{\mbox{MS}}$ mass one needs to know its relation to the pole mass to ${\cal{O}}(\alpha_s^{k+1})$. This is necessary because this is the only way in which the high order large perturbative corrections coming from infrared-sensitive terms are cancelled. \begin{figure}[t!] \begin{center} \leavevmode \epsfxsize=3.8cm \epsffile[200 400 400 530]{fig7a.ps} \hspace{4.2cm} \epsfxsize=3.8cm \leavevmode \epsffile[200 400 400 530]{fig7b.ps}\\[3cm] \leavevmode \epsfxsize=3.8cm \epsffile[200 400 400 530]{fig7c.ps} \vskip 2.7cm \caption{\label{figtotm1S} The total vector-current-induced cross section $Q_t^2 R^v$ for centre-of-mass energies $346\,\mbox{GeV}< \sqrt{q^2}< 354$~GeV in the $1S$ mass scheme. The dependence on the renormalization scale $\mu$ (a), on the cutoff $\Lambda$ (b) and on $\alpha_s(M_Z)$ (c) is displayed. More details and the choice of parameters are given in the text. } \end{center} \end{figure} \begin{figure}[t!] \begin{center} \leavevmode \epsfxsize=3.8cm \epsffile[200 400 400 530]{fig8a.ps} \hspace{4.2cm} \epsfxsize=3.8cm \leavevmode \epsffile[200 400 400 530]{fig8b.ps}\\[3cm] \leavevmode \epsfxsize=3.8cm \epsffile[200 400 400 530]{fig8c.ps} \vskip 2.7cm \caption{\label{figtotm1Sax} The total axial-vector-current-induced cross section $R^a$ for centre-of-mass energies $346\,\mbox{GeV}< \sqrt{q^2}< 354$~GeV in the $1S$ mass scheme. The dependence on the renormalization scale $\mu$ (a), on the cutoff $\Lambda$ (b) and on $\alpha_s(M_Z)$ (c) is displayed. More details and the choice of parameters are given in the text. } \end{center} \end{figure} \begin{figure}[t!] \begin{center} \leavevmode \epsfxsize=3.8cm \epsffile[200 400 400 530]{fig9a.ps} \hspace{4.4cm} \epsfxsize=3.8cm \leavevmode \epsffile[200 400 400 530]{fig9b.ps}\\[3cm] \leavevmode \epsfxsize=3.8cm \epsffile[200 400 400 530]{fig9c.ps} \vskip 2.7cm \caption{\label{figdistm1S} The three-momentum distribution of the vector-current-induced cross section $Q_t^2 R^v$ for centre-of-mass energies $\sqrt{q^2}=M_{peak}$ and $M_{peak}+5$~GeV in the $1S$ mass scheme. The dependence on the renormalization scale $\mu$ (a), on the cutoff $\Lambda$ (b) and on $\alpha_s(M_Z)$ (c) is displayed. More details and the choice of parameters are given in the text. } \end{center} \end{figure} \begin{figure}[t!] \begin{center} \leavevmode \epsfxsize=3.8cm \epsffile[200 400 400 530]{fig10a.ps} \hspace{4.4cm} \epsfxsize=3.8cm \leavevmode \epsffile[200 400 400 530]{fig10b.ps}\\[3cm] \leavevmode \epsfxsize=3.8cm \epsffile[200 400 400 530]{fig10c.ps} \vskip 2.7cm \caption{\label{figaxdistm1S} The three-momentum distribution of the axial-vector-current-induced cross section $R^a$ for centre-of-mass energies $\sqrt{q^2}=M_{peak}$ and $M_{peak}+5$~GeV in the $1S$ mass scheme. The dependence on the renormalization scale $\mu$ (a), on the cutoff $\Lambda$ (b) and on $\alpha_s(M_Z)$ (c) is displayed. More details and the choice of parameters are given in the text. } \end{center} \end{figure} In Figs.~\ref{figtotm1S} the total vector-current-induced cross section $Q_t^2 R^v$ is displayed in the $1S$ scheme for $346\,\mbox{GeV}<\sqrt{q^2}<354\,\mbox{GeV}$ at LO (dotted lines), NLO (dashed lines) and NNLO (solid lines). In all figures shown in this section the top quark width is chosen as $\Gamma_t=1.43$~GeV and the top quark $1S$ mass as $M_{1S}=175$~GeV. Figure~\ref{figtotm1S}a displays the dependence on the renormalization scale for $\mu=15$, $30$ and $60$~GeV for $\alpha_s(M_Z)=0.118$ and $\Lambda=175$~GeV. At LO and NNLO the choices $\mu=15$, $30$ and $60$~GeV correspond to the upper, middle and lower curves. At NLO the choices $\mu=15$, $30$ and $60$~GeV correspond to the lower, middle and upper curves for centre-of-mass energies below the peak position. In Fig.~\ref{figtotm1S}b the dependence of $Q_t^2 R^v$ on the choice of the cutoff $\Lambda$ is shown for $\alpha_s(M_Z)=0.118$, $\mu=30$~GeV and $\Lambda=90$ (lower curves), $175$ (middle curves) and $350$~GeV (upper curves). Figure~\ref{figtotm1S}c displays the dependence of $Q_t^2 R^v$ on the choice of $\alpha_s(M_Z)$ for $\alpha_s(M_z)=0.113$ (lower curves), $0.118$ (middle curves) and $0.123$ (upper curves) and $\Lambda=175$~GeV, $\mu=30$~GeV. Comparing the result with the curves displayed in Figs.~\ref{figtotpole}, the improvement of the stability of the peak position is evident. The strong dependence on the renormalization scale and the strong correlation with $\alpha_s(M_Z)$ have vanished. However, we also observe that the large corrections in the normalization of the curves are essentially not affected at all. In Figs.~\ref{figtotm1Sax}a,b,c the total axial-vector-current-induced cross section $R^a$ is displayed in the $1S$ mass scheme for the same input parameters as in Figs.~\ref{figtotm1S}. Figure~\ref{figtotm1Sax}a shows the dependence on the renormalization scale for $\mu=15$ (solid line), $30$ (dashed line) and $60$~GeV (dotted line), respectively. Figure~\ref{figtotm1Sax}b exhibits the dependence of the cutoff for $\Lambda=90$ (solid line), $175$ (dashed line) and $350$~GeV (dotted line). Figure~\ref{figtotpoleax}c shows $R^a$ for $\alpha_s(M_z)=0.113$ (solid line), $0.118$ (dashed line) and $0.123$ (dotted line). Compared to the plots in the pole mass scheme, we find a slightly smaller variation in the normalization with respect to the renormalization scale and the choice of $\alpha_s(M_Z)$. Clearly the effects of using the $1S$ scheme instead of the pole one are much smaller in the axial-vector case because no peak is visible there. In Figs.~\ref{figdistm1S}a,b and c the LO (dotted lines), NLO (dashed lines) and NNLO (solid lines) top-antitop vector-current-induced three-momentum distribution $Q_t^2 d R^v/d |\mbox{\boldmath$k$}_t|$ is shown for $0 < |\mbox{\boldmath$k$}_t| < 60$~GeV in the $1S$ mass scheme for centre-of-mass energies exactly on top of the visible peak, $\sqrt{q^2}=M_{peak}$ and for $\sqrt{q^2}=M_{peak}+5$~GeV. The input parameters have been chosen as in Figs.~\ref{figtotm1S}. Figure~\ref{figdistm1S}a shows the distributions for $\mu=15$ and $60$~GeV. At LO and NNLO $\mu=15$~GeV corresponds to the upper curves and $\mu=60$~GeV to the lower curves for centre-of-mass energies below the peak. At NLO $\mu=60$~GeV corresponds to the higher peak and $\mu=15$~GeV to the lower. Figure~\ref{figdistm1S}b displays the dependence of the distributions on the cutoff for $\Lambda=90$ (lower curves) and $350$~GeV (upper curves), and Fig.~\ref{figdistm1S}c exhibits the dependence of the distributions on the strong coupling for $\alpha_s(M_Z)=0.113$ and $0.123$. Below the peak the larger value of $\alpha_s(M_Z)$ always corresponds to the upper curve. In Figs.~\ref{figaxdistm1S}a,b and c the top-antitop axial-vector-current-induced three-momentum distribution $d R^a/d |\mbox{\boldmath$k$}_t|$ is shown in the $1S$ scheme for $0 < |\mbox{\boldmath$k$}_t| < 60$~GeV and both for centre-of-mass energies exactly on top of the visible peak, $\sqrt{q^2}=M_{peak}$, and for $\sqrt{q^2}=M_{peak}+5$~GeV. The input parameters have been chosen as before. Figure~\ref{figaxdistm1S}a shows the distribution for $\mu=15$ (solid curves), $30$ (dashed curves) and $60$~GeV (dotted curves). Figure~\ref{figaxdistm1S}b displays the dependence of the distribution on the cutoff for $\Lambda=90$ (solid curves), $175$ (dashed curves) and $350$~GeV (dotted curves), and Fig.~\ref{figaxdistm1S}c exhibits the dependence of the distribution on the strong coupling for $\alpha_s(M_Z)=0.113$ (solid curves) $0.118$ (dashed curves) and $0.123$ (dotted curves). The curves shown in Figs.~\ref{figaxdistm1S} are somewhat higher than in Figs.~\ref{figdistpoleax}, because the choice of $175$~GeV for the top quark mass corresponds to a higher value for $M_{peak}$ in the $1S$ scheme. From Figs.~\ref{figdistm1S} and \ref{figaxdistm1S} it is evident that the $1S$ scheme does not essentially affect at all the three-momentum distributions. Compared to the results in the pole mass scheme the variations of the peak position remain unchanged. This can be understood from the fact that a mass redefinition corresponds to a shift in the centre-of-mass energy, but leaves the definition of the off-shell top quark three-momentum unchanged. In Table~\ref{tabpeakposition1S} we have displayed the LO, NLO and NNLO corrections to the peak position with respect to $2 M_{1S}$: \begin{eqnarray} M_{peak} & = & 2M_{1S} \, + \, \delta M_{peak,1S}^{LO} \, + \, \delta M_{peak,1S}^{NLO} \, + \, \delta M_{peak,1S}^{NNLO} \nonumber \\[3mm] & = & 2M_{1S} \, + \, \delta M_{peak,1S} \,, \end{eqnarray} in the $1S$ mass scheme for $M_{1S}=175$~GeV, $\Gamma_t=1.43$~GeV, $\alpha_s(M_Z)=0.113, 0.118$ and $0.123$, and $\mu=15, 30$ and $60$~GeV for various choices of the renormalization scale $\mu$ and the strong coupling $\alpha_s(M_Z)$. \begin{table}[t] \vskip 7mm \begin{center} \begin{tabular}{|c||c||c|c|c|c|} \hline $\mu [\mbox{GeV}]$ & $\alpha_s(M_Z)$ & $\delta M_{peak,1S}^{LO} $ & $\delta M_{peak,1S}^{NLO} $ & $\delta M_{peak,1S}^{NNLO} $ & $\delta M_{peak,1S} $ \\ \hline\hline $15$ & $0.113$ & $0.21$ & $0.03$ & $-0.03$ & $0.20$ \\\hline $30$ & & $0.38$ & $-0.11$ & $-0.09$ & $0.17$ \\\hline $60$ & & $0.78$ & $-0.50$ & $-0.11$ & $0.17$ \\ \hline\hline $15$ & $0.118$ & $0.16$ & $0.02$ & $-0.00$ & $0.17$ \\\hline $30$ & & $0.30$ & $-0.09$ & $-0.08$ & $0.12$ \\\hline $60$ & & $0.54$ & $-0.32$ & $-0.10$ & $0.11$ \\\hline\hline $15$ & $0.123$ & $0.12$ & $-0.00$ & $0.04$ & $0.16$ \\\hline $30$ & & $0.23$ & $-0.07$ & $-0.08$ & $0.08$ \\\hline $60$ & & $0.42$ & $-0.26$ & $-0.10$ & $0.07$ \\ \hline \end{tabular} \caption{\label{tabpeakposition1S} LO, NLO and NNLO contributions to the peak position of the total vector-current-induced cross section $R^v$ in GeV in the $1S$ mass scheme for $M_{1S}=175$~GeV, $\Gamma_t=1.43$~GeV, $\alpha_s(M_Z)=0.113, 0.118$ and $0.123$, and $\mu=15, 30$ and $60$~GeV, respectively. For the strong coupling two-loop running has been employed. The results are insensitive to the choice of the cutoff scale $\Lambda\sim 175$~GeV. } \end{center} \vskip 3mm \end{table} Taking the size of the NLO and NNLO corrections as a measure for the present theoretical uncertainty in the peak position, and assuming that the latter can be used to estimate the theoretical uncertainty in the determination of $M_{1S}$, we find that this uncertainty is approximately $200$~MeV. If the effects of beamstrahlung and initial state radiation at a future $e^+e^-$ or muon pair collider do not spoil a precise determination of the $1S$ mass with an uncertainty of $200$~MeV one has to ask the question how $M_{1S}$ is related to the top mass parameters usually used for calculations of physical observables that are not related to the threshold regime. In principle, one could use $M_{1S}$ as a new top mass parameter in its own right. This would, of course, require that all formulae be expressed in terms of $M_{1S}$, using the Upsilon expansion discussed after Eq.~(\ref{1SpoleIR2}). A more economical way is to relate the $1S$ mass to the $\overline{\mbox{MS}}$ top quark mass, which is a common mass parameter for perturbative calculations involving heavy quarks and which, in a number of cases, even leads to improved convergence properties of the perturbative series.\footnote{ Prominent cases are the top quark QCD corrections to the $\rho$-parameter~\cite{Chetyrkin1} and the massive quark pair production cross section at large energies~\cite{HoangKT1}. } Because the $1S$ mass is a short-distance mass, its perturbative relation to the $\overline{\mbox{MS}}$ mass is much better behaved at large orders than the corresponding relation of the pole mass. The relation between $M_{1S}$ and $\bar m_t(\bar m_t)$ can be derived from Eq.~(\ref{1Sdef}) and the relation between pole mass and $\bar m_t(\bar m_t)$ using the Upsilon expansion discussed above. We emphasize that the three-loop relation between the pole and the $\overline{\mbox{MS}}$ mass is needed to relate $M_{1S}$ and $\bar m_t(\bar m_t)$ at NNLO accuracy. Assuming that those 3-loop corrections can be approximated by the known corrections in the large-$\beta_0$ limit~\cite{Beneke5}, we find the following numerical value for $\bar m_t(\bar m_t)$ for $M_{1S}=175\pm 0.2$~GeV and $\alpha_s(M_Z)=0.118\pm x\,0.001$, \begin{eqnarray} \bar m_t(\bar m_t) & = & \bigg[\, 175 - 7.58\, \epsilon (\mbox{LO}) - 0.96\, \epsilon^2 (\mbox{NLO}) - 0.23\, \epsilon^3 (\mbox{NNLO}) \nonumber \\[3mm] & & \hspace{3cm} \pm 0.2 (\delta M_{1S}) \pm x\,0.07 (\delta \alpha_s) \,\bigg]~\mbox{GeV} \,. \label{MSmassestimate} \end{eqnarray} For the numbers given in Eq.~(\ref{MSmassestimate}) we have assumed an uncertainty in the value of the strong coupling at $M_Z$ of $x \,0.001$ in order to demonstrate the importance of $\alpha_s$ for the determination of $\bar m_t(\bar m_t)$. This uncertainty is independent of the order to which the relation between the pole and the $\overline{\mbox{MS}}$ mass is known because it comes from the LO term. We note that this fact shows that the strong correlation of the peak position to the strong coupling, which was visible in the pole mass scheme, is not necessarily eliminated by adopting the $1S$ scheme. This correlation might come back whenever the $1S$ mass is related to another short-distance mass or is used as a parameter in other quantities. However, the use of the $1S$ mass has the advantage to free the process of the mass extraction from the total cross section close to threshold also from strong dependences on other parameters such as the renormalization scale or the order of approximation used. Therefore systematic uncertainties are expected to be smaller if the $1S$ scheme is used for the threshold calculations. Equation~(\ref{MSmassestimate}) shows that the knowledge of the 3-loop corrections in the relation of pole and $\overline{\mbox{MS}}$ mass and a small uncertainty in $\alpha_s(M_Z)$ are crucial for a determination of $\bar m_t(\bar m_t)$ with uncertainties comparable to $\delta M_{1S}$. In recent literature there have been two other proposals for alternative short-distance mass definitions, which can also be used for a measurement of the top quark mass from the total cross section. In Refs.~\cite{Voloshin1,Bigi2} the ``low scale running mass'' was proposed to subtract the infrared behaviour from the heavy quark self energy. The ``low scale running mass'' was devised in order to improve the convergence of the perturbative series describing the contributions leading in $1/M_b$ in inclusive $B$-meson decays. Due to the universality of the dominant infrared sensitive contribution, the low scale running mass can also serve as a top mass definition, which leads to an improved stability of the peak position in the total cross section. The low scale running mass depends on the cutoff $\mu_{LS}$, which limits the momenta that are subtracted from the self energy. At order $\alpha_s$ (i.e. at LO) its relation to the pole mass reads~\cite{Voloshin1,Bigi2} \begin{eqnarray} m_t^{LS}(\mu_{LS}) - M_t & = & -\,\frac{16}{9}\,\frac{\alpha_s}{\pi}\,\mu_{LS}\, \bigg[\, 1 + {\cal{O}}(\alpha_s) + {\cal{O}}\Big(\frac{\mu_{LS}}{M_t}\Big) \,\bigg] \,. \label{lowscalepole} \end{eqnarray} By adjusting the scale $\mu_{LS}$ in such a way that the RHS of Eq.~(\ref{lowscalepole}) is comparable in size to the RHS of Eq.~(\ref{1Sdef}) the position of the peak in the total cross section can be stabilized. In Ref.~\cite{Beneke4} the ``potential-subtracted'' mass was proposed. It subtracts the dominant infrared-sensitive contribution in the Schr\"odinger equation~(\ref{NNLOSchroedinger}), which is contained in the static potential $V_c$. The subtraction is in fact equal to the RHS of Eq.~(\ref{1SpoleIR2}). Like the low scale running mass, the potential-subtracted mass depends on a cutoff, $\mu_{PS}$. At order $\alpha_s$ (LO) the relation to the pole mass reads~\cite{Beneke4} \begin{eqnarray} m_t^{PS}(\mu_{PS}) - M_t & = & -\,\frac{4}{3}\,\frac{\alpha_s}{\pi}\,\mu_{PS}\, \bigg[\, 1 + {\cal{O}}(\alpha_s) \,\bigg] \,. \label{polesubtractedpole} \end{eqnarray} As for the low scale running mass, the scale $\mu_{PS}$ can be adjusted in such a way that the RHS of Eq.~(\ref{polesubtractedpole}) is comparable in size to the RHS of Eq.~(\ref{1Sdef}). To achieve this, $\mu_{PS}$ has to be chosen of the order of the inverse Bohr radius $\sim M_t\alpha_s$, which is much larger than the scale $\mu_f\ll M_t\alpha_s$ introduced in Eq.~(\ref{1SpoleIR2}). For $\mu_{PS}=\frac{4}{3}\mu_{LS}$ the low scale running and the potential-subtracted mass lead to approximately equivalent results. However, the stabilization of the peak position can be expected to be slightly worse than for the $1S$ mass if $\mu_{PS}$ or $\frac{4}{3}\mu_{LS}$ are not fine-tuned. In addition, the results that could finally be obtained for the $\overline{\mbox{MS}}$ top mass can depend on the value that is chosen for the cutoff scale $\mu_{LS}$ and $\mu_{PS}$. \vspace{1cm} \subsection{Normalization of the Total Cross Section} \label{subsectionnormalization} In the previous subsection we have demonstrated that a proper redefinition of the top quark mass leads to a considerable improvement in the stability of the peak position in the vector-current-induced total cross section $R^v$. However, there have been only marginal changes in the size of the NNLO corrections to the overall normalization of the line-shape. Compared to the NLO normalization of the total vector-current-induced cross section, the NNLO corrections are between $15$ and $25\%$, which is rather large if one recalls that the NNLO corrections are parametrically of order $v^2\sim\alpha_s^2$. In this section we try to find some answers to the question, whether the large NNLO corrections to the normalization of the total vector-current-induced cross section have to be interpreted as a sign that the non-relativistic expansion for the top-antitop cross section close to threshold breaks down. Clearly, this question can only be answered reliably after the complete N${^3}$LO corrections have been determined, which are, unfortunately, beyond the capabilities of present technology. We therefore analyse the NNLO corrections to the normalization of the total cross section with respect to their sensitivity to infrared momenta and carry out a comparison to the one- and two-loop cross section for energies far above the top-antitop threshold, where conventional perturbation theory in $\alpha_s$ is believed to be reliable. We provide arguments that the large NNLO corrections to the normalization are genuine ${\cal{O}}(v^2,\alpha_s^2)$ relativistic corrections, which cannot be removed by changing the definition of $\alpha_s$ or the top quark mass, and that their size does not necessarily indicate a breakdown of the non-relativistic expansion used in this work. As far as a redefinition of the top quark mass is concerned, it is quite obvious that it cannot significantly affect the normalization of the total cross section because the dominant effect in a mass shift is an energy shift of the entire line-shape. Nevertheless, it is quite interesting that the normalization is at all insensitive to the dominant infrared-sensitive terms in the Schr\"odinger equation~(\ref{NNLOSchroedinger})\footnote{ In our case the two issues are in fact connected to each other. But it is important to conceptually separate the issue of a simple energy shift from the more fundamental question of infrared sensitivity. }, which, in the pole mass scheme, would cause the corrections to the peak position to grow factorially at large orders of perturbation theory. To show this let us recall that the total vector-current-induced cross section is proportional to the absorptive part of the Green function, with both arguments evaluated at the origin in configuration space representation: \begin{eqnarray} R^v & \sim & \mbox{Im}\,\sum\hspace{-5mm}\int\limits_n\hspace{2mm}\, \frac{|\Phi_n(0)|^2}{E_n-E-i\,\Gamma_t} \,, \end{eqnarray} where the sum extends over discrete and continuum states with $S$ wave quantum numbers. Thus, for fixed energy the normalization only depends on the wave function. Repeating the steps following Eq.~(\ref{1Spolefull}) we find that the correction to the wave function coming from the dominant infrared-sensitive terms in the Schr\"odinger equation reads \begin{eqnarray} \bigg[\,\delta \Phi_n(0)\,\bigg]^{IR} & = & \bigg[\, \int\frac{d^3\mbox{\boldmath $p$}}{(2\pi)^3} \int\frac{d^3\mbox{\boldmath $q$}}{(2\pi)^3}\, \sum\hspace{-6.5mm}\int\limits_{m\neq n}\, \frac{\Phi_m(0)\,\tilde\Phi_m^*(\mbox{\boldmath $p$})}{E_m-E-i\Gamma_t} \,\delta {\cal{H}}(\mbox{\boldmath $p$},\mbox{\boldmath $q$})\, \tilde\Phi_{n}(\mbox{\boldmath $q$}) \,\bigg]^{IR} \nonumber\\ & \sim & \int\frac{d^3\mbox{\boldmath $p$}}{(2\pi)^3} \int\limits^{|\mbox{\boldmath $q$}|<\mu_f} \frac{d^3\mbox{\boldmath $q$}}{(2\pi)^3}\, \sum\hspace{-6.5mm}\int\limits_{m\neq n}\, \frac{\Phi_m(0)\,\tilde\Phi_m^*(\mbox{\boldmath $p$})}{E_m-E-i\Gamma_t}\, \tilde\Phi_{n}(\mbox{\boldmath $p$}) \,\delta\tilde V_c(\mbox{\boldmath $q$}) \nonumber \\[3mm] & = & 0 \,, \end{eqnarray} i.e. it vanishes because of the orthogonality of the wave functions. Therefore the large corrections in the normalization of the total cross section are not related to an infrared sensitivity of the corrections, in particular at large orders. To demonstrate that this is also the case for the NNLO corrections calculated in this work, we have displayed in Fig.~\ref{figcompareinfrared} the total vector-current-induced cross section $Q_t^2 R^v$ close to threshold for $M_{1S}=\Lambda=175$~GeV, $\mu=30$~GeV, $\alpha_s(M_Z)=0.118$ and $\Gamma_t=1.43$~GeV successively including various NNLO corrections. \begin{figure}[t!] \begin{center} \leavevmode \epsfxsize=4.5cm \epsffile[200 400 400 530]{fig11.ps} \vskip 2.7cm \caption{\label{figcompareinfrared} The total vector-current-induced cross section $Q_t^2 R^v$ close to threshold for $M_{1S}=\Lambda=175$~GeV, $\mu=30$~GeV, $\alpha_s(M_Z)=0.118$ and $\Gamma_t=1.43$~GeV at NLO (dotted curve) and NNLO (solid curve). The dash-dotted curve is NLO including also the NNLO corrections to the Coulomb potential $V_c$, and the dashed line contains, in addition, all Abelian NNLO corrections. The differences between the curves indicates the size of individual NNLO relativistic corrections. } \end{center} \end{figure} The dotted line represents the NLO cross section and the solid line the NNLO one. The dash-dotted line is the NLO cross section including also the NNLO corrections to the Coulomb potential $V_c$; the dashed line contains, in addition, all Abelian NNLO corrections, i.e. those that do not involve the SU(3) group theoretical factor $C_A$. The separation of the NNLO corrections into those coming from the Coulomb potential and from Abelian and non-Abelian relativistic corrections is gauge-invariant. The difference between the dashed and the solid curve represents the corrections of the non-Abelian NNLO effects originating from the potential $V_{\mbox{\tiny NA}}$ and those ${\cal{O}}(\alpha_s^2)$ contributions to the short-distance coefficient $C^v$ that are proportional to $C_A$. From the rather small difference between the dotted and the dash-dotted curves (2--4\%) we see that the large NNLO corrections to the normalization are not related to the corrections in the Coulomb potential. Because a redefinition of the strong coupling would mainly affect the size of the higher-order corrections in the Coulomb potential, we can conclude that using a different scheme for the strong coupling (such as the $V$-scheme~\cite{Brodsky1,Brodsky2}) will not significantly affect the size of the NNLO corrections. The curves plotted in Fig.~\ref{figcompareinfrared} demonstrate that the ${\cal{O}}(20\%)$ NNLO correction to the normalization is a sum of corrections, each of which positive and individually either smaller than or approximately equal to ${\cal{O}}(10\%)$. Although this observation, of course, cannot be taken as a proof that the still unknown N$^3$LO corrections are smaller than the NNLO ones, it indicates that the size of the latter does not necessarily have to be taken as an argument for the non-relativistic expansion to break down for the normalization of the total cross section. An interesting insight into the question of how to interpret the large normalization corrections can also be obtained by comparing the total cross section line-shape, which we have calculated in the threshold regime, with earlier calculations of the total cross section for higher energies, where a resummation of Coulomb singular terms is not yet necessary and perturbation theory in $\alpha_s$ is believed to be reliable.~\cite{Chetyrkin2} We would like to note that it is the large mass of the top quark that allows us to draw conclusions from a comparison of the threshold cross section with the one calculated for higher energies. To illustrate this we recall that our calculation of the threshold cross section is valid if the hierarchy $\alpha_s, v\ll 1$ is satisfied, where the scale of the strong coupling is of the order of the inverse Bohr radius, the kinetic energy, or the top width. This means that the threshold cross section represents a simultaneous expansion in $\alpha_s$ and $v$, where powers of $(\alpha_s/v)$ are resummed to all orders in $\alpha_s$. The high energy cross section, on the other hand, is valid if $\alpha_s\ll v, 1$, where the scale in the strong coupling if of order the top-antitop relative momentum or the centre-of-mass energy. Thus a comparison of the threshold results with the high energy perturbative ones is only sensible if there exists a kinematic regime where both hierarchies are satisfied at the same time, i.e. if $\alpha_s\ll v\ll 1$. In this regime the effects of the resummation of powers of $(\alpha_s/v)$ not contained in the high energy cross section should be small as well as the effects of velocity corrections beyond NNLO, which are not contained in the threshold cross section. Obviously this relation is difficult or impossible to satisfy for bottom or charm quarks, but it is possible for the top quark case. For $\alpha_s(M_t\alpha_s)\sim 0.13$ we can argue that a meaningful comparison between threshold and high energy cross section should be possible for $v\approx$~0.3--0.4, which corresponds to $\sqrt{s}\approx 365$~GeV. \begin{figure}[t!] \begin{center} \leavevmode \epsfxsize=5cm \epsffile[220 360 400 550]{matching.ps} \vskip 3.4cm \caption{\label{figthreshhigh} The vector-current-induced total cross section in the non-relativistic expansion at NLO (lower bunch of threshold curves), NNLO (upper bunch) and in conventional perturbation theory at ${\cal{O}}(\alpha_s)$ (lower bunch of high energy curves) and ${\cal{O}}(\alpha_s^2)$ (upper bunch). The pole mass scheme has been used. The curves have been plotted for $\alpha_s(M_z)=0.118$, $\Gamma_t=1.43$~GeV, $M_t=\Lambda=175$~GeV and $\mu=25$~GeV (dotted lines), $2 (p_0^4+M_t^2\,\Gamma_t^2)^{1/4}$ (solid lines), $175$~GeV (dashed lines) and $\sqrt{q^2}$ (dash-dotted lines). The formulae for the ${\cal{O}}(\alpha_s^2)$ high energy cross section have been taken from Ref.~\cite{Chetyrkin2}. } \end{center} \end{figure} In Fig.~\ref{figthreshhigh} we have plotted the threshold and the high energy cross sections at NLO/NNLO and ${\cal{O}}(\alpha_s)$/${\cal{O}}(\alpha_s^2)$, respectively, for the renormalization scales $\mu=25$~GeV (dotted lines), $2 (p_0^4+M_t^2\,\Gamma_t^2)^{1/4}$ (solid lines), $175$~GeV (dashed lines) and $\sqrt{q^2}$ (dash-dotted lines) for $M_t=\Lambda=175$~GeV, $\alpha_s(M_Z)=0.118$ and $\Gamma_t=1.43$. The lower bunch of threshold curves (characterized by the peak at around $\sqrt{q^2}=348$~GeV) is NLO and the upper bunch NNLO. Likewise, the lower bunch of high energy curves is ${\cal{O}}(\alpha_s)$ and the upper bunch ${\cal{O}}(\alpha_s^2)$. We note that we have not plotted the threshold curves for $\mu=\sqrt{q^2}$ and the high energy curves not for $\mu=25$~GeV, which seems to be a rather unnatural choice for each. The formulae for the ${\cal{O}}(\alpha_s^2)$ high energy cross section have been taken from Ref.~\cite{Chetyrkin2}. For convenience we have plotted the curves in Fig.~\ref{figthreshhigh} in the pole mass scheme. Because the choice of the mass definition does not alter the behaviour of the cross section normalization for energies above the peak position, this choice does not affect the conclusions drawn below. For the threshold (high energy) cross section, we observe that the NNLO (${\cal{O}}(\alpha_s^2)$) corrections decrease for energies further away from the threshold region. However, the ${\cal{O}}(\alpha_s^2)$ corrections to the high energy cross sections are much larger than the NNLO corrections to the threshold cross section at the same centre-of-mass energy. For $\sqrt{s}=\,$360--370~GeV the ${\cal{O}}(\alpha_s^2)$ corrections to the high energy cross section are, for equal choices of renormalization scales, between 10 and 20\% compared to only around $5\%$ for the NNLO corrections to the threshold cross section. We also see a much weaker renormalization scale dependence of the threshold cross sections. The curves show that the resummation of Coulomb singular terms contained in the threshold calculation leads to a considerable stabilization of the cross section determined in conventional perturbation theory in $\alpha_s$ for energies below $\sqrt{s}=365$~GeV. If we believe that conventional perturbation theory is reliable down to energies around $\sqrt{s}=360$~GeV, the results displayed in Fig.~\ref{figthreshhigh} indicate that the non-relativistic expansion does certainly not break down. However, the curves of Fig.~\ref{figthreshhigh} also make it evident that the small renormalization scale dependence of the NLO threshold cross section does certainly not reflect the true size of the remaining theoretical uncertainties at NLO. We believe that $10\%$ should be a fair estimate of the remaining theoretical uncertainties contained in the normalization of the NNLO total cross section close to threshold. As far as the top mass determination at a future electron-positron linear or muon pair collider is concerned, this rather large normalization uncertainty might in fact lead to uncertainties in the determination of the $1S$ mass that are larger than indicated in the previous subsection. This is due to the effects of beamstrahlung and initial state radiation that lead to a smearing of the effective centre-of-mass energy of about 1--2~GeV~\cite{Orange1}. Beamstrahlung and initial state radiation render the visible peak in the total cross section either smaller (at the muon pair collider) or completely invisible (at the linear collider), which makes it possible that the uncertainty in the normalization feeds into larger uncertainties in the determination of $M_{1S}$. It is the task of realistic simulation studies to determine how large this effect is for the various collider and detector designs and to devise optimized strategies to minimize it. If the effects of beamstrahlung and initial state radiation on the top quark mass determination are small, the uncertainty in the normalization will mainly affect the measurement of the strong coupling (see Figs.~\ref{figtotm1S}c and \ref{figalltotalm1S}b). \begin{figure}[t!] \begin{center} \leavevmode \epsfxsize=3.8cm \epsffile[200 400 400 530]{fig13a.ps} \hspace{4.4cm} \epsfxsize=3.8cm \leavevmode \epsffile[200 400 400 530]{fig13b.ps} \vskip 2.7cm \caption{\label{figalltotalm1S} The total cross section $\sigma_{tot}^{\gamma,Z}$, Eq.~(\ref{totalcrossfullQCD}), is plotted in the $1S$ scheme at NNLO for $M_{1S}=\Lambda=175$~GeV, $\Gamma_t=1.43$~GeV and $\alpha=1/125.7$. (a) shows the renormalization scale dependence for $\mu=15$ (solid line), $30$ (dashed line) and $60$~GeV (dotted line), and $\alpha_s(M_Z)=0.118$; (b) shows the dependence on the strong coupling for $\alpha_s(M_Z)=0.113$ (solid line), $0.118$ (dashed line) and $0.123$ (dotted line), and for $\mu=30$~GeV. } \end{center} \end{figure} In Figs.~\ref{figalltotalm1S} the total cross section $\sigma_{tot}^{\gamma,Z}(e^+e^-\to \gamma^*, Z^*\to t\bar t)$, Eq.~(\ref{totalcrossfullQCD}), is plotted at NNLO in the $1S$ scheme for $M_{1S}=\Lambda=175$~GeV, $\Gamma_t=1.43$~GeV and $\alpha=1/125.7$. Figure~\ref{figalltotalm1S}a shows the renormalization scale dependence for $\mu=15$ (solid line), $30$ (dashed line) and $60$~GeV (dotted line), and $\alpha_s(M_Z)=0.118$. Figure~\ref{figalltotalm1S}b displays the dependence on the strong coupling for $\alpha_s(M_Z)=0.113$ (solid line), $0.118$ (dashed line) and $0.123$ (dotted line), and for $\mu=30$~GeV. \vspace{1.5cm} \section{Summary and Conclusions} \label{sectionconclusion} Within the framework of the non-relativistic effective field theories NRQCD and PNRQCD, we have calculated the vector-current-induced total cross section of top-antitop pair production in electron-positron annihilation close to threshold at NNLO in the non-relativistic expansion. The corresponding NNLO QCD relativistic corrections have also been determined for the vector-current-induced top three-momentum distribution. In addition, the axial-vector-current-induced total cross section and the three-momentum distribution have been calculated to fully account for the $Z$-boson contributions in electron-positron annihilation. For the total cross section and the three-momentum distribution, the axial-vector-current-induced contributions are suppressed by $v^2$ with respect to the vector current contributions; they have therefore been determined in leading order in the non-relativistic expansion. The size of the axial-vector-current-induced contributions is smaller than the remaining theoretical uncertainties in the vector-current-induced cross section (for unpolarized electrons and positrons). In contrast with previous literature on the same subject, we have implemented the top quark width by including electroweak corrections into the (P)NRQCD matching conditions of the Lagrangian and the currents. This allows for a straightforward generalization of the Fadin-Khoze prescription ``$E\to E+i \Gamma_t$'' to implement the top quark width at NNLO in the non-relativistic expansion, where $\Gamma_t/M_t$ is counted as order $v^2$. We have shown that at NNLO this cannot be achieved by a simple shift of the centre-of-mass energy to complex values. Our calculations have been carried out using numerical techniques to solve the corresponding integral equations within a cutoff regularization scheme and using analytic methods for the matching procedure. We have addressed the question of large NNLO corrections to the peak position and the normalization of the total vector-current-induced cross section. The position of the peak, which is observable in the total vector-current-induced cross section, can be stabilized if the cross section is expressed in terms of the $1S$ mass, instead of the pole mass. The $1S$ mass, $M_{1S}$, is defined as half the mass of a fictitious ${}^3\!S_1$ toponium ground state for a stable top quark. The $1S$ mass is a short-distance mass and, by construction, reduces to a large extent the dependence of the peak position on theoretical parameters such as the renormalization scale of the strong coupling. We have also shown that the large NNLO corrections to the normalization of the total cross section, of order $20\%$, are genuine NNLO corrections, which cannot be removed by a redefinition of the top quark mass or the strong coupling. The large size of the corrections to the normalization originates from the fact that NNLO relativistic corrections from several sources have the same sign. We believe that the remaining theoretical uncertainties in the normalization are of order $10\%$. If the effects of beamstrahlung and initial state radiation at the $e^+e^-$ linear collider do not lead to a significant cross feed of the uncertainties in the normalization into $M_{1S}$, we expect that an uncertainty in the determination of $M_{1S}$ of less than $200$~MeV will be possible at the linear collider with an integrated luminosity of 50--100~$fb^{-1}$. In order to determine the $\overline{\mbox{MS}}$ top quark mass from the $1S$ mass with the same precision, the knowledge of the full three-loop relation between the pole and the $\overline{\mbox{MS}}$ mass, and a small uncertainty in $\alpha_s(M_Z)$ are crucial. After completion of this work, we received Refs.~\cite{Beneke6,Sumino3,Penin1}. In Ref.~\cite{Beneke6} the total vector-current-induced cross section has been calculated analytically, using the $\overline{\mbox{MS}}$ regularization scheme based on the Schr\"odinger equation~(\ref{NNLOSchroedinger}). The NLO and NNLO corrections have been treated perturbatively, supplemented by a resummation of the energy denominators for the $n=1$ and $n=2$ states in the spectral representation of the Green function. The renormalization scale dependence of the cross section line-shape is considerably larger in Ref.~\cite{Beneke6} than in our work. This might be a consequence of the perturbative treatment of the NLO and NNLO corrections. In addition, a next-to-leading logarithmic resummation of logarithms of the ratio $M_t/\mu$ in the short-distance coefficient $C^v_{\tiny \overline{\mbox{MS}}}$ has been carried out, taking the $\overline{\mbox{MS}}$ cutoff scale $\mu$ of order $M_t v$. The effect of this resummation is around $5\%$ for the normalization of the total cross section. In our cutoff scheme, where the regularization scale is of order $M_t$, the corresponding logarithm is contained in the non-relativistic current correlators. In Ref.~\cite{Sumino3} the vector-current-induced cross total section and the three-momentum distribution have been calculated at NNLO, based on the simplified Schr\"odinger equation~(\ref{NNLOSchroedingersimplified}), which we have discussed critically at the end of Sec.~\ref{sectionwidth}. For the three-momentum distribution the authors of Ref.~\cite{Sumino3} have included further corrections to account for the difference with the results of the correct Schr\"odinger equation~(\ref{NNLOSchroedingergamma}). In Refs.~\cite{Beneke6,Sumino3} the ``potential-subtracted'' mass has been tested in different ways as an alternative mass parameter for the total cross section. As far as the uncertainties in the top mass determination at a future linear collider are concerned, Ref.~\cite{Beneke6} arrives at conclusions similar to ours. In Ref.~\cite{Penin1} the techniques used in Refs.~\cite{Hoang2,Hoang3} have been employed to calculate the total cross section, the angular distribution and the top quark polarization for top quark pair production close to threshold in $e^+e^-$ and $\gamma\gamma$ collisions. The corrections originating from the higher-order contributions in the Coulomb potential have been calculated analytically. In Refs.~\cite{Beneke6,Sumino3,Penin1} the top quark width has been implemented by the replacement rule ``$E\to E+i\Gamma_t$'', where $E$ is the centre-of-mass energy with respect to two times the top mass. \vspace{1.5cm} \section*{Acknowledgement} We thank M.~Beneke for discussions and M.~Beneke, Z.~Ligeti and A.~V.~Manohar for reading the manuscript. The work of A.~H.~H. is in part supported by the EU Fourth Framework Program ``Training and Mobility of Researchers'', Network ``Quantum Chromodynamics and Deep Structure of Elementary Particles'', contract FMRX-CT98-0194 (DG12-MIHT). \vspace{2cm} \begin{appendix} \section{Calculation of the Short-Distance Coefficient $C^v$} \label{appendixshortdistance} In this appendix we present details of the calculation of the short-distance coefficient $C^v$ to order $\alpha_s^2$, assuming that the top quarks are stable ($\Gamma_t=0$). We recall that $C^v$ is the square of the short-distance coefficient $c_1^v$ of the ${}^3\!S_1$ NRQCD current ${\tilde \psi}^\dagger \mbox{\boldmath $\sigma$} \tilde \chi$ (see Eq.~(\ref{vectorcurrentexpansion1})); $c_1^v$ contains those contributions in the vector-current-induced top-antitop production diagrams, which come from loop momenta $p=(p^0,\vec{p})$ with $|\vec{p}|>\Lambda$ for $\sqrt{q^2}=2 M_t$. As explained in Sec.~\ref{sectionregularization}, we have to determine $C^v$ by employing the specific routing convention shown in Fig.~\ref{figroutingladder}. In principle, it would be possible to determine $C^v$ by calculating the diagrams for the vector-current-induced cross section in full QCD restricting the loop momenta such that the spatial components would be larger than $\Lambda$. However, in a cutoff scheme it is more economical to first calculate the vector-current-induced cross section in NRQCD up to order $\alpha_s^2$ and NNLO in the velocity expansion and then to adjust the coefficients of $C^v$ such that the cross section in NRQCD is equal to the cross section in full QCD, likewise calculated to order $\alpha_s^2$ and NNLO in the velocity expansion. The expression of the total vector-current-induced cross section in full QCD at order $\alpha_s^2$ and NNLO in the velocity expansion reads ($a\equiv C_F\alpha_s(\mu)$): \begin{eqnarray} \lefteqn{ R_{\mbox{\tiny 2loop QCD}}^{v,\mbox{\tiny NNLO}} \, = \, N_c\,\bigg\{\,\bigg[\, \frac{3}{2}\,\frac{p_0}{M_t}-\frac{5}{4}\,\frac{p_0^3}{M_t^3} \,\bigg] + \frac{a}{\pi}\,\bigg[\, \frac{3\,\pi^2}{4}-6\,\frac{p_0}{M_t}+ \frac{\pi^2}{2}\,\frac{p_0^2}{M_t^2} \,\bigg] } \nonumber\\[2mm] & & +\, a^2\,\bigg[\, \frac{\pi^2\,M_t}{8\,p_0} - \frac{3}{2}\,\bigg(\, 2 + \frac{1}{8\,C_F} \,\Big(\, \beta_0\,\ln\frac{4\,p_0^2}{\mu_{\rm hard}^2} - a_1 \,\Big) \,\bigg) \nonumber\\[2mm] & & \qquad + \bigg(\, \frac{13\,\pi^2}{48} + \frac{3}{2\,C_F^2}\,\kappa + \frac{3\,\beta_0}{2\,C_F\,\pi^2} \, \ln\frac{M_t^2}{\mu_{\rm hard}^2} - \Big(1 +\frac{3}{2}\, \frac{C_A}{C_F} \Big)\,\ln\Big(\frac{p_0}{M_t}\Big) \,\bigg)\,\frac{p_0}{M_t} \,\bigg] \,\bigg\} \,, \label{RphotonfullQCD} \end{eqnarray} where \begin{eqnarray} \kappa & = & C_F^2\,\bigg[\, \frac{1}{\pi^2}\,\bigg(\, \frac{39}{4}-\zeta_3 \,\bigg) + \frac{4}{3} \ln 2 - \frac{35}{18} \,\bigg] - C_A\,C_F\,\bigg[\, \frac{1}{\pi^2} \,\bigg( \frac{151}{36} + \frac{13}{2} \zeta_3 \,\bigg) + \frac{8}{3} \ln 2 - \frac{179}{72} \,\bigg] \nonumber\\[2mm] & & +\, C_F\,T\,\bigg[\, \frac{4}{9}\,\bigg(\, \frac{11}{\pi^2} - 1\,\bigg) \,\bigg] + C_F\,T\,n_l\,\bigg[\, \frac{11}{9\,\pi^2} \,\bigg] \,. \label{kappadef} \end{eqnarray} The Born and ${\cal{O}}(\alpha_s)$~\cite{Kallensabry1} contributions are standard. At order $\alpha_s^2$ the contributions in Eq.~(\ref{RphotonfullQCD}) that are proportional to $C_F^2$, $C_A C_F$, $C_F T n_l$ and $C_F T$ have been calculated in~\cite{Hoang4}, \cite{Melnikov4}, \cite{HoangKT1,Voloshin2} and \cite{HoangKT1,Karshenboim1}, respectively. (See also Refs.~\cite{Chetyrkin2,Chetyrkin3}.) To determine the corresponding total vector-current-induced cross section in NRQCD, we have to calculate the absorptive part of the correlator diagrams depicted in Figs.~\ref{fignonrelcurrentborn}, \ref{fignonrelcurrent1loop} and \ref{fignonrelcurrent2loop}. The various symbols are defined in Fig.~\ref{fignonrealcurrentsymbols}. We emphasize that we neglect multiple insertions of NNLO contributions. The results for the absorptive parts of the individual diagrams read ($a\equiv C_F \alpha_s$, $D(\mbox{\boldmath $k$})\equiv M_t/(\mbox{\boldmath $k$}^2-p_0^2-i\epsilon)$): \begin{figure}[t] \begin{center} \leavevmode \epsfxsize=1.7cm \epsffile[220 410 420 540]{curcor0n1.ps} \hspace{2.2cm} \leavevmode \epsfxsize=1.7cm \epsffile[220 410 420 540]{curcor0n2.ps} \hspace{2.2cm} \leavevmode \epsfxsize=1.7cm \epsffile[220 410 420 540]{curcor0n3.ps} \vskip 1.5cm \caption{\label{fignonrelcurrentborn} Graphical representation of the NRQCD vector-current correlators diagrams needed to determine the non-relativistic vector-current-induced cross section at the Born level and NNLO in the non-relativistic expansion. } \end{center} \end{figure} \begin{figure}[tb] \begin{center} \leavevmode \epsfxsize=1.7cm \epsffile[220 410 420 540]{curcor1n1.ps} \hspace{2.2cm} \leavevmode \epsfxsize=1.7cm \epsffile[220 410 420 540]{curcor1n2.ps} \hspace{2.2cm} \leavevmode \epsfxsize=1.7cm \epsffile[220 410 420 540]{curcor1n3.ps} \hspace{2.2cm} \leavevmode \epsfxsize=1.7cm \epsffile[220 410 420 540]{curcor1n4.ps} \vskip 1.5cm \caption{\label{fignonrelcurrent1loop} Graphical representation of the NRQCD vector-current correlators diagrams needed to determine the non-relativistic vector-current-induced cross section at ${\cal{O}}(\alpha_s)$ and NNLO in the non-relativistic expansion. } \end{center} \end{figure} \begin{figure}[htb] \begin{center} \leavevmode \epsfxsize=1.7cm \epsffile[220 410 420 540]{curcor2n1.ps} \hspace{2.2cm} \leavevmode \epsfxsize=1.7cm \epsffile[220 410 420 540]{curcor2n2.ps} \hspace{2.2cm} \leavevmode \epsfxsize=1.7cm \epsffile[220 410 420 540]{curcor2n3.ps} \\[1.3cm] \leavevmode \epsfxsize=1.7cm \epsffile[220 410 420 540]{curcor2n4.ps} \hspace{2.2cm} \leavevmode \epsfxsize=1.7cm \epsffile[220 410 420 540]{curcor2n5.ps} \hspace{2.2cm} \leavevmode \epsfxsize=1.7cm \epsffile[220 410 420 540]{curcor2n6.ps} \\[1.3cm] \leavevmode \epsfxsize=1.7cm \epsffile[220 410 420 540]{curcor2n7.ps} \hspace{2.2cm} \leavevmode \epsfxsize=1.7cm \epsffile[220 410 420 540]{curcor2n8.ps} \hspace{2.2cm} \leavevmode \epsfxsize=1.7cm \epsffile[220 410 420 540]{curcor2n9.ps} \vskip 1.5cm \caption{\label{fignonrelcurrent2loop} Graphical representation of the NRQCD vector-current correlators diagrams needed to determine the non-relativistic vector-current-induced cross section at ${\cal{O}}(\alpha_s^2)$ and NNLO in the non-relativistic expansion. } \end{center} \end{figure} \begin{figure}[t] \begin{center} \leavevmode \epsfxsize=2.3cm \epsffile[220 410 420 540]{interaction1.ps} \hspace{1cm} \leavevmode \epsfxsize=2.3cm \epsffile[220 410 420 540]{interaction2.ps} \hspace{1cm} \leavevmode \epsfxsize=2.3cm \epsffile[220 410 420 540]{interaction3.ps} \hspace{1cm} \leavevmode \epsfxsize=2.3cm \epsffile[220 410 420 540]{interaction4.ps}\\[1.5cm] $V_c^{\tiny \mbox{LO}}$ \mbox{\hspace{2.65cm}} $V_c^{\tiny \mbox{NLO}}$ \mbox{\hspace{2.55cm}} $V_c^{\tiny \mbox{NNLO}}$ \mbox{\hspace{2.55cm}} $V_{\mbox{\tiny BF}}$ \mbox{\hspace{1mm}}\\[.2cm] \hspace{1cm} \leavevmode \epsfxsize=2.3cm \epsffile[220 410 420 540]{interaction6.ps} \hspace{1cm} \leavevmode \epsfxsize=2.3cm \epsffile[220 410 420 540]{interaction5.ps} \hspace{1cm} \leavevmode \epsfxsize=2.3cm \epsffile[220 410 420 540]{interaction7.ps}\\[1.5cm] \mbox{\hspace{2.cm}} $V_{\mbox{\tiny NA}}$ \mbox{\hspace{2.65cm}} $\delta H_{\mbox{\tiny kin}}$ \mbox{\hspace{1.55cm}} ${\tilde \psi}^\dagger \sigma_i (\mbox{$-\frac{i}{2}$} \stackrel{\leftrightarrow}{\mbox{\boldmath $D$}})^2 \tilde \chi$ \mbox{\hspace{.25cm}} \vskip .2cm \caption{\label{fignonrealcurrentsymbols} Symbols describing the interactions potentials $V_c^{\tiny \mbox{LO}}$, $V_c^{\tiny \mbox{NLO}}$, $V_c^{\tiny \mbox{NNLO}}$, $V_{\mbox{\tiny BF}}$ and $V_{\mbox{\tiny NA}}$ and the kinetic energy correction $\delta H_{\mbox{\tiny kin}} = (p_0^4-\vec k^4)/(4 M_t^3)$. $V_c^{\tiny \mbox{LO}}$, $V_c^{\tiny \mbox{NLO}}$ and $V_c^{\tiny \mbox{NNLO}}$ refer to the Born, one-loop and two-loop contributions to the Coulomb potential presented in Eq.~(\ref{NNLOCoulomb}).} \end{center} \end{figure} \begin{eqnarray} I_1^{(0)} & = & \mbox{Im}\,\bigg[\, \int\frac{d^3 \mbox{\boldmath $k$}}{(2\pi)^3}\, D(\mbox{\boldmath $k$}) \,\bigg] \, = \, \frac{M_t^2}{4 \pi}\,\frac{p_0}{M_t} \,, \\[2mm] I_2^{(0)} & = & \mbox{Im}\,\bigg[\, 2\,\int\frac{d^3 \mbox{\boldmath $k$}}{(2\pi)^3}\, \bigg(-\frac{\mbox{\boldmath $k$}^2}{6\,M_t^2}\,\bigg)\, D(\mbox{\boldmath $k$}) \,\bigg] \, = \, - \frac{M_t^2}{4 \pi}\,\frac{p_0^3}{3\,M_t^3} \,, \\[2mm] I_3^{(0)} & = & \mbox{Im}\,\bigg[\, \int\frac{d^3 \mbox{\boldmath $k$}}{(2\pi)^3}\, \bigg(\,\frac{\mbox{\boldmath $k$}^2+p_0^2}{4\,M_t^2}\,\bigg)\, D(\mbox{\boldmath $k$}) \,\bigg] \, = \, \frac{M_t^2}{4 \pi}\,\frac{p_0^3}{2\,M_t^3} \,, \\[5mm] I_1^{(1)} & = & \mbox{Im}\,\bigg[\, \int\frac{d^3 \mbox{\boldmath $k_1$}}{(2\pi)^3}\, \int\frac{d^3 \mbox{\boldmath $k_2$}}{(2\pi)^3}\, D(\mbox{\boldmath $k_1$})\, \frac{4\pi a}{(\mbox{\boldmath $k_1$}-\mbox{\boldmath $k_2$})^2}\, D(\mbox{\boldmath $k_2$}) \,\bigg] \, = \, \frac{a\,M_t^2}{4\, \pi^2}\,\bigg[\, \frac{\pi^2}{2}-\frac{4\,p_0}{\Lambda} + {\cal{O}}\bigg( \frac{p_0^3}{M_t^3} \bigg) \,\bigg] \,, \\[2mm] I_2^{(1)} & = & \mbox{Im}\,\bigg[\,2\, \int\frac{d^3 \mbox{\boldmath $k_1$}}{(2\pi)^3}\, \int\frac{d^3 \mbox{\boldmath $k_2$}}{(2\pi)^3}\, \bigg(-\frac{\mbox{\boldmath $k_1$}^2}{6\,M_t^2}\,\bigg)\, D(\mbox{\boldmath $k_1$})\, \frac{4\pi a}{(\mbox{\boldmath $k_1$}-\mbox{\boldmath $k_2$})^2}\, D(\mbox{\boldmath $k_2$}) \,\bigg] \nonumber \\[2mm] & = & -\, \frac{a\,M_t^2}{4\, \pi^2}\,\bigg[\, \frac{2\,\Lambda\,p_0}{3\,M_t^2} + \frac{p_0^2\,\pi^2}{6\,M_t^2} + {\cal{O}}\bigg( \frac{p_0^3}{M_t^3} \bigg) \,\bigg] \,, \\[2mm] I_3^{(1)} & = & \mbox{Im}\,\bigg[\,2\, \int\frac{d^3 \mbox{\boldmath $k_1$}}{(2\pi)^3}\, \int\frac{d^3 \mbox{\boldmath $k_2$}}{(2\pi)^3}\, \bigg(\,\frac{\mbox{\boldmath $k_1$}^2+p_0^2}{4\,M_t^2}\,\bigg)\, D(\mbox{\boldmath $k_1$})\, \frac{4\pi a}{(\mbox{\boldmath $k_1$}-\mbox{\boldmath $k_2$})^2}\, D(\mbox{\boldmath $k_2$}) \,\bigg] \nonumber \\[2mm] & = & \frac{a\,M_t^2}{4\, \pi^2}\,\bigg[\, \frac{\Lambda\,p_0}{M_t^2} + \frac{p_0^2\,\pi^2}{2\,M_t^2} + {\cal{O}}\bigg( \frac{p_0^3}{M_t^3} \bigg) \,\bigg] \,, \\[2mm] I_4^{(1)} & = & \mbox{Im}\,\bigg[\, \int\frac{d^3 \mbox{\boldmath $k_1$}}{(2\pi)^3}\, \int\frac{d^3 \mbox{\boldmath $k_2$}}{(2\pi)^3}\, D(\mbox{\boldmath $k_1$})\, \bigg(\, 2\,\frac{\pi\,a}{M_t^2}\, \frac{\mbox{\boldmath $k_1$}^2+{\mbox{\boldmath $k_2$}}^2} {(\mbox{\boldmath $k_1$}-\mbox{\boldmath $k_2$})^2} - \frac{11}{3}\,\frac{\pi\,a}{M_t^2}\, \bigg)\, D(\mbox{\boldmath $k_2$}) \,\bigg] \nonumber \\[2mm] & = & \frac{a\,M_t^2}{4\, \pi^2}\,\bigg[\, - \frac{5\,\Lambda\,p_0}{3\,M_t^2} + \frac{p_0^2\,\pi^2}{2\,M_t^2} + {\cal{O}}\bigg( \frac{p_0^3}{M_t^3} \bigg) \,\bigg] \,, \\[5mm] I_1^{(2)} & = & \mbox{Im}\,\bigg[\, \int\frac{d^3 \mbox{\boldmath $k_1$}}{(2\pi)^3}\, \int\frac{d^3 \mbox{\boldmath $k_2$}}{(2\pi)^3}\, \int\frac{d^3 \mbox{\boldmath $k_3$}}{(2\pi)^3}\, D(\mbox{\boldmath $k_1$})\, \frac{4\pi a}{(\mbox{\boldmath $k_1$}-\mbox{\boldmath $k_2$})^2}\, D(\mbox{\boldmath $k_2$})\, \frac{4\pi a}{(\mbox{\boldmath $k_2$}-\mbox{\boldmath $k_3$})^2}\, D(\mbox{\boldmath $k_3$}) \,\bigg] \nonumber \\[2mm] & = & \frac{a^2\,M_t^2}{4\, \pi^3}\,\bigg[\, \frac{M_t\,\pi^4}{12\,p_0}-\frac{2\,M_t\,\pi^2}{\Lambda} + \frac{M_t\,p_0\,(12-\pi^2)}{2\,\Lambda^2} + {\cal{O}}\bigg( \frac{p_0^2}{M_t^2} \bigg) \,\bigg] \,, \\[2mm] I_2^{(2)} & = & \mbox{Im}\,\bigg[\,2\, \int\frac{d^3 \mbox{\boldmath $k_1$}}{(2\pi)^3}\, \int\frac{d^3 \mbox{\boldmath $k_2$}}{(2\pi)^3}\, \int\frac{d^3 \mbox{\boldmath $k_3$}}{(2\pi)^3}\, \bigg(-\frac{\mbox{\boldmath $k_1$}^2}{6\,M_t^2}\,\bigg)\, D(\mbox{\boldmath $k_1$})\, \frac{4\pi a}{(\mbox{\boldmath $k_1$}-\mbox{\boldmath $k_2$})^2}\, D(\mbox{\boldmath $k_2$})\, \frac{4\pi a}{(\mbox{\boldmath $k_2$}-\mbox{\boldmath $k_3$})^2}\, D(\mbox{\boldmath $k_3$}) \,\bigg] \nonumber \\[2mm] & = & \, - \, \frac{a^2\,M_t^2}{4\, \pi^3}\,\bigg[\, \frac{\Lambda\,\pi^2}{3\,M_t} - \frac{p_0\,(84+3\,\pi^2-\pi^4)}{36\,M_t} + {\cal{O}}\bigg( \frac{p_0^2}{M_t^2} \bigg) \,\bigg] \,, \\[2mm] I_3^{(2)} & = & \mbox{Im}\,\bigg[\,2\, \int\frac{d^3 \mbox{\boldmath $k_1$}}{(2\pi)^3}\, \int\frac{d^3 \mbox{\boldmath $k_2$}}{(2\pi)^3}\, \int\frac{d^3 \mbox{\boldmath $k_3$}}{(2\pi)^3}\, \bigg(\,\frac{\mbox{\boldmath $k_1$}^2+p_0^2}{4\,M_t^2}\,\bigg)\, D(\mbox{\boldmath $k_1$})\, \frac{4\pi a}{(\mbox{\boldmath $k_1$}-\mbox{\boldmath $k_2$})^2}\, D(\mbox{\boldmath $k_2$})\, \frac{4\pi a}{(\mbox{\boldmath $k_2$}-\mbox{\boldmath $k_3$})^2}\, D(\mbox{\boldmath $k_3$}) \,\bigg] \nonumber \\[2mm] & = & \, \frac{a^2\,M_t^2}{4\, \pi^3}\,\bigg[\, \frac{\Lambda\,\pi^2}{2\,M_t} - \frac{p_0\,(84+3\,\pi^2-2\,\pi^4)}{24\,M_t} + {\cal{O}}\bigg( \frac{p_0^2}{M_t^2} \bigg) \,\bigg] \,, \\[2mm] I_4^{(2)} & = & \mbox{Im}\,\bigg[\, \int\frac{d^3 \mbox{\boldmath $k_1$}}{(2\pi)^3}\, \int\frac{d^3 \mbox{\boldmath $k_2$}}{(2\pi)^3}\, \int\frac{d^3 \mbox{\boldmath $k_3$}}{(2\pi)^3}\, D(\mbox{\boldmath $k_1$})\, \frac{4\pi a}{(\mbox{\boldmath $k_1$}-\mbox{\boldmath $k_2$})^2}\, \bigg(\,\frac{\mbox{\boldmath $k_2$}^2+p_0^2}{4\,M_t^2}\,\bigg)\, D(\mbox{\boldmath $k_2$})\, \frac{4\pi a}{(\mbox{\boldmath $k_2$}-\mbox{\boldmath $k_3$})^2}\, D(\mbox{\boldmath $k_3$}) \,\bigg] \nonumber \\[2mm] & = & \, \frac{a^2\,M_t^2}{4\, \pi^3}\, \bigg[\, \frac{p_0}{24\,M_t}\,\bigg(\, 12\,\pi^2 + \pi^4- 42\,\zeta_3 - 12\,\pi^2\,\ln\Big(\frac{2\,p_0}{\Lambda}\Big) \,\bigg) + {\cal{O}}\bigg( \frac{p_0^2}{M_t^2} \bigg) \,\bigg] \,, \\[2mm] I_5^{(2)} & = & \mbox{Im}\,\bigg[\,2\, \int\frac{d^3 \mbox{\boldmath $k_1$}}{(2\pi)^3}\, \int\frac{d^3 \mbox{\boldmath $k_2$}}{(2\pi)^3}\, \int\frac{d^3 \mbox{\boldmath $k_3$}}{(2\pi)^3}\, D(\mbox{\boldmath $k_1$})\, \bigg(\, 2\,\frac{\pi\,a}{M_t^2}\, \frac{\mbox{\boldmath $k_1$}^2+{\mbox{\boldmath $k_2$}}^2} {(\mbox{\boldmath $k_1$}-\mbox{\boldmath $k_2$})^2} - \frac{11}{3}\,\frac{\pi\,a}{M_t^2}\, \bigg)\, \nonumber \\[2mm] & & \mbox{\hspace{3cm}} \times\, D(\mbox{\boldmath $k_2$})\, \frac{4\pi a}{(\mbox{\boldmath $k_2$}-\mbox{\boldmath $k_3$})^2}\, D(\mbox{\boldmath $k_3$}) \,\bigg] \nonumber \\[2mm] & = & \frac{a^2\,M_t^2}{4\, \pi^3}\,\bigg[\, -\,\frac{5\,\Lambda\,\pi^2}{6\,M_t} + \frac{p_0}{12\,M_t}\, \bigg(92+21\,\pi^2 + 2\,\pi^4 - 7\,\zeta_3 - 2\,\pi^2\,\ln\Big(\frac{2\,p_0}{\Lambda}\Big)\,\bigg) + {\cal{O}}\bigg( \frac{p_0^2}{M_t^2} \bigg) \,\bigg] \,, \\[2mm] I_6^{(2)} & = & \mbox{Im}\,\bigg[\, \int\frac{d^3 \mbox{\boldmath $k_1$}}{(2\pi)^3}\, \int\frac{d^3 \mbox{\boldmath $k_2$}}{(2\pi)^3}\, D(\mbox{\boldmath $k_1$})\, \bigg(\, \frac{C_A}{C_F}\,\frac{\pi^2\,a^2} {M_t\,|\mbox{\boldmath $k_1$}-\mbox{\boldmath $k_2$}|} \bigg)\, D(\mbox{\boldmath $k_2$}) \,\bigg] \nonumber \\[2mm] & = & \,-\, \frac{a^2\,M_t^2}{4\, \pi^3}\,\bigg[\, \frac{C_A\,p_0\,\pi^2}{C_F\,M_t}\, \bigg(\,-1 + \ln\Big(\frac{2\,p_0}{\Lambda}\Big)\,\bigg) + {\cal{O}}\bigg( \frac{p_0^2}{M_t^2} \bigg) \,\bigg] \,, \\[2mm] I_7^{(2)} & = & \mbox{Im}\,\bigg[\, \int\frac{d^3 \mbox{\boldmath $k_1$}}{(2\pi)^3}\, \int\frac{d^3 \mbox{\boldmath $k_2$}}{(2\pi)^3}\, D(\mbox{\boldmath $k_1$})\, \frac{4\pi a}{(\mbox{\boldmath $k_2$}-\mbox{\boldmath $k_3$})^2}\, \frac{a}{4\,C_F\,\pi} \bigg(\, -\beta_0\, \ln\Big(\frac{(\mbox{\boldmath $k_1$}-\mbox{\boldmath $k_2$})^2}{\mu^2}\Big) + a_1 \bigg)\, D(\mbox{\boldmath $k_2$}) \,\bigg] \nonumber \\[2mm] & = & \frac{a^2\,M_t^2}{4\,C_F\, \pi^3}\,\bigg[\, \beta_0\,\bigg(\, -\frac{\pi^2}{4}\,\ln\Big(\frac{2\,p_0}{\mu}\Big) +\frac{2\,p_0}{\Lambda}\, \bigg(1 + \ln\Big(\frac{\Lambda}{\mu}\Big)\bigg) \,\bigg) + a_1\,\bigg(\, \frac{\pi^2}{8} -\frac{p_0}{\Lambda} \,\bigg) + {\cal{O}}\bigg( \frac{p_0^2}{M_t^2} \bigg) \,\bigg] \,, \\[2mm] I_8^{(2)} & = & \mbox{Im}\,\bigg[\,2 \int\frac{d^3 \mbox{\boldmath $k_1$}}{(2\pi)^3}\, \int\frac{d^3 \mbox{\boldmath $k_2$}}{(2\pi)^3}\, \bigg(-\frac{\mbox{\boldmath $k_1$}^2}{6\,M_t^2}\,\bigg)\, \nonumber\\ & & \mbox{\hspace{3cm}} \times\, D(\mbox{\boldmath $k_1$})\, \frac{4\pi a}{(\mbox{\boldmath $k_2$}-\mbox{\boldmath $k_3$})^2}\, \frac{a}{4\,C_F\,\pi} \bigg(\, -\beta_0\, \ln\Big(\frac{(\mbox{\boldmath $k_1$}-\mbox{\boldmath $k_2$})^2}{\mu^2}\Big) + a_1 \bigg)\, D(\mbox{\boldmath $k_2$}) \,\bigg] \nonumber \\[2mm] & = & \frac{a^2\,M_t^2}{4\,C_F\, \pi^3}\,\bigg[\, \beta_0\,\bigg(\, \frac{\Lambda\,p_0}{3\,M_t^2}\, \bigg(-1 + \ln\Big(\frac{\Lambda}{\mu}\Big)\,\bigg) \,\bigg) - a_1\, \frac{\Lambda\,p_0}{6\,M_t^2} + {\cal{O}}\bigg( \frac{p_0^2}{M_t^2} \,\bigg) \,\bigg] \,, \\[2mm] I_9^{(2)} & = & \mbox{Im}\,\bigg[\,2 \int\frac{d^3 \mbox{\boldmath $k_1$}}{(2\pi)^3}\, \int\frac{d^3 \mbox{\boldmath $k_2$}}{(2\pi)^3}\, \bigg(\,\frac{\mbox{\boldmath $k_1$}^2+p_0^2}{4\,M_t^2}\,\bigg)\, \nonumber\\ & & \mbox{\hspace{3cm}} \times\, D(\mbox{\boldmath $k_1$})\, \frac{4\pi a}{(\mbox{\boldmath $k_2$}-\mbox{\boldmath $k_3$})^2}\, \frac{a}{4\,C_F\,\pi} \bigg(\, -\beta_0\, \ln\Big(\frac{(\mbox{\boldmath $k_1$}-\mbox{\boldmath $k_2$})^2}{\mu^2}\Big) + a_1 \bigg)\, D(\mbox{\boldmath $k_2$}) \,\bigg] \nonumber \\[2mm] & = & \frac{a^2\,M_t^2}{4\,C_F\, \pi^3}\,\bigg[\, \beta_0\,\bigg(\, \frac{\Lambda\,p_0}{2\,M_t^2}\, \bigg(1 - \ln\Big(\frac{\Lambda}{\mu}\Big)\bigg) \,\bigg) + a_1\, \frac{\Lambda\,p_0}{4\,M_t^2} + {\cal{O}}\bigg( \frac{p_0^2}{M_t^2} \bigg) \,\bigg] \,, \end{eqnarray} where the upper index of the functions $I_j^{(i)}$ corresponds to the power of the strong coupling of the diagrams and the lower index to the numeration given in Figs.~\ref{fignonrelcurrentborn}, \ref{fignonrelcurrent1loop} and \ref{fignonrelcurrent2loop}. Combinatorial factors are taken into account. We note that the results above have been expanded in $p_0/M_t, p_0/\Lambda \ll 1$; no condition has been assumed for the ratio $\Lambda/M_t$. Summing all terms leads to the total vector-current-induced cross section in NRQCD: \begin{eqnarray} R_{\mbox{\tiny NNLO}}^{v,\mbox{\tiny thr}} & = & \frac{6\,\pi\,N_c}{M_t^2\,(1+\frac{p_0^2}{M_t^2})}\,C^v\, \bigg[\, \sum\limits_{i=1}^{3}\,I^{(0)}_i + \sum\limits_{i=1}^{4}\,I^{(1)}_i + \sum\limits_{i=1}^{9}\,I^{(2)}_i \,\bigg] \label{RphotonfullNRQCD} \end{eqnarray} The short-distance coefficient $C^v$ has to be chosen such that the RHS of Eqs.~(\ref{RphotonfullQCD}) and (\ref{RphotonfullNRQCD}) are equal for all terms up to order $\alpha_s^2$ and NNLO in the non-relativistic expansion. The result reads \begin{eqnarray} C^v(\Lambda,\mu) \, = \, 1 \, + \, c^v_{\mbox{\tiny NLO}}(\Lambda,\mu) \, + \, c^v_{\mbox{\tiny NNLO}}(\Lambda,\mu) \,, \label{CVNNLOshortdistance} \end{eqnarray} where \begin{eqnarray} c^v_{\mbox{\tiny NLO}}(\Lambda,\mu) & = & \frac{4\,a}{\pi}\, \bigg[\, -1 + \frac{M_t}{\Lambda} \,\bigg] \,, \label{cvNLOshortdistance} \\ c^v_{\mbox{\tiny NNLO}}(\Lambda,\mu) & = & \frac{4\,a\,\Lambda}{3\,\pi\,M_t} + \frac{a^2}{\pi^2}\,\bigg[\, \frac{\beta_0}{C_F}\,\bigg(\, - \frac{\Lambda^2 + 12\,M_t^2}{6\,\Lambda\,M_t} + \frac{\Lambda^2 - 12\,M_t^2} {6\,\Lambda\,M_t}\,\ln\Big(\frac{\Lambda}{\mu}\Big) + 2\,\ln\Big(\frac{M_t}{\mu}\Big) \,\bigg) \nonumber \\[2mm] & & - \frac{a_1}{C_F}\,\frac{\Lambda^2 - 12\,M_t^2}{12\,\Lambda\,M_t} + \pi^2\,\bigg(\,\frac{2}{3}+\frac{C_A}{C_F}\,\bigg)\, \ln\Big(\frac{2\,M_t}{\Lambda}\,\Big) + \frac{\pi^2\,\kappa}{C_F^2} \nonumber \\[2mm] & & + \frac{16\,\Lambda^2}{9\,M_t^2} - \frac{16\,\Lambda}{3\,M_t} - \frac{16\,M_t}{\Lambda} + \frac{M_t^2\,(20 + \pi^2)}{2\,\Lambda^2} - \frac{53\,\pi^2}{24} - \frac{C_A\,\pi^2}{C_F} + \frac{25}{6} + \frac{7}{3}\zeta_3 \,\bigg] \,. \label{cvNNLOshortdistance} \end{eqnarray} In Eq.~(\ref{CVNNLOshortdistance}) we have displayed the NLO and NNLO short-distance contributions separately. We note that the NLO short-distance contributions in Eq.~(\ref{cvNLOshortdistance}) differ from the commonly quoted one by the term $4\frac{C_F\alpha_s}{\pi}\frac{M_t}{\Lambda}$. This is a consequence of our cutoff regularization scheme, which excludes loop momenta with spatial components larger than $\Lambda$, even if the corresponding integration is UV-convergent. We also point out that the NNLO short-distance contribution in Eq.~(\ref{cvNNLOshortdistance}) contains the term $\frac{4\alpha_s}{3\pi}\frac{\Lambda}{M_t}$, which is of order $\alpha_s$ only. This term is a manifestation of the power-counting breaking effects discussed in Sec.~\ref{sectionregularization}. The term exists because it subtracts the power-counting breaking terms originating from the linear UV-divergent behaviour of the Breit-Fermi potential, $V_{\mbox{\tiny BF}}$, the kinetic energy correction, $(p_0^4-\mbox{\boldmath $k$}^4)/(4 M_t^3)$, and the dimension-5 NRQCD vector-current in the non-relativistic current correlators. Thus it is important to consider this term as NNLO. It is a conspicuous fact that the NLO short-distance coefficient $c^v_{\mbox{\tiny NLO}}$ vanishes for the choice $\Lambda=M_t$. We emphasize, however, that this cancellation is purely accidental. Nevertheless, comparing the short-distance constant $C^v$ calculated in our cutoff regularization scheme with the corresponding coefficient obtained in the $\overline{\mbox{MS}}$ scheme~\cite{Melnikov4,Beneke2} \begin{eqnarray} C^v_{\tiny \overline{\mbox{MS}}}(\mu) \, = \, 1 \, + \, c^v_{{\tiny \overline{\mbox{MS}}, \tiny NLO}}(\Lambda,\mu) \, + \, c^v_{{\tiny \overline{\mbox{MS}}, \tiny NNLO}}(\Lambda,\mu) \,, \label{CVNNLOshortdistanceMSbar} \end{eqnarray} where ($\alpha_s\equiv\alpha_s(\mu)$) \begin{eqnarray} c^v_{\tiny \overline{\mbox{MS}}, \mbox{\tiny NLO}}(\mu) & = & -4\,C_F\,\frac{\alpha_s}{\pi} \,, \label{cvNLOshortdistanceMSbar} \\ c^v_{\tiny \overline{\mbox{MS}}, \mbox{\tiny NNLO}}(\mu) & = & \frac{\alpha_s^2}{\pi^2}\,\bigg[\, C_F^2\,\bigg(\,\frac{39}{4} - \frac{79\,\pi^2}{18} + 2\,\pi^2\,\ln 2 + \frac{\pi^2}{3}\,\ln\Big(\frac{M_t^2}{\mu^2}\Big) - \zeta_3 \,\bigg) \nonumber \\[2mm] & & + C_A\,C_F\,\bigg( -\frac{151}{36} + \frac{89\,\pi^2}{72} - \frac{5\,\pi^2}{3}\,\ln 2 + \frac{\pi^2}{2}\,\ln\Big(\frac{M_t^2}{\mu^2}\Big) - \frac{13}{2}\,\zeta_3 \,\bigg) \nonumber \\[2mm] & & + C_F\,T\,\bigg(\,\frac{44}{9} - \frac{4\,\pi^2}{9}\,\bigg) + C_F\,T\,n_f\,\frac{11}{9} \,\bigg] \,, \label{cvNNLOshortdistanceMSbar} \end{eqnarray} we find that the perturbative corrections are in general smaller in the cutoff scheme. \begin{table}[t] \vskip 7mm \begin{center} \begin{tabular}{|c||c|c|c||c||c|c|c||c|} \hline $\mu [\mbox{GeV}]$ & \multicolumn{3}{|c||} {$c^v_{\mbox{\tiny NLO}}(\Lambda [\mbox{GeV}])$} & $c^v_{\tiny \overline{\mbox{MS}}, \mbox{\tiny NLO}}$ & \multicolumn{3}{|c||} {$c^v_{\mbox{\tiny NNLO}}(\Lambda [\mbox{GeV}])$} & $c^v_{\tiny \overline{\mbox{MS}}, \mbox{\tiny NNLO}}$ \\ \hline & $90$ & $175$ & $350$ & & $90$ & $175$ & $350$ & \\ \hline\hline $15$ & $0.261$ & $0$ & $-0.138$ & $-0.276$ & $-0.018$ & $-0.101$ & $-0.040$ & $0.128$ \\\hline $30$ & $0.228$ & $0$ & $-0.120$ & $-0.241$ & $0.018$ & $-0.069$ & $-0.029$ & $0.025$ \\\hline $60$ & $0.202$ & $0$ & $-0.107$ & $-0.214$ & $0.039$ & $-0.048$ & $-0.022$ & $0.037$ \\\hline $175$ & $0.172$ & $0$ & $-0.091$ & $-0.182$ & $0.056$ & $-0.028$ & $-0.015$ & $-0.091$ \\ \hline\hline \end{tabular} \caption{\label{tabshortdistance} The NLO and NNLO contributions to the short-distance coefficient $C^v$ in our cutoff scheme and in the $\overline{\mbox{MS}}$ scheme for various choices of the cutoff $\Lambda$ and the renormalization scale $\mu$. We have chosen $\alpha_s(M_z)=0.118$, and two-loop running for the strong coupling has been employed. } \end{center} \vskip 3mm \end{table} In Table~\ref{tabshortdistance} we have displayed the NLO and NNLO short-distance corrections for exemplary choices of the renormalization scale $\mu$ and the cutoff $\Lambda$ for $\alpha_s(M_Z)=0.118$ and using two-loop running for the strong coupling. We are not aware of any principle reason why the short-distance corrections should be, in general, better convergent in our cutoff scheme than when using the $\overline{\mbox{MS}}$ regularization. We finally note that the potentially large logarithmic term $\alpha_s^2 C_A C_F \ln(M_t/\mu)$ in $C^v_{\tiny \overline{\mbox{MS}}}$, which corresponds to an anomalous dimension of the dimension-3 NRQCD vector-current ${\tilde \psi}^\dagger \mbox{\boldmath $\sigma$} \tilde \chi$, does not exist in $C^v$, since in our cutoff scheme the corresponding logarithmic divergence in the NRQCD diagrams is cut off at the scale $\Lambda\sim M_t$ rather than $\mu\sim M_t v$, as in the $\overline{\mbox{MS}}$ scheme. However, we emphasize that the absence of this logarithmic term in $C^v$ is traded for the existence of logarithms of the ratio $2p_0/\Lambda$ in the non-relativistic correlator, which are not present in the $\overline{\mbox{MS}}$ scheme. The logarithms of $\Lambda/\mu$ and $M_t/\mu$ in $c^v_{\mbox{\tiny NNLO}}$, Eq.~(\ref{cvNNLOshortdistance}), originate from the running of the strong coupling and are not related to an anomalous dimension. \end{appendix} \vspace{1.5cm} \sloppy \raggedright \def\app#1#2#3{{\it Act. Phys. Pol. }{\bf B #1} (#2) #3} \def\apa#1#2#3{{\it Act. Phys. Austr.}{\bf #1} (#2) #3} \defProc. LHC Workshop, CERN 90-10{Proc. LHC Workshop, CERN 90-10} \def\npb#1#2#3{{\it Nucl. Phys. }{\bf B #1} (#2) #3} \def\nP#1#2#3{{\it Nucl. Phys. }{\bf #1} (#2) #3} \def\plb#1#2#3{{\it Phys. Lett. }{\bf B #1} (#2) #3} \def\prd#1#2#3{{\it Phys. Rev. }{\bf D #1} (#2) #3} \def\pra#1#2#3{{\it Phys. Rev. }{\bf A #1} (#2) #3} \def\pR#1#2#3{{\it Phys. Rev. }{\bf #1} (#2) #3} \def\prl#1#2#3{{\it Phys. Rev. Lett. }{\bf #1} (#2) #3} \def\prc#1#2#3{{\it Phys. Reports }{\bf #1} (#2) #3} \def\cpc#1#2#3{{\it Comp. Phys. Commun. }{\bf #1} (#2) #3} \def\nim#1#2#3{{\it Nucl. Inst. Meth. }{\bf #1} (#2) #3} \def\pr#1#2#3{{\it Phys. Reports }{\bf #1} (#2) #3} \def\sovnp#1#2#3{{\it Sov. J. Nucl. Phys. }{\bf #1} (#2) #3} \def\sovpJ#1#2#3{{\it Sov. Phys. LETP Lett. }{\bf #1} (#2) #3} \def\jl#1#2#3{{\it JETP Lett. }{\bf #1} (#2) #3} \def\jet#1#2#3{{\it JETP Lett. }{\bf #1} (#2) #3} \def\zpc#1#2#3{{\it Z. Phys. }{\bf C #1} (#2) #3} \def\epj#1#2#3{{\it Eur. Phys. J. }{\bf C #1} (#2) #3} \def\ptp#1#2#3{{\it Prog.~Theor.~Phys.~}{\bf #1} (#2) #3} \def\nca#1#2#3{{\it Nuovo~Cim.~}{\bf #1A} (#2) #3} \def\ap#1#2#3{{\it Ann. Phys. }{\bf #1} (#2) #3} \def\hpa#1#2#3{{\it Helv. Phys. Acta }{\bf #1} (#2) #3} \def\ijmpA#1#2#3{{\it Int. J. Mod. Phys. }{\bf A #1} (#2) #3} \def\ZETF#1#2#3{{\it Pis'ma Zh. Eksp. Teor. Fiz. }{\bf #1} (#2) #3} \def\jmp#1#2#3{{\it J. Math. Phys. }{\bf #1} (#2) #3} \def\yf#1#2#3{{\it Yad. Fiz. }{\bf #1} (#2) #3} \def\aspn#1#2#3{{\it Arch. Sci. Phys. Nat. }{\bf #1} (#2) #3}
\section{Introduction} We describe an experiment to measure the cross sections for the disintegration of deuterons by neutral- and charged-current interactions with low energy electron-antineutrinos. Data were taken at the Centrale Nucleaire de Bugey in France, at 18~m from the core of Reactor~5. Improvements were made to the cosmic-ray shielding of the detector which we previously used in a similar experiment at the Savannah River Plant in South Carolina in the late 1970s~\cite{Pas79,Rei80}. An outer layer of active cosmic ray veto detectors was added which completely surrounds the lead and steel gamma ray shield. These improvements reduced the neutron background due to cosmic rays by a factor of six to $\sim 25$ day$^{-1}$. There are two reactions of interest in this experiment --- the Neutral Current disintegration of the Deuteron (NCD), \[\bar{\nu}_e + d \rightarrow \bar{\nu}_e + p + n, \] and the Charged Current disintegration of the Deuteron (CCD), \[\bar{\nu}_e + d \rightarrow e^+ + n + n.\] The experiment was designed to probe these reactions at low energies ($\sim$1 MeV). In particular, it measures the square of the isovector-axial vector coupling constant ($\beta^2$). The neutrino-induced disintegration of the deuteron is an ideal reaction for this purpose since, at reactor neutrino energies, all other coupling constants make negligible contributions to the cross section. Other coupling constants depend on the value of the Weinberg mixing angle, $\theta_W$, which is an unspecified parameter of the theory, while $\beta$ is predicted to be -1.0, independent of $\theta_W$. In addition, it does not suffer from ambiguities arising from the presence of vector interactions, nor from momentum-transfer-dependent form factors, to which high-energy experiments are subject. The deuteron disintegration experiment is unique, then, in being able to measure the contribution of a single coupling constant with an unambiguous theoretical value. \section{The Detector} \label{sec:elem} \subsection{Location} The detector was installed at Reactor~5 of the Centrale Nucleaire de Bugey, near Lyon, France. It is located in a room about 10 m below ground level with an overburden of 25~mwe. The distance from the center of the reactor core to the center of the detector was 18.5 meters. \subsection{The Target} Schematics of the detector and shielding are shown in Figures~\ref{topviewout}, \ref{sideview}, and~\ref{fig:target}. The target detector, labeled D$_{2}$O in Figures~\ref{topviewout} and \ref{sideview} and shown in more detail in Figure ~\ref{fig:target}, consists of a cylindrical stainless steel tank, 54 cm in diameter, 122 cm in height, and a wall thickness of 0.18 cm, containing 267~kg of 99.85\% pure D$_{2}$O and ten tubular proportional chambers, equally spaced in two concentric rings of 10.16 cm and 20.37 cm radius and offset from each other by $36^{\circ}$. Running down the center of the target tank is a stainless steel tube that allows the placement of radioactive sources inside the tank for calibration purposes. \begin{figure} \centerline{\psfig{figure=fig1.ps,height=5in}} \caption{Plan view of detector and shielding in the opened configuration at the Bugey site.} \label{topviewout} \end{figure} \begin{figure} \centerline{\psfig{figure=fig2.ps,height=4in}} \caption{Side view of detector and shielding configuration.} \label{sideview} \end{figure} \begin{figure} \centerline{\psfig{figure=fig3.ps,height=3in}} \caption{Top view of target tank.} \label{fig:target} \end{figure} Immediately surrounding the target tank is 10~cm of lead shielding and a 1~mm layer of cadmium to absorb thermal neutrons. These are contained in an outer steel tank that sits on a small pedestal inside the large, inner veto detector tank (Tank 2 of Figures~\ref{topviewout} and \ref{sideview}). The proportional counters are 5.08 cm in diameter, 122 cm in height, have a wall thickness of 0.025 cm, and are filled with 1~atm of $^3$He and 1.7 atm of Ar as a buffer. They are essentially black to thermal neutrons, with a capture cross section of $\sim$5300 barns per $^{3}$He nucleus. The neutron capture in the counters proceeds via the (n,p) reaction: \[ ^{3}\mbox{He} + n \rightarrow \: ^{3}\mbox{H} + p + 764 \mbox{ keV.}\] The energy resolution of the counters was measured to be 3\% at the 764~KeV neutron capture peak. A typical neutron spectrum obtained with a $^{252}$Cf neutron source is shown in Fig.~\ref{fig:He}. A discussion of the neutron detection efficiency is given in Section~\ref{sec:neff}. A more detailed description of the construction and testing of the $^{3}$He proportional tubes can be found in Reference~\cite{Ref29}. \begin{figure} \vspace*{2in} \centerline{\psfig{figure=fig4.ps,height=1.5in}} \caption{Neutron-response spectrum of a $^3$He counter with a $^{252}$Cf source.} \label{fig:He} \end{figure} \subsection{Detection Technique} The neutral-current and charged-current events in the D$_{2}$O target are recognized solely by the neutrons they produce: the neutral-current reaction releases a single neutron and the charged-current releases two. Consequently, the quantities of interest are the rates of single and double neutron captures. \subsection{The Shielding and Anticoincidence System} Due to the detector's close proximity to the reactor core, there can be a significant reactor associated gamma flux. Gamma rays of $>$2.2~MeV which reach the target detector can photodisintegrate the deuterons, leading to single neutron signals. In the previous version of the experiment, this background was reduced by surrounding the inner layer of active cosmic-ray veto detectors with a layer of lead and water shielding. Unfortunately, cosmic rays interacting in the surrounding lead shield, but not reaching the inner veto counters, were a significant source of neutrons in the target detector. It was concluded that the shielding could be improved by an additional layer of active cosmic ray veto detectors outside the lead shielding. In this way, cosmic rays interacting in the lead would be seen by the outer veto detectors. Simulations showed that this would reduce the cosmic ray neutron background by a factor of three to four. In the current configuration, the target tank is in the center of a large liquid scintillator detector (the ``inner'' veto) composed of Tank~2 and Tank~4, shown in Figure~\ref{sideview}. Immediately surrounding Tanks 2 and 4 is a layer of each lead and steel. Surrounding this layer of passive shielding is an outer layer of cosmic ray veto detectors (the ``outer'' veto). Slabs of plastic scintillator cover the north and south sides and the bottom face, while larger tanks of mineral oil scintillator (Tanks~1, 3, \& 5) cover the east, west, and top faces. The liquid scintillator used in all five tanks is mineral oil based with a high flash point. Five-inch hemispherical photomultiplier tubes (PMTs) are used to view the liquid scintillator tanks and three inch tubes are employed on the plastic slabs. As noted above, the inner veto system consists of two liquid scintillator tanks, Tank~2 and Tank~4. As indicated in Figures~\ref{topviewout} and \ref{sideview}, there is a string of fifteen evenly spaced PMTs along each vertical corner of Tank~2 . Alternating tubes are offset in direction by $90^{\circ}$. Along the east and west walls on the floor of the tank is a row of PMTs which view the space underneath the target tank. Tank~4 has three PMTs in each vertical corner that are configured like those of Tank~2. There are eight signal lines coming from the inner veto system. The PMTs in each vertical string are ganged onto a single line, as are each row along the floor. The signals from the east corners of Tank~4 are fanned together as are the signals from the west corners. The inner veto detector is primarily a ``soft'' veto --- its signals are recorded at each trigger and analyzed off line. However, it also triggers an on-line veto in the event that all four corner strings see a large pulse simultaneously. Such a signal is likely to be produced by a throughgoing muon. Figure~\ref{fig:electronics} shows some details of our electronics configuration. \begin{figure} \centerline{\psfig{figure=fig5_1.ps,height=8in}} \caption{Some details of our electronics configuration.} \label{fig:electronics} \end{figure} \begin{figure} \centerline{\psfig{figure=fig5_2.ps,height=8in}} \begin{center} Figure \protect\ref{fig:electronics} continued. \end{center} \end{figure} \subsection{Data collection system} The data-collection program, based on a 80486DX processor and the software package LabVIEW, has a fast, graphical interface to the electronics. The software takes advantage of the multitasking capabilities of the operating system, allowing the transfer and processing of data without interrupting data collection. A trigger is generated under the following conditions: \begin{enumerate} \item A neutron-like pulse is detected in one of the $^3$He proportional counters. \item No pulses above hardware thresholds were detected in any of the inner or outer veto detectors in the preceeding $\sim 900\mu$s. This value was chosen to reduce background from muon-induced neutrons arising in the inner-anti scintillator which had a neutron capture time of about 200$\mu$s. \end{enumerate} When a trigger occurs, the contents of waveform digitizers and scalers are read by the computer and written to disk. The contents of the digitizers give a pulse history of all detectors for a period of 4~milliseconds before and after the event. More details on the detector and data-collection system can be found in Reference~\cite{steveT}. \section{Data Analysis} \subsection{Selection criteria} After the data are collected, they are further reduced by offline selection according to the following criteria. \subsubsection{Target cuts} The purpose of the target cuts is to remove any events that do not appear to be valid neutron captures. \begin{description} \item[\it No neutron in pulse-height window.] An event is removed if there is no target pulse in the pulse-height acceptance window within 5$\mu$sec\ of the trigger time. The pulse-height acceptance values were determined from the neutron calibrations and varied slightly from run to run. The total number of target peaks in the pulse-height window during the 782$\mu$sec\ (three times the neutron capture time in the target) following the trigger is taken to be the number of neutrons in the event. This time interval was selected to maximize the signal to background. Shorter time windows yield consistent results with larger statistical errors. \item [\it Remove ``early'' neutrons.] Remove event if it has a pulse in the neutron pulse-height acceptance window before the trigger time. \end{description} \subsubsection{Outer-anti cut} This cut removes cosmic-ray muons that might create neutrons that would subsequently be detected by the target. \begin{description} \item[\it Remove muons.] If during the 1800 $\mu$sec\ preceeding the trigger a signal is detected in any outer anti which exceeds the threshold for that counter, the event is removed. The pulse-height threshold values were determined run by run. The values were chosen at the lowest value for which the events removed by this criterion were at least twice the number of the ``background'' peaks at the same pulse height. \end{description} \subsubsection{Inner-anti cuts} The inner-antis provide additional protection against cosmic-ray muons that sneak through the outer anti. However, this large volume of liquid scintillator also provides a large target for inverse-beta events on hydrogen ($\bar{\nu_e} p \rightarrow e^+ n$). A small fraction ($\sim 0.08\%$) of the neutrons thus produced diffuse into the target area and are recorded by the $^3$He tubes. To reduce this number, a low-energy cut is applied to the inner antis, thus using the light produced by the positron to veto the event. \begin{description} \item[\it Low-energy cut.] The main purpose of this cut is to remove the inverse-beta events. Any event with $>$0.8 MeV in either Tank 2 or Tank 4 within 900$\mu$sec\ before to 200$\mu$sec\ after the trigger was removed. This energy-threshold value was chosen in order to remove the maximum number of inverse-beta events, while not suffering too much dead time from the many low-energy background pulses. From the Monte Carlo, the mean time between production of a neutron by the inverse-beta process in Tanks 2 or 4 and its subsequent capture by a $^3$He tube was about 230$\mu$sec. Thus the period in which this cut is active is from about four such mean times before to one after the trigger. \item[\it High-energy cut.] Events having a pulse of total energy exceeding 8~MeV in Tank 2 or 6~MeV in Tank 4 from 2400 to 900$\mu$sec\ before the trigger are removed. This cut removes cosmic-ray events that are recorded before the beginning of the hardware anti block. Extending the times earlier than 2400$\mu$sec\ has little effect. \end{description} The fraction of events removed by each of the above cuts is shown in Table~\ref{tab:cuts}. Figure~\ref{fig:cuts} shows the effects of the cuts on the data. As a result of the cuts, the number of candidate neutrino events is reduced from roughly 60,000 per day to about 25 per day, with the reactor off. \begin{figure} \vspace*{2in} \centerline{\psfig{figure=fig6.ps,width=7in}} \vspace{-1in} \caption{All $^3$He detector signals, above the 25-count hardware threshold, occuring in the 782~$\mu$sec (three neutron capture times) following the event trigger. a) and b) are before any software cuts were applied; c) and d) after all cuts were applied.} \label{fig:cuts} \end{figure} \begin{table} \begin{center}\begin{tabular}{lcccc} \hline\hline & \multicolumn{2}{c}{Reactor ON} & \multicolumn{2}{c}{Reactor OFF}\\ Cut & Random & Normal & Random & Normal \\ & Triggers & Triggers & Triggers & Triggers \\ \hline Early neutron & 0.5\% & 27.9\% & 0.5\% & 28.7\% \\ No neutron in window & 0\% & 17.2\% & 0\% & 17.5\% \\ Outer-anti cut & 26.0\% & 38.9\%& 26.4\% & 38.7\% \\ Inner-anti (high-energy) & 9.2\% & 8.5\% & 9.5\% & 9.3\% \\ Inner-anti (low-energy) & 17.2\% & 6.5\% & 19.2\% & 5.1\% \\ \hline\hline \end{tabular}\end{center} \caption{Fractions of total numbers of all events removed by each data cut when applied in the order indicated. Since many events satisfy more than one cut criterion, these values would change if the ordering of the cuts were changed.} \label{tab:cuts} \end{table} \subsection{Monte Carlo calculations} {\sc geant} with the {\sc gcalor} interface was used for all Monte Carlo simulations. The {\sc gcalor} interface handles neutron transport from 20 MeV down to thermal energies. {\sc geant} handles the transport of all other particles. One comparison of the data and the Monte Carlo is given in Figs.~\ref{fig:HeA} and \ref{fig:HeB}. The former shows the capture-time spectrum of neutrons detected by the $^3$He counters from a simulated $^{252}$Cf source at the center of the target detector. The mean capture time is $265\pm3$ $\mu$secs. Fig.~\ref{fig:HeB} is for the same configuration, but for real data. The mean capture time is $267\pm4$ $\mu$secs. \begin{figure} \centerline{\psfig{figure=fig7.ps,height=5in}} \caption{Capture-time spectrum of neutrons detected by the $^3$He counters from a simulated $^{252}$Cf source in the center of the target detector. The mean capture time is 265$\pm$3 $\mu$secs. Results are from 100,000 generated neutrons.} \label{fig:HeA} \end{figure} \begin{figure} \centerline{\psfig{figure=fig8.ps,height=5in}} \caption{Capture-time spectrum of the second neutron detected by the $^3$He counters from a real $^{252}$Cf source in the center of the target detector. The mean capture time is 267$\pm$4~$\mu$secs.} \label{fig:HeB} \end{figure} \subsection{Neutron Detection Efficiency} \label{sec:neff} Special neutron-calibration runs were periodically made with a $^{252}$Cf source in the center of the target detector. Data from these runs were processed thru the same programs used to analyze the neutrino events. In particular, the same target cuts (as described above) were used. The resulting pulse-height spectra from the $^3$He tubes were histogrammed for each calibration run, and the peaks fitted to Gaussians. Only those pulses within 2 standard deviations of the peak value are finally accepted as neutrons. The numbers of events with 1, 2, 3, 4, and 5 neutrons within a given time window were tallied. The time window chosen was 3 neutron capture times. Based on the known neutron multiplicity from $^{252}$Cf fissions, one can calculate the neutron detection efficiency by assuming various efficiencies and comparing the observed number distribution with the calculated distribution. Our procedure took into account: \begin{itemize} \item The neutron-number distribution from $^{252}$Cf fission. \item The neutron acceptance time window of 3 capture times. \item The probability of an ``extra'' fission from the Cf source during the acceptance time window, which is a function of the source activity. \end{itemize} This procedure yielded a mean efficiency of 0.41$\pm$0.01 for a neutron source at the center of the target. This value agreed well with the value derived from the Monte Carlo. As a result we were able to use the Monte Carlo value of 0.29${\pm}$0.01 as the mean efficiency for single neutrons generated isotropically throughout the D$_2$O of the target volume. The efficiency for two neutrons is the square of the single-neutron efficiency (0.084${\pm}$0.006). And the efficiency for seeing only 1 neutron, if 2 were produced is $2 \times 0.29 \times (1.0-0.29) = 0.41\pm0.01$. \subsection{Energy calibration of inner antis} Since we desired to base the inner anti cut criteria on energy, both Tanks 2 and 4 must be energy calibrated. Periodic runs were made over the course of the experiment with a $^{60}$Co source placed at various known positions in Tank 2, and beneath the center of Tank 4. (Tank 4 was also calibrated with a $^{252}$Cf source in that same position.) The data were compared with Monte Carlo simulations. Several algorithms were tested to find the best estimates of the energy. The best measures found were: Tank 4, sum the signals from the two PMT strings; Tank 2, sum the signals from the 4 vertical corner strings. Results are shown in Table~\ref{tab:calib}. \begin{table} \begin{center}\begin{tabular}{lcc} \hline\hline & Tank 2 & Tank 4\\ \hline Uncertainty in peak position & ${\pm}$15\% & ${\pm}$10\% \\ Standard deviation & 0.25E$^{\frac{1}{2}}$ & 0.25E$^{\frac{1}{2}}$ \\ Max. deviation of peak over entire run & ${\pm}$20\% & ${\pm}$10\% \\ \hline\hline \end{tabular}\end{center} \caption{Calibration results for inner anti using a $^{60}$Co source.} \label{tab:calib} \end{table} \section{Results} \subsection{Event rates} The 1- and 2-neutron event rates for both the reactor up and down data are given in Table~\ref{tab:updown}. Subtracting the reactor down rates from the up rates yields the data shown in Table~\ref{tab:onetwo}, where we have also given the corresponding neutron detection efficiencies. \begin{table} \begin{center}\begin{tabular}{lcc} \hline\hline & Reactor Up & Reactor Down\\ \hline Raw 1-neutron rate (events/day) & 44.62${\pm}$0.59 & 25.28${\pm}$0.68\\ Raw 2-neutron rate (events/day) & 2.69${\pm}$0.14 & 1.45${\pm}$0.16 \\ Software efficiency & 0.471${\pm}$0.003 & 0.444${\pm}$0.005 \\ Corrected 1-neutron rate & 94.66${\pm}$1.24 & 57.00${\pm}$1.53 \\ Corrected 2-neutron rate & 5.71${\pm}$0.31 & 3.26${\pm}$0.36 \\ \hline\hline \end{tabular}\end{center} \caption{Reactor up and down event rates.} \label{tab:updown} \end{table} \begin{table} \begin{center}\begin{tabular}{lcc} \hline\hline & 1 neutron & 2 neutron \\ \hline Up minus down rate & 37.7${\pm}$2.0 & 2.45${\pm}$0.48 \\ Neutron efficiency & 0.29${\pm}$0.01 & 0.084${\pm}$0.006\\ \hline\hline \end{tabular}\end{center} \caption{1- and 2- neutron event rates and detection efficiencies.} \label{tab:onetwo} \end{table} The 2-neutron rate (per day) is $ (2.45\pm0.48)/(0.084\pm0.006) = 29.2\pm6.1$. To get the CCD rate from this value we need only correct for the effect of a nearby reactor, Reactor \#4. It is located about 80~m from our detector. While taking data with Reactor \#5 up, the mean power of reactor \#4 was 1925 MW; while \#5 was down, it was 2246 MW. This gives a correction factor of +0.6\% to our final rates. Thus the CCD daily rate is \[R_{CCD} = (29.2\pm6.1) \times (1.006) = 29.4\pm6.1 \] To get the NCD rate from the 1-neutron rate, two corrections must first be applied to the 1-neutron rate. \begin{itemize} \item The number of CCD reactions in which only 1, instead of 2, neutrons was observed must be subtracted. This number is the CCD rate times the efficiency of seeing only one out of the two neutrons: \[ (29.2\pm6.1)\times(0.41\pm0.01) = 12.0\pm2.5 \] \item The number of inverse-beta decays in the inner detector that leak into the target volume and create a single neutron must also be subtracted. From the Monte Carlo we estimate 22.0${\pm}$0.5 inverse-beta events per day enter the target volume. Also from the Monte Carlo we estimate that only 5${\pm}$1\% of those events survive the 0.8 MeV inner-anti cut. Thus the number of events to be subtracted from the 1-neutron rate is: \[ (22.0\pm0.5) \times (0.05\pm0.01) \times (0.29\pm0.01) = 0.3\pm0.1 \] \end{itemize} The corrected 1-neutron event rate is then \[(37.7\pm2.0) - (12.0\pm2.5) - (0.3\pm0.1) = 25.4\pm3.2\] Applying the single-neutron detection efficiency correction and the Reactor \#4 correction from above, yields the daily NCD rate: \[R_{NCD} = (25.4\pm3.2)\times(1.006)/(0.29\pm0.01) = 88.1 \pm 11.1 \] \subsection{Systematic uncertainties} The significant systematic uncertainties are given in Table~\ref{tab:syst}. Other possible sources of systematic effects which were considered, but found to be insignificant were: the calculated neutrino energy spectrum and the energy-calibration effects on data cuts. \begin{table} \begin{center}\begin{tabular}{lcc} \hline\hline Parameter & Value & \%\\ \hline Detector-reactor distance & 18.5${\pm}$0.1 m & 1.1\\ Mass of D$_2$O & 267.0${\pm}$2.0 kg & 0.8\\ No. MeV per fission & 205.0${\pm}$0.7& 0.3 \\ \multicolumn{2}{l}{Total systematic uncertainty} & 1.4 \\ \hline\hline \end{tabular}\end{center} \caption{Parameters which have significant contributions to the systematic uncertainties in the data rates. The last column shows the contribution of each parameter to the systematic uncertainty in the final event rates.} \label{tab:syst} \end{table} \subsection{Theoretically-expected event rates} The rates (events per day) are given by: \begin{equation} R = \frac{N_D}{4\pi r^2} \int \bar{N}_{\nu}(E_{\nu}) \sigma(E_{\nu}) dE_{\nu} \label{eqn:R} \end{equation} where $E_{\nu}$\ is the neutrino energy, $\bar{N}_{\nu}(E_{\nu})$ the daily average neutrino energy spectrum per MeV, $N_D$ the total number of deuterons in the target, $\sigma(E_{\nu})$ the cross section for the process, and $r$ is the distance from the reactor to the detector. The mean neutrino energy spectrum was determined from the reactor power and the core ``burn up,'' i.e.\ the isotopic composition of the fuel, as a function of time. The reactor power was obtained from reactor monitoring devices several times per day. The isotopic composition of the fuel rods was given to us at the beginning and ending of each reactor cycle of about 11 months. The only four nuclei of importance are: $^{235}$U\cite{Schreckenbach}, $^{238}$U\cite{Klapdor}, $^{239}$Pu\cite{Hahn}, and $^{241}$Pu\cite{Hahn}. Combining the data in those references with the reactor power as a function of time, both the neutrino energy spectrum and the conversion factor from MW-hours to total number of neutrinos was calculated for each day. The energy per fission and the mean number of fissions per day are given in Table~\ref{tab:fissions} for each isotope. \begin{table} \begin{center}\begin{tabular}{ccc} \hline\hline Isotope & MeV/fission & N$_{fiss}$ \\ \hline $^{235}$U & 201.7${\pm}$0.6 & 4.17$\times10^{24}$ \\ $^{238}$U & 205.0${\pm}$0.9 & 5.61$\times10^{23}$ \\ $^{239}$Pu & 210.0${\pm}$0.9 & 2.10$\times10^{24}$ \\ $^{241}$Pu & 212.4${\pm}$1.0 & 3.46$\times10^{23}$ \\ \hline\hline \end{tabular}\end{center} \caption{The energy per fission and the mean number of fissions per day for each isotope.} \label{tab:fissions} \end{table} The data-collection MW-hours was calculated for every day by combining the data collection times with the reactor power level at that time. The number of deuterons was 1.605$\times10^{28}$. Combining all these factors and dividing by the number of live days yields the mean neutrino spectrum (neutrinos/MeV/day) as shown in Table~\ref{tab:spectrum}. \begin{table} \begin{center}\begin{tabular}{lclc} \hline\hline Energy & N$_{\nu}$ & Energy & N$_{\nu}$ \\ \hline 2.0& 3.56$\times10^{25}$ & 6.0& 8.98$\times10^{23}$ \\ 2.25& 2.99$\times10^{25}$ & 6.25& 6.49$\times10^{23}$ \\ 2.5& 2.47$\times10^{25}$ & 6.5& 4.84$\times10^{23}$ \\ 2.75& 2.09$\times10^{25}$ & 6.75& 3.55$\times10^{23}$ \\ 3.0& 1.75$\times10^{25}$ & 7.0& 2.47$\times10^{23}$ \\ 3.25& 1.45$\times10^{25}$ & 7.25& 1.58$\times10^{23}$ \\ 3.5& 1.18$\times10^{25}$ & 7.5& 1.01$\times10^{23}$ \\ 3.75& 9.47$\times10^{24}$ & 7.75& 6.13$\times10^{22}$ \\ 4.0& 7.54$\times10^{24}$ & 8.0& 3.27$\times10^{22}$ \\ 4.25& 5.93$\times10^{24}$ & 8.25& 1.35$\times10^{22}$ \\ 4.5& 4.52$\times10^{24}$ & 8.5& 8.11$\times10^{21}$ \\ 4.75& 3.44$\times10^{24}$ & 8.75& 4.89$\times10^{21}$ \\ 5.0& 2.68$\times10^{24}$ & 9.0& 2.77$\times10^{21}$ \\ 5.25& 2.07$\times10^{24}$ & 9.25& 1.65$\times10^{21}$ \\ 5.5& 1.57$\times10^{24}$ & 9.5& 1.17$\times10^{21}$ \\ 5.75& 1.21$\times10^{24}$ & \\ \hline\hline \end{tabular}\end{center} \caption{Time-averaged number of neutrinos per day per MeV. Energies are at lower bin edge.} \label{tab:spectrum} \end{table} There has been considerable work done on the CCD and NCD cross sections in the past few years. Kubodera and Nozawa review the field in Ref.~\cite{Kubodera}. In their Table 1, they give the cross sections for both the CCD and NCD reactions from threshold to 170 MeV. They state that the uncertainties in the values are 5\%. Using the data of Ref.~\cite{Kubodera} with Eqn.~\ref{eqn:R}, yields $R_{NCD} = 87.2\pm4.4$ and $R_{CCD} = 30.4\pm1.5$. \subsection{Experimental cross sections} The average cross section per neutrino is given by \[ \bar{\sigma} = \frac{ \int \bar{N}_{\nu}(E_{\nu}) \sigma(E_{\nu}) dE_{\nu}} { \int \bar{N}_{\nu}(E_{\nu}) dE_{\nu}} \] where the integrals go from the threshold for the reaction to infinity. Combining this with Eqn.~\ref{eqn:R}, we get \[ \bar{\sigma} = \frac{4\pi r^2 R} {N_D \int \bar{N}_{\nu}(E_{\nu}) dE_{\nu}}. \] The values obtained for the NCD and CCD cross sections are given in Table~\ref{tab:sigma}. \begin{table} \begin{center}\begin{tabular}{lcc} \hline\hline & NCD & CCD \\ \hline Rate (events per day) & 88.1 ${\pm}$ 11.1 ${\pm}$ 1.2 & 29.4 ${\pm}$ 6.1 ${\pm}$ 0.4 \\ Reaction threshold (MeV) & 2.23 & 4.03 \\ Neutrinos (per day) & 3.86 $\times 10^{25}$ & 8.01 $\times 10^{24}$ \\ Average cross section (10$^{-45}$ cm$^2$ per $\bar{\nu_e}$) & 6.08 $\pm$ 0.77 & 9.83 $\pm$ 2.04 \\ \hline\hline \end{tabular}\end{center} \caption{The rates (with statistical and systematic uncertainties), reaction thresholds, total numbers of neutrinos above threshold, and cross sections for the NCD and CCD reactions as measured in this experiment}. \label{tab:sigma} \end{table} \subsection{Improved NCD cross section} As stated above, the CCD events create a significant background for the NCD events, and this background must be subtracted. The large uncertainty in our measured CCD rate makes a significant contribution to the uncertainty in the NCD rate. However, we note that our experimentally determined rates and the theoretically-expected rates agree within one standard deviation of each the experimental and theoretical uncertainties. Given this excellent agreement, we feel that an improved value for the NCD cross section may be calculated by using the theoretically-expected CCD daily rate (30.4 $\pm$ 1.52) rather than our observed rate (29.4 $\pm$ 6.1). Repeating the procedure described above, this yields an improved NCD rate of 86.7 $\pm$ 7.9, and a corresponding cross section of $5.98 \pm 0.54 \times 10^{-45}$ cm$^2$ per neutrino. \subsection{Calculation of $\beta^2$} The value of $\beta^2$ is given by the ratio of the measured neutral current rate to the theoretically expected rate. Thus we find \[ \beta^2 = \frac{88.1\pm11.1\pm1.2}{87.2\pm4.4} = 1.01\pm0.16 \] Using the improved NCD cross section determined above, we get an improved $\beta^2$ of 0.99 $\pm$ 0.10. \subsection{Neutrino oscillations} Another aspect of this experiment is its ability to explore neutrino oscillation by measuring the ratio of the CCD to NCD rates. At reactor neutrino energies, there is insufficient energy to create leptons more massive than the electron. Therefore, if neutrino oscillation occurs at a significant level, a deficit of charged-current events compared to neutral-current events should be seen. This leads us to define the ratio $R$, where \begin{equation} R = \frac{ \frac{ {\rm CCD_{exp}}}{ {\rm NCD_{exp}}}} { \frac{ {\rm CCD_{th}}}{ {\rm NCD_{th}}}}, \end{equation} a ratio of ratios of experimentally determined reaction rates to theoretically expected reaction rates. A deficit of charged current reactions could imply that some electron antineutrinos have oscillated to a different flavor or helicity state, either of which would imply new physics. We find \[ R = \frac{\frac{29.4\pm6.1}{88.1\pm11.1}}{\frac{30.4}{87.2}} = \frac{0.334\pm0.080}{0.348\pm0.004} = 0.96\pm0.23 \] The error of 1\% in the theoretical ratio is taken from reference~\cite{Kubodera}. The neutrino-oscillation exclusion plot resulting from this value of R is shown in Fig.~\ref{fig:exclusion}. \begin{figure} \centerline{\psfig{figure=fig9.ps,height=8in}} \vspace{-1.5in} \caption{The neutrino-oscillation exclusion plot corresponding to our value of $R$, the ratio of the observed to expected ratios of the CCD to NCD rates. The solid line is the 90\% confidence level contour; the dashed, the 95\%.} \label{fig:exclusion} \end{figure} \subsection{Possible extension of this technique} Since the theoretical error in the ratio is quite small, a high statistics, good precision measurement of R should be possible. This measurement has the potential of reaching small values of $\sin^{2}2\theta$. In the current experiment, the CCD measurement is handicapped by the requirement that we observe two neutrons. The efficiency for observing this goes as the square of the single-neutron detection efficiency and so is necessarily small. Another method, which we explored but did not pursue, employs the addition of a small amount (approximately 10\%) of light water into the heavy water target. This small addition does not effect the neutron detection efficiency appreciably and gives one the opportunity to observe the charged current reaction on the proton. Since the CCP reaction has a much larger cross-section than the CCD reaction, a threshold of 1.8 MeV, closer to the CCD threshold and since it can be detected by searching for a single neutron, one can determine the ratio of NCD to CCP with higher precision. \section{Discussion} This experiment was an improved version of our experiment done at Savannah River in the late 1970s. The primary improvements were in the cosmic-ray shielding, which cut that background by a factor of six, and an improved data-collection system. During the past 20 years great progress has been made in calculating the CCD and NCD cross sections, and they agree well with the results of this experiment. \section*{Acknowledgments} The authors would like to acknowledge the operators of the Bugey Nuclear Plant, and the contributions of our technicians, Thomasina Godbee, Herb Juds, Eric Juds, and Butch Juds. This work was supported by the U.S. Department of Energy.
\section{Introduction} Supersymmetric sigma models in 2+1 dimensions with a K\"ahler target space generally admit static soliton-like `lump' solutions with energy $E= |T|$, where $T$ is the topological charge $\int \omega $ obtained by integrating the K\"ahler 2-form $\omega$ over the image in target space of the 2-dimensional space (see e.g. \cite{ruback}). If the K\"ahler target space admits a holomorphic Killing vector field $k$ then one can perform a `Scherk-Schwarz' (SS) dimensional reduction to arrive at a `massive' supersymmetric sigma model in 1+1 dimensions with a scalar potential $V \sim k^2$. This theory admits `Q-kink' solutions \cite{AT,paptown} with an energy \begin{equation} E = \sqrt {Q_0^2 + Q^2}\, , \end{equation} where $Q_0$ is the Noether charge associated with $k$, and $Q = \int i_k\omega$, the integral being taken over the image in target space of the 1-dimensional space. Because $k$ is holomorphic the 1-form $i_k\omega $ is closed, so $Q$ is a topological charge. When $Q_0 \ne 0$ the Q-kink is a {\sl time-dependent} solution of the sigma-model field equations. When $Q_0 = 0$ it becomes a standard static kink solution. A 2+1 dimensional supersymmetric sigma model with a K\"ahler target space has an N=2 supersymmetry and the topological charge $T$ appears as a central charge in the supersymmetry algebra. This implies the bound $E \ge |T|$, which is saturated by the sigma-model lumps. Similarly, 1+1 dimensional massive supersymmetric sigma models obtained by SS dimensional reduction actually have (2,2) supersymmetry, and both $Q_0$ and $Q$ appear in the supersymmetry algebra as central charges. This implies the bound $E \ge \sqrt {Q_0^2 + Q^2}$, which is saturated by the Q-kinks. If the K\"ahler target space is actually hyper-K\"ahler then the topological charge $T$ of the 2+1 dimensional model is just one of a triplet of topological charges \begin{equation} {\bf T} = \int\! \mbox{\boldmath $\omega$}\, , \end{equation} where $\mbox{\boldmath $\omega$}$ is the triplet of K\"ahler 2-forms. The number of supersymmetries is also doubled to N=4, and the triplet of charges $\bf T$ appear as central charges in the N=4 supersymmetry algebra. If the hyper-K\"ahler space admits a tri-holomorphic Killing vector field $k$ then SS dimensional reduction along its orbits yields a (4,4) supersymmetric massive sigma model in 1+1 dimensions, again with $V \sim k^2$. The topological charge $Q$ is now one of a triplet of topological charges \begin{equation} {\bf Q} = \int i_k \mbox{\boldmath $\omega$}\, , \end{equation} and the four charges $(Q_0, {\bf Q})$ appear as central charges in the (4,4) supersymmetry algebra. This implies the bound \begin{equation} E \ge \sqrt{Q_0^2 + {\bf Q\cdot Q}}\, , \end{equation} which is saturated by the (hyper-K\"ahler) Q-kinks. There is a close analogy here to N=2 and N=4 supersymmetric Yang-Mills (SYM) theories in 4+1 and 3+1 dimensions \cite{AT,dorey}. The lumps of the 2+1 dimensional sigma model are similar to the instantonic solitons of the 4+1 SYM theory; for example, they have no fixed scale. The Q-kinks of the 1+1 dimensional sigma model are similar to the dyons of 3+1 SYM theory; for example the sigma model has a vacuum angle and Q-kinks generally have fractional $Q_0$-charge, just as SYM dyons generally carry fractional electric charge for nonzero vacuum angle. The scale introduced by the potential term in the 1+1 dimensional sigma model is analogous to the scale introduced by the Higgs mechanism in the SYM case. The N=2 and N=4 SYM theories have interpretations in IIB string theory as effective field theories describing the fluctuations of D-branes around some `vacuum' brane configuration. The dyon solutions are the field theory realization of (p,q) strings, or string webs, stretched between the D-branes. A feature of the brane interpretation of the SYM theories is that in a limit in which the individual branes become widely separated the dyon solutions must transmute into a solution of the equations governing the dynamics of a ${\sl single}$ brane. This is an Abelian SYM theory, although not of a conventional type because the brane action involves higher derivative interactions. These `DBI solitons', were found in \cite{CM,G}; the supersymmetric solutions are worldvolume `spikes' of infinite total energy per unit length equal to the tension of a (p,q) string. Solutions with finite ${\sl total}$ energy can be found by considering the DBI action in an appropriate supergravity background \cite{UK}. These considerations motivate us to seek an interpretation of sigma-model lumps and Q-kinks as solitons on the worldvolume of the eleven-dimensional supermembrane \cite{supermem}, otherwise known as the M2-brane. An M2-brane in a vacuum background has supersymmetric, but infinite energy, vortex solutions that can be interpreted as intersections with other M2-branes \cite{CM,G,GGT}. In a non-vacuum K{\" a}hler background we may have the option of wrapping the `other' M2-branes on finite area holomorphic 2-cycles of the background. These are finite energy solitons that provide the brane realization of K{\" a}hler sigma-model lumps. We shall concentrate here on the hyper-K\"ahler case; specifically, we shall consider the supermembrane in a background for which the 4-form field-strength vanishes and the 11-metric takes the form \begin{equation} \label{11m} ds^2 = ds^2(\bb{E}^{(1,5)} \times S^1) + ds^2_4\, , \end{equation} where $ds^2_4$ is the Kaluza-Klein (KK) monopole metric \begin{equation}\label{tfour} ds^2_4= V^{-1}\left (d\varphi - {\bf A}\right )^2 + V\, ds^2(\bb{E}^3) \, . \end{equation} The 1-form $\bf A$ on $\bb{E}^3$ satisfies $\mbox{\boldmath $\nabla$} V = \mbox{\boldmath $\nabla$} \times {\bf A}$, which implies that $V$ is harmonic on $\bb{E}^3$. The vector field $\partial/\partial\varphi$ is Killing and triholomorphic. We take it to be the vector field $k$ of the previous discussion, i.e. \begin{equation} \label{KV} k = \partial/\partial\varphi\, . \end{equation} The orbits of $k$ are Kaluza-Klein (KK) circles which shrink to points at singularities of $V$. Let $\bf X$ be Cartesian coordinates on $\bb{E}^3$ and ${\bf X}_0$ a constant 3-vector from the origin. The simplest choice of $V$ that serves our purposes is \begin{equation}\label{vee} V = 1 + {1\over |{\bf X}+{\bf X}_0|} + {1\over |{\bf X}-{\bf X}_0|}\, , \end{equation} which describes a two-centre KK-monopole of M-theory. Upon reduction on orbits of $k$, the KK-monopole acquires an interpretation as two parallel IIA D6-branes separated in $\bb{E}^3$ by the constant vector $2{\bf X}_0$. The two centres of the metric at ${\bf X}=\pm {\bf X}_0$ can be considered as the poles of a 2-sphere parametrized by $\varphi$ and the distance from one D6-brane along the line joining the two of them. A membrane wrapped on this 2-sphere has a IIA interpretation as a string stretched between the two D6-branes \cite{sen}. Now consider a D2-brane parallel to the two D6-branes. In general it will not be colinear in $\bb{E}^3$ with the two D6-branes and so will not intersect the string joining them. However, we may move it until it does intersect. From the D=11 perspective we then have a pointlike intersection of two M2-branes, one an infinite planar one and the other one wrapped on a finite area 2-cycle of the background. The singular intersection point may be desingularized so that we have a single M2-brane with a non-singular lump soliton on it of some finite size $L$. From the IIA perspective this corresponds to separating the points at which the strings from each of the two D6-branes meet the D2-brane. In the case of the lump, the vacuum is an infinite planar M2-brane. To find a brane interpretation of the hyper-K\"ahler Q-kink we will need to wrap this M2-brane on some one-cycle of the background space. This corresponds to SS reduction on some Killing vector field with closed orbits. The dimensional reduction will preserve all supersymmetries only if this Killing vector field is triholomorphic. The Killing vector field $k$ of (\ref{KV}) is therefore an obvious candidate, but SS reduction on orbits of $k$ does ${\it not}$ yield a potential $V \sim k^2$ as one might have expected from our earlier summary of the results of SS reduction in sigma models. Rather, it yields a non-vanishing, and non-uniform, IIA string tension. The non-uniformity of the tension creates an attractive force between the string and the D6-brane but on reaching the D6-brane core the string can simply dissolve into Born-Infeld flux. To get the potential term in the dimensionally reduced action one must suppose that the 11-metric (\ref{11m}) has another tri-holomorphic Killing vector field with closed orbits. We may take this to be a vector field generating the U(1) isometry of the $S^1$ factor in this metric. Let us call this vector field $\ell$. Dimensional reduction on orbits of $k+\ell$ leads to a bound state of the IIA string discussed above with a D2-brane wrapped on orbits of $\ell$. This bound state is itself bound to the D6-brane. The effective string action is the desired brane version of the massive hyper-K\"ahler sigma model, admitting Q-kink solutions. T-dualizing in the (compact) $\ell$ direction yields a (1,1) IIB string bound to a D5-brane. As we shall see, the Q-kink solution can then be interpreted as a (1,1) string that migrates from one D5-brane to another. Although lump and Q-kink solutions are known to minimise the energy of the relevant sigma model it does not immediately follow that they minimise the energy on the M2-brane because of the nonlinearities of the Dirac membrane action. By means of the brane version of the Bogomol'nyi argument \cite{GGT}, we show that the energy of the M2-brane is indeed minimised by these solutions. We consider the lumps first, as these are static, and then generalize to the Q-kinks. Both configurations are then shown to preserve some fraction of the worldvolume supersymmetry. Again, this is known in the sigma-model case, but the supersymmetry transformations of the supermem\-brane are different. They can be deduced from a combination of the target space supersymmetry and the kappa-symmetry of the supermembrane, and this leads to a simple condition for a worldvolume field configuration to preserve some fraction of supersymmetry \cite{singleton,bbs,BKOP}. For a vacuum background this condition is easily interpreted as a constraint on the 32 independent constant Killing spinors of the background, but its interpretation is less direct in a non-vacuum background in which the Killing spinors are not constant and span a space of lower dimension. Here we present a more geometrical derivation of the conditions for preservation of supersymmetry and we discuss some subtleties of the non-vacuum case that have been passed over previously. \section{Energy bounds} Our starting point for finding soliton solutions as minimum energy configurations of the supermembrane will be its Hamiltonian formulation \cite{yosh}. Let $\xi^i = (t, \sigma^a)$ be the worldvolume coordinates, with $\sigma^a$ the worldspace coordinates, and let $X^m$ be the D=11 spacetime coordinates. The supermembrane Lagrangian, omitting fermions, can then be written as \begin{equation} {\cal L} = P_m \dot {X}^m - s^a P_m\partial_a X^m - {\textstyle{1\over 2}} v \left [ P^2 + {\rm det} (g_{ab}) \right ]\, , \end{equation} where \begin{equation} g_{ab} = \partial_a X^m \partial _b X^n g_{mn} \end{equation} is the induced worldspace metric, $P_m$ is the 11-momentum conjugate to $X^m$, and $s^a$ and $v$ are Lagrange multipliers. Let $X^m = (Y^i, X^I)$ ($i=0,1,2$) so that \begin{equation} ds^2_{11} = dY^i dY^j \eta_{ij} + dX^I dX^J g_{IJ}\, , \end{equation} where $\eta$ is the 3-dimensional Minkowski metric. We make the gauge choice $Y^i(\xi) = \xi^i$. This implies that \begin{equation} g_{ab} = \eta_{ab} + \partial_a X^I \partial_b X^J g_{IJ}\, . \end{equation} It also implies that \begin{equation} P_m = \left ( -\varepsilon -1, -P_I\partial_a X^I, P_I\right ) \end{equation} where $\varepsilon$ is the energy density relative to the brane vacuum (which is taken to have unit tension). The Hamiltonian constraint imposed by $v$ can be solved for $\varepsilon$ \begin{eqnarray}\label{16} (\varepsilon +1)^2 &=& 1 + \nabla X^I \cdot \nabla X^J g_{IJ} + (g^{IJ} + \nabla X^I \cdot \nabla X^J)P_I P_J\nonumber\\ &&+ {\textstyle{1\over 2}} (\nabla X^I \times \nabla X^J)(\nabla X^K\times \nabla X^L)g_{IK}g_{JL}\, , \end{eqnarray} where we have used standard 2D vector calculus notation for worldspace derivatives. This expression differs in several respects from the corresponding expression for the sigma-model energy density. Firstly, the supermembrane expression is quadratic in $\varepsilon$; this is because the sigma-model approximation is a kind of non-relativistic approximation to the supermembrane (they differ in the same way that the energies of a relativistic and non-relativistic particles differ). Secondly, the supermembrane expression involves terms quartic in derivatives that are absent in the sigma-model case. \subsection{Lumps} We now aim to rewrite the above expression for the energy density in the form \begin{eqnarray} \label{e2} (\varepsilon+1)^2 &=& \left ( 1 \pm {\textstyle {1\over 2}}\nabla X^I \times \nabla X^J \omega_{IJ}\right )^2 \nonumber\\ &&+ {\textstyle {1\over 2}}\left ( \nabla X^I \pm *\nabla X^K I_K{}^I\right) \left ( \nabla X^J \pm *\nabla X^L I_L{}^J\right ) g_{IJ}\\ &&+ {\textstyle{1\over 4}}\sum_{r=1}^6 \left (\nabla X^I \times \nabla X^J \Omega_{IJ}^{(r)}\right )^2\, , \nonumber \end{eqnarray} where we have set $P_I =0$ and $*\nabla = (\partial_2, -\partial_1)$ if $\nabla = (\partial_1, \partial_2)$. We assume that $I_I{}^J$ is a complex structure, that the 8-metric $g_{IJ}$ is Hermitian with respect to it and that $\omega_{IJ}=I_I{}^Kg_{KJ}$ is the corresponding closed K\"ahler 2-form. For the moment we leave unspecified the six 2-forms $\Omega^{(r)}$. These conditions are already sufficient to ensure that all but the quartic terms in $\nabla X$ of (\ref{16}) are reproduced. To reproduce the quartic terms too we require that \begin{equation} \label{XIJ} X^{IJ} X^{KL} \left [ \omega_{IJ} \omega_{KL} + \sum_{r=1}^6 \Omega^{(r)}_{IJ} \Omega^{(r)}_{KL} -2 g_{KI}g_{JL}\right ] = 0\, , \end{equation} where \begin{equation} X^{IJ} \equiv \nabla X^I \times \nabla X^J\, . \end{equation} Note that $X^{IJ}$ is an antisymmetric $8\times 8$ matrix. If none of its 4 skew-eigenvalues vanish, then (\ref{XIJ}) implies that \begin{equation} \label{15} \omega_{I(J} \omega_{K)L} + \sum_{r=1}^6 \Omega^{(r)}_{I(J} \Omega^{(r)}_{K)L} = g_{I(K}g_{J)L} - g_{KJ}g_{IL}\, . \end{equation} For a membrane in flat space this condition is satisfied by taking the matrices \begin{equation} I_I{}^J \equiv \omega_{IK}g^{KJ}\, ,\qquad (J^{(r)})_I{}^J \equiv \Omega^{(r)}_{IK}g^{KJ} \end{equation} to be the seven complex structures of $\bb{E}^8$. For every vanishing skew eigenvalue of $X^{IJ}$ the dimension of the transverse space is effectively reduced by two. In this reduced space, we must again have (\ref{15}) but it may now be possible to choose some of the six $J$ matrices to vanish. For example, if $X^{IJ}$ has two vanishing skew-eigenvalues then the transverse space is effectively 4-dimensional; in other words, there are four `active scalars'. We may now set all but two of the $J$ matrices to zero. The other two, together with $I$ can be taken to be the three almost complex structures of the transverse 4-manifold (these will be covariantly constant if this transverse 4-space is hyper-K\"ahler, but we need not assume any special properties at this point). If $X^{IJ}$ has three vanishing skew-eigenvalues, corresponding to two active scalars, then the transverse space is effectively two-dimensional, and we may take all the $J$ matrices to vanish. Given (\ref{XIJ}) we deduce that \begin{equation} \varepsilon \ge {\textstyle{1\over 2}}|X^{IJ} \omega_{IJ}| \end{equation} with equality when \begin{equation}\label{17} \nabla X^I = \mp *\nabla X^J I_J{}^I \end{equation} and \begin{equation} \label{18} X^{IJ} \Omega_{IJ}^{(r)} = 0\hskip 1truecm r = 1,\cdots ,6\, . \end{equation} The condition (\ref{17}) is the statement that in complex coordinates $Z^\alpha$, adapted to the complex structure $I$, the functions $Z^\alpha(z)$ are holomorphic on worldspace, with $z=\sigma^1 + i\sigma^2$. The conditions (\ref{18}) are implied by (\ref{17}) if the matrices $J^{(r)}$ are such that \begin{equation} \label{19} IJ^{(r)} + J^{(r)} I = 0\, ,\hskip 1.5truecm r = 1,\cdots ,6\, . \end{equation} This is true when $I,J^{(r)}$ are the 7 complex structures of $\bb{E}^8$. It is also satisfied if $I,J^{(1)}, J^{(2)}$ are the three almost complex structures of a 4-dimensional space, with the other $J$ matrices vanishing. This is the case of most interest here because we may obviously reduce the transverse 8-space to an effective transverse 4-space by requiring all scalars to vanish except those associated with the $ds^2_4$ metric in (\ref{11m}). This restriction still allows configurations with either two or four active scalars. In the case of a flat background, a solution of (\ref{17}) with $2n$ real `active scalars' has the interpretation as the (orthogonal) intersection with the worldvolume of $n$ M2-branes, corresponding to a spacetime intersection of $n+1$ M2-branes. The spacetime configuration is known to preserve the fraction $1/2^{n+1}$ of the spacetime supersymmetry \cite{mbrane} so we may expect the fraction of worldvolume supersymmetry preserved to be $1/2^n$. This can be confirmed directly from a consideration of $\kappa$-symmetry of the supermembrane \cite{gibpap,GLW}. The lump solution of (\ref{17}) for the KK-monopole background is also one with two `active scalars' and preserves half the worldvolume supersymmetry but the total number of worldvolume supersymmetries is half what it would be in a flat spacetime. The fraction of supersymmetry of the M-theory vacuum that is preserved by the total system is therefore 1/8 (1/2 for the solution, 1/2 for the brane and 1/2 for the background). We shall examine the question of supersymmetry in more detail in section 3. \subsection{Q-Kinks} We now set \begin{equation}\label{ssred} \partial_2 X^I = k^I\, , \end{equation} where $k$ is a holomorphic Killing vector field. The holomorphicity condition ensures that the dimensionally reduced 1+1 dimensional theory preserves the $N=2$ supersymmetry of the (2+1)-dimensional model. Any additional supersymmetries will be associated with additional complex structures; if $k$ is holomorphic with respect to them too then the reduction will preserve these additional supersymmetries. For the KK-monopole background we may take $k$ to be the triholomorphic Killing vector of (\ref{KV}). Using (\ref{ssred}) in (\ref{16}) we have \begin{eqnarray}\label{24} (\varepsilon+1)^2 &=& 1 + \left ( g^{IJ} + \partial X^I\partial X^J + k^I k^J\right ) P_I P_J + |\partial X|^2 + |k|^2\nonumber\\ && + \ 2 \partial X^{[I} k^{J]} \partial X^{[K} k^{L]} g_{IK} g_{JL}\, , \end{eqnarray} where $\partial X = \partial_1 X$. Restricting to static $(P=0)$ and uniform $(\partial X = 0)$ configurations yields $\epsilon = \sqrt{1 + |k|^2} - 1 \approx {1\over 2} |k|^2$, which is the membrane version of the scalar potential that leads to Q-kink solutions interpolating between its minima at fixed points of $k$ where $|k|$ vanishes. Under the same conditions as before, the expression (\ref{24}) for the energy density can be rewritten as \begin{eqnarray} (\varepsilon+1)^2 &=& \left [ 1 + v k\cdot P + \sqrt {1-v^2}\ \partial X^I k^J \omega_{IJ}\right ]^2 + \left |P - vk\right |^2\nonumber\cr &&+\ \left |\partial X^I + \sqrt {1-v^2}\ k^J I_J{}^I\right |^2 + \left (P\cdot \partial X\right )^2\cr &&+\ \left [v\ \partial X^I k^J \omega_{IJ} - \sqrt {1-v^2}\ k\cdot P\right ]^2 + \sum_r \left ( \partial X^I k^J \Omega_{IJ}^{(r)}\right )^2\, . \end{eqnarray} for arbitrary constant $v$ with $|v|<1$. We deduce that \begin{equation} \varepsilon \ge vk\cdot P + \sqrt {1-v^2}\ \partial X^I k^J \omega_{IJ}\, \end{equation} with equality when \begin{eqnarray}\label{kinkbps} P^I &=& v k^I\, ,\cr \partial X^I &=& - \sqrt {1- v^2}\ k^J I_J{}^I\, , \end{eqnarray} since these equations imply the vanishing of the remaining terms. Setting $P^I=\dot X^I$ in (\ref{kinkbps}) we recover the equations found in \cite{AT}, the solutions of which are Q-kinks. The explicit Q-kink solution of \cite{AT} was given for the two-centre metric with $V$ as in (\ref{vee}) but without the constant term (i.e. for the Eguchi-Hanson metric \cite{prasad}). The explicit solution when $V$ includes a constant term has been found by Opfermann \cite{andreas}. \section{Supersymmetry} The supermembrane is invariant under all isometries of the background. Supersymmetries correspond to Grassmann odd Killing vector superfields $\chi = \chi^A E_A$, where $E_A = E_A{}^M\partial_M$. The (Grassman even) spinor component $\chi^\alpha$ is a Killing spinor in the standard sense, at least in a purely bosonic background. The (Grassman odd) vector component $\chi^a$ is a superfield satisfying the constraint ${\cal D}_\alpha \chi^a = (\Gamma^a\chi)_\alpha$. Let $\{\chi\}$ be the complete set of these Killing vector superfields and let $\{\epsilon\}$ be a corresponding set of anticommuting parameters. The supersymmetry transformations of the worldvolume fields $Z^M$ are then \begin{equation} \delta_\epsilon Z^M = \epsilon\cdot \chi^M\, , \end{equation} where $\epsilon\cdot\chi$ is used to denote the sum over the $(\epsilon,\chi)$ pairs. Defining $\delta E^A = \delta Z^M E_M{}^A$, we then have \begin{equation} \delta_\epsilon E^A = \epsilon\cdot \chi^A\, . \end{equation} The $\kappa$-symmetry variation $\delta_\kappa Z^M$ can be similarly expressed in the form \begin{equation} \delta_\kappa E^\alpha = \kappa^\beta (1+\Gamma)_\beta{}^\alpha\, , \hskip 1truecm \delta_\kappa E^a = 0\, , \end{equation} where \begin{equation} \Gamma = {1\over 6\sqrt{-g}}\epsilon^{ijk} E_i{}^a E_j{}^b E_k{}^c \Gamma_{abc}\, , \end{equation} with $E_i{}^a = \partial_i Z^M E_M{}^a$, and $g$ is the determinant of the induced worldvolume metric $g_{ij} = E_i{}^a E_j{}^b\eta_{ab}$. To fix $\kappa$-symmetry, we make the gauge choice \cite{singleton} \begin{equation} E^\alpha (1+\Gamma)_\alpha{}^\beta = 0\, . \end{equation} This restricts only $dZ^M$, but this is sufficient. Note that this gauge choice is invariant under supersymmetry, at least in a bosonic background and for vanishing worldvolume fermions; under these conditions we may neglect the variation of $\Gamma$, while \begin{equation} \delta_\epsilon (dZ^M E_M{}^\alpha) = D(\epsilon\cdot\chi)^\alpha - \epsilon\cdot\chi^\beta E^\gamma T_{\gamma\beta}{}^\alpha \, , \end{equation} which vanishes by the Killing spinor equation (the $T_{\gamma\beta}{}^\alpha$ component of the torsion tensor is proportional to the 4-form field strength of D=11 supergravity). The remaining physical variables are such that their variations are $\delta E^\alpha (1-\Gamma)_\alpha{}^\beta$. The condition that the worldvolume configuration preserves some supersymmetry is therefore \begin{equation} \label{30} \epsilon \cdot \chi^\alpha (1-\Gamma)_\alpha{}^\beta = 0\, . \end{equation} For flat superspace, $\chi_I{}^\alpha = \delta_I{}^\alpha$, so $\epsilon\cdot\chi=\epsilon$, a constant 32-component spinor. We thus recover the flat space condition \cite{singleton} \begin{equation}\label{flat} \epsilon^\alpha (1-\Gamma)_\alpha{}^\beta = 0\, . \end{equation} More generally, we must take into account the fact that $\epsilon\cdot\chi$ is neither constant nor a spinor with 32 independent components. For the simplest backgrounds, including the KK-monopole background considered here, we have \begin{equation}\label{first} \epsilon\cdot \chi = f_\chi\, \epsilon\, , \end{equation} where $f_\chi$ is an ordinary function, and $\epsilon$ is a constant 32-component spinor satisfying \begin{equation}\label{chicon} P_\chi\, \epsilon =0\, , \end{equation} with $P_\chi$ a {\sl constant} projection matrix. For the KK-monopole background the matrix $P_\chi$ is just the product of four constant Dirac matrices, one for each of the four dimensions of the 4-metric, and it has the property (associated with the fact that this background preserves 1/2 of the spacetime supersymmetry) that ${\rm tr} P_\chi=16$. The fraction of spacetime supersymmetry preserved by the brane plus background configuration is therefore determined by the number of simultaneous solutions to (\ref{flat}) and (\ref{chicon}). Note that the function $f_\chi$ of (\ref{first}) is irrelevant to the final result. We now fix worldvolume diffeomorphisms by the `static gauge' choice \begin{equation} X^m = \left (\xi^i, X^I(\xi)\right )\, . \end{equation} With this gauge choice the condition (\ref{flat}) becomes \begin{eqnarray}\label{gaugefixed} \sqrt {-g}\, \epsilon &=& \big[\Gamma_* + \Gamma^k\partial_k X^I\Gamma_I \Gamma_* + {\textstyle{1\over 2}}\Gamma_k \epsilon^{ijk}\partial_iX^I\partial_jX^J \Gamma_{IJ} \nonumber\\ && +\ {\textstyle{1\over 6}}\varepsilon^{ijk}\partial_i X^I\partial_j X^J \partial_k X^K \Gamma_{IJK}\big]\,\epsilon\, , \end{eqnarray} where \begin{equation} \Gamma_* = \Gamma_{012}\, . \end{equation} In addition \begin{equation} g = {\rm det}\left (\eta_{ij} + {\tilde g}_{ij}\right )\, , \end{equation} where \begin{equation} {\tilde g}_{ij} = \partial_i X^I\partial_j X^J g_{IJ}\, . \end{equation} The condition (\ref{gaugefixed}) for preservation of supersymmetry can now be expanded in a power series in $\partial X$. We assume here that each term in the series must vanish separately\footnote{For a flat space background this amounts to the assumption that the worldspace is the contact set of a K{\"a}hler calibration. K{\"a}hler calibrations are only ones of relevance here, although for the M5-brane there are other calibrations for which the assumption would be false. See \cite{gibpap,GLW,QMW} for a discussion of calibrations in relation to branes.}. At zeroth order in this expansion we learn that \begin{equation} \Gamma_* \epsilon = \epsilon\, . \end{equation} Because the projector $P_\chi$ involves only the $\Gamma^I$ matrices, this equation tells us that the worldvolume vacuum preserves half the supersymmetries of the supergravity background, i.e. that the M2--brane is $1/2$ supersymmetric. At first order in the $\partial X$ expansion we learn that \begin{equation}\label{key} \Gamma^k \partial_k X^I \Gamma_I\, \epsilon = 0\, . \end{equation} This implies various higher-order identities. In particular it implies that ${\rm det}\, {\tilde g}_{ij}$ vanishes and that \begin{equation} \label{38} \eta^{im} \eta^{jn} {\tilde g}_{jm} {\tilde g}_{in} = {\textstyle{1\over 2}}(\eta^{ij}{\tilde g}_{ij})^2\, . \end{equation} Using these identities, and the constraints on $\epsilon$ quadratic and cubic in $\partial X$ that also follow from (\ref{key}), one can show that the ${\sl full}$ constraint $\Gamma\epsilon = \epsilon$ is satisfied. Thus, (\ref{key}) is the only condition (apart from $P_\chi\epsilon=0$) that we need analyse to determine the fraction of supersymmetry preserved by lump and Q-kink soliton solutions. Having found the conditions for partial preservation of worldvolume supersymmetry, we are now in a position to verify that the lump and Q-kink solitons are supersymmetric and to determine the fraction of supersymmetry they preserve. We need not discuss the lump and Q-kink cases separately because the formalism to follow will apply equally to both. For lumps we just set $v=0$ while for Q-kinks we set $\partial_2 X^I = k^I$. We begin with the observation that the equations \begin{equation} \label{39} \dot {X}^I = v\partial_2 X^I\, ,\hskip 1.0truecm \partial_1 X^I + \sqrt {1-v^2}\ \partial_2X^J\, I_J{}^I=0\, , \end{equation} imply that \begin{equation} {\tilde g} = \tilde g_{22} \times \pmatrix{v^2&0&v\cr 0&1-v^2&0\cr v&0&1}\, , \end{equation} which manifestly has vanishing determinant and solves (\ref{38}). Using (\ref{39}) in (\ref{key}) and $\Gamma_*\epsilon = \epsilon$, we have \begin{equation}\label{susycon} \partial_2 X^I\Gamma^J \left ( g_{IJ} + {\tilde \Gamma} \omega_{IJ}\right )\epsilon = 0\,, \end{equation} where \begin{equation} {\tilde\Gamma} = {1\over \sqrt{1-v^2}}\left (\Gamma^0 + v \Gamma^2 \right )\, . \end{equation} Note that $\tilde \Gamma^2=-1$. The matrices $\Gamma^I$ are space-dependent. We can write them in terms of the constant complex matrices \begin{equation} \Gamma^\alpha = \Gamma^I e_I{}^\alpha \end{equation} and their complex conjugates $\bar\Gamma^{\bar\alpha}$, where $e_I{}^\alpha$ is a complex target space vielbein (with complex conjugate $\bar e_I{}^{\bar\alpha}$) chosen such that \begin{equation} \{\Gamma^\alpha,\Gamma^\beta\}=0 \qquad \{\Gamma^\alpha,\bar\Gamma^{\bar\beta}\} = \delta^{\alpha\bar\beta}\, . \end{equation} Using the fact that $\omega_{\alpha\bar\beta}= i\delta_{\alpha\bar\beta}$ in this basis, we now have \begin{equation} \sum_{\alpha=\bar\alpha} \left[ e^\alpha \bar\Gamma^{\bar\alpha} + \bar e^{\bar\alpha} \Gamma^\alpha + i(e^\alpha\bar\Gamma^{\bar\alpha} - \bar e^{\bar\alpha}\Gamma^\alpha)\tilde \Gamma \right ]\epsilon =0 \end{equation} where \begin{equation} e^\alpha = \partial_2 X^I e_I{}^\alpha\, . \end{equation} Each term in the sum must vanish separately. This leads to a set of equations, each of which can be written in the form \begin{equation} (e\bar\Gamma + \bar e\Gamma)(1-i\tilde \Gamma[\Gamma,\bar\Gamma])\epsilon =0\, . \end{equation} It follows that {\it either} $e=0$, which effectively requires one complex worldvolume scalar to be constant, {\it or} $\epsilon$ must satisfy the constraint \begin{equation} (1-i\tilde \Gamma[\Gamma,\bar\Gamma])\epsilon =0\, , \end{equation} which reduces the fraction of supersymmetry preserved by two, unless it is already satisfied by virtue of the $P_\chi$ projection imposed by the background. We briefly discussed the flat background case in section 2.1. Solutions with $2n$ active (real) scalars preserve $1/2^n$ of the worldvolume supersymmetry and hence $1/2^{n+1}$ of the spacetime supersymmetry; their spacetime interpretation is as $n+1$ intersecting M2-branes. The computation of the fraction of worldvolume supersymmetry preserved by the finite energy lumps and Q-kinks is slightly more involved because the effects of the $P_\chi$ projection must be taken into account. However this just reduces the initial number of supersymmetries by a factor of two. The M2-brane breaks half of that and the lump and Q-kink solitons halve it again, exactly as in the flat space case. These results could be anticipated from the central charge structure of the supermembrane worldvolume superalgebra. In the KK-monopole background we would need to consider the N=4 D=3 worldvolume supersymmetry algebra. For simplicity we concentrate here on the N=8 D=3 algebra relevant to a supermembrane in a flat space background. As we are considering only scalar central charges, the supersymmetry algebra is \cite{BGT} \begin{equation} \{Q_\alpha^{\tilde I} ,Q_\beta^{\tilde J}\} = \delta^{\tilde I \tilde J}P_{\alpha\beta} + \varepsilon_{\alpha\beta}\tilde Z^{\tilde I \tilde J}\, , \end{equation} where the 8 supersymmetry charges $Q^{\tilde I}$ transform as a chiral $SO(8)$ spinor. The antisymmetric central charge matrix $\tilde Z$ has four skew eigenvalues $\zeta_k$ ($k=1,2,3,4$). The positivity of the $\{Q,Q\}$ anticommutator implies the bound \begin{equation} M^2 \ge {\rm sup} (\zeta_1,\zeta_2,\zeta_3,\zeta_4)\, . \end{equation} The fraction of worldvolume supersymmetry preserved is $2^{n-5}$ where $n$ is the number of factors of $\det \{Q,Q\}$ of the form $(M^2-\zeta)^2$ that simultaneously vanish. For example, states for which all four skew eigenvalues are equal, but non-zero, preserve half of the worldvolume supersymmetry. Note that since $\tilde I$ is a {\sl spinor} index the central charge $\tilde Z$ cannot be directly interpreted as the two-form topological charge $Z$ associated with a membrane in a given 2-plane; the relation between the two is such that equal skew-eigenvalues of $\tilde Z$ corresponds to three vanishing skew-eigenvalues of $Z$, and vice-versa. \section{IIB interpretation of Q-kinks} In the introduction we explained briefly the IIA superstring interpretation of the supermembrane lump solutions. As mentioned there, the most natural superstring interpretation of Q-kinks is in terms of IIB superstring theory. We now return to this point. It was implicit in our discussion of the Q-kink in section 2.2 that $\xi^2=\rho$ is periodically identified; otherwise we do not have a genuine compactification. Since we had already made the static gauge choice $Y^2= \rho$, it follows that we must take $Y^2$ to be an angular variable, i.e. the coordinate of the $S^1$ factor in (\ref{11m}). In fact, the Killing vector field $\partial/\partial Y^2$ can be identified as a multiple of the triholomorphic Killing vector field $\ell$ mentioned in the introduction. A standard dimensional reduction on orbits of the triholomorphic Killing vector field $k$ would imply that $Y^2$ is the only field depending on $\rho$. However, the SS reduction ansatz of (\ref{ssred}) means that $Y^2$ is not the only $\rho$-dependent worldvolume field. In fact, given (\ref{KV}), the condition (\ref{ssred}) implies that $\partial_\rho\varphi=1$, or $\varphi=\rho$ up to a constant. If we introduce the new coordinates \begin{equation} X^0 = \varphi - Y^2 \qquad \tilde Y = {1\over2}\left(Y^2 + \varphi\right) \end{equation} then the combination of the static gauge choice and the SS reduction imply that $(X^0,{\bf X})$ are $\rho$-independent while $\tilde Y=\rho$. In other words, we are wrapping the membrane on the $\tilde Y$ direction, i.e. on the $k+\ell$ cycle. We can consider this to be a non-marginal bound state of a membrane wrapped on the $k$ cycle with one wrapped on the $\ell$ cycle \cite{george}. The IIA interpretation of this bound state (with the $k$ cycle interpreted as the KK circle) was explained briefly in the introduction: a membrane wrapped on the $k$ cycle yields a IIA string in the D6-brane while a membrane wrapped on the $\ell$ cycle yields a D2-brane, so we end up with a IIA string bound to a D2-brane in a D6-brane. We now consider the IIB interpretation obtained by T-duality in the $\ell$ direction. The IIB dual of a membrane wrapped once on each of the two cycles of the torus relating the IIB theory to M-theory is a (1,1) string \cite{jhs}. In our case the (1,1) string is bound to the D5-brane that is the IIB dual of the KK-monopole. The binding is due to the fact that the D5-brane attracts (1,0) strings and is neutral to (0,1) strings. Thus, there is effectively a potential confining the (1,1) string to the D5-brane (as expected from the $V\sim k^2$ potential relevant to the IIA description; in fact, the potential is T-duality invariant \cite{lambert}). Given sufficient energy, the (1,1) string could migrate from one D5-brane to another one at some position in the transverse 4-space specified by a 4-vector. In fact the supermembrane Q-kinks discussed earlier correspond to strings which begin on one D5-brane but then jump over to another one. The charge 4-vector $(Q_0,{\bf Q})$ is just the position 4-vector of the other D5-brane, as we now explain. The triplet of K\"ahler 2-forms associated with the 4-metric (\ref{tfour}) is \begin{equation} {\mbox{\boldmath $\omega$}} = (d\varphi + {\bf A} \cdot d{\bf X}) d{\bf X} - V d{\bf X} \times d{\bf X} \end{equation} where the wedge product of forms is understood. Hence the triplet of topological charges ${\bf Q}$ is given by \begin{equation} {\bf Q} = \int i_k\mbox{\boldmath $\omega$} = \int d{\bf X} \end{equation} where the integral is over the (1,1) string worldspace. For $V$ as given in (\ref{vee}), the potential $k^2$ has minima at ${\bf X}=\pm {\bf X}_0$, so a string that starts at one minimum and ends at the other one has a 3-vector kink charge ${\bf Q} = 2{\bf X}_0$. This is the same charge as in the IIA interpretation. However, the Noether charge in the IIA interpretation becomes a fourth topological charge in the IIB interpretation (cf. \cite{lambert}). To see this it is simplest to get to the IIB theory by first compactifying on the $\ell$ cycle followed by T-duality on the $k$ cycle. This leads to the S-dual of the configuration obtained from performing these operations in the reverse order (i.e. a (1,1) string in a NS-5-brane), but the result we are aiming at is unaffected by S-duality. Having compactified on the $\ell$ cycle, T-duality on the $k$ cycle takes $\dot\varphi$ to $\partial \tilde\varphi$, where $\tilde\varphi$ is the T-dual coordinate, and hence takes the Noether charge $Q_0 = \int V^{-1} \dot\varphi$ to the topological charge \begin{equation} \tilde Q_0 = \int d\tilde\varphi\, . \end{equation} This result is to be expected from the fact that the transverse space of the IIB D5-brane is 4-dimensional. Thus, in the IIB theory the Q-kink charges $(Q_0,{\bf Q})$ become a single topological 4-vector charge $Q=(\tilde Q_0,{\bf Q})$. A configuration for which this charge is non-zero represents a (1,1) string that starts at one D5-brane and then migrates to another one positioned at some distance $|Q|$ from the first in the direction given by $Q$. \vskip .5truecm \noindent {\bf Acknowledgements} \vspace{.5truecm} One of us (PKT) would like to thank Groningen university for its hospitality, and Neil Lambert and George Papadopoulos for discussions. The work of E.B.~is supported by the European Commission TMR program ERBFMRX-CT96-0045, in which E.B.~is associated to the University of Utrecht. \vspace{.5truecm}
\section{#1}\setcounter{equation}{0}} \renewcommand{\baselinestretch}{1.3} \renewcommand{\thefootnote}{\fnsymbol{footnote}} \renewcommand{\textfraction}{0.01} \renewcommand{\topfraction}{0.99} \setcounter{topnumber}{2} \renewcommand{\bottomfraction}{0.99} \setcounter{bottomnumber}{2} \renewcommand{\floatpagefraction}{0.99} \textwidth16.cm \textheight22.5cm \topmargin-1.0cm \oddsidemargin-0.0cm \evensidemargin-0.0cm \begin{document} \author{ {\large\bf S. Bosch${}^{1}$, A.J.~Buras${}^{1}$, M. Gorbahn${}^{1}$, S. J\"ager${}^{1}$,} \\ {\large\bf M. Jamin${}^{2}$, M.E.~Lautenbacher${}^{1}$ and L. Silvestrini${}^{1}$} \\ \ \\ {\small\bf ${}^{1}$ Physik Department, Technische Universit\"at M\"unchen,} \\ {\small\bf D-85748 Garching, Germany} \\ {\small\bf ${}^{2}$ Sektion Physik, Universit\"at M\"unchen,} \\ {\small\bf Theresienstrasse 37, D-80333 M\"unchen, Germany} } \date{} \title{ {\normalsize\sf \rightline{TUM-HEP-347/99} \rightline{LMU 06/99} } \bigskip {\LARGE\bf Standard Model Confronting\\ New Results for $\varepsilon'/\varepsilon$ }} \maketitle \thispagestyle{empty} \phantom{xxx} \vspace{-6mm} \begin{abstract} We analyze the CP violating ratio $\varepsilon'/\varepsilon$ in the Standard Model in view of the new KTeV results. We review the present status of the most important non-perturbative parameters $B_6^{(1/2)}$, $B_8^{(3/2)}$, $\hat B_K$ and of the strange quark mass $m_{\rm s}$. We also briefly discuss the issues of final state interactions and renormalization scheme dependence. Updating the values of the CKM parameters, of $m_{\rm t}$ and $\Lambda_{\overline{\rm MS}}^{(4)}$ and using Gaussian errors for the experimental input and flat distributions for the theoretical parameters we find $\varepsilon'/\varepsilon$ substantially below the NA31 and KTeV data: $\varepsilon'/\varepsilon= ( 7.7^{~+6.0}_{~-3.5}) \cdot 10^{-4}$ and $\varepsilon'/\varepsilon= ( 5.2^{~+4.6}_{~-2.7}) \cdot 10^{-4}$ in the NDR and HV renormalization schemes respectively. A simple scanning of all input parameters gives on the other hand $1.05 \cdot 10^{-4} \le \varepsilon'/\varepsilon \le 28.8 \cdot 10^{-4}$ and $0.26 \cdot 10^{-4} \le \varepsilon'/\varepsilon \le 22.0 \cdot 10^{-4}$ respectively. Analyzing the dependence on various parameters we find that only for extreme values of $B_6^{(1/2)}$, $B_8^{(3/2)}$ and $m_{\rm s}$ as well as suitable values of CKM parameters and $\Lambda_{\overline{\rm MS}}^{(4)}$, the ratio $\varepsilon'/\varepsilon$ can be made consistent with data. We analyze the impact of these data on the lower bounds for ${\rm Im} V_{td}V_{ts}^*$, $Br(K_L\to\pi^0\nu\bar\nu)$, $Br(K_L\to\pi^0e^+e^-)_{\rm dir}$ and on $\tan\beta$ in the Two Higgs Doublet Model II. \end{abstract} \newpage \setcounter{page}{1} \setcounter{footnote}{0} \section{Introduction} \setcounter{equation}{0} One of the most fascinating phenomena in particle physics is the violation of CP symmetry in weak interactions. In the Standard Model CP violation is supposed to originate in a single complex phase $\delta$ in the charged current interactions of quarks \cite{KM}. This picture is consistent, within theoretical hadronic uncertainties, with CP violation in $K^0-\bar K^0$ mixing (indirect CP violation) discovered in $K_L\to\pi\pi$ decays already in 1964 \cite{CRONIN} and described by the parameter $\varepsilon$ \cite{PDG}: \begin{equation} \varepsilon=(2.280\pm 0.013)\cdot 10^{-3} \exp(i \Phi_\varepsilon), \qquad \Phi_\varepsilon\approx \frac{\pi}{4}. \label{eexp} \end{equation} It is also consistent with the recent measurement of $\sin 2\beta$ from $B\to \psi K_S$ at CDF \cite{CDF99}, although the large experimental error precludes any definite conclusion. It should be emphasized that the agreement of the Standard Model with the experimental value of $\varepsilon$ is non-trivial as $|\sin\delta|\le 1 $. Indeed in the Standard Model \begin{equation} \varepsilon= \hat B_K~ {\rm Im}\lambda_t \cdot F_\varepsilon(m_{\rm t},{\rm Re}\lambda_t) \exp(i \pi/4) \label{epsm} \end{equation} where $\hat B_K$ is a non-pertubative parameter ${\cal O}(1)$ and $\lambda_t=V_{td}V_{ts}^*$ with $V_{ij}$ being the elements of the CKM matrix \cite{KM,CAB}. The function $F_\varepsilon$ results from well known box diagrams with $W^\pm,~ t,~c,~u$ exchanges \cite{IL} and includes NLO QCD corrections \cite{BJW90,HNab}. An explicit expression for $F_\varepsilon$ can be found in (10.42) of \cite{AJBLH}. It is an increasing function of the top quark mass $m_{\rm t}$ and of ${\rm Re}\lambda_t$ . The QCD scale ($\Lambda_{\overline{\rm MS}}$) dependence of $F_\varepsilon$ is very weak. Now, ${\rm Im}\lambda_t$ is an important quantity as it plays a central role in the phenomenology of CP violation in $K$ decays and is furthermore closely related to the Jarlskog parameter $J_{CP}$ \cite{CJ}, the invariant measure of CP violation in the Standard Model: $J_{CP}=\lambda\sqrt{1-\lambda^2}\, {\rm Im}\lambda_t$ with $\lambda=0.221$ denoting one of the Wolfenstein parameters \cite{WO}. To an excellent approximation one has \begin{equation} {\rm Im}\lambda_t = |V_{ub}||V_{cb}|\sin\delta. \label{imt} \end{equation} As can be inferred from (\ref{epsm}) and (\ref{imt}) only for sufficiently large values of $|V_{ub}|$, $|V_{cb}|$, $m_{\rm t}$ and $\hat B_K$ can $\varepsilon$ in (\ref{epsm}) and consequently the indirect CP violation in the Standard Model be consistent with the one observed experimentally. It turns out that using the known values of $|V_{ub}|$, $|V_{cb}|$, $m_{\rm t}$ (see Section 3) and taking $\hat B_K=0.80\pm 0.15$ in accordance with lattice and large-N calculations (see Section 2), the experimental value of $\varepsilon$ can be reproduced in the Standard Model provided $\sin\delta \ge 0.69$. This determination of $\sin\delta$ includes constraints from $B^0_{d,s}-\bar B^0_{d,s}$ mixings. We also find \begin{equation} 1.04\cdot 10^{-4}\le{\rm Im}\lambda_t\le 1.63\cdot 10^{-4}~. \label{imte} \end{equation} It should be noticed that $\sin\delta={\cal O}(1)$ and that the extracted range for ${\rm Im}\lambda_t$ is not far from the upper limit of $1.73\cdot 10^{-4}$ following from the unitarity of the CKM matrix. It should also be emphasized that the large top quark mass plays an important role in obtaining the experimental value for $\varepsilon$. Had $m_{\rm t}$ been substantially lower than it is, the theoretical value of $\varepsilon$ would be below the experimental one. While indirect CP violation in $K_L\to\pi\pi$ reflects the fact that the mass eigenstates in the $K^0-\bar K^0$ system are not CP eigenstates, the so-called direct CP violation is realized via direct transitions between states of different CP parities: CP violation in the decay amplitude. In $K_L\to\pi\pi$ decays this type of CP violation is characterized by the parameter $\varepsilon'$. In the Standard Model one has \begin{equation} \frac{\varepsilon'}{\varepsilon}= {\rm Im}\lambda_t \cdot F_{\varepsilon'}(m_{\rm t},\Lambda_{\overline{\rm MS}}^{(4)},m_{\rm s},B_6^{(1/2)},B_8^{(3/2)},\Omega_{\eta+\eta'}) \label{epeth} \end{equation} where the function $F_{\varepsilon'}$ results from the calculation of QCD penguin and electroweak penguin diagrams. Here $B_6^{(1/2)}$ and $B_8^{(3/2)}$ are non-perturbative parameters related to the dominant QCD penguin and electroweak penguin contributions respectively, $\Lambda_{\overline{\rm MS}}^{(4)}$ is the QCD scale and $\Omega_{\eta+\eta'}$ represents isospin breaking effects. The expression (\ref{epeth}) has been obtained by calculating $\varepsilon'$ and dividing it by the experimental value of $\varepsilon$ in (\ref{eexp}) in order to be able to compare with the experimental value of $\varepsilon'/\varepsilon$. This procedure exhibits the nature of $\varepsilon'$ which representing direct CP violation is proportional to ${\rm Im}\lambda_t$. However, one could also proceed differently and ignoring the constraint (\ref{eexp}) calculate $\varepsilon'/\varepsilon$ fully in theory. In this case (\ref{epeth}) is replaced by \begin{equation} \frac{\varepsilon'}{\varepsilon}= \frac{\bar F_{\varepsilon'}(m_{\rm t},\Lambda_{\overline{\rm MS}}^{(4)},m_{\rm s},B_6^{(1/2)},B_8^{(3/2)},\Omega_{\eta+\eta'})} {\hat B_K F_\varepsilon(m_{\rm t},{\rm Re}\lambda_t)} \label{epeth1} \end{equation} where $\bar F_{\varepsilon'}=|\varepsilon_{\rm exp}|F_{\varepsilon'}$ is independent of $\varepsilon$. One should notice that ${\rm Im}\lambda_t$ cancelled out in $\varepsilon'/\varepsilon$ calculated in this manner and $\varepsilon'/\varepsilon$ is actually a function of ${\rm Re}\lambda_t$ and not of ${\rm Im}\lambda_t$. However, once the constraint (\ref{eexp}) has been taken into account (\ref{epeth1}) reduces to (\ref{epeth}). We will return to this point in Section 3. There is a long history of calculations of $\varepsilon'/\varepsilon$ in the Standard Model. The first calculation of $\varepsilon'/\varepsilon$ for $m_{\rm t} \ll M_{\rm W}$ without the inclusion of renormalization group effects can be found in \cite{EGN}. Renormalization group effects in the leading logarithmic approximation have been first presented in \cite{GW79}. For $m_{\rm t} \ll M_{\rm W}$ only QCD penguins play a substantial role. First extensive phenomenological analyses in this approximation can be found in \cite{BSS}. Over the eighties these calculations were refined through the inclusion of QED penguin effects for $m_{\rm t} \ll M_{\rm W}$ \cite{BW84,donoghueetal:86,burasgerard:87}, the inclusion of isospin breaking in the quark masses \cite{donoghueetal:86,burasgerard:87,lusignoli:89}, and through improved estimates of hadronic matrix elements in the framework of the $1/N$ approach \cite{bardeen:87}. This era of $\varepsilon'/\varepsilon$ culminated in the analyses in \cite{flynn:89,buchallaetal:90}, where QCD penguins, electroweak penguins ($\gamma$ and $Z^0$ penguins) and the relevant box diagrams were included for arbitrary top quark masses. The strong cancellation between QCD penguins and electroweak penguins for $m_t > 150~\, {\rm GeV}$ found in these papers was confirmed by other authors \cite{PW91}. During the nineties considerable progress has been made by calculating complete NLO corrections to $\varepsilon'$ \cite{BJLW1}-\cite{ROMA2}. Together with the NLO corrections to $\varepsilon$ and $B^0-\bar B^0$ mixing \cite{BJW90,HNab,Dresden}, this allowed a complete NLO analysis of $\varepsilon'/\varepsilon$ including constraints from the observed indirect CP violation ($\varepsilon$) and $B_{d,s}^0-\bar B_{d,s}^0$ mixings ($\Delta M_{d,s}$). The improved determination of the $V_{ub}$ and $V_{cb}$ elements of the CKM matrix, the improved estimates of hadronic matrix elements using the lattice approach as well as other non-perturbative approaches and in particular the determination of the top quark mass $m_{\rm t}$ had of course also an important impact on $\varepsilon'/\varepsilon$. In a crude approximation (not to be used for any serious analysis) \begin{equation}\label{ap} F_{\varepsilon'}\approx 13\cdot \left[\frac{110\, {\rm MeV}}{m_{\rm s}(2~\, {\rm GeV})}\right]^2 \left[B_6^{(1/2)}(1-\Omega_{\eta+\eta'})-0.4\cdot B_8^{(3/2)}\left(\frac{m_{\rm t}}{165\, {\rm GeV}}\right)^{2.5}\right] \left(\frac{\Lambda_{\overline{\rm MS}}^{(4)}}{340~\, {\rm MeV}}\right) \end{equation} where $\Omega_{\eta+\eta'}\approx 0.25$. This formula exhibits very clearly the dominant uncertainties in $F_{\varepsilon'}$ which reside in the values of $m_{\rm s}$, $B_6^{(1/2)}$, $B_8^{(3/2)}$, $\Lambda_{\overline{\rm MS}}^{(4)}$ and $\Omega_{\eta+\eta'}$. Because of the accurate value $m_{\rm t}(m_{\rm t})=165\pm 5~\, {\rm GeV}$, the uncertainty in $\varepsilon'/\varepsilon$ due to the top quark mass amounts only to a few percent. A more accurate formula for $F_{\varepsilon'}$ will be given in Section 2. A comparison of the formulae (\ref{epsm}) and (\ref{epeth}) reveals that the analysis of $\varepsilon$ is theoretically cleaner. Indeed, $\varepsilon$ depends on a single non-perturbative parameter $\hat B_K$, whereas $\varepsilon'/\varepsilon$ is a sensitive function of $B_6^{(1/2)}$, $B_8^{(3/2)}$, $m_{\rm s}$, $\Lambda_{\overline{\rm MS}}^{(4)}$ and $\Omega_{\eta+\eta'}$. Moreover, the partial cancellation between QCD penguin ($B_6^{(1/2)}$) and electroweak penguin ($B_8^{(3/2)}$) contributions requires accurate values of $B_6^{(1/2)}$ and $B_8^{(3/2)}$ for an acceptable estimate of $\varepsilon'/\varepsilon$. Until recently the experimental situation on $\varepsilon'/\varepsilon$ was rather unclear. While the result of the NA31 collaboration at CERN with Re$(\varepsilon'/\varepsilon) = (23.0 \pm 6.5)\cdot 10^{-4}$ \cite{barr:93} clearly indicated direct CP violation, the value of E731 at Fermilab, Re$(\varepsilon'/\varepsilon) = (7.4 \pm 5.9)\cdot 10^{-4}$ \cite{gibbons:93}, was compatible with superweak theories \cite{wolfenstein:64} in which $\varepsilon'/\varepsilon = 0$. This controversy is now settled with the very recent measurement by KTeV at Fermilab \cite{KTEV} \begin{equation}\label{eprime1} {\rm Re}(\frac{\varepsilon'}{\varepsilon}) = (28.0 \pm 4.1)\cdot 10^{-4} \end{equation} which together with the NA31 result confidently establishes direct CP violation in nature. The grand average including NA31, E731 and KTeV results reads \begin{equation} {\rm Re}(\frac{\varepsilon'}{\varepsilon}) = (21.8\pm 3.0)\cdot 10^{-4} \label{ga} \end{equation} very close to the NA31 result but with a smaller error. The error should be further reduced once the first data from NA48 collaboration at CERN are available and complete data from both collaborations have been analyzed. It is also of great interest to see what value for $\varepsilon'/\varepsilon$ will be measured by KLOE at Frascati, which uses a different experimental technique than KTeV and NA48. Does the direct CP violation observed in $K_L\to\pi\pi$ decays agree with the Standard Model expectations? Before entering the details let us take a set of ``central" values for the parameters entering $F_{\varepsilon'}$. Together with $\hat B_K=0.80$, $m_{\rm t}(m_{\rm t})=165~\, {\rm GeV}$, $|V_{ub}|=3.56\cdot 10^{-3}$ and $|V_{cb}|=0.040$ needed for the $\varepsilon$-analysis we set \begin{equation} B_6^{(1/2)}=1.0, \qquad B_8^{(3/2)}=0.8, \qquad {m_{\rm s}}(2~\, {\rm GeV})=110~\, {\rm MeV}, \qquad \Omega_{\eta+\eta'}=0.25 \label{central} \end{equation} and $\Lambda_{\overline{\rm MS}}^{(4)}=340~\, {\rm MeV}$. Using the formula (\ref{eq:3b}) for $F_{\varepsilon'}$, we find $F_{\varepsilon'}= 5.2$. On the other hand the $\varepsilon$-analysis gives ${\rm Im} \lambda_t=1.34\cdot 10^{-4}$. Consequently \begin{equation} \left(\frac{\varepsilon'}{\varepsilon}\right)^{\rm central} = 7.0 \cdot 10^{-4} \label{cth} \end{equation} well below the experimental findings in (\ref{ga}). Equivalently, with $F_{\varepsilon'}=5.2 $, the experimental value in (\ref{ga}) implies ${\rm Im}\lambda_t=(4.2\pm0.6)\cdot 10^{-4} $ which lies outside the range (\ref{imte}) extracted from the standard analysis of the unitarity triangle. Moreover it violates the upper bound ${\rm Im}\lambda_t=1.73\cdot 10^{-4}$ following from the unitarity of the CKM matrix. The fact that for central values of the input parameters the size of $\varepsilon'/\varepsilon$ in the Standard Model is well below the NA31 value of $(23.0\pm6.5)\cdot 10^{-4}$ has been known for some time. The extensive NLO analyses with lattice and large-N estimates of $B_6^{(1/2)}\approx 1$ and $B_8^{(3/2)}\approx 1$ performed first in \cite{BJLW,ROMA1} and after the top discovery in \cite{ciuchini:95}-\cite{ciuchini:96} have found $\varepsilon'/\varepsilon$ in the ball park of $(3-7)\cdot 10^{-4}$ for $m_{\rm s}(2~\, {\rm GeV})\approx 130~\, {\rm MeV}$. On the other hand it has been stressed repeatedly in \cite{AJBLH,BJL96a} that for extreme values of $B_6^{(1/2)}$, $B_8^{(3/2)}$ and $m_{\rm s}$ still consistent with lattice, QCD sum rules and large-N estimates as well as sufficiently high values of ${\rm Im}\lambda_t$ and $\Lambda_{\overline{\rm MS}}^{(4)}$, a ratio $\varepsilon'/\varepsilon$ as high as $(2-3)\cdot 10^{-3}$ could be obtained within the Standard Model. Yet, it has also been admitted that such simultaneously extreme values of all input parameters and consequently values of $\varepsilon'/\varepsilon$ close to the NA31 result are rather improbable in the Standard Model. Different conclusions have been reached in \cite{paschos:96}, where values $(1-2)\cdot 10^{-3}$ for $\varepsilon'/\varepsilon$ can be found. Also the Trieste group \cite{BERT98}, which calculated the parameters $B_6^{(1/2)}$ and $B_8^{(3/2)}$ in the chiral quark model, found $\varepsilon'/\varepsilon=(1.7\pm 1.4)\cdot 10^{-3}$. On the other hand using an effective chiral lagrangian approach, the authors in \cite{BELKOV} found $\varepsilon'/\varepsilon$ consistent with zero. The purpose of the present paper is to update the analyses in \cite{AJBLH,BJL96a} and to confront the Standard Model estimates of $\varepsilon'/\varepsilon$ with the experimental findings in (\ref{ga}). Other very recent discussions of $\varepsilon'/\varepsilon$ can be found in \cite{Nierste}-\cite{MM99}. We will comment on them below. In the present paper we address in particular the following questions: \begin{itemize} \item What is the maximal value of $\varepsilon'/\varepsilon$ in the Standard Model consistent with the usual analysis of the unitarity triangle as a function of $B_6^{(1/2)}$, $B_8^{(3/2)}$, $m_{\rm s}$ and $\Lambda_{\overline{\rm MS}}^{(4)}$ ? \item What is the lowest value of $B_6^{(1/2)}$ as a function of $B_8^{(3/2)}$ for fixed values of $m_{\rm s}$ and $\Lambda_{\overline{\rm MS}}^{(4)}$ for which the Standard Model is simultaneously compatible with (\ref{ga}) and the analysis of the unitarity triangle? \item What is the sensitivity of the analysis of $\varepsilon'/\varepsilon$ to the values of $\Omega_{\eta+\eta'}$ and $\hat B_K$? \item What is the impact of the experimental value for $\varepsilon'/\varepsilon$ on ${\rm Im}\lambda_t$, on the usual analysis of the unitarity triangle and in particular on Standard Model expectations for the rare decays $K_L\to\pi^0\nu\bar\nu$ and $K_L\to\pi^0e^+e^-$ in which direct CP violation plays an important role? \item What are the general implications of (\ref{ga}) for physics beyond the Standard Model? In particular, what is the impact on the allowed range in the space $({\rm M_H}, \tan\beta)$ in the so called two Higgs doublet model II (2HDMII) \cite{Abbott}? \end{itemize} While addressing these questions we would like to emphasize that it is by no means the purpose of our paper to fit $B_6^{(1/2)}$, $B_8^{(3/2)}$, $m_{\rm s}$, $\Lambda_{\overline{\rm MS}}^{(4)}$, $\Omega_{\eta+\eta'}$ and $\hat B_K$ in order to make the Standard Model compatible simultaneously with experimental values on $\varepsilon'/\varepsilon$, $\varepsilon$ and the analysis of the unitarity triangle. Such an approach would be against the whole philosophy of searching for new physics with the help of loop induced transitions as represented by $\varepsilon'/\varepsilon$ and $\varepsilon$. Moreover it should be kept in mind that: \begin{itemize} \item $B_6^{(1/2)}$, $B_8^{(3/2)}$ and $\hat B_K$, in spite of carrying the names of non-perturbative parameters, are really not parameters of the Standard Model as they can be calculated by means of non-perturbative methods in QCD. The same applies to $\Omega_{\eta+\eta'}$. \item $m_{\rm s}$, $\Lambda_{\overline{\rm MS}}^{(4)}$, $m_{\rm t}$, $|V_{cb}|$ and $|V_{ub}|$ are parameters of the Standard Model but there are better places than $\varepsilon'/\varepsilon$ to determine them. In particular the usual determinations of these parameters can only marginally be affected by physics beyond the Standard Model, which is not necessarily the case for $\varepsilon$ and $\varepsilon'/\varepsilon$. \end{itemize} Consequently, the only parameter to be fitted by direct CP violation is $\sin\delta$ or ${\rm Im}\lambda_t$. The numerical analysis of $\varepsilon'/\varepsilon$ as a function of $B_6^{(1/2)}$, $B_8^{(3/2)}$, $m_{\rm s}$, $\Lambda_{\overline{\rm MS}}^{(4)}$, $\Omega_{\eta+\eta'}$ and $\hat B_K$ should only give a global picture for which ranges of parameters the presence of new physics in $\varepsilon'/\varepsilon$ and $\varepsilon$ should be expected. Our paper is organized as follows. In Section 2 we recall briefly the basic formulae for $\varepsilon'/\varepsilon$ in the Standard Model. We also review the existing methods for estimating hadronic matrix elements of relevant local operators and we present a rather accurate analytic formula for $F_{\varepsilon'}$. In Section 3 we address several of the questions listed above. In Section 4 we discuss briefly general implications for physics beyond the Standard Model. In particular we investigate the lower bound on $\tan\beta$ as a function of the charged Higgs mass in the 2HDMII. Conclusions and outlook are given in Section 5. \section{Basic Formulae} \setcounter{equation}{0} \subsection{Formulae for $\varepsilon'/\varepsilon$} The parameter $\varepsilon'$ is given in terms of the isospin amplitudes $A_I$ as follows \begin{equation}\label{first} \varepsilon'=\frac{1}{\sqrt{2}}{\rm Im}\left(\frac{A_2}{A_0}\right) \exp(i\Phi_{\varepsilon'}), \qquad \Phi_{\varepsilon'}=\frac{\pi}{2}+\delta_2-\delta_0, \end{equation} where $\delta_I$ are final state interaction phases. Then, the basic formula for $\varepsilon'/\varepsilon$ is given by \begin{equation} \frac{\varepsilon'}{\varepsilon} = {\rm Im} \lambda_t\cdot F_{\varepsilon'}, \label{eq:epe1} \end{equation} where \begin{equation} F_{\varepsilon'} = \left[ P^{(1/2)} - P^{(3/2)} \right] \exp(i\Phi), \label{eq:epe2} \end{equation} with \begin{eqnarray} P^{(1/2)} & = & r \sum y_i \langle Q_i\rangle_0 (1-\Omega_{\eta+\eta'})~, \label{eq:P12} \\ P^{(3/2)} & = &\frac{r}{\omega} \sum y_i \langle Q_i\rangle_2~.~~~~~~ \label{eq:P32} \end{eqnarray} Here \begin{equation} r = \frac{G_{\rm F} \omega}{2 |\varepsilon| {\rm Re} A_0}~, \qquad \langle Q_i\rangle_I \equiv \langle (\pi\pi)_I | Q_i | K \rangle~, \qquad \omega = \frac{{\rm Re} A_2}{{\rm Re} A_0}. \label{eq:repe} \end{equation} Since \begin{equation} \Phi=\Phi_{\varepsilon'}-\Phi_\varepsilon \approx 0, \label{Phi} \end{equation} $F_{\varepsilon'}$ and $\varepsilon'/\varepsilon$ are real to an excellent approximation. The operators $Q_i$ are given explicitly as follows: {\bf Current--Current :} \begin{equation}\label{OS1} Q_1 = (\bar s_{\alpha} u_{\beta})_{V-A}\;(\bar u_{\beta} d_{\alpha})_{V-A} ~~~~~~Q_2 = (\bar s u)_{V-A}\;(\bar u d)_{V-A} \end{equation} {\bf QCD--Penguins :} \begin{equation}\label{OS2} Q_3 = (\bar s d)_{V-A}\sum_{q=u,d,s}(\bar qq)_{V-A}~~~~~~ Q_4 = (\bar s_{\alpha} d_{\beta})_{V-A}\sum_{q=u,d,s}(\bar q_{\beta} q_{\alpha})_{V-A} \end{equation} \begin{equation}\label{OS3} Q_5 = (\bar s d)_{V-A} \sum_{q=u,d,s}(\bar qq)_{V+A}~~~~~ Q_6 = (\bar s_{\alpha} d_{\beta})_{V-A}\sum_{q=u,d,s} (\bar q_{\beta} q_{\alpha})_{V+A} \end{equation} {\bf Electroweak--Penguins :} \begin{equation}\label{OS4} Q_7 = {3\over 2}\;(\bar s d)_{V-A}\sum_{q=u,d,s}e_q\;(\bar qq)_{V+A} ~~~~~ Q_8 = {3\over2}\;(\bar s_{\alpha} d_{\beta})_{V-A}\sum_{q=u,d,s}e_q (\bar q_{\beta} q_{\alpha})_{V+A} \end{equation} \begin{equation}\label{OS5} Q_9 = {3\over 2}\;(\bar s d)_{V-A}\sum_{q=u,d,s}e_q(\bar q q)_{V-A} ~~~~~Q_{10} ={3\over 2}\; (\bar s_{\alpha} d_{\beta})_{V-A}\sum_{q=u,d,s}e_q\; (\bar q_{\beta}q_{\alpha})_{V-A} \,. \end{equation} Here, $\alpha,\beta$ are colour indices and $e_q$ denotes the electric quark charges reflecting the electroweak origin of $Q_7,\ldots,Q_{10}$. The Wilson coefficient functions $ y_i(\mu)$ were calculated including the complete next-to-leading order (NLO) corrections in \cite{BJLW1}-\cite{ROMA2}. The details of these calculations can be found there and in the review \cite{BBL}. Their numerical values for $\Lambda_{\overline{\rm MS}}^{(4)}$ corresponding to $\alpha_{\overline{MS}}^{(5)}(M_{\rm Z})=0.119\pm 0.003$ and two renormalization schemes (NDR and HV) are given in table \ref{tab:wc10smu13}. There we also give the coefficients $z_{1,2}$ relevant for the discussion of hadronic matrix elements. \begin{table}[htb] \caption[]{$\Delta S=1 $ Wilson coefficients at $\mu=m_{\rm c}=1.3\, {\rm GeV}$ for $m_{\rm t}=165\, {\rm GeV}$ and $f=3$ effective flavours. $y_1 = y_2 \equiv 0$. \label{tab:wc10smu13}} \begin{center} \begin{tabular}{|c|c|c||c|c||c|c|} \hline & \multicolumn{2}{c||}{$\Lambda_{\overline{\rm MS}}^{(4)}=290\, {\rm MeV}$} & \multicolumn{2}{c||}{$\Lambda_{\overline{\rm MS}}^{(4)}=340\, {\rm MeV}$} & \multicolumn{2}{c| }{$\Lambda_{\overline{\rm MS}}^{(4)}=390\, {\rm MeV}$} \\ \hline Scheme & NDR & HV & NDR & HV & NDR & HV \\ \hline $z_1$ & --0.393 & --0.477 & --0.425 & --0.521 & --0.458 & --0.570 \\ $z_2$ & 1.201 & 1.256 & 1.222 & 1.286 & 1.244 & 1.320 \\ \hline $y_3$ & 0.027 & 0.030 & 0.030 & 0.034 & 0.033 & 0.038 \\ $y_4$ & --0.054 & --0.056 & --0.059 & --0.061 & --0.064 & --0.067 \\ $y_5$ & 0.006 & 0.015 & 0.005 & 0.016 & 0.003 & 0.017 \\ $y_6$ & --0.082 & --0.074 & --0.092 & --0.083 & --0.105 & --0.093 \\ \hline $y_7/\alpha$ & --0.038 & --0.037 & --0.037 & --0.036 & --0.037 & --0.034 \\ $y_8/\alpha$ & 0.118 & 0.127 & 0.134 & 0.143 & 0.152 & 0.161 \\ $y_9/\alpha$ & --1.410 & --1.410 & --1.437 & --1.437 & --1.466 & --1.466 \\ $y_{10}/\alpha$ & 0.496 & 0.502 & 0.539 & 0.546 & 0.585 & 0.593 \\ \hline \end{tabular} \end{center} \end{table} It is customary in phenomenological applications to take ${\rm Re} A_0$ and $\omega$ from experiment, i.e. \begin{equation} {\rm Re} A_0 = 3.33 \cdot 10^{-7}\, {\rm GeV}, \qquad \omega = 0.045, \label{eq:ReA0data} \end{equation} where the last relation reflects the so-called $\Delta I=1/2$ rule. This strategy avoids to a large extent the hadronic uncertainties in the real parts of the isospin amplitudes $A_I$. In order to be consistent the constraint (\ref{eq:ReA0data}) should also be incorporated in the matrix elements $\langle Q_i\rangle_I$ necessary for the evaluation of $\varepsilon'/\varepsilon$. This in fact has has been done in \cite{BJLW} and we will return to this approach briefly below. Studies of the $\Delta I=1/2$ rule can be found in \cite{DI12,kilcup:99,DORT99}. The sum in (\ref{eq:P12}) and (\ref{eq:P32}) runs over all contributing operators. $P^{(3/2)}$ is fully dominated by electroweak penguin contributions. $P^{(1/2)}$ on the other hand is governed by QCD penguin contributions which are suppressed by isospin breaking in the quark masses ($m_u \not= m_d$). The latter effect is described by \begin{equation} \Omega_{\eta+\eta'} = \frac{1}{\omega} \frac{({\rm Im} A_2)_{\rm I.B.}}{{\rm Im} A_0}\,. \label{eq:Omegaeta} \end{equation} For $\Omega_{\eta+\eta'}$ we will first set \begin{equation} \Omega_{\eta+\eta'} = 0.25\,, \label{eq:Omegaetadata} \end{equation} which is in the ball park of the values obtained in the $1/N$ approach \cite{burasgerard:87} and in chiral perturbation theory \cite{donoghueetal:86,lusignoli:89}. $\Omega_{\eta+\eta'}$ is independent of $m_{\rm t}$. We will investigate the sensitivity of $\varepsilon'/\varepsilon$ to $\Omega_{\eta+\eta'}$ in Section 3. \subsection{Hadronic Matrix Elements} The main source of uncertainty in the calculation of $\varepsilon'/\varepsilon$ are the hadronic matrix elements $\langle Q_i \rangle_I$. They generally depend on the renormalization scale $\mu$ and on the scheme used to renormalize the operators $Q_i$. These two dependences are canceled by those present in the Wilson coefficients $y_i(\mu)$ so that the resulting physical $\varepsilon'/\varepsilon$ does not (in principle) depend on $\mu$ and on the renormalization scheme of the operators. Unfortunately, the accuracy of the present non-perturbative methods used to evalutate $\langle Q_i \rangle_I$ is not sufficient to have the $\mu$ and scheme dependences of $\langle Q_i \rangle_I$ fully under control. We believe that this situation will change once the lattice calculations and QCD sum rule calculations improve. A brief review of the existing methods including most recent developments will be given below. In view of this situation it has been suggested in \cite{BJLW} to determine as many matrix elements $\langle Q_i \rangle_I$ as possible from the leading CP conserving $K \to \pi\pi$ decays, for which the experimental data is summarized in (\ref{eq:ReA0data}). To this end it turned out to be very convenient to determine $\langle Q_i \rangle_I$ in the three-flavour effective theory at a scale $\mu \approx m_c$. With this choice of $\mu$ the operators $Q_{1,2}^c$, being present only for $\mu>m_{\rm c}$, are integrated out and the contribution of penguin operators to ${\rm Re} A_I$ turns out to be very small. Unfortunately, since the charm mass is not much larger than the scale $M_K$ of the process we are studying, the matching procedure between the four- and three-flavour effective theories contains an ambiguity related to the choice of external momenta in the matching \cite{BJLW1,BJLW}. Furthermore, as pointed out in \cite{ciuchini:95}, there is an ambiguity due to the contribution of higher dimensional operators which are unsuppressed for $\mu \approx m_c$. However, all these ambiguities are of $O(\alpha_s)$ and one can easily verify that their possible contribution to ${\rm Re} A_I$ is at the level of a few percent at most. Consequently, they have only a minor impact on our determination of $\langle Q_i \rangle_I$ at $\mu=m_{\rm c}$ from ${\rm Re} A_I$. Using the renormalization group evolution one can then find $\langle Q_i \rangle_I$ at any other scale $\mu \not= m_{\rm c}$. The details of this procedure can be found in \cite{BJLW}. As we will see below this method allows to determine only the matrix elements of the $(V-A)\otimes(V-A)$ operators. For the central value of ${\rm Im}\lambda_t$ these operators give a negative contribution to $\varepsilon'/\varepsilon$ of about $-2.5\cdot 10^{-4}$. This shows that these operators are only relevant if $\varepsilon'/\varepsilon$ is below $1 \cdot 10^{-3}$. Unfortunately the matrix elements of the dominant $(V-A)\otimes(V+A)$ operators cannot be determined by the CP conserving data and one has to use non-perturbative methods to estimate them. Before giving the results for $\langle Q_i\rangle_I$ in our approach we would like to emphasize why it is reasonable to extract hadronic parameters from ${\rm Re} A_I$, while this would not be the case for ${\rm Im} A_I$, which govern $\varepsilon'/\varepsilon$. The point is that ${\rm Re} A_I$, in contrast to ${\rm Im} A_I$, are not expected to be affected by new physics contributions. It is customary to express the matrix elements $\langle Q_i \rangle_I$ in terms of non-perturbative parameters $B_i^{(1/2)}$ and $B_i^{(3/2)}$ as follows: \begin{equation} \langle Q_i \rangle_0 \equiv B_i^{(1/2)} \, \langle Q_i \rangle_0^{\rm (vac)}\,, \qquad \langle Q_i\rangle_2 \equiv B_i^{(3/2)} \, \langle Q_i \rangle_2^{\rm (vac)} \,. \label{eq:1} \end{equation} The label ``vac'' stands for the vacuum insertion estimate of the hadronic matrix elements in question for which $B_i^{(1/2)}=B_i^{(3/2)}=1$. Then the approach in \cite{BJLW} gives at $\mu=m_{\rm c}$: \begin{eqnarray} \langle Q_1(m_{\rm c}) \rangle_0 &=& \frac{0.187\, {\rm GeV}^{3}}{z_1(m_{\rm c})} - \frac{z_2(m_{\rm c})}{z_1(m_{\rm c})} \langle Q_2(m_{\rm c}) \rangle_0\, , \label{eq:Q10} \\ \langle Q_2(m_{\rm c}) \rangle_0 &=& \frac{5}{9} X B_2^{(1/2)} (m_{\rm c}) \, , \label{eq:Q20} \\ \langle Q_3(m_{\rm c}) \rangle_0 &=& \frac{1}{3} X B_3^{(1/2)} (m_{\rm c}) \, , \label{eq:Q30} \\ \langle Q_4(m_{\rm c}) \rangle_0 &=& \langle Q_3(m_{\rm c})\rangle_0 + \langle Q_2 (m_{\rm c}) \rangle_0 -\langle Q_1 (m_{\rm c}) \rangle_0 \, , \label{eq:Q40} \\ \langle Q_5 (m_{\rm c}) \rangle_0 &=& \frac{1}{3} B_5^{(1/2)}(m_{\rm c}) \langle \overline{Q_6(m_{\rm c})} \rangle_0 \, , \label{eq:Q50} \\ \langle Q_6 (m_{\rm c}) \rangle_0 &=& -\,4 \sqrt{\frac{3}{2}} \left[ \frac{m_{\rm K}^2}{m_{\rm s}(m_{\rm c}) + m_{\rm d}(m_{\rm c})}\right]^2 \frac{F_\pi}{\kappa} \,B_6^{(1/2)}(m_{\rm c}) \, , \label{eq:Q60} \\ \langle Q_7 (m_{\rm c}) \rangle_0 &=& - \left[ \frac{1}{6} \langle \overline{Q_6(m_{\rm c})} \rangle_0 (\kappa + 1) - \frac{X}{2} \right] B_7^{(1/2)} (m_{\rm c}) \, , \label{eq:Q70} \\ \langle Q_8 (m_{\rm c}) \rangle_0 &=& - \left[ \frac{1}{2} \langle \overline{Q_6(m_{\rm c})} \rangle_0 (\kappa + 1) - \frac{X}{6} \right] B_8^{(1/2)} (m_{\rm c}) \, , \label{eq:Q80} \\ \langle Q_9 (m_{\rm c}) \rangle_0 &=& \frac{3}{2} \langle Q_1 (m_{\rm c}) \rangle_0 - \frac{1}{2} \langle Q_3(m_{\rm c}) \rangle_0 \, , \label{eq:Q90} \\ \langle Q_{10}(m_{\rm c}) \rangle_0 &=& \langle Q_2(m_{\rm c}) \rangle_0 + \frac{1}{2} \langle Q_1(m_{\rm c}) \rangle_0 - \frac{1}{2} \langle Q_3(m_{\rm c}) \rangle_0 \, , \label{eq:Q100} \end{eqnarray} \begin{eqnarray} \langle Q_1(m_{\rm c}) \rangle_2 &=& \langle Q_2(m_{\rm c}) \rangle_2 = \frac{8.44 \cdot 10^{-3}\, {\rm GeV}^3}{z_+(m_c)} \, , \label{eq:Q122} \\ \langle Q_i \rangle_2 &=& 0 \, , \qquad i=3,\ldots,6 \, , \label{eq:Q362} \\ \langle Q_7(m_{\rm c}) \rangle_2 &=& -\left[ \frac{\kappa}{6 \sqrt{2}} \langle \overline{Q_6(m_{\rm c})} \rangle_0 + \frac{X}{\sqrt{2}} \right] B_7^{(3/2)}(m_{\rm c}) \, , \label{eq:Q72} \\ \langle Q_8(m_{\rm c}) \rangle_2 &=& -\left[ \frac{\kappa}{2 \sqrt{2}} \langle \overline{Q_6(m_{\rm c})} \rangle_0 + \frac{\sqrt{2}}{6} X \right] B_8^{(3/2)}(m_{\rm c}) \, , \label{eq:Q82} \\ \langle Q_9 (m_{\rm c}) \rangle_2 &=& \langle Q_{10}(m_{\rm c}) \rangle_2 = \frac{3}{2} \langle Q_1(m_{\rm c}) \rangle_2 \, , \label{eq:Q9102} \end{eqnarray} where \begin{equation} \kappa = \frac{F_\pi}{F_{\rm K} - F_\pi} \, , \qquad X = \sqrt{\frac{3}{2}} F_\pi \left( m_{\rm K}^2 - m_\pi^2 \right) \, , \label{eq:XQi} \end{equation} and \begin{equation} \langle \overline{Q_6(m_{\rm c})} \rangle_0 = \frac{\langle Q_6(m_{\rm c}) \rangle_0}{B_6^{(1/2)}(m_{\rm c})} \,, \qquad z_+=z_1+z_2. \label{eq:Q60bar} \end{equation} The equality of the matrix elements in (\ref{eq:Q122}) follows from isospin symmetry of strong interactions. Finally, by making the very plausible assumption, valid in known non-perturbative approaches, that $\langle Q_-(m_{\rm c}) \rangle_0 \ge \langle Q_+(m_{\rm c}) \rangle_0 \ge 0$, where $Q_\pm=(Q_2\pm Q_1)/2$, $B_2^{(1/2)} (m_{\rm c})$ can be determined as well. This gives for $\Lambda_{\overline{\rm MS}}^{(4)}=340\, {\rm MeV}$ \begin{equation} B_{2,NDR}^{(1/2)}(m_{\rm c}) = 6.5 \pm 1.0, \qquad B_{2,HV}^{(1/2)}(m_{\rm c}) = 6.1 \pm 1.0 \, . \label{eq:B122mc} \end{equation} The actual numerical values used for $m_{\rm K}$, $m_\pi$, $F_{\rm K}$, $F_\pi$ are collected in the appendix of \cite{BBL}. In particular $F_\pi=131~\, {\rm MeV}$. It should be noted that this method allows to determine not only the size but also the renormalization scheme dependence of those matrix elements which can be fixed in this manner. This dependence enters through $z_{1,2}(m_{\rm c})$ and the scheme dependence of $B_2^{(1/2)} (m_{\rm c})$. In obtaining the results above one also uses operator relations valid for $\mu\lem_{\rm c}$ which allow to express $Q_4$, $Q_9$ and $Q_{10}$ in terms of $Q_1$, $Q_2$ and $Q_3$. Theoretical issues related to these relations in the presence of NLO QCD corrections and the case of matrix elements for $\mu>m_{\rm c}$ are discussed in detail in \cite{BJLW}. In order to proceed further one has to specify the remaining $B_i$ parameters in the formulae above. As the numerical analysis in \cite{BJLW} shows $\varepsilon'/\varepsilon$ is only weakly sensitive to the values of the parameters $B_3^{(1/2)}$, $B_5^{(1/2)}$, $B_7^{(1/2)}$, $B_8^{(1/2)}$ and $B_7^{(3/2)}$ as long as their absolute values are not substantially larger than 1. As in \cite{BJLW} our strategy is to set \begin{equation} B_{3,7,8}^{(1/2)}(m_{\rm c}) = 1, \qquad B_5^{(1/2)}(m_{\rm c}) = B_6^{(1/2)}(m_{\rm c}), \qquad B_7^{(3/2)}(m_{\rm c}) = B_8^{(3/2)}(m_{\rm c}) \label{eq:B1278mc} \end{equation} and to treat $B_6^{(1/2)}(m_{\rm c})$ and $B_8^{(3/2)}(m_{\rm c})$ as free parameters. The approach in \cite{BJLW} allows then in a good approximation to express $\varepsilon'/\varepsilon$ or equivalently $F_{\varepsilon'}$ in terms of $\Lambda_{\overline{\rm MS}}^{(4)}$, $m_{\rm t}$, $m_{\rm s}$ and the two non-perturbative parameters $B_6^{(1/2)}\equiv B_6^{(1/2)}(m_{\rm c})$ and $B_8^{(3/2)}\equiv B_8^{(3/2)}(m_{\rm c})$ which cannot be fixed by the CP conserving data. \subsection{The Issue of Final State Interactions} In (\ref{first}) and (\ref{Phi}) the strong phases $\delta_0\approx 37^\circ$ and $\delta_2\approx -7^\circ$ are taken from experiment. They can also be calculated from NLO chiral perturbation for $\pi\pi$ scattering \cite{JGUM}. However, generally non-perturbative approaches to hadronic matrix elements are unable to reproduce them at present. As $\delta_I$ are factored out in (\ref{first}), in non-perturbative calculations in which some final state interactions are present in $\langle Q_i\rangle_I$ one should make the following replacements in (\ref {eq:P12}) and (\ref{eq:P32}): \begin{equation}\label{FS0} \langle Q_i\rangle_I \to \frac{{\rm Re}\langle Q_i\rangle_I} {(\cos\delta_I)_{\rm th}} \end{equation} in order to avoid double counting of final state interaction phases. Here $(\cos\delta_I)_{\rm th}$ is obtained in a given non-perturbative calculation. In leading large-N calculations and in quenched lattice calculations the phases $\delta_I$ vanish and this replacement is ineffective. When loop corrections in the large-N approach \cite{bardeen:87,DORT99,DORT98} and in the chiral quark model \cite{BERT98} are included an absorptive part and related non-vanishing phases are generated. Yet, in most calculations the phases are substantially smaller than found in experiment. For instance in the chiral quark model $(\cos\delta_0)_{\rm th}\approx 0.94$ to be compared to the experimental value $(\cos\delta_0)_{\rm exp}\approx 0.8$. Even smaller phases are found in \cite{bardeen:87,DORT99,DORT98}. The above point has been first discussed by the Trieste group \cite{BERT98} who suggested that in models in which at least the real part of $\langle Q_i\rangle_I$ can be calculated reliably, one should make the following replacements in (\ref {eq:P12}) and (\ref{eq:P32}): \begin{equation}\label{FS} \langle Q_i\rangle_I \to \frac{{\rm Re}\langle Q_i\rangle_I} {(\cos\delta_I)_{\rm exp}} \end{equation} where this time the experimental value of $\delta_I$ enters the denominator. As $(\cos\delta_0)_{\rm exp}\approx 0.8$ and $(\cos\delta_2)_{\rm exp}\approx 1$ this modification enhances $P^{(1/2)}$ by $25\%$ leaving $P^{(3/2)}$ unchanged. The same procedure has been adopted in \cite{DORT99}. To our knowledge there is no method for hadronic matrix elements which can provide $\delta_0\approx 37^\circ$ and consequently the replacement (\ref{FS}) may lead to an overestimate of the matrix elements. As in our paper the matrix elements of $(V-A)\otimes(V-A)$ operators are extracted from the data, the replacements in (\ref{FS0}) and (\ref{FS}) are ineffective for the determination of the corresponding contributions. They merely change the definition of the $B_i$ parameters in the matrix elements of $(V-A)\otimes(V-A)$ operators. The situation is different with the matrix elements of $(V-A)\otimes(V+A)$ operators which are taken from theory. Yet in view of the remarks made above, in our analysis we will use exclusively (\ref {eq:P12}) and (\ref{eq:P32}) including possible effects of this sort in the uncertainties in $B_6^{(1/2)}$ and $B_8^{(3/2)}$. \subsection{An Analytic Formula for $\varepsilon'/\varepsilon$} \label{subsec:epeanalytic} As shown in \cite{buraslauten:93}, it is possible to cast the formal expressions for $\varepsilon'/\varepsilon$ in (\ref{eq:epe1})--(\ref{eq:P32}) into an analytic formula which exhibits the $m_{\rm t}$ dependence together with the dependence on $m_{\rm s}$, $\Lambda_{\overline{\rm MS}}^{(4)}$, $B_6^{(1/2)}$ and $B_8^{(3/2)}$. To this end the approach for hadronic matrix elements presented above is used and $\Omega_{\eta+\eta'}$ is set to $0.25$. The analytic formula given below, while being rather accurate, exhibits various features which are not transparent in a pure numerical analysis. It can be used in phenomenological applications if one is satisfied with a few percent accuracy. Needless to say, in our numerical analysis in Section 3 we have used exact expressions. In this formulation the function $F_{\varepsilon'}$ is given simply as follows ($x_t=m_{\rm t}^2/M_{\rm W}^2$): \begin{equation} F_{\varepsilon'} = P_0 + P_X \, X_0(x_t) + P_Y \, Y_0(x_t) + P_Z \, Z_0(x_t) + P_E \, E_0(x_t). \label{eq:3b} \end{equation} Exact expressions for the $m_{\rm t}$-dependent functions in (\ref{eq:3b}) can be found for instance in \cite{AJBLH,BBL}. In the range $150\, {\rm GeV} \le m_{\rm t} \le 180\, {\rm GeV}$ one has to an accuracy much better than 1\% \begin{equation} X_0(x_t)=1.51~\left(\frac{m_{\rm t}}{165\, {\rm GeV}}\right)^{1.13}, \quad\quad Y_0(x_t)=0.96~\left(\frac{m_{\rm t}}{165\, {\rm GeV}}\right)^{1.55}, \end{equation} \begin{equation} Z_0(x_t)=0.66~\left(\frac{m_{\rm t}}{165\, {\rm GeV}}\right)^{1.90},\quad\quad E_0(x_t)= 0.27~\left(\frac{m_{\rm t}}{165\, {\rm GeV}}\right)^{-1.08}. \end{equation} In our numerical analysis we use exact expressions. The coefficients $P_i$ are given in terms of $B_6^{(1/2)} \equiv B_6^{(1/2)}(m_{\rm c})$, $B_8^{(3/2)} \equiv B_8^{(3/2)}(m_{\rm c})$ and $m_{\rm s}(m_{\rm c})$ as follows: \begin{equation} P_i = r_i^{(0)} + r_i^{(6)} R_6 + r_i^{(8)} R_8 \, . \label{eq:pbePi} \end{equation} where \begin{equation}\label{RS} R_6\equiv B_6^{(1/2)}\left[ \frac{137\, {\rm MeV}}{m_{\rm s}(m_{\rm c})+m_{\rm d}(m_{\rm c})} \right]^2, \qquad R_8\equiv B_8^{(3/2)}\left[ \frac{137\, {\rm MeV}}{m_{\rm s}(m_{\rm c})+m_{\rm d}(m_{\rm c})} \right]^2. \end{equation} The $P_i$ are renormalization scale and scheme independent. They depend, however, on $\Lambda_{\overline{\rm MS}}^{(4)}$. In table~\ref{tab:pbendr} we give the numerical values of $r_i^{(0)}$, $r_i^{(6)}$ and $r_i^{(8)}$ for different values of $\Lambda_{\overline{\rm MS}}^{(4)}$ at $\mu=m_{\rm c}$ in the NDR renormalization scheme. This table differs from the ones presented in \cite{AJBLH,BJL96a} in the values of $\Lambda_{\overline{\rm MS}}^{(4)}$ and the central value of $m_{\rm s}(m_{\rm c})$ in $R_s$ which has been lowered from $150\, {\rm MeV}$ to $130\, {\rm MeV}$. The coefficients $r_i^{(0)}$, $r_i^{(6)}$ and $r_i^{(8)}$ depend only very weakly on $m_{\rm s}(m_{\rm c})$ as the dominant $m_{\rm s}$ dependence has been factored out. The numbers given in table~\ref{tab:pbendr} correspond exactly to $m_{\rm s}(m_{\rm c})=130\,\, {\rm MeV}$. However, even for $m_{\rm s}(m_{\rm c})\approx100\, {\rm MeV}$ or $m_{\rm s}(m_{\rm c})\approx160\, {\rm MeV}$, the analytic expressions given here reproduce the numerical calculations of $\varepsilon'/\varepsilon$ given in Section 3 to better than $4\%$. For different scales $\mu$ the numerical values in the tables change without modifying the values of the $P_i$'s as it should be. The values of $B_6^{(1/2)}$ and $B_8^{(3/2)}$ should also be modified, in principle, but as a detailed numerical analysis in \cite{BJLW} showed, it is a good approximation to keep them $\mu$-independent for $1~\, {\rm GeV} \le\mu\le 2~\, {\rm GeV}$. We will return to this point below. \begin{table}[thb] \caption[]{Coefficients in the formula (\ref{eq:pbePi}) for various $\Lambda_{\overline{\rm MS}}^{(4)}$ in the NDR scheme. The last row gives the $r_0$ coefficients in the HV scheme. \label{tab:pbendr}} \begin{center} \begin{tabular}{|c||c|c|c||c|c|c||c|c|c|} \hline & \multicolumn{3}{c||}{$\Lambda_{\overline{\rm MS}}^{(4)}=290\, {\rm MeV}$} & \multicolumn{3}{c||}{$\Lambda_{\overline{\rm MS}}^{(4)}=340\, {\rm MeV}$} & \multicolumn{3}{c| }{$\Lambda_{\overline{\rm MS}}^{(4)}=390\, {\rm MeV}$} \\ \hline $i$ & $r_i^{(0)}$ & $r_i^{(6)}$ & $r_i^{(8)}$ & $r_i^{(0)}$ & $r_i^{(6)}$ & $r_i^{(8)}$ & $r_i^{(0)}$ & $r_i^{(6)}$ & $r_i^{(8)}$ \\ \hline 0 & --2.771 & 9.779 & 1.429 & --2.811 & 11.127 & 1.267 & --2.849 & 12.691 & 1.081 \\ $X_0$ & 0.532 & 0.017 & 0 & 0.518 & 0.021 & 0 & 0.506 & 0.024 & 0 \\ $Y_0$ & 0.396 & 0.072 & 0 & 0.381 & 0.079 & 0 & 0.367 & 0.087 & 0 \\ $Z_0$ & 0.354 & --0.013 & --9.404 & 0.409 & --0.015 & --10.230 & 0.470 & --0.017 & --11.164 \\ $E_0$ & 0.182 & --1.144 & 0.411 & 0.167 & --1.254 & 0.461 & 0.153 & --1.375 & 0.517 \\ \hline 0 & --2.749 & 8.596 & 1.050 & --2.788 & 9.638 & 0.871 & --2.825 & 10.813 & 0.669 \\ \hline \end{tabular} \end{center} \end{table} The inspection of table~\ref{tab:pbendr} shows that the terms involving $r_0^{(6)}$ and $r_Z^{(8)}$ dominate the ratio $\varepsilon'/\varepsilon$. Moreover, the function $Z_0(x_t)$ representing a gauge invariant combination of $Z^0$- and $\gamma$-penguins grows rapidly with $m_{\rm t}$ and due to $r_Z^{(8)} < 0$ these contributions suppress $\varepsilon'/\varepsilon$ strongly for large $m_{\rm t}$ \cite{flynn:89,buchallaetal:90}. \subsection{Renormalization Scheme Dependence}\label{rsd} Concerning the renormalization scheme dependence only the coefficients $r_0^{(0)}$, $r_0^{(6)}$ and $r_0^{(8)}$ are scheme dependent at the NLO level. Their values in the HV scheme are given in the last row of table~\ref{tab:pbendr}. We note that the parameter $r_0^{(0)}$ is essentially the same in both schemes as the dominant scheme independent contributions to $r_0^{(0)}$ have been determined by the data on ${\rm Re} A_I$. Since $P_0$ must be scheme independent and $r_0^{(6)}$ and $r_0^{(8)}$ are scheme dependent, we conclude that $B_6^{(1/2)}$ and $B_8^{(3/2)}$ must be scheme dependent. Indeed the matrix elements in the NDR and HV schemes are related by a finite renormalization which can be found in equation (3.7) of \cite{BJLW}. Using this equation together with the approach to matrix elements presented above, we find approximate relations between the values of $(B_6^{(1/2)},B_8^{(3/2)})$ in the NDR scheme and the corresponding values in the HV scheme: \begin{equation}\label{NDRHV} (B_6^{(1/2)})_{\rm HV}\approx 1.2 (B_6^{(1/2)})_{\rm NDR}, \qquad (B_8^{(3/2)})_{\rm HV}\approx 1.2 (B_8^{(3/2)})_{\rm NDR}. \end{equation} One can check that the scheme dependence of $(B_6^{(1/2)},B_8^{(3/2)})$ cancels to a very good approximation the one of $r_0^{(6)}$ and $r_0^{(8)}$ so that $P_0$ is scheme independent. On the other hand the coefficients $r_i$, $i=X, Y, Z, E$ are scheme independent at NLO. This is related to the fact that the $m_{\rm t}$ dependence in $\varepsilon'/\varepsilon$ enters first at the NLO level and consequently all coefficients $r_i$ in front of the $m_{\rm t}$ dependent functions must be scheme independent. Strictly speaking then the scheme dependence of $B_6^{(1/2)}$ and $B_8^{(3/2)}$ inserted into $P_i$ with $i \not= 0$ is really a part of higher order contributions to $\varepsilon'/\varepsilon$ and should be dropped at the NLO level. Formally this can be done by not performing the finite renormalization when going from the NDR to the HV scheme. Then the coefficients $P_i$ with $i \not= 0$ are clearly scheme independent. In practice the situation is more complicated. The present non-perturbative methods used to evaluate $B_6^{(1/2)}$ and $B_8^{(3/2)}$ like the large-N approach are not sensitive to the renormalization scheme dependence and we do not know which renormalization scheme the resulting values for these parameters correspond to. Lattice calculations, QCD sum rule calculations and the chiral quark model can in principle give us the scheme dependence of $B_6^{(1/2)}$ and $B_8^{(3/2)}$ but the accuracy of these methods must improve before they could be useful in this respect. In view of this situation our strategy will be to use the same values for $B_6^{(1/2)}$ and $B_8^{(3/2)}$ in the NDR and HV schemes. This will introduce a scheme dependence in $P_0$ and consequently in $\varepsilon'/\varepsilon$ but will teach us something about the uncertainty in $\varepsilon'/\varepsilon$ due to the poor sensitivity of present methods to renormalization scheme dependence. It should also be noted that even if we knew the scheme dependence of $B_6^{(1/2)}$ and $B_8^{(3/2)}$ without the ability of separating a scheme independent part in these parameters, the resulting $\varepsilon'/\varepsilon$ would be scheme dependent at the NLO level. This time the scheme dependence would enter through the scheme dependence of $P_i$ with $i \not= 0$. The latter scheme dependence could only be reduced by including the next order of perturbation theory in the Wilson coefficients: a formidable task. We should also stress \cite{BJLW} that the scheme dependences discussed here apply not only to QCD corrections but also to QED corrections. That is QED corrections to the matrix elements of operators have to be also known. For similar reasons the NLO analysis of $\varepsilon'/\varepsilon$ is still insensitive to the precise definition of $m_{\rm t}$. In view of the fact that the NLO calculations needed to extract ${\rm Im} \lambda_t$ have been performed with $m_{\rm t}=m_{\rm t}(m_{\rm t})$ we will also use this definition in calculating $F_{\varepsilon'}$. \subsection{Status of the Strange Quark Mass} At this point it seems appropriate to summarize the present status of the value of the strange quark mass. Since different methods provide $m_{\rm s}$ at different values of $\mu$ we give in table~\ref{tab:ms} a dictionary between the $m_{\rm s}$ values at $\mu=1\, {\rm GeV}$, $\mu=m_{\rm c}=1.3\, {\rm GeV}$ and $\mu=2\, {\rm GeV}$. In the case of quenched lattice QCD the present status has been summarized recently by Kenway \cite{kenway98}. Averaging the results presented by him at LATTICE 98, we obtain $m_{\rm s}(2\, {\rm GeV})=(120\pm20)\,\, {\rm MeV}$. It is expected that unquenching will lower this value but it is difficult to tell by how much. Strange quark masses as low as $m_{\rm s}(2\, {\rm GeV})=80\,\, {\rm MeV}$ have been reported in the literature \cite{LOWMS}, although the errors on unquenched calculations are still large. Lacking more precise information on unquenched lattice calculations we take as the average lattice value \begin{equation}\label{mslat} m_{\rm s}(2\, {\rm GeV})=(110\pm20)\;\, {\rm MeV}. \end{equation} which is very close to the one given by Gupta \cite{GUPTA98}. A large number of determinations of the strange quark mass from QCD sum rules exist in the literature. Historically, QCD sum rule results for $m_{\rm s}$ are given at a scale $1\,\, {\rm GeV}$. Taking an average over recent results \cite{jaminmuenz:95}-\cite{dominguezetal:98} we find $m_{\rm s}(1\, {\rm GeV})=(170\pm30)\,\, {\rm MeV}$. This translates to $m_{\rm s}(2\, {\rm GeV})=(124\pm22)\,\, {\rm MeV}$, somewhat higher than the lattice result but compatible within the errors. QCD sum rules also allow to derive lower bounds on the strange quark mass. It was found that generally $m_{\rm s}(2\, {\rm GeV})~{}_{\textstyle\sim}^{\textstyle >}~ 100\,\, {\rm MeV}$ \cite{DERAF}-\cite{Dosch}. If these bounds hold, they would rule out the very low strange mass values found in unquenched lattice QCD simulations. \begin{table}[thb] \caption[]{The dictionary between the values of $m_s(\mu)$ in units of $\, {\rm MeV}$. $\Lambda_{\overline{\rm MS}}^{(4)}=340~\, {\rm MeV}$ and $m_{\rm c}=1.3~\, {\rm GeV}$ have been used. \label{tab:ms}} \begin{center} \begin{tabular}{|c|c|c|c|c|}\hline $m_{\rm s}(m_{\rm c})$& $ 105$& $130$& $155$ & $180$ \\ \hline $m_{\rm s}(2~\, {\rm GeV})$& $ ~90$& $111$& $132$ & $154$ \\ \hline $m_{\rm s}(1~\, {\rm GeV})$& $ 123$& $152$& $181$ & $211$ \\ \hline \end{tabular} \end{center} \end{table} Finally, one should also mention the very recent determination of the strange mass from the hadronic $\tau$-spectral function \cite{PP,aleph:99} which proceeds similarly to the determination of $\alpha_s$ from $\tau$-decays. Normalized at the $\tau$ mass, the ALEPH collaboration obtains $m_{\rm s}(m_\tau)=(176^{+46}_{-57})\,\, {\rm MeV}$ which translates to $m_{\rm s}(2\, {\rm GeV})=(170^{+44}_{-55})\,\, {\rm MeV}$. We observe that the central value is much larger than the corresponding results from lattice and sum rules although the error is still large. In the future, however, improved experimental statistics and a better understanding of perturbative QCD corrections should make the determination of $m_{\rm s}$ from the $\tau$-spectral function competitive to the other methods. We conclude that the error on $m_{\rm s}$ is still rather large. In our numerical analysis of $\varepsilon'/\varepsilon$, where $m_{\rm s}$ is evaluated at the scale $m_{\rm c}$, we will set \begin{equation}\label{msvalues} m_{\rm s}(m_{\rm c})=(130\pm25)\;\, {\rm MeV} \,, \end{equation} roughly corresponding to $m_{\rm s}(2~\, {\rm GeV})$ given in (\ref{mslat}). \subsection{Review of $\hat B_K$, $B^{(1/2)}_{6}$ and $B^{(3/2)}_{8}$ } \subsubsection{$\hat B_K$} The renormalization group invariant parameter $\hat B_K$ is defined through \begin{equation} \hat B_K = B_K(\mu) \left[ \alpha_s^{(3)}(\mu) \right]^{-2/9} \, \left[ 1 + \frac{\alpha_s^{(3)}(\mu)}{4\pi} J_3 \right], \label{eq:BKrenorm} \end{equation} \begin{equation} \langle \bar K^0| (\bar s d)_{V-A} (\bar s d)_{V-A} |K^0\rangle \equiv \frac{8}{3} B_K(\mu) F_K^2 m_K^2 \label{eq:KbarK} \end{equation} where $J_3=1.895$ and $J_3=0.562$ in the NDR and HV scheme respectively. There is a long history of evaluating $\hat B_K$ in various non-perturbative approaches. The status of quenched lattice calculations \cite{JLQCD,GKS,APE} as of 1998 has been reviewed by Gupta \cite{GUPTA98}. The most accurate result for $B_K(2~\, {\rm GeV})$ using lattice methods has been obtained by the JLQCD collaboration \cite{JLQCD}: $B_K(2~\, {\rm GeV})=0.628\pm0.042$. A similar result has been published by Gupta, Kilcup and Sharpe \cite{GKS} last year. The APE collaboration \cite{APE} found $B_K(2~\, {\rm GeV})=0.66\pm0.11$ which is consistent with \cite{JLQCD,GKS}. The final lattice value given by Gupta was then \begin{equation}\label{G1} (\hat B_K)_{\rm Lattice}=0.86\pm0.06\pm0.06 \end{equation} where the second error is attributed to quenching. The corresponding result from the APE collaboration \cite{APE} was $\hat B_K=0.93\pm0.16$. The most recent global analysis of lattice data including also the UKQCD results gives \cite{LL} \begin{equation} \hat B_K=0.89\pm0.13 \end{equation} in good agreement with (\ref{G1}). In the $1/N$ approach of \cite{bardeen:87} one finds $\hat B_K=0.70\pm 0.10$ \cite{BBG0,Bijnens}. The most recent analysis in this approach with a modified matching procedure and inclusion of higher order terms in momenta gives a bigger range $0.4<\hat B_K<0.7 $ \cite{DORT99} which results from a stronger dependence on the matching scale between short and long distance contributions than found in previous calculations. It is hoped that inclusion of higher resonances in the effective low energy theory will make the dependence weaker. QCD sum rules give results around $\hat B_K=0.5-0.6$ with errors in the range $0.2-0.3$ \cite{BKQCD}. Still lower values are found using the QCD Hadronic Duality approach ($\hat B_K=0.39\pm0.10$) \cite{Prades}, the SU(3) symmetry and PCAC ($\hat B_K=1/3$) \cite{Donoghue} or chiral perturbation theory at next-to-leading order ($\hat B_K =0.42\pm 0.06$) \cite{Bruno}. However, as stressed in \cite{Bijnens,Sonoda}, SU(3) breaking effects considerably increase these values. Finally, the analysis in the chiral quark model gives a value as high as $\hat B_K=1.1\pm 0.2$ \cite{BERT97}. In our numerical analysis presented below we will use \begin{equation}\label{BKT} \hat B_K=0.80\pm 0.15 \, \end{equation} which is in the ball park of various lattice and large-N estimates. We will, however, discuss what happens if values outside this range are used. \subsubsection{General Comments on $B_6^{(1/2)}$ and $B_8^{(3/2)}$} As the different methods for the evaluation of these parameters use different values of $\mu$, it is useful to say something about their $\mu$-dependence. As seen in (\ref{eq:Q60}) and (\ref{eq:Q82}) the $\mu$-dependences of $\langle Q_6 (\mu) \rangle_0$ and $\langle Q_8 (\mu) \rangle_2$ are governed by the known $\mu$-dependence of $m_{\rm s}$ and $m_{\rm d}$ and could also in principle be present in $B_6^{(1/2)}(\mu)$ and $B_8^{(3/2)}(\mu)$. Now, as can be demonstrated in the large-N limit, the $\mu$-dependence of $1/(m_{\rm s}(\mu)+m_{\rm d}(\mu))^2$ in $\langle Q_6\rangle_0$ and $\langle Q_8\rangle_2$ is exactly cancelled in the decay amplitude by the diagonal evolution (no operator mixing) of the Wilson coefficients $y_6(\mu)$ and $y_8(\mu)$ taken in the large-N limit. An explicit demonstration of this feature is given in \cite{AJBLH}. In the large-N limit one also finds \begin{equation}\label{LN} B^{(1/2)}_{6}=B^{(3/2)}_{8}=1, \quad\quad{\rm (Large-N~Limit)}. \end{equation} The $\mu$-dependence of $B_6^{(1/2)}$ and $B_8^{(3/2)}$ for $N=3$ and in the presence of mixing with other operators has been investigated in \cite{BJLW}. This analysis shows that $B_6^{(1/2)}$ and $B_8^{(3/2)}$ depend only very weakly on $\mu$, when $\mu\ge 1~\, {\rm GeV}$. In such a numerical renormalization study the factors $B_{6}^{(1/2)}$ and $B_{8}^{(3/2)}$ have been set to unity at $\mu=m_{\rm c}$. Subsequently the evolution of the matrix elements in the range $1\, {\rm GeV} \le \mu \le 2\, {\rm GeV}$ has been calculated showing that for the NDR scheme $B_{6}^{(1/2)}$ and $B_{8}^{(3/2)}$ were $\mu$ independent within an accuracy of $2\,\%.$ The $\mu$ dependence in the HV scheme has been found to be stronger but still below $6\,\%$. Similar weak $\mu$-dependences have been found for $B_{5}^{(1/2)}$ and $B_{7}^{(3/2)}$. These findings simplify the comparison of results for $B_{5,6}^{(1/2)}$ and $B_{7,8}^{(3/2)}$ obtained by different methods. \subsubsection{$B^{(1/2)}_{6}$ and $B^{(3/2)}_{8}$ from the Lattice} The lattice calculations of $B_{5,6}^{(1/2)}$ and $B_{7,8}^{(3/2)}$ have been reviewed by Gupta \cite{GUPTA98} and the APE collaboration \cite{APE}. They are all given at $\mu=2\, {\rm GeV}$ and in the NDR scheme. The most reliable results are found for $B^{(3/2)}_{7,8}$. The ``modern" quenched estimates for these parameters are collected in table \ref{tab:317} \cite{GUPTA98}. The errors given there are purely statistical. The first three calculations use perturbative matching between lattice and continuum, the last one uses non-perturbative matching. All three groups agree within perturbative matching that $B_{7,8}^{(3/2)}$ are suppressed below unity: $B_{7}^{(3/2)}\approx 0.6$ and $B_{8}^{(3/2)}\approx 0.8$. The non-perturbative matching seems to increase these results by about $20\%$. It is important to see whether this feature will be confirmed by other groups. Concerning the lattice results for $B^{(1/2)}_{5,6}$ the situation is worse. The old results read $B^{(1/2)}_{5,6}(2~\, {\rm GeV})=1.0 \pm 0.2$ \cite{kilcup:91,sharpe:91}. More accurate estimates for $B^{(1/2)}_{6}$ have been given in \cite{kilcup:98}: $B^{(1/2)}_{6}(2~\, {\rm GeV})=0.67 \pm 0.04\pm 0.05$ (quenched) and $B^{(1/2)}_{6}(2~\, {\rm GeV})=0.76 \pm 0.03\pm0.05$ ($f=2$). However, as stressed by Gupta \cite{GUPTA98}, the systematic errors in this analysis are not really under control. A recent work of Pekurovsky and Kilcup \cite{kilcup:99}, in which $B_6^{(1/2)}$ is even found to be negative, unfortunately supports this criticism. We have to conclude that there are no solid predictions for $B^{(1/2)}_{5,6}$ from the lattice at present. \begin{table}[thb] \caption[]{ Lattice results for $B^{(3/2)}_{7,8} (2~\, {\rm GeV})$ obtained by various groups. \label{tab:317}} \begin{center} \begin{tabular}{|c|c|c|c|}\hline { Fermion type}& $B^{(3/2)}_7$& $B^{(3/2)}_8$ & Matching \\ \hline Staggered\cite{GKS}& $0.62(3)(6)$ &$0.77(4)(4)$ & 1-loop \\ Wilson\cite{G67}& $0.58(2)(7)$ &$0.81(3)(3)$ & 1-loop \\ Clover\cite{APE}& $0.58(2)$ &$0.83(2)$ & 1-loop \\ Clover\cite{APE}& $0.72(5)$ &$1.03(3)$ & Non-pert. \\ \hline \end{tabular} \end{center} \end{table} \subsubsection{$B^{(1/2)}_{6}$ and $B^{(3/2)}_{8}$ from the 1/N Approach} The 1/N approach to weak hadronic matrix elements was introduced in \cite{bardeen:87}. In this approach the 1/N expansion becomes a loop expansion in an effective meson theory. In the strict large-N limit only the tree level matrix elements of $Q_6$ and $Q_8$ contribute and one finds (\ref{LN}) while $B^{(1/2)}_{5}=B^{(3/2)}_{7}=0$. The latter fact is not disturbing, however, as the operators $Q_5$ and $Q_7$ having small Wilson coefficients are unimportant for $\varepsilon'/\varepsilon$. In view of the fact that for $B^{(1/2)}_{6}=B^{(3/2)}_{8}=1$ and the known value of $m_{\rm t}$ there is a strong cancellation between gluon and electroweak penguin contributions to $\varepsilon'/\varepsilon$, it is important to investigate whether the $1/N$ corrections significantly affect this cancellation. This has been investigated in \cite{DORT98}, where a calculation of $\langle Q_6\rangle_0$ and $\langle Q_8\rangle_2$ in the twofold expansion in powers of external momenta $p$, and in $1/N$ has been presented. The final results for $\langle Q_6\rangle_0$ and $\langle Q_8\rangle_2$ in \cite{DORT98} include the orders $p^2$ and $p^0/N$. For $\langle Q_8\rangle_2$ also the term $p^0$ contributes. Of particular interest are the ${\cal O}(p^0/N)$ contributions resulting from non-factorizable chiral loops which are important for the matching between long- and short-distance contributions. The cut-off scale $\Lambda_c$ in these non-factorizable diagrams is identified with the QCD renormalization scale $\mu$ which enters the Wilson coefficients. \begin{table}[thb] \caption[]{ Results for $B_6^{(1/2)}$ and $B_8^{(3/2)}$ obtained in the $1/N$ approach. \label{tab:318}} \begin{center} \begin{tabular}{|c||c|c|c|c|}\hline & $\Lambda_c=0.6~\, {\rm GeV}$ & $\Lambda_c=0.7~\, {\rm GeV}$ & $\Lambda_c=0.8~\, {\rm GeV}$ & $\Lambda_c=0.9~\, {\rm GeV}$ \\ \hline\hline $B_{6}^{(1/2)}$ & $1.10$ &$0.96$ & $0.84$ & $ 0.72 $ \\ & $(1.30)$ &$(1.19)$ & $(1.09)$ & $(0.99) $ \\ $B_{8}^{(3/2)}$ & $0.64$ &$0.56$ & $0.49$ & $ 0.42 $ \\ & $(0.71)$ &$(0.65)$ & $(0.59)$ & $(0.53) $ \\ \hline \end{tabular} \end{center} \end{table} In table \ref{tab:318}, taken from \cite{DORT98,DORT99}, we show the values of $B_{6}^{(1/2)}$ and $B_{8}^{(3/2)}$ as functions of the cut-off scale $\Lambda_c$. The results depend on whether $F_\pi$ or $F_K$ is used in the calculation, the difference being of higher order. The results using $F_K$ are shown in parentheses. The decrease of both B-factors with $\Lambda_c=\mu$ is qualitatively consistent with their $\mu$-dependence found for $\mu\ge 1$ in \cite{BJLW}, but it is much stronger. Clearly one could also expect a stronger $\mu$-dependence in the analysis of \cite{BJLW} for $\mu\le 1~\, {\rm GeV}$, but in view of large perturbative corrections for such small scales a meaningful test of the dependence in table \ref{tab:318} cannot be made. We note that for $\Lambda_c=0.7~\, {\rm GeV}$ the value of $B_{6}^{(1/2)}$ is close to unity as in the large-N limit. However, $B_{8}^{(3/2)}$ is considerably suppressed. An interesting feature of these results is the near $\Lambda_c$ independence of the ratio $B_6^{(1/2)}/B_8^{(3/2)}$. Consequently the results in \cite{DORT98,DORT99} can be summarized by \begin{equation} \label{DOR} \frac{B_6^{(1/2)}}{B_8^{(3/2)}} \approx 1.72~(1.84), \qquad 0.72~(0.99)\leB_6^{(1/2)}\le 1.10~(1.30)~. \end{equation} It is difficult to decide which value should be used in the phenomenology of $\varepsilon'/\varepsilon$. On the one hand, for $\Lambda_c\ge 0.7~\, {\rm GeV}$ neglected contributions from vector mesons in the loops should be included. On the other hand for $\Lambda_c=\mu= 0.6~\, {\rm GeV}$ the short distance calculations are questionable. Probably the best thing to do at present is to vary $\Lambda_c=\mu$ in the full range shown in table \ref{tab:318}. This has been done in a recent analysis \cite{DORT99b} in which $\varepsilon'/\varepsilon$ has been found to be a decreasing function of $\Lambda_c$. Finally, we would like to mention that the first non-trivial $1/N$ corrections to the matrix elements of $Q_7$ have been calculated in \cite{DER1} using the methods developed in \cite{DER2}. In particular it has been found that $B^{(3/2)}_7$ is a rather strongly increasing function of $\mu$ with negative values for $\mu\lem_{\rm c}$, $B^{(3/2)}_7(m_{\rm c})=0$ and positive values for $\mu>m_{\rm c}$. This strong $\mu$-dependence of $B^{(3/2)}_7$ is rather surprising as the numerical renormalization group analysis in \cite{BJLW} has shown a rather weak dependence of this parameter. We suspect that the inclusion of the full mixing between $Q_7$ and other operators in the analysis of \cite{DER1} would weaken the $\mu$-dependence of $B^{(3/2)}_7$ considerably. While this issue requires an additional investigation, the value of $B^{(3/2)}_7$ has fortunately only a minor impact on $\varepsilon'/\varepsilon$. Setting $B^{(3/2)}_7(m_{\rm c})=0$ instead of $B^{(3/2)}_7(m_{\rm c})=B_8^{(3/2)}(m_{\rm c})$ used here would change our results for $\varepsilon'/\varepsilon$ only by a few percent. \subsubsection{$B^{(1/2)}_{6}$ and $B^{(3/2)}_{8}$ from the Chiral Quark Model} Effective Quark Models of QCD can be derived in the framework of the extended Nambu-Jona-Lasinio model of chiral symmetry breaking \cite{NJL}. For kaon decays and in particular for $\varepsilon'/\varepsilon$, an extensive analysis of this model including chiral loops, gluon and ${\cal O}(p^4)$ corrections has been performed over the last years by the Trieste group \cite{TR96,TR97}. The crucial parameters in this approach are a mass parameter $M$ and the condensates $\langle\bar q q\rangle $ and $\langle\alpha_s GG\rangle$. They can be constrained by imposing the $\Delta I=1/2$ rule. Since there exists a nice review \cite{BERT98} by the Trieste group, we will only quote here their estimates of the relevant $B_i$ parameters. They are given in the HV scheme as follows \begin{equation}\label{TRIESTE} B^{(1/2)}_{6}=1.6\pm0.3, \quad\quad B^{(3/2)}_{8}=0.92\pm 0.02~, \quad\quad{(\rm Chiral~QM)}. \end{equation} Translating these values into the NDR scheme by means of (\ref{NDRHV}) one finds \begin{equation}\label{TRHV} B^{(1/2)}_{6}=1.33\pm0.25, \quad\quad B^{(3/2)}_{8}=0.77\pm 0.02~, \quad\quad{(\rm NDR)}. \end{equation} We observe a substantial enhancement of $B^{(1/2)}_{6}$ in the chiral quark model, not found in other calculations, and a moderate suppression of $B^{(3/2)}_{8}$. The errors given above arise from the variation of $m_{\rm s}$. We will return to this point in subsection \ref{SL}. It should be remarked that the definitions of the $B_i$ parameters used in \cite{BERT98} agree with our definitions only if in the vacuum insertion formulae in \cite{BERT98} the $\langle\bar q q\rangle$ condensate is given in terms of $m_{\rm s}$ as follows: \begin{equation} \langle\bar q q\rangle^2 =\frac{F_\pi^4}{4} \left[\frac{m_{\rm K}^2}{m_{\rm s} + m_{\rm d}}\right]^2. \end{equation} This means that in the usual PCAC relation one has to set $F_K=F_\pi$. It is interesting to observe that in this method $B_6^{(1/2)}/B_8^{(3/2)}=1.74\pm0.33$ in the ball park of (\ref{DOR}). It will be of interest to see whether future lattice calculations will confirm this correlation between $B_6^{(1/2)}$ and $B_8^{(3/2)}$. \subsubsection{$B_6^{(1/2)}$ and the $\Delta I=1/2$ Rule} In one of the first estimates of $\varepsilon'/\varepsilon$, Gilman and Wise \cite{GW79} used the suggestion of Vainshtein, Zakharov and Shifman \cite{PENGUIN} that the amplitude ${\rm Re} A_0$ is dominated by the QCD-penguin operator $Q_6$. Estimating $\langle Q_6 \rangle_0$ in this manner they predicted a large value of $\varepsilon'/\varepsilon$. Since then it has been understood \cite{DI12,kilcup:99,DORT99} that as long as the scale $\mu$ is not much lower than $1~\, {\rm GeV}$ the amplitude ${\rm Re} A_0$ is dominated by the operators $Q_1$ and $Q_2$, rather than by $Q_6$. Indeed, at least in the HV scheme the operator $Q_6$ does not contribute to ${\rm Re} A_0$ for $\mu=m_{\rm c}$ at all, as its coefficient $z_6(m_{\rm c})$ relevant for this amplitude vanishes. Also in the NDR scheme $z_6(m_{\rm c})$ is negligible. For decreasing $\mu$ the coefficient $z_6(\mu)$ increases and the $Q_6$ contribution to ${\rm Re} A_0$ is larger. However, if the analyses in \cite{DI12,kilcup:99,DORT99} are taken into account, the operators $Q_1$ and $Q_2$ are responsible for at least $90\%$ of ${\rm Re} A_0$ if the scale $\mu=1~\, {\rm GeV}$ is considered. Therefore in our opinion there is no strict relation between the large value of $\varepsilon'/\varepsilon$ and the $\Delta I=1/2$ rule as sometimes stated in the literature. Moreover, if the $90\%$ contribution of the operators $Q_1$ and $Q_2$ to ${\rm Re} A_0$ is taken into account and $z_6(1~\, {\rm GeV})$ is calculated in the NDR scheme, $B_6^{(1/2)}$ cannot exceed 1.5 if $m_{\rm s}(1~\, {\rm GeV})=150\, {\rm MeV}$. Consequently we do not think that values of $B_6^{(1/2)}$ in the NDR scheme as high as 4.0 suggested in \cite{Nierste} are plausible. Unfortunately, due to the very strong $\mu$ and renormalization scheme dependences of $z_6(\mu)$, general definite conclusions about $B_6^{(1/2)}$ cannot be reached in this manner at present. Similarly, we cannot exclude the possibility that $B_6^{(1/2)}$ is substantially higher than unity if it turned out that the present methods overestimate the role of $Q_1$ and $Q_2$ in ${\rm Re} A_0$. \subsubsection{Summary} We have seen that most non-perturbative approaches discussed above found $B_8^{(3/2)}$ below unity. The suppression of $B_8^{(3/2)}$ below unity is rather modest (at most $20\%$) in the lattice approaches and in the chiral quark model. In the $1/N$ approach $B_8^{(3/2)}$ is rather strongly suppressed and can be as low as 0.5. Concerning $B_6^{(1/2)}$ the situation is worse. As we stated above there is no solid prediction for this parameter in the lattice approach. On the other hand while the average value of $B_6^{(1/2)}$ in the $1/N$ approach is close to $1.0$, the chiral quark model gives at $\mu=0.8~\, {\rm GeV}$ and in the NDR scheme the value for $B_6^{(1/2)}$ as high as $1.33\pm 0.25$. Interestingly both approaches give the ratio $B_6^{(1/2)}/B_8^{(3/2)}$ in the ball park of 1.7. Guided by the results presented above and biased to some extent by the results from the large-N approach and lattice calculations, we will use in our numerical analysis below $B_6^{(1/2)}$ and $B_8^{(3/2)}$ in the ranges: \begin{equation}\label{bbb} B_6^{(1/2)}=1.0\pm0.3, \qquad B_8^{(3/2)}=0.8\pm 0.2 \end{equation} keeping always $B_6^{(1/2)}\ge B_8^{(3/2)}$. In our 1996 analysis \cite{BJL96a} we have used $B_6^{(1/2)}=1.0\pm0.2$ and $B_8^{(3/2)}=1.0\pm0.2$ without the constraint $B_6^{(1/2)}\ge B_8^{(3/2)}$. The decrease of $B_8^{(3/2)}$ below unity is motivated by the recent results discussed above. The increase in the range of $B_6^{(1/2)}$ is supposed to take effectively into account the uncertainty in $\Omega_{\eta+\eta'}$ which we estimate to be at most $\pm 30\%$ i.e $\Omega_{\eta+\eta'}=0.25\pm0.08$. We will return to this point in Section 3. \section{Numerical Results in the Standard Model}\label{sec:standard} \subsection{Input Parameters} In order to make predictions for $\varepsilon'/\varepsilon$ we need the value of ${\rm Im} \lambda_t$. This can be obtained from the standard analysis of the unitarity triangle which uses the data for $|V_{cb}|$, $|V_{ub}|$, $\varepsilon$, $\Delta M_d$ and $\Delta M_s$, where the last two measure the size of $B^0_{d,s}-\bar B^0_{d,s}$ mixings. Since this analysis is very well known we do not list the relevant formulae here. They can be found for instance in \cite{AJBLH,UT99}. The input parameters needed to perform the standard analysis of the unitarity triangle are given in table \ref{tab:inputparams}, where $m_{\rm t}$ refers to the running current top quark mass defined at $\mu=m_{\rm t}^{Pole}$. It corresponds to $m_{\rm t}^{Pole}=174.3\pm 5.1\, {\rm GeV}$ measured by CDF and D0 \cite{CDFD0}. We also recall that the lower bound on $\Delta M_s$ together with $\Delta M_d$ puts the following constraint on the ratio $|V_{td}|/|V_{ts}|$: \begin{equation}\label{107b} \frac{|V_{td}|}{|V_{ts}|}< \xi\sqrt{\frac{m_{B_s}}{m_{B_d}}} \sqrt{\frac{\Delta M_d}{\Delta M^{\rm min}_s}}, \qquad \xi = \frac{F_{B_s} \sqrt{B_{B_s}}}{F_{B_d} \sqrt{B_{B_d}}}. \end{equation} The range for $\Lambda_{\overline{\rm MS}}^{(4)}$ in table~\ref{tab:inputparams} corresponds roughly to $\alpha_s(M_{\rm Z})=0.119\pm 0.003$. \begin{table}[thb] \caption[]{Collection of input parameters. We impose $B_6^{(1/2)}\geB_8^{(3/2)}$. \label{tab:inputparams}} \vspace{0.4cm} \begin{center} \begin{tabular}{|c|c|c|c|} \hline {\bf Quantity} & {\bf Central} & {\bf Error} & {\bf Reference} \\ \hline $|V_{cb}|$ & 0.040 & $\pm 0.002$ & \cite{PDG} \\ $|V_{ub}|$ & $3.56\cdot 10^{-3}$ & $\pm 0.56\cdot 10^{-3} $ & \cite{STOCCHI} \\ $\hat B_K$ & 0.80 & $\pm 0.15$ & See Text \\ $\sqrt{B_d} F_{B_{d}}$ & $200\, {\rm MeV}$ & $\pm 40\, {\rm MeV}$ & \cite{BF} \\ $m_{\rm t}$ & $165\, {\rm GeV}$ & $\pm 5\, {\rm GeV}$ & \cite{CDFD0} \\ $\Delta M_d$ & $0.471~\mbox{ps}^{-1}$ & $\pm 0.016~\mbox{ps}^{-1}$ & \cite{LEPB}\\ $\Delta M_s$ & $>12.4~\mbox{ps}^{-1}$ & $ 95\% {\rm C.L.}$ & \cite{LEPB}\\ $\xi$ & $1.14$ & $\pm 0.08$ & \cite{BF} \\ $\Lambda_{\overline{\rm MS}}^{(4)}$ & $340 \, {\rm MeV}$ & $\pm 50\, {\rm MeV}$ & \cite{PDG,BETKE} \\ $m_{\rm s}(m_{\rm c})$ & $130\, {\rm MeV}$ & $\pm 25\, {\rm MeV}$ & See Text\\ $B_6^{(1/2)} $ & 1.0 & $\pm 0.3$ & See Text\\ $B_8^{(3/2)} $ & 0.8 & $\pm 0.2$ & See Text\\ \hline \end{tabular} \end{center} \end{table} \subsection{Monte Carlo and Scanning Estimates of $\varepsilon'/\varepsilon$} In what follows we will present two types of numerical analyses of ${\rm Im}\lambda_t$ and $\varepsilon'/\varepsilon$: \begin{itemize} \item Method 1: The experimentally measured numbers are used with Gaussian errors and for the theoretical input parameters we take a flat distribution in the ranges given in table~\ref{tab:inputparams}. \item Method 2: Both the experimentally measured numbers and the theoretical input parameters are scanned independently within the ranges given in table~\ref{tab:inputparams}. \end{itemize} Using the first method we find the probability density distributions for ${\rm Im}\lambda_t$ and $\varepsilon'/\varepsilon$ in figs.~\ref{g3} and \ref{g1} respectively. From the distributions in figs.~\ref{g3} and \ref{g1} we deduce the following results: \begin{equation} {\rm Im}\lambda_t =( 1.33\pm 0.14) \cdot 10^{-4} \label{eq:imfinal} \end{equation} \begin{equation} ~~~~~~\varepsilon'/\varepsilon= ( 7.7^{~+6.0}_{~-3.5}) \cdot 10^{-4}\qquad {\rm (NDR)} \label{eq:eperangefinal} \end{equation} Since the probability density in fig.~\ref{g3} is rather symmetric we give only the mean and the standard deviation for ${\rm Im}\lambda_t$. On the other hand, the resulting probability density distribution for $\epsilon'/\epsilon$ is very asymmetric with a very long tail towards large values. Therefore we decided to quote the median and the $68\%(95\%)$ confidence level intervals. This means that $68\%$ of our data can be found inside the corresponding error interval and that $50\%$ of our data has smaller $\epsilon'/\epsilon$ than our median. We observe that negative values of $\epsilon'/\epsilon$ can be excluded at $95\%$ C.L. For completeness we quote the mean and the standard deviation for $\epsilon'/\epsilon$: \begin{equation}\label{mean} \varepsilon'/\varepsilon=9.1\pm6.2 \qquad {\rm (NDR)} \end{equation} \begin{figure} \begin{center} \input{imlamt.tex} \end{center} \vspace{-6mm} \caption{Probability density distributions for ${\rm Im}\lambda_t$ without (solid line) and with (dashed line) the $\varepsilon'/\varepsilon$-constraint.} \label{g3} \end{figure} Using the second method and the parameters in table~\ref{tab:inputparams} we find : \begin{equation} 1.04 \cdot 10^{-4} \le {\rm Im}\lambda_t \le 1.63 \cdot 10^{-4} \label{eq:imnew} \end{equation} \begin{equation} ~~~~~1.05 \cdot 10^{-4} \le \varepsilon'/\varepsilon \le 28.8 \cdot 10^{-4}\qquad {\rm (NDR)}. \label{eq:eperangenew} \end{equation} The above results for $\varepsilon'/\varepsilon$ apply to the NDR scheme. $\varepsilon'/\varepsilon$ is generally lower in the HV scheme if the same values for $B_6^{(1/2)}$ and $B_8^{(3/2)}$ are used in both schemes. As discussed in subsection \ref{rsd}, such treatment of $B_6^{(1/2)}$ and $B_8^{(3/2)}$ is the proper way of estimating scheme dependences at present. Using the two error analyses we find respectively: \begin{equation} ~~~~~~\varepsilon'/\varepsilon= ( 5.2^{~+4.6}_{~-2.7}) \cdot 10^{-4}\qquad {\rm (HV)} \label{hv:final} \end{equation} and \begin{equation} ~~~~~0.26 \cdot 10^{-4} \le \varepsilon'/\varepsilon \le 22.0 \cdot 10^{-4}\qquad {\rm (HV)}. \label{hv:eperangenew} \end{equation} Moreover, the mean and the standard deviation read \begin{equation}\label{hmean} \varepsilon'/\varepsilon =6.3\pm4.8 \qquad {\rm (HV)}. \end{equation} The corresponding probability density distribution for $\varepsilon'/\varepsilon$ is compared to the one obtained in the NDR scheme in fig.~\ref{g1}. Assuming, on the other hand, that the values in (\ref{bbb}) correspond to the NDR scheme and using the relation (\ref{NDRHV}), we find for the HV scheme the range $0.58 \cdot 10^{-4} \le \varepsilon'/\varepsilon \le 26.9 \cdot 10^{-4}$ which is much closer to the NDR result in (\ref{eq:eperangenew}). This exercise shows that it is very desirable to have the scheme dependence under control. \begin{figure} \begin{center} \input{monte.tex} \end{center} \vspace{-6mm} \caption{Probability density distributions for $\varepsilon'/\varepsilon$ in NDR and HV schemes.} \label{g1} \end{figure} We observe that the most probable values for $\varepsilon'/\varepsilon$ in the NDR scheme are in the ball park of $1 \cdot 10^{-3}$. They are lower by roughly $30\%$ in the HV scheme if the same values for $(B_6^{(1/2)},B_8^{(3/2)})$ are used. On the other hand the ranges in (\ref{eq:eperangenew}) and (\ref{hv:eperangenew}) show that for particular choices of the input parameters, values for $\varepsilon'/\varepsilon$ as high as $(2-3)\cdot 10^{-3}$ cannot be excluded at present. Let us study this in more detail. \subsection{Anatomy of $\varepsilon'/\varepsilon$} \subsubsection{Global Analysis} In table~ \ref{tab:31731} we show the values of $\varepsilon'/\varepsilon$ in units of $10^{-4}$ for specific values of $B_6^{(1/2)}$, $B_8^{(3/2)}$ and $m_{\rm s}(m_{\rm c})$ as calculated in the NDR scheme. The corresponding values in the HV scheme are lower as discussed above. The fourth column shows the results for central values of all remaining parameters. The comparison of the the fourth and the fifth column demonstrates how $\varepsilon'/\varepsilon$ is increased when $\Lambda_{\overline{\rm MS}}^{(4)}$ is raised from $340~\, {\rm MeV}$ to $390~\, {\rm MeV}$. As stated in (\ref{ap}) $\varepsilon'/\varepsilon$ is roughly proportional to $\Lambda_{\overline{\rm MS}}^{(4)}$. Finally, in the last column maximal values of $\varepsilon'/\varepsilon$ are given. To this end we have scanned all parameters relevant for the analysis of ${\rm Im}\lambda_t$ within one standard deviation and have chosen the highest value of $\Lambda_{\overline{\rm MS}}^{(4)}=390\, {\rm MeV}$. Comparison of the last two columns demonstrates the impact of the increase of ${\rm Im}\lambda_t$ from its central to its maximal value and of the variation of $m_{\rm t}$. Table~\ref{tab:31731} gives a good insight in the dependence of $\varepsilon'/\varepsilon$ on various parameters which is roughly described by (\ref{ap}). We observe the following hierarchies: \begin{itemize} \item The largest uncertainties reside in $m_{\rm s}$, $B_6^{(1/2)}$ and $B_8^{(3/2)}$. $\varepsilon'/\varepsilon$ increases universally by roughly a factor of 2.3 when $m_{\rm s}(m_{\rm c})$ is changed from $155 \, {\rm MeV}$ to $105 \, {\rm MeV}$. The increase of $B_6^{(1/2)}$ from 1.0 to 1.3 increases $\varepsilon'/\varepsilon$ by $(55\pm 10)\%$, depending on $m_{\rm s}$ and $B_8^{(3/2)}$. The corresponding changes due to $B_8^{(3/2)}$ are approximately $(40\pm 15)\%$. \item The combined uncertainty due to ${\rm Im}\lambda_t$ and $m_{\rm t}$, present both in ${\rm Im}\lambda_t$ and $F_{\varepsilon'}$, is approximately $\pm 25\%$. The uncertainty due to $m_{\rm t}$ alone is only $\pm 5\%$. \item The uncertainty due to $\Lambda_{\overline{\rm MS}}^{(4)}$ is approximately $\pm 16\%$. \end{itemize} The large sensitivity of $\varepsilon'/\varepsilon$ to $m_{\rm s}$ has been known since the analyses in the eighties. In the context of the KTeV result this issue has been analyzed in \cite{Nierste}. It has been found that provided $2B_6^{(1/2)}-B_8^{(3/2)}\le 2$ the consistency of the Standard Model with the KTeV result requires the $2\sigma$ bound $m_{\rm s}(2\, {\rm GeV})\le 110\, {\rm MeV}$. Our analysis is compatible with these findings. It is of interest to investigate the impact of the relation (\ref{DOR}) on our results. Scanning all parameters in the ranges given in table~\ref{tab:inputparams} and imposing $B_6^{(1/2)}=1.7\cdot B_8^{(3/2)}$ we find \begin{equation} 3.7 \cdot 10^{-4} \le \varepsilon'/\varepsilon \le 26.2 \cdot 10^{-4} \label{eperangenew} \end{equation} which is somewhat reduced with respect to (\ref{eq:eperangenew}). Finally we would like to comment on formula (\ref{epeth1}) in which ${\rm Re}\lambda_t$ appears instead of ${\rm Im}\lambda_t$. Since $F_{\varepsilon}$ decreases with decreasing ${\rm Re}\lambda_t$ one can come closer to the experimental data for $\varepsilon'/\varepsilon$ by choosing ${\rm Re}\lambda_t$ sufficiently small. In the Wolfenstein parametrization ${\rm Re}\lambda_t$ is proportional to $1-\varrho$ and a small ${\rm Re}\lambda_t$ corresponds to a sufficiently large positive value of the parameter $\varrho$. Yet it is known from analyses of the unitarity triangle that $\varrho$ is bounded from above by the ratio $|V_{ub}/V_{cb}|$ and even stronger by the value of $\varepsilon$. If these constraints are taken into account the analysis using (\ref{epeth1}) reduces to the one presented above. \begin{table}[thb] \caption[]{ Values of $\varepsilon'/\varepsilon$ in units of $10^{-4}$ for specific values of $B_6^{(1/2)}$, $B_8^{(3/2)}$ and $m_{\rm s}(m_{\rm c})$ and other parameters as explained in the text. \label{tab:31731}} \begin{center} \begin{tabular}{|c|c|c|c|c||c|}\hline $B^{(1/2)}_6$& $B^{(3/2)}_8$ & $m_{\rm s}(m_{\rm c})[\, {\rm MeV}]$ & Central & $\Lambda_{\overline{\rm MS}}^{(4)}=390\, {\rm MeV} $ & Maximal \\ \hline & & $105$ & 20.2 & 23.3 & 28.8\\ $1.3$&$0.6$& $130$ & 12.8 & 14.8 & 18.3\\ & & $155$ & 8.5 & 9.9 & 12.3 \\ \hline & & $105$ & 18.1 & 20.8 & 26.0\\ $1.3$&$0.8$ & $130$ & 11.3 & 13.1 & 16.4\\ & & $155$ & 7.5 & 8.7 & 10.9\\ \hline & & $105$ & 15.9 & 18.3 & 23.2\\ $1.3$&$1.0$ & $130$ & 9.9 & 11.5 & 14.5\\ & & $155$ & 6.5 & 7.6 & 9.6\\ \hline\hline & & $105$ & 13.7 & 15.8 & 19.7\\ $1.0$&$0.6$ & $130$ & 8.4 & 9.8& 12.2 \\ & & $155$ & 5.4 & 6.4 & 7.9 \\ \hline & & $105$ & 11.5 & 13.3 & 16.9\\ $1.0$&$0.8$ & $130$ & 7.0 & 8.1 & 10.4\\ & & $155$ & 4.4 & 5.2 & 6.6\\ \hline & & $105$ & 9.4 & 10.9 & 14.1 \\ $1.0$&$1.0$ & $130$ & 5.5 & 6.5 & 8.5 \\ & & $155$ & 3.3 & 4.0 & 5.2\\ \hline \end{tabular} \end{center} \end{table} \subsubsection{Parametric vs. Hadronic Uncertainties} One should distinguish between parametric and hadronic uncertainties. Parametric uncertainties are related to $m_{\rm t}$, $|V_{ub}|$, $|V_{cb}|$ and $\Lambda_{\overline{\rm MS}}^{(4)}$. One should in principle include $m_{\rm s}$ in this list. However, in order to extract $m_{\rm s}$ from the kaon mass one encounters large non-perturbative uncertainties. Clearly such uncertainties are also present in the determination of $|V_{cb}|$ and in particular in the determination of $|V_{ub}|$, but they are substantially smaller. Hence the hadronic uncertainties discussed below are related to $\hat B_K$, $B_6^{(1/2)}$, $B_8^{(3/2)}$ and $m_{\rm s}$. In table~\ref{tab:3} we show ranges for $\varepsilon'/\varepsilon$ related to various uncertainties. The parametric uncertainties have been obtained for central values of $\hat B_K$ and $m_{\rm s}$ and two choices of $(B_6^{(1/2)},B_8^{(3/2)})$. The hadronic uncertainties due to $\hat B_K$, $B_6^{(1/2)}$ and $B_8^{(3/2)}$ have been found by setting all the remaining parameters at their central values. The uncertainty due to $m_{\rm s}$ has been shown for two choices of $(B_6^{(1/2)},B_8^{(3/2)})$ and all other parameters set at their central values. The last row in table~\ref{tab:3} shows the total hadronic uncertainty. It is evident from this table that hadronic uncertainties dominate, although the reduction of parametric uncertainties is very desirable. \begin{table}[thb] \caption[]{ Uncertainties in $\varepsilon'/\varepsilon$ in units of $10^{-4}$ as explained in the text. \label{tab:3}} \begin{center} \begin{tabular}{|c|c|c|c|c|}\hline Uncertainties & $B_6^{(1/2)}$ & $B^{(3/2)}_8$ & $(\varepsilon'/\varepsilon)_{\rm min}$& $(\varepsilon'/\varepsilon)_{\rm max}$ \\ \hline Parametric & 1.0 & 0.8 & 5.0 & 9.5 \\ Parametric & 1.3 & 0.8 & 8.4 & 15.1 \\ \hline Hadronic ($B_i$) & -- & -- & 3.0 & 13.6 \\ Hadronic ($m_{\rm s}$) & 1.0 & 0.8 & 4.5 & 11.3 \\ Hadronic ($m_{\rm s}$) & 1.3 & 0.8 & 7.6 & 17.9 \\ Hadronic (full ) & -- & -- & 1.7 & 21.3 \\ \hline \end{tabular} \end{center} \end{table} \subsection{$B_6^{(1/2)}$-$B_8^{(3/2)}$ Plot} In fig.~\ref{g2} we show the minimal value of $B_6^{(1/2)}$ for two choices of $m_{\rm s}(m_{\rm c})$ and $\Lambda_{\overline{\rm MS}}^{(4)}$ as a function of $B_8^{(3/2)}$ for which the theoretical value of $\varepsilon'/\varepsilon$ is higher than $2.0 \cdot 10^{-3}$. To obtain this plot we have varied all other parameters in the ranges given in table~\ref{tab:inputparams}. We show also the line corresponding to the relation (\ref{DOR}). We observe that as long as $B_8^{(3/2)}\ge0.6$, the parameter $B_6^{(1/2)}$ is required to be larger than unity. This plot should be useful when our knowledge of $B_6^{(1/2)}$, $B_8^{(3/2)}$, $m_{\rm s}$ and $\Lambda_{\overline{\rm MS}}^{(4)}$ improves. \begin{figure} \begin{center} \input{b6.tex} \end{center} \vspace{-6mm} \caption{Minimal value of $B_6^{(1/2)}$ consistent with $\varepsilon'/\varepsilon\ge 2.0\cdot 10^{-3}$.} \label{g2} \end{figure} \subsection{Approximate Scaling Laws for $\varepsilon'/\varepsilon$}\label{SL} \subsubsection{Preliminaries} Table~\ref{tab:31731} contains a lot of information on $\varepsilon'/\varepsilon$. This information can be further extended by noting that $\varepsilon'/\varepsilon$ depends to a very good approximation on certain combinations of the input parameters. This is seen in (\ref{ap}) and (\ref{eq:pbePi}). Here we want to provide scaling laws based on these formulae which allow to obtain from table~\ref{tab:31731} values for $\varepsilon'/\varepsilon$ for different sets of input parameters. \subsubsection{$B_6^{(1/2)}$, $B_8^{(3/2)}$ and $m_{\rm s}$} As seen in (\ref{eq:pbePi}), $\varepsilon'/\varepsilon$ depends on these important three parameters only through $R_6$ and $R_8$ defined in (\ref{RS}). Using this property one can for instance immediately find that the values for $\varepsilon'/\varepsilon$ in the tenth row of table~\ref{tab:31731} can also be obtained for the set \begin{equation} B_6^{(1/2)}=1.50,\qquad B_8^{(3/2)}=0.90,\qquad m_{\rm s}(m_{\rm c})=130~\, {\rm MeV}~. \end{equation} This set of parameters is similar to the input parameters used by the Trieste group \cite{BERT98}. At this point we would like to remark that in principle the determination of $B_6^{(1/2)}$ and $B_8^{(3/2)}$ in a given non-perturbative framework could depend on the value of $m_{\rm s}$. This turns out not to be the case in the large-N approach \cite{bardeen:87,DORT98,DORT99}. In the lattice approach this question has still to be investigated. On the other hand there are results in the literature showing a strong $m_s$-dependence of the $B_i$ parameters. This is the case for $B_6^{(1/2)}$ in the chiral quark model where $B_6^{(1/2)}$ scales like $m_{\rm s}$ \cite{BERT98}. Similarly values for $B_7^{(3/2)}$ calculated in \cite{DER1} show a strong $m_{\rm s}$-dependence. In the present paper we have varied $(B_6^{(1/2)},B_8^{(3/2)})$ and $m_{\rm s}$ independently which is in accordance with large-N calculations. This resulted in the following ranges for $R_6$ and $R_8$ \begin{equation} 0.5\le R_6\le 1.95, \qquad 0.4\le R_8\le 1.5 \end{equation} which are correlated through their common dependence on $m_{\rm s}$. If $(B_6^{(1/2)},B_8^{(3/2)})$ depend on $m_{\rm s}$ these ranges could change. In this context one should remark that in the chiral quark model \cite{BERT98} the highest value of $R_6$ corresponds to the minimal value of $B_6^{(1/2)}$ and consequently the comparison of the results from the chiral quark model and the large-N approach has to be made with care. Finally, it should be remarked that the decomposition of the relevant hadronic matrix elements of penguin operators into a product of $B_i$ factors times $1/m_s^2$ although useful in the $1/N$ approach will become unnecessary in the lattice approach, once matrix elements of dimension three will be calculable with improved accuracy. \subsubsection{$\Lambda_{\overline{\rm MS}}^{(4)}$ and ${\rm Im}\lambda_t$} For $\alpha_s(M_{\rm Z})=0.119\pm0.003$, the ratio $\varepsilon'/\varepsilon$ is within a few percent proportional to $\Lambda_{\overline{\rm MS}}^{(4)}$. On the other hand $\varepsilon'/\varepsilon$ is exactly proportional to ${\rm Im}\lambda_t$ at fixed $m_{\rm t}$. However, if $m_{\rm t}$ is varied the correlation in $m_{\rm t}$ between ${\rm Im}\lambda_t$ extracted from $\varepsilon$ and $F_{\varepsilon'}$ has to be taken into account. Consequently the simple rescaling of $\varepsilon'/\varepsilon$ with the values of ${\rm Im}\lambda_t$ is only true within a few percent. \subsubsection{Sensitivity to $\Omega_{\eta+\eta'}$} The dependence of $\varepsilon'/\varepsilon$ on $\Omega_{\eta+\eta'}$ can be studied numerically by using the formula (\ref{eq:P12}) or incorporated approximately into the analytic formula (\ref{eq:3b}) by simply replacing $B_6^{(1/2)}$ with an effective parameter \begin{equation}\label{eff} (B_6^{(1/2)})_{\rm eff}=B_6^{(1/2)}\frac{(1-0.9~\Omega_{\eta+\eta'})}{0.775} \end{equation} A numerical analysis shows that using $(1-\Omega_{\eta+\eta'})$ overestimates the role of $\Omega_{\eta+\eta'}$. In our numerical analysis we have incorporated the uncertainty in $\Omega_{\eta+\eta'}$ by increasing the error in $B_6^{(1/2)}$ from $\pm 0.2$ to $\pm 0.3$. The last estimates of $\Omega_{\eta+\eta'}$ have been done more than ten years ago \cite{donoghueetal:86}-\cite{lusignoli:89} and it is desirable to update these analyses which can be summarized by \begin{equation} \Omega_{\eta+\eta'}=0.25\pm 0.08~. \end{equation} The uncertainty in $\varepsilon'/\varepsilon$ due to $\Omega_{\eta+\eta'}$ alone is approximately $\pm 12\%$ and is slightly lower than the one originating from $\Lambda_{\overline{\rm MS}}^{(4)}$. \subsubsection{Sensitivity to $\hat B_K$} As ${\rm Im}\lambda_t$ extracted from $\varepsilon$ increases with decreasing $\hat B_K$, there is a possibility of increasing $\varepsilon'/\varepsilon$ by decreasing $\hat B_K$ below the range considered in table~\ref{tab:inputparams}. It should be remarked that $\varepsilon'/\varepsilon$ is not simply proportional to $1/\hat B_K$ as the extraction of ${\rm Im}\lambda_t$ from $\varepsilon$ involves also ${\rm Re}\lambda_t$ (see (\ref{epsm})). For the phase $\delta$ in the first quadrant as favoured by the analyses of the unitarity triangle \cite{UT99}, the dependence of $\varepsilon'/\varepsilon$ on $\hat B_K$ is weaker than $1/\hat B_K$ \cite{PBE0}. Now, the highest value of ${\rm Im}\lambda_t$ consistent with the unitarity of the CKM matrix is $1.73\cdot 10^{-4}$. It is obtained from $\varepsilon$ for $\hat B_K=0.52$. This increase of ${\rm Im}\lambda_t$ beyond the range in (\ref{eq:imnew}) would increase the maximal values in table~\ref{tab:31731} by approximately $6\%$. On the other hand it should be emphasized that for $\hat B_K=0.8-0.9$, as indicated by lattice calculations, $\varepsilon'/\varepsilon$ is generally smaller than found in our paper unless $B_6^{(1/2)}$ is substantially increased. This is what happens in the chiral quark model \cite{BERT98} where on the one hand $\hat B_K=1.1\pm 0.2$ and on the other hand $B_6^{(1/2)}=1.6\pm0.3$. \subsection{Impact on ${\rm Im}\lambda_t$ and the Unitarity Triangle} As we stressed at the beginning of this paper the main new parameter to be fitted by means of $\varepsilon'/\varepsilon$ is ${\rm Im}\lambda_t$. Our analysis indicates that the Standard Model estimates of $\varepsilon'/\varepsilon$ are generally below the data. If the parameters $m_{\rm s}$, $B_6^{(1/2)}$, $B_8^{(3/2)}$, $\Lambda_{\overline{\rm MS}}^{(4)}$ and $\Omega_{\eta+\eta'}$ are such that ${\rm Im}\lambda_t$ consistent with $\varepsilon$ (see (\ref{eq:imnew})) cannot accomodate the experimental value of $\varepsilon'/\varepsilon$, one has to conclude that new contributions from new physics are required. On the other hand if the data on $\varepsilon'/\varepsilon$ can be reproduced within the Standard Model, then generally a lower bound on ${\rm Im}\lambda_t$ excluding a large fraction of the range (\ref{eq:imnew}) can be obtained. Unfortunately, the strong dependence of the lower bound on the parameters involved precludes any firm conclusions. Similar comments apply to the possible impact of $\varepsilon'/\varepsilon$ on the analysis of the unitarity triangle: the presently allowed area in the $(\bar\varrho,\bar\eta)$ plane \cite{UT99} can be totally removed or an improved lower limit on $\bar\eta$ from $\varepsilon'/\varepsilon$ will decrease the allowed region considerably. As an illustration we show in table~\ref{tab:355} the lower bound on ${\rm Im}\lambda_t$ from $\varepsilon'/\varepsilon$ as a function of $B_8^{(3/2)}$ for $m_{\rm s}(m_{\rm c})=105~\, {\rm MeV}$, $B_6^{(1/2)}=1.3$ and $\Lambda_{\overline{\rm MS}}^{(4)}=390~\, {\rm MeV}$. To this end we have used the formula (\ref{eq:epe1}) with $\varepsilon'/\varepsilon\ge 2.0\cdot 10^{-3}$. Comparing with (\ref{eq:imnew}) we indeed observe that the lower bound on ${\rm Im}\lambda_t$ has been improved. The impact of $\varepsilon'/\varepsilon$-data as given in (\ref{ga}) on ${\rm Im}\lambda_t$ can also be investigated by the method 1 which was used to obtain (\ref{eq:imfinal}) and (\ref{eq:eperangefinal}). We find \begin{equation} {\rm Im}\lambda_t =( 1.38\pm0.14) \cdot 10^{-4}. \label{mfinal} \end{equation} The corresponding distribution is compared with the one without $\varepsilon'/\varepsilon$-constraint in fig.~\ref{g3}. We observe a very modest but visible shift towards higher values for ${\rm Im}\lambda_t$. \begin{table}[thb] \caption[]{ Minimal values of ${\rm Im}\lambda_t$, $Br(K_L\to\pi^0\nu\bar\nu)$ and $Br(K_L\to\pi^0 e^+e^-)_{\rm dir}$ for $m_{\rm s}(m_{\rm c})=105~\, {\rm MeV}$, $B_6^{(1/2)}=1.3$ and $\Lambda_{\overline{\rm MS}}^{(4)}=390~\, {\rm MeV}$ and specific values of $B_8^{(3/2)}$ assuming $\varepsilon'/\varepsilon\ge 2.0\cdot 10^{-3}$. \label{tab:355}} \begin{center} \begin{tabular}{|c|c|c|c|}\hline $B^{(3/2)}_8$ & $({\rm Im}\lambda_t)^{\rm min}$ & $Br(K_L\to\pi^0\nu\bar\nu)^{\rm min}$& $Br(K_L\to\pi^0e^+e^-)_{\rm dir}^{\rm min}$ \\ \hline 0.6 & $ 1.14\cdot 10^{-4}$ &$1.8\cdot 10^{-11}$ &$3.0\cdot 10^{-12}$ \\ 0.8 & $ 1.27\cdot 10^{-4}$ &$2.2\cdot 10^{-11}$ & $3.7\cdot 10^{-12}$\\ 1.0 & $ 1.42\cdot 10^{-4}$ &$2.7\cdot 10^{-11}$ &$4.7\cdot 10^{-12}$ \\ \hline \end{tabular} \end{center} \end{table} \subsection{Impact on $K_L\to\pi^0\nu\bar\nu$ and $K_L\to\pi^0e^+e^-$} The rare decay $K_L\to\pi^0\nu\bar\nu$ is the cleanest decay in the field of K-decays. It proceeds almost entirely through direct CP violation \cite{littenberg:89} and after the inclusion of NLO QCD corrections \cite{BB2} the theoretical uncertainties in the branching ratio are at the level of $1-2\%$. Similarly the contribution of direct CP-violation to the decay $K_L\to\pi^0e^+e^-$ is very clean. Using the known formulae for these decays \cite{AJBLH,BB2} and scanning the parameters given in table~\ref{tab:inputparams} we find: \begin{equation}\label{KL0} 1.6\cdot 10^{-11}\le Br(K_L\to\pi^0\nu\bar\nu)\le 3.9\cdot 10^{-11} \end{equation} \begin{equation}\label{KLE} 2.8\cdot 10^{-12}\le Br(K_L\to\pi^0 e^+e^-)_{\rm dir} \le 6.5\cdot 10^{-12} \end{equation} Since these branching ratios are proportional to $({\rm Im}\lambda_t)^2$ any impact of $\varepsilon'/\varepsilon$ on the latter CKM factor will also modify these estimates. We illustrate this in table~\ref{tab:355} where an improved lower bound on ${\rm Im}\lambda_t$ implies improved lower bounds on the branching ratios in question. With decreasing $B_6^{(1/2)}$ and increasing $m_{\rm s}$ these lower bounds continue to improve excluding a large fraction of the ranges in (\ref{KL0}) and (\ref{KLE}). In obtaining the results in table~ \ref{tab:355} correlations in $m_{\rm t}$ and the CKM parameters between $\varepsilon'/\varepsilon$, $\varepsilon$ and the branching ratios for the decays considered have been taken into account. Unfortunately, due to large hadronic uncertainties in $\varepsilon'/\varepsilon$, no strong conclusions can be reached at present. In the future the situation will be reversed. As pointed out in \cite{BB96} the cleanest measurement of ${\rm Im}\lambda_t$ is offered by $Br(K_{\rm L}\rightarrow\pi^0\nu\bar\nu)$: \begin{equation}\label{imlta} {\rm Im}\lambda_t=1.41\cdot 10^{-4} \left[\frac{165\, {\rm GeV}}{m_{\rm t} (m_{\rm t})}\right]^{1.15} \left[\frac{Br(K_{\rm L}\rightarrow\pi^0\nu\bar\nu)}{3\cdot 10^{-11}}\right]^{1/2}\,. \end{equation} Once ${\rm Im}\lambda_t$ is extracted in this manner it can be used in $\varepsilon'/\varepsilon$ thereby somewhat reducing the uncertainties in the estimate of this ratio. \section{Implications for Physics Beyond the Standard Model} \subsection{General Comments} We have seen that the Standard Model estimates of $\varepsilon'/\varepsilon$ are generally below the experimental results from NA31 and KTeV. In view of the large theoretical uncertainties it is, however, impossible at present to conclude that new physics is signaled by the $\varepsilon'/\varepsilon$-data. Still, we can make a few general comments on the extensions of the Standard Model with respect to $\varepsilon'/\varepsilon$: \begin{itemize} \item In models where the phase of the CKM matrix is the only source of CP violation, the modifications with respect to the Standard Model come through new loop contributions to $\varepsilon$ and $\varepsilon'/\varepsilon$. If the new contributions to $\varepsilon$ are positive and the contributions to $F_{\varepsilon'}$ are negative, then ${\rm Im} \lambda_t$, $F_{\varepsilon'}$ and consequently $\varepsilon'/\varepsilon$ are smaller than in the Standard Model putting these models into difficulties. An example of this disfavoured situation is the two-Higgs doublet model II in which $\varepsilon'/\varepsilon$ has been analysed a long time ago \cite{BBHLS}. We will update this analysis below. \item In the Minimal Supersymmetric Standard Model, the last analysis of $\varepsilon'/\varepsilon$ after the top quark discovery has been performed in \cite{GG95}. Here in addition to charged Higgs exchanges in loop diagrams, also charginos contribute. The chargino contribution to $\varepsilon$ has always the effect of decreasing ${\rm Im}\lambda_t$. However, depending on the choice of the supersymmetric parameters, the chargino contribution to $F_{\varepsilon'}$ can have either sign. Consequently, $\varepsilon'/\varepsilon$ in the MSSM can be enhanced with respect to the Standard Model expectations for a suitable choice of parameters, low values of chargino (stop) masses and high charged Higgs masses. Yet, as stressed in \cite{GG95}, generally $F_{\varepsilon'}$ is further suppressed by chargino contributions and the most conspicuous effect of minimal supersymmetry is a depletion of $\varepsilon'/\varepsilon$. \item The situation can be different in more general models in which there are more parameters than in the two Higgs doublet model II and in the MSSM, in particular new CP violating phases. As an example, in more general supersymmetric models $\varepsilon'/\varepsilon$ can be made consistent with experimental findings \cite{MM99,GMS}. Unfortunately, in view of the large number of free parameters such models are not very predictive. Similar comments apply to models with anomalous gauge couplings \cite{HE} and models with additional fermions and gauge bosons \cite{Frampton} in which new positive contributions to $\varepsilon'/\varepsilon$ are in principle possible. A recent discussion of new physics effects in $\varepsilon'/\varepsilon$ can also be found in \cite{Nierste}. In the past, there have of course been several other analyses of $\varepsilon'/\varepsilon$ in the extensions of the Standard Model but a review of these analyses is clearly beyond the scope of this paper. \item Finally, models with an enhanced $\bar s d Z$ vertex, considered in \cite{ISI}, can give rise to large contributions to $\varepsilon'/\varepsilon$ as pointed out in \cite{BSII}. As analyzed in the latter paper, in these models there exist interesting connections between $\varepsilon'/\varepsilon$ and rare K decays. \end{itemize} \subsection{An Update on $\varepsilon'/\varepsilon$ in the Two-Higgs Doublet Model II} A detailed renormalization group analysis of $\varepsilon'/\varepsilon$ in the Two-Higgs Doublet Model II \cite{Abbott} has been presented in \cite{BBHLS}. It has been found that due to additional {\it positive} charged Higgs contributions to $\varepsilon$ and corresponding {\it negative} contributions to $F_{\varepsilon'}$ through the increase of the importance of $Z^0$-penguin diagrams, the ratio $\varepsilon'/\varepsilon$ is suppressed with respect to the Standard Model expectations. Since this analysis goes back to 1990 and several input parameters, in particular $m_{\rm t}$, have been modified we would like to update this analysis. We recall that the two new parameters relevant for our analysis are the charged Higgs mass (${\rm M_H}$) and $\tan\beta$, the ratio of the two vacuum expectation values. The expressions for the new contributions with charged Higgs exchanges to $\varepsilon$ are rather complicated and will not be repeated here. They can be found in Section 3 of \cite{BBHLS}. The QCD corrections to these contributions are given there in the leading logarithmic approximation. As of 1999 only NLO corrections to box diagram contributions with internal top-quark exchanges to $B^0-\bar B^0$ mixing are known \cite{Dresden}. Unfortunately the NLO QCD analysis for $\varepsilon$ in the 2HDMII is still lacking. For this reason we have used the leading order expressions for the Higgs contributions to $\varepsilon$ \cite{BBHLS} except for the box diagram contributions with internal top-quark exchanges where we took the NLO QCD factor obtained in the Standard Model. While such a treatment is clearly an approximation, it is sufficient for our purposes. The analysis of $F_{\varepsilon'}$ on the other hand can be done fully at the NLO level. We only have to add to the functions $X_0(x_t)$, $Y_0(x_t)$, $Z_0(x_t)$ and $E_0(x_t)$ the contributions from charged Higgs exchanges. They are given as follows: \begin{equation} \Delta X_0=\Delta Y_0 = \frac{x_t}{\tan^2\beta} \left[\frac{y}{8(y-1)}-\frac{y}{8(x-1)^2}\log{y}\right] \end{equation} \begin{equation} \Delta Z_0= \Delta X_0 +\frac{1}{4} \frac{1}{\tan^2\beta}D_H(y) \end{equation} \begin{equation} \Delta E_0 = \frac{1}{\tan^2\beta} \left[\frac{y\left(7y^2-29y+16\right)}{36\left(y-1\right)^3}+ \frac{y\left(3y-2\right)}{6\left(y-1\right)^4} \log{y}\right] \end{equation} where \begin{equation} D_{H}(y)=\frac{y\left(47y^2-79y+38\right)}{108\left(y-1\right)^3}+ \frac{y\left(-3y^3+6y-4\right)}{18\left(y-1\right)^4} \log{y} \end{equation} with $y=m_{\rm t}^2/{\rm M_H}^2$. We observe that all new contributions to $F_{\varepsilon'}$ are inversely proportional to $\tan^2 \beta$. In $\varepsilon$ they are inversely proportional to $\tan^2 \beta$ and $\tan^4 \beta$. This should be contrasted with the case of $B\to X_s\gamma$ where there are new contributions with charged Higgs exchanges, which do not involve $\tan\beta$. Thus $\varepsilon'/\varepsilon$ is more sensitive to $\tan\beta$ than $B\to X_s\gamma$. This implies that in principle a better constraint for $\tan\beta$ could be obtained from $\varepsilon'/\varepsilon$ than from the latter decay. \begin{figure} \begin{center} \input{tbmh.tex} \end{center} \vspace{-6mm} \caption{Lower bound on $\tan\beta$ as a function of ${\rm M_H}$ consistent with $\varepsilon'/\varepsilon\ge 2.0\cdot 10^{-3}$.} \label{higgs} \end{figure} It is obvious from this discussion that $\varepsilon'/\varepsilon$ in the 2HDMII is lower than in the Standard Model for any choice of input parameters. Consequently, for low $\rm{M_H}$ and $\tan\beta$, the ratio $\varepsilon'/\varepsilon$ is generally well below the experimental data. On the other hand if the Standard Model is consistent with the experimental value of $\varepsilon'/\varepsilon$, it is possible to put a lower bound on $\tan\beta$ as a function of $\rm{M_H}$. In fig.~\ref{higgs} we show the result of such an analysis for $B_8^{(3/2)}=0.8$ and selected values of $B_6^{(1/2)}$ and $m_{\rm s}$. The remaining parameters have been scanned in the ranges given in table~\ref{tab:inputparams}. We require $\varepsilon'/\varepsilon\ge 2.0\cdot 10^{-3}$. We observe that for the lowest values of ${\rm M_H}\approx 200\, {\rm GeV}$ allowed by the $B\to X_s\gamma$ decay \cite{GAMB}-\cite{strum}, $m_{\rm s}(m_{\rm c})=105\, {\rm MeV}$ and $B_6^{(1/2)}=1.3$ the lower bound on $\tan\beta$ is similar to the one obtained from $B\to X_s\gamma$. For higher values of $m_{\rm s}(m_{\rm c})$ and lower values of $B_6^{(1/2)}$ the bound on $\tan\beta$ becomes stronger than from $B\to X_s\gamma$. \section{Conclusions and Outlook} We have presented a new analysis of $\varepsilon'/\varepsilon$ in the Standard Model in view of the recent KTeV measurement of this ratio, which together with the previous NA31 result firmly establishes direct CP violation in nature. Compared with our 1996 analysis \cite{BJL96a}, the present analysis uses improved values of $|V_{ub}|$, $|V_{cb}|$, $m_{\rm t}$, $\Lambda_{\overline{\rm MS}}^{(4)}$ and $m_{\rm s}$ as well as new insights in the hadronic parameters $B_6^{(1/2)}$, $B_8^{(3/2)}$ and $\hat B_K$. Our findings are as follows: \begin{itemize} \item The estimates of $\varepsilon'/\varepsilon$ in the Standard Model are typically below the experimental data. Our Monte Carlo analysis gives \begin{equation}\label{eperangefinal} \varepsilon'/\varepsilon =\left\{ \begin{array}{ll} ( 7.7^{~+6.0}_{~-3.5}) \cdot 10^{-4} & {\rm (NDR)} \\ ( 5.2^{~+4.6}_{~-2.7}) \cdot 10^{-4} & {\rm (HV)} \end{array} \right. \end{equation} The difference between these two results indicates the left over renormalization scheme dependence. \item On the other hand a simple scanning of all input parameters gives \begin{equation} 1.05 \cdot 10^{-4} \le \varepsilon'/\varepsilon \le 28.8 \cdot 10^{-4}\qquad {\rm (NDR)} \label{eper} \end{equation} and \begin{equation} 0.26 \cdot 10^{-4} \le \varepsilon'/\varepsilon \le 22.0 \cdot 10^{-4}\qquad {\rm (HV)} \label{eperh} \end{equation} This means that for suitably chosen parameters, $\varepsilon'/\varepsilon$ in the Standard Model can be made consistent with data. However, this happens only if all relevant parameters are simultaneously close to their extreme values. This is clearly seen in table~\ref{tab:31731} and fig.~\ref{g2}. Moreover, the probability density distributions for $\varepsilon'/\varepsilon$ in fig.~\ref{g1} indicates that values of $\varepsilon'/\varepsilon$ in the ball park of NA31 and KTeV results are rather improbable. \item Unfortunately, in view of very large hadronic and substantial parametric uncertainties, it is impossible to conclude at present whether new physics contributions are indeed required to fit the data. Similarly it is difficult to conclude what is precisely the impact of the $\varepsilon'/\varepsilon$-data on the CKM matrix. However, there are indications as seen in table~\ref{tab:355} that the lower limit on ${\rm Im}\lambda_t$ is improved. The same applies to the lower limits for the branching ratios for $K_L\to\pi^0\nu\bar\nu$ and $K_L\to\pi^0 e^+ e^-$ decays. \item Finally, we have pointed out that the $\varepsilon'/\varepsilon$ data puts models in which there are new positive contributions to $\varepsilon$ and negative contibutions to $\varepsilon'$ in serious difficulties. In particular we have analyzed $\varepsilon'/\varepsilon$ in the 2HDMII demonstrating that with improved hadronic matrix elements this model can either be ruled out or a powerful lower bound on $\tan\beta$ can be obtained from $\varepsilon'/\varepsilon$. \end{itemize} The fact that one cannot firmly conclude at present that the data for $\varepsilon'/\varepsilon$ requires new physics is rather unfortunate. In an analogous situation in the very clean rare decays $K\to\pi\nu\bar\nu$ a departure of the experimental result from the Standard Model expectations by only $30\%$ would give a clear signal for new physics. This will indeed be the case if the improved measurements of the $Br(K^+\to\pi^+\nu\bar\nu)$ from BNL787 collaboration at Brookhaven \cite{Adler} find this branching ratio above $1.5 \cdot 10^{-10}$. All efforts should be made to measure this branching ratio and the branching ratio for $K_L\to\pi^0\nu\bar\nu$, which while being directly CP violating is almost free of theoretical uncertainties. The future of $\varepsilon'/\varepsilon$ in the Standard Model and in its extensions depends on the progress in the reduction of parametric and hadronic uncertainties. We have analyzed these uncertainties in detail in Section 3 with the results given in table~\ref{tab:3}. Concerning parametric uncertainties related to $|V_{ub}|$, $|V_{cb}|$, $m_{\rm t}$ and $\Lambda_{\overline{\rm MS}}^{(4)}$, we expect that they should be reduced considerably in the coming years. This will, however, result only in a modest reduction of the total uncertainty in $\varepsilon'/\varepsilon$. In this respect a measurement of ${\rm Im}\lambda_t$ in a very clean decay like $K_L\to\pi^0\nu\bar\nu$ would be very useful. A real progress in estimating $\varepsilon'/\varepsilon$ will only be made if the non-perturbative parameters $\hat B_K$, $B_6^{(1/2)}$, $B_8^{(3/2)}$ and $\Omega_{\eta+\eta'}$ as well as the strange quark mass $m_{\rm s}$ will be brought under control. In particular the sensitivity of non-perturbative methods to $\mu$ and renormalization scheme dependences of $B_6^{(1/2)}$ and $B_8^{(3/2)}$ is clearly desirable. We expect that considerable progress on $\hat B_K$ and $B_8^{(3/2)}$ should be made in the coming years through improved lattice calculations. Progress on $\Omega_{\eta+\eta'}$ should also be possible in the near future. Moreover, as various estimates of $\hat B_K$, $B_8^{(3/2)}$ and $\Omega_{\eta+\eta'}$ by means of several non-perturbative methods are compatible with each other we do not expect big surprises here. Similar comments apply to $|V_{ub}|$, $|V_{cb}|$, $m_{\rm t}$ and $\Lambda_{\overline{\rm MS}}^{(4)}$. On the other hand, it appears that it will take longer to obtain acceptable values for $m_{\rm s}$ and $B_6^{(1/2)}$. In view of the bounds \cite{DERAF}-\cite{Dosch}, it is difficult to imagine that $m_{\rm s}(m_{\rm c})\le 105\, {\rm MeV}$. Consequently we expect that future improved estimates of $m_{\rm s}$ will most probably exclude the lowest values of $m_{\rm s}$ considered in this paper. This would simultaneously exclude the highest values for $\varepsilon'/\varepsilon$ obtained by us unless $B_6^{(1/2)}$ is found to be higher than used here. In this respect improved estimates of $B_6^{(1/2)}$, if found substantially higher than unity, could have considerable impact on our analysis. Finally, it should be stressed that future lattice calculations will give the full matrix elements without the necessity to use separately $(B_6^{(1/2)},B_8^{(3/2)})$ and $m_{\rm s}$. In any case $\varepsilon'/\varepsilon$ already played a decisive role in establishing direct CP violation in nature and its rather large value gives additional strong motivation for searching for this phenomenon in cleaner K decays like $K_L\to\pi^0\nu\bar\nu$ and $K_L\to\pi^0 e^+ e^-$, in B decays, in D decays and elsewhere. {\bf Acknowledgements}\\ We would like to thank S. Bertolini, G. Buchalla, M. Ciuchini, E. Franco, P. Gambino, T. Hambye, L. Littenberg, G. Martinelli, P. Soldan and J. Urban for discussions. This work has been supported by the German Bundesministerium f\"ur Bildung und Forschung under contract 06 TM 874 and DFG Project Li 519/2-2.
\section{\@startsection {section}{1}{\zeta@}{3.ex plus 1ex minus .2ex}{2.ex plus .2ex}{\large\bf}} \def\subsection{\@startsection{subsection}{2}{\zeta@}{2.75ex plus 1ex minus .2ex}{1.5ex plus .2ex}{\bf}} \def\subsect#1{\par\penalty1000{\noindent \bf #1}\par\penalty500} \def
\section{Introduction} The Skyrme model presents an opportunity to understand nuclear physics as a low energy limit of quantum chromodynamics (QCD). The model was initially proposed as a theory of strong interactions of hadrons \cite{Skyr1}, but recently, it was shown to be the low energy limit of QCD in the large $N_c$ limit \cite{Wit}. Since then further work has suggested that topologically nontrivial solutions of this model, known as skyrmions, can be identified with classical ground states of light nuclei. However, a thorough understanding of the structure and dynamics of multi-skyrmion configurations is required before a more qualitative assessment of the validity of this application of the model can be made. The $SU(N)$ Skyrme model involves fields which take values in $SU(N)$; {\it ie } are described by $SU(N)$ valued functions of $\vec{x}$ and $t$. Its static solutions correspond to field configurations describing multi-skyrmions. In this paper new solutions have been obtained for fields whose energy density is spherically symmetric. Multi-skyrmions are stationary points (maxima or saddle points) of the static energy functional, which is given in topological charge units by \begin{equation} E=\frac{1}{12\pi^2}\int_{R^3}\left\{-\frac{1}{2}\mbox{tr}\left(\partial_iU\, U^{-1}\right)^2-\frac{1}{16}\mbox{tr}\left[\partial_iU\, U^{-1},\;\partial_j U\, U^{-1}\right]^2\right\}d^3\vec{x}, \label{gene} \end{equation} where $U(\vec{x})\in SU(N)$. In this case multi-skyrmions are solutions of the equation \begin{equation} \partial_i\left(\partial_iU\, U^{-1}-\frac{1}{4}\,[\partial_jU\, U^{-1},[\partial_j U\, U^{-1}, \partial_iU\, U^{-1}]]\right)=0. \label{geq} \end{equation} We have, for simplicity, set the mass terms to zero. This has been done for convenience, since the conventional mass terms introduce only small changes and, as we will see later, affect only the profile functions. Therefore, all our discussion can be easily generalised to include such mass terms. Finiteness of the energy functional requires that $U(\vec{ x})$ approaches a constant matrix at spatial infinity, which can be chosen to be the identity matrix by a global $SU(N)$ transformation. So, without any loss of generality, we can impose the following boundary condition on $U$: $U \rightarrow {\it I}$ as $|\vec{ x}|\rightarrow \infty$. Since $U \rightarrow {\it I}$ as $|\vec{ x}|\rightarrow \infty$ is a mapping from $S^3\rightarrow SU(N)$, it can be classified by the third homotopy group $\pi_3(SU(N))\equiv Z$ or, equivalently, by the integer valued winding number \begin{equation} B=\frac{1}{24\pi^2}\int_{R^3} \varepsilon_{ijk}\,\mbox{tr}\left(\partial_i U\, U^{-1}\partial_j U\, U^{-1} \partial_k U\, U^{-1}\right)d^3\vec{x},\label{bar} \end{equation} which is a topological invariant. This winding number classifies the solitonic sectors in the model, and as Skyrme has argued \cite{Skyr1}, $B(U)$ may be identified with the baryon number of the field configuration. Up to now most of the studies involving the Skyrme model have concentrated on the $SU(2)$ version of the model and its embeddings into $SU(N)$. The simplest nontrivial classical {\it solution} involves a single skyrmion ($B=1$) and has already been discussed by Skyrme \cite{Skyr1}. The energy density of this solution is radially symmetric and, as a result, using the so-called hedgehog ansatz one can reduce (\ref{geq}) to an ordinary differential equation, which then has to be solved numerically. Many solutions with $B > 1$ of the $SU(2)$ model have also been computed numerically and, in all cases, the solutions are very symmetrical (cf. Battye et al. \cite{BS} and references therein). However, since the model is not integrable, with few exceptions, explicit solutions (even) for spherical symmetric $SU(N)$ skyrmions are not known. The first example of a {\it non-embedded} solution for a higher group was the $SO(3)$ soliton, corresponding to a bound system of two skyrmions, which was found by Balachandran et al. \cite{BBLRS}. Another solution, with a large $SU(3)$ strangeness content, was found by Kopeliovich et al. \cite{Kop}. However, all other known multi-skyrmion configurations seem to be the embeddings of the solutions of the $SU(2)$ model. Recently, we have showed in \cite{IPZ} how to construct low energy states of the $SU(N)$ model by using $CP^{N-1}$ harmonic maps. Our discussion involved only one projector. In this paper, we extend our method to more projectors. We show that, for the $SU(N)$ model, when we take $N-1$ projectors which lead to spherically symmetric energy densities, the full equations of the model separate and the problem of finding exact solutions is reduced to having to solve $N-1$ coupled nonlinear ODE's for $N-1$ profile functions. This way we obtain a whole family of new spherical symmetric multi-baryon solutions of the $SU(N)$ models. Our solutions include the $SU(3)$ dibaryon configuration of Balachandran et al. \cite{BBLRS} and the non-topological $SU(3)$ four baryon configuration of \cite{IPZ}. \section{Harmonic Maps} In \cite{IPZ} we generalised the $SU(2)$ ansatz of Houghton et al. \cite{HMS} to $SU(N)$. This generalisation involved re-writing the expression of Houghton et al. as a projector from $S\sp2$ to $CP\sp{N-1}$. It gave us a new way of interpreting old results and of deriving expressions for the low energy $SU(N)$ field configurations which are {\it not} simple embeddings of $SU(2)$ fields. In particular, the energy distributions exhibit very different symmetries from those of the embeddings. The method also gave us a new solution of the $SU(3)$ model, which lies in the topologically trivial sector of the model ({\it ie} it has zero baryon number) and so, obviously, is not stable. The method of \cite{IPZ} can be generalised further, to involve more projectors. In fact, we can exploit here some ideas taken from the theory of harmonic maps of $S\sp2\rightarrow CP\sp{N-1}$ \cite{Zak, DinZak}; since they play an important role in our construction. Recall (cf. \cite{DinZak}) that in $N$-dimensional space there is a ``natural" set of projectors: $S\sp2\rightarrow CP\sp{N-1}$ maps, which are constructed as follows: Write each projector $P$ as \begin{equation} P(V)=\frac{V \otimes V^\dagger}{|V|^2}, \label{for} \end{equation} where $V$ is a $N$-component complex vector of two variables $\xi$ and $\bar{\xi}$ which locally parametrise $S\sp2$. In terms of the more familiar $\theta$ and $\varphi$, they are given by $\xi=\tan(\theta/ 2)\,e\sp{i\varphi}$. The first projector is obtained by taking $V=f(\xi)$, {{\it ie }} an analytic vector of $\xi$; while the other projectors are obtained from the original $V$ by differentiation and Gramm-Schmidt orthogonalisation. If we define an operator $P_+$ by its action on any vector $v \in {\rm C}^N$ \cite{DinZak} as \begin{equation} P_+ v=\partial_\xi v- v\,\frac{v^\dagger \,\partial_\xi v}{|v|^2}, \end{equation} then the further vectors $P^k_+ v$ can be defined by induction: $P^k_+ v=P_{+}(P^{k-1}_+ v)$. Therefore, in general, we can consider projectors $P_k$ of the form (\ref{for}) corresponding to the family of vectors $V\equiv V_k=P\sp{k}_+f$ (for $f=f(\xi)$) as \begin{equation} P_k=P(P^k_+ f), \hspace{5mm} \hspace{5mm} k=0,\dots,N-1, \label{maps} \end{equation} where, due to the orthogonality of the projectors, we have $\sum_{k=0}^{N-1}P_k=1$. The orthogonality properties of our projectors follow from the following properties of vectors $P^k_+ f$ which hold when $f$ is holomorphic: \begin{eqnarray} \label{bbb} &&(P^k_+ f)^\dagger \,P^l_+ f=0, \hspace{5mm} \hspace{5mm} \hspace{5mm} k\neq l,\\[3mm] &&\partial_{\bar{\xi}}\left(P^k_+ f\right)=-P^{k-1}_+ f \frac{|P^k_+ f|^2}{|P^{k-1}_+ f|^2}, \hspace{5mm} \hspace{5mm} \partial_{\xi}\left(\frac{P^{k-1}_+ f}{|P^{k-1}_+ f|^2}\right)=\frac{P^k_+ f}{|P^{k-1}_+f|^2}. \label{aaa} \end{eqnarray} Note that, for $SU(N)$, the last projector $P_{N-1}$ in the sequence corresponds to an anti-analytic vector; ({{\it ie }} the components of $V_{N-1}=P^{N-1}_+f$, up to an irrelevant overall factor which cancels in the projector, are functions of only $\bar{\xi}$). Our new $SU(N)$ generalisation of \cite{IPZ} involves the introduction of $N-1$ projectors, {\it ie } \begin{eqnarray} U&=&\exp\{ig_0 (P_0-\frac{{\it I}}{N})+ig_1(P_1-\frac{{\it I}}{N})-\dots+ig_{N-2}(P_{N-2}-\frac{{\it I}}{N})\} \nonumber\\ &=&e^{-ig_0/N}({\it I}+A_0P_0)\,\,e^{-ig_1/N}({\it I}+A_1P_1)\,\dots\, \,e^{-ig_{N-2}/N}({\it I}+A_{N-1}P_{N-2}), \label{SUN} \end{eqnarray} where $g_k=g_k(r)$, for $k=0, \dots, N-2$, are the profile functions and $A_k=e^{ig_k}-1$. Note that the projector $P_{N-1}$ is not included in the above formula since it is the linear combination of the others. [Our previous ansatz given in \cite{IPZ} corresponds to putting all the profile functions, but the first one, equal to zero.] The spherically symmetric maps into $CP^{N-1}$ are given by \begin{eqnarray} f = (f_0, \,f_1, \dots, \, f_{N-1})^t, \hspace{5mm} \mbox{where} \hspace{5mm} f_k = \xi^k \sqrt{C_{k+1}^{N-1}}, \label{vec1} \end{eqnarray} where $C_{k+1}^{N-1}$ denote the binomial coefficients. Furthermore, as we prove in the appendix, the modulus of the corresponding vector $P_+^k f$ for $f$ of the above form is \begin{equation} \vert P\sp{k}_+f\vert\sp2=\alpha(1+\vert \xi\vert\sp2)\sp{N-2k-1}, \end{equation} where $\alpha$ depends on $N$ and $k$. \section{Constructing the Skyrmion Solutions} In this section we construct a family of exact spherical symmetric solutions of the $SU(N)$ Skyrme models. In fact, we show that for each $SU(N)$ model the Skyrme field involving $N-1$ projectors leads to an exact solution involving $N-1$ profile functions. \subsection{Skyrme Equations} The Skyrme equations (\ref{geq}), when re-written in spherical coordinates, take the form: \begin{eqnarray} \partial_r\!\left[r\sp2R_r+{1\over 4}\left(A_{\theta r \theta}+{1\over \sin\sp2 \theta}A_{\varphi r\varphi}\right)\right] \!\!+\!\!{1\over \sin\theta}\partial_{\theta}\!\left[\sin\theta\left\{R_\theta +{1\over 4}\left(A_{r\theta r}+{1\over r\sp2\sin\sp2\theta}A_{\varphi \theta\varphi}\right)\right\}\right]&&\nonumber \\ +{1\over \sin\sp2\theta} \partial_\varphi\!\left[R_\varphi+{1\over 4} \left(A_{r\varphi r}+{1\over r\sp2}A_{\theta\varphi\theta}\right)\right]=0,&& \label{eqqq} \end{eqnarray} where $R_i=U\sp{-1}U_i$ and $A_{\alpha\beta\gamma}\equiv [R_\alpha,\,[R_\beta,\,R_\gamma]].$ It is easy to see, using (\ref{SUN}), that \begin{equation} \label{rr} R_r=i\sum_{j=0}\sp{N-2}\dot g_j\left(P_j-{{\it I}\over N}\right), \end{equation} where $\dot g_j(r)$ denotes the derivative of $g_j(r)$ with respect to its argument; and that, in terms of the holomorphic variables $\xi$ and $\bar \xi$, \begin{eqnarray} R_\xi&=&e^{(-i\sum_{k=0}\sp{N-2}g_kP_k)}\, \partial_\xi\!\left[e^{(i\sum_{i=0}\sp{N-2}g_iP_i)}\right]\nonumber\\ &=&\left[1+\sum_{k=0}\sp{N-2}(e\sp{-ig_k}-1)P_k\right] \left[\sum_{l=0}\sp{N-2}(e\sp{ig_l}-1)P_{l\xi}\right]\nonumber\\ &=& \sum_{i=1}\sp{N-1}\left[e\sp{i(g_i-g_{i-1})}-1\right]{V_i\,V\sp{\dagger}_{i-1} \over \vert V_{i-1}\vert\sp2}, \label{rxi} \end{eqnarray} where the last line follows from the identity $e^{(-i\sum_{k=0}\sp{N-2}g_kP_k)}=1+\sum_{k=0}\sp{N-2}(e\sp{-ig_k}-1)P_k$. Here, $g_{N-1}=0$ and $R_{\bar\xi}=-(R_\xi)\sp{\dagger}.$ Next we note that \begin{equation} \partial_{\theta}=\frac{1+|\xi|^2}{2\sqrt{|\xi|^2}}\left(\xi \,\partial_\xi\,+\bar\xi\,\partial_{\bar\xi}\right),\hspace{5mm} \hspace{5mm} \partial_{\varphi}=i\left(\xi \,\partial_\xi\,-\bar\xi\,\partial_{\bar\xi}\right), \label{mig} \end{equation} and re-write all the terms in (\ref{eqqq}) in terms of $R_\xi$, $R_{\bar \xi}$ and $R_r$ and their commutators, {\it ie} \begin{eqnarray} &&\hspace{-13mm}\partial_{\theta} \left(\sin \theta\, R_{\theta}\right)+\frac{1}{\sin \theta}\partial_{\phi} R_{\phi}=(1+|\xi|^2)\sqrt{|\xi|^2}\left((R_\xi)_{\bar{\xi}} +(R_{\bar{\xi}})_{\xi}\right),\\ &&\label{En}\hspace{-13mm}A_{\theta r \theta}+{1\over \sin\sp2\theta}A_{\varphi r\varphi} =\frac{(1+|\xi|^2)^2}{2}\left\{[R_{\bar\xi},[R_r,R_\xi]] +[R_\xi,[R_r,R_{\bar\xi}]]\right\},\\[3mm] &&\hspace{-13mm}\sin \theta \,\partial_{\theta} \left(\sin \theta\,A_{r\theta r}\right)+\partial_{\phi}(A_{r\varphi r})=2|\xi|^2 \left([R_r,[R_{\xi},R_r]]_{\bar{\xi}} +[R_r,[R_{\bar{\xi}},R_r]]_{\xi}\right),\\ &&\hspace{-13mm}\partial_{\theta}\!\!\left(\!\frac{A_{\varphi \theta\varphi}}{\sin \theta}\!\right)\!\!+\!\!\frac{1}{\sin \theta}\partial_{\phi}\!\!\left(A_{\theta\varphi\theta}\right)\!=\! \frac{(1+|\xi|^2)\,\sqrt{|\xi|^2}}{2}\!\!\left[\partial_{\bar{\xi}}\left((1+|\xi|^2)^2 [R_\xi,[R_\xi,R_{\bar\xi}]]\right)\!\!-\!\!\partial_{\xi}\!\left((1+|\xi|^2)^2 [R_{\bar\xi},[R_{\xi},R_{\bar\xi}]]\right)\!\right].\nonumber\\ \end{eqnarray} Thus equation (\ref{eqqq}), when re-written in the holomorphic variables, becomes \begin{eqnarray} &&\partial_r\left[r\sp2R_r+ \frac{(1+|\xi|^2)^2}{8}\left([R_{\bar\xi},[R_r,R_\xi]] +[R_\xi,[R_r,R_{\bar\xi}]]\right)\right] +\frac{(1+|\xi|^2)^2}{2}\left((R_{\bar{\xi}})_\xi+ (R_{\xi})_{\bar{\xi}}\right)+\nonumber\\[1.5mm] &&\frac{(1+\vert\xi\vert\sp2)\sp3}{8r^2}\left(\xi\, [R_\xi,[R_\xi,R_{\bar\xi}]]-\bar\xi\, [R_{\bar\xi},[R_\xi,R_{\bar\xi}]]\right) +{(1+\vert \xi\vert\sp2)\sp4\over 16r\sp2}\left([R_\xi,[R_{\xi},R_{\bar\xi}]]_{\bar\xi}-[R_ {\bar\xi},[R_\xi,R_{\bar\xi}]]_{\xi}\right)\nonumber\\[1.5mm] &&+{(1+\vert\xi\vert\sp2)\sp2\over 8} \left([R_{r},[R_{\bar\xi},R_{r}]]_\xi +[R_r,[R_\xi,R_{r}]]_{\bar\xi}\right)=0. \label{Eqq} \end{eqnarray} Using (\ref{SUN}) we observe that \begin{eqnarray} &&[R_\xi,\,R_{\bar{\xi}}]=2\,P_0{\vert V_1\vert\sp2\over \vert V_0\vert\sp2}\left(1-\cos(g_1-g_0)\right)\,-2\,P_{N-1}{\vert V_{N-1} \vert\sp2\over \vert V_{N-2}\vert\sp2}\left(1-\cos(g_{N-2})\right)+\nonumber\\ &&\hspace{20mm}\,2\,\sum_{i=1}\sp{N-2}\,P_i\, \left[{\vert V_{i+1}\vert\sp2\over \vert V_i\vert\sp2}\left(1-\cos(g_{i+1}-g_i)\right) -{\vert V_i\vert\sp2\over \vert V_{i-1}\vert\sp2} \left(1-\cos(g_i-g_{i-1})\right)\right]\!\!,\label{ncomm}\hspace{5mm}\\[3mm] &&[R_{\xi},[R_\xi,\,R_{\bar{\xi}}]]= \sum_{i=1}\sp{N-1} \left(\mu_i{\vert V_{i+2}\vert\sp2\over \vert V_{i+1}\vert\sp2} +\nu_i{\vert V_{i+1}\vert\sp2\over \vert V_{i}\vert\sp2} +\rho_i{\vert V_{i}\vert\sp2\over \vert V_{i-1}\vert\sp2}\right) {V_iV\sp{\dagger}_{i-1} \over \vert V_{i-1}\vert\sp2},\\[3mm] &&[R_r,[R_{\bar\xi},R_r]]=\sum_{i=1} \sp{N-1}s_i\,{V_{i-1}V\sp{\dagger}_i\over\vert V_{i-1}\vert\sp2}, \end{eqnarray} where $\mu$, $\nu$ and $\rho$ are functions of $g_k(r)$, only; while $s_i$ are functions of $g_k(r)$ and their derivatives. Since $V_k=P_+^k f$, one can show that ${\vert V_{i}\vert\sp2\over \vert V_{i-1}\vert\sp2}\propto (1+\vert \xi\vert\sp2)^{-2}$; while $\partial_\xi\,(1+\vert \xi\vert\sp2)^{-2}=-2\bar\xi (1+\vert\xi\vert\sp2)^{-3}$ and thus, the derivative terms involving $[R_\xi,[R_\xi,R_{\bar\xi}]]$ in (\ref{Eqq}) cancel leaving us with derivatives of ${V_iV\sp{\dagger}_{i-1} \over \vert V_{i-1}\vert\sp2}$ -- which are proportional to $\sum_{i=1}\sp{N-1}(1+\vert\xi\vert\sp2)^{-2}(P_i-P_{i-1})\,h_i$, where $h_i$ involve functions of $g_k(r)$ (due to (\ref{aaa})), multiplied by terms of the form ${\vert V_{i}\vert\sp2\over\vert V_{i-1}\vert\sp2}$. So the factors $(1+\vert\xi\vert\sp2)^{-4}$ in ({\ref{Eqq}) cancel -- leaving us with a sum of differences of two successive projectors multiplied by functions dependent only on $r$. Following the above argument and using the properties of $R_r$, {\it etc } one can show that the terms $[R_r,[R_{\xi},R_r]]_{\bar\xi}$ in (\ref{Eqq}), are proportional to $\sum_{i=1}\sp{N-1}S_i\, (1+\vert\xi\vert\sp2)^{-2}(P_i-P_{i-1})$, where $S_i$ are functions of $g_k(r)$ and their derivatives -- leaving us, once again, with a sum of differences of two successive projectors multiplied by functions dependent only on $r$. Finally, the contribution of the terms $(1+|\xi|^2)\,R_{\xi\bar\xi}$ is given by $\sum_{i=1}\sp{N-1} (P_i-P_{i-1}) H_i,$ where $H_i$ are only functions of $g_k$; while the commutators in (\ref{En}) are equal to a sum of projectors mutlipied by $(1+\vert\xi\vert\sp2)^{-2}$, which cancel out in (\ref{Eqq}). In addition, $\partial_r( r\sp2 R_r)=i \sum_{i=0}\sp{N-2}\left(P_i-{{\it I}\over N}\right)(2r\dot g_i+r\sp2\ddot g_i).$ We note that, for our choice of the vectors $V_k$, all the dependence on $\xi$ and $\bar\xi$ in (\ref{Eqq}) resides only in the projectors (the rest of it cancels out). The terms involving $\partial_r(r\sp2 R_r)$ give us expressions involving ${1\over N}-P_i$ while all the other terms give us expressions involving $P_i-P_{i-1}$. Although $N$ projectors arise in (\ref{Eqq}), the projector $P_{N-1}$ can be re-expressed in terms of the previous ones -- giving $N-1$ factors involving the harmonic maps $P_i-{1\over N}$ (for $i=0,... N-2$). To satisfy (\ref{Eqq}) the coefficients of such factors have to vanish leaving us with $N-1$ equations for the $N-1$ profile functions $g_i$. Hence, if these equations have solutions then they correspond to exact solutions of the $SU(N)$ Skyrme models. Notice that (\ref{vec1}) implies that these solutions have a covariant axial symmetry, {\it ie} any rotation by an angle $\alpha$ around the $z$-axis is equivalent to the gauge transformation $U \rightarrow A^{\dagger}U A$ where $A = \mbox{diag}(1,e^{i\alpha}, e^{2 i \alpha}, .. e^{(N-1)i\alpha})$. On the other hand, as will be shown below, the energy density for these solution is radially symmetric. The $N-1$ equations for the profile functions can be obtained either from (\ref{Eqq}) -- which is a hard task; or from the variation of the energy (\ref{gene}) -- using (\ref{SUN}) and integrating out $\xi$ and $\bar\xi$ variables. Clearly, the two methods give the same equations. Let us stress that our procedure hinges on having $N-1$ profile functions and on the very special form of our vectors $V_k$. Had we taken other vectors $V_k$, we would have got some $\xi$ and $\bar\xi$ dependence outside the projectors; while had we taken less than $N-1$ profile functions and projectors we would have got too many equations for our functions. It is only in the case of $N-1$ projectors that we get the right number of equations. \subsection{Energy Dependence on Profile Functions} The energy (\ref{gene}), when written in the holomorphic variables, becomes \begin{equation} \label{energy} E\!=\!-\frac{i}{12\pi^2}\!\!\int\! r^2 dr\,d\xi d\bar{\xi} \,\mbox{tr}\left(\frac{1}{(1+\vert \xi\vert\sp2)\sp2}R\sp2_r+ \frac{1}{r\sp2} \vert R_\xi\vert\sp2 +\frac{1}{4r^2}[R_r,R_\xi][R_r, R_{\bar\xi}]-\frac{(1+\vert \xi\vert\sp2)\sp2}{16r\sp4}[R_{\bar\xi},R_\xi]\sp2\right). \label{pen} \end{equation} Using (\ref{rr}) and (\ref{rxi}) we find that \begin{eqnarray} &&\mbox{tr}\,R_r\sp2={1\over N}\left(\sum_{i=0}\sp{N-2}\dot g_i\right)\sp2\,-\,\sum_{i=0}\sp{N-2}\dot g_i\sp2,\\ &&\mbox{tr}\vert R_\xi\vert\sp2\,=\,-2\sum_{i=1}\sp{N-1} B_i,\\ &&\mbox{tr}[R_r,R_\xi][R_r,R_{\bar\xi}]=-2\sum_{k=1}\sp{N-1} B_k(\dot g_k-\dot g_{k-1})\sp2,\\ &&\mbox{tr}[R_{\bar\xi},R_\xi]\sp2=4\left(B_1\sp2 +\sum_{i=1}\sp{N-2} (B_i-B_{i+1})\sp2+B_{N-1}\sp2\right), \end{eqnarray} where $B_i={\vert V_i\vert\sp2\over \vert V_{i-1}\vert\sp2}(1-\cos(g_i-g_{i-1}))$ and $[R_{\bar\xi},R_\xi]=2\sum_{l=1}\sp{N-1}(P_{l-1}-P_l)B_l$. Since ${\vert V_k\vert\sp2\over \vert V_{k-1}\vert\sp2}=k(N-k)(1+\vert \xi\vert\sp2)^{-2}$ (see appendix) all terms in (\ref{pen}) have a factor $(1+\vert \xi\vert\sp2)^{-2}$ and the integration over $\xi$ and $\bar\xi$ is a topological constant, {\it ie} $i\int d\xi d\bar{\xi} (1+\vert \xi\vert\sp2)^{-2}\,=\,2\pi$. Thus we get \begin{eqnarray} E\!\!\!\!&=&\!\!\!\frac{1}{6\pi}\int\!\! r\sp2 dr\{-{1\over N}\left(\sum_{i=0}\sp{N-2} \dot g_i\right)\sp2+\sum_{i=0}\sp{N-2}\dot g_i\sp2 +{1\over 2r^2}\sum_{k=1} \sp{N-1}\left(\dot g_k-\dot g_{k-1}\right)\sp2D_k +{2\over r\sp2} \sum_{k=1}\sp{N-1}D_k\nonumber\\ &&\hspace{17mm}+{1\over 4r\sp4} \left(D_1\sp2+\sum_{k=1}\sp{N-2}(D_k-D_{k+1})\sp2+D_{N-1}\sp2\right)\}, \end{eqnarray} where $D_k=k(N-k)(1-\cos(g_k-g_{k-1}))$. Let us, for simplicity, take $F_k=g_k-g_{k+1}$, ($k=0,..N-2)$ with $F_{N-2}=g_{N-2}$. Then, the variation of the integrand of the energy $\tilde E$ with respect to the functions $\dot F_l$ (for $l=0,..N-2$) is \begin{equation} {\partial \tilde E\over \partial \dot F_l}=\left[-{2(l+1)\over N}\sum_{i= 0} \sp{N-2}(i+1)\dot F_i+2\sum_{k=0}\sp{l}\left(\sum_{i=k}\sp{N-2}\dot F_i\right) +\frac{1}{r^2}\dot F_l D_{l+1}\right]r\sp2, \end{equation} where $D_k=k(N-k)\left(1-\cos F_{k-1}\right)$. Therefore, the equations of motion for the functions $F_i$, and thus for the profile functions, are \begin{eqnarray} &&-{2(l+1)\over N}\sum_{i=0}\sp{N-2}(i+1)\ddot F_i+2\sum_{k=0}\sp{l} \sum_{i=k}\sp{N-2}\ddot F_i+\frac{1}{r^2}\ddot F_l(l+1)(N\!-\!l\!-\!1)(1-\cos F_l)+\nonumber\\ \label{profs} &&{1\over 2r\sp2}\dot F_l\sp2(l\!+\!1)(N\!-\!l\!-\!1)\sin F_l\!+\!{2\over r}\left( \!-{2(l\!+\!1)\over N}\sum_{i=0} \sp{N-2}(i+1)\dot F_i\!+\!2\sum_{k=0}\sp{l}\left(\sum_{i=k}\sp{N-2}\dot F_i\right)\right)\nonumber\\ &&-{2\over r\sp2}\,(l+1)(N-l-1)\,\sin F_l- {1\over r\sp4}\,(l+1)\sp2(N-l-1)\sp2\left(1-\cos F_l\right)\sin F_l\nonumber\\ &&+{1\over 2r\sp4}(l\!+\!1)(N\!-\!l\!-\!1)\sin F_l\left[l(N\!-\!l)(1-\cos F_{l-1})+(l\!+\!2)(N\!-\!l\!-\!2)(1-\cos F_{l+1})\right]=0.\nonumber\\ \end{eqnarray} These equations can be solved numerically by imposing the appropriate boundary conditions on the profile functions. To do this we have to specialise to a particular model, {\it ie } for specific $N$ and diagonalise the terms involving the second derivatives. The simplest cases: $N=2$, $N=3$ and $N=4$ involve $1$, $2$ or $3$ functions and will be discussed in the next sections. \section{Topological Properties and Symmetries} Before we discuss special cases, let us first investigate the general topological properties of our fields. The topological charge (\ref{bar}), which in many applications of the Skyrme model is identified with the baryon number, is given by \begin{equation} B=\frac{1}{8\pi^2}\int dr\, d\xi d\bar{\xi} \,\mbox{tr}\, \left(R_r\,[R_{\bar\xi},R_\xi]\right), \end{equation} when written in the complex coordinates. Due to (\ref{ncomm}) and (\ref{rr}) the terms involving $\dot g_i/N$ in $R_r$ after taking the trace cancel and the expression for the baryon number simplifies to \begin{eqnarray} B&=&-\frac{i}{4\pi^2}\int dr\, d\xi d\bar{\xi} \,\sum_{i=0}\sp{N-2}\, \dot F_i\,(1-\cos F_i){\vert V_{i+1}\vert\sp2\over \vert V_i\vert\sp2}\nonumber\\ &=&\frac{1}{2\pi}\int dr \sum_{i=0}\sp{N-2}\dot F_i\,(1-\cos F_i)(i+1)(N-i-1)\nonumber \\ &=& \frac{1}{2\pi}\sum_{i=0}\sp{N-2}(i+1)(N-i-1)\,\left(F_i-\sin F_i\right)_{r=0}\sp{r=\infty}. \label{BofF} \end{eqnarray} As $g_i(\infty)=0$ (required for the finiteness of the energy) the only contributions to the topological charge come from $F_i(0)$. The $N-1$ equations for the profile functions and their differences given in (\ref{profs}) have many symmetries. These symmetries can be used to derive special skyrmion solutions which involve a smaller number of profile functions and projectors. The main symmetry of our expressions, are the independent interchanges \begin{equation} F_k\,\leftrightarrow\,F_{N-k-2},\qquad \mbox{for}\hspace{5mm} k=0,\cdots, N-2. \end{equation} This symmetry follows from the fact the terms $D_k=k(N-k)(1-\cos F_{k-1})$ which appear in the energy are symmetric under the interchange: $D_k\leftrightarrow D_{N-k}$ when $F_{k-1}\leftrightarrow F_{N-k-1}$. In addition, all the other terms in the energy also exhibit this symmetry since they are combinations of $F_i$ and their derivatives. \section{Spherical Skyrmions} The simplest case corresponds to the $SU(2)$ spherically symmetric skyrmion. This is the solution which was found thirty years ago by Skyrme and is usually presented in terms of the well-known hedgehog ansatz. \subsection{$SU(3)$ Skyrme Model} In this case $N=3$ and we have two functions: $F_0$ and $F_1$. The energy density ${\cal E}$, such that $E = (6\pi)^{-1}\int {\cal E} r^2 dr$, is given by \begin{eqnarray} {\cal E}\!\!\!\!&=&\!\!\!\! {2\over 3}(\dot F_0\sp2+\dot F_1\sp2+\dot F_0\dot F_1) + \frac{1}{r^2}\left[(\dot F_0\sp2+4)(1-\cos F_0)+ (\dot F_1\sp2+4)(1-\cos F_1)\right]+\nonumber\\ \!\!\!\!&& \qquad\frac{2}{r^4}\left[(1-\cos F_0)^2-(1-\cos F_0)\,(1-\cos F_1)+ (1-\cos F_1)^2\right],\nonumber\\ \end{eqnarray} and the equations for the profile functions are \begin{eqnarray} &&\hspace{-10mm} \ddot F_0 \left(\!1\!+\!\frac{3}{2r^2}(1\!-\!\cos F_0)\!\right)+ \frac{\ddot F_1}{2}\!+\!\frac{2\dot F_0\!+\!\dot F_1}{r}+ \frac{3\sin F_0}{4r^2}\left[\dot F_0^2\!-4\!-\!\frac{4(1\!-\!\cos F_0)}{r^2}\!+\! \frac{2(1\!-\!\cos F_1)}{r^2}\right]\!\!=\!0,\nonumber\\ &&\hspace{-10mm} \ddot F_1 \left(\!1\!+\!\frac{3}{2r^2}(1\!-\!\cos F_1)\!\right)+ \frac{\ddot F_0}{2}+ \frac{2\dot F_1\!+\!\dot F_0}{r}\!+\! \frac{3\sin F_1}{4r^2}\left[\dot F_1^2\!-4\!-\!\frac{4(1\!-\!\cos F_1)}{r^2}\!+\! \frac{2(1\!-\!\cos F_0)}{r^2}\right]\!\!=\!0.\nonumber\\ \label{SU3} \end{eqnarray} These equations can be solved numerically when the right boundary conditions have been imposed. However, by letting $F_0=F_1=F$ ({\it ie} using the symmetry) they reduce to \begin{equation} \ddot F \left(1+\frac{1-\cos F}{r^2}\right)+\frac{2\dot F}{r}+ \frac{\sin F}{2r^2}\left[\dot F^2-4-\frac{2(1-\cos F)}{r^2}\right]=0. \label{max} \end{equation} This equation coincides with the equation for the profile function of a single $SU(2)$ skyrmion. Here we note that as $F_0(0)=F_1(0)=2\pi$ the topological charge of our solution is four. Thus the energy of this configuration, which corresponds to four skyrmions is $E_{B=4}=4\times\,1.232$, {\it ie} is exactly four times the energy of one skyrmion. We see that we have a static solution corresponding to four non-interacting skyrmions, placed on top of each other in such a way that their energy density is spherically symmetric. In addition, there is a further symmetry which allows us to set $F_0=-F_1=G$. In this case the equations reduce to \begin{equation} \ddot G \left(\frac{1}{2}+\frac{3}{2r^2}(1-\cos G)\right)+\frac{\dot G}{r}+ \frac{3\sin G}{4r^2}\left[\dot G^2-4-\frac{2(1-\cos G)}{r^2}\right]=0. \end{equation} Let us note that, since $F_0=g_0-g_1$ and $F_1=g_1$, this case corresponds to $g_0=0$ and thus, the field (\ref{SUN}) involves only one projector, namely $P_1$. This solution is the topologically trivial solution discussed in \cite{IPZ} and its energy is $3.861$. Finally, Balachandran et al. skyrmion solution can be obtained from (\ref{SU3}) by imposing the following boundary conditions: $g_0(0)=2\pi$, $g_1(0)=0$ and $g_0(\infty)=0$, $g_1(\infty)=0$; its energy is $E_{B=2}=2.3764$. \subsection{$SU(4)$ Skyrme Model} In this case the energy density becomes \begin{eqnarray} &&{\cal E}= {1\over 4}\left(3\dot F_0\sp2+4\dot F_1\sp2+3\dot F_2^2+4\dot F_0\dot F_1+ 4\dot F_1\dot F_2+2\dot F_0\dot F_2\right)+\nonumber\\ &&\hspace{8mm}{1\over 2 r^2}\left[3(\dot F_0\sp2+4)(1\!-\!\cos F_0)\!+ \!4(\dot F_1\sp2+4)(1\!-\!\cos F_1)\!+ \!3(\dot F_2^2+4)(1\!-\!\cos F_2)\right]\nonumber\\ && \hspace{8mm}+\frac{1}{2r^4}\,\{9\,(1-\cos F_0)^2+16\,(1-\cos F_1)^2+ 9\,(1-\cos F_2)^2\nonumber\\ &&\hspace{8mm} - 12 (1-\cos F_0) (1-\cos F_1)-12(1-\cos F_1)(1-\cos F_2)\}, \end{eqnarray} while the equations for $F_0$, $F_1$ and $F_2$ are more complicated: \begin{eqnarray} &&\hspace{-20mm} \ddot F_0\left(\!1\!\!+\!\!\frac{2(1-\cos F_0)}{r^2}\!\right)\!+ \!\frac{2 \ddot F_1\!+\!\ddot F_2}{3}\!+\!\frac{3\dot F_0\!+ \!4\dot F_1\!+\!2\dot F_2}{3r}\!+ \!\frac{\sin F_0}{r^2} \left[\dot F_0^2\!-\!4\!-\!\frac{6(1\!\!-\!\!\cos F_0)}{r^2}\!+ \!\frac{4(1\!\!-\!\!\cos F_1)}{r^2}\right]\!\!=\!0,\nonumber\\ &&\hspace{-20mm} \ddot F_1\!\left(\!1\!\!+\!\!\frac{2(1\!\!-\!\!\cos F_1)}{r^2}\!\right)\!\!+ \!\frac{\ddot F_0\!+\!\ddot F_2}{2}\!+ \!\frac{2\dot F_1\!+\!\dot F_0\!+\!\dot F_2}{r}\!+ \!\frac{\sin F_1}{r^2}\! \left[\!\dot F_1^2\!-\!4\!-\!\frac{8(1\!\!-\!\!\cos F_1)}{r^2}\!+ \!\frac{3(1\!\!-\!\!\cos F_0)}{r^2}\!+ \!\frac{3(1\!\!-\!\!\cos F_2)}{r^2}\!\right]\!\!=\!0,\nonumber\\ &&\hspace{-20mm} \ddot F_2\left(\!1\!\!+\!\!\frac{2(1\!\!-\!\!\cos F_2)}{r^2}\!\right)\!+ \!\frac{\ddot F_0\!+\!2\ddot F_1}{3}\!+ \!\frac{6\dot F_2\!+\!2\dot F_0\!+\!4\dot F_1}{3r}\!+ \!\frac{\sin F_2}{r^2} \left[\dot F_2^2\!-\!4\!-\!\frac{6(1\!\!-\!\!\cos F_2)}{r^2}\!+ \!\frac{4(1\!\!-\!\!\cos F_1)}{r^2}\right]\!\!=\!0.\nonumber\\ \label{SU4eq} \end{eqnarray} These equations have the previously mentioned symmetry $F_k \leftrightarrow F_{N-k-2}$ which allows us to set $F_0=F_2$ by keeping $F_1$ arbitrary. In addition, letting $F_0=F_1=F_2=F$ the above system reduces to equation (\ref{max}) and therefore, the configuration which consists of ten skyrmions (as $B={3F_0(0)+4F_1(0)+3F_2(0)\over 2\pi}=10$) is $E_{B=10}=10\times\,1.232$, {\it ie } is exactly ten times the energy of one skyrmion. Once again we have a solution describing many skyrmions, which are non-interacting and whose energy density is spherically symmetric. In addition, letting $F_0=-F_2=G$ we spot that when $F_1=0$, we have a solution of the form \begin{equation} \ddot G\left(1+\frac{3(1-\cos G)}{r^2}\right)+\frac{2\dot G}{r} +\frac{3\sin G}{2r^2}\left[\dot G^2-4-\frac{6(1-\cos G)}{r^2}\right]=0, \end{equation} which corresponds to a non-topological solution, {\it ie } its baryon number is zero; however the corresponding configuration consists of three skyrmions and three anti-skyrmions. [Recall, that the profile functions are $g_0=0$ and $g_1=g_2$, {\it ie } the field (\ref{SUN}) involves only one projector of rank two -- namely $P_1+P_2$.] In general, however, the solutions depend on more functions. We can always assume that the functions $F_i$ go to zero at infinity, so the topological charge of a solution is determined, using (\ref{BofF}), by the boundary value of each $F_i$ at the origin. When each of these values is positive the solution is a mixture of skyrmions. When some $F_i$'s take positive and some $F_i$'s take negative values at the origin the solution corresponds to a mixture of skyrmions and anti-skyrmions. This is very similar to what happens in the two-dimensional Grasmannian sigma model \cite{Zak}. We have solved numerically equations (\ref{SU4eq}) taking all combinations, modulo the exchange of $F_0$ and $F_2$, of $0$, $2\pi$ and $-2\pi$ for the value of $F_i$ at the origin. The results are summarised in the Table below. \begin{center} \begin{tabular}{r|r|r|r|l|l|l} \vbox to 5mm {}$F_0(0)$ & $F_1(0)$ & $F_2(0)$ & B & Energy & E/baryon & SU(2) En \\ &&&&&&\\ \hline \hline \vbox to 4mm{}$2\pi$ & 0 & 0 & 3 & 3.518 & 1.173 & 3.429 \\ \vbox to 4mm{}0 & $2\pi$ & 0 & 4 & 4.788 & 1.197 & 4.464 \\ \vbox to 4mm{}$2\pi$ & 0 & $2\pi$ & 6 & 7.22553 & 1.204 & 6.654 \\ \vbox to 4mm{}$2\pi$ & 2$\pi$ & 0 & 7 & 8.45219 & 1.207 & 7.7693 \\ \vbox to 4mm{}$2\pi$ & $2\pi$ & $2\pi$ & 10 & 12.32 & 1.232 & - \\ &&&&&&\\ \hline \vbox to 4mm{}$2\pi$ & $-2\pi$ & $2\pi$ & 6-4 & 8.852 & 0.8852 & - \\ \vbox to 4mm{}$2\pi$ & $2\pi$ & $-2\pi$ & 7-3 & 9.896 & 0.9896 & - \\ \vbox to 4mm{}$2\pi$ & 0 & $-2\pi$ & 3-3 & 6.63422 & 1.106 & - \\ \vbox to 4mm{}$-2\pi$ & $2\pi$ & 0 & 4-3 & 6.61478 & 0.945 & - \\ \end{tabular} \end{center} The first five solutions are bound states of skyrmions with energies larger than the energies of the corresponding $SU(2)$ solutions \cite{BS}. Moreover, the excitation energy of the first two solutions is very small. As mentioned above, the energy of the $B=10$ solution is exactly ten times the energy of a single skyrmion solution. These solutions are all axially symmetric (but their energy densities are radially symmetric) and thus they are all more symmetrical than the corresponding $SU(2)$ solutions. The last four solutions are bound states of skyrmions and anti-skyrmions. Although their energies are very small, we know that these solutions must be unstable. \section{Conclusions} In this paper we have shown how to construct radially symmetric solutions of the $SU(N)$ Skyrme models. In the general case these solutions depend on $N-1$ profile functions which have to be determined numerically. In some cases we can exploit symmetries of our expressions and reduce the number of functions. Thus in the case of $SU(3)$ we can recover the topologically trivial solution discussed in \cite{IPZ}. We have not discussed the derived solutions in much detail. Their properties and their relation to physics deserve further study and these topics are currently under investigation. It is worth pointing out that there is a rather close connection between $SU(N)$ BPS monopoles and skyrmions. Both monopole and skyrmion fields can be constructed in terms of harmonic maps between Riemann spheres. Thus, recently, it has been shown \cite{Sut} that the monopoles fields can also be constructed using the projector ansatz. Its generalisation to multi-projectors and the construction in \cite{Sut} provide explicit examples of spherically symmetric $SU(N)$ monopoles with various symmetry breaking patterns. In the monopole case, the Bogomolny equation is the analogue of our Skyrme equation and the monopole number corresponds to our baryon number. \section{ Acknowledgments} We thank C. Houghton, V. Kopeliovich and P. Sutcliffe for their interest.
\section{Introduction} \label{introduction} \PARstart{B}{ased} on vast experience watching the fruit of my hard work fall apart time and again, I feel highly qualified to discuss the manner in which systems break. In particular, the goal of this paper is to define and analyze systems which exhibit brittle behavior. This behavior is characterized by a sudden and steep decline in performance as system state changes as shown by point D along curve $P_h$ in Figure \ref{bintro}. $P_h$ is the performance curve for a high performance system with brittle characteristics, $P_l$ is a lower performance system with less brittle characteristics. Clearly the slope from point $D$ along curve $P_h$ is much steeper than that of point $E$ along curve $P_l$. The steep decline of performance along $P_h$ can be due to input parameters which exceed a specified tolerance, or environmental conditions which exceed specified operating boundaries. This is equivalent to material fracture. Materials science provides a terminology which is apropos and flexible enough to describe the characteristics of this work. A table of materials science terms and their corresponding brittle system definitions is shown in Table \ref{terms}. Toughness \cite{Vlack} is the amount of energy absorbed by a material prior to failure. A brittle fracture occurs with very little energy absorption while a ductile fracture is accompanied by much energy absorption. Clearly toughness is the analog of the robustness of a system. To carry the analogy further, ductility is quantified as the amount of permanent strain prior to fracture. A system which does not exhibit brittle behavior will be called ductile \cite{Vlack}. Strain is unit-less and refers to the amount of deformation per unit length of a material and is caused by stress which is the force per unit area. Material deformation is analogous to degradation in a brittle system. In our work, stress is the distance by which a parameter exceeds its specified operating tolerance. There are two forms of strain, reversible and permanent. Reversible strain is called elastic strain and is characterized by Young's modulus: the ratio of the stress over the strain. Permanent strain leaves the shape of a material permanently changed and is known as plastic strain. In a brittle system, plastic strain will be degradation from which the system cannot recover, while a brittle system can recover from reversible strain. \begin{figure*}[htbp] \centerline{\psfig{file=figures/intro.eps,width=6.0in}} \caption{A Brittle versus Ductile System.} \label{bintro} \end{figure*} Increasing both hardness and ductility increases the toughness of a material. Hardness is increased by deforming the crystal structure, either by adding impurities to a homogeneous material or by rapid cooling of the material after processing. In this work, increasing hardness of a material is analogous to increasing the gain of the sub-components of a system. Previous work has focused on the hardness of a system, but relatively little on the ductility. For example, in choosing design parameters for a system \cite{Montgomery}, one examines the effect of high and low parameter values within the utility of normal operation of system performance ($U(normalOperation)$) and chooses those values which result in the best performance ($P_h$). However, the behavior and utility of the system when tolerance is exceeded ($U(robust)$) have rarely been examined. Certainly if time is considered, then based on simple reliability theory the utility is shown in Equation \ref{util} where $H$ and $D$ are shown in Figure \ref{bdef}, $x$ is a design parameter, $P[]$ is the probability of the event in the brackets, and $U(normalOperation)$ is the utility to the user of the system in normal operation, and $U(robust)$ is the utility to the user of the system outside normal operation. As graceful degradation becomes a more desirable feature, the utility of area $D$ increases. Let us define brittleness as the ratio of the hardness over the ductility which is the area H over D in Figure \ref{bdef}. \begin{figure*}[htbp] \centerline{\psfig{file=figures/bdef.eps,width=6.0in}} \caption{A Definition of System Brittleness.} \label{bdef} \end{figure*} \begin{eqnarray} Utility = P[x \in T] H U(normalOperation) + P[x \notin T] D U(robust) \label{util} \end{eqnarray} \begin{table*}[htbp] \centering \begin{tabular}{||l|l||} \hline \textbf{Materials Science} & \textbf{Brittle Systems} \\ \hline \hline Stress & amount parameter exceeds its tolerance \\ \hline Toughness & system robustness \\ \hline Hardness & level of performance within tolerance \\ \hline Ductility & level of performance outside tolerance \\ \hline Plastic Strain & system cannot recover from degradation \\ \hline Reversible Strain & system can recover from degradation \\ \hline Brittle Fracture & sudden steep decline in performance \\ \hline Ductile Fracture & graceful degradation in performance \\ \hline Brittleness & ratio of hardness over ductility \\ \hline Deformation & degradation in performance \\ \hline Young's Modulus & amount tolerance exceeded over degradation \\ \hline \end{tabular} \caption{\label{terms} Terminology.} \end{table*} At this point it must be mentioned that some causes of brittle fracture may be more difficult to deal with than others. For example the sudden loss of performance can be due to a catastrophe \cite{Saunders,IB-D84160}. Catastrophe Theory is essentially the study of singularities; in this work it would be one of many causes for brittle behavior. The connection between Catastrophe Theory and Brittle Systems is only one of the many areas that need to be explored in this new research area. \section{An Example of Ductility in a Communications Network} \label{example} The following applications which exhibit brittle behavior have been chosen as simple examples so that the ideas presented in this work, rather than the details of the applications, can be investigated. These examples will be examined in more detail as this work progresses. \subsection{Adaptive Multimedia} Current network applications, especially multimedia applications, have performance which degrades rapidly after bandwidth is reduced beyond a certain point. In \cite{Lee} it is suggested that if applications can be developed which degrade gracefully with respect to loss in bandwidth as shown in Figure \ref{qbplot}, then the network can be designed to maintain bandwidth within the required bounds on a best effort basis. A solution recommended in \cite{Lee} is for the network to keep a certain amount of bandwidth in reserve. However, the more bandwidth kept in reserve, the less that remains to support the network as a whole. Thus the amount of reserve bandwidth is the greatest factor affecting ductility in this example. As the value of reserve bandwidth increases, the number of users which can be supported is reduced, but fewer calls in progress are disconnected. \begin{figure*}[htbp] \centerline{\psfig{file=figures/qbplot.eps,width=2.5in}} \caption{Ductile Network Applications (Top Graph).} \label{qbplot} \end{figure*} \subsection{Packet Recovery: Stop And Wait System} The second example is recovery from packet loss in an Automatic Repeat Request (ARQ) link shown in Figure \ref{saw}. We consider two types of packets: packets with a large delay and packets which are lost. Setting a high time-out value results in better performance for packets which have a high ratio of delay to loss, but degrades rapidly as the ratio approaches zero. \begin{figure*}[htbp] \centerline{\psfig{file=graphs/StopAndWaitDelay.ps,width=4.0in}} \caption{Delay, Probability of Error, and Timeout in a Stop And Wait System.} \label{saw} \end{figure*} \subsection{TDMA Reservation System} Figure \ref{tdma} shows transmission rate versus probability of transmission for two values of retransmission. The lower valued setting for retransmission has higher performance, however, the higher valued retransmission setting is slightly more robust around a probability of 0.013. \begin{figure*}[htbp] \centerline{\psfig{file=graphs/TDMA.ps,width=4.0in}} \caption{TDMA Probability of Transmission and Transmission Rate.} \label{tdma} \end{figure*} \subsection{Mobile Cellular Telephone System} Figure \ref{mobile} shows grade of service versus channels per base station. \begin{figure*}[htbp] \centerline{\psfig{file=graphs/Mobile.ps,width=4.0in}} \caption{Grade of Service versus number of Channels per Base Station.} \label{mobile} \end{figure*} \subsection{Buffer Capacity} Another example of a brittle system involves choosing buffer capacity in a data communications system. \subsection{Backlogged Packets in Slotted ALOHA} In Slotted ALOHA, data packet transmission occurs using equal sized packets within equally divided time slots. If two or more users transmit within a given time slot, a collision occurs; the packet will be retransmitted in a following time slot with a given probability. This example of a brittle system exhibits catastrophic behavior \cite{Nelson:1987:SCT}. Let $p_0$ be the probability that a packet to be transmitted finds an empty cell, and $p_1$ be the probability that after a collision, the cell attempts retransmission. The design parameters are $p_0$ and $p_1$ and the number of packets waiting for retransmission is the state. A graph of $p_0$ and $p_1$ forms a cusp and all the classic symptoms of catastrophe are present, namely, bifurcation, sudden jumps, hysteresis, inaccessibility, and divergence. \subsection{Variable Window Flow Control} \begin{figure*}[htbp] \centerline{\psfig{file=graphs/VariableWindow.ps,width=4.0in}} \caption{Variable Window Throughput Comparison.} \label{varwin} \end{figure*} \subsection{Flow Control} Also in \cite{Nelson:1987:SCT}, it is suggested that flow control, shown in Figure \ref{varwin}, in a communications network exhibits not only brittle behavior, but catastrophic behavior. The specific model of flow control considered in \cite{Nelson:1987:SCT} is to divide available buffer space into classes and allow packets which have passed through $i$ hops to occupy buffers assigned to class $i$. \section{Techniques for Handling Brittle Systems} There are a variety of techniques for controlling and enhancing the ductility of a system. The first is to assign values to design parameters which influence ductility in a static manner, that is before the system becomes operational. The next involves dynamically changing the ductility as the system operates. This would be analogous to a material which could automatically trade-off hardness for ductility whenever necessary in order to maximize its performance. The remaining techniques involve methods of attempting to avoid brittle fracture, by design or by rolling back from a fracture. \subsection{Ductility Setting of System Sub-Components} Now that ductility has been defined and the design parameters controlling ductility identified, a natural question to ask is how should the sub-component parameters be set. Within normal operation, the performance requirements must be met, and in addition we would like the system to be tough (robust) outside the normal operating range as well. Is there a benefit to how ductility is distributed among subsystem components? As an example, in network and transport level data communications systems, if the system is going to fail, it is beneficial for low level system components to fail early in the transmission process rather than transporting a packet close to its destination and finding that the entire packet/frame has to be retransmitted later. Thus, it would be better to set $X_1$, in Figure \ref{bnet}, so that sub-component $S1$, which performs its processing early, has a lower ductility than components later in the process. A highly brittle component, as illustrated in Figure \ref{onoff}, would appear to have the characteristics of an on-off constant bit rate (on-off CBR) source. These types of sources have been used to model ATM \cite{Prycker} traffic sources. Queue fill distribution has been analyzed in \cite{Anick1982} for on-off CBR models. These results could be used in a buffer solution for such highly brittle components. \begin{figure*}[htbp] \centerline{\psfig{file=figures/onoff.eps,width=2.5in}} \caption{An On-Off Brittle Sub-Component.} \label{onoff} \end{figure*} \subsection{Adaptation} As mentioned previously, there are two forms of strain, reversible and permanent. Reversible strain is called elastic strain and is characterized by Young's modulus: the ratio of the stress over the strain. Permanent strain leaves the shape of a material permanently changed and is known as plastic strain. In brittle systems, an analog to plastic strain is adaptation. Once we know the parameters which affect ductility, that is, having determined $\psi$, the values of the parameters can be changed dynamically as stress causes the system to approach a brittle fracture. \subsection{Rollback} Another possibility is that once the system approaches a brittle fracture, the system has the capability to rollback to a safe state and choose another gradient which attempts to remain in a safe state of operation. Rollback techniques within a communications network environment have been described in \cite{BushThesis}. \section{Brittle Sub-Components} Consider a system whose sub-components exhibit various degrees of ductility as defined above. Just as adding impurities to a pure metal causes it to become stronger but more brittle, the addition of more efficient but also more sensitive components to a system causes the system to increase performance within its operating range, but become less ductile. How do the effects of ductility propagate among the sub-components to influence the ductility of the entire system? Assume the performance response curve is known for each sub-component and that the output from one component feeds into the input of the next component as shown in Figure \ref{bnet}. Assume that the sub-component output performance cannot be better than any of its inputs. Then the performance curve for the output of each sub-component is the minimum of the input sub-component performance curve and the current component performance curve. \begin{figure*}[htbp] \centerline{\psfig{file=figures/bnet.eps,width=4.0in}} \caption{Brittle Subsystem Components.} \label{bnet} \end{figure*} The hardness component of the brittleness enhances the performance when values are within tolerance and low ductility degrades the performance when values are out of tolerance. The amount of degradation depends on the amount by which the tolerance was exceeded. This is illustrated in Figure \ref{poutdiagram} and is stated in Equation \ref{pout}, where $b$ is the brittleness, $P_{in}$ is the input performance, $T$ is the set of in-tolerance values, $x$ is a state parameter, $E[]$ is the expected value, and $P_{out}$ is the output performance. The result of Equation \ref{pout} is plotted in Figure \ref{analpcurve}. \begin{figure*}[htbp] \centerline{\psfig{file=figures/pout.eps,width=6.0in}} \caption{Distribution and Performance.} \label{poutdiagram} \end{figure*} \begin{equation} \label{pout} P_{out} = (P_{in}+P_{in} b)Prob[x \in T] + (P_{in} - P_{in} b E[x-max[T]])Prob[x \notin T] \end{equation} \begin{figure*}[htbp] \centerline{\psfig{file=graphs/analpcurve.ps,width=6.0in}} \caption{Analytical Brittle Subcomponent Result.} \label{analpcurve} \end{figure*} As $b$ decreases in a non-brittle system, $T$ increases. It is this relationship between $b$ and $T$ which is the principal focus of brittle systems analysis. Assume the simple case of a normally distributed performance distribution, then Equation \ref{normpout} shows how Equation \ref{pout} can be refined. $N(\eta,\sigma)$ is a normal distribution with an average of $\eta$ and variance of $\sigma$ and $R.V._{N(\eta,\sigma)}$ is a random variable with distribution $N(\eta,\sigma)$. \begin{equation} \label{normpout} P_{out} = (P_in +P_{in} b) Prob[R.V._{N(\eta,\sigma)} \le T] + (P_in - b \eta) 1.0 - Prob[R.V._{N(\eta,\sigma)} \le T] \end{equation} As $b$ decreases we assume that the system is non-brittle so that $T$ increases. Assume that $b$ is linear, then $T$ increases as shown in Figure \ref{linbfig} and Equation \ref{linbeq}. \begin{figure*}[htbp] \centerline{\psfig{file=figures/linbfig.ps,width=6.0in}} \caption{Relationship Between $b$ and $T$.} \label{linbfig} \end{figure*} \begin{equation} \label{linbeq} T = p_x + {1 \over b} p_y \end{equation} A BONeS model has been developed to examine brittle sub-components as shown in Figure \ref{bmodel} which models Figure \ref{bnet}. A BONeS data structure contains the performance or quality of the input to a component. A normal random number generator produces a value with a specified mean and variance, in this case 10.0 and 3.0 respectively. The difference between the random number and the upper limit (11.0) is computed. If the normal random number is greater than the upper limit, then the performance value of the input data structure is reduced by the brittleness multiplied by the amount by which the tolerance was exceeded. If the normal random number is within tolerance then the input data structure is increased by an amount proportional to the brittleness. \begin{figure*}[htbp] \centerline{\psfig{file=figures/Bsys2.ps,width=6.0in}} \caption{BONeS Model.} \label{bmodel} \end{figure*} The results are shown in Figure \ref{sysperf} for the normal random number values, the upper limit, and the system performance as a function of the consecutive order in which each of the values were sampled. The brittleness is varied from zero to 0.8 and the results are averaged. Clearly the performance degrades when the normal values exceed the upper limit. Figure \ref{pcurve} shows the performance results for the sub-components and the entire system from Figure \ref{bmodel}. Components 1 and 3 generate data structures with a performance value of one. An intermediate component, Component 2, has a brittleness which varies from zero to 0.8. The final output component, Component 4, has a brittleness of 0.3. The analytical results from Figure \ref{analpcurve} and the simulated system performance curve from Figure \ref{pcurve} are in close agreement. Although Component 2 performance improves when the brittleness is between 0.2 and 0.5, Component 4, which is the system performance, declines. This is because Component 4 performance depends on the minimum performance input which comes from Component 3, an initial input component that always generates a performance of one. \begin{figure*}[htbp] \centerline{\psfig{file=graphs/bsys2.ps,width=6.0in}} \caption{System Performance.} \label{sysperf} \end{figure*} \begin{figure*}[htbp] \centerline{\psfig{file=graphs/pcurve.ps,width=6.0in}} \caption{Performance Curve.} \label{pcurve} \end{figure*} If the ductility of sub-components can be controlled, how should the brittleness be adjusted among the sub-components? One line of reasoning yields the result that in systems run near the maximum operating tolerance, better performance will be achieved with highly brittle components placed near the outputs of the system. This is because there is then less chance for the highly brittle components to effect the other sub-components. The next simulation, shown in Figure \ref{dloc} examines this question. The brittleness of the first component is varied from zero to one and the second component remains at a brittleness of 0.5. The results are shown in Figure \ref{fvss}. The results are also shown in the same figure for the first component brittleness of 0.5 and the second component brittleness varying from zero to one. Figure \ref{fvss} indicates that the best performance curve results when the more highly brittle component is the last component in the chain. \begin{figure*}[htbp] \centerline{\psfig{file=figures/Bsys3.ps,width=6.0in}} \caption{Brittleness Location Experiment.} \label{dloc} \end{figure*} \begin{figure*}[htbp] \centerline{\psfig{file=graphs/bvsp.ps,width=6.0in}} \caption{Brittleness Location Results.} \label{fvss} \end{figure*} \section{Sensitivity} As a first step, design parameters, $X_p$, which affect ductility must be identified. The sensitivity of ductility to a particular parameter is characterized by $\psi$, as shown in Equation \ref{pdeff} and \ref{peff}. In Equation \ref{pdeff}, $\gamma$ is the shaded area $A-B$ in Figure \ref{bintro}, which is a function of two values, $x_1$ and $x_2$, of a single design parameter, $X_p$. In Equation \ref{peff}, the sensitivity of ductility is defined as the rate of change of the difference of $A-B$. Figure \ref{esys} shows two curves for the same system, one curve which is brittle, the other robust. A function which returns the value of the ductile sensitivity is implemented in Mathematica \cite{Wolfram:MSD91} in Figure \ref{ddef}. In Figure \ref{ddef}, $\psi$ takes two arguments, $s1$ and $s2$ which are two values of a single design parameter, $x_1$ and $x_2$. The Mathematica module returns the partial derivative of $\gamma$ as shown in the bottom of Figure \ref{ddef}. In Figure \ref{duc}, the value of the ductility sensitivity is shown for the system from Figure \ref{esys} as a function of the difference between $x_1$ and $x_2$. In Figure \ref{duc}, $x_2$ is constant and $x_1$ varies. As $x_1$ and $x_2$ become equal, $\Psi$ goes to zero. This is because the performance curves become the same and the area $A-B$ disappears. Also, when the values of $x_1$ and $x_2$ are far apart, the area $A-B$ becomes large and the rate of change of the area becomes large. Note that because of the implementation of the Mathematica module which computes $\gamma$, the order of the arguments to the Mathematica function in Figure \ref{ddef} is significant. \begin{equation} \label{pdeff} \gamma(x_1,x_2) = {A - B} \end{equation} \begin{eqnarray} \psi(x_1, x_2) = {{\partial \gamma(x_1,x_2)} \over {\partial x_1}} & & \label{peff} \end{eqnarray} \begin{figure*}[htbp] \centerline{\psfig{file=graphs/examp.ps,width=6.0in}} \caption{An Example System.} \label{esys} \end{figure*} \begin{figure*}[htbp] \centerline{\psfig{file=figures/ductility.ps,width=6.0in}} \caption{Ductility Sensitivity Definition.} \label{ddef} \end{figure*} \begin{figure*}[htbp] \centerline{\psfig{file=graphs/phi2.ps,width=4.0in}} \caption{Ductile Sensitivity.} \label{duc} \end{figure*} \section{System Energy} As a digital system approaches the edges of it operating tolerance the energy required to maintain the performance increases as shown in Figure \ref{energy}. Consider quality of service on a router in a communications network. The energy required to forward a packet is routinely modeled as directly proportional to the length of the packet. As the load increases, input queues begin to fill to capacity and packets are dropped because computational energy is not sufficient to keep up with the load. In this case, performance is the probability of not dropping a packet and energy is the processing power which is directly proportional to the packet service rate, $\mu$. In an M/M/1 queue, a direct relation between performance and energy is shown in Equation \ref{perfprob} where $n$ is the expected queue size, $N$ is the maximum queue capacity and $\rho = {\lambda \over \mu}$. The result is graphed in Figure \ref{epgraph}. \begin{equation} P[n<N] = \sum_0^{N-1} \rho^n(1-\rho) \label{perfprob} \end{equation} \begin{figure*}[htbp] \centerline{\psfig{file=figures/energy.eps,width=4.0in}} \caption{System Performance and Energy.} \label{energy} \end{figure*} \begin{figure*}[htbp] \centerline{\psfig{file=graphs/energy.ps,width=4.0in}} \caption{Performance and Energy.} \label{epgraph} \end{figure*}
\section{Introduction} \mbox{}\indent The standard deep inelastic scattering picture applies when the four-momentum transfer squared from the lepton line to the hadron line ($Q^2$) is large. When virtual photon wave length increases and reaches the size of the nucleon one may expect a transition to another regime where the standard partonic model is no longer valid. In this region a kinematical constraint \cite{BK_lowq2} guarantees the vanishing of the $F_2(x,Q^2)$ structure function. This requirement is not embodied in the perturbative parton distributions. A phenomenological fit based on parton screening was proposed in \cite{MRS_lowQ2} to satisfy this condition by introducing an extra form factor. The recent low-$Q^2$ data from HERA \cite{H1_lowQ2,ZEUS_lowQ2} have triggered many phenomenological analyses. Especially interesting is the unexplored transition region. At present there is no consensus on the details of the transition mechanism. Here we concentrate on the region of somewhat larger $x$ rather than the new HERA data. We shall demonstrate that also at larger $x$ a similar transition due to vanishing partonic components at small $Q^2$ takes place, although it is not directly seen from experimental data. It is common wisdom that the vector dominance model applies at low $Q^2$ while the parton model describes the region of large $Q^2$, leading at lowest order to Bjorken scaling, and to logarithmic scaling violation in higher orders of QCD. A proposal was made in Ref.\cite{BK_model} to unify both the limits in a consistent dispersion method approach. In the traditional formulation of the VDM one is limited to large lifetimes of hadronic fluctuations of the virtual photon, i.e. small Bjorken $x < 0.1$ for the existing data. It is a purpose of this paper to generalize the model to a full range of $Q^2$ and $x$ by introducing extra phenomenological form factors to be adjusted to the experimental data. Some authors believe that it is old-fashioned to talk about VDM contribution in the QCD era. However, VDM effects appear naturally in the time-like region in the production of vector mesons in $e^+ e^-$ collisions. These effects cannot be described in terms of perturbative QCD, as in the production of resonances many complicated nonperturbative effects take place. The physics must be similar in the space-like region. We shall demonstrate that it is essential to include this contribution explicitly in order to describe the structure functions at low Q$^2$. In the next section we outline our model and discuss how to choose its basic parameters in a model independent way. In section 3 we discuss a fit of the remaining parameters to the fixed target data and present results of the fitting procedure for the proton and deuteron structure functions. In addition we compare the result of our model for $F_2^p - F_2^n$ and some subtle isovector higher-twist effects with another low-$Q^2$ model. Finally we discuss some interesting predictions which could be verified in the future. \section{The model} \mbox{}\indent As in Ref.\cite{BK_model} the total nucleon structure function is represented as a sum of the standard vector dominance part, important at small $Q^2$ and/or small Bjorken $x$, and partonic ({\it part}) piece which dominates over the vector dominance (VDM) part at large $Q^2$: \begin{equation} F_2^N(x,Q^2) = F_2^{N,VDM}(x,Q^2) + F_2^{N,part}(x,Q^2) \; . \label{F2_decomposition} \end{equation} The standard range of applicability of vector dominance contribution is limited to large invariant masses of the hadronic system ($W$), i.e. small values of $x$. In the target (nucleon) reference frame the time of life of the hadronic fluctuation is given according to the uncertainty principle as $\tau \sim 1/\Delta E$ with \begin{equation} \Delta E = \sqrt{M_V^2 + |{\bf q}|^2 } - \sqrt{q^2 + |{\bf q}|^2 } \; , \end{equation} where $M_V$ is the mass of the hadronic fluctuation (vector meson mass). In terms of the photon virtuality and Bjorken $x$ this can be expressed as \begin{equation} \Delta E = \frac{M_V^2 + Q^2}{Q^2} \cdot M_N x \; . \end{equation} As the energy transfer $\nu \rightarrow \infty$ the time of life of the hadronic fluctuaction becomes $\tau \sim \frac{Q^2}{M_V^2 m_N x}$. It is natural to expect small VDM contribution when the time of life of the hadronic fluctuation is small. We shall model this fact by introducing a form factor $\Omega(\tau) = \Omega(x,Q^2)$. Then the modified vector dominance contribution can be written as: \begin{equation} F_2^{N,VDM}(x,Q^2) = \frac{Q^2}{\pi} \sum_V \frac{M_V^4 \cdot \sigma_{VN}^{tot}(s^{1/2})} {\gamma_V^2 (Q^2 + M_V^2)^2} \cdot \Omega_V(x,Q^2) \; . \label{F2_VDM} \end{equation} In the present paper we take $\gamma$'s calculated from the leptonic decays of vector mesons which include finite width corrections \cite{IKL} $\gamma_{\rho}^2/4\pi$ = 2.54, $\gamma_{\omega}^2/4\pi$ = 20.5, $\gamma_{\phi}^2/4\pi$ = 11.7. \footnote{Please note different normalization of $\gamma$'s in comparison to \cite{BK_model}.} In general one can try different functional forms for $\Omega$. In the present analysis we shall use only exponential and Gaussian form factors \begin{eqnarray} \Omega(x,Q^2) = \exp(-\Delta E / \lambda_E) \;, \nonumber \\ \Omega(x,Q^2) = \exp(-(\Delta E / \lambda_G)^2) \; . \label{time_formfactor} \end{eqnarray} As in Ref.\cite{BK_model} we take the partonic contribution as \begin{equation} F_2^{N,part}(x,Q^2) = \frac{Q^2}{Q^2 + Q_0^2} \cdot F_2^{asymp}(\bar x,\bar Q^2) \; . \label{F2_part} \end{equation} where $\bar x = \frac{Q^2 + Q_2^2}{W^2 - m_N^2 + Q^2 + Q_2^2}$ and $\bar Q^2 = Q^2 + Q_1^2$. The $F_2^{asymp}(x,Q^2)$ above denotes the standard partonic structure function which in the leading order can be expressed in terms of the quark distributions: \begin{center} $F_2^{asymp}(x,Q^2) = x \cdot \sum_f \; e_f^2 \cdot \left[ q_f(x,Q^2) + \bar q_f(x,Q^2) \right] $. \end{center} The extra factor in front of Eq.(\ref{F2_part}) assures a correct kinematic beheviour in the limit $Q^2 \rightarrow 0$. In general $Q_0^2$, $Q_1^2$ and $Q_2^2$ can be slightly different. In the following section we shall consider different options. At large Bjorken $x$ one has to include also the so-called target mass corrections. Their origin is mainly kinematic \cite{GP76}. In our approximate treatment we substitute the Bjorken variable $x$ in the partonic distributions by the Nachtmann variable $\xi$ \cite{Nachtmann} given by: \begin{equation} \xi = \frac{2 x} {1 + \sqrt{(1 + \frac{4 M_N^2 x^2}{Q^2})} } \; , \label{Nachtmann_variable} \end{equation} which is the dominant modification. In principle $F_2^{asymp}(x,Q^2)$ could be obtained in any realistic model of the nucleon combined with QCD evolution. We leave the rather difficult problem of modeling the partonic distributions for future studies. We expect that at not too small $x >$ 0.01, the region of the interest of the present paper, the leading order Gl\"uck-Reya-Vogt (GRV) parametrization of $F_2^{p,asymp}(x,Q^2)$ and $F_2^{n,asymp}(x,Q^2)$ should be adequate. Furthermore in our opinion the parametrization \cite{GRV95} with the valence-like input for the sea quark distributions and $\bar d$ - $\bar u$ asymmetry built in incorporates in a phenomenological way nonperturbative effects caused by the meson cloud in the nucleon \cite{meson_cloud}. The total cross section for (vector meson) -- (nucleon) collision is not well known. Above meson-nucleon resonances, one may expect the following approximation to hold: \begin{eqnarray} \sigma_{\rho N}^{tot} &=& \sigma_{\omega N}^{tot} = \frac{1}{2} \left[ \sigma_{\pi^{+} p}^{tot} + \sigma_{\pi^{-} p}^{tot} \right] \; , \nonumber \\ \sigma_{\phi N}^{tot} &=& \sigma_{K^+ p}^{tot} + \sigma_{K^- p}^{tot} - \frac{1}{2} \left[ \sigma_{\pi^{+} p}^{tot} + \sigma_{\pi^{-} p}^{tot} \right] \; . \end{eqnarray} Using a simple Regge-inspired parametrizations by Donnachie-Landshoff \cite{DL92} of the total $\pi N$ and $K N$ cross sections we get simple and economic parametrizations for energies above nucleon resonances \begin{eqnarray} \sigma_{\rho N}^{tot} &=& \sigma_{\omega N}^{tot} = 13.63 \cdot s^{0.0808} + 31.79 \cdot s^{-0.4525} \; , \nonumber \\ \sigma_{\phi N}^{tot} &=& 10.01 \cdot s^{0.0808} + 2.72 \cdot s^{-0.4525} \; , \end{eqnarray} where the resulting cross sections are in mb. We expect that our model should be valid in a broad range of $x$ and $Q^2$ except for very small $x < 0.001$, where genuine effects of BFKL pomeron physics could show up, and except for very large $x$, where the energy ($s^{1/2}$ in Eq.(\ref{F2_VDM})) is small and the behaviour of the total $VN$ cross section is essentially unknown. Because our main interest is in the transition region, the large $Q^2$ data were not taken into account in the fit. There the partonic contribution is by far dominant and the GRV parametrization \cite{GRV95} is known to provide a reliable description of the data. \section{Comparison to experimental data} \mbox{}\indent Most of the previous parametrizations in the literature centered on the proton structure function. In the present analysis we are equally interested in both proton and neutron structure functions. Achieving this goal requires special selection of the experimental data with similar statistics and similar range in ($x$,$Q^2$) for proton and deuteron structure function. In Fig.1 we display the experimental data for proton (left panel) and deuteron (right panel) structure functions chosen in our fit. We have selected only NMC, E665 and SLAC sets of data \cite{F2_data} for both proton and deuteron structure functions, together amounting to 1833 experimental points: 901 for the proton structure function and 932 for the deuteron structure function. According to the arguments presented above, we have omitted BCDMS and HERA data in the fitting procedure but these will be compared to our parametrization when discussing the quality of the fit. The deuteron structure function has been calculated as \begin{equation} F_2^d(x,Q^2) = {1 \over 2} [ F_2^p(x,Q^2) + F_2^n(x,Q^2) ] \; , \label{F2d} \end{equation} i.e. we have neglected all nuclear effects such as shadowing, antishadowing due to excess mesons, Fermi motion, binding, etc, which are known to be relatively small for the structure function of the deuteron \cite{BK92,Z92,MT93}, one of the most loosely bound nuclear systems. In addition we have assumed isospin symmetry for the proton and neutron quark distributions, i.e. $u_n(x,Q^2) = d_p(x,Q^2)$, $d_n(x,Q^2) = u_p(x,Q^2)$ and $s_n(x,Q^2) = s_p(x,Q^2)$. The charm contribution, which in the GRV parametrization \cite{GRV95} is due to the photon-gluon fusion, is in practice negligible in the region of $x$ and $Q^2$ taken in the fit, and therefore is omitted throughout the present analysis. The results of the fit are summarised in Table 1. Because in general $Q_0^2$, $Q_1^2$ and $Q_2^2$ can be different, there are 4 independent free parameters of the model. In order to limit the number of parameters we have imposed extra conditions as specified in Table 1. A series of seven fits has been performed. In all cases considered the number of free parameters has been reduced to two: the cut-off parameter of the form factor and $Q_0^2$. Statistical and systematic errors were added in quadrature when calculating $\chi^2$. Only data with $Q^2 >$ 0.25 GeV$^2$ were taken in the fit which is connected with the domain of applicability of the GRV parametrization. The values of the parameters found are given in each case in parentheses below the value of $\chi^2$ per degree of freedom. In addition to the combined fit, which includes both proton $F_2^p$ and deuteron $F_2^d$ structure function data, we show the result of the fit separately for proton and deuteron structure functions. As can be seen from the table fairly similar values of parameters are found for the proton and deuteron structure function and the $\chi^2$ per degree of freedom is slightly worse in the latter case which can be due to the omission of nuclear effects as mentioned above. The best fit (No 1 in the table) is obtained with $Q_1^2 = Q_2^2 = 0$ (fits of similar quality can be obtained with very small values of $Q_1^2 \sim$ 0.1 GeV$^2$ and $Q_2^2 \sim$ 0.1 GeV$^2$). While the value of $\chi^2$ does not practically depend on the type of the form factor (exponential vs. Gaussian), much larger value of $Q_0^2$ is found for the Gaussian ($Q_0^2$ = 0.84 GeV$^2$) than for the exponential ($Q_0^2$ = 0.52 GeV$^2$) parametrization. The value of $Q_0^2$ found here is smaller than in the original Bade{\l}ek-Kwieci\'nski model \cite{BK_model} but larger than that found by H1 collaboration in the fit to low-$x$ data \cite{H1_lowQ2}. Although the resulting $\chi^2$ is similar in both cases, the $F_2^n(x) / F_2^p(x)$ ratio for $x \rightarrow 1$ prefers the Gaussian form factor. While the vector meson contribution with the exponential form factor survives up to relatively large $x$, with the Gaussian form factor it is negligible at large $x$.. For comparison the GRV parametrization of quark distributions alone yields:\\ $\chi^2/N_{dof}$ = 9.74 (21.48) (proton structure functions),\\ $\chi^2/N_{dof}$ = 13.73 (32.99) (deuteron structure functions),\\ $\chi^2/N_{dof}$ = 11.77 (27.33) (combined data),\\ where the first numbers include target mass corrections and for illustration in parentheses their counterparts without target mass corrections are given. Clearly the inclusion of the target mass effects is essential and only such results will be discussed in the course of the present paper. The agreement of the CKMT parametrization is comparable to that obtained in our model. For instance for parametrization (b) in Table 2 in the second paper \cite{CKMT}, which includes new HERA data:\\ $\chi^2/N_{dof}$ = 2.22 (1.00) (proton structure functions),\\ $\chi^2/N_{dof}$ = 3.54 (3.59) (deuteron structure functions),\\ $\chi^2/N_{dof}$ = 2.89 (2.33) (combined data),\\ where in the parentheses we present $\chi^2$ for $Q^2 <$ 4 GeV$^2$ i.e. in the region of applicability of the CKMT parametrization. \footnote{The number of experimental points is reduced then to 354 and 373 for proton and deuteron structure functions, respectively} We note much better description of the proton data in comparison to the deuteron data. The agreement of the Donnachie-Landshoff parametrization \cite{DL94} with the proton structure function data is of similar quality. In Fig.2 we present for completeness a map of $\chi^2$ for our best fit as a function of model parameters $Q_0^2$ and $\lambda$. A well defined minimum of $\chi^2$ for $\lambda_G \approx$ 0.5 GeV and $Q_0^2 \approx$ 0.85 GeV$^2$ can be seen. The experimental statistical uncertainty of the obtained parameters $\lambda_G$ and $Q_0^2$ is less than 1 \%. Some examples of the fit quality can be seen in Fig.3 (x-dependence for different values of $Q^2$ = 0.585, 1.1, 2.0, 3.5 GeV$^2$) and in Figs.4, 5 ($Q^2$-dependence for different values of Bjorken $x$ = 0.00127, 0.0125, 0.05, 0.10, 0.18, 0.35, 0.55, 0.75). Shown are experimental data \cite{F2_data} which differ from the nominal $Q^2$ or Bjorken $x$ specified in Fig.3, 4, 5 by less than $\pm$ 3 \%. An excellent fit is obtained for $Q^2 >$ 4 GeV$^2$ (not shown in Fig.3), although the VDM contribution stays large up to 10 GeV$^2$. In comparison to the GRV parametrization (dashed line) our model describes much better the region of small $Q^2 <$ 3 GeV$^2$, especially at intermediate Bjorken $x$: 0.05 $< x <$ 0.4. The CKMT model (long-dashed line), shown according to the philosophy in \cite{CKMT} for $Q^2 <$ 10 GeV$^2$ gives a better fit at very small Bjorken $x$. However, one may expect here a few more effects which will be discussed below. \footnote{The use of the next-to-leading order structure functions \cite{GRV95} in our model would improve the description of low-$x$ data, discussion of which is left for a separate analysis.} It is however slightly worse as far as isovector quantities are considered, as will be discussed later. There seems to be a systematically small (up to about 5 \%) discrepancy between our model and the data for $Q^2 <$ 2 GeV$^2$ and $x$ = 0.1 - 0.3. This is caused by some higher-twist effects due to the production of the $\pi N$ \cite{S72} and $\pi \Delta$ exclusive channels and will be discussed elsewhere. A fit of similar quality is obtained in our model for the proton (left panels) and deuteron (right panels) structure functions. Rather good agreement of our model with the BCDMS data can be observed in Figs.4 and 5 in spite of the fact that the data were not used in the fitting procedure. At very small $x <$ 0.01 the description of the data becomes worse. This is partially due to the use of the leading order approximation. The fit to the fixed-target data prefers $\bar x \approx x$ and $\bar Q^2 \approx Q^2$ (see Table 1 and the discussion therein). On the other hand, the HERA data would prefer $Q_1^2 \ne $ 0 and $Q_2^2 \ne $ 0. If we included the HERA data in the fit the description of the fixed target data would become worse. There are no fundamental reasons for the parameters in both regions to be identical. In addition at very small $x$ other effects of isoscalar character, not included here, such as heavy long-lived fluctuactions of the incoming photon \cite{Shaw} and/or BFKL pomeron effects \cite{BFKL}, may become important. For illustration a VDM contribution modified by a form factor (\ref{time_formfactor}) is shown separately by the short-dashed line. The partonic component can be obtained as a difference between the solid and VDM line. It can be seen from Figs.3-5 that the partonic component decreases towards small $Q^2$. This decrease is faster than one could directly infer from the failure of the GRV parametrization at low $Q^2$ because in our model a part of the strength resides in the VDM contribution. The modified VDM contribution is sizeable for small values of Bjorken $x$ and not too large $Q^2$ and survives up to relatively large $Q^2$. At $Q^2 >$ 3.5 GeV$^2$ the structure functions in our model almost coincide with those in the GRV parametrization despite that the VDM term is still not small. For $Q^2 \rightarrow \infty$ only the partonic contribution survives and $F_2(x,Q^2) \rightarrow F_2^{part}(x,Q^2) \rightarrow F_2^{GRV}(x,Q^2)$. The deviations from the partonic model can be also studied in the language of higher-twist corrections. Then the structure function can be written formally as \begin{equation} F_2^{p/n}(x,Q^2) = F_2^{p/n,LT}(x,Q^2) \left[ 1 + \frac{ c_2^{p/n}(x) }{Q^2} + \frac{c_4^{p/n}(x)}{Q^4} + ... \right] \; . \label{ht_expansion} \end{equation} However, in empirical analyses one usually includes only one term in (\ref{ht_expansion}) \begin{equation} F_2^{p/n}(x,Q^2) = F_2^{p/n,LT}(x,Q^2) \left[ 1 + \frac{c^{p/n}(x) }{Q^2} \right] \; . \label{ht_trunc} \end{equation} In our model for $Q^2 \sim M_V^2, Q_0^2$ there are an infinite number of active terms in the expansion of the structure function (\ref{ht_expansion}). Therefore the coefficient $c^{p/n}$ (the same is true for the deuteron counterpart $c^d$) in (\ref{ht_trunc}) becomes effectively $Q^2$-dependent $c^{p/n}(x) = c^{p/n}(x,Q^2)$. As an example in Fig.6 we show $c^p$ and $c^d$ as a function of Bjorken $x$ for three different values of $Q^2$ = 2, 4, 8 GeV$^2$ in the range relevant for the analysis in \cite{VM92}. In order to correctly compare our results for $c^p$ and $c^d$ with the results of the analysis in \cite{VM92} the structure function $F_2^{p/n,LT}(x,Q^2)$ in Eq.(\ref{ht_trunc}) will include complete target mass corrections calculated according to Ref.\cite{GP76}. A fairly similar pattern is obtained for $c^p$ and $c^d$ especially at small $x$. The rise of $c^p$ or $c^d$ for $x \rightarrow$ 1 is caused by our treatment of the target mass corrections and partially by the VDM contribution which survives in our model up to relatively large x. We obtain small negative $c^p$ and $c^d$ for $x <$ 0.3 in agreement with \cite{VM92}. The smallnest of $c^p$ and $c^d$ in our model for x $<$ 0.3 is due to the cancellation of the positive VDM contribution and a negative contribution caused by the external damping factor $\frac{Q^2}{Q^2+Q_0^2}$ of the partonic contribution in Eq.(\ref{F2_part}). The CKMT parametrization \cite{CKMT} (shown only in its applicability range for $Q^2$ = 2, 4 GeV$^2$) provides a very good explanation of $c^{p}$. It predicts, however, somewhat larger negative $c^{d}$ for x $<$ 0.3. This will have some unwanted consequences for $c^p - c^n$ discussed below. In Fig.7 we show $c^p-c^n$ for $Q^2$ = 2, 4 GeV$^2$ together with empirical results from \cite{NMC_ht}. Rather strong $Q^2$-dependence of $c^p-c^n$ is observed. Our model correctly describes the trend of the experimental data. For comparison we show also the result obtained by means of the CKMT parametrization (long-dashed lines) of the structure functions which somewhat fails to reproduce the details of the empirical results from \cite{NMC_ht}, especially for small Bjorken $x$. To our best knowledge no other model in the literature is able to describe quantitatively this subtle higher-twist effect. Our model seems to provide a very good description of some isovector quantities. As an example in Fig.8 we present $F_2^p(x,Q^2) - F_2^n(x,Q^2)$ at $Q^2$ = 4 GeV$^2$ obtained in our model (solid lines for different form factors), as well as the results obtained with the GRV parametrization (dashed line) and in the CKMT model (long-dashed line).\footnote{ No evolution of the CKMT quark distributions was done, but it is negligible between 2 $\rightarrow$ 4 GeV$^2$ for the nonsinglet quantity. } The NMC data \cite{NMC} prefer rather our model. As a consequence of the imperfect description of the deuteron data the CKMT model fails to describe the difference $F_2^p(x) - F_2^n(x)$ for $x <$ 0.3. The success of our model is related to the violation of the Gottfried Sum Rule and/or $\bar d - \bar u$ asymmetry which is included in our model explicitly. In contrast to our model in the CKMT model for $Q^2 >$ 2 GeV$^2$ the Gottfried Sum Rule $S_G = {1 \over 3}$. \section{Conclusions and discussion} \mbox{}\indent We have constructed a simple model incorporating nonperturbative structure of the nucleon and photon. Our model is a generalization of the well known and successful Bade{\l}ek-Kwieci\'nski model \cite{BK_model}. While the original Bade{\l}ek-Kwieci\'nski model is by construction limited to a small-$x$ region, our model is intended to be valid in much broader range. The original VDM model assumes implicitly a large coherence length for the photon-hadron fluctuation, i.e. assumes that the hadronic fluctuation is formed far upstreem of the target. When the fluctuation length becomes small the VDM is expected to break-down. This effect has been modelled by introducing an extra form factor. As a result we have succeeded in constructing a physically motivated parametrization of both proton and deuteron structure functions. In comparison to the pure partonic models with QCD evolution our model leads to a much better agreement at low $Q^2$ in a broad range of $x$. With only two free parameters we have managed to describe well the transition from the high- to low-$Q^2$ region simultaneously for the proton and deuteron structure functions. Our analysis of the experimental data indicates that the QCD parton model begins to fail already at $Q^2$ as high as about 3 GeV$^2$. This value is larger than commonly believed. In our discussion we have omitted the region of the HERA data. In our opinion the physics there may be slightly more complicated. Other effects of isoscalar character, not included in our analysis, for example the BFKL pomeron effects \cite{BFKL}, may become important. In contrast to other models in the literature we obtain a very good description of the NMC $F_2^p - F_2^n$ data \cite{NMC} where the previously mentioned isoscalar effects cancel. Recently an intriguing, although small, difference between $\bar d - \bar u$ asymmetry obtained from recent E866 Drell-Yan data \cite{E866} and muon DIS NMC data \cite{NMC} has been observed. At least part of the difference can be understood in our model. We expect for $Q^2$ smaller than about 4 GeV$^2$ an extra $Q^2$ dependence of some parton model sum rules. We predict a rather strong $Q^2$ effect for the integrand of the Gottfried Sum Rule where in the first approximation the VDM contribution cancels. Recently in the literature there has been sizeable activity towards a better understanding of higher-twist effects. Some were estimated within the operator product expansion, some in terms of the QCD sum rules. It is, however, rather difficult to predict the absolute normalization of the higher-twist effects. Our model leads to relatively large higher-twist contributions. For some observables, like structure functions, they almost cancel. For other observables, like $F_2^p - F_2^n$, the cancellation is not so effective. Our model provides specific higher-twist effects not discussed to date in the literature. This will be a subject of a future separate analysis. \vskip 1cm {\bf Acknowledgments:} \\ We are especially indebted to J. Kwieci\'nski for valuable discussions and suggestions and J. Outhwaite for careful reading of the manuscript. We would also like to thank C. Merino for the discussion of the details of the CKMT model. This work was supported partly by the German-Polish exchange program, grant No. POL-81-97. \newpage
\section*{Acknowledgments} U.T. is a TUBITAK M\"{u}nir Birsel Foundation Fellow and acknowledges partial support from Ege University Research Fund under the Project Number 97 FEN 025. During the course of this research, D.F.T. was a Chevening Scholar of the British Council Foundation and acknowledges partial support from CONICET and Fundaci\'on Antorchas.
\section{Introduction} The old problem of magnetic structure of rare-earth metals and their compounds is still a subject of experimental and theoretical investigations. These substances have complicated phase diagrams and demonstrate a number of orientational phase transitions. In particular, such transitions take place in the orthoferrites and practically important intermetallic compounds RCo$% _5 $ (R=Pr,Nd,Tb,Dy,Ho), see, e.g., Ref. \cite{Levitin}. Qualitative explanation of these transitions has been obtained many years ago within the Heisenberg model with inclusion of magnetic anisotropy \cite{Coqblin}. In a number of systems, lattice (magnetoelastic) effects are important. The standard description is usually performed within mean-field approaches. However, quantitative comparison with experimental data requires a more detailed treatment. Provided that the orientational transition temperature is low (in comparison with the magnetic ordering point), spin-wave theory is applicable \cite {Coqblin}. In the simplest case of the second-order anisotropy the magnetization lies either along the easy axis or in the easy plane. Inclusion of higher-order anisotropy constants can lead to cone phases where magnetization makes the angle $\theta $ with the $z$-axis. The case $\theta =\pi /2$ was considered in Refs. \cite{Tb,TbDy1,GdTbDy,TbDy2} where the temperature renormalization of anisotropy constants and the spin-wave spectrum in Tb and Dy within the standard spin-wave theory were calculated. In the present paper we consider the cone phase with arbitrary $0\leq \theta \leq \pi /2$. The situation here is analogous to that for the field-induced orientational phase transitions, e.g., in the transverse-field Ising model (see Ref.\cite{OurITF} and references therein). Unlike the latter model, one can expect that at low enough temperatures and small value of anisotropy the spin-wave theory is applicable for an arbitrary relation between anisotropy parameters, not only close to the orientational phase transition. Even for the second-order easy-plane anisotropy, the Holstein-Primakoff representation for spin operators used in Refs. \cite{Tb,GdTbDy,TbDy2} leads to so-called kinematical inconsistencies because of incorrect treating on-site kinematical relations. To avoid this difficulty, we use the technique of spin coherent states. Our approach is to some extent similar to the operator approach used in Ref. \cite{TbDy1}, but gives a possibility to treat more simply higher-order anisotropy constants, as well as to calculate higher-order terms of $1/S$-expansion. The anisotropic Heisenberg model used is formulated in Sect 2. In Sect 3 we develop a special form of the $1/S$-expansion which gives a possibility to take into account exactly on-site kinematical relations. In Sect 4 we treat the cone-plane transition owing to the temperature dependence of the cone angle $\theta $ and discuss experimental data on the rare earth metals. \section{The model and mean-field approximation} We start from the Hamiltonian of the Heisenberg model with inclusion of single-site anisotropy \begin{equation} {\cal H}=-\frac J2\sum_{\langle ij\rangle }{\bf S}_i{\bf S}% _j+B_2^0\sum_i(O_2^0)_i+B_4^0\sum_i(O_4^0)_i \label{H} \end{equation} where $J>0$ is the exchange parameter, \begin{eqnarray} O_2^0 &=&3(S^z)^2-S(S+1) \nonumber \\ O_4^0 &=&35(S^z)^4-30S(S+1)(S^z)^2+25(S^z)^2 \nonumber \\ &&\ +3S^2(S+1)^2-6S(S+1) \label{Tens} \end{eqnarray} are the irreducible tensor operators of second and fourth orders, $B_l^m$ are the corresponding anisotropy constants. Up to unimportant constant we can rewrite the Hamiltonian (\ref{H}) in the form \begin{equation} {\cal H}=-\frac J2\sum_{\langle ij\rangle }{\bf S}_i{\bf S}% _j+D\sum_i(S_i^z)^2+D^{\prime }\sum_i(S_i^z)^4 \end{equation} where \begin{eqnarray} D &=&3B_2^0-[30S(S+1)-25]B_4^0 \label{rel} \\ D^{\prime } &=&35B_4^0 \nonumber \end{eqnarray} For $D,D^{\prime }>0$ spins of magnetic ions lie in the easy plane $xy$, while for $D,D^{\prime }<0$ we have the easy axis $z.$ For $D>0,\;D^{\prime }<0$ a first-order transition takes place between the easy plane (which is favored by second-order anisotropy) and easy axis (which is favoured by large $|D^{\prime }|$). We consider only the case $D<0,\;D^{\prime }>0$ where the cone phase occurs at intermediate values of $D/(2D^{\prime }S^2)$, so that spin orientation direction makes the angle $\theta $ with the $z$% -axis and the orientational phase transitions are of the second order. This is the case for Gd and also for Ho, Er in low-temperature phases. In the phenomenological approach it is supposed (see, e.g., Refs. \cite {Levitin,Coqblin}) \begin{equation} F=F_{\text{is}}+D(T)(S\cos \theta )^2+D^{\prime }(T)(S\cos \theta )^4 \end{equation} where $F_{\text{is}}$ is the isotropic ($\theta $-independent) part of the free energy. Then we obtain by minimization of $F$% \begin{equation} \cos ^2\theta (T)=-\frac{D(T)}{2D^{\prime }(T)S^2} \label{ct} \end{equation} so that at the point where $D(T)=0$ the spins become directed in the $xy$ plane while at $|D(T)|\geq 2D^{\prime }(T)S^2$ they are aligned along the $z$% -axis. The temperature dependence of $D(T)$ is supposed to have the form \begin{equation} D(T)=2D^{\prime }S^2(T_1-T)/(T_2-T_1) \label{PhT} \end{equation} with $D^{\prime }(T)>0.$ Thus at $T=T_1$ the transition from the easy-plane to cone structure takes place, while at $T=T_2$ the transition from the cone to easy-axis structure occurs. At the same time, Zener's \cite{Zener} resut for the temperature dependence of anisotropy constants in an axially symmetric state with $\theta =0$ has the form \begin{equation} B_l^0(T)=B_l^0M^{l(l+1)/2} \label{Zener} \end{equation} where $M=\langle \widetilde{S}^z\rangle /S$ is the relative magnetization, and $D(T),$ $D^{\prime }(T)$ are determined by the same relations (\ref{rel}% ) with $B_l^0\rightarrow B_l^0(T)$. As pointed in Refs. \cite {Tb,TbDy1,GdTbDy,TbDy2}, the temperature dependences of anisotropy constants have a more complicated form for the cone structures with $\theta >0$ (in fact, only the case $\theta =\pi /2$ is discussed in Refs. \cite {Tb,TbDy1,GdTbDy,TbDy2}). A systematic way of calculating temperature dependences of anisotropy constants is the $1/S$-expansion which is considered in the next section. \section{The $1/S$ expansion of the partition function} The $1/S$-expansion developed here is slightly different from the standard scheme of $1/S$-expansion \cite{Tb,GdTbDy,TbDy2} since it gives a possibility to take into account exactly the kinematical relations between powers of spin operators on each site. We use the coherent state approach (see, e.g., Ref. \cite{Arovas}) to write down the partition function in the form \begin{equation} {\cal Z}=\int D\mbox {\boldmath $\pi $}\exp \left\{ iS\int\limits_0^\beta d\tau (1-\cos \vartheta )\frac{\partial \varphi }{\partial \tau }-\langle \pi |{\cal H}|\pi \rangle \right\} \label{Zp} \end{equation} where $\mbox {\boldmath $\pi $}$ is the unit-length vector, $\vartheta $ and $\varphi $ are its polar and azimuthal angles respectively, $|\pi \rangle =\exp (i\vartheta S^y+i\varphi S^z)|S\rangle $ are the coherent states ($% S^z|S\rangle =S|S\rangle $). To construct the $1/S$-expansion we rotate the coordinate system around the $y$-axis through the angle $\theta .$ The Hamiltonian (\ref{H}) takes the form \begin{eqnarray} {\cal H} &=&-\frac J2\sum_{\langle ij\rangle }{\bf S}_i{\bf S}_j \\ &&\ \ \ +\sum_i\sum_{l,m}\sum_{m^{\prime }=-l}^lB_l^m\sqrt{\frac{% (l+m)!(l-|m^{\prime }|)!}{(l-m)!(l+|m^{\prime }|)!}}\frac{A_l^m}{% A_l^{|m^{\prime }|}}d_{mm^{\prime }}^l(\theta )(\widetilde{O}_l^{|m^{\prime }|})_i \nonumber \end{eqnarray} where $d_{mm^{\prime }}^l(\theta )$ are the Wigner matrices of the rotation group irreducible representation, \begin{equation} A_l^m=\frac{(l-m)!}{(l+[m]-1)!!}\frac 1{K_l^m} \end{equation} ($[m]=m$ for $m$ even and $[m]=m+1$ for $m$ odd), for $l\leq 4$ we have $% K_l^m=1,$ and the tilde sign here and hereafter is referred to the rotated coordinate system. Since the partition function (\ref{Zp}) is invariant under rotation of the states $|\pi \rangle ,$ it is convenient to use the coherent states defined in the same coordinate system, i.e., $|\widetilde{% \pi }\rangle =\exp (i\vartheta \widetilde{S}^y+i\varphi \widetilde{S}^z)|% \widetilde{S}\rangle \;$with $\widetilde{S}^z|\widetilde{S}\rangle =S|% \widetilde{S}\rangle .$ The advantage of using the coherent states is the simple form of the averages of the tensor operators (\ref{Tens}) over $|% \widetilde{\pi }\rangle .$ By direct calculation we obtain \begin{equation} \langle \widetilde{\pi }|\widetilde{O}_l^m|\widetilde{\pi }\rangle =S_lA_l^mP_l^m(\cos \vartheta )\cos m\varphi \label{Av} \end{equation} where $P_l^m(x)$ are the associated Legendre polynomials, the factors $% S_l=S(S-1/2)...[S-(l-1)/2]$ take into account properly the kinematical relations on each site. In particular, the second-order anisotropy term vanishes for $S=1/2,$ and the fourth-order for $S=1/2,1,3/2,$ as it should be (unlike the results of boson representations in Refs.\cite {Tb,GdTbDy,TbDy2}). Using (\ref{Av}) we obtain for the case $% B_l^m=B_l^0\delta _{m0}$ under consideration the result \begin{eqnarray} \langle \widetilde{\pi }|{\cal H}|\widetilde{\pi }\rangle &=&-\frac{JS^2}% 2\sum_{\langle ij\rangle }\widetilde{\mbox {\boldmath $\pi $}}_i\widetilde{% \mbox {\boldmath $\pi $}}_j \label{Cs} \\ &&\ \ \ +\sum_i\sum_{l=2,4}\sum_{m=-l}^lS_lB_l^0A_l^0\frac{(l-|m|)!}{(l+|m|)!% }P_l^{|m|}(\cos \theta )P_l^{|m|}(\cos \vartheta )\cos m\varphi \nonumber \end{eqnarray} Further calculations are performed in the same line as in Ref. \cite{OurITF}% . Representing $\cos \vartheta =\sqrt{1-\sin ^2\vartheta }$ and expanding in $\sin \vartheta $ we obtain the $1/S$-expansion of the partition function. It should be stressed that we retain the factors $S_l,$ as well as $S$% -dependences in (\ref{rel}), non-expanded. By performing decouplings, terms of third order are reduced to linear ones, and terms of fourth order to quadratic ones. The requirement of absence of $\sin \vartheta $-linear terms leads to the result for the cone angle $\theta $ \begin{equation} \cos ^2\theta =\frac 37\left[ 1-X+Y-\frac 1{10}\frac{B_2^0S_2}{B_4^0S_4}% \left( 1-\frac 7{2S}+6X+Y\right) \right] \label{Teta} \end{equation} where \begin{eqnarray} X &=&\langle \pi _{xi}^2\rangle \equiv \langle \sin ^2\vartheta \cos ^2\varphi \rangle =\sum_{{\bf q}}\frac{J_0-J_{{\bf q}}}{2E_{{\bf q}}}\coth \frac{E_{{\bf q}}}{2T}, \nonumber \\ \;Y &=&\langle \pi _{yi}^2\rangle \equiv \langle \sin ^2\vartheta \sin ^2\varphi \rangle =\sum_{{\bf q}}\frac{J_0-J_{{\bf q}}+\Delta _0/S}{2E_{{\bf % q}}}\coth \frac{E_{{\bf q}}}{2T}, \end{eqnarray} and the ``bare'' magnon spectrum reads \begin{eqnarray} E_{{\bf q}} &=&S\sqrt{(J_0-J_{{\bf q}})(J_0-J_{{\bf q}}+\Delta _0/S)} \label{Eq} \\ \Delta _0 &=&2\left[ 3B_2^0S_2\cos 2\theta -10B_4^0S_4(28\cos ^4\theta -27\cos ^2\theta +3)\right] , \nonumber \end{eqnarray} $\Delta _0$ being the energy gap. The corrections in (\ref{Teta}) can be collected into powers in the same way as in Refs.\cite{Rast2,OurSSWT} to obtain the correct description of thermodynamics at not too low temperatures. (In the presence of higher-order anisotropy this is essential since the coefficients at $X,Y$ increase as $\sim l^2/2$ with anisotropy order.) Then we have \begin{equation} \cos ^2\theta =\frac 37\frac{Z_X}{Z_Y}\left[ 1-\frac 1{10}\frac{B_2^0(T)S_2}{% B_4^0(T)S_4}\right] \label{Teta1} \end{equation} where \begin{equation} B_2^0(T)=Z_X^2Z_YB_2^0,\;B_4^0(T)=Z_X^9Z_YB_4^0 \label{AnisRen} \end{equation} are the temperature-renormalized anisotropy constants, \begin{equation} Z_X=1+\frac 1{2S}-X,\;Z_Y=1+\frac 1{2S}-Y \end{equation} The relations (\ref{AnisRen}) extend the results of Refs. (\cite {Tb,TbDy1,GdTbDy,TbDy2}) to the case where spins make a non-zero angle with the $z$-axis. The renormalized gap has the form \begin{eqnarray} \Delta &=&6\cos 2\theta \,B_2^0S_2-20B_4^0S_4\left[ 3(1-7\widetilde{X}% )-3\,\cos ^2\theta \,\left( 9-56\widetilde{X}-7\widetilde{Y}\right) \right. \nonumber \\ &&\ \ \ \ \ \ \left. +28\,\cos ^4\theta \,\left( 1-6\widetilde{X}-\widetilde{% Y}\right) \right] -196\sin ^2\theta \cos ^2\theta \nonumber \\ &&\ \ \ \ \ \ \ \times \sum_{k,\omega _n}\left[ \frac{3B_2^0S_2(J_0-J_{{\bf k% }})-10B_4^0(S_4/S)\Delta _0\cos ^2\theta }{\omega _n^2+S^2(J_0-J_{{\bf k}% })(J_0-J_{{\bf k}}+\Delta _0/S)}\right] ^2 \label{SpRen} \end{eqnarray} where $\widetilde{X}=X-1/(2S),$ $\widetilde{Y}=Y-1/(2S)$. After introducing the temperature-renormalized second- and fourth order anisotropy parameters $% D(T)$ and $D^{\prime }(T),$ \begin{eqnarray} D(T)S^2 &=&3B_2^0(T)S_2-30B_4^0(T)S_4, \label{rel1} \\ D^{\prime }(T)S^4 &=&35B_4^0(T)S_4\left( Z_Y/Z_X\right) , \nonumber \end{eqnarray} the expression for $\cos \theta $ coincides with that of the phenomenological theory (\ref{ct}). Collecting again corrections in (\ref {SpRen}) into powers, we obtain for the renormalized gap in the notations (% \ref{rel1}) the expression \begin{eqnarray} \Delta &=&2D(T)S^2\cos 2\theta +2D^{\prime }(T)S^4\cos ^4\theta +6D^{\prime }(T)S^4\sin ^2\theta \cos ^2\theta \nonumber \\ &&\ \ \ \ -196\sin ^2\theta \cos ^2\theta \sum_{k,\omega _n}\left[ \frac{% 3B_2^0S_2(J_0-J_{{\bf k}})-10B_4^0(S_4/S)\Delta _0\cos ^2\theta }{\omega _n^2+S^2(J_0-J_{{\bf k}})(J_0-J_{{\bf k}}+\Delta _0/S)}\right] ^2 \end{eqnarray} which also coincides with that obtained in the phenomenological theory except for the last term. Note that at $\theta >0$ the renormalizations (\ref {rel1}) are present even at $T=0,$ which should be taken into account when treating experimental data. \section{Orientational phase transitions} Now we can pass to description of possible orientational phase transitions. Consider first the case of a small enough constant $B_4^0$ (or, equivalently, $D^{\prime }$), so that $\cos ^2\theta (0)\;$ is close to unity. Then $\cos ^2\theta (T)$ increases with temperature and there occurs a transition to the easy-axis phase at the point determined by \begin{equation} \frac 37\frac{Z_X}{Z_Y}\left[ 1-\frac 1{10}\frac{B_2^0(T)S_2}{B_4^0(T)S_4}% \right] =1 \label{PhT1} \end{equation} In the opposite case of a large enough $B_4^0,$ $\cos ^2\theta (0)$ is small, and $\cos ^2\theta (T)$ decreases with temperature, so that at the point where \begin{equation} B_2^0(T)S_2=10B_4^0(T)S_4 \label{PhT2} \end{equation} a phase transition to the easy-plane phase occurs. Thus one can expect that there exists the critical value $\theta _c$: for $\theta _0=\theta (0)<\theta _c$ we have a decrease of $\theta (T)$ with $T$ and the phase transition from cone to easy-axis phase, while for $\pi /2>\theta _0>\theta _c$ we have an increase of $\theta (T)$ with $T$ and the phase transition from cone to easy-plane phase. The numerical computations for the simple cubic lattice (see Fig.1) yield $\theta _c\simeq 50^{\circ }$. Fig.2 shows the corresponding temperature dependences of the anisotropy constants $D(T),$ $D^{\prime }(T)$. For simplicity, $J_{{\bf q}}$ is taken for the simple cubic lattice. The phase transitions described by Eqs. (\ref{PhT1}) and (\ref{PhT2}) are analogous to those in the phenomenological theory of Ref. \cite{Levitin} that occur at $D(T)=0$ and $D(T)=-2D^{\prime }(T)S^2,$ respectively. However, unlike the phenomenological approach, microscopical consideration leads to either cone to easy-axis or cone to easy-plane transition with increasing temperature, depending on the zero-temperature value of $\theta .$ At the same time, the transition from the easy-plane to easy-axis structure (through the intermediate cone phase) cannot be explained by purely magnetic renormalizations of anisotropy constants. The result (\ref{Teta}) gives the mean-field values of the critical exponents (e.g., $\beta =1/2$) for both the ground-state and temperature orientational phase transitions. Unlike the systems discussed in Ref. \cite {OurITF}, the system under consideration has the dynamical critical exponent $z=2$ (i.e., single excitation mode with nearly quadratic dispersion is present). Thus the upper critical dimensionality for the ground-state QPT is $d_c^{+}=4-z=2.$ In this respect, the system is analogous to $XY$ model in the transverse magnetic field \cite{XY-TF}. A characteristic feature of such systems is the mean-field behavior of critical exponents both above and below the critical dimensionality. For (hypothetical) systems with $d=2$ logarithmic corrections to ground-state properties near QPT are present (see, e.g., Ref. \cite{Sachdev}). At the same time, the upper critical dimensionality for the temperature phase transition is $d_{cT}^{+}=4,$ and at $d<d_{cT}^{+}$ the temperature-transition critical exponents differ from their mean-field values. Now we discuss the experimental situation. In Gd (see, e.g., Refs. \cite {Coqblin,JensenBook}) the orientational phase transition from cone-phase to easy-axis phase is observed at $T_c=240$K. The temperature dependence of the cone angle at $T<T_c$ (and also of magnetic anisotropy constants) is non-monotonous, unlike the results obtained in Sect 2. This complicated situation is connected with the absence of orbital momentum and smallness of anisotropy in gadolinium. In holmium the low-temperature phase is conical spiral one, the angle of the cone changing from $\approx 80^{\circ }$ to $90^{\circ }$ in the temperature interval $0-20K.$ The spiral angle makes up about $30^{\circ }.$ Since the sixth-order anisotropy is important, we use the Hamiltonian \cite{Jensen} \begin{equation} {\cal H}_{\text{Ho}}={\cal H}+B_6^0\sum_i(O_6^0)_i+B_6^6\sum_i(O_6^6)_i \label{HHo} \end{equation} The hcp lattice is not of a Bravais type. However, if we neglect the optical mode (which is possible at $T\ll T_N=133$ K) one can put (see, e.g., Ref. \cite{Coqblin}) \begin{eqnarray} J_{{\bf q}} &=&2J\left[ \cos q_x+2\cos (q_x/2)\cos (\sqrt{3}q_y/2)\right] \nonumber \\ &&\ \ \ \ +2J^{\prime }\cos \frac{q_z}2\left| \exp (iq_y/\sqrt{3})+2\cos \frac{q_x}2\exp (-iq_y/2\sqrt{3})\right| \end{eqnarray} The parameters of the Hamiltonian were taken from Ref. \cite{Jensen}: $% J=0.65 $K$,$ $J^{\prime }=0.6J,\;B_2^0=0.35$K,\ $B_4^0=0,\;B_6^0=-1.1\cdot 10^{-5}$K, $B_6^6=1.07\cdot 10^{-4}$K (note that our value of $B_2^0$ includes also renormalization due to dipolar anisotropy). For simplicity, we restrict ourselves to a collinear magnetic structure (this is justified by that the spiral angle in the rare earths is small, especially at low temperatures). The calculations with the Hamiltonian (\ref{HHo})\ are completely analogous to those in the previous Section. Calculated dependence of the cone angle is compared with the result of the mean-field approximation and experimental data in Fig.3. One can see that our results improve somewhat those of the mean-field theory where the temperature dependence of the anisotropy constants is given by (\ref{Zener}). To conclude, we have formulated a consistent spin-wave approach to description of thermodynamic properties of anisotropic magnets at low temperatures. The renormalizations of the anisotropy constants and spin-wave spectrum for an arbitrary cone angle are calculated. This gives a possibility to describe the orientational phase transition between the cone and plane phases. We are grateful to J.Jensen for comments concerning the experimental situation in holmium. {\sc Figure captions} Fig.1. The theoretical temperature dependences of the cone angle $\theta (T)$ for $S=7/2$ and different values of second-order anisotropy: $D/J=0.004;$ $% 0.005;$ $0.006$ from upper to lower curve. The value of $D^{\prime }/J$ is $% 3.7\cdot 10^{-4}.$ Fig.2. The temperature dependences of the anisotropy constants $D(T),$ $% D^{\prime }(T)$ corresponding to Fig.1. Fig.3. Calculated dependences of the cone angle in the mean-field approximation (short-dashed line) and renormalized spin-wave theory (RSWT, long-dashed line) as compared with experimental points for holmium (Refs. \cite{Coqblin,HoEx1}).
\section{Introduction} Semiconductor superlattices continue to attract substantial interest both among fundamental and applied researchers. One motivating factor is the possibility of tailoring the miniband structure\cite{ESA70,SHI75,GRA95a} for device purposes. Furthermore, a large variety of other physical phenomena such as the formation of Wannier-Stark ladders \cite{MEN90}, negative differential conductance \cite{SIB90}, and Bloch oscillations\cite{LEO96} can be observed in superlattices. The presence of minibands has been probed directly by investigating the transmission of ballistic electrons through short semiconductor superlattices \cite{KUA92,RAU97}. In recent experiments the quenching of the miniband structure by an applied electric field was also demonstrated \cite{RAU97a}. Comparison of further experiments with theoretical calculations indicated a strong influence of scattering on the transmission, and it was argued that interface roughness might cause significant deviations from pure ballistic transmission through the sample\cite{RAU98,WAC99a}. A good understanding of the transmission characteristics through short superlattices is important as these structures are used as energy filters. For example, in quantum cascade lasers, superlattice filters are used to selectively populate the upper energy level of the active region \cite{CAP97}. The most straightforward way to calculate the transmission through a superlattice is the transfer matrix method \cite{TSU73}. Alternatively, the Schr{\"o}dinger equation of the superlattice can be solved directly. These methods typically assume homogeneity in the direction perpendicular [the $(x,y)$-plane] to the superlattice. The momenta in the $(x,y)$-plane are then good quantum numbers and decouple from the superlattice direction, reducing the problem to a one-dimensional calculation; see, {\it e.g.}, Ref.~\cite{GLY91}. The one-dimensional calculation can handle fluctuations \cite{DIE94} in the well or barrier thickness. However, real samples also exhibit a lack of periodicity in the $(x,y)$-plane due to the presence of impurities and interface roughness. This can change the transport properties essentially, as the states with different parallel momenta couple to each other. This $(x,y)$-plane inhomogeneity can be tackled by solving the Schr{\"o}dinger equation on a mesh for the full three dimensional structure \cite{TIN95}. Alternatively, the method of Green-functions, based on Ref.~\cite{CAR71}, may be used (see Ref.~\cite{DAT95} for an easily accessible presentation of the method). Recently such an approach has been presented for a full calculation of the current through a resonant tunneling diode where both interface roughness and phonon scattering have been taken into account \cite{LAK97}. However, these simulations use a fine grid and are hence unsuitable for longer structures such as superlattices consisting of many wells, since the number of grid points increases dramatically. In this paper we propose a new method for such calculations which significantly reduces the computational complexity. We treat the inhomogeneity in the $(x,y)$-plane by averaging over disorder configurations. The number of grid points in the $z$-direction is reduced significantly by restricting to the basis set to the Wannier functions localized in the wells. This Wannier approximation is shown to be valid near the resonance condition if the energy gap between the minibands is large compared to the bias and the widths of the minibands themselves. We compare our results to calculations on a finite grid and find good agreement. Our method has the advantage that it corresponds to infinitely large cross-sections and hence, unlike the finite-grid calculations, does not show configuration dependent fluctuations. The paper is organized as follows. We present the general model within which our calculations are performed in section \ref{SecGenformal}. In section \ref{SecApplication}, we describe the approximations which allow us to perform practical calculations of extended superlattice structures. Our results are presented in section \ref{Secresults} and we conclude with a summary. Appendix \ref{SecImpurityav} shows the equivalence of the Landauer-B\"uttiker transmission formalism with the approach by nonequilibrium Green functions for the case of impurity averaging. In appendix \ref{SecAccuracy} we justify the approximations used in section \ref{SecApplication}. As many different symbols appear in this paper, for easy reference we display the frequently used ones in Table 1. \section{General formalism} \label{SecGenformal} In this paper, we study transport through a superlattice contacted to external voltage sources via leads. We model the superlattice as an active central region coupled to noninteracting lead regions. This is the general approach described in Refs.~\onlinecite{CAR71,DAT95,MEI92,JAU94} and \onlinecite{HAU96}. In this section we briefly review this approach, introduce our notation, and discuss the issue of impurity averaging. We divide the sample into a central region C and lead regions, indexed by $\ell$. The Hamiltonian is \begin{equation} \hat H = \hat H_C + \sum_{\ell} \hat H_\ell + \hat H_{{\ell}C} + \hat H_{{\ell}C}^\dagger \end{equation} Here $\hat H_C$ and $\hat H_\ell$ are the terms for the central structure and leads, respectively, and $\hat H_{{\ell}C}$ is the coupling term from the center to the lead $\ell$. In this paper, we ignore electron--electron interactions beyond Hartree so all the above terms are single-particle-like. The central structure has states with the wave function $\phi_{C,j}(\vec{r})$, where $j$ is the eigenstate index. We assume that each lead $\ell$ is disorder-free so that the eigenstates can be separated into transverse and longitudinal parts, $\phi_{{\ell}\alpha q}(\vec{r})=\chi_{{\ell}\alpha}({\bf r}) \varphi^{\ell}_{q}(z)$, where $z$ is the spatial coordinate in the direction towards the central structure and ${\bf r}$ is a two-dimensional vector perpendicular to $z$. The index $\alpha$ numbers the modes within a given lead. The index $q$ denotes the behavior far away from the central region where $\varphi^{\ell}_q(z) \sim e^{iqz}$ is assumed. \subsection{Green functions and current through structure} The current through a structure can be determined by the Green function of the structure {\sl in the presence of coupling to the leads}, given by a matrix {\bf G} with matrix elements \begin{mathletters} \begin{eqnarray} G^{<}_{ij}(t,t')&=& i\langle \hat{c}_j^\dagger(t') \hat{c}_i(t)\rangle\\ G^{\rm ret\atop adv}_{ij}(t,t') &=& \mp i \langle\{\hat{c}_i(t), \hat c_j^\dagger(t')\}\rangle\ \theta(\pm (t-t')\;) . \end{eqnarray} \end{mathletters} Here $\hat{c}_i^\dagger\;(\hat{c}_i)$ are fermion creation (annihilation) operators of states $\phi_{C,i}$ in the central region, and $\{\cdots,\cdots\}$ denote anticommutators. In the following we consider time-independent problems, so that ${\bf G}$ only depends on $t-t'$, and we work in energy-space by Fourier transforming ${\bf G}$ with respect to $t-t'$. The net current from mode $\alpha$ in lead $\ell$ into the structure is given by\cite{MEI92,HAU96} \begin{equation} \begin{split} J_{\ell\alpha} =2 \frac{ie}{\hbar}\int\frac{dE}{2\pi} {\rm Tr}\big\{&\boldsymbol{\Gamma}_{\ell\alpha}(E) \left[{\bf G}^{<}(E)\right.\\ &\left. +f_{\ell\alpha}(E)\left({\bf G}^{\rm ret}(E)-{\bf G}^{\rm adv}(E)\right) \right]\big\} \, .\label{EqCurrentalpha} \end{split}\end{equation} Here $f_{\ell\alpha}(E)$ gives the occupation of a state with energy $E$ in lead $\ell$ for the mode $\alpha$, $e<0$ is the charge of the electron, and $\boldsymbol{\Gamma}_{l,\alpha}$ is a parameter describing the coupling between the states in the central region and the leads [see Eq.\ (\ref{Eqdefine-Gamma}) below]. The factor of 2 is for spin. To describe transmission through the superlattice, we need to obtain expressions for the right hand side of Eq.\ (\ref{EqCurrentalpha}). We do so as follows. We first define $\hat H_{C,0}$ and $\hat H_{C}'$, as the ordered, solvable part and the disordered part of the central region Hamiltonian, respectively, and $\hat H_C = \hat H_{C,0} + \hat H_{C}'$. The retarded Green function for the structure is determined by the equation (see Ref.~\cite{HAU96}, chapter 12) \begin{equation} \left( E-{\bf H}_{C,0}-\boldsymbol{\Sigma}_{C}^{\rm ret} - \sum_{\ell\alpha}\boldsymbol{\Sigma}^{\rm ret}_{\ell\alpha}(E)\right) {\bf G}^{\rm ret}(E)=\boldsymbol{1}\,. \label{EqGret} \end{equation} The term $\boldsymbol{\Sigma}^{\rm ret}_C$ is the irreducible self-energy due to $\hat H_{C}'$. In cases where $\hat H_C'$ contains interparticle interactions, $\boldsymbol{\Sigma}^{\rm ret}_C$ is often very difficult to calculate; however, for static disorder, simply $\boldsymbol{\Sigma}^{\rm ret}_C = {\bf H}_C'$. The term $\boldsymbol{\Sigma}^{\rm ret}_{\ell\alpha}(E)$ gives the self-energy contributions due to the coupling of the central region to lead $\ell$ and mode $\alpha$, \begin{eqnarray} \Sigma^{\rm ret}_{\ell\alpha,ij}(E)&=& \sum_q \langle \phi_{C,i}|\hat{H}_{\ell C}^\dagger|\phi_{\ell\alpha q}\rangle \langle \phi_{\ell\alpha q}|\hat{H}_{\ell C}|\phi_{C,j}\rangle\; g^{\rm ret}_{\ell\alpha q}(E)\nonumber \\ &=&\frac{L_{\ell}}{2\pi}\int_0^{\infty} dE_q\ \frac{2}{\hbar v_q} \langle \phi_{C,i}|\hat{H}_{\ell C}^\dagger|\phi_{\ell\alpha q}\rangle \nonumber \\ &&\phantom{\frac{L_{\ell}}{2\pi}\int_0^{\infty} dE_q} \langle \phi_{\ell\alpha q}|\hat{H}_{\ell C}|\phi_{C,j}\rangle\; g^{\rm ret}_{\ell\alpha q}(E) \label{Eqsigmaalpha} \end{eqnarray} where we have taken the continuum limit $\sum_q\to L_{\ell}/2\pi\int_{-\infty}^{\infty} dq$ ($L_\ell$ is the length of lead $\ell$). The factor $2$ results from the two possible values $\pm q$ for a given energy $E_q$, $g^{\rm ret}_{\ell\alpha q}(E)=1/(E-E_q-E_{\ell\alpha}+i0^+)$ is the free-particle Green function of the lead in absence of the central region, $E_q=\hbar^2q^2/2m^*$, $v_q=\hbar q/m^*$ and $E_{{\ell},\alpha}$ is the lateral energy of the mode $\alpha$. Here $m^*$ is the effective electron mass. Note that $G_{ij}^{\rm adv}(E) = [G_{ji}^{\rm ret}(E)]^*$, since we have a time-independent system. The coupling parameter $\boldsymbol{\Gamma}_{\ell\alpha}$ is defined by \begin{eqnarray} {\Gamma}_{\ell\alpha,ij}(E) &=&i\left[{\Sigma}^{\rm ret}_{\ell\alpha,ij}(E)- {\Sigma}_{\ell\alpha,ij}^{\rm adv}(E)\right]\nonumber \\ &\equiv&\frac{2 L_{\ell}\langle \phi_{C,i}|\hat{H}^\dagger_{\ell C}| \phi_{\ell\alpha\; q(E-E_{\ell\alpha})}\rangle} {\hbar v_{q(E-E_{\ell\alpha})}}\nonumber\\ &&\times \langle \phi_{\ell\alpha\; q(E-E_{\ell\alpha})}| \hat{H}_{\ell C}|\phi_{C,j}\rangle \; \Theta(E-E_{{\ell}\alpha}) \label{Eqdefine-Gamma} \end{eqnarray} where $q({\cal E})=\sqrt{2m^*{\cal E}}/\hbar$. $\Theta(x)$ is the Heavyside function with $\Theta(x)=1$ for $x\ge 0$ and $\Theta(x)=0$ for $x< 0$. Note that $\Gamma_{\ell\alpha,ij}(E)=0$ for $E<E_{\ell\alpha}$ since there are no propagating states into which the central-region states can tunnel. ${\bf G}^{<}(E)$ can be obtained by the Keldysh relation\cite{KEL65} \begin{equation} {\bf G}^{<}(E)={\bf G}^{\rm ret}(E) \boldsymbol{\Sigma}^<(E){\bf G}^{\rm adv}(E)\label{EqKeldysh} \end{equation} where \begin{equation} \boldsymbol{\Sigma}^<(E)=\boldsymbol{\Sigma}^{<}_{C}(E)+\sum_{{\ell}\alpha} \boldsymbol{\Sigma}_{\ell\alpha}^{<}(E). \label{EqSigmaless} \end{equation} Here, $\boldsymbol\Sigma_C^<$ is the self-energy resulting from scattering inside the structure. For a fixed disorder potential, this term is identically zero. The term $\boldsymbol\Sigma_{\ell\alpha}^<(E)$ is the self-energy due to the presence of the coupling to the leads, \begin{eqnarray} \Sigma_{\ell\alpha,ij}^<(E)&=& \sum_q \langle \phi_{C,i}|\hat{H}_{\ell C}^\dagger|\phi_{\ell\alpha q}\rangle \langle \phi_{\ell\alpha q}|\hat{H}_{{\ell}C}|\phi_{C,j}\rangle g^{<}_{\ell\alpha q}(E)\nonumber \\ &=&i\,\Gamma_{\ell\alpha,ij}(E)\, f_{\ell\alpha}(E)\label{EqSigmalessL} \end{eqnarray} where we have used $g^{<}_{\ell\alpha q}(E)= -2if_{\ell\alpha}(E)\,{\rm Im}\left\{g^{\rm ret}_{\ell\alpha q}(E)\right\}$. The occupation function $f_{\ell\alpha}(E)$ in lead $\ell$ and is given by the externally imposed conditions. Usually, the leads are assumed to be in thermal equilibrium and hence a Fermi distribution with chemical potential $\mu_{\ell}$, independent of $\alpha$, is used. In contrast, the different modes can be populated individually by injection, as discussed later, so that we want to keep the full function $f_{\ell\alpha}(E)$. \subsection{Relation to the Landauer-B{\"u}ttiker approach} The Landauer-B\"uttiker approach has been used extensively to study transmission through mesoscopic structures, and consequently many people are familiar with the formalism. As the Keldysh formulation is not as widely known, in this subsection we demonstrate the equivalence of the two approaches for transport through a system with static disorder. The retarded and advanced Green functions can be expressed in terms of $\Gamma$ via \begin{equation} {\bf G}^{\rm ret}(E)-{\bf G}^{\rm adv}(E)=-i{\bf G}^{\rm ret}(E) \boldsymbol{\Gamma}(E){\bf G}^{\rm adv}(E)\label{Eqspectral} \end{equation} where the total scattering rate $\boldsymbol{\Gamma}$ has two contributions \begin{equation} \boldsymbol{\Gamma}(E)= i\left[\boldsymbol{\Sigma}^{{\rm ret}}_{C}(E) -\boldsymbol{\Sigma}^{{\rm adv}}_{C}(E)\right]+ \sum_{\ell\alpha}\boldsymbol{\Gamma}_{\ell\alpha}(E) \end{equation} resulting from scattering inside the structure and transitions into the leads. If the scattering within the structure itself is purely elastic and is treated in a particular fixed configuration as a potential in Eq.~(\ref{EqGret}), then $\boldsymbol{\Sigma}^{\rm ret}_C = {\bf H}_C'$ and $\boldsymbol{\Sigma}^{<}_C(E)=0$; hence we may insert Eqs.~(\ref{EqKeldysh}) and (\ref{Eqspectral}) into Eq.~(\ref{EqCurrentalpha}) and find the Landauer-B{\"u}ttiker expression \cite{MEI92,BUE90} \begin{equation} J_{\ell\alpha}=2 \frac{e}{\hbar}\int\frac{dE}{2\pi} \sum_{\ell'\beta} T_{\ell\alpha\leftarrow \ell'\beta}(E) \left[f_{\ell\alpha}(E)-f_{\ell'\beta}(E)\right] \label{EqBuett} \end{equation} (factor of 2 for spin) with the transmission matrix \begin{equation} T_{\ell\alpha\leftarrow \ell'\beta}(E)= {\rm Tr}\left\{\boldsymbol{\Gamma}_{\ell\alpha}(E) {\bf G}^{\rm ret}(E) \boldsymbol{\Gamma}_{\ell'\beta}(E){\bf G}^{\rm adv}(E)\right\}. \label{Eqtransmat} \end{equation} (There are several alternate ways to derive this result; {\it e.g.}, Ref.~\cite{DAT95} uses spatial discretization.) Note that Eq.~(\ref{Eqtransmat}) does {\sl not} hold if the scattering process is inelastic or the elastic scattering by static disorder is described by a self-energy obtained by configuration averaging. In both cases $\boldsymbol{\Sigma}^{<}_C(E)\neq 0$ in contrast to the assumption leading to Eq.~(\ref{EqBuett}). \subsection{Impurity Averaging} \label{SubsecImpav} Eq.\ (\ref{Eqtransmat}) is exact for a given configuration of impurities and roughness, {\it i.e.}, for a specific ${\bf H}_C'$. However, obtaining the transmission by simulating individual configurations is not computationally efficient, and hence it is advantageous to average over impurity configurations. In particular, such a procedure reestablishes symmetries which are broken by specific impurity configurations, thus simplifying the calculation significantly. After impurity averaging, we obtain \begin{equation} \overline{\bf G}^{\rm ret}(E) = [E + i0^+ - {\bf H}_{C,0} - {\boldsymbol{\overline\Sigma}}_C^{\rm ret} (E) - \sum_{\ell} {\boldsymbol{\Sigma}}_{\ell}^{\rm ret}(E)]^{-1}, \label{EqGretav} \end{equation} where the overlines indicate averages over disorder configurations in the central region. Note that the disorder averaging introduces non-zero self-energies $\overline{\boldsymbol{\Sigma}}_C^{\rm ret}(E)$ and $\overline{\boldsymbol{\Sigma}}_C^{<}(E)$. As $\overline{\boldsymbol{\Sigma}}_C^{<}(E)\neq 0$ one cannot simply use Eq.\ (\ref{Eqtransmat}) with the $G$'s replaced by $\overline{G}$. In order to describe configuration-averaged elastic scattering within the transmission formalism, the averaging must be performed for the total transmission matrix in Eq.~(\ref{Eqtransmat}), and not just on the individual $G^{\rm ret}$ and $G^{\rm adv}$. This procedure is similar to the calculation of bulk conductivities using the Kubo formula, where it is crucial to include vertex corrections which fulfill the Ward-identity (see, e.g., Ref.~\cite{DON74}). We perform such a calculation in appendix \ref{SecImpurityav} for the superlattice structure discussed in section \ref{SecApplication}. We use the self-consistent Born-approximation for the scattering and therefore the appropriate vertex function is the so-called ladder approximation. The application of the more general Keldysh approach to calculate the current in the configuration averaged case is more straightforward, in that one {\em can} replace the $G$ by $\overline{G}$ in Eq.~(\ref{EqCurrentalpha}). Therefore, in order to evaluate the current, we need $\overline{\bf G}^<(E)$ and $\overline{\bf G}^{\rm ret}(E)$. The general iterative procedure for computing these is as follows.\label{Secprocedure} First the self-energies $\boldsymbol{\Sigma}_{\ell}^{\rm ret}(E)$ and $\boldsymbol{\Sigma}_{\ell}^<(E)$ due to the leads are evaluated by Eqs.~(\ref{Eqsigmaalpha}) and (\ref{EqSigmalessL}). As these terms are independent of disorder configuration, these need only be evaluated once and then stored. With $\boldsymbol{\overline{\Sigma}}_C^{{\rm ret},<}$ initially set equal to zero, $\overline{\bf G}^<$ and $\overline{\bf G}^{\rm ret}$ are calculated. These $\overline{\bf G}$'s are used to calculate the $\boldsymbol{\overline{\Sigma}}_C$'s, via an appropriate approximation scheme. The updated $\boldsymbol{\overline{\Sigma}}_C$'s are used to generate new $\overline{\bf G}$'s via Eqs.~(\ref{EqKeldysh}) and (\ref{EqGretav}), and the process is iterated until convergence is achieved. Finally, the current is evaluated with Eq.~(\ref{EqCurrentalpha}). In Appendix \ref{SecImpurityav} we show explicitly that the ladder approximation for the vertex function in the transmission formulation yields the same equations as the Keldysh approach within the self-consistent Born approximation, demonstrating the equivalence of the two methods for impurity scattering. Nevertheless the Keldysh approach seems to be conceptually easier as there is only one place within this formulation where an approximation is made; {\em i.e.}, in the self-energy. In contrast, with the transmission formalism, errors can occur if the vertex function does not fulfill the Ward identity, providing a pitfall to trap the uninitiated and unwary. \section{Application to a superlattice structure} \label{SecApplication} Let us consider the superlattice structure sketched schematically in Fig.~\ref{Figstruktur}. The superlattice consists of $N$ identical wells embedded in $N+1$ barriers. A bias $U$ is applied to the structure yielding constant potentials $U_L$ and $U_R=U_L+eU$ at the left and right contact, respectively. In order to perform calculations we now specify the basis states $\phi_{C,j}(\vec{r})$ and $\chi_{\ell\alpha}({\bf r})$ for our superlattice structure. The lead index $\ell$ takes two different values $L$ and $R$, for the left and right contact region, respectively. For superlattices with a large cross section $A$ it is natural to use a basis of plane waves $e^{i{\bf k}\cdot{\bf r}}/\sqrt{A}$ for the transverse coordinates $(x,y)$ both in the lead regions and in the superlattice itself. Then the index $\alpha$ of the states in the leads is replaced by ${\bf k}$ and we have $E_{(L/R,{\bf k})}=E_k+U_{L/R}$, where $E_k=\hbar^2k^2/2m$. \subsection{Wannier approximation for a superlattice\label{SecWA}} Let us now consider the central region; {\it i.e.,} the superlattice structure itself. In order to make a calculation tractable, we restrict ourselves to a subset of the basis functions of total Hilbert space, ignoring irrelevant high energy states. With respect to the $z$-direction inside the superlattice we use a basis of Wannier-functions $\Psi_n(z)$ ($n=1,\ldots N$) from the lowest miniband which are maximally localized in well $n$\cite{KOH59}. Such a basis has been successfully applied to superlattice transport \cite{WAC98}. This approximation, which we call the Wannier approximation (WA), neglects higher mini-bands, and its validity is discussed in the Appendix \ref{SecAccuracy}. There we demonstrate that this approximation gives good results for the transmission probability provided that the miniband width is smaller than the energy of the center of the miniband and the energy range of interest is sufficiently below the levels corresponding to the higher miniband states. The states $\phi_{C,j}$ within the superlattice are, within the WA, given by products $\Psi_n(z)e^{i{\bf k}\cdot{\bf r}}/\sqrt{A}$, which can be labeled by $(n,{\bf k})$. Within the superlattice, the Green-function is determined by Eq.~(\ref{EqGret}) which in the WA basis reads \begin{equation} \begin{split} \sum_{n'{\bf k}'} \left[(E-E^a-E_k-U_n)\delta_{{\bf k},{\bf k}'}\delta_{n,n'} -H_{n{\bf k},n'{\bf k}'}\right.\\ -T_1\delta_{{\bf k},{\bf k}'}\left(\delta_{n,n'+1}+\delta_{n,n'-1}\right) \left.-\sum_{\ell\alpha}\Sigma^{\rm ret}_{\ell\alpha;\;n{\bf k},\,n'{\bf k}'}(E) \right]\\ G^{\rm ret}_{n'{\bf k}',m{\bf k}_1}(E)=\delta_{{\bf k},{\bf k}_1}\delta_{n,m} \label{EqGreenWannk} \end{split} \end{equation} Here, $U_n$ denotes the potential in the well $n$ (see Fig.~\ref{Figstruktur}) which is due to an external bias. (The mean-field potential induced by the carriers in the structure can be added as well.) $T_1$ is the coupling between the wells and $E^a$ is the level energy of the Wannier state relative to the bottom of the well. For a given structure, we calculate $T_1$ and $E^a$ as follows. We consider first an infinite superlattice of the same composition. The eigenstates in the infinite superlattice are Bloch functions with the miniband dispersion $E^a(q)$. $E^a$ is then identified as the center of the miniband $d/(2\pi)\int dq E^a(q)$ and $T_1= d/(2\pi) \int dq E^a(q)\cos(qd)$, where $d$ is the period of the superlattice \cite{WAC98}; {\it i.e.}, $|T_1|$ is about a fourth of the miniband width. Finally, $H_{n{\bf k},n'{\bf k}'}$ is the disorder scattering matrix element. If we average over disorder configurations the translational invariance in the $(x,y)$-plane is restored, and consequently, all impurity-averaged quantities are diagonal in ${\bf k}$ parallel to the $(x,y)$-plane. Therefore we are able to use the notation $\overline{G}_{n{\bf k},\, m{\bf k}}(E) \equiv \overline{G}_{nm}({\bf k},E)$ and matrices ${\bf G}({\bf k},E)$ and $\boldsymbol{\Sigma}({\bf k},E)$ have the components $G_{nm}({\bf k},E)$ and $\Sigma_{nm}({\bf k},E)$, respectively. \subsection{Estimating the coupling and wide band limit} \label{SecCoupling} The coupling with the mode ${\bf k}$ in the left contact yields, from Eq.~(\ref{Eqsigmaalpha}), the self energy \begin{equation}\begin{split} &\Sigma^{\rm ret}_{L{\bf k};\;n{\bf k}_1,\, n'{\bf k}_2}(E)= \delta_{n,1}\delta_{n',1} \delta_{{\bf k}_1,{\bf k}}\delta_{{\bf k}_2,{\bf k}} \\ &\phantom{\Sigma}\times \frac{1}{2\pi} \int_0^{\infty} dE_q\ \frac{2L_L|V_q|^2}{\hbar v_q} \frac{1}{E-E_q-E_{\bf k}-U_L+i0^+} \label{EqSigmawannierk} \end{split}\end{equation} where $V_q=\langle \varphi_{q}^L(z)|\hat{H}_{LC}|\Psi_1(z)\rangle$ is the $z$-dependent part of the matrix element for the coupling to the leads. Here we neglect the coupling to the inner wells $(n\neq 1)$, which should be small. The right contact gives the same term except with replacements $\delta_{n,1}\rightarrow \delta_{n,N}$ and $U_L,L_L\rightarrow U_R,L_R$. If the transmission function is strongly determined by resonances, only a small energy range of $E\approx E^a+U_L+E_k$ contributes to the transmission. In this range we neglect the $q$ dependence of the coupling and extend the lower limit of the integration in Eq.~(\ref{EqSigmawannierk}) to $-\infty$. Then we obtain for the left lead \begin{eqnarray} \Sigma^{\rm ret}_{L{\bf k};\; n{\bf k}_1,\,n',{\bf k}_2}(E)&=& \delta_{n,1}\delta_{n',1} \delta_{{\bf k}_1,{\bf k}}\delta_{{\bf k}_2,{\bf k}} \frac{-i}{2} \Gamma_L \end{eqnarray} with \begin{equation} \Gamma_L=\frac{2L_L|V_{q(E^a)}|^2}{\hbar v_{q(E^a)}}. \end{equation} This approximation is often referred to as wide band limit. Note that this limit becomes problematic if the voltage drop across the first barrier becomes large, as this changes the relevant values of $E$ and it cannot be regarded as constant [see also Appendix \ref{SecAccuracy}]. Now we want to estimate the value of $|V_q|^2$. For $E_q\approx E^a$ the wavefunction $\varphi_{q}^L(z)$ in the left lead behaves like the Wannier function $\Psi_0(z)=\Psi_1(z+d)$ which is localized in a fictitious additional well on the left side of the structure. Now $\varphi_{q}^L(z)$ is normalized to $L_L$ while the spatial extension of the Wannier function is given by $w_{\rm eff}$, which should be slightly larger than the well width, as the function penetrates into the barriers. Therefore we may set $\varphi_{q}^L(z)\sim \sqrt{w_{\rm eff}/L_L}\Psi_0(z)$. Then we can estimate the matrix element \begin{equation} \langle \varphi_q^L|H|\Psi_1\rangle\approx \sqrt{\frac{w_{\rm eff}}{L_L}}\langle \Psi_0|H|\Psi_1\rangle= \sqrt{\frac{w_{\rm eff}}{L_L}}T_1, \end{equation} yielding \begin{equation} \Gamma_L\approx \frac{2w_{\rm eff}T_1^2}{\hbar v_{q(E^a)}}. \end{equation} For the right contact, $\Gamma_R$ is given by the same value. \subsection{Interface Roughness} For ideal structures the potential $H_{n{\bf k},n'{\bf k}'}$ in Eq.~(\ref{EqGreenWannk}) is zero due to the translational invariance within the $(x,y)$-plane. However, interface fluctuations leading to well width fluctuations $\xi_n({\bf r})$ in real samples break this translational invariance. If interwell scattering and well-width correlations between different wells can be neglected, the averaged square of the scattering matrix element is given by \cite{GOO85,RID98} \begin{equation} \langle |H_{n{\bf k}+{\bf p},n'{\bf k}}|^2\rangle= \frac{K^2}{A}S({\bf p}) \delta_{n,n'} \end{equation} where $K$ is equal to the change of energy $dE^a/dw$ per well width fluctuation \cite{RefTES} and $S({\bf p})$ is the Fourier transformation of the well width correlation function $\langle \xi_n({\bf r})\,\xi_n({\bf r'})\rangle=f({\bf r}-{\bf r'})$ which is assumed to be independent of the well index. The theory can be extended to accommodate interwell scattering and well-width correlations between different wells (which may result from a repetition of the microscopic interface structure over several superlattice periods) by the inclusion of the appropriate correlation functions $\langle H_{n_1{\bf k}-{\bf p},n_1'{\bf k}} H_{n_2{\bf k}+{\bf p},n_2'{\bf k}}\rangle $. We use an isotropic exponential distribution $f(r)=\eta^2 \exp(-r/\lambda)$ yielding \begin{equation} S({\bf p})= \eta^2 \lambda^2 \frac{2\pi}{\left(1+(p\lambda)^2\right)^{3/2}}\,, \end{equation} where $\eta$ denotes the standard deviation and $\lambda$ the in-plane correlation length of the well-width fluctuation. It is straightforward to implement more sophisticated distribution functions, which might be obtained from Monte-Carlo simulations of the growth conditions (see, {\it e.g.,} Ref.~\cite{GRO95}) or X-ray characterizations of the superlattice structure (see, {\it e.g.,} Ref.~\cite{GRE98a}). Within the self-consistent Born approximation we obtain the self energy $\overline{\boldsymbol{\Sigma}}_{C}$ \begin{equation} \overline{\Sigma}^{</{\rm ret}}_{C;\; nn}({\bf k},E)=\sum_{\bf k'} \langle |H_{{\bf k}',{\bf k}}|^2\rangle \overline{G}^{</{\rm ret}}_{nn}({\bf k}',E) \end{equation} which provides the functional needed in the procedure scetched in section \ref{Secprocedure}. \section{Results}\label{Secresults} Let us consider the transmission of ballistic electrons through the superlattices considered in recent experiments by Rauch {\it et al.} \cite{RAU97}. The structure consists of $N$ wells of $6.5$ nm GaAs and $N+1$ barriers of $2.5$ nm Al$_{0.3}$Ga$_{0.7}$As. We obtain the band parameters $E^a=54.5$ meV, $T_1=-5.84$ meV, $K=13.25$ meV/nm and use $w_{\rm eff}=10.7$ nm, where we obtained the best agreement with ``exact'' calculations; see appendix \ref{SecAccuracy}. This value is somewhat larger than the well width in good agreement with the discussion in Section~\ref{SecCoupling}. We assume thickness fluctuations of half a monolayer $\eta=0.14$ nm around the nominal value and a correlation length $\lambda=5$ nm, unless otherwise stated. Motivated by the relatively sharp electron distribution injected into the structure, we assume that the electrons occupy the mode ${\bf k}=0$ of the left contact at an energy $E=E_{in}$; {\it i.e.}, we have $f_{L{\bf k}}(E)=\delta_{{\bf k},0} \delta(E-E_{in})$ and $f_{R{\bf k}}(E)=0$. The total current through the right contact is then given by \begin{equation} J_R=\sum_{\bf k} J_{R\,{\bf k}}=-\frac{e}{\pi \hbar} \sum_{\bf k}\int dE\ {\rm Tr}\left\{-i\boldsymbol{\Gamma}_{R{\bf k}}(E) \, {\bf G}^{<}(E)\right\}. \label{EqJGless} \end{equation} This can be expressed via Eq.~(\ref{EqBuett}) by \begin{equation} J_R=-\frac{e}{\pi \hbar} \sum_{\bf k} T_{(R,{\bf k})\leftarrow (L,{\bf 0})}(E_{\rm in}). \label{EqJtrans} \end{equation} For illustrative purpose we calculate the effective transmission $T(E_{\rm in})=-J_R\pi \hbar/e$ in the following. Regarding the applied bias we assume a homogeneous electric field $F$ inside the superlattice and set $U_L=0$, $U_n=-(n-1/2)eFd-eFb/2$ and $U_R=-NeFd-eFb=eU$ where $b$ is the barrier width. In the experiments considered, there is no charge accumulation inside the structure as there is on average less than one electron inside the structure at a given time. If necessary such effects can be easily taken into account by solving the Poisson equation for the electron density given by \begin{equation} N_n=\frac{-i}{2\pi A}\sum_{\bf k}\int dE\ \overline{G}^<_{nn}({\bf k},E)\, . \end{equation} In Fig.~\ref{FigTrough} we show the effective transmission with and without scattering. In both cases we find a series of peaks, equal to the number of quantum wells, which reflect the eigenstates of the superlattice structure. For $U= 0$ the peak maxima reach the value 1 for the ideal superlattice. The broadening of these peaks results from the coupling to the leads and is of the order $(\Gamma_R+\Gamma_L)/N$. In contrast the maxima are lower and the widths are wider for the calculation including scattering. These effects becomes more pronounced with increasing superlattice length as the broadening due to scattering dominates with respect to the lead induced broadening. An important quantity is the integrated transmission for a given potential drop $U$ \begin{equation} T_{\rm int}(U)=\int dE_{\rm in}\ T(E_{\rm in}; U), \label{EqTint} \end{equation} where the integration is extended over the whole energy range of the band. This quantity was measured in Ref.~\cite{RAU97,RAU98}. Results are shown in Fig.~\ref{FigTintrough}. Let us compare the result of the calculations with (full line) and without roughness (dotted line) first. Without interface roughness, the function $T_{\rm int}(U)$ is always symmetric with respect to $U$. This can be understood from the symmetry property of the transmission matrix $T_{\ell\alpha\leftarrow \ell'\beta}(E)= T_{\ell'\beta \leftarrow \ell\alpha }(E)$ (see, e.g., Ref.~\cite{DAT95}). For an ideal structure, ${\bf k}$ is conserved within the structure and we find according to Eq.~(\ref{EqJtrans}): \begin{eqnarray} T_{\rm int}(U)&=& \int dE_{\rm in}\sum_{\bf k}\; T_{(R,{\bf k})\leftarrow (L,{\bf 0})}(E_{\rm in}; U) \nonumber\\ &=&\int dE_{\rm in}\ T_{(R,{\bf 0})\leftarrow (L,{\bf 0})}(E_{\rm in};U) \nonumber\\ &=&\int dE_{\rm in}\ T_{(L,{\bf 0})\leftarrow (R,{\bf 0})}(E_{\rm in};U) \end{eqnarray} Now $T_{(L,{\bf 0})\leftarrow (R,{\bf 0})}(E_{in};U)= T_{(R,{\bf 0})\leftarrow (L,{\bf 0})}(E_{in}+eU;-U)$ due to the symmetry of the structure and so we find $T_{\rm int}(U)=T_{\rm int}(-U)$. This argument does not hold for a superlattice with interface roughness as the scattering is able to transfer electrons from state ${\bf k}={\bf 0}$, where they are injected to a finite value of ${\bf k}$. In this case kinetic energy $E_k$ is transferred to the $(x,y)$-direction and the electrons leave the superlattice with a lower $z$-component of the energy $E_q$. This opens up new channels for new processes if $U>0$; see also the discussion in Ref.~\cite{WAC99a}. Therefore the function $T_{int}(U)$ is asymmetric with respect to the bias $U$ as can be clearly seen in Figure~\ref{FigTintrough} (full line). These findings are in excellent agreement with recent measurements\cite{RAU98}. In Fig.~\ref{FigTintrough} we have also shown the transmission due to electrons traversing the superlattice without scattering (dashed line). This curve is obtained by neglecting the term $\boldsymbol{\Sigma}^{<}_{C}(E)$ in Eq.~(\ref{EqSigmaless}). It can be clearly seen that this curve is symmetric with respect to the bias and its magnitude is decreasing with increasing sample length. An alternative way of calculating the transmission has been performed in Refs.~\cite{WAC99a}. There the Green functions were calculated for a fixed interface potential following Ref.~\cite{TIN95}. For practical reasons the size of the samples is relatively small. The diamonds and crosses refer to two different random interface potentials as shown in Fig.~\ref{FigInterface} which both have approximately the same statistical features. The data obtained for the transmission are not smooth for $U>0$ and exhibit differences between each other. This indicates that significantly larger areas than $10\times 10$ or $15\times 15$ grid points must be used for reliable calculations utilizing this method, which is not practicable. In contrast the method using impurity averaging presented here gives a smooth behavior which, in effect, averages the scattered data points obtained from the previous calculations. In Fig.~\ref{Fig10ges}(a) we have shown the integrated transmission for different values of the correlation length for the roughness distributions. In the range considered we find that the asymmetry increases with the correlation length of the interface roughness. This indicates that larger islands lead to an enhancement of scattering even if the average coverage is identical. The reason is that scattering events with low momentum transfer is enhanced. Such scattering events dominate the transport characteristics of the superlattice due to the energy scales involved in the system. Fig.~\ref{Fig10ges}(b) shows the increase of the asymmetry with the fluctuation height. The strong dependence allows for an estimation of the interface quality by analyzing the experimental transmission data. \section{Summary and conclusions} We have presented a formalism to calculate the transmission of electrons through a finite superlattice in the presence of scattering processes. Due to impurity averaging the results are applicable to samples with large cross-sections. We have also shown that reasonable results can be obtained by restricting the calculation to a basis of Wannier-functions. Within this Wannier approximation all couplings are well defined and can be easily calculated from the superlattice parameters, with the only slight ambiguity being the effective normalization width $w_{\rm eff}$, which is typically a few nanometers larger than the well width. Although we have only presented results for interface roughness scattering, the formalism is easily applicable to other elastic scattering processes, such as impurity scattering, as well. With regard to inelastic phonon scattering, the formalism holds as well if Langreth rules \cite{LAN76a,HAU96} are taken into account, which provide the more complicated functionals for the retarded and lesser self-energies, see also Ref.~\cite{WAC99b}. Nevertheless, one encounters the problem that the Green functions at different energies couple to each other. Therefore the set of equations which has to be solved self-consistently becomes significantly larger. The inclusion of electron-electron interaction within the mean-field model is straightforward. Our results show that interface roughness gives an enhancement of the electron transmission for positive biases applied to the superlattice. The shape of the integrated transmissions depends strongly on the distribution of the well width fluctuations and allows us to study interface roughness in semiconductor heterostructures. This provides a a complementary approach to the usual method of characterization by luminescence spectra. \acknowledgements Helpful discussions with S. Bose, W. Boxleitner, F. Elsholz, E. Gornik, A.-P. Jauho, G. Kie{\ss}lich, and C. Rauch are acknowledged. This work has been supported by the Deutsche Forschungsgemeinschaft in the framework of SFB 296.
\section*{Acknowledgements} We would like to thank B. Ivanyi for lively and helpful discussions. We gratefully acknowledge financial support by DFG, BMBF and GSI.
\section{Introduction} It was long ago suggested that black holes should be treated as elementary particles \cite{tHooft}, because both are parametrized only by their mass, angular momentum (or spin) and gauge charges. String theory has made a significant contribution towards putting this assertion on a firm basis. A common feature of black holes and elementary string states is that the degeneracy of states with a given mass increases with it. Unfortunately, for elementary string states the logarithm of the degeneracy of states increases linearly with mass, whereas the Bekenstein-Hawking entropy of the black hole increases as the square of the mass. There are some cases in which the discrepancy between the two entropies can be removed appealing to the large renormalization of the mass of a black hole\cite{Sus1}. There are, however, some particular states in string theory, called BPS states, which do not receive any mass renormalization\cite{Witten}. Whereas the logarithm of the degeneracy of BPS states grows linearly with the mass, the area of the event horizon of a BPS or extremal black hole actually vanishes. This motivates the argument\cite{sen} that the entropy of an extremal black hole is not exactly equal to the area of the event horizon, but to the area of a surface close to the event horizon called ``stretched horizon''. By carefully defining the location of the stretched horizon in a consistent way, the Bekenstein-Hawking formula for the black hole entropy can receive corrections in such a way that it correctly reproduces the logarithm of the density of elementary string states. The stretched horizon of a black hole is defined as the surface where the space-time curvature (in the string metric) becomes large. It is also the surface where the local Unruh temperature for a stationary observer (constant $r$) is the Hagedorn temperature of the string theory. (The local Unruh temperature becomes infinite at the event horizon). Specifically, we define the stretched horizon as the surface where the scalar curvature $C=($Riemann$)^2$ is equal to one in Planck's units. \section{Stretched horizon for the classical solutions} Throughout, we shall be thinking of black holes as the classical description of a quantum object. The nature of this object is well approximated by (classical) general relativity far away from the horizon. At any rate, we should not expect the metric which solves Einstein's equations to make any sense at distances to the origin smaller than the de Broglie wavelength $\lambda_B= M^{-1}$ for a black hole of mass $M$. Since we are stringy, it is perhaps more appropriate to use Veneziano's generalized uncertainty principle\cite{Veneziano} $\Delta x \geq \frac{\ell_s^2}{2\hbar} \Delta p + \frac{\hbar}{ \Delta p}$, where $ \ell_s$ is the string scale, $g$ is the string coupling constant and $\ell_p= g \ell_s$ is Planck's length. So, in Planck units, the metric certainly is not expected to make any sense at radii smaller than \begin{equation}{\label {veneciano} } \lambda_V = {\frac1 2} M + {\frac1 M} \end{equation} Where is the stretched horizon? If the place where the curvature becomes big (unity in Planck units) is inside the event horizon, then clearly there is no need to stretch it at all. Similarly, if it falls at a radius smaller than $\lambda_V$, there is no point in talking about it. We are interested in finding under what circumstances the stretched horizon is a meaningful and useful concept for non-supersymmetric four-dimensional black holes. In other words, we must find when the stretched horizon is bigger than the event horizon and also bigger than $\lambda_V$ (or $\lambda_B$). Consider the general solution to the Einstein-Maxwell equations \begin{eqnarray}\label{metric} ds^2 = -\left (\frac { {\Delta} - a^2{\sin^2{\theta}}} {\Sigma}\right) dt^2 - \frac{2a \sin^2{\theta}\left( r^2 + a^2 - {\Delta} \right)}{\Sigma} dtd{\phi}\,\,\,\,\,\,\, \nonumber \\ + \left[ \frac {{\left( r^2 + a^2 \right)}^2 - {\Delta}a^2 \sin^2{\theta}}{\Sigma}\right] \sin^2 {\theta} d{\phi}^2 + \frac {\Sigma}{\Delta} dr^2 + {\Sigma}{d{\theta}}^2, \end{eqnarray} \begin{equation} A_{\mu} = - \frac {er}{\Sigma} \left[ {(dt)}_{\mu} - a\sin^2 {\theta}({d{\phi}}_{\mu}) \right], \end{equation} where \begin{equation} {\Sigma} = r^2 + a^2cos^2{\theta}, \,\,\,\,\,\, {\Delta} = r^2 + a^2 + e^2 - 2Mr, \end{equation} Using Mathematica, we have computed the square of the Riemann tensor for this solution, $C$, and evaluated where it becomes one. For the Schwarszchild black hole ($e=a=0 $) the scalar curvature is simply \begin{equation}{\label {sch} } C = \frac{48 M^2}{r^6} \end{equation} whereby the stretched horizon radius is \begin{equation}{\label {rsch} } r_s = 48^{\frac{1}{6}} M^{\frac{1}{3}} \end{equation} Thus, the stretched horizon is always inside the event horizon, except for such ridiculously small masses that the de Broglie wavelength is actually bigger than both. This continues to be the case for any charged static black hole: in Fig.~1 we plot the critical mass (in Planck units) at which the stretched horizon crosses the event horizon in terms of the charge $e$. Only for masses below the line is the stretched horizon bigger than the event horizon. The curve is essentially flat just below $M\sim 0.95$ when it begins to grow. When $e$ reaches its extremal value $e=1$, $M_c$ reaches 1.09. So both for Schwarzschild and for Reissner-Nordstr\"om black holes, the stretched horizon is useless. When $a\ne0$, i.e. for rotating black holes, the situation doesn't change much. The stretched horizon is again bigger than the event horizon only for small masses, but the critical mass below which the stretched horizon is relevant (bigger than the event horizon) is now a few times bigger and grows with $a$. Fig.~2 shows this critical mass when $e=0$ (the plot was computed in the axial direction $\theta=\pi$). In the extremal limit $a\to1$, the critical mass approaches the value $5.5$ corresponding to a stretched horizon radius of $11.0$, which is $3.7$ times bigger than Veneziano's wavelength and $63.9$ times bigger than de Broglie's wavelength. The charge, in the general Kerr-Newman case, remains rather irrelevant. \section{Conclusions} In this work we have shown two things. First, that outside the event horizon of a static black hole, the curvature is always small. Secondly, we have found numerically the critical mass above which the event horizon is bigger than the stretched horizon for a four dimensional black hole. In the extremal limit, it ranges from 1.09 when $a=0$, to $5.5$ when $e=0$. This results are in general consistent, since the temperature near the horizon of a classical black hole is always much smaller than the corresponding temperature of an extremal supersymmetric black hole. Since the critical mass is of order unity in all cases this mean that the stretched horizon is not a very useful concept for four-dimensional non-supersymmetric black holes. Physically, this indicates that only the massless modes of the string can be seen by an outside observer. \begin{figure}[t!] \centerline{\epsfig{file=acero.eps,height=2.0in,width=2.5in}} \vspace{10pt} \caption{Critical mass $Mc$ for a nonrotating charged black hole.} \label{fig1} \end{figure} \bigskip \bigskip \bigskip \bigskip \qquad \begin{figure}[t!] \centerline{\epsfig{file=ecero.eps,height=2.0in,width=2.5in}} \vspace{10pt} \caption{Critical mass $Mc$ for a rotating black hole.} \label{fig2} \end{figure} \bigskip \bigskip \bigskip \bigskip {\bf Acknowledgements}. This work is supported in part by CONACYT 25504E, DGAPA-UNAM IN103997. C.E. enjoys a scholarship from DGEP-UNAM.
\section{Introduction and Motivation} In \cite {Z1} we proposed a Lagrangian density for the topological part of the non-sypersymmetric M-theory using Polyakov's flat bundle description of non-linear $\sigma $ models. The new key ingredient from the geometric point of view was "characteristic classes for flat bundles". This idea about using characteristic classes of flat bundles came from the definition of a new invariant for Haefliger structures \cite{Z2}. This invariant can be defined for \emph{any} foliation in general; as far as physics is concerned however (and this includes the case of M-theory treated in \cite {Z1}), we are primarily concerned with a special kind of folia tions, called \emph{flat foliations of bundles}. This is due to the fact that a s Polyakov had noticed in \cite {polyakov}, $\sigma $ models can be thought of as flat principal bundles (se e \cite{Z1} for more details). Thus, hopefully, our invariant might be of some relevance whenever $\sigma $ models are met in physics. We organise this paper as follows: in section 2 we explain the strategy of the construction; in section 3 we provide all the details; in section 4 we give the invariant formula; in section 5 we calculate this invariant for the si mplest case of a principal bundle and in sections 6 and 7 we discuss some possible applications in physics. In section 3 we review some of the techniques from non-commutative topology which we shall use for defining the invariant. All other sections contain new original material. \section{Strategy} \subsection{Instantons} Let us recall some facts about instantons. We would like to think of our invariant as an \emph{analogue of the instanton number for foliations}. We consider a principal bundle $(P,\pi,M,G)$, where $M$ is the base manifold assumed to be compact and 4-dim for brevity, $G$ is SU(2) for simplici ty, $P$ is the total space of the bundle and $\pi$ is the projection. Assuming we have a connection $A$ on $P$ with curvature $F$, then the instanton number ignoring constants is simply $\int _{M}F\wedge F$, i.e. the second Chern number $c_2$ of the bundle $P$. We would like to think of this number slightly differently: moreorless by definition, any principal bundle $P$ over $M$ defines an element (K-class) of the group $K^0(M)$ (we forget equivariant K-theory for simplicity). Using the Chern-Weil homomorphism we get the Chern classes of $P$ which belong to the \emph{cohomology} groups $H^{2*}(M)$. Considering the (top dimensional) fundamental class $[M]$ of $M$ in the \emph{homology} group $H_*(M)$ of $M$ and taking the \emph{pairing} between homology and cohomology, which in this case is just integration over $M$, we get the instanton number. We can consider the Chern-Weil homomorphism from $K^0(M)\rightarrow H^{2*}(M)$ as a "black box" and forget all about cohomology for the moment; then the instanton number will be the result from pairings between K-theory and (singular) homology $H_*(M)$. Our construction since we are dealing with foliations (more accurately with the space of leaves of foliations) which provide a good example of non-commutative topological spaces, immitates the above picture: to each foliation we can associate a homology class which will be the analogue of the fundamental class $[M]$ above; this class however will belong to an appropriate homology theory called \emph{cyclic homology} and it is called \emph{transverse fundamental class} of the foliation. Moreover one can also construct a class in \emph{K-homology}, being the analogue of K-theory for our purpose. Then we use a formula for pairings between cyclic homology and K-homology to get our result. \subsection{Non-commutative Topology} In this subsection we would like to mention briefly what non-commutative topology is about. As its name suggests, this is one aspect of non-commutative geometry. Non-commutative geometry has appeared in physics literature some years ago mainly through the so called "quantised calculus". Anyway, the starting point of non-commutative topology is the fact that given any compact Hausdorff space $X$ say, the commutative ($C^*$)-algebra $C(X)$ of complex valued functions defined on $X$ can capture all the toplogical information of the space $X$ itself; in fact $X$ and $C(X)$ are completely equivalent, one can be uniquely constructed by the other. Conversely, given any commutative algebra $A$, say, there exists a compact Hausdorff space $X$ say, (called the spectrum of $A$) "realising" the commutative algebra $A$. Realising means that the commutative algebra $C(X)$ of complex valued functions on $X$ is essentially the algebra $A$. In mathematics terminology one says that the categories of compact Hausdorff spaces and commutative $C^*$-algebras are equivalent. This is the so called \emph{Gelfand's theorem}. We know however that there exist non-commutative $C^*$-algebras as well. The natural question then is whether one can find a "topological" realisation for them just like for the commutative ones. We are looking for a non-commutative analogue of Gelfand's theorem. This question is not fully answered in mathematics, it is related to the famous Boum-Connes conjecture. There are some things already known in mathematics and these are related to foliations. This is what we shall be using extensively in this paper. The appropriate framework is that of K-theory and various homology theories. During the '70s mathematicians (Baum, Douglas, Kasparov and others) developed a K-theory for arbitrary $C^*$-algebras (commutative or not) and it is a well-known theorem due to Serre and Swan that in the commutative case this K-theory reduces to Atiyah's original topological K-theory. Moreover in the 80's mathematicians (Connes, Loday, Quillen and others) developed a homology theory called \emph{cyclic homology} for arbitrary algebras which again in the commutative case gives in the limit the usual simplicial homology. So non-commutative topology, in terms of K-theory and various homology theories gives a generalisation of ordinary topology through Gelfand's theorem. A good example of a non-commutative topological space is the space of leaves of a foliation (see below for definitions). In general quotients of ordinary topological spaces by discrete groups give non-commutative (abreviated to \emph{"nc"} in the sequel) spaces. Good textbooks are \cite{wegge} and \cite{loday} for an introduction on K-theory of $C^*$-algebras and cyclic homology respectively. \subsection{The Invariant} In order to construct our invariant for \emph{any} foliation \cite{Z2}, we use some ideas from non-commutative geometry \cite{connes1}, \cite{connes2}, \cite{connes3}, \cite{conness}. The strategy is as follows: given any foliation $F$ of a manifold $V$, namely an integrable subbundle $F$ of $TV$, one can associate to it another manifold $\Gamma (F)$, called the \emph{graph} (or \emph{holonomy groupoid}) of the foliation introduced in \cite{win}. This is of dimension $dimV+dimF$. Using the complex line bundle $\Omega ^{1/2}(\Gamma (F))$ of $1/2$-densities defined on $\Gamma (F)$, we consider the set (actually vector space) of smooth sections of this line bundle equipped with a $*$ product, thus obtaining an algebra. We then complete this algebra in a "minimal" manner (in standard $C^*$-algebra theory this is called the \emph{reduced} $C^*$-algebra completion), thus we obtain a $C^*$-algebra denoted $C^{*}(F)$ which is naturally associated to our original foliation $F$. From now on one can forget the original foliation $F$ of $V$ alltogether and concentrate on its corresponding $C^*$-algebra $C^{*}(F)$. We are interested in the $K_0$ group of $C^{*}(F)$ and in its cyclic homology groups. If we pick a metric $g$ on the transverse bundle $t$ of $F$ we can construct in a natural way a $C^{*}(F)$-module $E(F)$, thus obtaining a class $[E(F)]$ in $K_0(C^{*}(F))$. Moreover to our foliation one can associate in a natural way a cyclic cocycle $[F]$ in the q-th cyclic homology group of the $C^*$-algebra $C^*(F)$, called the \emph{fundamental transverse cyclic cocycle} of the foliation, where $q$ is the codimension of the foliation $F$. Then we use the even pairing between K-homology and cyclic homology in this case, namely we consider the pairing $$\langle [E(F)],[F]\rangle :=(m!)^{-1}(F \# Tr)(E(F),...,E(F))$$ as was firstly introduced in the abstract algebraic context in \cite{connes3}. Hence we obtain a \emph{complex number} as a result from the above pairing and this complex number characterises our original foliation $F$. \section{The constructions in detail:} \subsection{Foliations} Let $V$ be a smooth manifold and $TV$ its tangent bundle. A smooth subbundle $F$ of $TV$ is called \emph{integrable} iff one of the following equivalent conditions is satisfied: 1. Every $x\in V $ is contained in a submanifold $W$ of $V$ such that $$T_y(W)=F_y$$ where $T_y$ denotes the tangent space over $y$. 2. Every $x\in V$ is in the domain $U\subset V$ of a submersion $p:U\rightarrow {\bf R}^q$ (q=codim $F$) with $F_y=ker(p_*)_y\forall y\in U$. 3. $C^{\infty }(V,F)={X\in C^{\infty }(V,TV); X_x\in F_x \forall x\in V}$ is a Lie subalgebra of the Lie algebra of vector fields on $V$. 4. The ideal $J(F)$ of smooth differential forms which vanish on $F$ is stable under differentiation:$d(J)\subset J$\\ The condition 3. is simply Frobenius' Theorem and 4. its dual.\\ \emph{Example:} Any 1-dimensional subbundle $F$ of $TV$ is integrable, but for $dimF\geq 2$ the condition is non-trivial; for instance if $V$ is the total space of a principal bundle with compact structure group, then we know that the subbundle of \emph{vertical} vectors is always integrable, but the \emph{horizontal} subbundle is integrable iff the connection is \textbf{flat}.\\ We shall make extensive use of this fact in this piece of work. A \emph{foliation} of $V$ is given by an integrable subbundle $F$ of $V$. The \emph{leaves} of the foliation are the maximal connected submanifolds $L$ of $V$ with $T_x(L)=F_x \forall x\in L$, and the partition of $V$ into leaves $V=\cup L_a$ where $a\in A$ is characterised geometrically by its "local triviality": every point $x\in V$ has a neighborhood $U$ and a system of local coordinates $(x^j)$, j=1,...,dimV which is called \emph{foliation chart} , so that the partition of $U$ into connected components of leaves, called \emph{plaques} (they are the leaves of the restriction of the foliation on $U$), corresponds to the partition of ${\bf R}^{dimV}={\bf R}^{dimF}\times {\bf R}^{codimF}$ into the parallel affine subspaces ${\bf R}^{dimF}\times pt$. Very simple examples indicate that the leaves $L$ may \textbf{not} be compact even if the manifold $V$ is and that the space of leaves $X:=V/F$ may \textbf{not} be Hausdorff for the quotient topology. The "rational torus" is such an example.\\ Throughout this paper we would mainly restrict our attention to \emph{two special kinds of foliations:} we consider a principal bundle $P$ with structure (Lie) group $G$ (assumed compact and connected) over a compact manifold $M$. The total space $P $ has automatically a foliation induced by the fibration: the leaves are the fibers which are isomorphic to the structure group $G$ and the space of leaves is just the base space $M$ with its manifold topology. We shall be refering to this foliation as the \emph{vertical} foliation of the principal bundle and it will be denoted $P_V$. Clearly, the dimension of this foliation is equal to the dimension of the group $G$, the integrable subbundle of $TP$ being in this case the vertical subbundle. The codimension is equal to the dimension of the base space $M$. Now if in addition a flat connection is given on our principal bundle, we have another foliation of the total space which we shall be referring to as the \emph{horizontal} or \emph{flat} foliation and it will be denoted $P_H$.. We shall study this foliation extensively in the following subsection. The dimension of this foliation equals the dimension of the base space and the codimension equals the dimension of the group. From this one can see that the vertical and the horizontal foliations of a principal bundle are \emph{transverse} to each other. Now the vertical foliation behaves very well; everything is compact and Hausdorff, as were the spaces we started with to build our bundle. In this case the general theory of foliations gives nothing more than the well-known theory of principal bundles. However, the horizontal foliation can suffer from various "pathological" deffects and for this reason it is interesting from the ncg point of view. Let us study it in greater detail. \subsection{Flat foliation of a principal bundle} To begin with, a flat connection on a principal bundle $P$ with structure group $G$ and base space $M$, corresponds to reduction of the structure group from $G$ to a subgroup isomorphic to a normal subgroup of the fundamental group of the base space $\pi_1(M)$. Moreover a (gauge equivalence class of a) flat conne ction also defines a (conjugacy class of a) representation $$H:\pi _1(M)\rightarrow G$$ If we identify the fundamental group with the group of covering translations of the universal covering $\tilde{M}$ of $M$ we get an action $\varkappa$ of $\pi _1(M)$ on $\tilde{M}\times G$ defined as follows: $$\varkappa :\pi _1(M)\times (\tilde{M}\times G)\rightarrow (\tilde{M}\times G)$$ $$(\gamma ,(\tilde{m},g))\mapsto (\gamma (\tilde{m}),H(\gamma )(g))$$ where we use the obvious notation $\gamma\in\pi _1(M), g\in G, \tilde{m}\in M$. This action gives a commutative diagram: \begin{equation} \begin{CD} \tilde{M}\times G@>pr>>\tilde{M}\\ @V\pi VV @VVqV\\ P'=(\tilde{M}\times G)/\varkappa @>>p>M\\ \end{CD} \end{equation} where $pr$ is the canonical projection, $\pi $ is the quotient map by $\varkappa$, $p$ is uniquely induced by $pr$ and $q$ is just the map from the universal covering space to the original space. This construction is called \emph{suspension of the representation $H$}. One can prove that the map $\pi $ is a covering map and that if $\digamma :=ImH$ is endowed with the induced topology, then $\xi _H=(P',p,M)$ is a fiber bundle with fiber $G$, total space $P'$, base $M$, projection $p$ and structure group $\digamma $. To study the geometric properties of suspensions we introduce a new topology on the total space $P'$ of $\xi _H$. We denote by $G^{\delta }$ the set $G$ supplied with the \textit{discrete} topology. Then the action $\varkappa$ of $\pi _1(M)$ on $\tilde{M}\times G^{\delta }$ remains continuous and the map $\pi :\tilde{M}\times G^{\delta }\rightarrow P'$ induces on $P'$ a new topology which is \textbf{finer} than its manifold topology. We denote by $P^{\delta }$ the set $P'$ supplied with this topology. The topology on $\tilde{M}\times G^{\delta }$ and the topology $P^{\delta }$ are called the \emph{leaf topologies}. Then the suspension diagram below is a commutative diagram of covering maps: \begin{equation} \begin{CD} \tilde{M}\times G^{\delta }@>pr>>\tilde{M}\\ @V\pi VV @VVqV\\ P^{\delta }@>>p>M\\ \end{CD} \end{equation} The topological space $P^{\delta }$ is not connected unless the fiber is contractible. A connected component of $P^{\delta }$ is called a leaf of $\xi _H$. Each point $x=\pi (\tilde{m},g)\in P'$ belongs to exactly one leaf which is denoted $L_x$ and equals $\pi (\tilde{M}\times {g})$. The leaves are injectively immersed submanifolds of $P'$ but in general not embedded. They are transverse to the fibers of $\xi _H$. Conjugate representations $H$ and $H'$ give suspension bundles $\xi _H$ and $\xi _{H'}$ which are isomorphic. Let now $x=\pi (\tilde{m},g)$. Then the representation $$H_x :\pi _1(L_x)\rightarrow G$$ with image $\digamma _g$, is called the \emph{holonomy} representation of the leaf $L_x$ at the point $x$. The group $\digamma _g$ is the \emph{holonomy group} of the leaf $L_x$ at the point $x$. $\digamma _g$ is the isotropy group of $\digamma $ in $g\in G$. Moreover $\pi _1(L_x)$ is isomorphic to the isotropy group of $\pi _1(M)$ in the point $g\in G$, namely $\pi _1(L_x)\cong \{\gamma\in\pi _1(M)|H(\gamma )g=g\}$. See also \cite{hirsch}.\\ There is a topological way to characterise these flat bundles which is by using \emph{classifying spaces for flat bundles} in a fashion analogous for ordinary bundles, namely:\\ Let $G$ be a connected Lie group and let $G^{\delta }$ denote the same group with the discrete topology. The \textit{Milnor join construction} for $G$ defines a connected space $BG$ which is the classifying space for principal G-bundles. The same construction applied to $G^{\delta }$ yields a connected topological space $BG^{\delta }$ which is an \emph{Eilenberg-Maclane space} $K(G,1)$, namely $\pi _1(BG)=G$ and $\pi _j(BG)=0$ for $j>1$. The inclusion $i:G^{\delta }\rightarrow G$ induces a continuous map $Bi:BG^{\delta }\rightarrow BG$. As sets these two spaces are the same with the source having \emph{finer} topology than the range. The difference in these two topologies is measured by introducing the homotopy fiber $BG'$. This is defined by first replacing $Bi$ with a homotopy equivalent weak fibration over $BG$, then take for $BG'$ the (homotopy class of the) fiber. The description then is just the construction of the \emph{Puppe Sequence} for $Bi$ (cf \cite{jcw}). Choose a base point in $BG^{\delta }$ and consider its image in $BG$. Then let $\Omega (BG)$ and $P(BG)$ denote the space of based loops and paths with initial point of $BG$ respectively. Let $e$ be the end point map of a path. Then one has a fibration $$\Omega (BG)\hookrightarrow P(BG)\rightarrow BG$$ where the second map is $e$. Then define $BG'$ via the homotopy pull-back diagram: \begin{equation} \begin{CD} \Omega (BG)@>>>\Omega (BG)\\ @VVV @VVV\\ BG'@>>>P(BG)\\ @VVV @VVeV\\ BG^{\delta }@>>Bi>BG\\ \end{CD} \end{equation} A principal G-bundle $P$ over a manifold $M$ is equivalent to giving an open covering for $M$ and the transition functions. This data defines a continuous map $g_P:M\rightarrow BG$. If the transition functions are locally constant, namely if the bundle $P$ is flat, then $g_P$ can be factored through $BG^{\delta }$ as a continuous map. A choice of transition functions which are locally constant is equivalent to specifying a flat G-structure on $P$. Hence $P$ has a horizontal foliation whose holonomy map $a:\pi _1(M)\rightarrow G$ defines the classifying map $Ba:M\rightarrow BG^{\delta }$. Conversely, given a continuous map $Ba:M\rightarrow BG^{\delta }$, there is induced a representation $a:\pi _1(M)\rightarrow G$ and a corresponding flat principal G-bundle $P_a=\tilde{M}\times _{\pi _1(M)}G$, where $\tilde{M}$ is the universal covering of $M$. The topological type of the G-bundle $P_a$ is determined by the composition $$g_a:M\rightarrow BG^{\delta }\rightarrow BG$$ The principal bundle is trivial iff $g_a$ is homotopic to the constant map $M\rightarrow pt$. The choice of the homotopy is equivalent to specifying a global section on $P_a$. \subsection{Groupoids and $C^{*}$-algebras associated to Foliations} The next step is to associate the \emph{holonomy groupoid} to any foliation. In general a groupoid is roughly speaking a small category with inverses, or more precisely \textbf{Definition 1:} A groupoid consists of a set $\Gamma $, a distinguished subset $\Gamma ^{(0)}$ of $\Gamma $, two maps $r,s:\Gamma \rightarrow \Gamma ^{(0)}$ and a law of composition $$\circ :\Gamma ^{(2)}:={(\gamma _1,\gamma _2)\in \Gamma \times \Gamma ; s(\gamma _1)=r(\gamma _2)}\rightarrow \Gamma $$ such that:\\ 1.$s(\gamma _1\circ \gamma _2)=s(\gamma _2)$, $r(\gamma _1\circ \gamma _2)=r(\gamma _1)$ $\forall (\gamma _1,\gamma _2)\in \Gamma ^{(2)}$ 2. $s(x)=r(x)=x \forall x\in \Gamma ^{(0)}$ 3.$\gamma \circ s(\gamma )=\gamma$, $r(\gamma )\circ \gamma =\gamma \forall \gamma \in \Gamma $ 4. $(\gamma _1\circ\gamma _2)\circ\gamma _3=\gamma _1\circ (\gamma _2\circ\gamma _3)$ 5. Each $\gamma $ has a two sided inverse $\gamma ^{-1}$, with $\gamma\gamma ^{-1}=r(\gamma )$ and $\gamma ^{-1}\gamma =s(\gamma )$\\ The maps $r$, $s$ are called \emph{range} and \emph{source} maps.\\ In the category theory terminology, $\Gamma ^{(0)}$ is the space of objects and $\Gamma ^{(2)}$ is the space of morphisms. \textbf{Definition 2:} A \emph{smooth} groupoid $\Gamma $ is a groupoid together with a differentiable structure on $\Gamma $ and $\Gamma ^{(0)}$ such that the maps $r$, $s$ are submersions and the object inclusion map $\Gamma ^{(0)}\rightarrow \Gamma $ is smooth, as is the composition map $\Gamma ^{(2)}\rightarrow \Gamma $.\\ The notion of a $\frac{1}{2}$-\emph{density} on a smooth manifold allows one to define in a canonical manner the \emph{convolution algebra} of a smooth groupoid $\Gamma $.\\ Specifically, given $\Gamma $, let $\Omega ^{1/2}$ be the line bundle over $\Gamma $ whose fiber $\Omega _{\gamma }^{1/2}$ at $\gamma\in \Gamma , r(\gamma )=x, s(\gamma )=y$, is the linear space of maps $$\rho :\wedge ^{k}T_{\gamma }(\Gamma ^x)\otimes \wedge ^{k}T_{\gamma }(\Gamma _y)\rightarrow \bf{C}$$ such that $$\rho (\lambda\nu )=|\lambda |^{1/2}\rho (\nu)\forall\lambda\in \bf{R}$$ Here $\Gamma _y={\gamma\in \Gamma ;s(\gamma )=y}, \Gamma ^x={\gamma\in \Gamma ;r(\gamma )=x}$, and $k=dimT_{\gamma }(\Gamma ^x)=dimT_{\gamma }(\Gamma _y)$ are the dimensions of the fibers of the submersions $r:\Gamma \rightarrow \Gamma ^{(0)}$ and $s:\Gamma \rightarrow \Gamma ^{(0)}$.\\ Then we endow the linear space $C_c^{\infty}(\Gamma ,\Omega ^{1/2})$ of smooth compactly supported sections of $\Omega ^{1/2}$ with the convolution product $$(a*b)(\gamma )=\int _{\gamma _1\circ\gamma _2=\gamma }a(\gamma _1)b(\gamma _2)$$ $\forall a,b\in C_c^{\infty }(\Gamma ,\Omega ^{1/2})$ where the integral on the RHS makes sense since it is the integral of a 1-density, namely $a(\gamma _1)b(\gamma _1^{-1}\gamma )$, on the manifold $\Gamma ^x$, $x=r(\gamma )$.\\ One then can prove that if $\Gamma $ is a smooth groupoid and $C_c^{\infty }(\Gamma ,\Omega ^{1/2})$ is the convolution algebra of smooth compactly supported $\frac{1}{2}$-densities with involution *, $f^*(\gamma )=\overline{f(\gamma ^{-1})}$. Then for each $x\in \Gamma ^{(0)}$, the following defines an involutive representation $\pi _x$ of $C_c^{\infty }(\Gamma ,\Omega ^{1/2})$ in the Hilbert space $L^2(\Gamma _x)$: $$(\pi _x(f)\xi )(\gamma )=\int f(\gamma _1)\xi (\gamma _1^{-1}\gamma )$$ $\forall\gamma\in \Gamma _x, \xi\in L^2(\Gamma _x)$ \\ The completion of $C_c^{\infty }(\Gamma ,\Omega ^{1/2})$ for the norm $||f||=sup_{x\in \Gamma ^{(0)}}||\pi _x(f)||$ is a $C^*$-algebra denoted $C^*_r(\Gamma )$. Moreover one defines the $C^*$-algebra $C^*(\Gamma )$ as the completion of the involutive algebra $C_c^{\infty }(\Gamma ,\Omega ^{1/2})$ for the norm \\ \\ $||f||_{max}=sup||\pi (f)||$; \{$\pi $ involutive Hilbert space representation of $C_c^{\infty }(\Gamma ,\Omega ^{1/2})$\} \\ \\ After this general introduction to groupoids and to $C^*$-algebras associated to them, now we pass to groupoids and $C^*$-algebras associated to foliations.\\ Let $(V,F)$ be a foliated manifold of codim $q$. Given any $x\in V$ and a small enough open set $W$ in $V$ containing $x$, the restriction of the foliation $F$ to $W$ has as its leaf space an open set of $\bf{R}^q$ which we shall call a transverse neighborhood of $x$. In other words, this open set $W/F$ is the set of plaques around $x$. Now given a leaf $L$ of $(V,F)$ and two points $x, y \in L$, any simple path $\gamma $ from $x$ to $y$ on $L$ uniquely determines a germ $h(\gamma )$ of a diffeomorphism from a transverse neighborhood of $x$ to one of $y$. This depends only on the homotopy class of $\gamma $ and is called the \emph{holonomy} of the the path $\gamma $. The \textit{holonomy groupoid of a leaf} $L$ is the quotient of its fundamental groupoid by the equivalence relation which identifies two paths $\gamma _1$, $\gamma _2$ from $x$ to $y$ both in $L$ iff $h(\gamma _1)=h(\gamma _2)$. Here by the fundamental groupoid of a leaf we mean the groupoid $\Gamma =L\times L$, $r$, $s$ are the two projections, $\Gamma ^{(0)}=L$ and the composition is $(x,y)\circ (y,z)=(x,z)$. (From this one can see that every space is a groupoid). The holonomy covering $\tilde{L}$ of a leaf $L$ is the covering of $L$ associated to the normal subgroup of its fundamental group $\pi _1(L)$ given by paths with trivial holonomy. The \textbf{holonomy groupoid} or \textbf{graph of the foliation} is the union $\Gamma $ of the holonomy groupoids of its leaves. Given an element $\gamma $ of $\Gamma $ we denote by $s(\gamma )=x$ the origin of the path $\gamma $ and by $r(\gamma )=y$ its end point, where $r$, $s$ are the range and source maps as in the general case. An element of $\Gamma $ is thus given by two points $x=s(\gamma )$ and $y=r(\gamma )$ of $V$ together with an equivalence class of smooth paths : the $\gamma (t)$, $t\in [0,1]$ with $\gamma (0)=x$ and $\gamma (1)=y$, tangent to the bundle $F$, namely with $\frac{d\gamma }{dt}\in F_{\gamma (t)}\forall t\in \bf{R}$, identifying $\gamma _1$ and $\gamma _2$ as equivalent iff the holonomy of the path $\gamma _2\gamma _1^{-1}$ at the point $x$ is the identity. The graph $\Gamma $ has an obvious composition law. For $\gamma _1$ and $\gamma _2$ in $\Gamma $, the composition $\gamma _1\circ\gamma _2$ makes sense if $s(\gamma _1)=r(\gamma _2)$. The groupoid $\Gamma $ is by construction a (not necessarily Hausdorff) manifold of dimension $dim\Gamma =dimV+dimF$. \textbf{Definition 3:} The \textbf{$C^*$-algebra of the foliation} is exactly the $C^*$-algebra of its graph, as described for arbitrary groupoids above.\\ For our foliations of interest, the graph $\Gamma $ is the following: for the ve rtical foliation is just the manifold $P\times G$ whereas for the horizontal foliation is $P\times _a\tilde{M}$, where $a$ is the representation from $\pi _1(M)$ to $G$ induced by the flat connection 1-form (via the holonomy). Moreover the distinguished subset $\Gamma ^{(0)}$ in both cases is the manifold we want to foliate, namely $P$, the total space of our bundle in our case.\\ The $C^*$-algebras associated to our foliations are: for the vertical foliation is $C(M)$ tensored with compact operators which act as smoothing kernels along the leaves which in turn is strongly Morita equivalent to just $C(M)$, whereas for the horizontal foliation is strongly Morita equivalent (abreviated to SME) to $C(P)\rtimes \pi _1(M)$. (Note: the representation of the fundamental group of the base onto the structure Lie group induced by the flat connection 1-form used enters the definition of the crossed product). The first algebra is \textbf{commutative} (up to SME), but the second \textbf{is not!} It is for this reason that we can see now that ncg has an important role to play, in fact we are deeply in the ncg setting. Obviously if the space is simply connected, i.e. $\pi _1$ vanishes, non-commutativity is lost. We would like to emphasise that in all cases in the literature where some "non-commutative" algebras were used, especially in connection to the well-known Connes-Lott model for electroweak theory (or even QCD), these algebras are in fact SME to commutative ones. Hence in terms of topology, this is not a real non-commutative case. \subsection{K-classes associated to foliations} We shall give the general construction for an arbitrary foliation.\\ Let $(V,F)$ be a foliated manifold and $t=TV/F$ the transverse bundle of the foliation. The holonomy groupoid $\Gamma $ of $(V,F)$ acts in a natural way on $t$ by the differential of the holonomy, thus for every $\gamma\in\Gamma$, $\gamma :x\rightarrow y$ determines a linear map $h(\gamma ):t_x\rightarrow t_y$. We denote this action by $h$. It is not in general possible to find a Euclidean metric on $t$ which is invariant under the above action of $\Gamma $. Let $g$ be an arbitrary smooth Euclidean metric on the real vector bundle $t$. Thus for $\xi\in t_x$ we let $||\xi ||_{g}=(\langle\xi ,\xi \rangle _{g})^{1/2}$ be the corresponding norms and inner products and drop subscript $g$ henceforth. Using $g$ we define a $C^*$-module $E$ on the $C^*$-algebra $C^*_r(V,F)$ of the foliation. Recall that $C^*_r(V,F)$ is the completion of the convolution algebra $C_c^{\infty }(\Gamma ,\Omega ^{1/2})$ which acts by right convolution on the linear space $C_c^{\infty }(\Gamma ,\Omega ^{1/2}\otimes r^*(t_{\bf{C}}))$ denoted $\Lambda $ for simplicity and $t_{\bf C}$ is the complexification of the transverse bundle $t$: $$(\xi f)(\gamma )=\int _{\Gamma ^y}\xi (\gamma _1)f(\gamma _1^{-1}\gamma )$$ where $y=r(\gamma )$. Endowing the complexified bundle $t_{\bf{C}}$ with the inner product associated to $g$ and anti-linear in the first variable, the following formula defines a $C_c^{\infty }(\Gamma ,\Omega ^{1/2})$-valued inner product $$\langle\xi ,n\rangle (\gamma )=\int _{\Gamma ^y}\langle\xi (\gamma _1^{-1}),n(\gamma _1^{-1}\gamma )\rangle $$ for any $\xi ,n\in C_c^{\infty }(\Gamma ,\Omega ^{1/2}\otimes r^*(t_{\bf{C}}))$ One then checks that the completion $E$ of the space $C_c^{\infty }(\Gamma ,\Omega ^{1/2}\otimes r^*(t_{\bf{C}}))$ for the norm $$||\xi ||=(||\langle\xi ,\xi \rangle ||_{C^*_r(V,F)})^{1/2}$$ becomes a $C^*$-module over $C^*_r(V,F)$. If one takes also the action $h$ of $\Gamma $ on $t$ into account, with some extra effort one can make $E$ into a $(\Lambda ,\Sigma )$-bimodule (for the definition of the algebra $\Sigma $ see below). The first construction thus gives us an element $E$ of $K_0(C^*_r(V,F))$ whereas the second gives $E$ as an element of $KK_0(\Lambda ,\Sigma )$, \emph{Kasparov's bivariant K-Theory}. (Recall that the 0th Kasparov's bivariant K-group in this case consists of stable isomorphism classes of $(\Lambda ,\Sigma )$-bimodules). We shall use this action $h$ to define a left action of $C_c^{\infty }(\Gamma ,\Omega ^{1/2})$ on $E$ by: $$(f\xi )(\gamma )=\int _{\Gamma ^y}f(\gamma _1)h(\gamma _1)\xi (\gamma _1^{-1}\gamma )$$ $\forall f\in C_c^{\infty }(\Gamma ,\Omega ^{1/2}), \xi\in E$ One then can prove that for any $f \in C_c^{\infty }(\Gamma ,\Omega ^{1/2})$ the above formula defines an endomorphism $\lambda (f)$ of the $C^*$-module $E$ whose adjoint $\lambda (f)^*$ is given by $$(\lambda (f)^*\xi )(\gamma )=\int _{\Gamma ^y}f^{\#}(\gamma _1)h(\gamma _1)\xi (\gamma _1^{-1}\gamma )$$ \\ where \\ $$f^{\#}(\gamma ):=\widetilde{f}(\gamma ^{-1})\Delta (\gamma )$$ \\ and \\ $$\Delta (\gamma )=(h(\gamma )^{-1})^{t}h(\gamma )^{-1}\in End(t_{\mathbf{C}}(r(\gamma )))$$ This shows that unless the metric on $t$ is $\Gamma $-invariant, the representation $\lambda $ is not a *-representation, the subtle difference between $\lambda (f)^*$ and $\lambda (f^*)$ being measured by $\Delta $. In particular $\lambda $ is not in general bounded for the $C^*$-algebra norms on both $End_{C^*_r(V,F)}E$ and $C^*_r(V,F)\supset C_c^{\infty }(\Gamma ,\Omega ^{1/2})$. However $\lambda $ is a closable homomorphism of $C^*$-algebras, namely, the closure of the graph of $\lambda $ is the graph of a densely defined homomorphism. Then with the graph norm $$||x||_{\lambda }=||x||+||\lambda(x)||$$ the domain $\Sigma $ of the closure $\widetilde{\lambda }$ of $\lambda $ is a Banach algebra which is dense in the $C^*$-algebra $\Lambda =C^*_r(V,F)$. The $C^*$-module $E$ is then a $(\Lambda ,\Sigma )$-bimodule. This particular module $E$ we constructed here will be the one of the two main ingredients which define the invariant we want and we shall denote it $E(F)$. \subsection{Cyclic classes associated to foliations (transverse fundamental cyclic cocycle)} We begin with some definitions from cyclic homology: \textbf{Definition 1:} 1. A \emph{cycle} of dimension $n$ is a triple $(\Omega ,d,\int )$ where $\Omega =\oplus _{j=0}^n \Omega ^j$ is a graded algebra over ${\bf C}$, $d$ is a differential on $\Omega $'s and $\int :\Omega ^n\rightarrow C$ is a closed graded trace on $\Omega $. 2. Let $A$ be an algebra over $\bf{C}$. Then a cycle over $A$ is given by a cycle $(\Omega ,d,\int )$ and a homomorphism $\rho :A \rightarrow \Omega ^0$. A cycle over $A$ of dimension $n$ is essentially determined by its \emph{character} which is the following $(n+1)$-linear functional on $A$: $$\tau (a^0,...,a^n)=\int \rho (a^0)d(\rho (a^1))d(\rho (a^2))...d(\rho (a^n))$$ \\ $\forall a^j\in A$ \\ One can then prove that this is a \emph{cyclic cocycle} of $A$, namely it defines a cohomology class in the cyclic homology of $A$ and that the above is a necessary and sufficient statement. We shall now describe the transverse fundamental class associated to foliations. There is a general construction for arbitrary foliations which is quite involving since one has to \emph{complete} the graded algebra. This is so because the transverse bundle of the foliation may not be integrable and in this case derivation along transverse directions will not be a differential. We, however, are primarily interested in our two special kinds of foliations, the vertical and the horizontal foliation of a principal bundle. These foliations are transverse, both are integrable so derivatives are differentials and hence one does not have to complete the graded algebras. We refer to \cite{conness} for the general construction. Here we shall only describe the classes wich are associated to our two foliations: The vertical and the horizontal (or flat) foliations of the total space of our principal bundle $P$ will be denoted $(P_V)$ and $(P_H)$ respectively. One then has that there is a natural cycle for the algebra of each foliation, namely: \textbf{Vertical Foliation}, cycle denoted $[P_V]$: The natural cycle canonically associated to the algebra $C_c^{\infty }(P\times G,\Omega ^{1/2})$ of the vertical foliation consists of: 1. The graded algebra $C_c^{\infty }(P\times G,\Omega ^{1/2}\otimes r^*(\wedge P_H^*))$ where $P_H$ is the \emph{horizontal} subbundle (i.e. the transverse bundle to the vertical foliation). 2. The differential $d=d_V+d_H+\theta $ where $$d_H:C^{\infty }(P,\wedge ^r P_V^*\otimes\wedge ^s P_H^*)\rightarrow C^{\infty }(P,\wedge ^r P_V^*\otimes\wedge ^{s+1}P_H^*)$$ $$d_V:C^{\infty }(P,\wedge ^rP_V^*\otimes\wedge ^s P_H^*)\rightarrow C^{\infty }(P,\wedge ^{r+1}P_V^*\otimes\wedge ^s P_H^*)$$ $\theta $ means contraction with the section $\theta\in C^{\infty }(P,P_V\otimes\wedge ^2P_H^*)$, where $\theta $ is defined by: $\theta (p_H(X),p_H(Y)=p_V([X,Y])$ for any pair of horizontal vector fields $X, Y \in C^{\infty }(P,P_H)$ and $(p_H,p_V)$ is the isomorphism $TP\rightarrow P_H\oplus P_V$ given by $P_H$. 3.The trace is defined via $$\tau (w)=\int _{\Gamma ^{(0)}} w $$ where $\Gamma $ is the graph of the vertical foliation $\Gamma =P\times G$.\\ \textit{Similarly one defines a fundamental class for the horizontal foliation}. One then can define the character of these cycles---essentially the trace---which is a class in the cyclic homology of the appropriate algebra for each foliation \cite{connes2}. \textbf{Note:} Since now we have a cyclic homology class, say $\phi $ of the algebra of the foliation, say $\Lambda $, we automatically have a map $$ K_i(\Lambda )\rightarrow \bf{C}$$ given by pairing it with K-group elements ($i=0,1$ above) to get index theorems for \emph{leafwise} elliptic operators. Let us mention here that the analytic \emph{Index} of an operator \emph{elliptic along the leaves of an arbitrary foliation} say $(V,F)$, is an element of $K_0(C^*(V,F))$, being in fact a generalisation of the index of families of elliptic operators considered by Atiyah and Singer. (In the Atiyah-Singer case of families of elliptic operators one is dealing with the foliation induced by the fibration, which is the commutative geometry case). The operator \emph{itself} which is elliptic along the leaves of the foliation is an element of $KK(C^*(V,F),C(V))$.\\ \section{Invariant for the nl$\sigma $m} The final step then is to make use of the general formula for pairings between K-homology and cyclic homology. In more concrete terms, one has: \textbf{Definition:} Let $A$ be an algebra. Then the following equality defines a bilinear pairing between K-theory and cyclic homology: 1. Even case: $K_0(A)$ and $HC^{ev}(A)$: $$\langle [e],[\phi ]\rangle :=(m!)^{-1}(\phi \# Tr)(e,...,e)$$ \\ for $e\in K_0(A)$ using the idempotents' description and $\phi\in HC^{2m}(A)$ and where $\#$ is the \emph{cup product} in cyclic homology (see for instance \cite{connes3} for the precise definition). 2. Odd case: $K_1(A)$ and $HC^{odd}(A)$: $$\langle [u],[\phi ]\rangle =\frac{1}{\sqrt{2i}}2^{-n}\Gamma (\frac{n}{2}+1)^{-1}(\phi \# Tr)(u^{-1}-1,u-1,u^{-1}-1,...,u-1)$$ \\ This is an important point because by pairing the $C^*$-module $E$ we constructed previously naturally associated to the foli ation considered with the cyclic cocycle naturally associated to the foliation, we get an invariant for arbitrary foliations. In particular if we apply this to the horizontal foliation, we get a \emph{complex number} which is an invariant for the nl$\sigma$m. Namely one has: $$\langle [E(P_H)],[P_H]\rangle =(m!)^{-1}((P_H)\#Tr)(E(P_H),...,E(P_H))\in \bf{C}$$ In more concrete terms, assuming that $E(P_H)\in M_k(C^{*}(P_H))$ for some $k$ (where $C^{*}(P_H)$ is the corresponding $C^{*}$-algebra to the horizontal foliation) and $[P_H]\in Z^{q}(C^{*}(P_H))$ where $Z$ denotes cyclic cocycles and $q$ is the codimension of the horizontal foliation, then $(P_H)\#Tr\in Z^{q}(M_k(C^{*}(P_H)))$ is defined by $$((P_H)\#Tr)(a^0\otimes m^0,...,a^q\otimes m^q)=(P_H)(a^0,...,a^q)Tr(m^0...m^q)$$ for any $a^i\in C^{*}(P_H), m^i\in M_k({\bf C})$, $i=1,...,q$.\\ \emph{Note:} Let us mention that the odd case formula is related to the $\eta $ invariant for leafwise elliptic operators, see \cite{douglas1}, \cite{douglas2} which in turn is related to global anomalies and to the Freedman-Townsend invariance (cf \cite{witten2}, \cite{Z2}, \cite{townsend}, \cite{henneaux}). \section{An example: principal fibre bundles} In order to get some more insight to this pairing we shall try to calculate it for the case of principal bundles (vertical foliation) which is the simplest example. We begin by describing the graph in detail: the set $\Gamma $ in this case is the manifold $P\times G$, the distinguished subset $\Gamma ^{(0)}=P\times \{e\}$ and denoting the action (on the right) of $g\ni G$ on $p\ni P$ simply by $(p,g)\mapsto pg$, one has that the range and source maps are respectively $r(p,g)=p$ and $s(p,g)=pg$, the inverse $(p,g)^{-1}=(pg,g^{-1})$ and the law of composition is $(p_{1},g_{1})\circ (p_{2},g_{2})=(p_{1},g_{1}g_{2})$ if $p_{1}g_{1}=p_{2}$. Obviously the set $\Gamma ^{(2)}=P\times G\times G$. Moreover we recall that the $C^*$-algebra for the vertical foliation is strongly Morita equivalent to $C(M)$, We now make use of two important facts: 1. $K_0(C(M))=K^0(M)$, namely \emph{Serre-Swan theorem}\\ and\\ 2. $H^{*}_{cont}(C(M))=H_{*}(M)$\\ where on the RHS we have the ordinary homology of $M$ (with complex coefficients) and by definition for the LHS we have $H^{*}(C(M)):=Lim_{\rightarrow} (HC^{n}(C(M)),S)$ (see \cite{connes3} for explanations of the notation), $HC^{*}$ denotes cyclic homology and "cont" means restriction to continuous linear functionals. The first fact says that for commutative $C^*$-algebras one gets Atiyah's topological K-theory for the underlying space (described in terms of stable isomorphism classes of complex vector bundles over the space considered) and the second says that in the commutative case again cyclic homology is "roughly speaking" the ordinary homology of the underlying space (and thus we see that non-commutative geometry reduces to ordinary geometry in the commutative case). Since we have these two results in our desposal, we shall try to reduce the whole discussion in terms of bundles and ordinary homology theory because this is more comprehensible. In order to describe the pairing then we need the transverse fundamental cyclic cocycle: we shall give a simple dimensional argument here; the exact computations are rather too technical to be presented in grater detail. The cyclic cocycle we will get from the vertical foliation will be of dimension equal to the codimension of the vertical foliation which is equal to the dimension of the base space of our bundle. Moreover as we mentioned above, in this case the $C^*$-algebra of this foliation is SME to the algebra of functions on the base $C(M)$. This is a commutative $C^*$-algebra whose cyclic homology is moreorless the de Rham cohomology of the base space. Hence it is not too hard to suspect that we get a top homology class, which in fact turns out to be the fundamental class of our base space [M]. For the module denoted $E$ above in this case one uses the following fact: it is a consequence of Serre-Swan theorem mentioned above that the link between topological K-theory and K-theory of commutative $C^*$-algebras is that given a complex vector bundle over $M$ (thus a topological K-class), one considers the corresponding $C(M)$-module of smooth sections of the given complex vector bundle. Thus in this case the module $E$ we get is $C^{\infty }_{c}(P\times G,\Omega ^{1/2}\otimes r^{*}(t_{\bf C}))$, hence we can recover the corresponding complex vector bundle over $M$ as follows: If we denote by $(P,\pi ,G,M)$ our original principal bundle and we consider the vertical foliation, then its normal bundle $t$ would be $\pi ^{*}(TM)$, where $TM$ is the tangent bundle of $M$. We prefer the topological K-theory description which in this case is rather easy to read: the bundle associated to this $C(M)$ module $E$ is: \begin{equation} \begin{CD} \Omega ^{1/2}\otimes pr_{1}^{*}\pi ^{*}(TM)@>>>\pi ^{*}(TM)@>>>TM\\ @VVV @VVV @VV\tau _{M}V\\ P\times G@>>pr_{1}>P@>>\pi >M\\ \end{CD} \end{equation} where the fibre of the line bundle $\Omega ^{1/2}$ is the linear space of maps $\rho :\wedge ^{dimG}T_{\gamma }(G)\otimes \wedge ^{dimG}T_{\gamma }(G)\rightarrow {\bf C}$ satisfying the well-known property for 1/2-densities. Hence in this case the result will be the number we get if we take the bundle $\Omega ^{1/2}\otimes pr_{1}^{*}\pi ^{*}(TM)$ \emph{seen as a bundle over $M$}, then apply the ordinary Chern character to it and integrate over $M$. The result will be a combination of the Pontryagin class of $TM$ and the second Chern class of $P$ (recall that we assumed $M$ to be 4-dim and $P$ is an SU(2) bundle) which is something expected. There are some subtleties though: we have the bundle $\Omega ^{1/2}\otimes pr_{1}^{*}\pi ^{*}(TM)$ over the graph which is $P\times G$. We want to see this as a bundle over $M$. We consider firstly the factor $pr_{1}^{*}\pi ^{*}(TM)$. This is indeed a bundle over $M$ with fibre $G\times G\times {\bf R}^{dim M}$ but this is neither a vector nor a principal bundle and in order to talk about characteristic classes one actually needs the one or the other. In order not to change the topology then, which is what we are mainly interested in, we can consider the vector bundle $TM\otimes adP$ instead, where $adP$ is the adjoint bundle to $P$. To study formally the classes of $TM\otimes adP$ is an exercise in mathematics (we forget the pull-backs since they can be treated easily). The point is that we shall get a combination of the Pontryagin classes of $TM$ (or Chern if we complexify) and of Chern classes of $P$. The later is known since the bundle is given whereas the former can be computed from topological information of $M$ itself. For example for simply connected closed 4-manifolds one has from the Hirzebruch signature formula that (see \cite{hirzebruch}): $$p_1 = 3\tau =3(b^{+} - b^{-})$$ where $\tau $ is the signature. As about the other factor, the 1/2-densities of the graph, seen as a rank 1 real bundle over the graph $P\times G$, is rather dull. It will be determined by $\omega _1$, the first Stiefel-Whitney class: either trivial ($\omega _1=0$) or it is non-orientable ($\omega _1=1$). (\emph{Note:} half densities over a complex manifold say $N$ with $dimN=k$ are slightly more complicated: in this case its class will be $\frac{1}{2}c_1(\wedge ^kT^*)$, where $\wedge ^kT^*$ is the canonical bundle, so it will correspond to $spin ^c$ structures on $N$). But we still have the same problem that over $M$ it is neither a vector nor a principal bundle. This can be overcome as before; the point is that since it is a dull bundle over $P\times G$, the projection does not change anything, so as a bundle over $M$ it will be determined by the topology of $P$, hence we also have Chern classes of $P$. What we gave above was a qualitative description and the lesson was that the invariant will be some combination of Chern numbers of $P$ and the Pontryagin number of the tangent bundle $TM$ of $M$. The key point is that $TM$ appears because it is the \emph{transverse bundle} of the vertical foliation. We expect then that characteristic classes of the transverse bundle should be important in general. What about the flat foliation then, which is the case which is related to nl$\sigma $m in general and to M-theory in particular in physics? This is a purely nc case. Computations are much harder and a picture involving bundles is impossible since we do not have Serre-Swan theorem. Moreover cyclic homology of nc algebras has no relation to the usual topology. We can still say something though: first of all, since we have a flat principal bundle, all its characteristic classes vanish, so we get nothing from them. Since the base space $M$ is not simply connected, gauge inequivalent classes of flat connections are characterised by their holonomy. So we expect the holonomy to play some role. Moreover, from the vertical foliation case we saw that the normal bundle of the foliation also plays a vital role. \textbf{Note:} It is not always true that in the commutative case cyclic homology identifies well with ordinary homology, in fact this is always essentially true only for $H_0$. There may be complications. However this point will not be treated in this article with greater detail. See \cite{connes5}. \section{Relation to Physics} There are three cases in physics where this invariant may play some role: \emph{1. Nl$\sigma $m.} As it is well-known, $\sigma $ models classically describe harmonic maps between two Riemannian manifolds (target and source spaces). From the genral remarks we made in the preceeding section, we said that we expect the invariant to include the holonomy of the flat connection (namely $\pi _1$ of the source space) plus the topology of the space of leaves (target space). Hence this invariant should contain information about the topology of both manifolds involved in $\sigma $ models. Characteristic classes of foliations may provide the way to calculate the invariant (analogue of Chern-Weil theory, see \cite{tondeur}). Moreover, in the flat foliation case the invariant describes the topological charge of the M-theory Lagrangian density suggested in \cite{Z1} Another application is the following (we thank Dr S. T. Tsou for pointing this out to us): we know from Polyakov (\cite{polyakov}) that Yang-Mills theories can be formulated as nl$\sigma $m on the \emph{loop space}. This point of view very recently exibited some very nice dualities of the Standard Model, see \cite{tsou}. Hence the invariant, for the flat foliation case (since nl$\sigma $m can be thought of as a flat bundle with structure group the isometries of the target space), maybe of some relevance also for Yang-Mills theories, the setting however will involve loop spaces now! We do not know exactly what its physical significance would be. Moreover doing K-theory on loop spaces (infinite dimensional manifolds) is considerably harder. There are however some path integral techniques. (see again \cite{tsou} and references therein). \emph{2. Instantons with non simply connected boundary}. Following largely the case of instantons we suggest that this invariant is related to interpolation between \emph{gauge inequivalent vacua} which exist due to the non simply connectedness of the space considered. Clearly there is \emph{extra} degeneracy of the vacuum coming from the fact the our space is \textsl{not simply connected}. This degeneracy is of different origin than that of instantons since as it is well-known for ordinary instantons the degeneracy comes from the \textsl{different topologies} of the bundle considered. In more concrete terms we suppose that this invariant will be relevant in the following case: let us assume that we try to follow the discussion in the BPST famous paper on instantons \cite{BPST}; if we assume that we have a space whose boundary is not just $S^3$ as in that case but a 3-manifold which has a non-trivial $\pi _{1}$. In this case we want the potential to become flat (pure gauge) on the boundary. However if the boundary is a 3-manifold with a non trivial fundamental group, then flat connections are not unique (up to gauge equivalence of course). We know more specifically that gauge equivalent classes of flat connections are in 1-1 correspondence with conjugate classes of representations of the fundamental group onto the structure group considered. Thus in this case the flat connection we choose will not be unique. This \emph{extra }degeneracy of the vacuum comes from the different possible choices of flat connection, which is something noticed for the first time. The invariant is related to interpolation between these extra vacua. We expect then some relation with the so called \emph{ALE gravitational instantons} which are important both in quantum gravity and in gauge theory \cite{barrett}, \cite{kronheimer}. \emph{3. Gravity, Non-Commutative Topological Quantum Field Theories (ncTQFT for brevity).} In ordinary Yang-Mills theory, gauge transformations are described as automorphisms of the bundle (namely fibre preserving maps) which \emph{induce the identity on the base space} (cf for example \cite{donald}). Sometimes these are called \emph{strong} bundle automorphisms. If one wants to generalise this picture and attempts to include the symmetry of general relativity, namely local diffeomorphisms of the base space, then there is a problem because there are local diffeomprphisms of the base space which can not be induced by bundle automorphisms (cf \cite{shard}). In simple words: there are "more" local diffeomorphisms of the base space than bundle automorphisms. The way that theoretical physicists usually try to go arround this problem is - to begin with, supersymmetry and finally, - supergravity. The origins of supersymmetry are actually quantum mechanical (multiplets with same number of bosons and fermions plus symmetry between particles of different spin which should exist, if all interactions are eventually unified since gauge particles have spin-1 whereas the graviton is supposed to have spin-2. Another point of view is to examine the largest possible symmetry of the S-matrix elements in the framework of relativistic quantum field theory on Minkowski space). To make the long story short, based on the Coleman-Mandula theorem (which is responsible for introducing anti-commuting coordinates), what one actually does (following the superspace formalism for N=1 supersymmetry coming from observation that Minkowski space is actually Poincare/Lorentz) is to enlarge the base manifold (assumed to be spacetime) by adding some fermionic dimensions (non-commuting coordinates), thus obtaining another space, the socalled superspace. This superspace, in an analogous fashion, can be seen as the quotient space superPoincare/Lorentz. Supersymmetric Yang-Mills theories are principal G-bundles over superspace with G some compact and connected Lie group (usually SU(N)) whereas supergravity can be seen as a principal G-bundle over superspace where G is the superPoincare group (generalisation of Einstein-Cartan theory). However the so-called Noether technique which makes local a rigid supersymmetry and which is used mainly to construct supersymmetric interacting Lagrangians actually suggests that supersymmetric Yang-Mills theories can be equivalently be seen as principal G-bundles over ordinary spacetime with G some Super Lie Group. Letting alone some severe criticism of supersymmetric theories (e.g. positive metric assumption), especially when the discussion comes to supergravity-- (the most important experimental problem of supersymmetric theories is the fact that none of the superparteners of particles has ever been observed, the way phenomenologists try to overcome this problem is to assume spontaneously breaking of supersymmetries; two of the main theoretical problems are: the aspect of supergravity as a local gauge theory which is not completely mathematically justified, for example in N=8 D=4 supergravity theory coming from D=11 N=1 supergravity which is supposed to be the best candidate for unification and according to recent progress one of the two low energy limits of M-theory, local diffeomorphisms are supposed to come from gauging the group O(8), an assumption which is based on the observation that ordinary gravity comes from gauging the Poincare group, something which is wrong because of the existence of the "shouldering" form on arbitrary curved manifolds; the second important problem is that most of the extended supersymmetric and supergravity theories are actually up to now formulated only "on-shell", namely they are essentially classical theories - for N=1 supergravity though there is another problem, one has more than one "off-shell" formulations, this in fact has now an explanation from the recently observed string/5-brane duality in D=10 (old brane-scan); - trying to be fair, we must mention that the good features of such theories are that they offer probably the only up to now known hope for unification plus the fact that they give "less divergent" theories, something essential for perturbative quantum fiels theories),-- we would like to propose here another approach; our approach is more in the spirit of non-perturbative quantum field theories, in fact topological quantum field theories: \emph{instead of enlarging the base manifold by considering anti-commuting coordinates, we chose to relax the fibre preserving condition}, meaning that now we allow "bundle" maps which are not fibre-preserving; in such a case, fibres may be "mixed up", for example they may be "tilted" or "broken". The resulting structure after applying these more general transformations to our original bundle may no longer be a fibre bundle, but it will still be a foliation. In this case however what we will get as the quotient space will not necessarily be the manifold we had originally in our bundle construction (supposed to be space-time) but another space of leaves with the same dimension and maybe very different topology. In this case the dimension of the leaves is kept fixed, equall to the dimension of the Lie algebra considered; had we changed that, the dimension of the space of leaves would have changed accordingly. This picture is quite close to the picture that string theorists patronise, namely that space-time is not fixed but it emerges as a ground state from some dynamical process. In fact there is a deep result due to Thurston which for a given manifold M, say, it relates the group of local diffeomorphisms of M with the group of foliations of M \cite{thurston}. In particular Thurston proves that one has an isomorphism in cohomology after some shift in the degrees between the classifying space of local diffeomorphisms and the classifying space of foliations of a closed manifold M. If we take M to be the total space of a principal bundle P over spacetime, then obviously local diffeomorphisms of the base space are included to local diffeomorphisms of the total space which in turn are very closely related to the group of foliations. Hence at least in principle looking at the group of foliations of the total space of a principal bundle provides a framework which is rich enough in order to encorporate local diffeomorphisms of the base space, something we need in order to relate general relativity with Yang-Mills theory symmetries and this framework is mathematically rigorous. The group then of all foliations of the total space with fixed codimension is huge. It definitely contains all foliations which are "regular" enough in order to get manifolds diffeomorphic to our original one. Yet foliations can be really nasty: in this case the quotient space may not be a manifold at all but a "quantum" topological space. All these cases need to be studied. For the moment we know that whenever the foliations have a corresponding $C^*$ algebra which is SME to a commutative one, then the space of leaves will be a compact Hausdorff topological space of the same dimension. If the $C^*$-algebras of two foliations can be related with a *-preserving homomorphism, then the corresponding quotient spaces will be homoeomorphic. What is the appropriate condition on the $C^*$ algebras in order to get diffeomorphic manifolds, we do not know (this point is of particular interest in 4-dim due to the existence of the so-called "exotic structures" for 4-manifolds). The main point here is that we can "control" how much non-commutativity we want in the $C^*$- algebra and then see what this means topologically. At this point we would like to recall that mathematically, going from classical physics to quantum is going from commuting algebras to non-commuting ones. The essence of Planck's constant then is that it tells us "how much" non-commutativity we want. Moreover there is the fundamental theorem for $C^*$- algebras representations, namely that for each $C^*$- algebra (commutative or not), there exists a Hilbert space whose space of bounded operators is actually "the same" as the original $C^*$- algebra. Hence for each foliation there exists a corresponding $C^*$- algebra (commutative or not), a corresponding topological space (space of leaves which may be a manifold or a quantum topological space respectively) and finally, a Hilbert space as a representation space! Let us now turn to something related to the above but a little more concrete: for the moment let us consider the case where the dimensions of the leaves and of the space of leaves are kept fixed. This situation has some similarities with quantum gravity seen as a TQFT (in fact we generalise that picture and we present a way to consider unified theories-namely gravity and Yang-Mills theories-as Non-Commutative Topological Quantum Field Theories). In \cite{barrett} it was argued that TQFT may provide a framework which is rich enough for the development of a quantum theory of gravity. In that aspect, space-time was treated as an unquantized object whereas the metric was quantum mechanical. The idea in TQFT framework is to find an invariant $Z(M)$ for a topological space $M$ and then one seeks for a Lagrangian density whose partition function yields the invariant $Z(M)$, see \cite{WIT}. One has to be a little more careful though in order for the Atiyah's axioms for TQFT to be satisfied \cite{AT}. In quantum mechanics one usually has a space of quantum states associated to a given system. Often this space of states refers to a particular instant of time, which can be represented in a 4-dimensional world by a space-like hypersurface. In TQFT this vector space appears as part of the definition, when the space-time $M$ has a boundary, i.e. a space of dimension 1 less. In more concrete terms, to a $(d-1)$-dim space $\Sigma $ we associate a vector space $V(\Sigma )$ and to each $d$-dim space $M$ with $\partial M=\Sigma $ we associate a vector $Z(M)$, the partition function of the space $M$. This point can be generalised, in fact $\Sigma $ can be any embedded submanifold of $M$ with dimension 1 less. One interpretation of these conditions is that $\Sigma $ represents the "present instant" of time and that the vectors in $V(\Sigma )$ which are determined by various choices of observables represent a memory of past facts. The primary problem nonetheless is the construction of invariants for spaces and the state spaces and partition functions for spaces with boundary are usually obtained as a by-product. Since by our proposal above one can end up with quantum topological spaces as spaces of leaves of foliations, one can call this theory \emph{non-commutative} topological quantum field theory and we believe that this can provide a framework for quantum unified theories (including Yang-Mills and gravity). The picture we have then is the following: we start with a G-bundle $P$ over a 4-manifold $M$. From symmetry considerations, namely we want to include local diffeomorphisms of the base space and relate them to bundle automorphisms (hence relating general relativity to gauge theory), we end up to consider all dimG-dim foliations of $P$. Automatically the 4-dim space which is the space of leaves is somehow "quantized", namely it is forced to have one of the leaf topologies. This is a difference with TQFT as explained in \cite{barrett} where the metric was quantised but space-time was unquantized (needless to say, in such a case as ours, the metric is quantised too automatically). Moreover, for each foliation we have a quotient space of leaves and hence an invariant $Z(M)$ which is a complex number. The boundary of course can be added with its vector space attached to it, one however has to examine what happens to it as foliations vary. The Lagrangian density whose partition function is this invariant for foliations is an open question. It should be related to characteristic classes for foliations. A good indication for that is the fact that the \emph{$\Gamma _q$ functor} of q-dim Haefliger structures or $\Gamma _q $ structures as they are known in topology (and hence foliations which is an example of a $\Gamma _q$ structure, where $q$ is the codimension of the foliation) \emph{is representable}, see \cite{james}. So for the moment we do not have a "full" specific ncTQFT since we have the invariant but not the appropriate Lagrangian density. We presented though a generalisation of TQFT for non-commutative cases. Another possible application might be the ability to construct \emph{deformed} Yang-Mills theories, see \cite{schwarz}. In that paper, some new compactifications of the IKKT matrix theory on non-commutative tori were introduced which, in a certain sense, could be realised as deformed Yang-Mills theories. Clearly in this case our invariant will be the "instanton number" of these deformed Yang-Mills theories. This picture also suggests that the above described non-commutative topological quantum field theories can be seen as emerging from M-theory compactified down to some non-commutative spaces (tori or other). \section{M-Theory} In this section we shall present an application to M-Theory. Since it is more extensive, we give it separately. We know that M-Theory consists of membranes and 5-branes living on an 11-manifold (\cite{duff}, \cite{west}) and it is non perturbative. This theory has a very intriguing feature: we can only extract information about it from its limiting theories, namely either from D=11 N=1 supergravity or from superstrings in D=10. This is so because this theory is \emph{genuinly} non perturbative for a reason which lies in the heart of manifold topology: Let us recall that in string theory, the path integral involves summation over all topologically distinct diagrams (same for point particles of course). Strings are 1-branes hence in time they swep out a 2-manifold. At the tree level then we need all topologically distinct simply connected 2-manifolds (actually there is only one, as topology tells us) and for loop corrections, topology again says that topologically distinct non simply connected 2-manifolds are classified by their genus, so we sum up over all Riemann surfaces with different genus. It is clear then that for a perturbative quantum field theory involving p-branes we have to sum upon all topologically distinct (p+1)-dim manifolds: simply connected ones for tree level and non simply connected ones for loop corrections. Thus \emph{we must know before hand the topological classification of manifolds} in the dimension of interest. That is the main problem of manifold topology in mathematics. But now we face a deep and intractable problem: geometry tells us, essentially via a no-go theorem which is due to Whitehead from late '40's, that: \emph{"we cannot classify non simply connected manifolds with dimension greater or equal to 4"}! Hence for p-branes with p greater or equal to 3, all we can do via perturbative methods is up to tree level! What happens for 3-manifolds then (hence for membranes)? The answer from mathematics is that we \emph{do not know} if all 3-manifolds can be classified! So even for 2-branes it is still unclear whether perturbative methods work (up to all levels of perturbation theory)!\\ The outlet from this situation that we propose here is not merely to look only at non perturbative aspects of these theories (i.e. the soliton part of the theory) and then apply S-duality, as was done up to now, but to abandon perturbative methods completely from the very beginning. There is only one way known up to now which can achieve this "radical" solution to our problem: \emph{formulate the theory as a Topological Quantum Field Theory} and hence get rid of all perturbations once and for all.\\ Let us explain how this can be acieved.\\ Our approach is based on one physical \emph{"principle"}:\\ \textsl{A theory containing \emph{p-branes} should be formulated on an m-dim manifold which \emph{admits $\Gamma _q$-structures}, where $q=m-p-1$}.\\ {\bf N.B.} Although we used in our physical principle $\Gamma _q$-structures which are more general than foliations, we shall use both these terms meaning essentially the same structure. The interested reader may refer to \cite{lawson} for example to see the precise definitions which are quite complicated. The key point however is that the difference between $\Gamma _q$-structures (or Haefliger structures as they are most commonly known in topology) and codim-q foliations is essentially the difference between \textsl{transverse} and \emph{normal}. This does not affect any of what we have to say, since Bott-Haefliger theory of characteristic classes is formulated for the most general case, namely $\Gamma $-structures. We would also like to mention the relation between $\Gamma $-structures and $\Omega $-spectra which is currently an active field in topology.\\ (For D-branes we need a variant of the above principle, namely we need what are called \emph{plane foliations} but we shall not elaborate on this point here).\\ One way of thinking about this principle is that it is analogous to the ``past histories'' approach of quantum mechanics. Clearly in quantum level one should integrate over all foliations of a given codim.\\ \textsl{A piece of warning here:} this principle does not imply that \emph{all} physical process between branes \emph{are} described by foliations. Although the group of foliations is huge, in fact comparable in size with the group of local diffeomorphisms \cite{thurston}, and foliations can be really ``very nasty'', we would not like to make such a strong statement. What is definitely true though is that \emph{some} physical process \emph{are indeed} described by foliations, hence \textsl{at least} this condition \emph{must} be satisfied because of them.\\ {\bf Note:} Before going on furter, we would like to make one crucial remark: this principle puts severe restrictions on the topology that the underlying manifold may have, in case of M-Theory this is an 11-manifold. It is also very important if the manifold is \emph{open} or \emph{closed}. This may be of some help, as we hope, for the compactification problem of string theory or even M-Theory, namely how we go from D=10 (or D=11) to D=4 which is our intuitive dimension of spacetime. We shall address this question in the next section. The final comment is this: this principle puts \emph{absolutely no restriction} to the usual quantum field theory for point particles in D=4, e.g. electroweak theory or QCD. This is so because in this case spacetime is just ${\bf R^4}$ which is non compact and we have 0-branes (point particles) and consequently 1-dim foliations for which the integrability condition is trivially satisfied (essentially this is due to a deep result of Gromov for foliations on open manifolds, which states that all open manifolds admit codim 1 foliations; in striking contrast, closed manifolds admit codim 1 foliations iff their Euler characteristic is zero, see for example in \cite{lawson}, \cite{gromov} or references therein).\\ If we believe this principle, then the story goes on as follows: we are on an 11-manifold, call it M for brevity and we want to describe a theory containing 5-branes for example (and get membranes from S-duality). Then M should admit 6-dim foliations or equivalently codim 5 foliations. We know from Haefliger that the $\Gamma _q $-functor, namely the functor of codim q Haefliger structures and in particular codim q foliations, is representable. Practically this means that we can have an analogue of Chern-Weil theory which characterises foliations of M up to homotopy using cohomology classes of M. (One brief comment for foliations: one way of describing Haefliger structures more generally is to say that they \emph{generalise fibre bundles in exactly the same way that fibre bundles generalise Cartesian product}. This observation is also important when mentioning \textsl{gerbes} later on).\\ In fact it is proved that the correct cohomology to classify Haefliger structures up to homotopy (and hence foliations which constitute a particular example of Haefliger structures) is the \emph{Gelfand-Fuchs} cohomology. This is a result of Bott and Haefliger, essentially generalising an earlier result due to Godbillon and Vey which was dealing only with codim 1 foliations, \cite{gv}.\\ Now we have a happy coincidence: the Bott-Haefliger class for a codim 5 foliation (which, recall, is what we want for 5-branes on an 11-manifold) is exactly an 11-form, something that fits well with using it as a Lagrangian density! The construction for arbitrary codim q foliations goes as follows: let $F$ be a codim q foliation on an m-manifold M and suppose its normal bundle $\nu (F)$ is orientable. Then $F$ is defined by a global decomposable q-form $\Omega $. Let $\{(U_{i},X_i)\}_{i\in I}$ be a locally finite cover of distinguished coordinate charts on M with a smooth partition of unity $\{\rho _i\}$. Then set $$ \Omega = \sum _{i\in I} \rho _{i}dx_{i}^{m-q+1}\wedge ... \wedge dx_{i}^{m} $$ Since $\Omega $ is integrable, \begin{equation} d\Omega =\theta \wedge \Omega , \label{eq1} \end{equation} where $\theta $ \emph{some} 1-form on M. The (2q+1)-form \begin{equation} \gamma = \theta \wedge (d\theta )^{q}, \label{eq2} \end{equation} is closed and its de Rham cohomology class is independent of all choices involved in defining it, depending only on homotopy type of $F$. That's the class we want. Clearly for our case we are on an 11-manifold dealing with 5-branes, hence 6-dim foliations, hence codim 5 and thus the class $\gamma $ is an 11-form. This construction can be generalised to arbitrary $\Gamma ^{r}_{q}$-structures as a mixed de Rham-Cech cohomology class and thus gives an element in $H^{2q+1}(B\Gamma ^{r}_{q};{\bf R})$, where $B\Gamma ^{r}_{q}$ is the classifying space for $\Gamma ^{r}_{q}$-structures. Note that in fact the BHGV class is a cobordism invariant of codim q foliations of compact (2q+1)-dim manifolds. This construction gives \emph{one} computable characteristic class for foliations. Optimally we would like a generalisation of the Chern-Weil construction for $GL_q$. That is we would like an abstract GDA with the property that for any codim q foliation $F$ on a manifold M there is a GDA homomorphism into the de Rham algebra on M, defined in terms of $F$ such that the induced map on cohomology factors through a universal map into $H^*(B\Gamma _{q}^{r};{\bf R})$. This algebra is nothing more than the \emph{Gelfand-Fuchs} Lie coalgebra of \emph{formal} vector fields in one variable. More concretely, let $\Gamma $ be a transitive Lie-pseudogroup acting on ${\bf R^n}$ and let $a(\Gamma )$ denote the \textsl{Lie algebra of formal $\Gamma $ vector fields} associated to $\Gamma $. Here a vector field defined on on $U\subset {\bf R^n}$ is called a $\Gamma $ vector field if the local 1-parameter group which it engenders is $\Gamma $ and $a(\Gamma )$ is defined as the inverse limit $$a(\Gamma )=lim_{\leftarrow}a^{k}(\Gamma )$$ of the k-jets at 0 of $\Gamma $ vector fields. In the pseudogroup $\Gamma $ let $\Gamma _0 $ be the set of elements of $\Gamma $ keeping 0 fixed and set $\Gamma _{0}^{k} $ equal to the k-jets of elements in $\Gamma _0 $.\\ Then the $\Gamma _{0}^{k} $ form an inverse system of Lie groups and we can find a subgroup $K\subset lim_{\leftarrow}\Gamma _{0}^{k}$ whose projection on every $\Gamma _{0}^{k} $ is a maximal compact subgroup for $k>0$. This follows from the fact that the kernel of the projection $\Gamma _{0}^{k+1}\rightarrow \Gamma _{0}^{k}$ is a vector space for $k>0$. The subgroup K is unique up to conjugation and its Lie algebra k can be identified with a subalgebra of $a(\Gamma )$. For our purposes we need the cohomology of basic elements rel K in $a(\Gamma )$, namely $H(a(\Gamma );K)$ which is defined as follows: Let $A\{a^k(\Gamma )\}$ denote the algebra of multilinear alternating forms on $a^k(\Gamma )$ and let $A\{a(\Gamma )\}$ be the direct limit of the $A\{a^k(\Gamma )\}$. The bracket in $a(\Gamma )$ induces a differential on $A\{a(\Gamma )\}$ and we write $H\{a(\Gamma )\}$ for the resulting cohomology group. The relative group $H^{*}(a(\Gamma );K)$ is now defined as the cohomology of the subcomplex of $A\{a(\Gamma )\}$ consisting of elements which are invariant under the natural action of K and annihilated by all inner products with elements of k. Then the result is: \textsl{Let $F$ be a $\Gamma $-foliation on M. There is an algebra homomorphism} $$\phi :H\{a(\Gamma );K\}\rightarrow H(M;{\bf R})$$ which is a natural transformation on the category $C(\Gamma )$. The construction of $\phi $ is as follows: Let $P^k(\Gamma )$ be the differential bundle of k-jets at the origin of elements of $\Gamma $. It is a principal $\Gamma _{0}^{k}$-bundle. On the other hand $\Gamma $ acts transitively on the left on $P^k(\Gamma )$. Denote by $A(P^{\infty}(\Gamma ))$ the direct limit of the algebras $A(P^{k}(\Gamma ))$ of differential forms on $P^k(\Gamma )$. The invariant forms wrt the action of $\Gamma $ constitute a differential subalgebra denoted $A_{\Gamma }$. One can then prove that it is actually isomorphic to $A(a(\Gamma ))$. Now let F be a foliation on M and let $P^k(F)$ be the differentiable bundle over M whose fibre at every point say $x\in M$ is the space of k-jets at this point of local projections that vanish on $x$. This is a $\Gamma _0^k $-principal bundle. Its restriction is isomorphic to the inverse image of the bundle $P^k(\Gamma )$, hence the differential algebra of $\Gamma $-invariant forms on $P^k(\Gamma )$ is mapped in the algebra $A(P^k(F))$ of differential forms on $P^k(F)$. If we denote by $A(P^{\infty}(F))$ the direct limit of $A(P^{k}(F))$ we get an injective homomorphism $\phi $ of $A(a(\Gamma ))$ in $A(P^{\infty}(F ))$ commuting with the differential. This homomorphism is compatible with the action of K, hence induces a homomorphism on the subalgebra of K-basic elements. But the algebra $A(P^{k}(F );K)$ of K-basic elements in $A(P^{k}(F))$ is isomorphic to the algebra of differential forms on $P^k(F)/K$ which is a bundle over M with contractible fibre $\Gamma _0^k/K$.Hence $H(A(P^k(F);K))$ is isomorphic via the de Rham theorem to $H(M;{\bf R})$. The homomorphism $\phi $ is therefore obtained as the composition $$H(a(\Gamma );K)\rightarrow H(A(P^{\infty }(F);K))=H(M;{\bf R})$$ But we think that is enough with \emph{abstract nonsense} formalism. Let us make our discussion more \textsl{down to earth}: Consider the GDA (over {\bf R}) $$WO_{q}=\wedge (u_1, u_3,..., u_{2(q/2)-1})\otimes P_q(c_1,..., c_q)$$ with $du_i=c_i $ for odd i and $dc_{i}=0$ for all i and $$W_{q}=\wedge (u_{1},u_{2},...,u_{q})\otimes P_{q}(c_1,...,c_q)$$ with $du_{i}=c_{i}$ and $dc_{i}=0$ for i=1,...,q where $deg u_{i}=2i-1$, $deg c_{i}=2i$ and $\wedge $ denotes exterior algebra, $P_q $ denotes the polynomial algebra in the $c_{i}$'s mod elements of total degree greater than 2q. The cohomology of $W_q $ is the Gelfand Fuchs cohomology of the Lie algebra of formal vector fields in q variables. We note that the ring structure at the cohomology level is trivial, that is all cup products are zero. Then the main result is that there are homomorphisms $$\phi :H^{*}(WO_{q})\rightarrow H^{*}(B\Gamma ^{r}_{q};{\bf R})$$ $$\tilde {\phi }:H^{*}(W_q)\rightarrow H^{*}(\tilde{B\Gamma _{q}^{r}};{\bf R})$$ for $r \geq 2$ with the following property ($\tilde{B\Gamma _{q}^{r}}$ denotes the classifying space for \emph{framed} foliations): If $F$ is a codim q $C^{r}$ foliation of a manifold M, there is a GDA homomorphism $$\phi _{F}:WO_{q}\rightarrow \wedge ^{*}(M)$$ into the de Rham algebra on M, defined in terms of the differential geometry of $F$ and unique up to chain homotopy, such that on cohomology we have $\phi _{F}=f^{*}\circ \phi $, where $f:M \rightarrow B\Gamma ^{r}_{q}$ classifies $F$. If the normal bundle of $F$ is trivial, there is a homomorphism $$\tilde {\phi _{F}}:W_q \rightarrow \wedge ^{*}(M)$$ with analogous properties.\\ Combining this result with the fact that $B\tilde{\Gamma ^{0}_{q}}$ is \emph{contractible}, we deduce that a foliation is essentially determined by the structure of its normal bundle; the \emph{Chern} classes of the normal bundle are contained in the image of the map $\phi $ above but we have \emph{additional} non trivial classes in the case of foliations (which are rather difficult to find though), one of which is this BHGV class which we constructed explicitly and it is the class we use as a Lagrangian density which is purely topological since its degree fits nicely for describing 5-branes. There is an alternative approach due to Simons \cite{simons} which avoids passing to the normal bundle using circle coefficients. What he actually does is to associate to a principal bundle with connection a family of characteristic homomorphisms from the integral cycles on a manifold to $S^{1}$ and then defining an extension denoted $K^{2k}_{q}$ of $H^{2k}(BGL_{q};{\bf Z})$. This approach is related to \emph{gerbes}. A gerbe over a manifold is a construction which locally looks like the Cartesian product of the manifold with a line bundle. Clearly it is a special case of foliations (remember our previous comment on foliations). However this approach actually suggests that they might be equivalent, if the approach of Bott-Haefliger is equivalent to that of Simons, something which is not known.\\ \textsl{Now the conjecture is that the \emph{partition function} of this Lagrangian is related to the invariant introduced in \cite{Z2}.}\\ In order to establish relation with physics, we must make some identifications. The 1-form $\theta $ appearing in the Lagrangian has no direct physical meaning. In physics it is assumed that a 5-brane gives rise to a 6-form gauge field denoted $A_6$ whose field strength is simply \begin{equation} dA_{6}=F_{7} \label{eq3} \end{equation} The only way we can explain geometrically this is that this 6-form is the Poincare dual of the 6-chain that the 5-brane sweps out as it moves in time. We know that since we have S-duality between membranes and 5-branes, in an obvious notation one has \begin{equation} F_{7}=*F_{4} \label{eq4} \end{equation} which is the S-duality relation, where \begin{equation} F_{4}=dA_{3} \label{eq5} \end{equation} Observe now that the starting point for 5-brane theory is $A_6$ where the starting point to construct the BHGV class was the 5-form $\Omega $. How are they related? There are three obvious possibilities: I. $d\Omega =A_6$ That would imply that $A_6 $ is pure gauge. II.$dF_{4}=\Omega $ This is trivial because it implies $d\Omega =0$, hence $d\Omega =\theta \wedge \Omega =0$. III. The only remaining possibility is \begin{equation} *A_{6}=\Omega \label{eq6} \end{equation} We call this \emph{``reality condition''}. So now in principle we can substitute equations (10) and (5) into (6) and get an expression for the Lagrangian which involves the gauge field $A_6$. The Euler-Lagrange equations which are actually analogous to D=11 N=1 supergravity Euler-Lagrange equations (see equation (12) below) read: \begin{equation} d*d\theta + \frac{1}{5}(d\theta )^{5}=0 \label{eq7} \end{equation} The on-shell relation with D=11 N=1 supergravity is established as follows: recall that the bosonic sector of this supergravity theory is $$\int F_{4}\wedge F_{4}\wedge A_{3}$$ where $F_{4}=dA_{3}$ with Euler-Lagrange equations \begin{equation} d*F_{4}+\frac{1}{2}F_{4}\wedge F_{4}=0 \label{eq8} \end{equation} Constraining $A_3 $ via (12), by (9), (8), (7), (10) and (5) we get a constraint for $\theta $ which can be added to the class $\gamma $ as a Lagrange multiplier. In order to calculate the partition function, some additional difficulties may arise because we do not know what notion of equivallence between foliations is the appropriate one for physics in order to fix the gauge and add Faddeev-Popov terms as constraints to kill-off the gauge freedom. There are actually four different notions of equivalence for foliations: conjugation, homotopy, integrable homotopy and foliated cobordism. In principle, one must end up with an equivallent theory starting with membranes (that's due to S-duality), provided of course a suitable class was found. Clearly the BHGV class for a membrane would be a \emph{17-form}. The final comment refers to \cite{sus}. In that article it was conjectured that the quantum mechanics of branes could be described as a matrix model. As it is well-known matrix models use point particle degrees of freedom. This is rather intriguing since we are talking about M-theory which contains various p-branes. In our approach though we propose a Lagrangian density which has as fundamental object a mysterious 1-form which, if seen as a gauge potential, that would imply the existence of some yet unknown underlying point particle! \subsection{Plane fields} We now pass on to the second question raised in this application, namely the restrictions on the topology of the underlying manifold of a theory containing p-branes via our physical principle. It is clear from the definition that the existence of a foliation of certain dim, say d (or equivalently codim q=n-d) on an n-manifold (closed) depends: a.) On the existence of a dim d subbundle of the tangent bundle b.) On this d-dim subbundle being integrable. The second question has been answered almost completely by Bott and in a more general framework by Thurston. Bott's result dictates that for a codim q subbundle of the tangent bundle to be integrable, the ring of Pontrjagin classes of the subbundle with degree $>2q$ must be zero. There is a secondary obstruction due to Shulman involving certain Massey triple products but we shall not elaborate on this. However Bott's result suggests nothing for question a.) above. Let us also mention that this result of Bott can be deduced by another theorem due to Thurston which states that the classifying space $B\tilde{\Gamma ^{\infty }_{q}}$ of smooth codim q framed foliations is (q+1)-connected. On the contrary, Thurston's result reduces the existence of codim $q>1$ foliations (at least up to homotopy) to the existence of \emph{q-plane fields}. This is a deep question in differential topology, related to the problem of classification of closed manifolds according to their \textsl{rank}. Now the problem of existence of q-plane fields has been answered \emph{only for some cases for spheres $S^n$ for various values of n,q} \cite{steenrod}. In particular we know everything for spheres of dimension 10 and less. We should however mention a theorem due to Winkelnkemper \cite{win1} which is quite general in nature and talks about simply connected compact manifolds of dim n greater than 5. If n is not 0 mod 4 then it admits a so-called \emph{Alexander decomposition} which under special assumptions can give a particular kind of a codim 1 foliation with $S^{1}$ as space of leaves and a surjection from the manifold to $S{1}$. If n is 0 mod 4 then the manifold admits an Alexander decomposition iff its signature is zero. Let us return to string theory now: String theory works in D=10 and in this case we have the old brane-scan suggesting the string/5-brane duality. The new brane-scan contains all p-branes for $p\leq 6$ and some D-7 and 8-branes are thought to exist. However topology says that for a sphere in dim 10 we can have only dim 0 and dim 10 plane fields (in fact this is true for all even dim spheres), hence by Thurston only dim 0 and dim 10 foliations and then our physical principle suggests that $S^{10}$ is ruled out as a possible underlying topological space for string theory.\\ What about M-Theory in D=11 then?\\ For the case of $S^{11}$ then it is known that $S^{11}$ admits a 3-plane field, hence by our physical principle a theory containing membranes \emph{can} be formulated on $S^{11}$. For $S^{11}$ nothing is known for the existence of q-plane fields for q greater than 3. But now we apply S-duality between membranes/5-branes and conjecture that:\\ \emph{$S^{11}$ should admit 5-plane fields}.\\ Let us close with two final remarks: 1. There is extensive work in foliations with numerous results which actually insert many extra parameters into their study, for example metric aspects, existence of foliations with compact leaves (all or at least one or exactly one), with leaves diffeomorphic to ${\bf R^n}$ for some n etc. We do not have a clear picture for the moment concerning imposing these in physics. Let us only mention one particularly strong result due to Wall generalising a result of Reeb \cite{reeb}: if a closed n-manifold admits a codim 1 foliation whose leaves are homeomorphic to ${\bf R^{n-1}}$, then by Thurston we know that its Euler characteristic must vanish, but in fact we have more: it has to be the n-torus!\\ The interesting point however is that although all these extended objects theories in physics are expressed as $\sigma $ models \cite{Z1}, hence they involve metrics on the manifold (target space) and on the worldvolumes ie on the leaves, in our approach the metric is only used in the reality condition (10) which makes connection with physical fields (that is some metric on the target space) where at the same time we do not use any metric on the source space (worldvolumes-leaves of the foliation).\\ 2. In \cite{Z1} another Lagrangian density was proposed. It is different from the one described here but they are related in an analogous way to the relation between the Polyakov and Nambu-Goto (in fact Dirac \cite{dirac}) actions for the free bosonic string: extended objects basically immitate string theory and we have two formalisms: the $\sigma $ model one which is the Lagrangian exhibited in \cite{Z1} using Polyakov's picture of $\sigma $ models as flat principal bundles with structure group the isometries of the metric on the target space \cite{polyakov}; yet we also have the \emph{embedded surface} picture which is the Dirac (Nambu-Goto) action and whose analogue is described in this work. In the light of a very recent work \cite{WIT}, we can also make some further comments: the first is that the Moyal algebra used in order to discuss noncommutative solitons is actually Morita equivalent to the usual commutative one. This fact can be further verified from the explicit construction of an algebra homomorphism between noncommutative and ordinary Yang-Mills fields based on Gelfand's theorem. Truly noncommutative situations appear when discussing noncommutative tori. The next comment refers to the last section of that paper: we already know that strings in a constant magnetic field can be described from that Moyal algebra-like spacetime structure and they also discuss what may happen to 5-branes in M-theory. Our point of view coincides with theirs in the following way: we here propose that this is indeed the case for 5-branes with a C-field turned on, namely that this situation can be described by 6-dim foliations which rather correspond to a "free" theory of branes but on a "noncommutative" topological space, which is actually, in our case, the space of leaves of the corresponding foliation. \begin {thebibliography}{50} \bibitem{bott}R. Bott and A. Haefliger: "Characteristic classes of $\Gamma $-foliations", Bull. Am. Math. Soc. 78.6, (1972)\\ \bibitem{lawson}H. B. Lawson: "Foliations", Bull. Am. Math. Soc. 80.3, 1974\\ \bibitem{thurston}W. Thurston: "Theory of foliations of codim greater than 1", Comment. Math. Helvetici 49 (1974), 214-231\\ W. Thurston: ``Foliations and groups of diffeomorphisms'', Bull. Am. Math. Soc. 80.2 (1974)\\ \bibitem{gromov}M. L. Gromov: ``Stable mappings of foliations into manifolds'', Izv. Akad. Nauk. USSR Ser. Mat. 33 (1969)\\ \bibitem{gv}C. Godbillon and J. Vey: ``Un invariant des feuilletages de codim 1'', C R Acad. Sci. Paris Ser AB 273 (1971)\\ \bibitem{steenrod}N. Steenrod: ``The topology of fibre bundles'', Princeton 1951\\ \bibitem{win1}H. E. Winkelnkemper: ``Manifolds as open books'', Bull. Am. Math. Soc. 79 (1973)\\ \bibitem{reeb}G. Reeb: ``Feuillages, resultats anciens et nouveaux'', Montreal 1982\\ \bibitem{simons}J. Simons: ``Characteristic forms and transgression'', preprint SUNY Stony Brook\\ \bibitem{dirac}P.A.M. Dirac: Proc. Roy. Soc. London A166 (1969)\\ \bibitem{duff}M. J. Duff et all: ``String Solitons'', Phys. Rep. 259 (1995) 213 M. J. Duff: ``Supermembranes'', hep-th 9611203\\ \bibitem{west}P. C. West: ``Supergravity, brane dynamics and string dualities'', hep-th 9811101\\ \bibitem{james} R. Bott:"Lectures on characteristic classes and foliations, Springer LNM 279, 1972.\\ \bibitem{AT} M. F. Atiyah: "Topological quantum field theories", Publ. Math. IHES 68, 175-186, (1989). "The geometry and the physics of knots", Cambridge University Press, 1990. \\ \bibitem{wegge} Wegge-Olsen: "K-Theory and $C^*$-algebras", Oxford University Press, 1992.\\ \bibitem{shard} G. Shardanashvili,O. Zakharov: "Gauge Gravitation Theory", World Scientific, 1991.\\ \bibitem{hirzebruch} Hirzebruch: "Geometrical methods in Topology", Springer, Berlin 1970.\\ \bibitem{polyakov} A. Polyakov, Nucl. Phys. B164 (1981), Phys. Lett. 82B (1979), hep-th/9607049 Princeton preprint 1996.\\ \bibitem{hirsch} G. Hector - U. Hirsch: "Introduction to the Geometry of Foliations", Vieweg 1981.\\ \bibitem{connes1} A. Connes: "A survey of foliations and operator algebras", Proc. Sympos. Pure Math., 38 AMS 1982.\\ \bibitem{baum}P. Baum, R. Douglas: "K-homology and Index theory", ibid.\\ \bibitem{connes2} A. Connes: "Non-Commutative Geometry", Academic Press 1994.\\ \bibitem{connes3} A. Connes: "Noncommutative differential Geometry I, II", IHES Publ. Math. 63 (1985).\\ \bibitem{connes5}A. Connes: ``Cyclic cohomology and the transverse fundamental class of a foliation'', Pitman Res. Notes in Math., 123, Longman Harlow 1986.\\ \bibitem{townsend} P. Townsend, D. Freedman, Nucl. Phys. B177 (1981) 282.\\ \bibitem{donald} S. K. Donaldson, P.B. Kronheimer, "The geometry of 4-manifolds", Oxford University Press, 1991.\\ \bibitem{conness} A. Connes, G. Skandalis, "The longitudinal index theorem for foliations", Publ. Res. Inst. Sci. Kyoto 20 (1984).\\ \bibitem{atiyah} M.F. Atiyah - V.K. Patodi - I.M. Singer: "Spectral assymetry and Riemannian geometry I, II, III", Math. Proc. Cambridge Philos. Soc. 77 (1975).\\ \bibitem{win} Winkelnkemper. The graph of a foliation. Ann. Global Anal. and Geom. 1 No3 (1983) 51.\\ \bibitem{jcw} G.W. Whitehead. Elements of Homotopy Theory. Springer Berlin 1978.\\ \bibitem{douglas1} R. Douglas - S. Harder - J. Kaminker: "Toeplitz operators and the eta invariant: the case of $S^1$" , Contemp. Math. 70 (1988).\\ \bibitem{douglas2} R. Douglas - S. Harder - J. Kaminker: "Cyclic cocycles, renormalisation and eta invariant", Invent. Math 103 (1991).\\ \bibitem{witten2} E. Witten: "Global Gravitation Anomalies", Commun. Math. Phys. 100 (1985).\\ \bibitem{WIT} E. Witten: "Quantum Field Theories and the Jones polynomial", Commun. Math. Phys. 121 (1989). "Topology-changing amplitudes in (2+1)-dim gravity", Nucl. Phys. B323, 113-140, (1989). "(2+1)-dim gravity as an exactly soluble system", Nucl. Phys. B311, 46-78, (1988). "String Theory and Noncommutative Geometry", hep-th/9908142.\\ \bibitem{tondeur} F. W. Kamber - P. Tondeur: "Foliated Bundles and Characteristic Classes", LNM 493, Springer 1975.\\ \bibitem{BPST} A.A. Belavin, A.M. Polyakov, A.S. Schwartz, Y.S. Tyupkin. Pseudoparticle Solutions of the Yang-Mills Equations. Phys. Lett. 59B (1975).\\ \bibitem{Z1} I.P. Zois: "Search for the M-theory Lagrangian". Phys. Lett. B 402 (1997) 33-35.\\ \bibitem{Z2} I.P. Zois: "The duality between two-index potentials and the non-linear $\sigma $ model in field theory", D.Phil Thesis, Oxford, Michaelmas 1996.\\ \bibitem{barrett} J.W. Barrett. Quantum Gravity as a Topological Quantum Field Theory. Nottingham Mathematics Preprint 1995.\\ \bibitem{kronheimer} P.B. Kronheimer, H. Nakajima. Yang-Mills instantons on ALE gravitational instantons. Mathematische Annalen 288 (1990) 263.\\ \bibitem{henneaux} M. Henneaux. Uniqueness of the Freedman-Townsend interaction vertex for 2-form gauge fields. Preprint Universite Libre de Bruxelles 1996.\\ \bibitem{loday} J.-L. Loday: "Cyclic Homology", Springer, Berlin, 1991.\\ \bibitem{schwarz} A. Connes, M.R. Douglas and A. Schwarz: "Noncommutative Geometry and Matrix Theory: Compactification on Tori", J. High Energy Phys. 02 (1998) 003.\\ \bibitem{tsou} S. T. Tsou et all: "Features of Quark and Lepton Mixing from Differential Geometry of Curves on Surfaces", Phys. Rev. D58 (1998) 053006.\\ S. T. Tsou et all: "The Dualised Standard Model and its Applications", talk given at the International Conference of High Energy Physics 1998 (ICHEP 98), Vancouver, Canada.\\ \bibitem{sus}T. Banks et all: "M-Theory as a Matrix Model: A Conjecture", Phys. Rev. D55 (1997) 5112.\\ \end{thebibliography} \end{document}
\section{Nonabelian Duality and Basic Ingredients of DSM} To illustrate how nonabelian duality enters in the Standard Model, we begin with a short resum\'e of (abelian) duality in electromagnetism. The Maxwell theory has long been known to possess a symmetry under the interchange of electricity and magnetism: \begin{equation} F_{\mu\nu}(x) \longleftrightarrow {}^*\!F_{\mu\nu}(x), \label{dualsymm} \end{equation} where the *-operation (Hodge star): \begin{equation} {}^*\!F_{\mu\nu}(x) = -\frac{1}{2} \epsilon_{\mu\nu\rho\sigma} F^{\rho\sigma}(x), \label{dualtransf} \end{equation} we may call the abelian dual transform. Electric charges are sources of the field $F$ but monopoles of the field ${}^*\!F$, while magnetic charges are monopoles of $F$ but sources of ${}^*\!F$, where the strength of the quantized electric charges $e$ is related to that of the also quantized magnetic charges $\tilde{e}$ by the famous Dirac condition: \begin{equation} e \tilde{e} = 2 \pi. \label{diraccond} \end{equation} At any point $x$ in space-time free (locally) of electric and magnetic charges, both $F_{\mu\nu}(x)$ and its dual ${}^*\!F_{\mu\nu}(x)$ are, by virtue of the Maxwell equations, `gauge fields' derivable from potentials, thus: \begin{equation} F_{\mu\nu}(x) = \partial_\nu A_\mu(x) - \partial_\mu A_\nu(x), \label{fina} \end{equation} \begin{equation} {}^*\!F_{\mu\nu}(x) = \partial_\nu \tilde{A}_\mu(x) - \partial_\mu \tilde{A}_\nu(x). \label{dftinda} \end{equation} It follows then that the theory is invariant under the 2 independent gauge transformations: \begin{equation} A_\mu(x) \longrightarrow A_\mu(x) + \partial_\mu \alpha(x), \label{gaugetransf} \end{equation} \begin{equation} \tilde{A}_\mu(x) \longrightarrow \tilde{A}_\mu(x) + \partial_\mu \tilde{\alpha}(x). \label{dgaugetransf} \end{equation} In other words, the theory is invariant under the doubled gauge symmetry $U(1) \times \tilde{U}(1)$, although obviously, given (\ref{dualtransf}), the dual fields $F$ and ${}^*\!F$ represent just the same physical degrees of freedom. Notice that this `doubled' gauge symmetry is inherent in the Maxwell theory itself and not an additional assertion imposed from outside. The only reason we are less familiar with, and have not made much used of, the dual gauge symmetry $\tilde{U}(1)$, which theoretically is on an equal footing with the other gauge symmetry $U(1)$, is just that in the physical world we have observed electric charges but not so far (apparently) their magnetic counterparts. Nonabelian Yang-Mills theory is not symmetric under the *-operation of (\ref{dualtransf}) \cite{Guyang} so that it was not known for some time whether the dual symmetry of electromagnetism generalizes to Yang-Mills theory. But it has now been shown in \cite{dualsymm} that there is a generalized nonabelian dual transform \ $\tilde{\ }$\ , reducing to * in the abelian case, under which Yang-Mills theory is symmetric. Unfortunately, the explicit formula for the generalized transform\ $\tilde{\ }$\ is known at present only in the language of loop space \cite{Polyakov,book} and is for that reason a little cumbersome. However, for our present discussion, we need only to note the consequences of the symmetry it implies, as follows. Dual to a Yang-Mills field $F_{\mu\nu}$ derivable from a potential $A_\mu$: \begin{equation} F_{\mu\nu}(x) = \partial_\nu A_\mu(x) -\partial_\mu A_\nu(x) + ig\, [A_\mu(x), A_\nu(x)], \label{FinA} \end{equation} there is a field $\tilde{F}_{\mu\nu}$ (the relation of which to $F_{\mu\nu}$ is known though complicated) which is also derivable from a potential, $\tilde{A}_\mu$: \begin{equation} \tilde{F}_{\mu\nu}(x) = \partial_\nu \tilde{A}_\mu(x) - \partial_\mu \tilde{A}_\nu(x) + i \tilde{g}\, [\tilde{A}_\mu(x), \tilde{A}_\nu(x)], \label{dFindA} \end{equation} where the coupling strengths $g$ and $\tilde{g}$ are related by a generalized Dirac condition \cite{dualcomm}: \begin{equation} g \tilde{g} = 4 \pi. \label{Diraccond} \end{equation} `Colour' (electric) charges appear as sources of the field $F$ but monopoles of the dual field $\tilde{F}$, while `dual colour' (magnetic) charges appear as monopoles of $F$ but sources of $\tilde{F}$ \cite{dualsymm,dualsym}. The symmetry claimed in \cite{dualsymm} then means that Yang-Mills theory is invariant under the 2 independent gauge transformations: \begin{equation} A_\mu(x) \longrightarrow A_\mu(x) + \partial_\mu \Lambda(x) + ig\, [\Lambda(x), A_\mu(x)], \label{Gaugetransf} \end{equation} and \begin{equation} \tilde{A}_\mu(x) \longrightarrow \tilde{A}_\mu(x) + \partial_\mu \tilde{\Lambda}(x) + i \tilde{g}\, [\tilde{\Lambda}(x), \tilde{A}_\mu(x)], \label{dGaugetransf} \end{equation} giving thus to a theory with structure group $G$ a `doubled' gauge symmetry $G \times \tilde{G}$, in close analogy to the situation in electromagnetism. This doubling of the gauge symmetry, as in the abelian case, is claimed to be an inherent property of the gauge theory, not an additional input. Again, there is no doubling in the physical degrees of freedom. Given this nonabelian duality let us now examine its implications in the Standard Model with the structure group $SU(3) \times SU(2) \times U(1)$.\footnote{Here, for ease of presentation, we ignore the subtle differences between gauge groups with the same algebra, although of course global properties of the gauge group are of primary importance in considerations of monopole charges. (See e.g. \cite{book}.)} From the discussion in the preceding paragraph, it follows that the theory will have, in addition to the familiar gauge symmetry $SU(3) \times SU(2) \times U(1)$, also the dual gauge symmetry $\widetilde{SU}(3) \times \widetilde{SU}(2) \times \tilde{U}(1)$. This dual gauge symmetry being, according to our observations above, an inherent property of the gauge theory, what we call the Dualized Standard Model \cite{dualcons} which utilizes this dual gauge symmetry is, as at present understood, no different in principle as a gauge theory from the Standard Model itself. As physical schemes, however, they differ, but only in the DSM scheme's recognition of the existence of and its assignment of physical meanings to the dual gauge symmetry which is ignored in the usual SM treatment. In other words, the new DSM results are deduced just by exploring how this `pre-existing' dual gauge symmetry is likely to manifest itself in the physical world. The exploration is based on two main assumptions, namely the identification of 2 ingredients inherent in the theory to 2 physical objects, as we shall now explain. The first assumption concerns the physical interpretation of dual colour. Indeed, the DSM results so far obtained are all from exploitations of only the $\widetilde{SU}(3)$ dual colour symmetry, and it is to this symmetry that our considerations in this paper will be restricted. Colour symmetry $SU(3)$ being known from experiment to be confined, 't~Hooft's famous arguments \cite{tHooft} then suggest that the dual colour symmetry $\widetilde{SU}(3)$ should be broken.\footnote{That the definition of duality given in \cite{dualsymm} is consistent with 't~Hooft's definition given in \cite{tHooft} is shown in \cite{dualcomm}.} Now the idea has long been toyed with that fermion generations may be considered as a broken `horizontal' symmetry. In most schemes, this new symmetry will have to be introduced {\it ad hoc}. Here, however, given the fact that because of duality and the arguments of 't~Hooft, there is inherent in the SM gauge theory already a broken symmetry $\widetilde{SU}(3)$, it seems natural to explore the possibility of making it into the `horizontal' symmetry of generations. The idea is made particularly attractive by the fact that recent LEP experiments have determined the number of generations of light neutrinos to be 3 to a high accuracy \cite{Lep3}. Hence, the first basic assumption one makes in the DSM scheme is that dual colour is to be identified with generations. This means, first, that there will be just 3 generations, no more no less, and second, that any particle carrying a generation index will carry a dual colour charge, implying, as explained above, that it is a colour monopole of the Yang-Mills field $F$. In particular a quark will be a colour `dyon' with both colour (electric) and dual colour (magnetic) charges. Now, 't~Hooft's arguments suggest that the dual colour symmetry is broken, but they do not tell us explicitly how this breaking will occur. The interesting thing is that within the DSM gauge theory, there are quantities which can figure as Higgs fields, and if so identified will imply a particular symmetry breaking pattern. The candidates as Higgs fields in the DSM are the frame vectors in internal symmetry space, which in the case of dual colour is the space of $\widetilde{SU}(3)$. The idea of using frame vectors as dynamical variables is made familiar already in the theory of relativity where in the Palatini treatment or the Einstein-Cartan-Kibble-Sciama formalism \cite{Heyl} the space-time frame vectors or vierbeins are used as dynamical variables. In gauge theory, frame vectors in internal symmetry space are not normally given a dynamical role, but it turns out that in the dualized framework they seem to acquire some dynamical properties, in being patched, for example, in the presence of monopoles \cite{dualsym}. Moreover, they are space-time scalars belonging to the fundamental representations of the internal symmetry group, i.e. doublets in electroweak $SU(2)$ and triplets in dual colour $\widetilde{SU}(3)$, and have finite lengths (as vev's). They thus seem to have just the right properties to be Higgs fields, at least as borne out by the familiar example of the Salam-Weinberg breaking of the electroweak theory. Hence, one makes the second basic assumption in the DSM scheme, namely that these frame vectors are indeed the physical Higgs fields required for the spontaneously broken symmetries. Identifying dual colour with generations and frame vectors with Higgs fields are of course assumptions. It is worth noting, however, that even if one does not choose to do so, the niches in the form of dual colour and frame vectors will still exist and manifest themselves physically in some way which will need to be accounted for in another manner. Opting for the identification, on the other hand, not only offers a hope of determining some of the SM parameters but also gives a much desired geometric significance both to generations and to Higgs fields which has been sadly lacking to an otherwise highly geometric theory. Having made these basic assumptions, let us now explore the consequences. First, making the frame vectors in internal symmetry space into dynamical variables and identifying them with Higgs fields mean that for dual colour $\widetilde{SU}(3)$, we introduce 3 triplets of Higgs fields $\phi^{(a)}_a$, where $(a) = 1, 2, 3$ labels the 3 triplets and $a = 1, 2, 3$ their 3 dual colour components. Further, the 3 triplets having equal status, it seems reasonable to require that the action be symmetric under their permutations, although the vacuum need not be so symmetric. An example of a Higgs potential which breaks both this permutation symmetry and also the $\widetilde{SU}(3)$ gauge symmetry completely is as follows \cite{dualcons}: \begin{equation} V[\phi] = -\mu \sum_{(a)} |\phi^{(a)}|^2 + \lambda \left\{ \sum_{(a)} |\phi^{(a)}|^2 \right\}^2 + \kappa \sum_{(a) \neq (b)} |\bar{\phi}^{(a)} .\phi^{(b)}|^2, \label{Vofphi} \end{equation} a vacuum of which can be expressed without loss of generality in terms of the Higgs vev's: \begin{eqnarray} \phi^{(1)} = \zeta \left( \begin{array}{c} x \\ 0 \\ 0 \end{array} \right), \,\,\, \phi^{(2)} = \zeta \left( \begin{array}{c} 0 \\ y \\ 0 \end{array} \right) , \,\,\, \phi^{(3)} = \zeta \left( \begin{array}{c} 0 \\ 0 \\ z \end{array} \right), \label{vevs} \end{eqnarray} with \begin{equation} x^2 + y^2 + z^2 = 1, \label{normvevs} \end{equation} and \begin{equation} \zeta = \sqrt{\mu/2 \lambda}, \label{zeta} \end{equation} $x, y, z$, and $\zeta$ being all real and positive. Indeed, this vacuum breaks not just the symmetry $\widetilde{SU}(3)$ but even the larger symmetry $\widetilde{SU}(3) \times \tilde{U}(1)$ completely giving rise to 9 massive dual gauge bosons. And of the 18 real components in $\phi^{(a)}_a$, 9 are thus eaten up, leaving just 9 (dual colour) Higgs bosons. Next, we turn to the fermion fields. In analogy to the electroweak theory, the left-handed fermion fields are assigned the triplet {\bf 3} and the right-handed fermions the singlet {\bf 1} representation in dual colour, so that one can construct their Yukawa couplings as: \begin{equation} \sum_{(a)[b]} Y_{[b]} \bar{\psi}^a_L \phi^{(a)}_a \psi^{[b]}_R, \label{Yukawa} \end{equation} which, by the same reasons as for the Higgs potential, we have made symmetric under permutations of the Higgs fields $\phi^{(a)}$. As a result of this permutation symmetry, the tree-level fermion mass matrix takes the following factorized form: \begin{equation} m = \zeta \left( \begin{array}{c} x \\ y \\ z \end{array} \right) (a, b, c), \label{mtree} \end{equation} where $a, b, c$ are just abbreviations for the Yukawa couplings $Y_{[b]}$. Written in the hermitized convention of e.g. \cite{Weinberg}, which is basically $\sqrt{m m^{\dagger}}$ in terms of the $m$ above and is the matrix of relevance to the mass spectrum, it becomes: \begin{equation} m = m_T \left( \begin{array}{c} x \\ y \\ z \end{array} \right) (x,y,z), \label{fermass0} \end{equation} where $m_T$ is a mass scale dependent on the fermion type $T = U, D, L, N$, i.e. whether $U$-type quarks ($U$), $D$-type quarks ($D$), charged leptons ($L$) or neutrinos ($N$), but the vector $(x, y, z)$ given in terms of the Higgs vev's is independent of the fermion type. One immediate consequence of such a factorized mass matrix is that at tree-level there is only one state with mass, represented by the eigenvector $(x,y,z)$ with eigenvalue $m_T$, the other 2 states having zero eigenvalues. This we interpret as embryo fermion mass hierarchy, with one generation much heavier than the other two generations. Such an arrangement is not a bad first approximation to the experimental situation where the highest generation states of $U, D, L$ are all more than one order of magnitude heavier than the lower states. Another immediate consequence is that at tree-level the CKM matrix is the identity, since the CKM matrix is the matrix giving the relative orientations between the physical states of the $U$- and $D$-type quarks, and these orientations at tree-level are both given in (\ref{fermass0}) by the vector $(x,y,z)$. Again, this is not a bad first approximation to the experimental CKM matrix whose off-diagonal elements are at most of order 20 percent. Though reasonable as a first approximation, this tree-level description is obviously too degenerate to be realistic. We shall see, however, that the degeneracy will be lifted at higher orders where nonzero masses for the two lower generations and nonzero off-diagonal elements for the CKM matrix will both result from loop corrections. \setcounter{equation}{0} \section{Quark Mixing and Light Masses as Loop Corrections} Radiative corrections to the tree-level mass matrix (\ref{fermass0}) have already been calculated to 1-loop level \cite{ourckm}. There are in principle many diagrams to calculate. There are first the usual diagrams of the Standard Model with loops of gluons, $\gamma$, $Z$, $W$, and electroweak Higgses. Then there are new diagrams with loops of dual gluons and dual Higgses. All these diagrams, however, share the common feature that they preserve the factorizability of the fermion mass matrix, namely that after the corrections the mass matrix $m'$ is of the form: \begin{equation} m' = m_T' \left( \begin{array}{c} x' \\ y' \\ z' \end{array} \right) (x',y',z'). \label{fermassc} \end{equation} The reason is that only those bosons carrying dual colour can affect the factorizability, and of these the dual gluon couples only to the left-handed fermion while the dual Higgses have themselves factorizable couplings. Indeed, it is believed, though not formally proved, that factorizability of $m'$ will remain intact to all orders. Only a small number of these diagrams, however, need be considered for the purpose of this paper. The reason is as follows. The value of the normalization factor $m_T'$ in (\ref{fermassc}) is actually not calculable perturbatively since it receives contributions from diagrams with a dual gauge boson loop, and these are proportional to the square of the dual gauge coupling $\tilde{g}$ which is large according to the Dirac condition (\ref{Diraccond}), given the known empirical value of the colour coupling $g$. The normalization $m'_T$ has thus to be treated in any case as an empirical parameter. In other words, the calculation of any diagram which affects only this normalization will not add to our present level of understanding and might as well be ignored. What one can calculate perturbatively, on the other hand, is the orientation of the vector $(x',y',z')$ and this is affected by only a small subset of the diagrams, namely those shown in Figure~\ref{oneloopdiag}. For example, it is clear that diagrams with only loops of gluons or $W$-bosons will not affect the orientation of $(x', y', z')$ since these bosons do not carry dual colour. Next, of the remaining 3 diagrams as in Figure~\ref{oneloopdiag}, the first 2 can give only negligible effects. This last conclusion is arrived at as follows. The diagrams (a) and (b) both give rotations to $(x', y', z')$ but these are of the order $\tilde{g}^2 m_T^2/\mu_N^2$, where $\mu_N$ represents the masses of the dual gauge bosons which are constrained by present experimental bounds on flavour-changing neutral current (FCNC) effects to be rather large. The identification of dual colours to generations means that particles carrying generation indices can interact by exchanging dual gluons, which would lead to generation-changing or FCNC effects, such as $K_L \rightarrow e \mu$ decay or an anomalous $K_L - K_S$ mass difference, and such effects are strongly bounded by experiment. An analysis using the latest data \cite{Cahnrari,ourfcnc} which will be outlined in Section 5 below, shows that a violation to present experimental bounds can be avoided only if $\mu_N^2/\tilde{g}^2$ is of order at least a few hundred TeV, which means contributions from diagrams (a) and (b), even for the top quark, are of at most $10^{-6}$. Neglecting then contributions of this order, the calculation becomes rather simple depending on only the (dual) Higgs loop diagram of Figure \ref{oneloopdiag} (c) \cite{ourckm}. \begin{figure}[htb] \vspace{1cm} \centerline{\psfig{figure=oneloopdiag.eps,width=0.9\textwidth}} \vspace{.5cm} \caption{One loop diagrams rotating the fermion mass matrix.} \label{oneloopdiag} \end{figure} The rotation given to the normalized vector $(x',y',z')$ by the remaining Higgs loop diagram Figure \ref{oneloopdiag} (c) has been calculated. It is found to depend on the energy scale $\mu$ as follows: \begin{equation} \frac{d}{d(\ln \mu^2)} \left( \begin{array}{c} x' \\ y' \\ z' \end{array} \right) = \frac{5}{64 \pi^2} \rho^2 \left( \begin{array}{c} \tilde{x}'_1 \\ \tilde{y}'_1 \\ \tilde{z}'_1 \end{array} \right), \label{runxyz} \end{equation} with \begin{equation} \tilde{x}'_1 = \frac{x'(x'^2-y'^2)}{x'^2+y'^2} + \frac{x'(x'^2-z'^2)} {x'^2+z'^2}, \ \ \ {\rm cyclic}, \label{x1tilde} \end{equation} and $\rho^2 = |a|^2 + |b|^2 + |c|^2$ being the Yukawa coupling strength. As $\mu$ varies the normalized vector traces out a curve on the unit sphere. It is easily seen from (\ref{runxyz}) and (\ref{x1tilde}) that the points $(1,0,0)$ and $\frac{1}{\sqrt{3}}(1,1,1)$ are fixed points under scale changes, and for decreasing energy, the point $(x',y',z')$ runs away from $(1,0,0)$ towards $\frac{1}{\sqrt{3}}(1,1,1)$. The trajectory along which $(x',y',z')$ runs depends in principle on the Higgs boson masses logarithmically, but because of a `happy accident' that we shall make clear later, even this weak dependence can be ignored. The only parameters remaining are then just the initial values of the components of the normalized vector $(x',y',z')$ identifiable with the Higgs vev's which are the same for all fermion types, and the Yukawa coupling strength $\rho_T$, one for each fermion type $T$. The information being now encoded in the vector $(x',y',z')$ running along a trajectory on the unit sphere, the remaining question is just how to extract the fermion mass and mixing parameters from this rotating vector. Let us first consider quarks of the $U$-type and ask what is the physical state vector the of $t$ quark in dual colour or generation space. Now, according to (\ref{fermassc}), $(x',y',z')$ is the vector with the single nonzero eigenvalue of the loop-corrected fermion mass matrix at the energy scale where $(x',y',z')$ is evaluated. It ought therefore to be identified with the state vector of the heaviest generation, except that $(x',y',z')$ has to be evaluated at the scale $\mu$ equal to the top mass $m_t$. Inputting then the empirical value of $m_t \sim 176\ {\rm GeV}$ we run $(x',y',z')$ to the scale $\mu = m_t$. This gives us the physical state vector $|{\bf v}_t \rangle$ and also fixes the value of the parameter $m_U'$ at the same scale. This calculation, however, still depends on the starting value $(x,y,z)$ of the rotating vector $(x',y',z')$ and also on the Yukawa coupling strength $\rho_U$. Next, we ask what is the physical state vector of the $c$ quark. Being an independent physical entity, the $c$ quark ought to have a state vector orthogonal to $|{\bf v}_t \rangle$. This means it must have a zero eigenvalue of the mass matrix (\ref{fermassc}) evaluated at the scale $\mu = m_t$. But there are two linearly independent such vectors and we do not as yet know which linear combination of these should correspond to the $c$ quark. Nor can we identify the mass $m_c$ as the zero eigenvalue for this is evaluated at the wrong scale $m_t$. Let us first extract the mass submatrix in the 2-dimensional subspace orthogonal to $|{\bf v}_t \rangle$ and run it down to lower scale. Being a submatrix of a rank 1 matrix, it is of rank at most 1 at this lower scale but need no longer be a zero matrix since the nonzero eigenvector $(x',y',z')$ of the full mass matrix (\ref{fermassc}) has already rotated from the direction $|{\bf v}_t \rangle$ and can have thus a component in the subspace orthogonal to $|{\bf v}_t \rangle$ which contains $|{\bf v}_c \rangle$. For consistency with the way we fixed the vector $|{\bf v}_t \rangle$ before, the procedure to determine $|{\bf v}_c \rangle$ is now clear. We ought to run the $2 \times 2$ mass submatrix down in energy until the scale $\mu$ equals the empirical value $m_c$ of the $c$ quark mass. Then its only nonzero eigenvector at that scale is to be identified with $|{\bf v}_c \rangle$. This vector is, of course, by definition orthogonal to $|{\bf v}_t \rangle$ as it should be. Its eigenvalue of the mass submatrix, however, will not in general be the same as the input value of $m_c$. But our calculation, we recall, still depends on $(x,y,z)$ and $\rho_U$, and by adjusting $\rho_U$ we can make the output eigenvalue of $|{\bf v}_c \rangle$ the same as the input value of $m_c$. This fixes $\rho_U$, leaving now only $(x,y,z)$ as parameters. Once the state vectors of the $t$ and $c$ quarks are determined, then obviously the state vector of the $u$ quark is also fixed as the vector orthogonal to both. We have thus the whole triad of physical state vectors of the $U$-type quarks in terms of the 2 remaining parameters in the normalised vector $(x,y,z)$ representing the Higgs vev's. The above procedure can now be repeated for the $D$-type quarks and for the charged leptons to determine the triad of their physical state vectors. (The problem for neutrinos is somewhat different as we shall make clear later.) To do so, we shall have to input the empirical masses of the two highest generations in each case, namely $b$ and $s$ for the $D$-type quarks and $\tau$ and $\mu$ for the charged leptons, and then, as before for the $U$-type quarks, to adjust the Yukawa coupling strengths $\rho_D$ and $\rho_L$ to obtain consistency. The values of the $\rho$'s determined in this way need not of course be the same for $D$ and $L$ nor as the value obtained before for the $U$-type quarks. Again the triads so obtained still depend on $(x,y,z)$, which is the same for all fermion types. Now the matrix relating the orientations in generation space of the two triads of physical state vectors of respectively the $U$-type and $D$-type quarks is what is known in the literature as the Cabibbo-Kobayashi-Moskawa (CKM) matrix. These orientations being now known, the CKM matrix can be calculated in terms of the two remaining parameters in $(x,y,z)$. By adjusting these, we can then try to fit the empirical CKM matrix. There are actually 4 independent degrees of freedom in the CKM matrix, but in our treatment up to 1-loop order, there is no $CP$-violating phase, the vector $(x',y',z')$ being real. We are thus left with 3 real quantities to fit with our two remaining parameters, which is still nontrivial but has been achieved rather well. For the matrix of absolute values $|V_{rs}|$, for $r = u,c,t$ and $s = d,s,b$, we obtained \cite{ourckm} \footnote{This fit was obtained using data given in the databook of 1996 \cite{databook}. For this reason the comparison with experiment given in this paper refers for consistency also to data from the same source. Although much of the data have since been updated, the changes are not large. For a parallel fit to the latest data and comparison with them, see \cite{phenodsm}, which arrives at a similar conclusion to that presented in this paper.}: \begin{equation} |V_{rs}| = \left( \begin{array}{ccc} 0.9755 & 0.2199 & 0.0044 \\ 0.2195 & 0.9746 & 0.0452 \\ 0.0143 & 0.0431 & 0.9990 \end{array} \right), \label{calckmq} \end{equation} as compared with the experimental values \cite{databook}: \begin{equation} |V_{rs}| = \left( \begin{array}{lll} 0.9745-0.9757 & 0.219-0.224 & 0.002-0.005 \\ 0.218-0.224 & 0.9736-0.9750 & 0.036-0.046 \\ 0.004-0.014 & 0.034-0.046 & 0.9989-0.9993 \end{array} \right). \label{expckmq} \end{equation} The above fit to the CKM matrix fixes all the parameters in the problem, the values of which so obtained show two very intriguing features. First, the vev's of the Higgs fields $(x,y,z)$ turn out to have values very close to the high energy fixed point $(1,0,0)$. Second, and even more intriguingly, the Yukawa coupling strengths $\rho_T$ turn out to be $T$-independent to a surprising accuracy \cite{ourckm}. Indeed, for fermion masses taken at the (geometric) median of experimental values, the fitted values of $\rho_T$ for $T = U, D, L$ differ by only 1.5 parts in a thousand, while varying the masses within the experimental error bars still give differences of at most a few percent. This last `happy accident' seems to indicate some hidden symmetry in the scheme, the reason for which we have at present only some ideas not yet fully understood. In practical terms, on the other hand, it is a bonus since it simplifies greatly not only the calculation but also the presentation of the results. It means, first, that the 3 originally different parameters $\rho_T$ can now be treated as just a single parameter $\rho$; second, that all fermion types run on the same trajectory at the same speed; and third, even the originally already weak dependence of the calculation on the Higgs boson masses can now be completely ignored. Next, we turn to the masses of the lowest generation. Since we know already the physical state vectors of the lowest generation states for all the 3 fermion types $U, D, L$, namely $u, d, e$, we can simply evaluate their expectation values of the mass matrix $m'$ at any scale. The actual masses of the 3 states are given by these expectation values evaluated at the scales equal to the values themselves, and can also be readily calculated. However, the calculation of these lowest generation masses is an extrapolation on a logarithmic scale and depends also on the normalization $m'_T$ of the mass matrix $m'$ in (\ref{fermassc}), whose variation with scale, as we have already explained, is not calculable perturbatively. It is thus expected to be far less reliable than the above calculation for the mixing parameters. Nevertheless, assuming simply that the normalization $m'_T$ is scale-independent, one may hope to get rough order-of-magnitude estimates for the lowest generation masses. The result of such a calculation is shown in Table \ref{masstable}. We notice that the mass of the electron is within about an order of magnitude of the experimental value, which we regard as reasonable. As for the quarks, the $d$-quark comes out about right but the $u$ quark is too large by nearly two orders of magnitude. However, we should note that light quark masses are notoriously difficult to define, being sensistive to nonperturbative QCD corrections below around 2 GeV. Moreover, the $u$ and $d$ masses given in the table are defined each at the scale equal to its value whereas the quoted experimental values are defined at the scale of 1 GeV. The two sets of values are therefore not directly comparable. Indeed, if the expectation value in the $u$-state of the running mass matrix $m'$ is taken at 1 GeV, a value of order only 1 MeV is obtained, although it is also unclear whether this should be compared with the empirical value quoted. In all cases, at least, the masses are hierarchical as they should be. \begin{table} \begin{eqnarray*} \begin{array}{||c|c|c||} \hline \hline & Calculation & Experiment \\ \hline m_c & 1.327 {\rm GeV} & 1.0-1.6 {\rm GeV} \\ m_s & 0.173 {\rm GeV} & 100-300 {\rm MeV} \\ m_\mu & 0.106 {\rm GeV} & 105.7 {\rm MeV} \\ m_u & 209 {\rm MeV} & 2-8 {\rm MeV} \\ m_d & 15 {\rm MeV} & 5-15 {\rm MeV} \\ m_e & 6 {\rm MeV} & 0.511 {\rm MeV} \\ m_{\nu_1} & 10^{-15} {\rm eV} & < 10 {\rm eV} \\ B & 400 {\rm TeV} & ? \\ \hline \hline \end{array} \end{eqnarray*} \caption{Predicted fermion masses compared with experiment. Notice, however, that for the $u$- and $d$-quarks, the calulated masses are defined each at the scale equal to its value, and are not directly comparable to the quoted experimental values defined at the scale of 1 GeV.} \label{masstable} \end{table} \setcounter{equation}{0} \section{Neutrino Oscillations} Next, we turn our attention to neutrinos. The case for neutrino oscillations has recently been much strengthened by the atmospheric neutrino data from SuperKamiokande \cite{superk} confirming earlier results of the last decade \cite{kamioka}--\cite{soudan}. These have not only given convincing evidence for the phenomena, but have even provided quite restrictive bounds on the relevant parameters which would be a challenge for theoreticians to explain. If the basic idea in the DSM scheme of identifying generations with dual colour is adhered to, then there will be 3 and only 3 generations of neutrinos as for any other fermion type. In principle then, there is nothing to stop us applying the same procedure as that applied above to quarks and charged leptons to determine also the masses and physical state vectors of neutrinos. Indeed, in strict adherence to the scheme, it would be incumbent upon us to do so. Since the Higgs vev's $(x,y,z)$ are already known, all we need to do so would be to input the (Dirac) masses of the two heaviest neutrinos as we did for quarks and charged leptons. In fact, since the Yukawa couplings $\rho$ turned out to be so close for all the other fermion types, it seems reasonable to assume the same value also for neutrinos. We shall need then to input only one (Dirac) mass. However, neutrinos differ from the other fermions in that they can also have Majorana masses, and it is from these together with their Dirac masses that one obtains their physical masses via the well-known seesaw mechanism \cite{Seesaw} by diagonalizing the matrix: \begin{equation} {\bf M}_r = \left( \begin{array}{cc} 0 & M_r \\ M_r & B \end{array} \right), \label{seesaw} \end{equation} where it turns out that in the DSM scheme as at present understood, all 3 generations of neutrinos will have to have the same Majorana mass $B$ for consistency \cite{ournuos}. With then $B$ as an extra parameter, we need to input two masses to perform the proposed calculation. Information on the (physical) mass of the heaviest neutrino, usually denoted by $m_3$, is obtained from the muon anomaly in atmospheric neutrinos. From \cite{kamioka,superk}, we have an estimate of the difference between the physical masses of the two heaviest neutrinos. Since in the DSM scheme, masses are supposed to be hierarchical, meaning $m_3 \gg m_2 \gg m_1$, we can put the mass itself equal to the difference $m_3^2 \sim 10^{-3}-10^{-2}\ {\rm eV}^2$, which we can take as one input, leaving thus just one more mass to be determined. To do so, we draw on the information from solar neutrino data. There one has 2 estimates for the (physical) mass of the second generation neutrino, again taking the masses to be hierarchical. From the LWO solution \cite{LWOfit}, one has $m_2^2 \sim 10^{-10}\ {\rm eV}^2$, and from the MSW solution \cite{MSWfit} $m_2^2 \sim 10^{-5}\ {\rm eV}^2$. With either of these as input, one has in principle enough information to determine the Dirac mass $M_3$ and hence complete the DSM calculation of the leptonic CKM matrix. It turns out, however, that inputting the MSW estimate for $m_2$ and the estimate of \cite{kamioka,superk} for $m_3$, one obtains no sensible DSM solution for neutrinos. The reason is that in the DSM scheme, lower generation masses come only as a `leakage' from the mass of the highest generation and this leakage mechanism does not easily admit a ratio $m_2/m_3$ as large as that wanted by the MSW solution. One concludes thus that the DSM scheme, as at present understood, disfavours the MSW solution to the solar neutrino problem.\footnote{It is interesting to note in this context that the latest SuperKamiokande data on day-night variations and flux reported at the Vancouver ICHEP'98 conference and at the APS meeting (DPF) at UCLA in fact also favour the LWO over the MSW solution for solar neutrinos \cite{Vagins,superk}.} On the other hand, inputting the LWO estimate for $m_2^2 \sim 10^{-10}\ {\rm eV}^2$ and the estimate of $m_3^2 \sim 10^{-3}- 10^{-2}\ {\rm eV}^2$ from \cite{kamioka,superk}, a solution is readily found. The state vectors of the neutrinos so determined then allow one immediately to calculate the leptonic CKM matrix. For $m_3^2 = 10^{-3} {\rm eV}^2$, one obtains \cite{ournuos}: \begin{equation} \left( \begin{array}{ccc} U_{e1} & U_{e2} & U_{e3} \\ U_{\mu1} & U_{\mu2} & U_{\mu3} \\ U_{\tau1} & U_{\tau2} & U_{\tau3} \end{array} \right) = \left( \begin{array}{ccc} 0.97 & 0.24 & 0.07 \\ 0.22 & 0.71 & 0.66 \\ 0.11 & 0.66 & 0.74 \end{array} \right), \label{calckml} \end{equation} all elements being real at the 1-loop level that we are working. The result is insensitive to the actual values (in the above range) of $m_2$ and $m_3$ used. Notice that apart from inputting the rough values of $m_2$ and $m_3$ from experiment in the way explained, the calculation involves no adjustment of parameters which have all been fixed by our earlier calculation of the quark CKM matrix \cite{ourckm}. The result (\ref{calckml}) should be compared with the leptonic CKM matrix extracted from experiment by, for example, \cite{Giunkimno}\footnote{This analysis \cite{Giunkimno} was done with the Kamiokande not with the SuperKamiokande data, for which no parallel analysis as far as we know has yet been done. The result, however, is expected to be similar, with a somewhat tighter bound but roughly the same central value for the $\mu 3$ element, but a looser bound on the $e 3$ element. For a comparison with the latest data from SuperKamiokande, see \cite{phenodsm}.} where the bounds on $U_{\mu3}$ comes mainly from atmospheric neutrino data and the bounds on $U_{e3}$ comes mainly from reactor data such as \cite{chooz} and the estimate for $U_{e2}$ comes from the solar neutrino data as interpreted by either the large angle MSW \cite{MSWfit} or the LWO \cite{LWOfit} scenario: \begin{equation} = \left( \begin{array}{ccc} \ast & 0.4-0.7 & 0.0-0.2 \\ \ast & \ast & 0.5-0.8 \\ \ast & \ast & \ast \end{array} \right). \label{expckml} \end{equation} Since we are ignoring for the present the CP-violating phase, there are only 3 independent elements of the matrix we need consider. We notice that the theoretical predictions for both the angles $U_{\mu 3}$ and $U_{e 3}$ fall neatly into the middle of the experimental range. A more detailed comparison of our prediction with experiment based again on the analysis in \cite{Giunkimno} is shown in Figures \ref{comp1} and \ref{comp2} for a range of $m_3$ values and for $m_2$ within the range allowed by \cite{LWOfit}. One sees that the agreement is consistently good. \begin{figure}[htb] \vspace{-3cm} \centerline{\psfig{figure=ue3.eps,width=0.75\textwidth}} \caption{90 \% CL limits on the CKM element $U_{e 3}$ compared with the result of our calculation.} \label{comp1} \end{figure} \begin{figure}[htb] \vspace{-3cm} \centerline{\psfig{figure=umu3.eps,width=0.75\textwidth}} \caption{90 \% CL limits on the CKM element $U_{\mu 3}$ compared with the result of our calculation.} \label{comp2} \end{figure} The prediction, however, for the other angle $U_{e 2}$ relevant to solar neutrinos does not score so well, being about a factor 2 too small and lying some way outside the experimental limits. We shall see later the reason why this element is particularly hard for the DSM scheme to get correct. In addition to the mixing angles, the calculation gives predictions also for the masses of the lightest and the `right-handed' neutrinos, namely $m_{\nu_1}$ and $B$, as listed in Table \ref{masstable}. The present experimental bound on $m_{\nu_1}$ is too weak to be a test. As for $B$, there is no direct information. However, given $B$, one can estimate within the scheme a value for the life-time of neutrinoless double beta decays. One notes that the value of $B$ we obtained is several orders of magnitude lower than that usually given, say for example, from grand unified theory models. The reason is that one usually assumes for the heaviest neutrino a Dirac mass similar in value to the mass of the $\tau$ or $t$, i.e. of order GeV or higher, whereas the value we obtain above by fitting $m_3$ and $m_2$ in the DSM scheme gives a value only of order MeV (and $B$ is proportional to the square of this estimate). Such a big difference between the (Dirac) masses of the charged leptons and neutrinos need not be a worry since the same is known already to occur between the $U$- and $D$-type quarks. But as a result of this lower value for $B$, the rate for neutrinoless double beta decays predicted here will be much more accessible to experiment. In particular, a rough estimate shows that the predicted $0\nu$ half-life for ${}^{76}Ge$ is only about 2--3 orders longer than the present experimental limit. To conclude, one sees that the DSM with no freedom left after fitting the quark CKM matrix, reproduces quite well the general features of neutrino oscillations as observed in experiment, and gives in addition some interesting and in principle testable predictions. \setcounter{equation}{0} \section{Features of Mixing from Differential Geometry} It is instructive to compare the quark (\ref{expckmq}) and leptonic (\ref{expckml}) CKM matrices as now experimentally known. We note in particular the following outstanding features: \begin{description} \item{(a)} The corner elements 13 and 31 are much smaller than the others for both quarks and leptons. \item{(b)} All off-diagonal elements for quarks are much smaller than the diagonal elements. \item{(c)} The 23 element is much smaller for quarks than for leptons. \end{description} These will need to be accounted for in any scheme aiming to explain the fermion mixing phenomenon. In the preceding sections we have already shown that the DSM scheme is able quantitatively to reproduce these features in terms of just a few parameters without explaining why it should be so. What we shall do in this section is to gain an intuitive understanding why the CKM matrices have the qualitative features they do and to show that they emerge from the basic structure of the scheme as simple consequences of classical differential geometry and can be deduced, almost quantitively in some cases, without a detailed calculation \cite{features}. We recall that the fermion mass matrix in the DSM scheme is factorized even after loop corrections, and all the information needed for our consideration of the CKM matrix is encoded in the normalized vector $(x',y',z')$. This vector rotates with the energy scale, tracing out a trajectory on the unit sphere, which trajectory is the same for all fermion types $T$, i.e. whether $U$- or $D$-type quarks, charged leptons or neutrinos. The various physical states, however, differ in the locations they occupy on this trajectory. Figure \ref{runtraj} shows the actual trajectory and locations of the 12 fermion states obtained in the fit of \cite{ourckm,ournuos}. \begin{figure}[htb] \vspace{-5cm} \centerline{\psfig{figure=sphere.eps,width=0.85\textwidth}} \vspace{0cm} \caption{The trajectory traced out by $(x',y',z')$ and the locations on it of the 12 fermion states.} \label{runtraj} \end{figure} The state vectors of the various physical states are given in terms of the rotating vector $(x',y',z')$ as follows. (i) Evaluated at the scale of the top mass, $(x',y',z')$ is the state vector $|{\bf v}_1 \rangle$ of $t$, as shown in Figure \ref{v1v2v3}. (ii) At the scale of the charm mass, the vector $(x',y',z')$ is rotated to another direction, say $|{\bf \tilde{v}}_1 \rangle$ in Figure \ref{v1v2v3}, with thus a zonzero component orthogonal to $|{\bf v}_1 \rangle$, the direction of which gives the state vector $|{\bf v}'_2 \rangle$ of $c$. (iii) The state vector of the $u$-quark is ${\bf v}'_3 = {\bf v}_1 \wedge {\bf v}'_2$. \begin{figure}[htb] \centerline{\psfig{figure=curve1.eps,width=0.6\textwidth}} \vspace{0cm} \caption{The state vectors of the 3 physical states belonging to the 3 generations of the $U$-type quark.} \label{v1v2v3} \end{figure} Similar constructions apply to the other 3 fermion types $D, L, N$. If we make the approximation that the locations of the $t$ and $c$ quarks on the trajectory are close together (as is seen to be true in Figure \ref{runtraj}), then the 3 state vectors of $t,c,u$ of Figure \ref{v1v2v3} form in that limit an orthonormal triad at the $t$ position. If we do the same for the $D$-type quarks, we have another such triad at the $b$ position, as illustrated in Figure \ref{2triads}. The entries of the \begin{figure}[htb] \centerline{\psfig{figure=curve2.eps,width=0.6\textwidth}} \vspace{0cm} \caption{Two triads of state vectors for two fermion types transported along a common trajectory.} \label{2triads} \end{figure} (quark) CKM matrix are nothing but the direction cosines between the vectors of these 2 triads. The leptonic CKM matrix is similar. Since the trajectory lies on the unit sphere, the $c$ vector is the tangent {\bf T} to the trajectory and the $t$ vector the normal {\bf N} to the surface, so that they form, together with the $u$ vector ${\bf B}={\bf N} \wedge {\bf T}$, what is known in elementary differential geometry as the `Darboux triad'. Differentiating then with respect to the arc-length, we get the following formulae similar to the well-known Serret--Frenet formulae for space curves \cite{docarmo}: \begin{eqnarray} {\bf N}' & = & -\kappa_n {\bf T} - \tau_g {\bf B}, \nonumber\\ {\bf T}' & = & \kappa_g {\bf B} + \kappa_n {\bf N}, \nonumber \\ {\bf B}' & = & -\tau_g {\bf N} - \kappa_g {\bf T}. \label{SFDarboux} \end{eqnarray} Here $\kappa_g$ is the geodesic curvature, $\kappa_n$ the normal curvature, and $\tau_g$ the geodesic torsion of the curve on the surface. Equivalently, to first order in arc-length $\Delta s$, (\ref{SFDarboux}) can be rewritten in the form of a CKM matrix with entries arranged in the conventional order: \begin{equation} \left( \begin{array}{ccc} 1 & -\kappa_g \Delta s & -\tau_g \Delta s \\ \kappa_g \Delta s & 1 & \kappa_n \Delta s \\ \tau_g \Delta s & -\kappa_n \Delta s & 1 \end{array} \right). \label{ckmdg} \end{equation} In our case of the unit sphere, $\tau_g=0$ and $\kappa_n=1$. It follows then from (\ref{ckmdg}) that \cite{features}: \begin{description} \item{(a)} The corner elements of both the quark CKM matrix ($V_{ub}, V_{td}$) and the leptonic CKM matrix ($U_{e3}, U_{\tau 1}$) are small since they vanish to first order in the separation between the corresponding fermion types. \item{(b)} The 4 other off-diagonal elements of the quark CKM matrix are small compared to the diagonal elements since they are of first order in the separation between the $t$ and $b$ quarks, which is small as seen in Figure~\ref{runtraj}. \item{(c)} The elements $V_{cb}, V_{ts}$ for quarks are much smaller than their counterparts $U_{\mu 3}, U_{\tau 2}$ for leptons, since they are to first order proportional to the separation, which is much smaller for quarks than for leptons as seen in Figure~\ref{runtraj}. \end{description} These 3 points are all borne out by experiment as already noted above. Indeed, it is amusing to note that even the approximate values for the 4 elements in (c), as quoted above from either experiment (\ref{expckmq}), (\ref{expckml}) or the DSM calculation (\ref{calckmq}), (\ref{calckml}), can simply be read off by measuring the separations between $t$ and $b$ and between $\tau$ and $\mu$ on the trajectory in Figure~\ref{runtraj}! Further, we note that since the geodesic curvature $\kappa_g$, in contrast to the geodesic torsion $\tau_g$ and the normal curvature $\kappa_n$ on a sphere, depends both on the location and on the trajectory, so do the values of the remaining pair of off-diagonal elements of the CKM matrix, namely the `Cabibbo angles' ($V_{us}, V_{cd}$ for quarks, and $U_{e2}, U_{\mu 1}$ for leptons). This explains why the Cabbibo angle is so large even though the separation between $t$ and $b$ is small. It also means that the 12 elements are much more sensitive to the details of the fit and explains why our calculation has been less successful in predicting $U_{e2}$ than with the other leptonic mixing angles. That all these features in the mixing matrices echoing experimental data can be derived {\em without} detailed calculations is very encouraging, for it means that the agreement with experiment reported in the 2 sections before are much less likely to be just numerical accidents of the calculation. \setcounter{equation}{0} \section{FCNC Effects from Dual Gluon Exchange} Besides explaining the features of the Standard Model, any scheme which attempts to go beyond has of course also to examine its own predictions for the possibility of their violating already some known experimental limits, and if not, for the feasibility of their being tested by future experiment. For the DSM scheme, one obvious direction to probe in this respect is the new interactions arising from exchanges of the dual colour gauge bosons. Dual colours in DSM having been identified with generations, it follows that any particle carrying a generation index can acquire a new interaction by exchanging these bosons, leading to generation-changing or flavour-changing neutral current (FCNC) effects. These gauge bosons are presumably quite heavy or otherwise they would already have been discovered. There are thus two areas where one can look for their influence. One can either look for effects at energies low compared with their mass where the effects of their exchange would be suppressed, or else for sizeable effects at ultra-high energies. We shall consider examples of both, at low energies in this section, and at high energies in the next. At low energy, flavour-changing neutral currents can manifest themselves in rare decays and in mass differences between charge conjugate neutral meson pairs. For DSM, as for other `horizontal gauge symmetry' models \cite{buchmuller}, these effects arise already at the level of one-(FCNC) gauge boson exchange and can thus be estimated once given the masses of the gauge bosons and their couplings to the fermions involved. What distinguishes DSM, however, is that the scheme has been made so restrictive by what has gone before that detailed estimates can now be given for all these FCNC effects at the one-boson exchange level in terms of only one additional parameter. In view both of the intrinsic structure built into the scheme and of the calculations already performed which are summarized above, most of the `fundamental' parameters of DSM as at present formulated are now known. First, by virtue of the Dirac quantization condition \cite{dualcomm}: \begin{equation} g_3 \tilde{g}_3 = 4\pi, \;\;\; g_2 \tilde{g}_2 = 4\pi, \;\;\; g_1 \tilde{g}_1 = 2\pi, \label{couplings} \end{equation} the coupling strengths $\tilde{g}_i$ of the dual gauge bosons are derivable from the coupling strengths $g_i$ of the ordinary colour and electroweak gauge bosons measured in present experiments. Secondly, the branching of these couplings $\tilde{g}_i$ into the various physical fermion states are given by the rotation matrices relating these physical states to the `gauge states' in generation or dual colour space, thus: \begin{equation} \psi_{gauge,L}^T = S^T \psi_{physical,L}^T \label{rotationf} \end{equation} where the index $T$ runs over the four types of fermions $U, D, L$ and $N$. These matrices $S^T$ were already determined in the calculation of fermion mixing matrices \cite{ourckm,ournuos}, where for example the (quark) CKM matrix was obtained as $(S^U)^\dagger S^D$ in Section 2 and there found to be in excellent agreement with experiment. Finally, in tree-level approximation, the masses of the dual gauge bosons are given in terms of the vacuum expectation values of the dual colour Higgs fields, the ratios $x, y, z$ between which are among the parameters determined in the calculation \cite{ourckm} by fitting the CKM matrix. Thus the only remaining unknown among the quantities required is the actual strength $\zeta$ of the vev's, which, though also entering in principle in the calculations of Standard Model parameters outlined in Section 2, turns out to be hardly restricted there. That being the case, one can now calculate in the DSM scheme all one-dual gauge boson exchange diagrams between any two fermions in terms of this single mass parameter $\zeta$. There are altogether 9 dual gauge bosons (including that corresponding to $\tilde{U}(1)$) which can be exchanged, whose masses at tree-level are given by diagonalizing a mass matrix dependent on the dual couplings $\tilde{g}_3, \tilde{g}_1$ and on $\zeta$ and $(x, y, z)$. Given that, as mentioned in Section 2, the value obtained for $(x, y, z)$ from fitting the quark CKM matrix \cite{ourckm} is very close to the fixed point $(1,0,0)$, the mass matrix for the dual gauge bosons can be readily diagonalized \cite{ourfcnc} yielding one particular state with mass: \begin{equation} M^2 = \zeta^2 z^2 \frac{3}{4} \frac{\tilde{g}_3^2} {1 + \frac{3}{16}\frac{\tilde{g}_3^2}{\tilde{g}_1^2}}, \label{gmass} \end{equation} which is much lower than the rest. As a result, the calculation for FCNC effects becomes quite simple, being dominated by the exchange of just this one boson. At energies much lower than the mass of this dual gauge boson, the effects can then be summarized in terms of an effective Lagrangian thus \cite{ourfcnc}: \begin{equation} L_{eff}= \frac{1}{2 \zeta^2 z^2} \sum_{T,T'} f^{T,T'}_{\alpha,\beta;\alpha',\beta'} (J^{\mu \dagger}_T)^{\alpha,\beta} (J_{\mu,T'})^{\alpha',\beta'}\,, \label{laeff1} \end{equation} with currents of the usual $\;V-A\;$ form: \begin{equation} (J^{T}_\mu)_{\alpha,\beta} = \bar{\psi}_{L,\alpha}^T \gamma_\mu \psi_{L,\beta}^T, \label{current2} \end{equation} and a group factor which, for reactions involving changes of flavour, reduces to: \begin{equation} f^{T,T'}_{\alpha,\beta;\alpha',\beta'} = S^{T*}_{3,\alpha} S^{T}_{3,\beta} S^{T'*}_{3,\alpha'} S^{T'}_{3,\beta'}, \label{groupfactor1} \end{equation} which is given entirely in terms of the matrices $S^T_{\alpha,\beta}$, so that the only remaining free parameter in (\ref{laeff1}) is the mass scale $\zeta z$. However, the effective Lagrangian (\ref{laeff1}) describes only interactions between quarks and leptons. To make contact with actual experiment on hadrons, we follow the usual procedures adopted in these contexts. For example, the effective action gives a contribution to the $K_L-K_S$ mass difference of the form: \begin{equation} \Delta m_K = \frac{1}{\zeta^2 \,\, z^2} |f^{D,D}_{2,3;2,3}| \langle K^0| \left[\bar{s}_L \gamma^\mu d_L \right]^2 |\bar{K}^0 \rangle. \end{equation} Evaluating the matrix elements in the vacuum insertion approximation one obtains: \begin{equation} \Delta m_K = \frac{1}{\zeta^2 \,\, z^2} |f^{D,D}_{2,3;2,3}|\frac{f_K^2 \, \, m_K}{3}, \label{delmK} \end{equation} where $f_K$ is the $K$ decay constant and $m_K$ is the $K$ mass. Mass differences between other charge conjugate neutral mesons are treated similarly. On the other hand, for hadron decays, in order to minimize the uncertainties in the hadron structure we take quotients between the rare and Standard Model-allowed processes which contain the same or similar hadronic matrix elements. For instance, for $K^+$ decays we take: \begin{equation} \frac {Br \left(K^+ \rightarrow \pi^+ l_{\alpha} l_{\beta} \right)} {Br \left(K^+ \rightarrow \pi^0 \nu_{\mu} \mu^+ \right)} = |f^{D,L}_{2,3;\alpha,\beta}|^2 \left( \frac{v}{\zeta \, \, z} \right)^4 \frac{2}{\sin^2 \theta_c}, \label{kplus} \end{equation} where $v=\frac{0.246}{\sqrt{2}}$ TeV and $\theta_c$ is the Cabibbo angle, $\sin \theta_c=0.23$. Similarly, for the leptonic decays of the neutral $K$-mesons, we take: \begin{equation} \frac {\Gamma \left(K^0_{L(S)} \rightarrow l_\alpha l_\beta \right)} {\Gamma \left(K^+ \rightarrow \mu^+ \nu_\mu \right)} = |f^{D,L}_{2,3;\alpha,\beta}|^2 \left( \frac{v}{\zeta \, \, z} \right)^4 \frac{1}{\sin^2 \theta_c}. \label{kls} \end{equation} Then from these formulae, given the total widths of the $K$'s and their widths in the Standard Model-allowed modes as measured in experiment, one can easily calculate the branching ratios of the various rare modes of $K$-decay. These procedures for dealing with the complexities of hadronic effects are of course far from foolproof but are likely to give the rough order of magnitudes correctly. All predictions we obtain in this way are still dependent on the single parameter $\zeta z$, so that without any further input we can give no numerical value for the predicted quantities. So long as $\zeta z$ remains undetermined, our predictions will of course lead to no violation of present experiment. However, given the experimental bound on any one quantity, a bound on $\zeta z$ is implied, which will then allow us to give the correlated bound on all the others. The most stringent lower bound on $\zeta z$ obtained in this way turns out to be that from the experimentally measured $K_L-K_S$ mass difference, namely $\Delta m_K(K_L-K_S) =3.5 \,\times \, 10^{-12}\mbox{ MeV}$ which is of roughly the right order of magnitude expected from second order weak interactions. Requiring that the FCNC effect due to dual gauge boson exchange be smaller than this value gives the bound \cite{ourfcnc}: \begin{equation} \zeta \,\, z \geq 400\mbox{ TeV}. \label{bound} \end{equation} The correlated (upper) bounds on other FCNC effects due to dual gauge boson exchange can then be estimated. As will be seen in the next section, there is a possible upper bound on the parameter $\zeta z$ coming from a rather unexpected angle which turns out to be similar to the lower bound quoted in (\ref{bound}). If that is the case, then the above bounds for FCNC effects can be treated as actual order of magnitude estimates. For $\zeta z =$ 400 TeV, the predicted branching ratios \cite{ourfcnc} of some rare $K$-decay modes are given in Table \ref{Kdecaytab} and compared with the experimental limits/measurements \cite{databook}. One notes that most of the predictions are way beyond the present experimental sensitivity, while some others, such as $K_L \rightarrow e^+ e^-, \mu^+ \mu^-$, can also go by second order weak interactions which are expected to give similar or even somewhat larger contributions and so will overshadow the present predicted effects. Only the mode $K_L \rightarrow e^\pm \mu^\mp$, which is inaccessible to second order weak interactions unless neutrinos mix, is interesting in having a predicted branching ratio less than two orders of magnitude down from the present experimental limits \cite{brookhaven} and so may be accessible in the near future. Its observation at this level may be considered as a confirmation either of the DSM prediction or that neutrinos mix and hence of interest in either case. \begin{table} \begin{eqnarray*} \begin{array}{||l|l|l||} \hline \hline & Theory & Experiment \\ \hline Br(K^+ \rightarrow \pi^+ e^+ e^-) & 4 \times 10^{-15} & 2.7 \times 10^{-7} \\ Br(K^+ \rightarrow \pi^+ \mu^+ \mu^-) & 2 \times 10^{-15} & 2.3 \times 10^{-7} \\ Br(K^+ \rightarrow \pi^+ e^+ \mu^-) & 2 \times 10^{-15} & 7 \times 10^{-9} \\ Br(K^+ \rightarrow \pi^+ e^- \mu^+) & 2 \times 10^{-15} & 2.1 \times 10^{-10} \\ Br(K^+ \rightarrow \pi^+ \nu {\bar \nu}) & 2 \times 10^{-14} & 2.4 \times 10^{-9} \\ Br(K_L \rightarrow e^+ e^-) & 2 \times 10^{-13} & 4.1 \times 10^{-11} \\ Br(K_L \rightarrow \mu^+ \mu^-) & 7 \times 10^{-14} & 7.2 \times 10^{-9} \\ Br(K_L \rightarrow e^{\pm} \mu^{\mp}) & 1 \times 10^{-13} & 5.1 \times 10^{-12} \\ Br(K_S \rightarrow \mu^+ \mu^-) & 1 \times 10^{-16} & 3.2 \times 10^{-7}\\ Br(K_S \rightarrow e^+ e^-) & 3 \times 10^{-16} & 2.8 \times 10^{-6} \\ \hline \hline \end{array} \end{eqnarray*} \caption{Branching ratios for rare leptonic and semileptonic $K$ decays. The first column shows the DSM predictions from one-dual gauge boson exchange with the lowest v.e.v. $\zeta z$ of the Higgs fields taken at 400 TeV. The second column gives either the present experimental limits on that process if not yet observed or the actual measured value for that process. In the latter case, it means that the process can go by other mechanisms such as second-order weak so that our predictions with dual gauge boson exchange will appear as corrections to these. Except for the entry for the decay $K_L \rightarrow e \mu$ from \cite{brookhaven} mentioned in the text, the other entries are from the databook \cite{databook}.} \label{Kdecaytab} \end{table} Similar tables have been compiled for rare $D$ and $B$ meson decays but the predicted branching ratios are in all cases much below the present experimental sensitivity and therefore not of immediate interest. Mass differences between the conjugate neutral $D$ and $B$ meson pairs are given in a similar way to that for the $K$'s. The contribution of one-dual gauge boson exchange to the mass splitting in $D$ is \cite{ourfcnc}: \begin{equation} {\Delta m_D} = \frac{m_D}{\zeta^2 z^2} \frac{f_D^2}{3} |f^{U,U}_{2,3;2,3}|. \label{Dmassdiff} \end{equation} Taking the values $f_D^2=10^{-8}\mbox{ TeV}^2$ for the $D$-meson coupling and $m_D=1.865$ GeV for the mass we have: \begin{equation} {\Delta m_D} = 5 \times 10^{-12}\mbox{ MeV}. \label{Dmassnos} \end{equation} This is one-and-a-half orders of magnitude off the present experimental limits ${\Delta m_D}\leq 1.4 \times 10^{-10}$ MeV and could be accessible to planned experiments in the near future. Applying the same procedure to the mass-splitting between the neutral $B$-mesons, one finds that the contribution from dual gauge boson exchange is 6 orders of magnitude smaller than that from the Standard Model and thus not likely to be accessible. We conclude therefore, for this section, that the DSM predictions on low energy FCNC effects so far made do not seem in contradiction with any existing experiment, and that for a couple of cases, namely $K_L \rightarrow e^\pm \mu^\mp$ and $\Delta m(D-\bar{D})$, apart of course from $\Delta m(K_L-K_S)$ itself, they may be testable in the near future. \setcounter{equation}{0} \section{Air Showers beyond the GZK Cut-off} At energies higher than the mass scale of the dual gauge bosons, FCNC effects will no longer be suppressed and, given the strength of their couplings $\tilde{g_i}$ as governed by the Dirac quantization conditions (\ref{couplings}), the interactions due to their exchange will become strong. Hence, DSM would predict a new strong interaction at ultra-high energies for all particles carrying a generation index. In particular, even the neutrinos corresponding to the usual 3 generations of charged leptons will acquire strong interactions at high energies. At first sight, this seems alarming until one recalls from the estimate given in the last section for the mass scale involved of order $\zeta z > 400\ {\rm TeV}$, which is way beyond anything that has been achieved in terrestial experiments or can be achieved in the foreseeable future. Nor are such energies accessible to astrophysical or cosmological considerations except in the very early universe. There is in fact only one instance known to us that energies of that order have been experimentally observed, namely in air showers produced by cosmic rays with energy beyond the Greisen-Zatsepin-Kuz'min (GZK) cutoff. Air showers with energy $E > 10^{20}\ {\rm eV}$ \cite{Volcano}--\cite{Auger}, though rare in occurrence, pose a long-standing and intriguing question of fundamental physical interest. High energy air showers are usually thought to be due to protons, but protons with such an energy would interact with the 2.7 K microwave background via, for example, the reaction: \begin{equation} p + \gamma_{2.7 K} = \Delta + \pi, \label{piprod} \end{equation} and degrade quickly in energy. Indeed, according to Greisen, Zatsepin and Kuz'min \cite{Greisen,Zatsemin}, the cosmic ray spectrum for protons should cut off sharply at around $5 \times 10^{19}\ {\rm eV}$ (the GZK cut-off) if they come from further than 50 Mpc away. And within such distances, there does not seem to be any likely source for producing particles of so high an energy. One possible solution would be that these post-GZK air showers are produced not by protons but by some other (stable, electrically neutral) particles which would not be so degraded in energy by the microwave background. The possibility can thus be considered that they are produced by neutrinos, which is feasible, of course, only if for some reason neutrinos acquire at high energy a new strong interaction, for otherwise they would not interact sufficiently with air nuclei to produce air showers. But this is exactly what is predicted by DSM as proposed in the preceding paragraph. So, it would appear that this prediction not only may escape contradiction with existing experiment as one might at first have feared, but may even offer an explanation for the long-standing puzzle of air showers beyond the GZK cutoff \cite{airshower1,airshower2,airshower3}. However, a strong interaction for neutrinos, though necessary, is by itself insufficient to guarantee a large cross section with air nuclei, which is needed for them to produce air showers in the atmosphere, since whatever the strength of the interaction, the cross section will remain small if the interaction is short-ranged. Now, the dual gauge bosons in DSM being supposedly very heavy, it looks at first sight that the interactions they mediate will be short-ranged and therefore not lead to large high energy cross sections for neutrinos. But this need not be the case for the following reason. The relation of dual colour to colour is similar to that of magnetism to electricity in electrodynamics which has only the physical degrees of freedom corresponding to the one photon, despite having two separate, electric {\it and} magnetic, gauge symmetries, as explained in Section 1. Hence, dual colour, though a different gauge degree of freedom to colour, represents just the same physical degree of freedom as colour \cite{dualsymm}. As a result, it is argued \cite{airshower2} that a dual gluon can `metamorphose' into a gluon in hadronic matter, thus giving neutrinos at high energy an interaction of hadronic range inside the nucleus. They will then interact coherently with the air nucleus and acquire with it a hadronic size cross section.\footnote{There has appeared a paper by Burdman, Halzen and Gandhi \cite{Halzen} subsequent to \cite{airshower2} claiming, among other things, that neutrinos cannot on general grounds acquire hadronic size cross sections. We think that their claim is ill-founded. Their arguments used only first order perturbation theory which is far from adequate for hadron reactions which are notoriously nonperturbative in character. Indeed, with their arguments, one would be unable to deduce that protons have hadronic cross sections. They also claimed their conclusions to be a consequence of s-wave unitarity but gave neither justifications nor references to substantiate this claim, and a repeated effort by one of us (CHM) in correspondence with Francis Halzen has not produced any clarification. We do not think s-wave unitarity can constrain the cross section the way they claim it does since high energy cross sections involve all partial waves. For more details of this discussion see \cite{airshower3,ourfcnc}.} This assertion is admittedly rather conjectural. If it fails, then the suggested explanation for post-GZK showers no longer stands, but the assertion that the prediction of strong neutrino interactions at high energy contradicts no existing experiment still remains valid. Since this suggested explanation for post-GZK air showers depends crucially on the assumed identification of generations with dual colour, it is worth examining its feasibility in some detail. Suppose that neutrinos do acquire strong interactions and a large enough cross section with air nuclei to produce air showers at high energy. The energy at which this begins to happen, according to DSM, is given by the scale estimated before to be $>$ 400 TeV in the centre of mass. For a neutrino impinging on a nucleon at rest in the atmosphere, this corresponds to a primary energy of around $8 \times 10^{19}\ \rm{eV}$, namely just above the GZK cut-off, exactly the sort of energy one wants. That being the case, let us now examine in more detail whether the hypothesis can accommodate the few observed facts known about the post-GZK showers, most of which have difficulties in being explained by protons as the primary particle. (A) First, one asks how neutrinos at such a high energy can be produced. One does not actually know at present a truly realistic mechanism even for protons, but according to Hillas \cite{Hillas} one can at least write down the condition that a source must satisfy in order to produce such energetic particles: \begin{equation} BR > E/Z, \label{Hillas} \end{equation} where $B$ is the magnetic field in $\mu$G, $R$ the size of the source in kpc, $E$ the energy in EeV = $10^{18}\ {\rm eV}$ and $Z$ the charge of the particle. There are only a few types of objects known which satisfy this condition, namely neutron stars, radio galaxies and active galactic nuclei. Of these, both the neutron stars and active galactic nuclei are surrounded by strong electromagnetic fields. The difficulty with protons is that even if the source can accelerate them to the required high energy, they would not be able to escape from the intense fields surrounding the source. However, there does not seem to be the same difficulty with neutrinos. By hypothesis based on the DSM, neutrinos interact strongly at high energy so that any source capable of accelerating protons to these energies will be able also copiously to produce neutrinos by say proton-proton collisions. Once produced, however, neutrinos will not be deterred by the intense e.m. fields and will be able to escape where protons fail. (B) Secondly, one asks whether neutrinos will be able to survive a long journey through the 2.7 K microwave background. There is no problem, for in colliding with a (massless) photon at this temperature, a neutrino even at $10^{20}\ {\rm eV}$ will produce only about 200 MeV C.M. energy, and at this energy a neutrino has still only weak interactions. The same applies also to its collision with the neutrino background in the intervening space. (C) Thirdly, one asks whether a neutrino when it arrives on earth will be able to produce air showers with the observed angular and depth distributions. Neutrinos with only weak interactions will have immense penetrating power, and even if an enormous neutrino flux is assumed sufficient to produce air showers at these rare occasions, the showers will be mostly horizontal and have a near constant distribution in depth. This is in contradiction to what is observed for post-GZK showers which are mostly near vertical and occur in the upper atmosphere. However, once neutrinos are ascribed a hadronic cross section with air nuclei, they will interact like hadrons and both the angular and depth distributions will automatically fall into place. (D) It was noted \cite{Hayashida} that out of the few post-GZK events seen, three pairs coincide in incident angles to within the experimental error of about $2^\circ$. The probability of this occurring at random is very small and the obvious conclusion would be that the two members of each pair originate from the same source. They have however different energies and if they are protons should be deflected differently by the intervening magnetic fields and hence arrive at different angles, contrary to what is observed. If they are neutrinos, on the other hand, they will not be deflected by magnetic fields and will arrive on earth in the same direction they started out. (E) The highest energy event at $3 \times 10^{20}\ {\rm eV}$ observed by the Fly's Eye \cite{Flyseye} was noted to point in the direction of a very powerful Seyfert galaxy 900 Mpc away \cite{Elmers}. If that is taken to be its source and if it is due to a proton, one would wonder why many more showers with lower energies are not observed from the same direction, for a powerful source capable of accelerating protons to such a high energy would surely also produce protons at lower energies as well. For neutrinos interacting strongly only at high energy, this is not a problem. At low energies, neutrinos are weakly interacting and would first of all not be produced at source, and even if produced would not give rise to air showers when they arrive on earth. It thus seems that the neutrino hypothesis has survived all the above tests (A)--(E) on post-GZK showers which pose difficulties for their having been produced by protons. In spite of this, however, the hypothesis must clearly be subjected to many more quantitative tests before it can be taken seriuosly. Fortunately, some such tests \cite{airshower2,ourfcnc} are available, as follows. (I) If we accept our previous argument that a neutrino at high energy would interact not only strongly but coherently with an air nucleus, it is easy to deduce from a geometric picture that the cross section of a neutrino with the nucleus would be about half that of a proton. To both the neutrino and the proton, the nucleus would appear as a black disc, say of radius $r_A$. The neutrino, with as yet no known internal structure, would appear to the nucleus still as a point, but the proton will appear again as a black disc of radius, say, $r_p$. One concludes thus that the (geometric) cross section of the neutrino with the nucleus is roughly: \begin{equation} \sigma_T(\nu A) = \pi r_A^2, \label{sigmanu} \end{equation} while that of the proton with the nucleus is roughly: \begin{equation} \sigma_T(p A) = \pi (r_p + r_A)^2. \label{sigmap} \end{equation} Assuming that: \begin{equation} r_A \sim A^{1/3} r_p, \label{rA} \end{equation} with $A$ around 15 for an air nucleus, one easily deduces the above estimate: \begin{equation} \frac{\sigma_T(\nu A)}{\sigma_T(p A)} \sim \frac{1}{2}. \label{sigmaratio} \end{equation} This means that neutrinos at post-GZK energies are expected to be about twice as penetrating as protons, and hence that neutrino-produced air showers will occur at a lower depth on the average than proton-produced air showers. Folding in the air density as a function of height, it is easy to evaluate the penetration probability as a function of depth. This was done in \cite{airshower2} which finds that: \begin{eqnarray} {\rm Most \; probable \; height \; of} \; p{\rm -produced \; showers} & \sim & {\rm 21 \; km}, \nonumber \\ {\rm Most \; probable \; height \; of} \; \nu{\rm -produced \; showers} & \sim & {\rm 15 \; km}. \label{showerdepth} \end{eqnarray} Hence, one predicts that post-GZK air showers which are supposedly produced by neutrinos would occur most probably at a height of only around 15 km, in contrast to lower energy showers produced by protons which would occur most probably at a height of around 21 km in our atmosphere. Present detectors do not locate the primary vertices of air showers readily. For this reason, we have found up to the present only one tentative piece of information for testing this prediction. The development profile of the highest energy event obtained by the Fly's Eye shows that light began to be observed at a (vertical equivalent) height of around 12 km. If we interpret this as the primary vertex for the event, then it is much more likely, according to the preceding arguments, to be a neutrino-produced shower than a proton-produced one, for which the probability is estimated to be less than 5 \%. This conclusion should not as yet be taken too seriously, but with new projects such as Auger \cite{Auger}, capable of collecting sizeable statistics, this could be a very useful test for the hypothesis that post-GZK showers are neutrino-produced. (II) As far as particle physics proper is concerned, the post-GZK air shower events, if interpreted as due to neutrinos, are useful in providing a rough upper bound to the dual gauge boson mass. Translated to the language of Section 5, this means an upper bound on the parameter $\zeta z$ of around 500 TeV \cite{ourfcnc}, remarkably close to the lower bound of around 400 TeV obtained there from the $K_L-K_S$ mass difference. Acceptance of this upper bound then converts the bounds estimated in Section 5 on rare meson decays and mass differences into actual order of magnitude predictions, and hence affords a second test for the neutrino hypothesis here for post-GZK air showers. \setcounter{equation}{0} \section{Concluding Remarks} The basic tenets and applications to-date of the DSM scheme are summarized in the flow chart of Figure \ref{flowchart}, which shows that starting from a previously derived result of nonabelian duality, one is led on the one hand to a calculation of some of the Standard Model's fundamental parameters, and on the other to new testable predictions ranging from FCNC effects at low energy to air showers from cosmic rays at the extreme end of the detected energy scale. \begin{figure}[htb] \center \begin{picture}(240,400) \put(60,350){\framebox(120,18){\shortstack{{\footnotesize NONABELIAN DUALITY}\\ {\tiny{Chan--Faridani--Tsou \cite{dualsymm,dualsym}}}}}} \put(116,340){\line(1,0){8}} \put(120,344){\line(0,-1){8}} \put(-10,320){\framebox(120,10){{\footnotesize STANDARD MODEL}}} \put(130,320){\framebox(130,10){{\footnotesize 'T HOOFT THEOREM \cite{tHooft}}}} \put(55,320){\line(0,-1){8}} \put(185,320){\line(0,-1){8}} \put(55,312){\line(1,0){130}} \put(120,312){\vector(0,-1){12}} \put(40,282){\framebox(160,18){\shortstack{{\footnotesize DUALIZED STANDARD MODEL}\\ {\tiny{Chan--Tsou \cite{dualcons}}}}}} \put(120,282){\line(0,-1){10}} \put(55,272){\line(1,0){130}} \put(55,272){\vector(0,-1){12}} \put(185,272){\vector(0,-1){12}} \put(-10,240){\framebox(120,20){\shortstack{{\footnotesize HIGGS FIELDS}\\ {\footnotesize (AS FRAME VECTORS)}}}} \put(130,240){\framebox(120,20){\shortstack{{\footnotesize 3 GENERATIONS}\\ {\footnotesize (AS DUAL COLOUR)}}}} \put(30,240){\vector(0,-1){18}} \put(160,240){\line(0,-1){6}} \put(80,234){\line(1,0){80}} \put(80,234){\vector(0,-1){12}} \put(-10,202){\framebox(120,20){\shortstack{{\footnotesize SYMMETRY}\\ {\footnotesize BREAKING PATTERN}}}} \put(55,202){\vector(0,-1){16}} \put(-10,146){\framebox(166,40){\shortstack{{\footnotesize FERMION MASS HIERARCHY:}\\ e.g. $m_t \gg m_c \gg m_u$\\{\footnotesize FERMION MASSES AND MIXINGS}\\ {\footnotesize CALCULABLE PERTURBATIVELY}}}} \put(55,146){\vector(0,-1){16}} \put(-10,84){\framebox(166,46){\shortstack{{\footnotesize (1-LOOP, 1ST ORDER)}\\ {\footnotesize QUARK AND LEPTON MASSES}\\{\footnotesize QUARK CKM MATRIX}\\ {\tiny{Bordes--Chan--Faridani--Pfaudler--Tsou \cite{ourckm}}}}}} \put(106,84){\line(0,-1){10}} \put(106,74){\line(1,0){74}} \put(180,74){\vector(0,-1){19}} \put(55,84){\vector(0,-1){54}} \put(-10,0){\framebox(166,30){\shortstack{{\footnotesize (EXTENDED TO NEUTRINOS)}\\ {\footnotesize NEUTRINO OSCILLATIONS}\\ {\tiny{Bordes--Chan--Pfaudler--Tsou \cite{ournuos,features}}}}}} \put(210,240){\vector(0,-1){24}} \put(176,176){\framebox(84,40){\shortstack{{\footnotesize NEUTRINOS}\\ {\footnotesize INTERACT}\\{\footnotesize STRONGLY AT}\\ {\footnotesize HIGH ENERGY}}}} \put(210,176){\vector(0,-1){16}} \put(174,90){\framebox(86,70){\shortstack{{\footnotesize `EXPLANATION'}\\ {\footnotesize FOR PUZZLE OF}\\{\footnotesize AIR SHOWERS}\\ {\footnotesize AT $E>10^{20}$eV}\\ {\tiny{Bordes--Chan--}}\\{\tiny{Faridani--Pfaudler--}}\\ {\tiny{Tsou \cite{airshower3,airshower2}}}}}} \put(210,90){\vector(0,-1){35}} \put(176,0){\framebox(84,55){\shortstack{{\footnotesize PREDICTIONS}\\ {\footnotesize OF B.R.\ FOR}\\{\footnotesize FCNC DECAYS}\\ {\tiny{Bordes--Chan--}}\\{\scriptsize{Faridani--Pfaudler--}}\\ {\tiny{Tsou \cite{ourfcnc}}}}}} \end{picture} \caption{Summary flow-chart} \label{flowchart} \end{figure} One of the most attractive features of DSM is undoubtedly its offer of a possible explanation for the existence both of exactly 3 fermion generations and of scalar Higgs fields. In the conventional formulation of the Standard Model, the necessity to introduce by hand both of these, neither having any known geometrical significance, must be regarded as rather a blotch on a gauge theory otherwise so beautifully founded on geometry. The identification thus of generations as dual colour and of Higgs fields as frame vectors in internal symmetry space, giving each a geometrical significance, seems very attractive. Besides, according to \cite{dualsymm}, nonabelian duality is an intrinsic property of the Standard Model (as of any gauge theory) which then brings with it automatically a 3-fold broken dual colour symmetry and frame vectors in internal symmetry space playing a dynamical role. In other words, the niches for 3 fermion generations and Higgs fields already `pre-exist' in the Standard Model. Hence, it seems appropriate to assign them to just these features we see in nature, because even if we do not, we shall still have to account for them physically in some other way. Implementing these identifications with some seemingly natural assumptions as detailed in Section 1, one is then led to a scheme with a hierarchical fermion mass spectrum where the mixings between fermion types and lower generation masses are calculable as loop corrections in terms of a few parameters. The present score from the 1-loop calculation carried out to-date is as follows. By adjusting 3 parameters, namely the (common) Yukawa coupling strength $\rho$ and the 2 ratios between the 3 vev's $(x,y,z)$ of the dual colour Higgs fields, one has calculated the following 14 among the 26 or so of the Standard Model's fundamental parameters: the 3 independent parameters in the quark CKM matrix $|V_{rs}|$, the 3 corresponding parameters in the leptonic CKM matrix $|U_{rs}|$, and the 8 masses $m_c, m_s, m_\mu, m_u, m_d, m_e, m_{\nu_1}, B$. Of these 14 calculated quantities, 2 (i.e.\ $m_u$, $U_{e2}$) compare unsatisfactorily at present with experiment, and another 2 (i.e. the mass of the lightest neutrino $m_{\nu_1}$ and that of the right-handed neutrino $B$) are untested being experimentally yet unknown. The other 10, however, (namely, $|V_{rs}|, |U_{e3}|, |U_{\mu3}|, m_c, m_s, m_\mu, m_d, m_e$), all agree as well as can be expected with their known empirical values. This, we think, is not a bad score for a first attempt based on some rather crude approximations, such as taking $\rho$ and $m_T$ as scale-independent constants. With more experience and sophistication, the score can possibly be improved. However, even if considered successful, this score by itself does not constitute a stringent test for the basic assumption of DSM that dual colour is generations. As emphasized in a recent paper \cite{phenodsm}, the same result can be obtained just by assuming generations to be a broken $U(3)$ symmetry independent of whether it is identified with dual colour. The only physical consequence considered in this paper which relies crucially on that identification is the explanation suggested in Section 6 for air showers beyond the GZK cut-off, which is still far from established. An urgent task for this scheme is thus to device some further tests for the dual colour hypothesis. Besides this, there are many further questions needing answers for checking the consistency of DSM, both within itself and with nature. Of these we list in particular the following. First, there is the question of CP-violation which, though known experimentally, has not yet made an appearance in DSM at the 1-loop level and might indicate a deficiency. Secondly, there is the question of the rotating mass matrix, which one has made use of in the calculation of mass and mixing parameters, and this might have other physical consequences yet waiting to be explored. Thirdly, there is the intriguing question of the `accidental' near equality between Yukawa coupling strengths $\rho$ for all fermion types, and the proximity of the Higgs vev's $(x, y, z)$ to the fixed point $(1,0,0)$, which presumably reflect a deeper intricacy in the problem than we have yet understood. Fourthly, there is the question of linkage between the breaking of the dual colour symmetry, studied so far in isolation, with the breaking of the electroweak symmetry, which may be related to the point raised immediately above. Fifthly, there is the subtle question of whether the duality assertion that gauge and dual gauge bosons represent the same physical degrees of freedom might give rise to a new class of phenomena, called `metamorphosis' by us in \cite{dualcons}, of which post-GZK air showers are but one example of many possible manifestations. Sixthly, going even further afield, there is the question of the symmetry $\widetilde{SU}(2)$ dual to electroweak $SU(2)$, which by the same logic adopted for $SU(3)$ colour here, ought to give rise to another level of confinement deeper than colour, and this should be amenable to experimental investigation, but only by deep inelastic scattering at ultra-high energies. And there will be other questions too which we have not yet learned even to formulate. One has thus the feeling that what has been attempted so far is but scratching the surface of a possibly very rich vein. \vspace{.5cm} \noindent {\large {\bf Acknowledgement}} \vspace{.2cm} It is a pleasure for us to thank Jos\'e Bordes, Jacqueline Faridani and Jakov Pfaudler for a most enjoyable and fruitful collaboration producing most of the work reported above. \clearpage
\chapter{Whittaker model}\label{Witt} In this chapter we recall the Whittaker model of the center of the universal enveloping algebra $U({\frak g})$, where ${\frak g}$ is a complex simple Lie algebra. \section{Notation}\label{notation} Fix the notation used throughout of the text. Let $G$ be a connected simply connected finite--dimensional complex simple Lie group, $% {\frak g}$ its Lie algebra. Fix a Cartan subalgebra ${\frak h}\subset {\frak % g}\ $and let $\Delta $ be the set of roots of $\left( {\frak g},{\frak h}% \right) .$ Choose an ordering in the root system. Let $\alpha_i,~i=1,\ldots l,~~l=rank({\frak g})$ be the simple roots, $\Delta_+=\{ \beta_1, \ldots ,\beta_N \}$ the set of positive roots. Denote by $\rho$ a half of the sum of positive roots, $\rho=\frac 12 \sum_{i=1}^N\beta_i$. Let $H_1,\ldots ,H_l$ be the set of simple root generators of $\frak h$. Let $a_{ij}$ be the corresponding Cartan matrix. Let $d_1,\ldots , d_l$ be coprime positive integers such that the matrix $b_{ij}=d_ia_{ij}$ is symmetric. There exists a unique non--degenerate invariant symmetric bilinear form $\left( ,\right) $ on ${\frak g}$ such that $(H_i , H_j)=d_j^{-1}a_{ij}$. It induces an isomorphism of vector spaces ${\frak h}\simeq {\frak h}^*$ under which $\alpha_i \in {\frak h}^*$ corresponds to $d_iH_i \in {\frak h}$. We denote by $\alpha^\vee$ the element of $\frak h$ that corresponds to $\alpha \in {\frak h}^*$ under this isomorphism. The induced bilinear form on ${\frak h}^*$ is given by $(\alpha_i , \alpha_j)=b_{ij}$. Let $W$ be the Weyl group of the root system $\Delta$. $W$ is the subgroup of $GL({\frak h})$ generated by the fundamental reflections $s_1,\ldots ,s_l$, $$ s_i(h)=h-\alpha_i(h)H_i,~~h\in{\frak h}. $$ The action of $W$ preserves the bilinear form $(,)$ on $\frak h$. We denote a representative of $w\in W$ in $G$ by the same letter. For $w\in W, g\in G$ we write $w(g)=wgw^{-1}$. Let ${{\frak b}_+}$ be the positive Borel subalgebra and ${\frak b}_-$ the opposite Borel subalgebra; let ${\frak n}_+=[{{\frak b}_+},{{\frak b}_+}]$ and $% {\frak n}_-=[{\frak b}_-,{\frak b}_-]$ be their nil-radicals. Let $H=\exp {\frak h},N_+=\exp {{\frak n}_+}, N_-=\exp {\frak n}_-,B_+=HN_+,B_-=HN_-$ be the Cartan subgroup, the maximal unipotent subgroups and the Borel subgroups of $G$ which correspond to the Lie subalgebras ${\frak h},{{\frak n}_+},% {\frak n}_-,{\frak b}_+$ and ${\frak b}_-,$ respectively. We identify $\frak g$ and its dual by means of the canonical invariant bilinear form. Then the coadjoint action of $G$ on ${\frak g}^*$ is naturally identified with the adjoint one. We also identify ${{\frak n}_+}^*\cong {\frak n}_-,~{{\frak b}_+}^*\cong {\frak b}_-$. Let ${\frak g}_\beta$ be the root subspace corresponding to a root $\beta \in \Delta$, ${\frak g}_\beta=\{ x\in {\frak g}| [h,x]=\beta(h)x \mbox{ for every }h\in {\frak h}\}$. ${\frak g}_\beta\subset {\frak g}$ is a one--dimensional subspace. It is well--known that for $\alpha\neq -\beta$ the root subspaces ${\frak g}_\alpha$ and ${\frak g}_\beta$ are orthogonal with respect to the canonical invariant bilinear form. Moreover ${\frak g}_\alpha$ and ${\frak g}_{-\alpha}$ are non--degenerately paired by this form. Root vectors $X_{\alpha}\in {\frak g}_\alpha$ satisfy the following relations: $$ [X_\alpha,X_{-\alpha}]=(X_\alpha,X_{-\alpha})\alpha^\vee. $$ If $V$ is a finite--dimensional complex vector space, $S(V)$ will denote the symmetric algebra over $V$ and $S_k(V)$ denotes the homogeneous subspace of degree k. If $V^*$ is the dual space to $V$ then $S(V^*)$ is regarded as the algebra of polynomial functions on $V$. Let $U({\frak g})$ be the universal enveloping algebra of $\frak g$, and $U_k({\frak g})$ the standard filtration in $U({\frak g})$. From the Poincar\'{e}--Birkhoff--Witt theorem it follows that the associated graded algebra $GrU({\frak g})$ is isomorphic to the symmetric algebra $S({\frak g})$ of the linear space $\frak g$. Equip $S({\frak g})$ with a Poisson structure as follows. For each $s_k\in S_k({\frak g})$ choose a representative $u_k\in U_k({\frak g})$ such that $u_k/ U_{k-1}({\frak g})=s_k$. We shall denote $s_k=Gru_k$. Given two such elements $s_i$ and $s_j$ with chosen representatives $u_i$ and $u_j$, the commutativity of $S({\frak g})$ implies that $$ [u_i,u_j]\in U_{i+j-1}({\frak g}). $$ Define \begin{equation}\label{KK} \{ s_i,s_j\} =[u_i,u_j]/ U_{i+j-2}({\frak g}). \end{equation} It is easy to see that this bracket is independent of the choice of representatives $u_i,~u_j$ and equips $S({\frak g})$ with the structure of a Poisson algebra, i.e. it is a derivation of the multiplication in $S({\frak g})$. We refer to the procedure described above as the graded limit. \section{The Whittaker model}\label{whitt} \setcounter{equation}{0} \setcounter{theorem}{0} In this section we introduce the Whittaker model of the center of the universal enveloping algebra $U({\frak g})$. We start by recalling the classical result of Chevalley which describes the structure of the center. Let $Z({\frak g})$ be the center of $U({\frak g})$. The standard filtration $U_k({\frak g})$ in $U({\frak g})$ induces a filtration $Z_k({\frak g})$ in $Z({\frak g})$. The following important theorem may be found for instance in \cite{Bur1}, Ch.8, \S 8, no. 3, Corollary 1 and no.5, Theorem 2. \vskip 0.3cm \noindent {\bf Theorem (Chevalley)} {\em One can choose elements $I_k\in Z_{m_k+1}({\frak g}),~~k=1,\ldots l$, where $m_k$ are called the exponents of $\frak g$, such that $Z({\frak g})={\Bbb C}[I_1,\ldots , I_l]$ is a polynomial algebra in $l$ generators.} \vskip 0.3cm The adjoint action of $G$ on $\frak g$ naturally extends to $S({\frak g})$. Let $S({\frak g})^G$ be the algebra of $G$--invariants in $S({\frak g})$. Clearly, $GrZ({\frak g})\cong S({\frak g})^G$. In particular $S({\frak g})^G\cong {\Bbb C}[\widehat I_1,\ldots , \widehat I_l]$, where $\widehat I_i=Gr I_i,~i=1,\ldots ,l$. The elements $\widehat I_i,~i=1,\ldots ,l$ are called fundamental invariants. Following Kostant we shall realize the center $Z({\frak g})$ of the universal enveloping algebra $U({\frak g})$ as a subalgebra in $U({\frak b}_-)$. Let $$ \chi :{{\frak n}_+} \rightarrow {\Bbb C} $$ be a character of ${{\frak n}_+}$. Since ${{\frak n}_+}=\sum_{i=1}^l{\Bbb C}X_{\alpha_i} \oplus[{{\frak n}_+},{{\frak n}_+}]$ it is clear that $\chi$ is completely determined by the constants $c_i=\chi (X_{\alpha_i}),~i=1,\ldots ,l$ and $c_i$ are arbitrary. In \cite{K} $\chi$ is called non--singular if $c_i\neq 0$ for all $i$. Let $f=\sum_{i=1}^l X_{-\alpha_i}\in {\frak n}_-$ be a regular nilpotent element. From the properties of the invariant bilinear form (see Section \ref{notation}) it follows that $(f,[{{\frak n}_+},{{\frak n}_+}])=0,~~(f,X_{\alpha_i})=(X_{-\alpha_i},X_{\alpha_i})$, and hence the map $x\mapsto (f,x),~~x\in {{\frak n}_+}$ is a non--singular character of ${{\frak n}_+}$. Recall that in our choice of root vectors no normalization was made. But now given a non--singular character $\chi :{{\frak n}_+}\rightarrow {\Bbb C}$ we will say that $f$ corresponds to $\chi$ in case $$ \chi (X_{\alpha_i}) =(X_{-\alpha_i},X_{\alpha_i}). $$ Conversely if $\chi$ is non--singular there is a unique choice of $f$ so that $f$ corresponds to $\chi$. In this case $\chi (x)=(f,x)$ for every $x\in {{\frak n}_+}$. Naturally, the character $\chi$ extends to a character of the universal enveloping algebra $U({{\frak n}_+})$. Let $U_\chi ({{\frak n}_+})$ be the kernel of this extension so that one has a direct sum $$ U({{\frak n}_+})={\Bbb C}\oplus U_\chi ({{\frak n}_+}). $$ Since ${\frak g}={\frak b}_-\oplus {{\frak n}_+}$ we have a linear isomorphism $U({\frak g})=U({\frak b}_-)\otimes U({{\frak n}_+})$ and hence the direct sum \begin{equation}\label{maindec} U({\frak g})=U({\frak b}_-) \oplus I_\chi, \end{equation} where $I_\chi=U({\frak g})U_\chi ({{\frak n}_+})$ is the left--sided ideal generated by $U_\chi ({{\frak n}_+})$. For any $u\in U({\frak g})$ let $u^\chi\in U({\frak b}_-)$ be its component in $U({\frak b}_-)$ relative to the decomposition (\ref{maindec}). Denote by $\rho_\chi$ the linear map $$ \rho_\chi : U({\frak g}) \rightarrow U({\frak b}_-) $$ given by $\rho_\chi (u)=u^\chi$. Let $W({\frak b}_-)=\rho_\chi (Z({\frak g}))$. \vskip 0.3cm \noindent {\bf Theorem A (\cite{K}, Theorem 2.4.2)} {\em The map \begin{equation}\label{map} \rho_\chi : Z({\frak g}) \rightarrow W({\frak b}_-) \end{equation} is an isomorphism of algebras. In particular $$ W({\frak b}_-)={\Bbb C}[I_1^\chi ,\ldots , I_l^\chi ], ~~I_i^\chi=\rho_\chi(I_i),~~i=1,\ldots ,l $$ is a polynomial algebra in $l$ generators.} \vskip 0.3cm \noindent {\em Proof.} First, we show that the map (\ref{map}) is an algebra homomorphism. If $u,v\in Z({\frak g})$ then $u^\chi v^\chi \in U({\frak b}_-)$ and $$ uv-u^\chi v^\chi =(u-u^\chi )v+u^\chi (v-v^\chi ). $$ Since $(u-u^\chi )v=v(u-u^\chi )$ the r.h.s. of the last equality is an element of $I_\chi$. This proves $u^\chi v^\chi =(uv)^\chi$. By definition the map (\ref{map}) is surjective. We have to prove that it is injective. Let $U({\frak g})^{\frak h}$ be the centralizer of $\frak h$ in $U({\frak g})$. Clearly $Z({\frak g})\subseteq U({\frak g})^{\frak h}$. From the Poincar\'{e}--Birkhoff--Witt theorem it follows that every element $z\in U({\frak g})^{\frak h}$ may be uniquely written as $$ z=\sum_{p,q\in{\Bbb N}^N,<p>=<q>}X_{-\beta_1}^{p_1}\ldots X_{-\beta_N}^{p_N}\varphi_{p,q} X_{\beta_1}^{q_1}\ldots X_{\beta_N}^{q_N}, $$ where $<p>=\sum_{i=1}^r p_i \beta_i \in {\frak h}^*$ and $\varphi_{p,q} \in U({\frak h})$. Now recall that $\chi (X_{\beta_i})=0$ if $\beta_i$ is not a simple root, and we easily obtain $$ \rho_\chi (z)=\sum_{p,q\in{\Bbb N}^l,<p>=<q>\neq 0}X_{-\alpha_{k_1}}^{p_{j_1}}\ldots X_{-\alpha_{k_l}}^{p_{j_l}}\varphi_{p,q} \prod_{i=1}^lc_{k_i}^{q_{j_i}}+\varphi_{0,0}. $$ Let $z\in Z({\frak g})$. One knows that the map $$ Z({\frak g})\rightarrow U({\frak h}),~~z\mapsto \varphi_{0,0}, $$ called the Harich-Chandra homomorphism, is injective (see (c), p. 232 in \cite{Dix}). It follows that the map (\ref{map}) is also injective. \vskip 0.3cm \noindent {\bf Definition A} {\em The algebra $W({\frak b}_-)$ is called the Whittaker model of $Z({\frak g})$.} \vskip 0.3cm Next we equip $U({\frak b}_-)$ with a structure of a left $U({{\frak n}_+})$ module in such a way that $W({\frak b}_-)$ is realized as the space of invariants with respect to this action. Let $Y_\chi$ be the left $U({\frak g})$ module defined by $$ Y_\chi =U({\frak g})\otimes_{U({{\frak n}_+})}{\Bbb C}_\chi , $$ where ${\Bbb C}_\chi$ denotes the 1--dimensional $U({{\frak n}_+})$--module defined by $\chi$. Obviously $Y_\chi$ is just the quotient module $U({\frak g})/I_\chi$. From (\ref{maindec}) it follows that the map \begin{equation}\label{iso2} U({\frak b}_-)\rightarrow Y_\chi;~~v\mapsto v\otimes 1 \end{equation} is a linear isomorphism. It is convenient to carry the module structure of $Y_\chi$ to $U({\frak b}_-)$. For $u\in U({\frak g}),~~ v\in U({\frak b}_-)$ the induced action $u\circ v$ has the form \begin{equation}\label{indact} u\circ v=(uv)^\chi. \end{equation} The restriction of this action to $U({{\frak n}_+})$ may be changed by tensoring with 1--dimensional $U({{\frak n}_+})$--module defined by $-\chi$. That is $U({\frak b}_-)$ becomes an $U({{\frak n}_+})$ module where if $x\in {U({\frak n}_+)},~~v\in U({\frak b}_-)$ one puts \begin{equation}\label{mainact} x\cdot v=x\circ v-\chi (x)v. \end{equation} \vskip 0.3cm \noindent {\bf Lemma A (\cite{K}, Lemma 2.6.1.)} {\em Let $v\in U({\frak b}_-)$ and $x\in {U({\frak n}_+)}$. Then} $$ x\cdot v =[x,v]^\chi. $$ \vskip 0.3cm \noindent {\em Proof.} By definition $x\cdot v=(xv)^\chi -\chi (x)v$. Then we have $xv=[x,v]+vx$ and hence $x\cdot v=([x,v])^\chi +(vx)^\chi -\chi (x)v$. But clearly $(vx)^\chi =v\chi (x)$. Thus $x\cdot v=([x,v])^\chi$. The action (\ref{mainact}) may be lifted to an action of the unipotent group $N_+$. Consider the space $U({\frak b}_-)^{N_+}$ of $N_+$ invariants in $U({\frak b}_-)$ with respect to this action. Clearly, $W({\frak b}_-)\subseteq U({\frak b}_-)^{N_+}$ . \vskip 0.3cm \noindent {\bf Theorem B (\cite{K}, Theorems 2.4.1, 2.6)} {\em Suppose that the character $\chi$ is non--singular. Then the space of $N_+$ invariants in $U({\frak b}_-)$ with respect to the action (\ref{mainact}) is isomorphic to $W({\frak b}_-)$, i.e.} \begin{equation}\label{inv} U({\frak b}_-)^{N_+}\cong W({\frak b}_-). \end{equation} \vskip 0.3cm We shall prove Theorem B in the next section. \section{Geometric approach to the Whittaker model}\label{geomappr} \setcounter{equation}{0} \setcounter{theorem}{0} In this section we establish a relation between the Whittaker model and the geometry of the adjoint action of the corresponding Lie group. Denote the character of $S({{\frak n}_+})$ that equals to $\chi(x)$ for every $x\in {\frak n}_+$ by the same letter. Similarly to (\ref{maindec}) we have the following decomposition for $S({\frak g})$: $$ S({\frak g})=S({\frak b}_-)\oplus I_\chi^0, $$ where $I_\chi^0$ is the ideal in $S({\frak g})$ generated by the kernel of $\chi$. For any $s\in S({\frak g})$ let $s^\chi$ be its component in $S({\frak b}_-)$ relative to this decomposition. Now using Lemma A we define the graded limit of the action (\ref{mainact}). For $x\in {\frak n}_+$ and $s\in S({\frak b}_-)$ we put \begin{equation}\label{classact} x\cdot s=(\{ x,s\} )^\chi. \end{equation} This action may be lifted to an action of the unipotent group $N_+$ on $S({\frak b}_-)$. For $a\in N_+,~s\in S({\frak b}_-)$ this action is given by $$ a\cdot s=({\rm Ad}(a)(s))^\chi. $$ Observe that $S({\frak b}_-)$ is naturally identified with the algebra of polynomial functions on ${\frak b}_+$. We shall describe the space of invariants $S({\frak b}_-)^{N_+}$ using the induced action of $N_+$ on ${\frak b}_+$. To calculate this action it suffices to consider the restriction of the action (\ref{classact}) to linear functions. Let $s\in {\frak b}_-$ be such a function. Then for $a\in N_+$ $$ a\cdot s=({\rm Ad}(a)(s))^\chi =P_{{\frak b}_-}({\rm Ad}(a)(s))+\chi(P_{{\frak n}_+}({\rm Ad}(a)(s))), $$ where $P_{{\frak b}_-}$ and $P_{{\frak n}_+}$ are the projection operators onto ${\frak b}_-$ and ${\frak n}_+$, respectively, in the direct sum ${\frak g}={\frak b}_- +{{\frak n}_+}$. By the definition of the induced action we have $$ a\cdot s(s')=s(a^{-1}\cdot s'),\mbox{ for every }s'\in {{\frak b}_+}. $$ On the other hand $$ \begin{array}{l} a\cdot s(s')=(P_{{\frak b}_-}({\rm Ad}(a)(s)),s')+\chi(P_{{\frak n}_+}({\rm Ad}(a)(s)))=\\ ({\rm Ad}(a)(s),s')+({\rm Ad}(a)(s),f), \end{array} $$ where $f\in {\frak n}_-$ corresponds to $\chi$. Since the canonical bilinear form $(,)$ is Ad-invariant the last formula may be rewritten as: $$ a\cdot s(s')=(s,{\rm Ad}(a)^{-1}(s'+f))=s(P_{{\frak b}_+}({\rm Ad}(a)^{-1}(s'+f))), $$ where $P_{{\frak b}_+}$ is the projector onto ${{\frak b}_+}$ in the direct sum ${\frak g}={{\frak b}_+}+{\frak n}_-$. Finally observe that the subspace $f+{{\frak b}_+}$ is stable under the adjoint action of $N_+$. Therefore $P_{{\frak b}_+}({\rm Ad}(a)^{-1}(s'+f))={\rm Ad}(a)^{-1}(s'+f)-f$, and the induced action of $N_+$ on ${{\frak b}_+}$ takes the form: \begin{equation}\label{dualact} a\cdot s'={\rm Ad}(a)(s'+f)-f. \end{equation} Now the algebra of invariants $S({\frak b}_-)^{N_+}$ may be identified with a certain subalgebra in the algebra of functions on the quotient ${{\frak b}_+}/{N_+}$. The space ${{\frak b}_+}/{N_+}$ has a nice geometric description. Observe that $[f,{{\frak n}_+}]\subset {{\frak b}_+}$. Moreover, $[f,{{\frak n}_+}]$ is an ${\rm ad}_\rho$ stable subspace of ${\frak b}_+$. Since $\rho$ is a semi--simple element there exists an ${\rm ad}_\rho$ invariant stable subspace ${\frak s}\subseteq {{\frak b}_+}$ such that ${{\frak b}_+}={\frak s}+[f,{{\frak n}_+}]$ is a direct sum. By Theorem 8 and Remark 19' in \cite{K1} $\frak s$ is an $l$--dimensional subspace in ${{\frak n}_+}$. \vskip 0.3cm \noindent {\bf Theorem C (\cite{K}, Theorem 1.2 )} {\em The map $$ {N_+}\times {\frak s}\rightarrow {{\frak b}_+} $$ given by $(a,x)\mapsto a\cdot x$ is an isomorphism of affine varieties. Therefore the quotient space ${{\frak b}_+}/{N_+}$ is isomorphic to ${\frak s}$.} \vskip 0.3cm The linear space $\frak s$ naturally appears in the study of regular elements in $\frak g$. Recall that an element of $\frak g$ is called regular if its centralizer in $\frak g$ is of minimal possible dimension. Let $R$ be the set of regular elements in $\frak g$. Clearly, $R$ is stable under the adjoint action of $G$ and in fact $R$ is the union of all $G$ orbits in $\frak g$ of maximal dimension. \vskip 0.3cm \noindent {\bf Theorem D (\cite{K}, Theorem 1.1; \cite{K1}, Theorem 8)} {\em The affine space $f+{\frak s}$ is contained in $R$ and is a cross--section for the action of $G$ on $R$. That is every $G$--orbit in $\frak g$ of maximal dimension intersects $f+{\frak s}$ in one and only one point. Let $\widehat I_1,\ldots , \widehat I_l\in S({\frak g})^G$ be the fundamental invariants. $\widehat I_1,\ldots , \widehat I_l$ may be viewed as polynomial functions on ${\frak g}^* \cong {\frak g}$. The restrictions of these functions to $f+{\frak s}$ define a global coordinate system on ${\frak s}$.} \vskip 0.3cm \noindent {\bf Theorem E (\cite{K}, Theorem 1.3)} {\em For any $\widehat I\in S({\frak g})^G$ one has $\widehat I^\chi\in S({\frak b}_-)^{N_+}$. Furthermore the map \begin{equation}\label{iso1} S({\frak g})^G\rightarrow S({\frak b}_-)^{N_+},~~\widehat I\mapsto \widehat I^\chi \end{equation} is an algebra isomorphism. In particular $$ S({\frak b}_-)^N={\Bbb C}[\widehat I_1^\chi ,\ldots , \widehat I_l^\chi ] $$ is a polynomial algebra in $l$ generators.} \vskip 0.3cm \noindent {\em Proof.} First observe that elements of $S({\frak g})$ may be viewed as polynomial functions on ${\frak g}^* \cong {\frak g}$. Note also that the ideal $I_\chi^0$ is generated by the elements $x-(x,f),~x\in {{\frak n}_+}$. Therefore $I_\chi^0$ is the ideal of polynomial functions vanishing on the subspace $f+{{\frak b}_+}$ and so for every $\widehat I\in S({\frak g})$ $\widehat I^\chi$ may be regarded as the restriction of the function $\widehat I$ to the subspace $f+{{\frak b}_+}$. For $\widehat I\in S({\frak g})^G,~s'\in {{\frak b}_+}$ and $a\in N_+$ one has $\widehat I({\rm Ad}(a)(f+s'))= \widehat I^\chi(a\cdot s')$. Since $\widehat I({\rm Ad}(a)(f+s'))=\widehat I(f+s')$ it follows that $\widehat I^\chi \in S({\frak b}_-)^{N_+}$. By Theorem C the map $$ S({\frak b}_-)^{N_+} \rightarrow S({\frak s}^*) $$ given by the restriction $v\mapsto v|_{\frak s}$ is an algebra isomorphism. Now by Theorem D the restrictions of the functions $\widehat I_i^\chi ,~i=1,\ldots ,l$ to ${\frak s}$ are a coordinate system. Therefore (\ref{iso1}) is an isomorphism. \vskip 0.3cm \noindent {\em Proof of Theorem B.} First observe that elements $\widehat I_i^\chi,~i=1,\ldots ,l$ are the graded limits of the elements $I_i^\chi \in U({\frak b}_-)^{N_+},~i=1,\ldots ,l$. Therefore $GrW({\frak b}_-)=S({\frak b}_-)^{N_+}$. Recall that $W({\frak b}_-)\subseteq U({\frak b}_-)^{N_+}$ is a linear subspace. Let $J\in U({\frak b}_-)^{N_+}\cap U_{k_1}({\frak b}_-)$ be an invariant element. Clearly, $GrJ\in S({\frak b}_-)^{N_+}$. Since $GrW({\frak b}_-)=S({\frak b}_-)^{N_+}$ one can find elements $I_1\in W({\frak b}_-)\cap U_{k_1}({\frak b}_-)$ and $J_1\in U({\frak b}_-)^{N_+}\cap U_{k_2}({\frak b}_-), ~~k_2<k_1$ such that $$ J-I_1=J_1. $$ Applying the same procedure to $J_1$ we obtain elements $J_2\in U({\frak b}_-)^{N_+}\cap U_{k_3}({\frak b}_-), ~k_3<k_2,~~I_2\in W({\frak b}_-)\cap U_{k_2}({\frak b}_-)$ such that $$ J_1-I_2=J_2. $$ We can continue this process. Since the standard filtration in $U({\frak b}_-)$ is bounded below we finally obtain that for some $i~~ J_i-I_i=c\in {\Bbb C}$. By construction the element $J$ is represented as $J=\sum_{j=1}^i I_j+c,~I_j\in W({\frak b}_-)\cap U_{k_j}({\frak b}_-)$. Therefore $J\in W({\frak b}_-)$. This concludes the proof. Now we make an important remark. \vskip 0.3cm {\bf Remark A} Observe that the space $U({\frak b}_-)^{N_+}$ may be interpreted as the zeroth cohomology space of the $U({{\frak n}_+})$ module $Y_\chi$, where $U({{\frak n}_+})$ is augmented by $\chi$. Indeed, for every associative algebra $B$ equipped with character $\chi$ and for every left $B$--module $V$ the cohomology module $H^*(V)$ is defined as the cohomology space of the complex (see \cite{carteil}) \begin{equation}\label{cohomol1} {\rm Hom}_B(X,V), \end{equation} where $X$ is a projective resolution of the one--dimensional $B$--module ${\Bbb C}_\chi$ defined by $\chi$. In homological algebra $\chi$ is called an augmentation of $B$. It is well--known that the graded vector space $H^*(V)$ does not depend on the resolution $X$ and the zeroth cohomology space $H^0(V)$ is isomorphic to the space of invariants ${\rm Hom}_B({\Bbb C}_\chi ,V)$ (see \cite{carteil}). Using the map $$ {\rm Hom}_B({\Bbb C}_\chi ,V)\rightarrow V;~~ \hat{v}\mapsto \hat{v}(1)=v $$ this space may be identified with the subspace in $V$ spanned by elements $v\in V$ such that $bv=\chi (b)v$ for every $b\in B$, i.e. $$ H^0(V)={\rm Hom}_B({\Bbb C}_\chi ,V)=\{ v\in V: bv=\chi (b)v \mbox{ for every } b\in B \}. $$ Now for $B=U({{\frak n}_+})$, $\chi$ as in Theorem B and $V=Y_\chi$ we have $H^0(Y_\chi)=\{ v\in Y_\chi: xv=\chi (x)v \mbox{ for every } x\in U({{\frak n}_+}) \}$. From (\ref{mainact}) and (\ref{iso2}) it follows that $H^0(Y_\chi)=U({\frak b}_-)^{N_+}$. Now recall that by Theorem B there exists a linear isomorphism $W({\frak b}_-)\cong U({\frak b}_-)^{N_+}$. Therefore the associative algebra $W({\frak b}_-)$ is isomorphic to $H^0(Y_\chi)$ as a linear space. In Section \ref{whitthom} we show that the multiplicative structure of $W({\frak b}_-)$ naturally appears in the context of homological algebra. \chapter{Hecke algebras}\label{Hom} In this section we give a homological definition of Hecke algebras (see Section \ref{Hecke}). Let $K$ be a ring with unit, $A$ an associative algebra over $K$, and $B$ a subalgebra of $A$ with augmentation, that is, a $K$--algebra homomorphism $\varepsilon : B\rightarrow K$. The Hecke algebra $Hk^*(A,B,\varepsilon)$ of the triple $(A,B,\varepsilon )$ is a natural generalization of the algebra ${\rm Hom}_A(A\otimes_BK,A\otimes_BK)$. For every left $A$ module $V$ and every right $A$ module $W$ the algebra $Hk^*(A,B,\varepsilon)$ acts in both the cohomology space $H^{*}(B,V)$ and the homology space $H_*(B,W)$ of $V$ and $W$ as $B$--modules. Hecke algebras are also closely related to the quantum BRST cohomology (see \cite{KSt}). To define Hecke algebras we study complexes of $A$--endomorphisms of graded left $A$ modules. Let $X$ be such a complex, ${\rm End}_{A}(X)$ be the corresponding complex of endomorphisms. Our main observation is that the natural multiplication in ${\rm End}_{A}(X)$ given by composition of endomorphisms induces a multiplicative structure on the cohomology space $H^*({\rm End}_{A}(X))$. Furthermore, the associative algebra $H^*({\rm End}_{A}(X))$ only depends on the homotopy class of the complex $X$. As an application of our construction we show that the Whittaker model $W({\frak b}_-)$ is the zeroth graded component of the Hecke algebra of the triple $(U({\frak g}),U({\frak n}),\chi )$. The exposition in this chapter follows \cite{S2}. \section{Endomorphisms of complexes}\label{end} \setcounter{equation}{0} \setcounter{theorem}{0} Let $A$ be an associative ring with unit, $X$ a graded complex of left $A$~modules equipped with a differential $d$ of degree $-1$. Recall the definition of the complex $Y={\rm End}_{A}(X)$ \cite{MacLane}. By definition $Y$ is a $\Bbb Z$--graded complex $$ Y=\bigoplus_{n=-\infty}^{\infty} Y^n $$ with graded components defined as $$ Y^n=\prod_{p+q=n} Y^{p,q}, $$ where $$ Y^{p,q}={\rm Hom}_{A}(X^{p},X^{-q}). $$ Clearly $Y$ is closed with respect to the multiplication given by composition of endomorphisms. Thus it is a graded associative algebra. We introduce a differential on $Y$ of degree +1 as follows: $$ \begin{array}{c} ({\bf d}f)^{p,q}=(-1)^{p+q}f^{p-1,q}\circ d +d\circ f^{p,q-1}, \\ f=\{ f^{p,q} \}, f^{p,q} \in Y^{p,q}, \end{array} $$ where $d$ is the differential of $X$. If $f$ is homogeneous then \begin{equation}\label{diff} {\bf d}f=d\circ f -(-1)^{{\rm deg} (f)}f\circ d . \end{equation} So that ${\bf d}$ is the supercommutator by $d$. We shall consider also the partial differentials $d'$ and $d''$ on Y : \begin{equation}\label{part} \begin{array}{cc} \mbox{for } f\in Y^{p,q}& \\ (d'f)(x)=(-1)^{p+q+1}f(dx),& x\in X^{p+1}; \\ (d''f)(x)=df(x),& x\in X^{p}. \end{array} \end{equation} It is easy to check that $$ d'^2=d''^2=d'd''+d''d'=0 $$ These conditions ensure that ${\bf d}^2=0$. The following property of ${\bf d}$ is crucial for the subsequent considerations. \begin{lemma} ${\bf d}$ is a superderivation of $Y$. \end{lemma} {\em Proof.} Let $f$ and $g$ be homogeneous elements of $Y$. Then ${\rm deg} (fg) = {\rm deg} (f) + {\rm deg} (g)$ and (\ref{diff}) yields: $$ \begin{array}{l} {\bf d}(fg)=d\circ fg - (-1)^{{\rm deg} (f) +{\rm deg} (g)} fg \circ d= \\ d\circ fg -(-1)^{{\rm deg} (f)} f \circ d \circ g +(-1)^{{\rm deg} (f)} f \circ d \circ g - (-1)^{{\rm deg} (f) +{\rm deg} (g)} fg \circ d= \\ ({\bf d}f)g+(-1)^{{\rm deg} (f)}f({\bf d}g). \end{array} $$ This completes the proof. The most important consequence of the lemma is \begin{proposition}\label{alg} The homology space $H^*(Y)$ inherits a multiplicative structure from Y. Thus $H^*(Y)$ is a graded associative algebra. \end{proposition} {\em Proof.} First, the product of two cocycles is a cocycle. For if $f$ and $g$ are homogeneous and ${\bf d}f={\bf d}g=0$ then $$ {\bf d}(fg)=({\bf d}f)g+(-1)^{{\rm deg} (f)}f({\bf d}g)=0. $$ Now we have to show that the product of homology classes is well--defined. It suffices to verify that the product of a homogeneous cocycle with a homogeneous coboundary is cohomologous to zero. For instance consider the product $f{\bf d}h$. Equation (\ref{diff}) gives \begin{eqnarray} f{\bf d}h=f\circ (d\circ h -(-1)^{{\rm deg} (h)}h\circ d)= \\ (-1)^{{\rm deg} (f)}d\circ f \circ h - (-1)^{{\rm deg} (h)}f\circ h\circ d = (-1)^{{\rm deg}(f)}{\bf d}(fh). \nonumber \end{eqnarray} This completes the proof. One of the principal statements of homological algebra says that homotopically equivalent complexes have the same homology. In particular the vector space $H^*(Y)$ depends only on the homotopy class of the complex $X$. It turns out that the same is true for the algebraic structure of $H^*(Y)$. Indeed we have the following \begin{theorem}\label{equiv} Let $X, X'$ be two homotopically equivalent graded complexes of left $A$--modules. Then $$ H^*(Y)\cong H^*(Y') $$ as graded associative algebras. \end{theorem} {\em Proof.} Let $F:X \rightarrow X' , F':X' \rightarrow X$ be two maps between the complexes such that $$ \begin{array}{lll} F'F-{\rm id}_X=d_Xs+sd_X,& s:X\rightarrow X ,& s \in Y^{-1},\\ FF'-{\rm id}_{X'}=d_{X'}s'+s'd_{X'},& s':X'\rightarrow X' ,& s' \in Y'^{-1}. \end{array} $$ Consider the induced mappings of the complexes $Y,~~Y'$: $$ \begin{array}{c} FF'^{*}:Y \rightarrow Y' ,\\ FF'^{*}f= F \circ f \circ F' , f\in Y ;\\ F'F^{*}:Y' \rightarrow Y ,\\ F'F^{*}g= F' \circ g \circ F , f\in Y' .\\ \end{array} $$ Their compositions are homotopic to the identity maps of $Y$ and $Y'$ (see Chap.~4, \cite{carteil} for a general statement about equivalences of functors). But this means that $FF'^{*}$ is inverse to $F'F^{*}$ when restricted to homology. Thus $H^*(Y)$ is isomorphic to $H^*(Y')$ as a vector space. We have to show that the restrictions of $FF'^{*}$ and $F'F^{*}$ to the homologies are homomorphisms of algebras. Let $f$ and $g$ be homogeneous elements of $Y$ and ${\bf d}_Xf={\bf d}_Xg=0$. By the definition of the induced maps we have $$ FF'^{*}(fg)=F \circ fg \circ F'. $$ On the other hand \begin{eqnarray}\label{hom} FF'^{*}(f)FF'^{*}(g)=F \circ f \circ F'F \circ g \circ F'= \\ F \circ f({\rm id}_X+d_Xs+sd_X)g \circ F'.\nonumber \end{eqnarray} Now recall that $f$ and $g$ are cocycles in $Y$. By (\ref{diff}) they supercommute with $d_X$: \begin{equation}\label{cocycle} d_X\circ f =(-1)^{{\rm deg} (f)}f\circ d_X. \end{equation} Using (\ref{cocycle}) and the fact that $F$ and $F'$ are morphisms of complexes we can rewrite (\ref{hom}) as follows: \begin{eqnarray}\label{hom1} F \circ f({\rm id}_X+d_Xs+sd_X)g \circ F'= F \circ fg \circ F' + \nonumber \\ +(-1)^{{\rm deg} (f)}d_{X'} \circ F \circ fsg \circ F' + (-1)^{{\rm deg} (g)}F \circ fsg \circ F' \circ d_{X'} = \\ =F \circ fg \circ F'+(-1)^{{\rm deg} (f)}{\bf d}_{X'}(F \circ fsg \circ F') . \nonumber \end{eqnarray} Finally observe that by (\ref{hom1}), $FF'^{*}(fg)$ and $FF'^{*}(f)FF'^{*}(g)$ belong to the same homology class in $H^{*}(Y')$. This completes the proof. \section{Hecke algebras}\label{Hecke} \setcounter{equation}{0} \setcounter{theorem}{0} Let $A$ be an associative algebra over a ring $K$ with unit, and $B$ a subalgebra of $A$ with augmentation, that is, a $K$--algebra homomorphism $\varepsilon : B\rightarrow K$. Let $X$ be a projective resolution of the left $B$--module $K$ defined by $\varepsilon$. Since $X$ is a complex of left $B$--modules, the space $A \otimes_B X$ is also a differential complex. Observe that this complex has the natural structure of a left $A$--module. Therefore we can apply Proposition \ref{alg} to define a graded associative algebra $$ Hk^*(A,B,\varepsilon)=H^*({\rm End}_A(A \otimes_B X)). $$ Note that all $B$--projective resolutions of $K$ are homotopically equivalent and so the complexes $A \otimes_B X$ are homotopically equivalent for different resolutions $X$. Hence by Theorem \ref{equiv} the associative algebra $Hk^*(A,B,\varepsilon)$ does not depend on the resolution $X$. We shall call it the {\it Hecke algebra} of the the triple $(A,B,\varepsilon )$. Now consider $A$ as a left $A$--module and a right $B$--module via multiplication. In this way $A$ becomes a left $A\otimes B^{opp}$--module. Let $X'$ be a projective resolution of this module. The complex $X' \otimes_B K$, where the $B$ module structure on $K$ is defined by $\varepsilon$, is a left $A$--module. Therefore one can define an associative algebra $$ \widehat{Hk}^*(A,B,\varepsilon)=H^*({\rm End}_A(X' \otimes_B K)) $$ independent of the resolution $X'$. \begin{proposition}\label{iso} $Hk^*(A,B,\varepsilon)$ is isomorphic to $\widehat{Hk}^*(A,B,\varepsilon)$ as a graded associative algebra. \end{proposition} {\em Proof.} We shall use the standard bar resolutions for computing $\widehat{Hk}^*(A,B,\varepsilon)$ and ${Hk}^*(A,B,\varepsilon)$ \cite{MacLane}, \cite{carteil}. Consider the complex $B\otimes T(I(B)) \otimes B$, where $I(B)=B/K$ and $T$ denotes the tensor algebra of the vector space. Elements of $B\otimes T(I(B)) \otimes B$ are usually written as $a[a_1,\ldots ,a_s]a'$. The differential is given by \begin{eqnarray} da[a_1,\ldots ,a_s]a'=aa_1[a_2,\ldots ,a_s]a'+ \\ \sum_{k=1}^{s-1}(-1)^{k}a[a_1,\ldots ,a_ka_{k+1},\ldots ,a_s]a' + (-1)^sa[a_1,\ldots ,a_{s-1}]a_sa'. \nonumber \end{eqnarray} Then $B\otimes T(I(B)) \otimes B \otimes_B K = B\otimes T(I(B)) \otimes K$ is a free resolution of the left $B$--module $K$. And $A \otimes_B B\otimes T(I(B)) \otimes B =A\otimes T(I(B)) \otimes B$ is a free resolution of $A$ as a right $B$--module. The complex $A\otimes T(I(B)) \otimes B$ is also a free left $A$--module via left multiplication by elements of $A$. Hence this is an $A\otimes B^{opp}$-- free resolution of $A$. Thus the complex ${\rm End}_A(A\otimes_B B\otimes T(I(B)) \otimes K)={\rm End}_A(A\otimes T(I(B)) \otimes K)$ for the computation of $Hk^*(A,B,\varepsilon)$ is canonically isomorphic to the complex ${\rm End}_A(A \otimes T(I(B)) \otimes B \otimes_B K)={\rm End}_A(A \otimes T(I(B)) \otimes K)$ for the computation of $\widehat{Hk}^*(A,B,\varepsilon)$. This establishes the isomorphism of the algebras. \section{Action in homology and cohomology spaces} \setcounter{equation}{0} \setcounter{theorem}{0} Recall that for every left $B$--module $V$ the cohomology modules are defined to be \begin{equation}\label{cohomol} H^*(B,V)={\rm Ext}_B^*(K,V)=H^*({\rm Hom}_B(X,V)), \end{equation} where $X$ is a projective resolution of $K$. On the other hand for every right $B$--module $W$ one can define the homology modules \begin{equation}\label{homol} H_*(B,W)={\rm Tor}_*^B(W,K)=H_*(W\otimes_B X). \end{equation} Now observe that for every left $A$--module $V$ the complex in (\ref{cohomol}) for calculating its cohomology as a right $B$--module may be represented as follows: \begin{equation}\label{cohomcompl} {\rm Hom}_B(X,V)={\rm Hom}_A(A\otimes_B X ,V). \end{equation} Endow the space ${\rm Hom}_A(A\otimes_B X ,V)$ with a right ${\rm End}_A(A\otimes_B X)$--action: \begin{equation}\label{act1} \begin{array}{ll} {\rm Hom}_A(A\otimes_B X ,V) \times {\rm End}_A(A\otimes_B X) \rightarrow {\rm Hom}_A(A\otimes_B X ,V) , &\\ \varphi \times f \mapsto \varphi \circ f, &\\ \varphi \in {\rm Hom}_A(A\otimes_B X ,V), f \in {\rm End}_A(A\otimes_B X).& \end{array} \end{equation} This action is well--defined since $f$ commutes with the left $A$--action. Clearly this action respects the gradings, i.e., it is an action of the graded associative algebra on the graded module. \begin{proposition}\label{cohomact} For every left $A$ module $V$ the action (\ref{act1}) gives rise to a right action \begin{eqnarray}\label{cohomact1} H^*(B,V)\times {Hk}^*(A,B,\varepsilon) \rightarrow H^*(B,V), \\ H^n(B,V)\times {Hk}^m(A,B,\varepsilon) \rightarrow H^{n+m}(B,V).\nonumber \end{eqnarray} \end{proposition} {\em Proof.} Let $\varphi \in {\rm Hom}_A(A\otimes_B X ,V)$ and $d\varphi =\varphi \circ d =0$. Let also $f \in {\rm End}_A(A\otimes_B X)$ be a homogeneous cocycle. By (\ref{cocycle}) $\varphi \circ f$ is a cocycle in ${\rm Hom}_A(A\otimes_B X ,V)$. Indeed $$ d(\varphi \circ f)=\varphi \circ f \circ d =(-1)^{{\rm deg} (f)} \varphi \circ d \circ f =0. $$ Next we need to show that the action does not depend on the choice of the representative $f$ in the homology class $[f]$, that is $\varphi \circ {\bf d}g$ is homologous to zero for every homogeneous $g\in {\rm End}_A(A\otimes_B X)$. This is a direct consequence of the definitions: $$ \varphi \circ {\bf d}g= \varphi \circ (d\circ g - (-1)^{{\rm deg}(g)} g\circ d)= -(-1)^{{\rm deg}(g)}d(\varphi \circ g), $$ since $\varphi \circ d =0$. Finally let us check that the action is independent of the representative in the homology class $[\varphi]$. For $\psi \in {\rm Hom}_A(A\otimes_B X ,V)$ $d\psi \circ f$ is always homologous to zero: $$ d\psi \circ f= \psi \circ d \circ f = (-1)^{{\rm deg}(f)}\psi \circ f \circ d =(-1)^{{\rm deg}(f)}d(\psi \circ f). $$ This concludes the proof. Similarly for every right $A$--module $W$ one can equip the homology module $H_*(B,W)$ with a structure of a left ${Hk}^*(A,B,\varepsilon)$--module. First the complex $W\otimes_B X=W\otimes_A A\otimes_B X$ has the natural structure of a left ${\rm End}_A(A\otimes_B X)$--module: \begin{equation}\label{act2} \begin{array}{ll} {\rm End}_A(A\otimes_B X)\times W\otimes_A A\otimes_B X \rightarrow W\otimes_A A\otimes_B X , &\\ f \times w\otimes x\mapsto w\otimes f(x), &\\ w\otimes x \in W\otimes_A (A\otimes_B X), f \in {\rm End}_A(A\otimes_B X).& \end{array} \end{equation} Observe that according to the convention of Section \ref{end} elements of ${\rm End}_A^n(A\otimes_B X)$ have degree -n as operators in the graded space $W\otimes_A A\otimes_B X$: $$ {\rm End}_A^n(A\otimes_B X)\times W\otimes_A A\otimes_B X_m \rightarrow W\otimes_A A\otimes_B X_{m-n}. $$ The following assertion is an analogue of Proposition \ref{cohomact} for homology. \begin{proposition} For every right $A$ module $W$ the action (\ref{act2}) gives rise to a left action \begin{eqnarray} Hk(A,B,\varepsilon)^* \times H_*(B,W) \rightarrow H_*(B,W),\\ Hk(A,B,\varepsilon)^n \times H_m(B,W) \rightarrow H_{m-n}(B,W).\nonumber \end{eqnarray} \end{proposition} \section{Structure of the Hecke algebras} \setcounter{equation}{0} \setcounter{theorem}{0} In this section we investigate the Hecke algebras under some technical assumptions. The main theorem here is \begin{theorem}\label{struct} Assume that $$ {\rm Tor}_n^B(A,K)=0 \mbox{ for } n>0. $$ Then $$ Hk^n(A,B,\varepsilon)={\rm Ext}^n_A(A\otimes_BK,A\otimes_BK)={\rm Ext}^n_B(K,A\otimes_BK). $$ In particular $$ Hk^n(A,B,\varepsilon)=0 ,~ n<0, $$ and $$ Hk^0(A,B,\varepsilon )={\rm Hom}_A(A\otimes_BK,A\otimes_BK) $$ as an associative algebra. \end{theorem} {\em Proof.} Equip the complex $Y={\rm End}_A(A\otimes T(I(B)) \otimes K)$, which we used in Proposition \ref{iso} for the computation of $Hk^*(A,B,\varepsilon)$, with the first filtration as follows: $$ F^kY=\sum_{n=-\infty}^{\infty}\prod_{p+q=n , p\geq k}Y^{p,q}. $$ The associated graded complex with respect to this filtration is the double direct sum $$ {\rm Gr}Y=\sum_{p,q=-\infty}^{\infty}Y^{p,q}. $$ One can show that the filtration is regular and the second term of the corresponding spectral sequence is \begin{equation}\label{spec} E_2^{p,q}=H^p_{d'}(H^q_{d''}({\rm Gr}Y)), \end{equation} where $H^*_{d'}$ and $H^*_{d''}$ denote the homologies of the complex with respect to the partial differentials (\ref{part}). Now observe that at the same time the complex $A\otimes T(I(B)) \otimes K$ is a complex for the calculation of ${\rm Tor}_n^B(A,K)$ because $A\otimes T(I(B)) \otimes B$ is a free resolution of $A$ as a right $B$--module. It is also free as a left $A$--module. Therefore the functor ${\rm Hom}_A(A\otimes T(I(B)) \otimes K, \cdot )$ is exact. By assumption $H^*(A\otimes T(I(B)) \otimes K)={\rm Tor}_0^B(A,K)=A\otimes_BK$. Using the last two observations we can calculate the cohomology of the complex ${\rm Gr}Y$ with respect to the differential $d''$ : \begin{equation}\label{degener} \begin{array}{l} H^*_{d''}({\rm Gr}Y)=H^*_{d''}({\rm Hom}_A(A\otimes T(I(B)) \otimes K,A\otimes T(I(B)) \otimes K))=\\ {\rm Hom}_A(A\otimes T(I(B)) \otimes K,A\otimes_BK). \end{array} \end{equation} Here ${\rm Hom}_A$ should be thought of as a direct sum of the double graded components. Now (\ref{degener}) provides that the spectral sequence (\ref{spec}) degenerates at the second term. Moreover, $$ E_2^{p,*}=H^p_{d'}(H^0_{d''}({\rm Gr}Y))=H^p_{d'}({\rm Hom}_A (A\otimes T(I(B)) \otimes K,A\otimes_BK)). $$ But the complex $A\otimes T(I(B)) \otimes K$ may be regarded as a free resolution of the left $A$--module $A\otimes_BK$. Therefore $$ E_2^{p,*}={\rm Ext}^p_A(A\otimes_BK,A\otimes_BK). $$ Finally by Theorem 5.12, \cite{carteil} we have: $$ Hk^n(A,B,\varepsilon)=H^n(Y)=E_2^{n,0}={\rm Ext}^n_A(A\otimes_BK,A\otimes_BK). $$ Since ${\rm Tor}_n^B(A,K)=0 \mbox{ for } n>0$ we can apply the Shapiro lemma (see Proposition 4.1.3 in \cite{carteil}) to simplify the last expression: $$ {\rm Ext}^n_A(A\otimes_BK,A\otimes_BK)={\rm Ext}^n_B(K,A\otimes_BK). $$ Clearly, $Hk^0(A,B,\varepsilon)={\rm Hom}_A(A\otimes_BK,A\otimes_BK)$ as an associative algebra. This completes the proof. \begin{remark}\label{multhk} In particular the conditions of the theorem are satisfied if $A$ is projective as a right $B$--module. For instance suppose that there exists a subspace $N \subset A$ such that multiplication in $A$ provides an isomorphism of vector spaces $A \cong N\otimes B$. Then $A$ is a free right $B$--module. \end{remark} \section{Comparison with the BRST complex} \setcounter{equation}{0} \setcounter{theorem}{0} Let $\frak g$ be a Lie algebra over a field $K$. For simplicity we suppose that $\frak g$ is finite--dimensional. However the arguments presented below remain true, with some technical modifications, for an arbitrary Lie algebra. We shall apply the construction of Section \ref{Hecke} in the following situation. Let $B=U({\frak g})$ and let $A$ be an associative algebra over $K$ containing $B$ as a subalgebra. Note that $U({\frak g})$ is naturally augmented. Consider the $U({\frak g})$--free resolution of the left $U({\frak g})$--module $K$ as follows: $$ \begin{array}{l} X=U({\frak g})\otimes \Lambda ({\frak g}),\\ d (u\otimes x_1 \wedge \ldots \wedge x_n)= \sum_{i=1}^n (-1)^{i+1} ux_i\otimes x_1 \wedge \ldots \wedge \widehat{x_i} \wedge \ldots \wedge x_n +\\ \sum_{1\leq i< j \leq n}(-1)^{i+j} u\otimes [x_i,x_j] \wedge x_1 \wedge \ldots \wedge \widehat{x_i} \wedge \ldots \wedge \widehat{x_j} \wedge \ldots \wedge x_n, \end{array} $$ where the symbol $\widehat{x_i}$ indicates that $x_i$ is to be omitted. Introduce operators of exterior and inner multiplication on $\Lambda ({\frak g})$ as follows. For every $x\in {\frak g}$ and $x^*\in {\frak g}^*$ we define $$ \begin{array}{l} \overline{x} x_1 \wedge \ldots \wedge x_n= x\wedge x_1 \wedge \ldots \wedge x_n,\\ \\ \overline{x^*}x_1 \wedge \ldots \wedge x_n= \sum_{i=1}^n (-1)^{i+1} x^*(x_i) x_1\wedge \ldots \wedge \widehat{x_i} \wedge \ldots \wedge x_n. \end{array} $$ Equip the linear space ${\frak g}+{\frak g}^*$ with a scalar product given by the canonical paring between ${\frak g}$ and ${\frak g}^*$. Using this scalar product we can construct the Clifford algebra $C({\frak g}+{\frak g}^*)$. The operators $\overline{x}, \overline{y^*}, x\in {\frak g},y^*\in{\frak g}^*$ satisfy the defining relations of this algebra, $$ \overline{x}\overline{y^*}+\overline{y^*}\overline{x}=y^*(x). $$ Therefore the algebra $C({\frak g}+{\frak g}^*)$ naturally acts in the space $\Lambda ({\frak g})$. Moreover, it is well--known that ${\rm End}_K(\Lambda ({\frak g}))=C({\frak g}+{\frak g}^*)$. Now the differential of the complex $A\otimes_{U({\frak g})} X = A\otimes \Lambda ({\frak g})$ may be explicitly described using the operators of exterior and inner multiplications, \begin{equation}\label{different} d=\sum_i e_i\otimes \overline{e_i^*}-\sum_{i,j} 1\otimes \overline{[e_i,e_j]}\overline{e_i^*}\overline{e_j^*}. \end{equation} Here $e_i$ is a linear basis of ${\frak g}$, $e_i^*$ is the dual basis, $e_i\otimes 1$ is regarded as the operator of right multiplication in $A$, $e_i\otimes 1\cdot u\otimes 1=ue_i\otimes 1$. Now consider the complex ${\rm End}_A(A\otimes_{U({\frak g})} X)={\rm End}_A(A\otimes \Lambda ({\frak g}))$ for the computation of the algebra $Hk(A,B,\varepsilon)$. Observe that $$ {\rm End}_A(A\otimes \Lambda ({\frak g})) = A^{opp}\otimes {\rm End}_K(\Lambda ({\frak g}))=A^{opp} \otimes C({\frak g}+{\frak g}^*). $$ Under this identification $A^{opp}$ acts on $A\otimes \Lambda ({\frak g})$ by multiplication in $A$ on the right and the Clifford algebra acts by the exterior and inner multiplication in $\Lambda ({\frak g})$. This allows to consider the differential (\ref{different}) as an element of the complex $A^{opp} \otimes C({\frak g}+{\frak g}^*)$. It is easy to see that the canonical $\Bbb Z$--grading of the complex $A^{opp} \otimes C({\frak g}+{\frak g}^*)$ coincides mod 2 with the ${\Bbb Z}_2$--grading inherited from the Clifford algebra. Therefore according to (\ref{diff}) the differential $\bf d$ is given by the supercommutator in $A^{opp} \otimes C({\frak g}+{\frak g}^*)$ by element (\ref{different}). Now recall that the complex $A^{opp} \otimes C({\frak g}+{\frak g}^*)$ with the differential given by the supercommutator by the element (\ref{different}) is the quantum BRST complex proposed in \cite{KSt}. This establishes \begin{proposition} The complex $({\rm End}_A(A\otimes_{U({\frak g})} X) , {\bf d})$ is isomorphic to the BRST one $A^{opp} \otimes C({\frak g}+{\frak g}^*)$ with the differential being the supercommutator by the element (\ref{different}). \end{proposition} \section{Whittaker model as a Hecke algebra}\label{whitthom} \setcounter{equation}{0} \setcounter{theorem}{0} In this section we use the notation introduced in Section \ref{notation}. Let $\frak g$ be a complex simple Lie algebra, ${\frak n}_+\subset {\frak g}$ the maximal nilpotent subalgebra, $\chi :{\frak n}_+\rightarrow {\Bbb C}$ a character. Let $W({\frak b}_-)$ be the Whittaker model of the center $Z({\frak g})$ of the universal enveloping algebra $U({\frak g})$. \begin{proposition}\label{hkhom} Suppose that the character $\chi$ is non--singular. Then $W({\frak b}_-)$ is isomorphic to $Hk^0(U({\frak g}),U({\frak n}_+),\chi )^{opp}$ as an associative algebra. \end{proposition} {\em Proof.} First observe that since ${\frak g}={\frak b}_-\oplus {\frak n}_+$ we have a linear isomorphism $U({\frak g})=U({\frak b}_-)\otimes U({\frak n}_+)$. Therefore from Remark \ref{multhk} and Theorem \ref{struct} it follows that $Hk^0(U({\frak g}),U({\frak n}_+),\chi )= {\rm Hom}_{U({\frak g})}(Y_\chi ,Y_\chi )$, where $Y_\chi =U({\frak g})\otimes_{U({\frak n}_+)}{\Bbb C}_\chi$. Now observe that the map $$ {\rm Hom}_{U({\frak g})}(Y_\chi ,Y_\chi )\rightarrow {\rm Hom}_{U({\frak n}_+)}({\Bbb C}_\chi ,Y_\chi);~~ \tilde{v}\mapsto \hat{v}, $$ where $\hat{v}$ is given by $\hat{v}(z)=\tilde{v}(1\otimes z)$ for every $z\in {\Bbb C}_\chi$, is a linear isomorphism. Note also that by Remark A and Theorem B there exists a linear isomorphism $$ {\rm Hom}_{U({\frak n}_+)}({\Bbb C}_\chi ,Y_\chi)\rightarrow W({\frak b}_-);~~\hat{v} \mapsto v, \mbox{ where }v\otimes 1=\hat{v}(1). $$ Therefore we have a linear isomorphism \begin{equation}\label{be} {\rm Hom}_{U({\frak g})}(Y_\chi ,Y_\chi )\rightarrow W({\frak b}_-);~~ \tilde{v}\mapsto v. \end{equation} We have to prove that (\ref{be}) is an antihomomorphism. Let $\tilde{v},\tilde{w}\in {\rm Hom}_{U({\frak g})}(Y_\chi ,Y_\chi )$ be two elements such that $\tilde{v}(1\otimes 1)=v\otimes 1,~\tilde{w}(1\otimes 1)=w\otimes 1$. Then $\tilde{v}(\tilde{w}(1\otimes 1))=\tilde{v}(w\otimes 1)$. Since $\tilde{v}$ is an $U({\frak g})$ endomorphism of $Y_\chi$ we have $\tilde{v}(\tilde{w}(1\otimes 1))=w\tilde{v}(1\otimes 1)=wv\otimes 1$ This completes the proof. \chapter{Quantum deformation of the Whittaker model}\label{qWitt} Let $\frak g$ be a complex simple Lie algebra, $U_h({\frak g})$ the standard quantum group associated with ${\frak g}$. In this section we construct a generalization of the Whittaker model $W({\frak b}_-)$ for $U_h({\frak g})$. Let $U_h({\frak n}_+)$ be the subalgebra of $U_h({\frak g})$ corresponding to the nilpotent Lie subalgebra ${\frak n}_+$. $U_h({\frak n}_+)$ is generated by simple positive root generators of $U_h({\frak g})$ subject to the quantum Serre relations. It is easy to show that $U_h({\frak n}_+)$ has no non--singular characters (taking nonvanishing values on all simple root generators). Our first main result is a family of new realizations of the quantum group $U_h({\frak g})$, one for each Coxeter element in the corresponding Weyl group (see also \cite{S1}). The counterparts of $U({\frak n}_+)$, which naturally arise in these new realizations of $U_h({\frak g})$, do have non--singular characters. Using these new realizations we can immediately formulate a quantum group version of Definition A. We also prove counterparts of Theorems A and B for $U_h({\frak g})$. Finally we define quantum group generalizations of the Toda Hamiltonians. In the spirit of quantum harmonic analysis these new Hamiltonians are difference operators. An alternative definition of these Hamiltonians has been recently given in \cite{Et}. \section{Quantum groups} \setcounter{equation}{0} \setcounter{theorem}{0} In this section we recall some basic facts about quantum groups. We follow the notation of \cite{ChP}. Let $h$ be an indeterminate, ${\Bbb C}[[h]]$ the ring of formal power series in $h$. We shall consider ${\Bbb C}[[h]]$--modules equipped with the so--called $h$--adic topology. For every such module $V$ this topology is characterized by requiring that $\{ h^nV ~|~n\geq 0\}$ is a base of the neighbourhoods of $0$ in $V$, and that translations in $V$ are continuous. It is easy to see that, for modules equipped with this topology, every ${\Bbb C}[[h]]$--module map is automatically continuous. A topological Hopf algebra over ${\Bbb C}[[h]]$ is a complete ${\Bbb C}[[h]]$--module $A$ equipped with a structure of ${\Bbb C}[[h]]$--Hopf algebra (see \cite{ChP}, Definition 4.3.1), the algebraic tensor products entering the axioms of the Hopf algebra are replaced by their completions in the $h$--adic topology. We denote by $\mu , \imath , \Delta , \varepsilon , S$ the multiplication, the unit, the comultiplication, the counit and the antipode of $A$, respectively. The standard quantum group $U_h({\frak g})$ associated to a complex finite--dimensional simple Lie algebra $\frak g$ is the algebra over ${\Bbb C}[[h]]$ topologically generated by elements $H_i,~X_i^+,~X_i^-,~i=1,\ldots ,l$, and with the following defining relations: \begin{equation}\label{qgrh} \begin{array}{l} [H_i,H_j]=0,~~ [H_i,X_j^\pm]=\pm a_{ij}X_j^\pm,\\ \\ X_i^+X_j^- -X_j^-X_i^+ = \delta _{i,j}{K_i -K_i^{-1} \over q_i -q_i^{-1}} , \\ \\ \mbox{where }K_i=e^{d_ihH_i},~~e^h=q,~~q_i=q^{d_i}=e^{d_ih}, \end{array} \end{equation} and the quantum Serre relations: $$ \begin{array}{l} \sum_{r=0}^{1-a_{ij}}(-1)^r \left[ \begin{array}{c} 1-a_{ij} \\ r \end{array} \right]_{q_i} (X_i^\pm )^{1-a_{ij}-r}X_j^\pm(X_i^\pm)^r =0 ,~ i \neq j ,\\ \\ \mbox{ where }\\ \\ \left[ \begin{array}{c} m \\ n \end{array} \right]_q={[m]_q! \over [n]_q![n-m]_q!} ,~ [n]_q!=[n]_q\ldots [1]_q ,~ [n]_q={q^n - q^{-n} \over q-q^{-1} }. \end{array} $$ $U_h({\frak g})$ is a topological Hopf algebra over ${\Bbb C}[[h]]$ with comultiplication defined by $$ \begin{array}{l} \Delta_h(H_i)=H_i\otimes 1+1\otimes H_i,\\ \\ \Delta_h(X_i^+)=X_i^+\otimes K_i+1\otimes X_i^+, \end{array} $$ $$ \Delta_h(X_i^-)=X_i^-\otimes 1 +K_i^{-1}\otimes X_i^-, $$ antipode defined by $$ S_h(H_i)=-H_i,~~S_h(X_i^+)=-X_i^+K_i^{-1},~~S_h(X_i^-)=-K_iX_i^-, $$ and counit defined by $$ \varepsilon_h(H_i)=\varepsilon_h(X_i^\pm)=0. $$ We shall also use the weight--type generators defined by $$ Y_i=\sum_{j=1}^l d_i(a^{-1})_{ij}H_j, $$ and the elements $L_i=e^{hY_i}$. They commute with the root vectors $X_i^\pm$ as follows: \begin{equation}\label{weight-root} L_iX_j^\pm L_i^{-1}=q_i^{\pm \delta_{ij}}X_j^\pm . \end{equation} The Hopf algebra $U_h({\frak g})$ is a quantization of the standard bialgebra structure on $\frak g$, i.e. $U_h({\frak g})/hU_h({\frak g})=U({\frak g}),~~ \Delta_h=\Delta~(\mbox{mod }h)$, where $\Delta$ is the standard comultiplication on $U({\frak g})$, and $$ {\Delta_h -\Delta_h^{opp} \over h}~(\mbox{mod }h)=\delta , $$ where $\delta: {\frak g}\rightarrow {\frak g}\otimes {\frak g}$ is the standard cocycle on $\frak g$. Recall that $$ \delta (x)=({\rm ad}_x\otimes 1+1\otimes {\rm ad}_x)2r_+,~~ r_+\in {\frak g}\otimes {\frak g}, $$ \begin{equation}\label{rcl} r_+=\frac 12 \sum_{i=1}^lY_i \otimes X_i + \sum_{\beta \in \Delta_+}(X_{\beta},X_{-\beta})^{-1} X_{\beta}\otimes X_{-\beta}. \end{equation} Here $X_{\pm \beta}\in {\frak g}_{\pm \beta}$ are root vectors of $\frak g$. The element $r_+\in {\frak g}\otimes {\frak g}$ is called a classical r--matrix. The following proposition describes the algebraic structure of $U_h({\frak g})$. \begin{proposition}{\bf (\cite{ChP}, Proposition 6.5.5)}\label{algq} Let $\frak g$ be a finite--dimensional complex simple Lie algebra, let $U_h({\frak h})$ be the subalgebra of $U_h({\frak g})$ topologically generated by the $H_i, i=1,\ldots l$. Then, there is an isomorphism of algebras $\varphi :U_h({\frak g})\rightarrow U({\frak g})[[h]]$ over ${\Bbb C}[[h]]$ such that $\varphi =id$ (mod $h$) and $\varphi|_{U_h({\frak h})}=id$. \end{proposition} \begin{proposition}{\bf (\cite{ChP}, Proposition 6.5.7)}\label{zq} If $\frak g$ is a finite--dimensional complex simple Lie algebra, the center $Z_h({\frak g})$ of $U_h({\frak g})$ is canonically isomorphic to $Z({\frak g})[[h]]$, where $Z({\frak g})$ is the center of $U({\frak g})$. \end{proposition} \begin{corollary}{\bf (\cite{ChP}, Corollary 6.5.6)}\label{rep} If $\frak g$ be a finite--dimensional complex simple Lie algebra, then the assignment $V\mapsto V[[h]]$ is a one--to--one correspondence between the finite--dimensional irreducible representations of $\frak g$ and indecomposable representations of $U_h({\frak g})$ which are free and of finite rank as ${\Bbb C}[[h]]$--modules. Furthermore for every such $V$ the action of the generators $H_i \in U_h({\frak g}),~~ i=1,\ldots l$ on $V[[h]]$ coincides with the action of the root generators $H_i \in {\frak h},~~ i=1,\ldots l$. \end{corollary} The representations of $U_h({\frak g})$ defined in the previous corollary are called finite--dimensional representations of $U_h({\frak g})$. For every finite--dimensional representation $\pi_V:{\frak g}\rightarrow {\rm End}V$ we denote the corresponding representation of $U_h({\frak g})$ in the space $V[[h]]$ by the same letter. $U_h({\frak g})$ is a quasitriangular Hopf algebra, i.e. there exists an invertible element ${\cal R}\in U_h({\frak g})\otimes U_h({\frak g})$, called a universal R--matrix, such that \begin{equation}\label{quasitr} \Delta^{opp}_h(a)={\cal R}\Delta_h(a){\cal R}^{-1}\mbox{ for all } a\in U_h({\frak g}), \end{equation} where $\Delta^{opp}=\sigma \Delta$, $\sigma$ is the permutation in $U_h({\frak g})^{\otimes 2}$, $\sigma (x\otimes y)=y\otimes x$, and \begin{equation}\label{rmprop} \begin{array}{l} (\Delta_h \otimes id){\cal R}={\cal R}_{13}{\cal R}_{23},\\ \\ (id \otimes \Delta_h){\cal R}={\cal R}_{13}{\cal R}_{12}, \end{array} \end{equation} where ${\cal R}_{12}={\cal R}\otimes 1,~{\cal R}_{23}=1\otimes {\cal R}, ~{\cal R}_{13}=(\sigma \otimes id){\cal R}_{23}$. From (\ref{quasitr}) and (\ref{rmprop}) it follows that $\cal R$ satisfies the quantum Yang--Baxter equation: \begin{equation}\label{YB} {\cal R}_{12}{\cal R}_{13}{\cal R}_{23}={\cal R}_{23}{\cal R}_{13}{\cal R}_{12}. \end{equation} For every quasitriangular Hopf algebra we also have (see Proposition 4.2.7 in \cite{ChP}): $$ (S\otimes id){\cal R}={\cal R}^{-1}, $$ and \begin{equation}\label{S} (S\otimes S){\cal R}={\cal R}. \end{equation} We shall explicitly describe the element ${\cal R}$. First following \cite{kh-t} we recall the construction of root vectors of $U_h({\frak g})$. We shall use the so--called normal ordering in the root system $\Delta_+=\{\beta_1,\ldots ,\beta_N\}$ (see \cite{Z1}). \begin{definition}\label{normord} An ordering of the root system $\Delta_+$ is called normal if all simple roots are written in an arbitrary order, and for any theree roots $\alpha,~\beta,~\gamma$ such that $\gamma=\alpha+\beta$ we have either $\alpha<\gamma<\beta$ or $\beta<\gamma<\alpha$. \end{definition} To construct root vectors we shall apply the following inductive algorithm. Let $\alpha , \beta , \gamma \in \Delta_+$ be positive roots such that $\gamma=\alpha+\beta,~\alpha<\beta$ and $[\alpha,\beta]$ is the minimal segment including $\gamma$, i.e. the segment has no other roots $\alpha',\beta'$ such that $\gamma=\alpha'+\beta'$. Suppose that $X_{\alpha}^\pm ,~X_{\beta}^\pm$ have already been constructed. Then we define \begin{equation}\label{rootvect} \begin{array}{l} X_{\gamma}^+=X_{\alpha}^+X_{\beta}^+ - q^{(\alpha,\beta)}X_{\beta}^+X_{\alpha}^+,\\ \\ X_{\gamma}^-= X_{\beta}^-X_{\alpha}^- - q^{-(\alpha,\beta)}X_{\alpha}^-X_{\beta}^-. \end{array} \end{equation} \begin{proposition}\label{rootprop} For $\beta =\sum_{i=1}^lm_i\alpha_i,~m_i\in {\Bbb N}$ $X_{\beta}^\pm $ is a polynomial in the noncommutative variables $X_i^\pm$ homogeneous in each $X_i^\pm$ of degree $m_i$. \end{proposition} The root vectors $X_{\beta}$ satisfy the following relations: $$ [X_\alpha^+,X_{\alpha}^-]=a(\alpha){e^{h\alpha^\vee}-e^{-h\alpha^\vee}\over q-q^{-1}}. $$ where $a(\alpha)\in {\Bbb C}[[h]]$. They commute with elements of the subalgebra $U_h({\frak h})$ as follows: \begin{equation}\label{roots-cart} [H_i,X_{\beta}^\pm]=\pm \beta(H_i)X_{\beta}^\pm,~i=1,\ldots ,l. \end{equation} Note that by construction $$ \begin{array}{l} X_\beta^+~(\mbox{mod }h)=X_\beta \in {\frak g}_\beta,\\ \\ X_\beta^-~(\mbox{mod }h)=X_{-\beta} \in {\frak g}_{-\beta} \end{array} $$ are root vectors of $\frak g$. This implies that $a(\alpha)~(\mbox{mod }h)=(X_\alpha,X_{-\alpha})$. Let $U_h({\frak n}_+),U_h({\frak n}_-)$ be the ${\Bbb C}[[h]]$--subalgebras of $U_h({\frak g})$ topologically generated by the $X_i^+$ and by the $X_i^-$, respectively. Now using the root vectors $X_{\beta}^\pm$ we can construct a topological basis of $U_h({\frak g})$. Define for ${\bf r}=(r_1,\ldots ,r_N)\in {\Bbb N}^N$, $$ (X^+)^{\bf r}=(X_{\beta_1}^+)^{r_1}\ldots (X_{\beta_N}^+)^{r_N}, $$ $$ (X^-)^{\bf r}=(X_{\beta_1}^-)^{r_1}\ldots (X_{\beta_N}^-)^{r_N}, $$ and for ${\bf s}=(s_1,\ldots s_l)\in {\Bbb N}^{~l}$, $$ H^{\bf s}=H_1^{s_1}\ldots H_l^{s_l}. $$ \begin{proposition}{\bf (\cite{kh-t}, Proposition 3.3)}\label{PBW} The elements $(X^+)^{\bf r}$, $(X^-)^{\bf t}$ and $H^{\bf s}$, for ${\bf r},~{\bf t}\in {\Bbb N}^N$, ${\bf s}\in {\Bbb N}^l$, form topological bases of $U_h({\frak n}_+),U_h({\frak n}_-)$ and $U_h({\frak h})$, respectively, and the products $(X^+)^{\bf r}H^{\bf s}(X^-)^{\bf t}$ form a topological basis of $U_h({\frak g})$. In particular, multiplication defines an isomorphism of ${\Bbb C}[[h]]$ modules: $$ U_h({\frak n}_-)\otimes U_h({\frak h}) \otimes U_h({\frak n}_+)\rightarrow U_h({\frak g}). $$ \end{proposition} An explicit expression for $\cal R$ may be written by making use of the q-exponential $$ exp_q(x)=\sum_{k=0}^\infty {x^k \over (k)_q!}, $$ where $$ (k)_q!=(1)_q\ldots (k)_q,~~(n)_q={q^n -1 \over q-1}. $$ Now the element $\cal R$ may be written as (see Theorem 8.1 in \cite{kh-t}): \begin{equation}\label{univr} {\cal R}=exp\left[ h\sum_{i=1}^l(Y_i\otimes H_i)\right]\prod_{\beta} exp_{q_{\beta}^{-1}}[(q-q^{-1})a(\beta)^{-1}X_{\beta}^+\otimes X_{\beta}^-], \end{equation} where $q_\beta =q^{(\beta,\beta)}$; the product is over all the positive roots of $\frak g$, and the order of the terms is such that the $\alpha$--term appears to the left of the $\beta$--term if $\alpha <\beta$ with respect to the normal ordering of $\Delta_+$. \begin{remark} The r--matrix $r_+=\frac 12 h^{-1}({\cal R}-1\otimes 1)~~(\mbox{mod }h)$, which is the classical limit of $\cal R$, coincides with the classical r--matrix (\ref{rcl}). \end{remark} \section{Non--singular characters and quantum groups} \setcounter{equation}{0} \setcounter{theorem}{0} In this section we construct quantum counterparts of the principal nilpotent Lie subalgebras of complex simple Lie algebras and of their non--singular characters. We mainly follow the exposition presented in \cite{S1}. First we would like to show that the algebra $U_h({\frak n}_+)$ spanned by $X_i^+ , i=1, \ldots , l$ does not admit characters which take nonvanishing values on all generators $X_i^+$, except for the case of $U_h(sl(2))$ when the quantum Serre relations do not appear. Suppose, $\chi_h$ is such a character, and $\chi_h(X_i^+)=c_i\in {\Bbb C}[[h]],~c_i\neq 0,~i=1,\ldots l$. By applying the character $\chi_h$ to the quantum Serre relations one obtains a family of identities, \begin{equation} \label{false} \sum_{r=0}^{1-a_{ij}}(-1)^r \left[ \begin{array}{c} 1-a_{ij} \\ r \end{array} \right]_{q_i} =0 , \, i \neq j. \end{equation} We claim that some of these relations fail for the quantized universal enveloping algebra $U_h({\frak g})$ of any simple Lie algebra $\frak g$ , with the exception of ${\frak g}=sl(2)$. In a more general setting, relations (\ref{false}) are analysed in the following lemma. \begin{lemma}\label{qbinom} The only solutions of equation \begin{equation}\label{c1} \sum_{k=0}^{m}(-1)^k \left[ \begin{array}{c} m \\ k \end{array} \right]_{t} t^{kc}=0 , \end{equation} where $t$ is an indeterminate, are of the form \begin{equation}\label{c2} c=-m+1,-m+2, \ldots ,m-2,m-1. \end{equation} \end{lemma} {\em Proof.} According to the q--binomial theorem \cite{GR}, \begin{equation}\label{z} \sum_{k=0}^{m}(-z)^k \left[ \begin{array}{c} m \\ k \end{array} \right]_{t} =\prod_{p=0}^{m-1}(1-t^{m-1-2p}z). \end{equation} Put $z=t^c$ in this relation. Then the l.h.s of (\ref{z}) coincides with the l.h.s. of (\ref{c1}). Now (\ref{z}) implies that $c=m-1-2p , p=0, \ldots ,m-1$ are the only solutions of (\ref{c1}). Now we return to identities (\ref{false}). Any Cartan matrix contains at least one off-diagonal element equal to $-1$. Then, $m= 1- a_{ij} = 2$ and $c=\pm 1$, and Lemma \ref{qbinom} implies that some of identities (\ref{false}) are false for any simple Lie algebra, except for $sl(2)$. Hence, subalgebras of $U_h({\frak g})$ generated by $X_i^+$ do not possess non--singular characters. It is our goal to construct subalgebras of $U_h({\frak g})$ which resemble the subalgebra $U({\frak n}_+) \subset U({\frak g})$ and possess non--singular characters. Denote by $S_l$ the symmetric group of $l$ elements. To any element $\pi \in S_l$ we associate a Coxeter element $s_{\pi}$ by the formula $s_\pi =s_{\pi (1)}\ldots s_{\pi (l)}$. For each Coxeter element $s_\pi$ we define an associative algebra $U_h^{s_\pi}({\frak n}_+) $ generated by elements $e_i ,~ i=1, \ldots l$ subject to the relations : \begin{equation}\label{fqpi} \sum_{r=0}^{1-a_{ij}}(-1)^r q^{r c_{ij}^{\pi}} \left[ \begin{array}{c} 1-a_{ij} \\ r \end{array} \right]_{q_i} (e_i )^{1-a_{ij}-r}e_j (e_i)^r =0 ,~ i \neq j , \end{equation} where $c_{ij}^{\pi}=\left( {1+s_\pi \over 1-s_\pi }\alpha_i , \alpha_j \right)$ are matrix elements of the Caley transform of $s_\pi$ in the basis of simple roots. \begin{proposition}\label{charf} The map $\chi_h^{s_\pi}:U_h^{s_\pi}({\frak n}_+) \rightarrow {\Bbb C}[[h]]$ defined on generators by $\chi_h^{s_\pi}(e_i)=c_i,~c_i\in {\Bbb C}[[h]],~c_i\neq 0$ is a character of the algebra $U_h^{s_\pi}({\frak n}_+) $. \end{proposition} To show that $\chi_h^{s_\pi}$ is a character of $U_h^{s_\pi}({\frak n}_+) $ it suffices to check that the defining relations (\ref{fqpi}) belong to the kernel of $\chi_h^{s_\pi}$, i.e. \begin{equation}\label{chifqpi} \sum_{r=0}^{1-a_{ij}}(-1)^r q^{r c_{ij}^{\pi}} \left[ \begin{array}{c} 1-a_{ij} \\ r \end{array} \right]_{q_i}=0 ,~ i \neq j . \end{equation} As a preparation for the proof of Proposition \ref{charf} we study the matrix elements of the Caley transform of $s_\pi$ which enter the definition of $U_h^{s_\pi}({\frak n}_+) $. \begin{lemma}\label{tmatrel} The matrix elements of ${1+s_\pi \over 1-s_\pi }$ are of the form : \begin{equation}\label{matrel} c_{ij}^{\pi}=\left( {1+s_\pi \over 1-s_\pi }\alpha_i , \alpha_j \right)= \varepsilon_{ij}^\pi b_{ij}, \end{equation} where $$ \varepsilon_{ij}^\pi =\left\{ \begin{array}{ll} -1 & \pi^{-1}(i) <\pi^{-1}(j) \\ 0 & i=j \\ 1 & \pi^{-1}(i) >\pi^{-1}(j) \end{array} \right . $$ \end{lemma} {\em Proof.} (compare \cite{Bur}, Ch. V, \S 6, Ex. 3). First we calculate the matrix of the Coxeter element $s_\pi$ with respect to the basis of simple roots. We obtain this matrix in the form of the Gauss decomposition of the operator $s_\pi$. Let $z_{\pi (i)}=s_\pi \alpha_{\pi (i)}$. Recall that $s_i(\alpha_j)=\alpha_j-a_{ji}\alpha_i$. Using this definition the elements $z_{\pi (i)}$ may be represented as: $$ z_{\pi (i)}=y_{\pi (i)} -\sum_{k \geq i} a_{\pi (k) \pi (i)}y_{\pi (k)}, $$ where \begin{equation}\label{y} y_{\pi (i)}=s_{\pi (1)}\ldots s_{\pi (i-1)}\alpha_{\pi (i)}. \end{equation} Using the matrix notation we can rewrite the last formula as follows: \begin{equation}\label{2*} \begin{array}{l} z_{\pi (i)}= (I+V)_{\pi (k) \pi (i)}y_{\pi (k)} , \\ \\ \mbox{ where } V_{\pi (k) \pi (i)}= \left\{ \begin{array}{ll} a_{\pi (k) \pi (i)} & k\geq i \\ 0 & k < i \end{array} \right . \end{array} \end{equation} To calculate the matrix of the operator $s_\pi$ with respect to the basis of simple roots we have to express the elements $y_{\pi (i)}$ via the simple roots. Applying the definition of simple reflections to (\ref{y}) we can pull out the element $\alpha_{\pi (i)}$ to the right: \[ y_{\pi (i)}=\alpha_{\pi (i)}-\sum_{k<i}a_{\pi (k) \pi (i)}y_{\pi (k)}. \] Therefore \[ \alpha_{\pi (i)}=(I+U)_{\pi (k) \pi (i)}y_{\pi (k)} ~, \mbox{ where } U_{\pi (k) \pi (i)}= \left\{ \begin{array}{ll} a_{\pi (k) \pi (i)} & k<i \\ 0 & k \geq i \end{array} \right . \] Thus \begin{equation}\label{1*} y_{\pi (k)}=(I+U)^{-1}_{\pi (j) \pi (k)}\alpha_{\pi (j)}. \end{equation} Summarizing (\ref{1*}) and (\ref{2*}) we obtain: \begin{equation}\label{**} s_\pi \alpha_i=\left( (I+U)^{-1}(I-V) \right)_{ki}\alpha_k . \end{equation} This implies: \begin{equation}\label{3*} {1+s_\pi \over 1-s_\pi}\alpha_i=\left( {2I+U-V \over U+V}\right)_{ki}\alpha_k . \end{equation} Observe that $(U+V)_{ki}=a_{ki}$ and $(2I+U-V)_{ij}=-a_{ij}\varepsilon_{ij}^\pi$. Substituting these expressions into (\ref{3*}) we get : \begin{eqnarray} \left( {1+s_\pi \over 1-s_\pi }\alpha_i , \alpha_j \right) = -(a^{-1})_{kp}\varepsilon_{pi}^\pi a_{pi}b_{jk}=\\ -d_ja_{jk}(a^{-1})_{kp}\varepsilon_{pi}^\pi a_{pi} = \varepsilon_{ij}^\pi b_{ij}. \end{eqnarray} This concludes the proof of the lemma. \noindent {\em Proof of Proposition \ref{charf} } Identities (\ref{chifqpi}) follow from Lemma \ref{qbinom} for $t=q_i,~~m=1-a_{ij},~~ c=\varepsilon_{ij}^\pi a_{ij}$ since set of solutions (\ref{c2}) always contains $\pm (m-1)$. Motivated by relations (\ref{fqpi}) we suggest new realizations of the quantum group $U_h({\frak g})$, one for each Coxeter element $s_\pi$. Let $U_h^{s_\pi}({\frak g})$ be the associative algebra over ${\Bbb C}[[h]]$ with generators $e_i , f_i , H_i,~i=1, \ldots l$ subject to the relations: \begin{equation}\label{sqgr} \begin{array}{l} [H_i,H_j]=0,~~ [H_i,e_j]=a_{ij}e_j, ~~ [H_i,f_j]=-a_{ij}f_j,\\ \\ e_i f_j -q^{ c^\pi _{ij}} f_j e_i = \delta _{i,j}{K_i -K_i^{-1} \over q_i -q_i^{-1}} , \\ \\ K_i=e^{d_ihH_i}, \\ \\ \sum_{r=0}^{1-a_{ij}}(-1)^r q^{r c_{ij}^\pi} \left[ \begin{array}{c} 1-a_{ij} \\ r \end{array} \right]_{q_i} (e_i )^{1-a_{ij}-r}e_j (e_i)^r =0 ,~ i \neq j , \\ \\ \sum_{r=0}^{1-a_{ij}}(-1)^r q^{r c_{ij}^\pi} \left[ \begin{array}{c} 1-a_{ij} \\ r \end{array} \right]_{q_i} (f_i )^{1-a_{ij}-r}f_j (f_i)^r =0 ,~ i \neq j . \end{array} \end{equation} It follows that the map $\tau_h^\pi :U_h^{s_\pi}({\frak n}_+) \rightarrow U_h^{s_\pi}({\frak g});~~~e_i\mapsto e_i$ is a {\em natural} embedding of $U_h^{s_\pi}({\frak n}_+) $ into $U_h^{s_\pi}({\frak g})$. From now on we identify $U_h^{s_\pi}({\frak n}_+) $ with the subalgebra in $U_h^{s_\pi}({\frak g})$ generated by $e_i ,i=1, \ldots l$. \begin{theorem} \label{newreal} For every solution $n_{ij}\in {\Bbb C},~i,j=1,\ldots ,l$ of equations \begin{equation}\label{eqpi} d_jn_{ij}-d_in_{ji}=c^\pi_{ij} \end{equation} there exists an algebra isomorphism $\psi_{\{ n\}} : U_h^{s_\pi}({\frak g}) \rightarrow U_h({\frak g})$ defined by the formulas: $$ \begin{array}{l} \psi_{\{ n\}}(e_i)=X_i^+ \prod_{p=1}^lL_p^{n_{ip}},\\ \\ \psi_{\{ n\}}(f_i)=\prod_{p=1}^lL_p^{-n_{ip}}X_i^- , \\ \\ \psi_{\{ n\}}(H_i)=H_i . \end{array} $$ \end{theorem} {\em Proof} is provided by direct verification of defining relations (\ref{sqgr}). The most nontrivial part is to verify deformed quantum Serre relations (\ref{fqpi}). The defining relations of $U_h({\frak g})$ imply the following relations for $\psi_{\{ n\}}(e_i)$, $$ \sum_{k=0}^{1-a_{ij}}(-1)^k \left[ \begin{array}{c} 1-a_{ij} \\ k \end{array} \right]_{q_i} q^{k({d_j}n_{ij}-d_in_{ji})}\psi_{\{ n\}}(e_i)^{1-a_{ij}-k}\psi_{\{ n\}}(e_j)\psi_{\{ n\}}(e_i)^k =0 , $$ for any $i\neq j$. Now using equation (\ref{eqpi}) we arrive to relations (\ref{fqpi}). \begin{remark} The general solution of equation (\ref{eqpi}) is given by \begin{equation}\label{eq3} n_{ji}=\frac 12 (\varepsilon_{ij}^\pi a_{ij} + \frac{s_{ij}}{d_i}), \end{equation} where $s_{ij}=s_{ji}$. \end{remark} We call the algebra $U_h^{s_\pi}({\frak g})$ the Coxeter realization of the quantum group $U_h({\frak g})$ corresponding to the Coxeter element $s_\pi$. \begin{remark} Let $n_{ij}$ be a solution of the homogeneous system that corresponds to (\ref{eqpi}), $$ d_in_{ji}-d_jn_{ij}=0. $$ Then the map defined by \begin{equation} \begin{array}{l} X_i^+ \mapsto X_i^+ \prod_{p=1}^lL_p^{n_{ip}},\\ \\ X_i^- \mapsto \prod_{p=1}^lL_p^{-n_{ip}}X_i^- , \\ \\ H_i \mapsto H_i \end{array} \end{equation} is an automorphism of $U_h({\frak g})$. Therefore for given Coxeter element the isomorphism $\psi_{\{ n\}}$ is defined uniquely up to automorphisms of $U_h({\frak g})$. \end{remark} Now we shall study the algebraic structure of $U_h^{s_\pi}({\frak g})$. Denote by $U_h^{s_\pi}({\frak n}_-) $ the subalgebra in $U_h^{s_\pi}({\frak g})$ generated by $f_i ,i=1, \ldots l$. From defining relations (\ref{sqgr}) it follows that the map $\overline \chi_h^{s_\pi}:U_h^{s_\pi}({\frak n}_-) \rightarrow {\Bbb C}[[h]]$ defined on generators by $\overline \chi_h^{s_\pi}(f_i)=c_i, c_i\in {\Bbb C}[[h]], c_i\neq 0$ is a character of the algebra $U_h^{s_\pi}({\frak n}_-)$. Let $U_h^{s_\pi}({\frak h})$ be the subalgebra in $U_h^{s_\pi}({\frak g})$ generated by $H_i,~i=1,\ldots ,l$. Define $U_h^{s_\pi}({\frak b}_\pm)=U_h^{s_\pi}({\frak n}_\pm)U_h^{s_\pi}({\frak h})$. We shall construct a Poincar\'{e}--Birkhoff-Witt basis for $U_h^{s_\pi}({\frak g})$. It is convenient to introduce an operator $K\in {\rm End}~{\frak h}$ such that \begin{equation}\label{Kdef} KH_i=\sum_{j=1}^l{n_{ij} \over d_i}Y_j. \end{equation} In particular, we have $$ {n_{ji} \over d_j}=(KH_j,H_i). $$ Equation (\ref{eqpi}) is equivalent to the following equation for the operator $K$: $$ K-K^* = {1+s_\pi \over 1-s_\pi}. $$ \begin{proposition}\label{rootss} (i)For any solution of equation (\ref{eqpi}) and any normal ordering of the root system $\Delta_+$ the elements $e_{\beta}=\psi_{\{ n\}}^{-1}(X_{\beta}^+e^{hK\beta^\vee})$ and $f_{\beta}=\psi_{\{ n\}}^{-1}(e^{-hK\beta^\vee}X_{\beta}^-),~\beta \in \Delta_+$ lie in the subalgebras $U_h^{s_\pi}({\frak n}_+)$ and $U_h^{s_\pi}({\frak n}_-)$, respectively. (ii)Moreover, the elements $e^{\bf r}=e_{\beta_1}^{r_1}\ldots e_{\beta_N}^{r_N},~~f^{\bf t}=e_{\beta_1}^{t_1}\ldots e_{\beta_N}^{t_N}$ and $H^{\bf s}=H_1^{s_1}\ldots H_l^{s_l}$ for ${\bf r},~{\bf t},~{\bf s}\in {\Bbb N}^N$, form topological bases of $U_h^{s_\pi}({\frak n}_+),~U_h^{s_\pi}({\frak n}_-)$ and $U_h^{s_\pi}({\frak h})$, and the products $f^{\bf t}H^{\bf s}e^{\bf r}$ form a topological basis of $U_h^{s_\pi}({\frak g})$. In particular, multiplication defines an isomorphism of ${\Bbb C}[[h]]$ modules$$ U_h^{s_\pi}({\frak n}_-)\otimes U_h^{s_\pi}({\frak h})\otimes U_h^{s_\pi}({\frak n}_+)\rightarrow U_h^{s_\pi}({\frak g}). $$ \end{proposition} {\em Proof.} Let $\beta=\sum_{i=1}^l m_i\alpha_i \in \Delta_+$ be a positive root, $X_{\beta}^+\in U_h({\frak g})$ the corresponding root vector. Then $\beta^\vee=\sum_{i=1}^l m_id_iH_i$, and so $K\beta^\vee=\sum_{i,j=1}^l m_in_{ij}Y_j$. Now the proof of the first statement follows immediately from Proposition \ref{rootprop}, commutation relations (\ref{weight-root}) and the definition of the isomorphism $\psi_{\{ n\}}$. The second assertion is a consequence of Proposition \ref{PBW}. Now we would like to choose a normal ordering of the root system $\Delta_+$ in such a way that $\chi_h^{s_\pi}(e_{\beta})=0$ and $\overline \chi_h^{s_\pi}(f_{\beta})=0$ if $\beta$ is not a simple root. \begin{proposition}\label{rootsh} Choose a normal ordering of the root system $\Delta_+$ such that the simple roots are written in the following order: $\alpha_{\pi (1)},\ldots ,\alpha_{\pi (l)}$. Then $\chi_h^{s_\pi}(e_{\beta})=0$ and $\overline \chi_h^{s_\pi}(f_{\beta})=0$ if $\beta$ is not a simple root. \end{proposition} {\em Proof.} We shall consider the case of positive root generators. The proof for negative root generators is similar to that for the positive ones. The root vectors $X_{\beta}^+$ are defined in terms of iterated q-commutators (see (\ref{rootvect})). Therefore it suffices to verify that for $i<j$ $$ \begin{array}{l} \chi_h^{s_\pi}(e_{\alpha_{\pi(i)}+\alpha_{\pi(j)}})=\\ \\ \chi_h^{s_\pi}(\psi_{\{ n\}}^{-1}( (X_{\pi(i)}^+X_{\pi(j)}^+ - q^{(\alpha_{\pi(i)},\alpha_{\pi(j)})}X_{\pi(j)}^+X_{\pi(i)}^+) e^{hK(d_{\pi(i)}H_{\pi(i)}+d_{\pi(j)}H_{\pi(j)})}))=0. \end{array} $$ From (\ref{Kdef}) and commutation relations (\ref{weight-root}) we obtain that \begin{equation}\label{bebe} \begin{array}{l} \psi_{\{ n\}}^{-1}((X_{\pi(i)}^+X_{\pi(j)}^+ - q^{(\alpha_{\pi(i)},\alpha_{\pi(j)})}X_{\pi(j)}^+X_{\pi(i)}^+) e^{hK(d_{\pi(i)}H_{\pi(i)}+d_{\pi(j)}H_{\pi(j)})})= \\ \\ q^{-d_{\pi(j)}n_{\pi(i)\pi(j)}}(e_{\pi(i)}e_{\pi(j)} - q^{b_{\pi(i)\pi(j)}+d_{\pi(j)}n_{\pi(i)\pi(j)}-d_{\pi(i)}n_{\pi(j)\pi(i)}}e_{\pi (j)}e_{\pi(i)}) \end{array} \end{equation} Now using equation (\ref{eqpi}) and Lemma \ref{tmatrel} the combination $b_{\pi(i)\pi(j)}+d_{\pi(j)}n_{\pi(i)\pi(j)}-d_{\pi(i)}n_{\pi(j)\pi(i)}$ may be represented as $b_{\pi(i)\pi(j)}+\varepsilon_{\pi(i)\pi(j)}^\pi b_{\pi(i)\pi(j)}$. But $\varepsilon_{\pi(i)\pi(j)}^\pi =-1$ for $i<j$ and therefore the r.h.s. of (\ref{bebe}) takes the form $$ q^{-d_{\pi(j)}n_{\pi(i)\pi(j)}}[e_{\pi(i)},e_{\pi(j)}]. $$ Clearly, $$ \chi_h^{s_\pi}(e_{\alpha_{\pi(i)}+\alpha_{\pi(j)}})= q^{-d_{\pi(j)}n_{\pi(i)\pi(j)}}\chi_h^{s_\pi}([e_{\pi(i)},e_{\pi(j)}])=0. $$ \section{Quantum deformation of the Whittaker model} \setcounter{equation}{0} \setcounter{theorem}{0} In this section we define a quantum deformation of the Whittaker model $W({\frak b}_-)$. Our construction is similar the one described in Section \ref{whitt}, the quantum group $U_h^{s_\pi}({\frak g})$, the subalgebra $U_h^{s_\pi}({\frak n}_+)$ and characters ${\chi_h^{s_\pi}}:U_h^{s_\pi}({\frak n}_+) \rightarrow {\Bbb C}[[h]]$ serve as natural counterparts of the universal enveloping algebra $U({\frak g})$, of the subalgebra $U({\frak n}_+)$ and of non--singular characters $\chi:U({\frak n}_+) \rightarrow {\Bbb C}$. Let ${U_h^{s_\pi}({\frak n}_+)}_{\chi_h^{s_\pi}}$ be the kernel of the character $\chi_h^{s_\pi}:U_h^{s_\pi}({\frak n}_+) \rightarrow {\Bbb C}[[h]]$ so that one has a direct sum $$ U_h^{s_\pi}({\frak n}_+)={\Bbb C}[[h]]\oplus {U_h^{s_\pi}({\frak n}_+)}_{\chi_h^{s_\pi}}. $$ From Proposition \ref{rootss} we have a linear isomorphism $U_h^{s_\pi}({\frak g})=U_h^{s_\pi}({\frak b}_-)\otimes U_h^{s_\pi}({\frak n}_+)$ and hence the direct sum \begin{equation}\label{maindecq} U_h^{s_\pi}({\frak g})=U_h^{s_\pi}({\frak b}_-) \oplus I_{\chi_h^{s_\pi}}, \end{equation} where $I_{\chi_h^{s_\pi}}=U_h^{s_\pi}({\frak g}){U_h^{s_\pi}({\frak n}_+)}_{\chi_h^{s_\pi}}$ is the left--sided ideal generated by ${U_h^{s_\pi}({\frak n}_+)}_{\chi_h^{s_\pi}}$. For any $u\in U_h^{s_\pi}({\frak g})$ let $u^{\chi_h^{s_\pi}}\in U_h^{s_\pi}({\frak b}_-)$ be its component in $U_h^{s_\pi}({\frak b}_-)$ relative to the decomposition (\ref{maindecq}). Denote by $\rho_{\chi_h^{s_\pi}}$ the linear map $$ \rho_{\chi_h^{s_\pi}} : U_h^{s_\pi}({\frak g}) \rightarrow U_h^{s_\pi}({\frak b}_-) $$ given by $\rho_{\chi_h^{s_\pi}} (u)=u^{\chi_h^{s_\pi}}$. Denote by $Z_h^{s_\pi}({\frak g})$ the center of $U_h^{s_\pi}({\frak g})$. From Proposition \ref{zq} and Theorem \ref{newreal} we obtain that $Z_h^{s_\pi}({\frak g})\cong Z({\frak g})[[h]]$. In particular, $Z_h^{s_\pi}({\frak g})$ is freely generated as a commutative topological algebra over ${\Bbb C}[[h]]$ by $l$ elements $I_1,\ldots , I_l$. Let $W_h({\frak b}_-)=\rho_{\chi_h^{s_\pi}} (Z_h^{s_\pi}({\frak g}))$. \vskip 0.3cm \noindent {\bf Theorem $\bf A_h$ } {\em The map \begin{equation}\label{mapq} \rho_{\chi_h^{s_\pi}} : Z_h^{s_\pi}({\frak g}) \rightarrow W_h({\frak b}_-) \end{equation} is an isomorphism of algebras. In particular, $W_h({\frak b}_-)$ is freely generated as a commutative topological algebra over ${\Bbb C}[[h]]$ by $l$ elements $I_i^{\chi_h^{s_\pi}}=\rho_{\chi_h^{s_\pi}}(I_i),~~i=1,\ldots ,l$.} \vskip 0.3cm \noindent {\em Proof} is similar to that of Theorem A in the classical case. \vskip 0.3cm \noindent {\bf Definition $\bf A_h$ } {\em The algebra $W_h({\frak b}_-)$ is called the Whittaker model of $Z_h^{s_\pi}({\frak g})$.} \vskip 0.3cm Next we equip $U_h^{s_\pi}({\frak b}_-)$ with a structure of a left $U_h^{s_\pi}({\frak n}_+)$ module in such a way that $W_h({\frak b}_-)$ is identified with the space of invariants with respect to this action. Following Lemma A in the classical case we define this action by \begin{equation}\label{mainactq} x\cdot v =[x,v]^{\chi_h^{s_\pi}}, \end{equation} where $v\in U_h^{s_\pi}({\frak b}_-)$ and $x\in U_h^{s_\pi}({\frak n}_+)$. Consider the space $U_h^{s_\pi}({\frak b}_-)^{U_h^{s_\pi}({\frak n}_+)}$ of $U_h^{s_\pi}({\frak n}_+)$ invariants in $U_h^{s_\pi}({\frak b}_-)$ with respect to this action. Clearly, $W_h({\frak b}_-)\subseteq U_h^{s_\pi}({\frak b}_-)^{U_h^{s_\pi}({\frak n}_+)}$. \vskip 0.3cm \noindent {\bf Theorem $\bf B_h$ } {\em Suppose that $\chi_h^{s_\pi}(e_i)\neq 0~(\mbox{mod }h)$ for $i=1,\ldots l$. Then the space of $U_h^{s_\pi}({\frak n}_+)$ invariants in $U_h^{s_\pi}({\frak b}_-)$ with respect to the action (\ref{mainactq}) is isomorphic to $W_h({\frak b}_-)$, i.e.} \begin{equation}\label{invq} U_h^{s_\pi}({\frak b}_-)^{U_h^{s_\pi}({\frak n}_+)}\cong W_h({\frak b}_-). \end{equation} \vskip 0.3cm \noindent {\em Proof.} Let $p: U_h^{s_\pi}({\frak g})\rightarrow U_h^{s_\pi}({\frak g})/hU_h^{s_\pi}({\frak g})=U({\frak g})$ be the canonical projection. Note that $p(U_h^{s_\pi}({\frak n}_+))=U({\frak n}_+),~ p(U_h^{s_\pi}({\frak b}_-))=U({\frak b}_-)$ and for every $x\in U_h^{s_\pi}({\frak n}_+)$ $\chi_h^{s_\pi}(x)~(\mbox{mod }h)=\chi(p(x))$ for some non--singular character $\chi: U({\frak n}_+)\rightarrow {\Bbb C}$. Therefore $p(\rho_{\chi_h^{s_\pi}}(x))=\rho_\chi(p(x))$ for every $x\in U_h^{s_\pi}({\frak g})$, and hence by Theorem ${\rm A}_q$ $p(W_h({\frak b}_-))=W({\frak b}_-)$. Using Lemma A and the definition of action (\ref{mainactq}) we also obtain that $p(U_h^{s_\pi}({\frak b}_-)^{U_h^{s_\pi}({\frak n}_+)})= U({\frak b}_-)^{N_+} = W({\frak b}_-)$. Now let $I\in U_h^{s_\pi}({\frak b}_-)^{U_h^{s_\pi}({\frak n}_+)}$ be an invariant element. Then $p(I)\in W({\frak b}_-)$, and hence one can find an element $K_0\in W_h({\frak b}_-)$ such that $I-K_0=hI_1,~I_1\in U_h^{s_\pi}({\frak b}_-)^{U_h^{s_\pi}({\frak n}_+)}$. Applying the same procedure to $I_1$ one can find elements $K_1\in W_h({\frak b}_-), ~I_2\in U_h^{s_\pi}({\frak b}_-)^{U_h^{s_\pi}({\frak n}_+)}$ such that $I_1-K_1=hI_2$, i.e. $I-K_0-hK_1=0~(\mbox{mod }h^2)$. We can continue this process. Finally we obtain an infinite sequence of elements $K_i\in W_h({\frak b}_-)$ such that $I-\sum_{i=0}^p h^pK_p=0~(\mbox{mod }h^{p+1})$. Since the space $U_h^{s_\pi}({\frak b}_-)$ is complete in the $h$--adic topology the series $\sum_{i=0}^\infty h^pK_p\in W_h({\frak b}_-)$ converges to $I$. Therefore $I\in W_h({\frak b}_-)$. This completes the proof. Similarly to Proposition \ref{hkhom} we have \begin{proposition} The algebra $W_h({\frak b}_-)$ is isomorphic to the zeroth graded component of the Hecke algebra of the triple $(U_h^{s_\pi}({\frak g}),U_h^{s_\pi}({\frak n}_+),\chi_h^{s_\pi})$ with the opposite multiplication, $$ W_h({\frak b}_-)=Hk^0(U_h^{s_\pi}({\frak g}),U_h^{s_\pi}({\frak n}_+),\chi_h^{s_\pi})^{opp}. $$ \end{proposition} \section{Coxeter realizations of quantum groups and Drinfeld twist} \setcounter{equation}{0} \setcounter{theorem}{0} In this section we show that the Coxeter realizations $U_h^{s_\pi}({\frak g})$ of the quantum group $U_h({\frak g})$ are connected with quantizations of some nonstandard bialgebra structures on $\frak g$. At the quantum level changing bialgebra structure corresponds to the so--called Drinfeld twist. We shall consider a particular class of such twists described in the following proposition. \begin{proposition}{\bf (\cite{ChP}, Proposition 4.2.13)}\label{twdef} Let $(A,\mu , \imath , \Delta , \varepsilon , S)$ be a Hopf algebra over a commutative ring. Let $\cal F$ be an invertible element of $A\otimes A$ such that \begin{equation}\label{twist} \begin{array}{l} {\cal F}_{12}(\Delta \otimes id)({\cal F})={\cal F}_{23}(id \otimes \Delta)({\cal F}),\\ \\ (\varepsilon \otimes id)({\cal F})=(id \otimes \varepsilon )({\cal F})=1. \end{array} \end{equation} Then, $v=\mu (id\otimes S)({\cal F})$ is an invertible element of $A$ with $$ v^{-1}=\mu (S\otimes id)({\cal F}^{-1}). $$ Moreover , if we define $\Delta^{\cal F}:A\rightarrow A\otimes A$ and $S^{\cal F}:A\rightarrow A$ by $$ \Delta^{\cal F}(a)={\cal F}\Delta(a){\cal F}^{-1},~~S^{\cal F}(a)=vS(a)v^{-1}, $$ then $(A,\mu , \imath , \Delta^{\cal F} , \varepsilon , S^{\cal F})$ is a Hopf algebra denoted by $A^{\cal F}$ and called the twist of $A$ by ${\cal F}$. \end{proposition} \begin{corollary}{\bf (\cite{ChP}, Corollary 4.2.15)} Suppose that $A$ and ${\cal F}$ as in Proposition \ref{twdef}, but assume in addition that $A$ is quasitriangular with universal R--matrix $\cal R$. Then $A^{\cal F}$ is quasitriangular with universal R--matrix \begin{equation}\label{rf} {\cal R}^{\cal F}={\cal F}_{21}{\cal R}{\cal F}^{-1}, \end{equation} where ${\cal F}_{21}=\sigma {\cal F}$. \end{corollary} Fix a Coxeter element $s_\pi\in W$, $s_\pi=s_{\pi (1)}\ldots s_{\pi (l)}$. Consider the twist of the Hopf algebra $U_h({\frak g})$ by the element \begin{equation}\label{Ftw} {\cal F}=exp(-h\sum_{i,j=1}^l {n_{ji} \over d_j}Y_i\otimes Y_j) \in U_h({\frak h})\otimes U_h({\frak h}), \end{equation} where $n_{ij}$ is a solution of the corresponding equation (\ref{eqpi}). This element satisfies conditions (\ref{twist}), and so $U_h({\frak g})^{\cal F}$ is a quasitriangular Hopf algebra with the universal R--matrix ${\cal R}^{\cal F}={\cal F}_{21}{\cal R}{\cal F}^{-1}$, where $\cal R$ is given by (\ref{univr}). We shall explicitly calculate the element ${\cal R}^{\cal F}$. Substituting (\ref{univr}) and (\ref{Ftw}) into (\ref{rf}) and using (\ref{roots-cart}) we obtain $$ \begin{array}{l} {\cal R}^{\cal F}=exp\left[ h(\sum_{i=1}^l(Y_i\otimes H_i)+ \sum_{i,j=1}^l (-{n_{ij} \over d_i}+{n_{ji} \over d_j})Y_i\otimes Y_j) \right]\times \\ \prod_{\beta} exp_{q_{\beta}^{-1}}[(q-q^{-1})a(\beta)^{-1}X_{\beta}^+e^{hK\beta^\vee} \otimes e^{-hK^*\beta^\vee}X_{\beta}^-], \end{array} $$ where $K$ is defined by (\ref{Kdef}). Equip $U_h^{s_\pi}({\frak g})$ with the comultiplication given by : $\Delta_{s_\pi}(x)=(\psi_{\{ n\}}^{-1}\otimes \psi_{\{ n\}}^{-1})\Delta_h^{\cal F}(\psi_{\{ n\}}(x))$. Then $U_h^{s_\pi}({\frak g})$ becomes a quasitriangular Hopf algebra with the universal R--matrix ${\cal R}^{s_\pi}=\psi_{\{ n\}}^{-1}\otimes \psi_{\{ n\}}^{-1}{\cal R}^{\cal F}$. Using equation (\ref{eqpi}) and Lemma \ref{tmatrel} this R--matrix may be written as follows \begin{equation}\label{rmatrspi} \begin{array}{l} {\cal R}^{s_\pi}=exp\left[ h(\sum_{i=1}^l(Y_i\otimes H_i)+ \sum_{i=1}^l {1+s_\pi \over 1-s_\pi }H_i\otimes Y_i) \right]\times \\ \prod_{\beta} exp_{q_{\beta}^{-1}}[(q-q^{-1})a(\beta)^{-1}e_{\beta} \otimes e^{h{1+s_\pi \over 1-s_\pi} \beta^\vee}f_{\beta}]. \end{array} \end{equation} The element ${\cal R}^{s_\pi}$ may be also represented in the form \begin{equation}\label{rmatrspi'} \begin{array}{l} {\cal R}^{s_\pi}=exp\left[ h(\sum_{i=1}^l(Y_i\otimes H_i)\right]\times \\ \prod_{\beta} exp_{q_{\beta}^{-1}}[(q-q^{-1})a(\beta)^{-1}e_{\beta}e^{-h{1+s_\pi \over 1-s_\pi}\beta^\vee}\otimes f_\beta] exp\left[ h(\sum_{i=1}^l {1+s_\pi \over 1-s_\pi }H_i\otimes Y_i)\right] . \end{array} \end{equation} The comultiplication $\Delta_{s_\pi}$ is given on generators by $$ \begin{array}{l} \Delta_{s_\pi}(H_i)=H_i\otimes 1+1\otimes H_i,\\ \\ \Delta_{s_\pi}(e_i)=e_i\otimes e^{hd_i{2 \over 1-s_\pi}H_i}+1\otimes e_i,\\ \\ \Delta_{s_\pi}(f_i)=f_i\otimes e^{-hd_i{1+s_\pi \over 1-s_\pi}H_i}+e^{-hd_iH_i}\otimes f_i. \end{array} $$ Note that the Hopf algebra $U_h^{s_\pi}({\frak g})$ is a quantization of the bialgebra structure on $\frak g$ defined by the cocycle \begin{equation}\label{cocycles} \delta (x)=({\rm ad}_x\otimes 1+1\otimes {\rm ad}_x)2r^{s_\pi}_+,~~ r^{s_\pi}_+\in {\frak g}\otimes {\frak g}, \end{equation} where $r^{s_\pi}_+=r_+ + \frac 12 \sum_{i=1}^l {1+s_\pi \over 1-s_\pi }H_i\otimes Y_i$, and $r_+$ is given by (\ref{rcl}). We shall also need the following property of the antipode $S^{s_\pi}$ of $U_h^{s_\pi}({\frak g})$. \begin{proposition}\label{sqant} The square of the antipode $S^{s_\pi}$ is an inner automorphism of $U_h^{s_\pi}({\frak g})$ given by $$ (S^{s_\pi})^2(x)=e^{2h\rho^\vee}xe^{-2h\rho^\vee}, $$ where $\rho^\vee=\sum_{i=1}^lY_i$. \end{proposition} {\em Proof.} First observe that by Proposition \ref{twdef} the antipode of $U_h^{s_\pi}({\frak g})$ has the form: $S^{s_\pi}(x)=\psi_{\{ n\}}^{-1}(vS_h(\psi_{\{ n\}}(x))v^{-1})$, where $$ v=exp(h\sum_{i,j=1}^l {n_{ji} \over d_j}Y_iY_j). $$ Therefore $(S^{s_\pi})^2(x)=\psi_{\{ n\}}^{-1}(vS_h(v^{-1})S_h^2(\psi_{\{ n\}}(x))S_h(v)v^{-1})$. Note that $S_h(v)=v$, and hence $(S^{s_\pi})^2(x)=\psi_{\{ n\}}^{-1}(S_h^2(\psi_{\{ n\}}(x)))$. Finally observe that from explicit formulas for the antipode of $U_h({\frak g})$ it follows that $S_h^2(x)=e^{2h\rho^\vee}xe^{-2h\rho^\vee}$. This completes the proof. In conclusion we note that using Corollary \ref{rep} and the isomorphism $\psi_{\{ n\}}$ one can define finite--dimensional representations of $U_h^{s_\pi}({\frak g})$. \section{Quantum deformation of the Toda lattice}\label{toda} \setcounter{equation}{0} \setcounter{theorem}{0} Recall that one of the main applications of the algebra $W({\frak b}_-)$ is the quantum Toda lattice \cite{K'}. Let $\overline \chi : {\frak n}_- \rightarrow {\Bbb C}$ be a non--singular character of the opposite nilpotent subalgebra ${\frak n}_-$. We denote the character of $N_-$ corresponding to $\overline \chi$ by the same letter. The algebra $U({\frak b}_-)$ naturally acts by differential operators in the space $C^\infty ({\Bbb C}_{\overline \chi}\otimes_{N_-}{B_-})$. This space may be identified with $C^\infty (H)$. Let $D_1,\ldots ,D_l$ be the differential operators on $C^\infty (H)$ which correspond to the elements $I_1^\chi ,\ldots , I_l^\chi\in W({\frak b}_-)$. Denote by $\varphi$ the operator of multiplication in $C^\infty (H)$ by the function $\varphi (e^h)=e^{\rho(h)}$, where $h\in {\frak h}$. The operators $M_i=\varphi D_i\varphi^{-1}, i=1,\ldots l$ are called the quantum Toda Hamiltonians. Clearly, they commute with each other. In particular if $I$ is the quadratic Casimir element then the corresponding operator $M$ is the well--known second--order differential operator: $$ M=\sum_{i=1}^l \partial_i^2 +\sum_{i=1}^l \chi(X_{\alpha_i})\overline \chi(X_{-\alpha_i})e^{-\alpha_i(h)}+(\rho,\rho), $$ where $\partial_i={\partial \over \partial y_i}$, and $y_i,~i=1,\ldots l$ is an ortonormal basis of $\frak h$. Using the algebra $W_h({\frak b}_-)$ we shall construct quantum group analogues of the Toda Hamiltonians. A slightly different approach has been recently proposed in \cite{Et}. Denote by $A$ the space of linear functions on ${\Bbb C}[[h]]_{\overline \chi_h^{s_\pi}}\otimes_{U_h^{s_\pi}({\frak n}_-)}U_h^{s_\pi}({\frak b}_-)$, where ${\Bbb C}[[h]]_{\overline \chi_h^{s_\pi}}$ is the one--dimensional $U_h^{s_\pi}({\frak n}_-)$ module defined by ${\overline \chi_h^{s_\pi}}$. Note that ${\Bbb C}[[h]]_{\overline \chi_h^{s_\pi}}\otimes_{U_h^{s_\pi}({\frak n}_-)}U_h^{s_\pi}({\frak b}_-)\cong U_h^{s_\pi}({\frak h})$ as a linear space. Therefore $A=U_h^{s_\pi}({\frak h})^*$. The algebra $U_h^{s_\pi}({\frak b}_-)$ naturally acts on ${\Bbb C}[[h]]_{\overline \chi_h^{s_\pi}}\otimes_{U_h^{s_\pi}({\frak n}_-)}U_h^{s_\pi}({\frak b}_-)$ by multiplications from the right. This action induces an $U_h^{s_\pi}({\frak b}_-)$--action in the space $A$. We denote this action by $L$, $L:U_h^{s_\pi}({\frak b}_-)\rightarrow {\rm End}A$. Clearly, this action generates an action of the algebra $W_h({\frak b}_-)$ on $A$. To construct deformed Toda Hamiltonians we shall use certain elements in $W_h({\frak b}_-)$. These elements may be described as follows. Let $\mu : U_h^{s_\pi}({\frak g}) \rightarrow {\Bbb C}[[h]]$ be a map such that $\mu(uv)=\mu(vu)$. By Proposition \ref{sqant} $(S^{s_\pi})^2(x)=e^{2h\rho^\vee}xe^{-2h\rho^\vee}$. Hence from Remark 1 in \cite{D} it follows that $(id\otimes \mu)({\cal R}_{21}^{s_\pi}{\cal R}^{s_\pi}(1\otimes e^{2h\rho^\vee}))$, where ${\cal R}_{21}^{s_\pi}=\sigma {\cal R}^{s_\pi}$, is a central element. In particular, for any finite--dimensional $\frak g$--module $V$ the element \begin{equation}\label{centrelv} C_V=(id\otimes tr_V)({\cal R}_{21}^{s_\pi}{\cal R}^{s_\pi}(1\otimes e^{2h\rho^\vee})), \end{equation} where $tr_V$ is the trace in $V[[h]]$, is central in $U_h^{s_\pi}({\frak g})$. Using formulas (\ref{rmatrspi}) and (\ref{rmatrspi'}) we can easily compute elements $\rho_{\chi_h^{s_\pi}}(C_V)\in W_h({\frak b}_-)$. For every finite--dimensional $\frak g$--module $V$ we have \begin{equation}\label{todah} \begin{array}{l} \rho_{\chi_h^{s_\pi}}(C_V)=(id\otimes tr_V)( e^{t_0}\prod_{\beta} exp_{q_{\beta}^{-1}}[(q-q^{-1})a(\beta)^{-1}f_\beta \otimes e_{\beta}e^{-h{1+s_\pi \over 1-s_\pi}\beta^\vee}]\times \\ \\ e^{t_0}\prod_{\beta} exp_{q_{\beta}^{-1}}[(q-q^{-1})a(\beta)^{-1}{\chi_h^{s_\pi}}(e_{\beta}) \otimes e^{h{1+s_\pi \over 1-s_\pi} \beta^\vee}f_{\beta}](1\otimes e^{2h\rho^\vee})), \end{array} \end{equation} where $t_0=h\sum_{i=1}^l(Y_i\otimes H_i)$. We denote by $W_h^{Rep}({\frak b}_-)$ the subalgebra in $W_h({\frak b}_-)$ generated by the elements $\rho_{\chi_h^{s_\pi}}(C_V)$, where $V$ runs through all finite--dimensional representations of $\frak g$. Note that for every finite--dimensional $\frak g$--module $V$ $\rho_{\chi_h^{s_\pi}}(C_V)$ is a polynomial in noncommutative elements $f_i,~e^{hx},~x\in {\frak h}$. Now we shall realize elements of $W_h^{Rep}({\frak b}_-)$ as difference operators. Let $H_h\in U_h^{s_\pi}({\frak h})$ be the subgroup generated by elements $e^{hx},~x\in {\frak h}$. A difference operator on $A$ is an operator $T$ of the form $T=\sum f_iT_{x_i}$ (a finite sum), where $f_i \in A$, and for every $y\in H_h~$ $T_{x}f(y)=(ye^{hx}),~x\in {\frak h}$. \begin{proposition}{\bf (\cite{Et}, Proposition 3.2)}\label{diffh} For any $Y \in U_h^{s_\pi}({\frak b}_-)$, which is a polynomial in noncommutative elements $f_i,~e^{hx},~x\in {\frak h}$, the operator $L(Y)$ is a difference operator on $A$. In particular, the operators $L(I),~I\in W_h^{Rep}({\frak b}_-)$ are mutually commuting difference operators on $A$. \end{proposition} {\em Proof.} It suffices to verify that $L(f_i)$ are difference operators on $H_h$. Indeed, $$ L(f_i)f(e^{hx})=f(e^{hx}f_i)=e^{-h\alpha_i(x)}f(f_ie^{hx})=\overline \chi_h^{s_\pi}(f_i)e^{-h\alpha_i(x)}f(e^{hx}). $$ This completes the proof. Let $\jmath : H_h\rightarrow U_h^{s_\pi}({\frak h})$ be the canonical embedding. Denote $A_h=\jmath^*(A)$. Let $T$ be a difference operator on $A$. Then one can define a difference operator $\jmath^*(T)$ on the space $A_h$ by $\jmath^*(T)f(y)=T(\jmath(y))$. Let $D_i^h=\jmath^*(L(\rho_{\chi_h^{s_\pi}}(C_{V_i})))$, where $V_i,~i=1,\ldots l$ are the fundamental representations of $\frak g$. Denote by $\varphi_h$ the operator of multiplication in $A_h$ by the function $\varphi_h (e^{hx})=e^{h\rho(x)}$, where $x\in {\frak h}$. The operators $M_i^h=\varphi_h D_i^h\varphi^{-1}_h, i=1,\ldots l$ are called the quantum deformed Toda Hamiltonians. From now on we suppose that $\pi=id$ and that the ordering of positive roots $\Delta_+$ is fixed as in Proposition \ref{rootsh}. We denote $s_{id}=s$. Now using formula (\ref{todah}) we outline computation of the operators $M_i^h$. This computation is simplified by the following lemma. \begin{lemma}{ \bf (\cite{Et}, Lemma 5.2)} Let $X=f_{\gamma_1}...f_{\gamma_n}$. If the roots $\gamma_1,...,\gamma_n$ are not all simple then $L(X)=0$. Otherwise, if $\gamma_i=\alpha_{k_i}$, then $$ \jmath^*(L(X))f(e^{hy})=e^{-h(\sum\alpha_{k_i},y)}f(e^{hy})\prod_i\overline \chi_h^{s}(f_{k_i}) $$ \end{lemma} {\em Proof } follows immediately from Proposition \ref{rootsh} and the arguments used in the proof of Proposition \ref{diffh}. Using this lemma we obtain that if $\beta$ is not a simple root then the term in (\ref{todah}) containing root vector $f_\beta$ gives a trivial contribution to the operators $L(\rho_{\chi_h^{s}}(C_{V_i}))$. Note also that by Proposition \ref{rootsh} ${\chi_h^{s}}(e_\beta)=0$ if $\beta$ is not a simple root. Therefore from formula (\ref{todah}) we have \begin{equation} \begin{array}{l} L(\rho_{\chi_h^{s}}(C_{V_i}))=\\ \\ L(id\otimes tr_V)( e^{t_0}\prod_{i} exp_{q^{-2d_i}}[(q_i-q_i^{-1})f_i \otimes e_ie^{-hd_i{1+s \over 1-s}H_i}]\times \\ \\ e^{t_0}\prod_{i} exp_{q^{-2d_i}}[(q_i-q_i^{-1}){\chi_h^{s}}(e_i) \otimes e^{hd_i{1+s \over 1-s}H_i}f_i](1\otimes e^{2h\rho^\vee})). \end{array} \end{equation} In particular, let ${\frak g}=sl(n)$, $V_1=V$ the fundamental representation of $sl(n)$. Then direct calculation gives $$ M_1f(e^{hy})=\left( \sum_{j=1}^n T_j^2- (q-q^{-1})^2\sum_{i=1}^{n-1}{\chi_h^{s}}(e_i){\overline \chi_h^{s}}(f_i) e^{-h(y,\alpha_i)}T_{i+1}T_i\right) f(e^{hy}), $$ where $T_i=T_{\omega_i}$, ${\omega_i}$ are the weights of $V$. The last expression coincides with formula (5.7) obtained in \cite{Et}. \chapter{Poisson--Lie groups and Whittaker model}\label{GWitt} In this Chapter we introduce another quantum version of the Whittaker model. We consider quantizations of algebras of regular functions on algebraic Poisson--Lie groups. We define the Whittaker model of the center of these quantum algebras. The algebraic structure of this model is related to the structure of the set of regular elements in the corresponding algebraic group. This relation is parallel to the one established by Kostant for Lie algebras (see Section \ref{geomappr}). Our main geometric result is an analog of Theorem C for algebraic groups. A generalization of this theorem for loop groups is contained in \cite{SS}. \section{Poisson--Lie groups} \setcounter{equation}{0} \setcounter{theorem}{0} Recall some notions concerned with Poisson--Lie groups (see \cite{Dm}, \cite{fact}, \cite{dual}, \cite{ChP}). Let $G$ be a finite--dimensional Lie group equipped with a Poisson bracket, $\frak g$ its Lie algebra. $G$ is called a Poisson--Lie group if the multiplication $G\times G \rightarrow G$ is a Poisson map. A Poisson bracket satisfying this axiom is degenerate and, in particular, is identically zero at the unit element of the group. Linearizing this bracket at the unit element defines the structure of a Lie algebra in the space $T^*_eG\simeq {\frak g}^*$. The pair (${\frak g},{\frak g}^{*})$ is called the tangent bialgebra of $G$. Lie brackets in $\frak{g}$ and $\frak{g}^{*}$ satisfy the following compatibility condition: {\em Let }$\delta: {\frak g}\rightarrow {\frak g}\wedge {\frak g}$ {\em be the dual of the commutator map } $[,]_{*}: {\frak g}^{*}\wedge {\frak g}^{*}\rightarrow {\frak g}^{*}$. {\em Then } $\delta$ {\em is a 1-cocycle on} $ {\frak g}$ {\em (with respect to the adjoint action of } $\frak g$ {\em on} ${\frak g}\wedge{\frak g}$). Let $c_{ij}^{k}, f^{ab}_{c}$ be the structure constants of ${\frak g}, {\frak g}^{*}$ with respect to the dual bases $\{e_{i}\}, \{e^{i}\}$ in ${\frak g},{\frak g}^{*}$. The compatibility condition means that $$ c_{ab}^{s} f^{ik}_{s} ~-~ c_{as}^{i} f^{sk}_{b} ~+~ c_{as}^{k} f^{si}_{b} ~-~ c_{bs}^{k} f^{si}_{a} ~+~ c_{bs}^{i} f^{sk}_{a} ~~= ~~0. $$ This condition is symmetric with respect to exchange of $c$ and $f$. Thus if $({\frak g},{\frak g}^{*})$ is a Lie bialgebra, then $({\frak g}^{*}, {\frak g})$ is also a Lie bialgebra. The following proposition shows that the category of finite--dimensional Lie bialgebras is isomorphic to the category of finite--dimensional connected simply connected Poisson--Lie groups. \begin{proposition}{\bf (\cite{ChP}, Theorem 1.3.2)} If $G$ is a connected simply connected finite--dimensional Lie group, every bialgebra structure on $\frak g$ is the tangent bialgebra of a unique Poisson structure on $G$ which makes $G$ into a Poisson--Lie group. \end{proposition} Let $G$ be a finite--dimensional Poisson--Lie group, $({\frak g},{\frak g}^{*})$ the tangent bialgebra of $G$. The connected simply connected finite--dimensional Poisson--Lie group corresponding to the Lie bialgebra $({\frak g}^{*}, {\frak g})$ is called the dual Poisson--Lie group and denoted by $G^*$. $({\frak g},{\frak g}^{*})$ is called a {\em factorizable Lie bialgebra }if the following conditions are satisfied (see \cite{fact} , \cite{Dm}): \begin{enumerate} \item ${\frak g}${\em \ is equipped with a non--degenerate invariant scalar product} $\left( \cdot ,\cdot \right)$. We shall always identify ${\frak g}^{*}$ and ${\frak g}$ by means of this scalar product. \item {\em The dual Lie bracket on }${\frak g}^{*}\simeq {\frak g}${\em \ is given by} \begin{equation} \left[ X,Y\right] _{*}=\frac 12\left( \left[ rX,Y\right] +\left[ X,rY\right] \right) ,X,Y\in {\frak g}, \label{rbr} \end{equation} {\em where }$r\in {\rm End}\ {\frak g}${\em \ is a skew symmetric linear operator (classical r-matrix).} \item $r${\em \ satisfies} {\em the} {\em modified classical Yang-Baxter identity:} \begin{equation} \left[ rX,rY\right] -r\left( \left[ rX,Y\right] +\left[ X,rY\right] \right) =-\left[ X,Y\right] ,\;X,Y\in {\frak g}{\bf .} \label{cybe} \end{equation} \end{enumerate} Define operators $r_\pm \in {\rm End}\ {\frak g}$ by \[ r_{\pm }=\frac 12\left( r\pm id\right) . \] We shall need some properties of the operators $r_{\pm }$. Denote by ${\frak b}_\pm$ and ${\frak n}_\mp$ the image and the kernel of the operator $r_\pm $: \begin{equation}\label{bnpm} {\frak b}_\pm = Im~r_\pm,~~{\frak n}_\mp = Ker~r_\pm. \end{equation} \begin{proposition}{\bf (\cite{BD}, \cite{rmatr})}\label{bpm} Let $({\frak g}, {\frak g}^*)$ be a factorizable Lie bialgebra. Then (i) ${\frak b}_\pm \subset {\frak g}$ is a Lie subalgebra, the subspace ${\frak n}_\pm$ is a Lie ideal in ${\frak b}_\pm,~{\frak b}_\pm^\perp ={\frak n}_\pm$. (ii) ${\frak n}_\pm$ is an ideal in ${\frak {g}}^{*}$. (iii) ${\frak b}_\pm$ is a Lie subalgebra in ${\frak {g}}^{*}$. Moreover ${\frak b}_\pm ={\frak {g}}^{*}/ {\frak n}_\pm$. (iv) $({\frak b}_\pm,{\frak b}_\pm ^*)$ is a subbialgebra of $({\frak {g}},{\frak {g}}^{*})$ and $({\frak b}_\pm,{\frak b}_\pm ^*)\simeq ({\frak b}_\pm,{\frak b}_\mp)$. The canonical paring between ${\frak b}_\mp$ and ${\frak b}_\pm$is given by \begin{equation} (X_\mp ,Y_\pm )_\pm=(X_\mp,r_\pm^{-1}Y_\pm ) ,~ X_\mp \in {\frak b}_\mp ;~ Y_\pm \in {\frak b}_\pm . \end{equation} \end{proposition} The classical Yang--Baxter equation implies that $r_{\pm }$ , regarded as a mapping from ${\frak g}^{*}$ into ${\frak g}$, is a Lie algebra homomorphism. Moreover, $r_{+}^{*}=-r_{-},$\ and $r_{+}-r_{-}=id.$ Put ${\frak {d}}={\frak g\oplus {g}}$ (direct sum of two copies). The mapping \begin{eqnarray}\label{imbd} {\frak {g}}^{*}\rightarrow {\frak {d}}~~~:X\mapsto (X_{+},~X_{-}),~~~X_{\pm }~=~r_{\pm }X \end{eqnarray} is a Lie algebra embedding. Thus we may identify ${\frak g^{*}}$ with a Lie subalgebra in ${\frak {d}}$. Naturally, embedding (\ref{imbd}) extends to an embedding $$ G^*\rightarrow G\times G,~~L\mapsto (L_+,L_-). $$ We shall identify $G^*$ with the corresponding subgroup in $G\times G$. \section{Poisson reduction}\label{poisred} \setcounter{equation}{0} \setcounter{theorem}{0} In this section we recall basic facts on Poisson reduction (see \cite{W}, \cite{RIMS}). These facts will be used in the proof of the group counterpart of Theorem E (see Section \ref{geomappr}). Let $M,~B,~B'$ be Poisson manifolds. Two Poisson surjections $$ \begin{array}{ccccc} & & M & & \\ & \stackrel{\pi^{\prime } }{\swarrow } & & \stackrel{\pi }{\searrow } & \\ B^{\prime } & & & & B \end{array} $$ form a dual pair if the pullback $\pi^{^{\prime }*}C^\infty(B^{\prime })$ is the centralizer of $\pi^* C^\infty (B)$ in the Poisson algebra $C^\infty (M)$. In that case the sets $B^{\prime }_b=\pi^{\prime } \left( \pi ^{-1}(b) \right),~b\in B$ are Poisson submanifolds in $B^{\prime }$ (see \cite{W}) called reduced Poisson manifolds. Fix an element $b\in B$. Then the algebra of functions $C^\infty (B^{\prime }_b)$ may be described as follows. Let $I_b$ be the ideal in $C^\infty (M)$ generated by elements ${\pi}^*(f),~f\in C^\infty (B),~f(b)=0$. Denote $M_b=\pi^{-1}(b)$. Then the algebra $C^\infty (M_b)$ is simply the quotient of $C^\infty (M)$ by $I_b$. Denote by $P_b:C^\infty (M)\rightarrow C^\infty (M)/I_b=C^\infty (M_b)$ the canonical projection onto the quotient. \begin{lemma}\label{redspace} Suppose that the map $f\mapsto f(b)$ is a character of the Poisson algebra $C^\infty (B)$. Then one can define an action of the Poisson algebra $C^\infty (B)$ on the space $C^\infty (M_b)$ by \begin{equation}\label{redact} f\cdot \varphi=P_b(\{ {\pi}^*(f), \tilde \varphi \}), \end{equation} where $f\in C^\infty (B)$, $\varphi \in C^\infty (M_b)$, $\tilde \varphi \in C^\infty (M)$ is a representative of $\varphi$ in $C^\infty (M)$ such that $P_b(\tilde \varphi)=\varphi$. Moreover, $C^\infty (B^{\prime }_b)$ is the subspace of invariants in $C^\infty (M_b)$ with respect to this action. \end{lemma} {\em Proof.} Let $\varphi \in C^\infty (M_b)$. Choose a representative $\tilde \varphi \in C^\infty (M)$ such that $P_b(\tilde \varphi)=\varphi$. Since the map $f\mapsto f(b)$ is a character of the Poisson algebra $C^\infty (B)$, Hamiltonian vector fields of functions ${\pi}^*(f),~f\in C^\infty (B)$ are tangent to the surface $M_b$. Therefore using the the definition of the dual pair we obtain that $\varphi={\pi^{\prime }}^*(\psi)$ for some $\psi \in C^\infty(B^{\prime }_b)$ if and only if $P_b(\{ {\pi}^*(f), \tilde \varphi\})=0$ for every $f\in C^\infty (B)$. Note also that the r.h.s. of (\ref{redact}) only depends on $\varphi$ but not on the representative $\tilde \varphi$, and hence formula (\ref{redact}) defines an action of the Poisson algebra $C^\infty (B)$ on the space $C^\infty (M_b)$. Finally we obtain that $C^\infty (B^{\prime }_b)$ is exactly the subspace of invariants in $C^\infty (M_b)$ with respect to this action. \begin{definition} The algebra $C^\infty (B^{\prime }_b)$ is called a reduced Poisson algebra. We also denote it by $C^\infty (M_b)^{C^\infty (B)}$. \end{definition} \begin{remark}\label{redpoisalg} Note that the description of the algebra $C^\infty (M_b)^{C^\infty (B)}$ obtained in Lemma \ref{redspace} is independent of both the manifold $B^{\prime }$ and the projection $\pi^{\prime }$. Observe also that the reduced space $B^{\prime }_b$ may be identified with a cross--section of the action of the Poisson algebra $C^\infty (B)$ on $M_b$ by Hamiltonian vector fields. In particular, $B^{\prime }_b$ may be regarded as a submanifold in $M_b$. \end{remark} An important example of dual pairs is provided by Poisson group actions. Recall that a Poisson group action of a Poisson--Lie group $A$ on a Poisson manifold $M$ is a group action $A\times M\rightarrow M$ which is also a Poisson map (as usual, we suppose that $A\times M$ is equipped with the product Poisson structure). In \cite{RIMS} it is proved that if the space $M/A$ is a smooth manifold, there exists a unique Poisson structure on $M/A$ such that the canonical projection $M\rightarrow M/A$ is a Poisson map. Let $\frak a$ be the Lie algebra of $A$. Denote by $\langle\cdot,\cdot\rangle$ the canonical paring between ${\frak a}^*$ and $\frak a$. A map $\mu :M\rightarrow A^*$ is called a moment map for a right Poisson group action $A\times M\rightarrow M$ if (\cite{Lu}) \begin{equation} L_{\widehat X} \varphi =\langle \mu^*(\theta_{A^*}) , X \rangle (\xi_\varphi ) , \end{equation} where $\theta_{A^*}$ is the universal right--invariant Maurer--Cartan form on $A^*$, $X \in {\frak a}$, $\widehat X$ is the corresponding vector field on $M$ and $\xi_\varphi $ is the Hamiltonian vector field of $\varphi \in C^\infty (M)$. By Theorem 4.9, \cite{Lu} one can always equip $A^*$ with a Poisson structure in such a way that $\mu$ becomes a Poisson mapping. Then from the definition of the moment map it follows that if $M/A$ is a smooth manifold then the canonical projection $M\rightarrow M/A$ and the moment map $\mu:M\rightarrow A^*$ form a dual pair (see \cite{Lu} for details). The main example of Poisson group actions is the so--called dressing action. The dressing action may be described as follows (see \cite{Lu}, \cite{RIMS}). \begin{proposition}\label{dressingact} Let $G$ be a connected simply connected Poisson--Lie group with factorizable tangent Lie bialgebra, $G^*$ the dual group. Then there exists a unique right Poisson group action $$ G^*\times G\rightarrow G^*,~~((L_+,L_-),g)\mapsto g\circ (L_+,L_-), $$ such that the identity mapping $\mu: G^* \rightarrow G^*$ is the moment map for this action. Moreover, let $q:G^* \rightarrow G$ be the map defined by $$ q(L_+,L_-)=L_-L_+^{-1}. $$ Then $$ q(g\circ (L_+,L_-))=g^{-1}L_-L_+^{-1}g. $$ \end{proposition} The notion of Poisson group actions may be generalized as follows. Let $A\times M \rightarrow M$ be a Poisson group action of a Poisson--Lie group $A$ on a Poisson manifold $M$. A subgroup $K\subset A$ is called {\em admissible} if the set $% C^\infty \left( M\right) ^K$ of $K$-invariants is a Poisson subalgebra in $% C^\infty \left( M\right)$. If space $M/K$ is a smooth manifold, we may identify the algebras $C^\infty(M/K)$ and $C^\infty \left( M\right) ^K$. Hence there exists a Poisson structure on $M/K$ such that the canonical projection $M\rightarrow M/K$ is a Poisson map. \begin{proposition}\label{admiss}{\bf (\cite{RIMS}, Theorem 6; \cite{Lu}, \S 2)} Let $\left( {\frak a},{\frak a}^{*}\right) $ be the tangent Lie bialgebra of $A.$ A connected Lie subgroup $K\subset A$ with Lie algebra ${\frak k}\subset {\frak a}$ is admissible if ${\frak k}^{\perp }\subset {\frak a}^{*}$ is a Lie subalgebra. \end{proposition} We shall need the following particular example of dual pairs arising from Poisson group actions. Let $A\times M \rightarrow M$ be a right Poisson group action of a Poisson--Lie group $A$ on a manifold $M$. Suppose that this action possesses a moment mapping $\mu : M\rightarrow A^*$. Let $K$ be an admissible subgroup in $A$. Denote by $\frak k$ the Lie algebra of $K$. Assume that ${\frak k}^\perp \subset {\frak a}^*$ is a Lie subalgebra in ${\frak a}^*$. Suppose also that there is a splitting ${\frak a}^*={\frak t}\oplus {\frak k}^\perp$, and that $\frak t$ is a Lie subalgebra in ${\frak a}^*$. Then the linear space ${\frak k}^*$ is naturally identified with $\frak t$. Assume that $A^*$ is the semidirect product of the Lie subgroups $K^\perp , T$ corresponding to the Lie algebras ${\frak k}^\perp , {\frak t}$ respectively. Suppose that $K^\perp$ is a connected subgroup in $A^*$. Fix the decomposition $A^*=K^\perp T$ and denote by $\pi_{K^\perp} , \pi_{T}$ the projections onto $K^\perp$ and $T$ in this decomposition. \begin{proposition}\label{QPmoment} Define a map $\overline{\mu}:M\rightarrow T$ by $$ \overline{\mu}=\pi_{T}\mu. $$ Then (i) $\overline{\mu}^*\left( C^\infty \left( T\right)\right)$ is a Poisson subalgebra in $C^\infty \left( M\right)$, and hence one can equip $T$ with a Poisson structure such that $\overline{\mu}:M\rightarrow T$ is a Poisson map. (ii)Moreover, the algebra $C^\infty \left( M\right) ^K$ is the centralizer of $\overline{\mu}^*\left( C^\infty \left( T\right)\right)$ in the Poisson algebra $C^\infty \left( M\right)$. In particular, if $M/K$ is a smooth manifold the maps \begin{equation}\label{dp} \begin{array}{ccccc} & & M & & \\ & \stackrel{\pi }{\swarrow } & & \stackrel{\overline{\mu}}{\searrow } & ,\\ M/K & & & & T \end{array} \end{equation} form a dual pair. \end{proposition} {\em Proof.} (i)First, by Theorem 4.9 in \cite{Lu} there exists a Poisson bracket on $A^*$ such that $\mu :M\rightarrow A^*$ is a Poisson map. Moreover, we can choose this bracket to be the sum of the standard Poisson--Lie bracket of $A^*$ and of a left invariant bivector on $A^*$. Denote by $A^*_M$ the manifold $A^*$ equipped with this Poisson structure. Now observe that $T$ is identified with the quotient $K^\perp \setminus A^*_M$, where $K^\perp$ acts on $A^*_M$ by multiplications from the left. Therefore to prove part (i) of the proposition it suffices to show that $K^\perp$--invariant functions on $A^*_M$ form a Poisson subalgebra in $C^\infty(A^*_M)$. Observe that since $A^*$ is a Poisson--Lie group and the Poisson structure of $A^*_M$ is obtained from that of $A^*$ by adding a left--invariant term, the action of $A^*$ on $A^*_M$ by multiplications from the left is a Poisson group action. Note also that $K^\perp$ is a connected subgroup in $A^*$ and $({\frak k}^\perp)^\perp \cong {\frak k}$ is a Lie subalgebra in $\frak a$. Therefore by Proposition \ref{admiss} $K^\perp$ is an admissible subgroup in $A^*$. Therefore $K^\perp$--invariant functions on $A^*_M$ form a Poisson subalgebra in $C^\infty(A^*_M)$, and hence $\overline{\mu}^*\left( C^\infty \left( T\right)\right)$ is a Poisson subalgebra in $C^\infty \left( M\right)$. This proves part (i). (ii)By the definition of the moment map we have: \begin{equation}\label{X5} L_{\widehat X} \varphi =\langle \mu^*(\theta_{A^*}) , X \rangle (\xi_\varphi ) , \end{equation} where $X \in {\frak a} , \widehat X$ is the corresponding vector field on $M$ and $\xi_\varphi $ is the Hamiltonian vector field of $\varphi \in C^\infty (M)$. Since $A^*$ is the semidirect product of $K^\perp$ and $T$ the pullback of the right--invariant Maurer--Cartan form $\mu^*(\theta_{A^*})$ may be represented as follows: $$ \mu^*(\theta_{A^*})= {\rm Ad}(\pi_{K^\perp}\mu )({\overline{\mu}}^*\theta_{T})+(\pi_{K^\perp}\mu )^*\theta_{K^\perp}, $$ where ${\rm Ad}(\pi_{K^\perp}\mu )({\overline{\mu}}^*\theta_{T})\in {\frak t},~(\pi_{K^\perp}\mu )^*\theta_{K^\perp}\in {\frak k}^\perp$. Now let $X \in {\frak k}$. Then $\langle (\pi_{K^\perp}\mu )^*\theta_{K^\perp}),X\rangle =0$ and formula (\ref{X5}) takes the form: \begin{equation}\label{+} \begin{array}{l} L_{\widehat X} \varphi = \langle {\rm Ad}(\pi_{K^\perp}\mu )({\overline{\mu}}^*\theta_{T}),X \rangle (\xi_\varphi )=\\ \\ \langle {\rm Ad}(\pi_{K^\perp}\mu )(\theta_{T}),X \rangle ({\overline{\mu}}_*(\xi_\varphi )) . \end{array} \end{equation} Since ${\rm Ad}(\pi_{K^\perp}\mu )$ is a non--degenerate transformation, $L_{\widehat X} \varphi =0$ for every $X\in {\frak k}$ if and only if ${\overline{\mu}}_*(\xi_\varphi )=0$, i.e. a function $\varphi \in C^\infty (M)$ is $K$--invariant if and only if $\{ \varphi ,\overline{\mu}^*(\psi) \}=0$ for every $\psi \in C^\infty (T)$. This completes the proof. \begin{remark}\label{remred} Let $t\in T$ be as in Lemma \ref{redspace}. Assume that $\pi(\overline{\mu}^{-1}(t))$ is a smooth manifold ($M/K$ does not need to be smooth). Then the algebra $C^\infty(\pi(\overline{\mu}^{-1}(t)))$ is isomorphic to the reduced Poisson algebra $C^\infty(\overline{\mu}^{-1}(t))^{C^\infty(T)}$. \end{remark} \section{Quantization of Poisson--Lie groups and Whittaker model} \setcounter{equation}{0} \setcounter{theorem}{0} Let $\frak g$ be a finite--dimensional complex simple Lie algebra. Observe that cocycle (\ref{cocycles}) equips $\frak g$ with the structure of a factorizable Lie bialgebra. For simplicity we suppose that $\pi =id$, and denote $s_{id}=s$. Using the identification ${\rm End}~{\frak g}\cong {\frak g}\otimes {\frak g}$ the corresponding r--matrix may be represented as $$ r^{s}=P_+-P_-+{1+s \over 1-s}P_0, $$ where $P_+,P_-$ and $P_0$ are the projection operators onto ${\frak n}_+,{\frak n}_-$ and $\frak h$ in the direct sum $$ {\frak g}={\frak n}_+ +{\frak h} + {\frak n}_-. $$ Let $G$ be the connected simply connected simple Poisson--Lie group with the tangent Lie bialgebra $({\frak g},{\frak g}^*)$, $G^*$ the dual group. Observe that $G$ is an algebraic group (see \S 104, Theorem 12 in \cite{Z}). Note also that $$ r^{s}_+=P_+ + {1 \over 1-s}P_0,~~r^{s}_-=-P_- + {s \over 1-s}P_0, $$ and hence the subspaces ${\frak b}_\pm$ and ${\frak n}_\pm$ defined by (\ref{bnpm}) coincide with the Borel subalgebras in $\frak g$ and their nil--radicals, respectively. Therefore every element $(L_+,L_-)\in G^*$ may be uniquely written as \begin{equation}\label{fact} (L_+,L_-)=(h_+,h_-)(n_+,n_-), \end{equation} where $n_\pm \in N_\pm$, $h_+=exp({1 \over 1-s}x),~h_-=exp({s \over 1-s}x),~x\in {\frak h}$. In particular, $G^*$ is a solvable algebraic subgroup in $G\times G$. For every algebraic variety $V$ we denote by ${\cal F}(V)$ the algebra of regular functions on $V$. Our main object will be the algebra of regular functions on $G^*$, ${\cal F}(G^*)$. This algebra may be explicitly described as follows. Let $\pi_V$ be a finite--dimensional representation of $G$. Then matrix elements of $\pi_V(L_\pm)$ are well--defined functions on $G^*$, and ${\cal F}(G^*)$ is the subspace in $C^\infty(G^*)$ generated by matrix elements of $\pi_V(L_\pm)$, where $V$ runs through all finite--dimensional representations of $G$. The elements $L^{\pm,V}=\pi_V(L_\pm)$ may be viewed as elements of the space ${\cal F}(G^*)\otimes {\rm End}V$. For every two finite--dimensional ${\frak g}$ modules $V$ and $W$ we denote ${r^s_+}^{VW}=(\pi_V\otimes \pi_W)r^s_+$, where $r^s_+$ is regarded as an element of ${\frak g}\otimes {\frak g}$. \begin{proposition}{\bf (\cite{dual}, Section 2)} ${\cal F}(G^*)$ is a Poisson subalgebra in the Poisson algebra $C^\infty(G^*)$, the Poisson brackets of the elements $L^{\pm,V}$ are given by \begin{equation}\label{pbf} \begin{array}{l} \{L^{\pm,W}_{1},L^{\pm,V}_{2}\}~=~ 2[{r_+^s}^{VW},L^{\pm,W}_{1}L^{\pm,V}_{2}],\\ \\ \{L^{-,W}_{1},L^{+,V}_{2}\}~=2[{r_{+}^s}^{VW},L^{-,W}_{1}L^{+,V}_{2}], \end{array} \end{equation} where $$ L^{\pm,W}_1=L^{\pm,W}\otimes I_V,~~L^{\pm,V}_2=I_W\otimes L^{\pm,V}, $$ and $I_X$ is the unit matrix in $X$. Moreover, the map $\Delta:{\cal F}(G^*)\rightarrow {\cal F}(G^*)\otimes {\cal F}(G^*)$ dual to the multiplication in $G^*$, \begin{equation}\label{comultcl} \Delta(L^{\pm,V}_{ij})=\sum_k L^{\pm,V}_{ik}\otimes L^{\pm,V}_{kj}, \end{equation} is a homomorphism of Poisson algebras. \end{proposition} \begin{remark} Recall that a Poisson--Hopf algebra is a Poisson algebra which is also a Hopf algebra such that the comultiplication is a homomorphism of Poisson algebras. According to Proposition \ref{pbf} ${\cal F}(G^*)$ is a Poisson--Hopf algebra. \end{remark} Now we describe a quantization of the Poisson--Hopf algebra ${\cal F}(G^*)$. Let $\tilde U_h^{s}({\frak g})$ be the subalgebra in $U_h^{s}({\frak g})$ topologically generated, in the sense of formal power series over ${\Bbb C}[[h]]$, by elements $\tilde H_i=hH_i,~i=1,\ldots l,~\tilde e_{\beta}=he_{\beta},~\tilde f_{\beta}=hf_{\beta},~\beta\in \Delta_+$. In fact $\tilde U_h^{s_\pi}({\frak g})$ is a Hopf subalgebra in $U_h^{s}({\frak g})$, explicit formulas for the comultiplication may be obtained using Proposition 8.3 in \cite{kh-t}. \begin{proposition} $\tilde U_h^{s}({\frak g})$ is a quantum formal series Hopf algebra (or QFSH algebra), i.e. $\tilde U_h^{s}({\frak g})$ is isomorphic as a ${\Bbb C}[[h]]$--module to $Map(I,{\Bbb C}[[h]])$ for some set $I$, and $\tilde U^{s}({\frak g})=\tilde U_h^{s}({\frak g})/h\tilde U_h^{s}({\frak g})\cong {\Bbb C}[[\xi_1,\xi_2,\ldots ]]$ as a topological algebra, for some (possibly infinite) sequence of indeterminates $\xi_1,\xi_2,\ldots $. \end{proposition} {\em Proof} is similar to the proof of the same result for $U_h({\frak g})$ (see Section 8.3 C in \cite{ChP}). Note that $\tilde U^{s}({\frak g})$ is naturally a Poisson--Hopf algebra, the Poisson bracket is given by \begin{equation}\label{quasipb} \{x_1,x_2\}={[a_1,a_2] \over h}~(\mbox{mod }h), \end{equation} if $a_1,a_2\in \tilde U_h^{s}({\frak g})$ reduce to $x_1,x_2\in \tilde U^{s}({\frak g})~(\mbox{mod }h)$. For any finite--dimensional $U_h^{s}({\frak g})$ module $V[[h]]$ we denote by ${^h{L^{\pm,V}}}$ the following elements of $U_h^{s}({\frak g})\otimes {\rm End}V[[h]]$ (see \cite{FRT}): $$ {^h{L^{+,V}}}=(id\otimes \pi_V){{\cal R}_{21}^{s}}^{-1}=(id\otimes \pi_VS^{s}){\cal R}_{21}^{s} ,~~ {^h{L^{-,V}}}=(id\otimes \pi_V){\cal R}^{s}. $$ We also denote $R^{VW}=(\pi_V\otimes \pi_W){\cal R}^{s}$. Observe that from formula (\ref{rmatrspi}) it follows that actually ${^h{L^{\pm,V}}}\in \tilde U_h^{s}({\frak g})\otimes {\rm End}V[[h]]$. If we fix a basis in $V[[h]]$, ${^h{L^{\pm,V}}}$ may be regarded as matrices with matrix elements $({^h{L^{\pm,V}}})_{ij}$ being elements of $\tilde U_h^{s}({\frak g})$. From the Yang--Baxter equation for $\cal R$ we get relations between $L^{\pm,V}$: \begin{equation}\label{ppcomm} \begin{array}{l} R^{VW}{^h{L^{\pm,W}_1}}{^h{L^{\pm,V}_2}}={^h{L^{\pm,V}_2}}{^h{L^{\pm,W}_1}}R^{VW }, \end{array} \end{equation} \begin{equation}\label{pmcomm} R^{VW}{^h{L^{-,W}_1}}{^h{L^{+,V}_2}}={^h{L^{+,V}_2}}{^h{L^{-,W}_1}}R^{VW}. \end{equation} By ${^h{L^{\pm,W}_1}},~{^h{L^{\pm,V}_2}}$ we understand the following matrices in $V\otimes W$: $$ {^h{L^{\pm,W}_1}}={^h{L^{\pm,W}}}\otimes I_V,~~{^h{L^{\pm,V}_2}}=I_W\otimes {^h{L^{\pm,V}}}, $$ where $I_X$ is the unit matrix in $X$. From (\ref{rmprop}) we can obtain the action of the comultiplication on the matrices ${^h{L^{\pm,V}}}$: \begin{equation}\label{comult} \Delta_s({^h{L^{\pm,V}_{ij}}})=\sum_k {^h{L^{\pm,V}_{ik}}}\otimes {^h{L^{\pm,V}_{kj}}}. \end{equation} We denote by ${\cal F}_h(G^*)$ the Hopf subalgebra in $\tilde U_h^{s}({\frak g})$ generated in the sense of $h$--adic topology by matrix elements of ${^h{L^{\pm,V}}}$, where $V$ runs through all finite--dimensional representations of $\frak g$. \begin{proposition}\label{quantreg} Denote by $p:\tilde U_h^{s}({\frak g})\rightarrow \tilde U^{s}({\frak g})$ the canonical projection. Then $p({\cal F}_h(G^*))$ is isomorphic to ${\cal F}(G^*)$ as a Poisson--Hopf algebra. \end{proposition} {\em Proof.} Denote ${\cal F}(G^*)'=p({\cal F}_h(G^*)),~{\tilde L^{\pm,V}}=p({^h{L^{\pm,V}}})\in {\cal F}(G^*)'\otimes {\rm End}V$. First observe that the map $$ \imath :{\cal F}(G^*)'\rightarrow {\cal F}(G^*),~~(\imath \otimes id){\tilde L^{\pm,V}}={L^{\pm,V}} $$ is a well--defined linear isomorphism. Indeed, consider, for instance, element ${\tilde L^{-,V}}$. From (\ref{rmatrspi}) it follows that \begin{equation} \begin{array}{l} {\tilde L^{-,V}}_{ij}=\{ exp\left[ \sum_{i=1}^l-2p(hY_i)\otimes \pi_V({s \over 1-s}H_i)\right]\times \\ \prod_{\beta} exp[2(X_\beta,X_{-\beta})^{-1}p(he_{\beta}) \otimes \pi_V(X_{-\beta})]\}_{ij}. \end{array} \end{equation} On the other hand (\ref{fact}) implies that every element $L_-$ may be represented in the form \begin{equation} \begin{array}{l} L_- = exp\left[ \sum_{i=1}^lb_i{s \over 1-s}H_i\right]\times \\ \prod_{\beta} exp[b_{\beta}X_{-\beta}],~b_i,b_\beta\in {\Bbb C}, \end{array} \end{equation} and hence \begin{equation} \begin{array}{l} L^{-,V}_{ij}=\{ exp\left[ \sum_{i=1}^lb_i\otimes \pi_V({s \over 1-s}H_i)\right]\times \\ \prod_{\beta} exp[b_{\beta} \otimes \pi_V(X_{-\beta})]\}_{ij}. \end{array} \end{equation} Therefore $\imath$ is a linear isomorphism. We have to prove that $\imath$ is an isomorphism of Poisson--Hopf algebras. Recall that ${\cal R}^{s}=1\otimes 1 +2hr_+^{s}$ (mod $h^2$). Therefore from commutation relations (\ref{ppcomm}), (\ref{pmcomm}) it follows that ${\cal F}(G^*)'$ is a commutative algebra, and the Poisson brackets of matrix elements ${\tilde L^{\pm,V}}_{ij}$ (see (\ref{quasipb})) are given by (\ref{pbf}), where $L^{\pm,V}$ are replaced by ${\tilde L^{\pm,V}}$. From (\ref{comult}) we also obtain that the action of the comultiplication on the matrices ${\tilde L^{\pm,V}}$ is given by (\ref{comultcl}), where $L^{\pm,V}$ are replaced by ${\tilde L^{\pm,V}}$. This completes the proof. We shall call the map $p:{\cal F}_h(G^*) \rightarrow {\cal F}(G^*)$ the quasiclassical limit. Now using the Hopf algebra ${\cal F}_h(G^*)$ we shall define another quantum version of the Whittaker model $W({\frak b}_-)$. Let ${\cal F}_h(N_\pm)$ be the subalgebras in ${\cal F}_h(G^*)$ generated by matrix elements of the matrices $N^{-,V}=(id\otimes \pi_V){\cal R}^{s}_\Delta,~N^{+,V}=(id\otimes \pi_V){{\cal R}_{\Delta}^{s}}^{-1}_{21}$, where $$ {\cal R}^{s}_\Delta= \prod_{\beta}exp_{q_{\beta}^{-1}}[(q-q^{-1})a(\beta)^{-1} e_{\beta} \otimes e^{h{1+s \over 1-s} \beta^\vee}f_{\beta}]. $$ Suppose that the ordering of the root system $\Delta_+$ is fixed as in Proposition \ref{rootsh}. Then by Proposition \ref{rootsh} the map $\chi_h^{s}:{\cal F}_h(N_-)\rightarrow {\Bbb C}$ defined by \begin{equation}\label{charq} (\chi_h^{s}\otimes id)(N^{-,V})= \prod_{i=1}^lexp_{q_{\alpha_i}^{-1}}[{(q_i-q_i^{-1})\over h}c_{i} \otimes \pi_V(e^{hd_i{1+s \over 1-s}H_i}f_i)], c_i\in {\Bbb C}[[h]],~c_i\neq 0 \end{equation} is a character of ${\cal F}_h(N_-)$. We also denote by ${\cal F}_h(H)$ the intersection $U_h^{s}({\frak h})\cap {\cal F}_h(G^*)$. Clearly, ${\cal F}_h(H)$ is a commutative subalgebra in ${\cal F}_h(G^*)$. From commutation relations (\ref{pmcomm}) one can obtain the following weak version of the Poincar\'{e}--Birkhoff--Witt theorem for ${\cal F}_h(G^*)$. \begin{proposition}\label{Pbw} Multiplication defines an isomorphism of ${\Bbb C}[[h]]$--modules $$ {\cal F}_h(N_+)\otimes {\cal F}_h(H)\otimes {\cal F}_h(N_-)\rightarrow {\cal F}_h(G^*). $$ \end{proposition} Define ${\cal F}_h(B_\pm)={\cal F}_h(N_\pm){\cal F}_h(H)$. Let ${{\cal F}_h(N_-)}_{\chi_h^{s}}$ be the kernel of the character $\chi_h^{s}$ so that one has a direct sum \begin{equation}\label{ker} {\cal F}_h(N_-)={\Bbb C}[[h]]\oplus {{\cal F}_h(N_-)}_{\chi_h^{s}}. \end{equation} From Proposition \ref{Pbw} and formula (\ref{ker}) we obtain also the direct sum \begin{equation}\label{maindecqg} {\cal F}_h(G^*)={\cal F}_h(B_+)\oplus I_{\chi_h^{s}}, \end{equation} where $I_{\chi_h^{s}}={\cal F}_h(G^*){{\cal F}_h(N_-)}_{\chi_h^{s}}$ is the left--sided ideal generated by ${{\cal F}_h(N_-)}_{\chi_h^{s}}$. Denote by $\rho_{\chi_h^{s}}$ the projection onto ${\cal F}_h(B_+)$ in the direct sum (\ref{maindecqg}). Let $Z({\cal F}_h(G^*))$ be the center of ${\cal F}_h(G^*)$. Similarly to the classical case we define a subspace $W_h(B_+)$ in ${\cal F}_h(B_+)$ by $W_h(B_+)=\rho_{\chi_h^{s}}(Z({\cal F}_h(G^*)))$. To formulate the quantum version of Theorem A for $W_h(B_+)$ we recall that for any finite--dimensional $\frak g$--module $V$ the element $$ C_V=(id\otimes tr_V)((S^{s}\otimes id)(L^{+,V})L^{-,V}(1\otimes e^{2h\rho^\vee})), $$ where $tr_V$ is the trace in $V[[h]]$, is central in ${\cal F}_h(G^*)$ (see formulas (\ref{centrelv}) and (\ref{S})). \vskip 0.3cm \noindent {\bf Theorem $\bf A_q$} {\em (i)The map \begin{equation}\label{qgrho} \rho_{\chi_h^{s_\pi}}:Z({\cal F}_h(G^*))\rightarrow W_h(B_+) \end{equation} is an isomorphism of algebras. (ii) The algebra $W_h(B_+)$ is freely generated as a commutative topological algebra over ${\Bbb C}[[h]]$ by the elements $C_{V_i}^{\rho_{\chi_h^{s}}}=\rho_{\chi_h^{s}}(C_{V_i}),~i=1,\ldots ,l$, where $V_i,~i=1,\ldots l$ are the fundamental representations of $\frak g$.} \vskip 0.3cm \noindent {\em Proof} of (i) is similar to that of Theorem A in the classical case. Part (ii) will be proved in Section \ref{cross}. \begin{corollary} The algebra $Z({\cal F}_h(G^*))$ is freely generated as a commutative topological algebra over ${\Bbb C}[[h]]$ by the elements $C_{V_i}$, where $V_i,~i=1,\ldots l$ are the fundamental representations of $\frak g$. \end{corollary} \vskip 0.3cm \noindent {\bf Definition $\bf A_q$} {\em The algebra $$ W_h(B_+)=\rho_{\chi_h^{s}}(Z({\cal F}_h(G^*))). $$ is called the Whittaker model of the center $Z({\cal F}_h(G^*))$.} \vskip 0.3cm Now following Section \ref{whitt} (see Lemma A) we equip ${\cal F}_h(B_+)$ with a structure of a left ${\cal F}_h(N_-)$ module in such a way that $W_h(B_+)$ is realized as the space of invariants with respect to this action. For every $v\in {\cal F}_h(B_+)$ and $x\in {\cal F}_h(N_-)$ we put \begin{equation}\label{mainactqg} x\cdot v =\rho_{\chi_h^{s}}([x,v]). \end{equation} Consider the space ${\cal F}_h(B_+)^{{\cal F}_h(N_-)}$ of ${\cal F}_h(N_-)$ invariants in ${\cal F}_h(B_+)$ with respect to this action. Clearly, $W_h(B_+)\subseteq {\cal F}_h(B_+)^{{\cal F}_h(N_-)}$. To formulate the quantum version of Theorem B for $W_h(B_+)$ we have to impose a restriction on the coefficients $c_i$ in (\ref{charq}). Define an element $u\in N_-$ by \begin{equation}\label{u} u=\prod_{i=1}^lexp[2d_ic_{i}^0 X_{-\alpha_i}],~c_i^0=c_i~(\mbox{mod }h), \end{equation} where the terms in the product are ordered as in (\ref{charq}). The motivation for this definition will be explained in the next section. \vskip 0.3cm \noindent {\bf Theorem $\bf B_q$ } {\em Suppose that $u\in N_+sN_+\cap N_-$, where $s$ stands for a representative of the Coxeter element in $G$. Then the space of ${\cal F}_h(N_-)$ invariants in ${\cal F}_h(B_+)$ with respect to the action (\ref{mainactqg}) is isomorphic to $W_h(B_+)$, i.e.} \begin{equation}\label{invqg} {\cal F}_h(B_+)^{{\cal F}_h(N_-)}\cong W_h(B_+). \end{equation} \vskip 0.3cm The proof of this theorem occupies two next sections. \begin{remark} The following lemma shows that the set $N_+sN_+\cap N_-$ is not empty. \begin{lemma}{\bf (\cite{st}, Lemma 4.5)}\label{f} Let $w_0\in W$ be the longest element; let $\tau \in Aut$ $\Delta _{+}$ be the automorphism defined by $\tau \left( \alpha \right) =-w_0\alpha ,\alpha \in \Delta _{+}.$ Let $N_i\subset N_+$ be the 1-parameter subgroup generated by the root vector $X_{\tau \left( \alpha _i\right) },i=1,\ldots l$. Choose an element $u_i\in N_i,u_i\neq 1.$ Then we have $w_0u_iw_0^{-1}\in B_+s_iB_+.$ We may fix $u_i$ in such a way that $w_0u_iw_0^{-1}\in N_+s_iN_+.$ Set $x=u_1u_{2}...u_l$. Then $f=w_0xw_0^{-1}\in N_+sN_+\cap N_-.$ \end{lemma} \end{remark} Similarly to Proposition \ref{hkhom} we also have the following homological description of $W_h(B_+)$. \begin{proposition} Suppose that the conditions of Theorem $B_q$ are satisfied. Then $W_h(B_+)$ is isomorphic to $Hk^0({\cal F}_h(G^*),{\cal F}_h(N_-),\chi_h^{s})^{opp}$ as an associative algebra. \end{proposition} \section{Poisson reduction and the Whittaker model} \setcounter{equation}{0} \setcounter{theorem}{0} In this section we start the proof of Theorem ${\rm B}_q$. We shall analyse the quasiclassical limit of the algebra ${\cal F}_h(B_+)^{{\cal F}_h(N_-)}$. Using results of Section \ref{poisred} we realize this limit algebra as the algebra of functions on a reduced Poisson manifold. Denote ${\cal F}(N_\pm)=p({\cal F}_h(N_\pm)),~{\cal F}(B_\pm)=p({\cal F}_h(B_\pm)), ~{\cal F}(H)=p({\cal F}_h(H))$. We denote by $\chi_h^{s}$ the character of the Poisson subalgebra ${\cal F}(N_-)$ such that $\chi^{s}(p(x))=\chi_h^{s}(x)~(\mbox{mod }h)$ for every $x\in {\cal F}_h(N_-)$. From (\ref{charq}) we have \begin{equation}\label{charcl} (\chi^{s}\otimes id)p(N^{-,V})= \prod_{i=1}^lexp[2d_ic_{i}^0 \otimes \pi_V(X_{-\alpha_i})],~~c_i^0=c_i~(\mbox{mod }h). \end{equation} Let ${{\cal F}(N_-)}_{\chi^{s}}$ be the kernel of the character $\chi^{s}$ so that one has a direct sum \begin{equation} {\cal F}(N_-)={\Bbb C}\oplus {{\cal F}(N_-)}_{\chi^{s}}. \end{equation} Similarly to (\ref{maindecqg}) we have the direct sum \begin{equation}\label{maindeccl} {\cal F}(G^*)={\cal F}(B_+)\oplus I_{\chi^{s}}, \end{equation} where $I_{\chi^{s}}={\cal F}(G^*){{\cal F}(N_-)}_{\chi^{s}}$ is the left--sided ideal generated by ${{\cal F}(N_-)}_{\chi^{s}}$. Denote by $\rho_{\chi^{s}}$ the projection onto ${\cal F}(B_+)$ in the direct sum (\ref{maindeccl}). Using Lemma ${\rm A}_q$ we define the quasiclassical limit of action (\ref{mainactqg}) by \begin{equation}\label{mainactcl} x\cdot v =\rho_{\chi^{s}}(\{x,v\} ), \end{equation} where $v\in {\cal F}(B_+)$ and $x\in {\cal F}(N_-)$. We shall describe the space of invariants ${\cal F}(B_+)^{{\cal F}(N_-)}$ with respect to this action by analysing ``dual geometric objects''. First observe that algebra ${\cal F}(B_+)^{{\cal F}(N_-)}$ is a particular example of the reduced Poisson algebra introduced in Lemma \ref{redspace}. Indeed, define a map $\mu_{N_-}:G^* \rightarrow N_-$ by \begin{equation}\label{mun} \mu_{N_+}(L_+,L_-)=n_-, \end{equation} where $n_-$ is given by (\ref{fact}). $\mu_{N_-}$ is a morphism of algebraic varieties. We also note that by definition ${\cal F}(N_-)=\{ \varphi\in {\cal F}(G^*):\varphi= \varphi(n_-)\}$. Therefore ${\cal F}(N_-)$ is generated by the pullbacks of regular functions on $N_-$. Since ${\cal F}(N_-)$ is a Poisson subalgebra in ${\cal F}(G^*)$, and regular functions on $N_-$ are dense in $C^\infty(N_-)$ on every compact subset, we can equip the manifold $N_-$ with the Poisson structure in such a way that $\mu_{N_+}$ becomes a Poisson mapping. Let $u$ be the element defined by (\ref{u}), \begin{equation} u=\prod_{i=1}^lexp[2d_ic_{i}^0 X_{-\alpha_i}]~ \in N_-. \end{equation} Then from (\ref{charcl}) it follows that $\chi^s(\varphi)=\varphi (u)$ for every $\varphi \in {\cal F}(N_-)$. $\chi^s$ naturally extends to a character of the Poisson algebra $C^\infty(N_-)$. Now applying Lemma \ref{redspace} for $M=G^*,~B=N_-,~\pi=\mu_{N_+},~b=u$ we can define the reduced Poisson algebra $C^\infty(\mu_{N_+}^{-1}(u))^{C^\infty(N_-)}$ (see also Remark \ref{redpoisalg}). Denote by $I_u$ the ideal in $C^\infty(G^*)$ generated by elements $\mu_{N_+}^*\psi,~\psi \in C^\infty(N_-), ~\psi(u)=0$. Let $P_u:C^\infty(G^*)\rightarrow C^\infty(G^*)/I_u=C^\infty(\mu_{N_+}^{-1}(u))$ be the canonical projection. Then the action (\ref{redact}) of $C^\infty(N_-)$ on $C^\infty(\mu_{N_+}^{-1}(u))$ takes the form: \begin{equation}\label{actred} \psi\cdot \varphi=P_u(\{ \mu_{N_+}^*\psi, \tilde \varphi\}), \end{equation} where $\psi \in C^\infty(N_-),~\varphi \in C^\infty(\mu_{N_+}^{-1}(u))$ and $\tilde \varphi \in C^\infty(G^*)$ is a representative of $\varphi$ such that $P_u\tilde \varphi=\varphi$. \begin{lemma}\label{redreg} $\mu_{N_+}^{-1}(u)$ is a subvariety in $G^*$. Moreover, the algebra ${\cal F}(B_+)^{{\cal F}(N_-)}$ is isomorphic to the algebra of regular functions on $\mu_{N_+}^{-1}(u)$ which are invariant with respect to the action (\ref{actred}) of $C^\infty(N_-)$ on $C^\infty(\mu_{N_+}^{-1}(u))$, i.e. $$ {\cal F}(B_+)^{{\cal F}(N_-)}={\cal F}(\mu_{N_+}^{-1}(u))\cap C^\infty(\mu_{N_+}^{-1}(u))^{C^\infty(N_-)}. $$ \end{lemma} {\em Proof.} By definition $\mu_{N_+}^{-1}(u)$ is a subvariety in $G^*$. Next observe that $I_{\chi^{s}}= {\cal F}(G^*)\cap I_u$. Therefore the algebra ${\cal F}(B_+)={\cal F}(G^*)/I_{\chi^{s}}$ is identified with the algebra of regular functions on $\mu_{N_+}^{-1}(u)$. Since ${\cal F}(N_-)$ is dense in $C^\infty(N_-)$ on every compact subset in $N_-$ we have: $$ C^\infty(\mu_{N_+}^{-1}(u))^{C^\infty(N_-)}\cong C^\infty(\mu_{N_+}^{-1}(u))^{{\cal F}(N_-)}. $$ Finally observe that action (\ref{actred}) coincides with action (\ref{mainactcl}) when restricted to regular functions. We shall realize the algebra $C^\infty(\mu_{N_+}^{-1}(u))^{C^\infty(N_-)}$ as the algebra of functions on a reduced Poisson manifold. In the spirit of Lemma \ref{redspace} we shall construct a map that forms a dual pair together with the mapping $\mu_{N_+}$. In this construction we use the dressing action of the Poisson--Lie group $G$ on $G^*$ (see Proposition \ref{dressingact}). Consider the restriction of the dressing action $G^*\times G \rightarrow G^*$ to the subgroup $N_+\subset G$. Note that by Proposition \ref{bpm} (i), (iii) and Proposition \ref{admiss} $N_+$ is an admissible subgroup in $G$. Therefore $C^\infty (G^*)^{N_+}$ is a subalgebra in the Poisson algebra $C^\infty (G^*)$. \begin{proposition} The algebra $C^\infty (G^*)^{N_+}$ is the centralizer of $\mu_{N_+}^*\left( C^\infty \left( N_-\right)\right)$ in the Poisson algebra $C^\infty (G^*)$. \end{proposition} {\em Proof.} We shall prove the proposition in a few steps. First we restrict the dressing action of $G$ on $G^*$ the the Borel subgroup $B_+$. According to part (iii) of Proposition \ref{bpm} $({\frak b}_+,{\frak b}_-)$ is a subbialgebra of $({\frak g},{\frak g}^*)$. Therefore $B_+$ is a Poisson--Lie subgroup in $G$. By Proposition \ref{dressingact} for $X \in {\frak b}_+$ we have: \begin{equation} L_{\widehat X} \varphi(L_+,L_-) =( \theta_{G^*}(L_+,L_-) , X ) (\xi_\varphi )= (r_-^{-1}\mu_{B_+}^*(\theta_{B_-}) , X) (\xi_\varphi ), \end{equation} where $\widehat X$ is the corresponding vector field on $G^*$, $\xi_\varphi $ is the Hamiltonian vector field of $\varphi \in C^\infty (G^*)$, and the map $\mu_{B_+}:G^*\rightarrow B_-$ is defined by $\mu_{B_+}(L_+,L_-)=L_-$. Now from Proposition \ref{bpm} (iv) and the definition of the moment map it follows that $\mu_{B_+}$ is a moment map for the dressing action of the subgroup $B_+$ on $G^*$. Observe that the orthogonal complement of the Lie subalgebra ${\frak n}_+ \subset {\frak b}_+$ in the dual space ${\frak b}_-$ coincides with the Lie subalgebra ${\frak h}\subset {\frak b}_-$. Hence by Proposition \ref{admiss} $N_+$ is an admissible subgroup in the Lie--Poisson group $B_+$. Moreover the dual group $B_-$ is the semidirect product of the Lie groups $H$ and $N_-$ corresponding to the Lie algebras ${\frak n}_+^\perp ={\frak h}$ and ${\frak n}_+^*={\frak n}_-$ , respectively. We conclude that all the conditions of Proposition \ref{QPmoment} are satisfied with $A=B_+ , K=N_+ , A^*=B_-, T=N_- , K^\perp = H, \mu =\mu_{B_+}$. It follows that the algebra $C^\infty (G^*)^{N_+}$ is the centralizer of $\mu_{N_+}^*\left( C^\infty \left( N_-\right)\right)$ in the Poisson algebra $C^\infty (G^*)$. This completes the proof. Let $G^*/N_+$ be the quotient of $G^*$ with respect to the dressing action of $N_+$, $\pi:G^* \rightarrow G^*/N_+$ the canonical projection. Note that the space $G^*/N_+$ is not a smooth manifold. However, in the next section we will see that the subspace $\pi(\mu_{N_+}^{-1}(u))\subset G^*/N_+$ is a smooth manifold. Therefore by Remark \ref{remred} the algebra $C^\infty(\pi(\mu_{N_+}^{-1}(u)))$ is isomorphic to $C^\infty(\mu_{N_+}^{-1}(u))^{C^\infty(N_-)}$. Moreover we will see that $\pi(\mu_{N_+}^{-1}(u))$ has a structure of algebraic variety. Using Lemma \ref{redreg} we will obtain that the algebra ${\cal F}(B_+)^{{\cal F}(N_-)}$ is the algebra of regular functions on this variety. \section{Cross--section theorem}\label{cross} \setcounter{equation}{0} \setcounter{theorem}{0} In this section we describe the reduced space $\pi(\mu_{N_+}^{-1}(u))\subset G^*/N_+$ and the algebra ${\cal F}(B_+)^{{\cal F}(N_-)}$. We also complete the proof of Theorem ${\rm B}_q$. First observe that using the embedding $q:G^*\rightarrow G$ (see Proposition \ref{dressingact}) one can reduce the study of the dressing action to the study of the action of $G$ on itself by conjugations. This simplifies many geometric problems. In particular, consider the restriction of this action to the subgroup $N_+$. Denote by $\pi_q:G\rightarrow G/N_+$ the canonical projection onto the quotient with respect to this action. Then we can identify the reduced space $\pi(\mu_{N_+}^{-1}(u))$ with the subspace $\pi_q(q(\mu_{N_+}^{-1}(u)))$ in $G/N_+$. Using this identification we shall explicitly describe the reduced space $\pi(\mu_{N_+}^{-1}(u))$. We start with description of the image of the ``level surface'' $\mu_{N_+}^{-1}(u)$ under the embedding $q$. \begin{proposition}\label{constrt} Let $q:G^*\rightarrow G$ be the map introduced in Proposition \ref{dressingact}, $$ q(L_+,L_-)=L_-L_+^{-1}. $$ Then $q(\mu_{N_+}^{-1}(u))$ is a subvariety in $N_+sN_+$. \end{proposition} {\em Proof.} First, using definition (\ref{mun}) of the map $\mu_{N_+}$ we can describe the space $\mu_{N_+}^{-1}(u)$ as follows: \begin{equation}\label{mun1} \mu_{N_+}^{-1}(u)=\{(h_+n_+,s(h_+)u) | n_+ \in N_+ , h_+ \in H \}, \end{equation} since by (\ref{fact}) $h_-=s(h_+)$. Therefore \begin{equation}\label{dva} q(\mu_{N_+}^{-1}(u))= \{ s(h_+)un_+^{-1}h_+^{-1}| n_+ \in N_+ , h_+ \in H \}. \end{equation} Now recall that $u \in N_+sN_+\cap N_-$, and hence \begin{equation}\label{ras} un_+^{-1}\in N_+sN_+. \end{equation} Next, the space $N_+sN_+$ is invariant with respect to the following action of $H$: \begin{equation}\label{tri} h\circ L= s(h)Lh^{-1}. \end{equation} Indeed, let $L=vsu,~ v,u \in N_+$ be an element of $N_+sN_+$. Then \begin{equation} h\circ L=s(h)vs(h)^{-1}s(h)sh^{-1}huh^{-1}=s(h)vs(h)^{-1}shuh^{-1}. \end{equation} The r.h.s. of the last equality belongs to $N_+sN_+$ because $H$ normalizes $N_+$. Comparing action (\ref{tri}) with (\ref{dva}) and adding (\ref{ras}) we obtain that $q(\mu_{N_+}^{-1}(f)) \subset N_+sN_+$. Since $q$ is an embedding, $q(\mu_{N_+}^{-1}(f))$ is a subvariety in $N_+sN_+$. This concludes the proof. We identify $\mu_{N_+}^{-1}(u)$ with the subvariety in $N_+sN_+$ described in the previous proposition. As we observed in the beginning of this section the reduced space $\pi(\mu_{N_+}^{-1}(u))$ is isomorphic to $\pi_q(q(\mu_{N_+}^{-1}(u)))$. Note that by Proposition \ref{constrt} $q(\mu_{N_+}^{-1}(u))\subset N_+sN_+$. But the variety $N_+sN_+$ is stable under the action of $N_+$ by conjugations. Therefore to describe the reduced space $\pi_q(q(\mu_{N_+}^{-1}(u)))$ we have to study the structure of the quotient $N_+sN_+/N_+$. Our main geometric result is \vskip 0.3cm \noindent {\bf Theorem $\bf C_q$ (\cite{SS}, Theorem 3.1)} {\em Let $N_+'=\{ v \in N_+|s^{-1}(v)\in N_- \}$. Then the action of $N_+$ on $N_+sN_+$ by conjugations is free, and $N_+'s$ is a cross--section for this action, i.e. for each $L\in N_+sN_+$ there exists a unique element $n\in N_+$ such that $n L n^{-1}\in N_+'s$. Moreover, the projection $\pi_q: N_+sN_+\rightarrow N_+'s$ is a morphism of varieties.} \begin{lemma}{\bf (\cite {Ch}, Theorem 8.4.3, \cite{st} Lemma 7.2)} $N_+'\subset N$ is an abelian subgroup, $\dim N_+'=l.$ Moreover, every element $L\in N_+sN_+$ may be uniquely represented in the form $L=vsu , v\in N_+' , u\in N_+$. \end{lemma} \noindent {\em Proof of Theorem $ C_q$.} Denote by $h$ the Coxeter number of $\frak g$. By definition $h$ is the order of the Coxeter element, $s^h=id$. Note that $h={2N \over l}$. Let $C_s\subset W$ be the cyclic subgroup generated by the Coxeter element. $% C_s$ has exactly $l$ different orbits in $\Delta$. The proof depends on the structure of these orbits. For this reason we have to distinguish several cases\footnote{ The proofs given below do not apply when $\frak g$ is the simple Lie algebra of type $E_6$.}. 1.Let ${\frak g}$ be of type $ A_l.$ The following lemma is checked by straightforward calculation. \begin{lemma} (i) Each orbit of $C_s$ in $\Delta$ consists of exactly $h$ elements. One can order these orbits in such a way that $k$-th orbit contains all positive roots of height $k$ and all negative roots of height $h-k.$ \end{lemma} Put \[ {\frak n}_k=\bigoplus_{\left\{ \alpha \in \Delta _{+},\;ht\,\alpha =k\right\} }{\frak n}_\alpha ,N_k=\exp {\frak n}_k. \] For each $k$ we can choose $\gamma _k\in \Delta _{+}$ in such a way that \[ {\frak n}_k=\bigoplus_{p=0}^{h-k-1}{\frak n}_{s^{-p}\left( \gamma_k\right) }. \] Put ${\frak n}_k^p={\frak n}_{s^{-p}\left( \gamma_k\right) },N_k^p=\exp {\frak % n}_k^p.$ Let $L=vsu,v\in N_+',u\in N_+$. We must find $n\in N_+$ such that \begin{equation} nvsu=v_0sn,~v_0\in N_+'. \label{nx} \end{equation} For any $n\in N_+$ there exists a factorization \[ n=n_1n_2\ldots n_l,\mbox{ where } n_k\in N_k. \] Moreover, each $n_k$ may be factorized as \[ n_k=n_k^0n_k^1\ldots n_k^{h-k-1},\;n_k^p\in N_k^p. \] For any $n\in N_+$ the element $nvsu$ admits a representation \[ nvsu=\tilde vs\tilde u,\;\tilde v\in N_+',~ \tilde u\in N_+. \] Let \[ \tilde u=\overrightarrow{\prod_{k=1}^l}\overrightarrow{\,\prod_{p=0}^{h-k-1}}% \tilde u_k^p,\;\tilde u_k^p\in N_k^p, \] be the corresponding factorization of $\tilde u.$ \begin{lemma} We have $\tilde u_k^p= s^{-1}\left( n_k^{p-1}\right) V_k^p,$ where the factors $V_k^p\in N_k^p$ depend only on $u,v$ and on $n_j^q$ with $% j<k.$ \end{lemma} Assume now that $n$ satisfies (\ref{nx}). Then we have $\tilde v=v_0,\tilde u=n.$ This leads to the following relations: \begin{equation} s^{-1}\left( n_k^{p-1}\right) V_k^p=n_k^p, \label{recur} \end{equation} where we set formally $n_k^{-1}=1.$ \begin{lemma} The system (\ref{recur}) may be solved recursively starting with $k=1,$ $% p=0. $ \end{lemma} Clearly, the solution is unique. This concludes the proof for ${\frak g}$ of type $A_l.$ 2. Let now ${\frak g}$ be a simple Lie algebra of type other than $A_l$ and $% E_6$. \begin{lemma} (i) The Coxeter number $h$ is even. (ii) Each orbit of $C_s$ in $\Delta$ consists of exactly $% h$ elements and contains an equal number of positive and negative roots. (iii) Put \[ \Delta _{+}^p=\{\alpha \in \Delta _{+};s^{p}\alpha \notin \Delta _{+}\},\;{\frak n}^p=\bigoplus_{\alpha \in \Delta _{+}^p}{\Bbb C}\cdot X_\alpha ; \] then ${\frak n}^p\subset {\frak n}$ is an abelian subalgebra, $\dim {\frak n}% ^p=l.$ \end{lemma} If $\frak g$ is not of type $D_{2k+1}$ this assertion follows from Proposition 33, Chap.6, no. 1.11 and Corollary 3, Chap.5, no. 6.2 in \cite{Bur}. For $\frak g$ of type $D_{2k+1}$ it may be checked directly. Put $N^p=\exp {\frak n}^p$. Let $N^p$ be the corresponding subgroup of $G.$ Let $L=vsu,~v\in N_+',~u\in N_+$. We must find $n\in N_+$ such that \[ vsu=nv_0sn^{-1},~v_0\in N_+'. \] Put \begin{equation} n=n_1n_2\ldots n_{\frac h2},~~n_p\in {N}_p. \label{n} \end{equation} The elements $n_p$ will be determined recursively. We have \begin{equation} vs\left( u\right) =\overrightarrow{\prod_p }n_p v_0 s\left( \overleftarrow{\prod_p}n_p^{-1}\right). \label{nnn} \end{equation} We shall say that an element $x\in G$ is in the big cell in $G$ if $x \in B_+N_-\subset G.$ \begin{lemma} $vs\left( u\right) $ is in the big cell in $G$ and admits a factorization \[ vs\left( u\right) =x_{+}^1 x_{-}^1,\;x_{+}^1\in N_+% ,\;x_{-}^1\in N_-. \] \end{lemma} Indeed, let $u=u_{h/2}u_{h/2-1}\ldots u_1,~u_p\in N^p$ be a similar decomposition of $u$. Then we have simply $x_{-}=s\left( u_1\right) .\ $(It is clear that $x_{+}^1\in B_+$ actually does not have an $H$-component and so belongs to $N_+$ A comparison of the r.h.s in (\ref{nnn}) with the Bruhat decomposition of the l.h.s. immediately yields that the first factor in (\ref{n}) is given by $n_1=s^{-1}\left( x_{-}\right) ^{-1}.$ Assume that $n_1,n_2,\ldots ,n_{k-1}$ are already computed. Put \[ m_k=n_1n_2\ldots n_{k-1} \] and consider the element \begin{equation} L^k:=s^{k-1}\left( m_k^{-1}vs(u)s(m_k)\right) . \label{Lk} \end{equation} \begin{lemma} $L^k$ is in the big cell in $G$ and admits a factorization \begin{equation} L^k=x_{+}^kx_{-}^k,\;x_{+}^k\in N_+,\;x_{-}^k\in N_-. \label{fk} \end{equation} \end{lemma} The elements $x_{\pm }^k$ are computed recursively from the known quantities. By applying a similar transform to the r.h.s. of (\ref{nnn}) we get \begin{eqnarray} L^k &=&s^{k-1}\left( m_k^{-1} \overrightarrow{\prod_p}n_p v_0s\left( \overleftarrow{\prod_p}n_p^{-1}\right) s(m_k) \right) = \label{Lrhs} \end{eqnarray} \[ s^{k-1}\left( \overrightarrow{\prod_{p\geq k}}n_p v_0\right) s^{k}\left( \overleftarrow{\prod_{p\geq k+1}} n_p^{-1}\right)s^{k}\left( n_k^{-1}\right) . \] Comparison of (\ref{Lrhs}) and (\ref{fk}) yields $x_{-}^k=s^{k}\left( n_k^{-1}\right)$. Hence $n_k=s^{-k}\left( x_{-}^k\right)^{-1}$, which concludes the induction. Finally observe that by construction the map $\pi_q: N_+sN_+\rightarrow N_+'s$ is a morphism of varieties. \begin{corollary}\label{var} The space $\pi(\mu_{N_+}^{-1}(u))$ is a subvariety in $N_+'s$. The algebra ${\cal F}(B_+)^{{\cal F}(N_-)}$ is isomorphic to the algebra of regular functions on $\pi(\mu_{N_+}^{-1}(u))$. \end{corollary} {\em Proof.} First observe that by construction $\pi(\mu_{N_+}^{-1}(u))\cong \pi_q(q(\mu_{N_+}^{-1}(u)))$ is a subvariety in $N_+'s$. In particular, $\pi(\mu_{N_+}^{-1}(u))$ is a smooth manifold. Hence by Remark \ref{remred} the map $$ C^\infty(\pi(\mu_{N_+}^{-1}(u)))\rightarrow C^\infty(\mu_{N_+}^{-1}(u))^{C^\infty(N_-)},~~\psi \mapsto \pi^*\psi $$ is an isomorphism. Now observe that by construction the map $\pi: \mu_{N_+}^{-1}(u)\rightarrow \pi(\mu_{N_+}^{-1}(u))$ is a morphism of varieties. Therefore if $\psi \in {\cal F}(\pi(\mu_{N_+}^{-1}(u)))$ then $\pi^*\psi$ is a regular function on $\mu_{N_+}^{-1}(u)$. Conversely, suppose that $\varphi \in {\cal F}(\mu_{N_+}^{-1}(u))\cap C^\infty(\mu_{N_+}^{-1}(u))^{C^\infty(N_-)}$. Note that $\pi(\mu_{N_+}^{-1}(u))$ may be regarded as a subvariety in $\mu_{N_+}^{-1}(u)$ (see Remark \ref{redpoisalg}). Then the restriction of $\varphi$ to $\pi(\mu_{N_+}^{-1}(u))\subset \mu_{N_+}^{-1}(u)$ is a regular function. Therefore the map $$ {\cal F}(\pi(\mu_{N_+}^{-1}(u)))\rightarrow {\cal F}(\mu_{N_+}^{-1}(u))\cap C^\infty(\mu_{N_+}^{-1}(u))^{C^\infty(N_-)},~~\psi \mapsto \pi^*\psi $$ is an isomorphism. Finally observe that by Lemma \ref{redreg} the algebra ${\cal F}(\mu_{N_+}^{-1}(u))\cap C^\infty(\mu_{N_+}^{-1}(u))^{C^\infty(N_-)}$ is isomorphic to ${\cal F}(B_+)^{{\cal F}(N_-)}$. This completes the proof. Theorem ${\rm C}_q$ is a group counterpart of Theorem C. Moreover the space $N_+'s$ naturally appears in the study of regular elements in $G$. Recall that an element of $G$ is called regular if its centralizer in $G$ is of minimal possible dimension. Let $R$ be the set of regular elements in $G$. Clearly, $R$ is stable under the action of $G$ on itself by conjugations and in fact $R$ is the union of all $G$ orbits in $G$ of maximal dimension. A function $\psi$ on $G$ is called a class function if $f(x)=f(y)$ whenever $x$ and $y$ are conjugate points of definition of $\psi$. We denote by ${\cal F}^G(G)$ the algebra of regular class functions on $G$. \vskip 0.3cm \noindent {\bf Theorem $\bf D_q$ (\cite{st}, Theorems 1.4 and 6.1)} {\em Let $G$ be a complex connected simply connected simple algebraic group. Then The space $N_+'s$ is contained in $R$ and is a cross--section for the action of $G$ on $R$. That is every $G$--orbit in $G$ of maximal dimension intersects $N_+'s$ in one and only one point. The algebra of regular class functions on $G$ is freely generated as a commutative algebra over $\Bbb C$ by the characters of fundamental representations of $G$, $\chi_1,\ldots , \chi_l$. Moreover, $N_+'s$ is an algebraic variety, and the algebra of regular functions on $N_+'s$ is freely generated as a commutative algebra over $\Bbb C$ by the restrictions of the characters $\chi_1,\ldots , \chi_l$ to $N_+'s$ .} \vskip 0.3cm \noindent {\bf Theorem $\bf E_q$} {\em For any $\psi \in {\cal F}^G(G)$ one has $\rho_{\chi^{s}}(p^*\psi) \in {\cal F}(B_+)^{{\cal F}(N_-)}$. Furthermore the map \begin{equation}\label{isoreg} {\cal F}^G(G)\rightarrow {\cal F}(B_+)^{{\cal F}(N_-)}, \psi \mapsto \rho_{\chi^{s}}(p^*\psi) \end{equation} is an algebra isomorphism. In particular, $$ {\cal F}(B_+)^{{\cal F}(N_-)}={\Bbb C}[\rho_{\chi^{s}}(p^*\chi_1),\ldots , \rho_{\chi^{s}}(p^*\chi_l)] $$ is a polynomial algebra in $l$ generators.} \vskip 0.3cm \noindent {\em Proof.} Let $\psi$ be an element of ${\cal F}^G(G)$. The restriction of $\psi$ to the subvariety $\pi(\mu_{N_+}^{-1}(u))\cong \pi_q(q(\mu_{N_+}^{-1}(u)))\subset N_+'s \subset G$ is a regular function. Using the isomorphism ${\cal F}(\pi(\mu_{N_+}^{-1}(u)))\cong {\cal F}(B_+)^{{\cal F}(N_-)}$ (see Corollary \ref{var}) this restriction may be identified with $\rho_{\chi^{s}}(q^*\psi)\in {\cal F}(B_+)^{{\cal F}(N_-)}$. By Theorem ${\rm D}_q$ the algebra ${\cal F}(N_+'s)$ is freely generated as a commutative algebra over $\Bbb C$ by the restrictions of the fundamental characters $\chi_1,\ldots , \chi_l$. Since $\pi(\mu_{N_+}^{-1}(u))$ is a subvariety in $N_+'s$ the algebra ${\cal F}(\pi(\mu_{N_+}^{-1}(u)))$ is generated by the restrictions of the fundamental characters $\chi_1,\ldots , \chi_l$. Therefore the map (\ref{isoreg}) is surjective. We have to prove that it is injective. Let $\chi_i$ be a fundamental character. Consider the restriction of the function $\rho_{\chi^{s}}(q^*\chi_i)$ to the subspace in $\mu_{N_+}^{-1}(u)$ formed by elements (see (\ref{mun1})): $$ (h_+,s(h_+)u),~h_+\in H. $$ Then $\rho_{\chi^{s}}(q^*\chi_i)(h_+,s(h_+)u)=\chi_i(s(h_+)uh_+^{-1})$. Since $\chi_i$ is a character we have $\chi_i(s(h_+)uh_+^{-1})=\chi_i(h_+^{-1}s(h_+)u)$. The element $u$ is unipotent, and hence $\chi_i(h_+^{-1}s(h_+)u)=\chi_i(h_+^{-1}s(h_+))$. Now recall that the restrictions of the fundamental characters to the Cartan subgroup are algebraically independent (they are given by the well--known Weyl formula). Therefore (\ref{isoreg}) is an isomorphism. This completes the proof. \vskip 0.3cm \noindent {\em Proof of Theorem $B_q$.} Let $p:{\cal F}_h(G^*)\rightarrow {\cal F}(G^*)$ be the map defined in Proposition \ref{quantreg}. Let $W_h^{Rep}(B_+)$ be the subalgebra in $W_h(B_+)$ topologically generated by the elements $C_{V_i}^{\rho_{\chi_h^{s}}}=\rho_{\chi_h^{s}}(C_{V_i}),~i=1,\ldots ,l$. From the the definition of the elements $C_{V_i}^{\rho_{\chi_h^{s}}}$ it follows that $p(C_{V_i}^{\rho_{\chi_h^{s}}})=\rho_{\chi^{s}}(p^*\chi_i)$. Therefore by Theorem ${\rm E}_q$ $p(W_h^{Rep}(B_+))={\cal F}(B_+)^{{\cal F}(N_-)}$, and $W_h^{Rep}(B_+)$ is freely generated as a commutative topological algebra over ${\Bbb C}[[h]]$ by the elements $C_{V_i}^{\rho_{\chi_h^{s}}}=\rho_{\chi_h^{s}}(C_{V_i}),~i=1,\ldots ,l$. On the other hand using the definitions of the algebras ${\cal F}_h(B_+)^{{\cal F}_h(N_-)}$ and ${\cal F}(B_+)^{{\cal F}(N_-)}$ it is easy to see that $p({\cal F}_h(B_+)^{{\cal F}_h(N_-)})= {\cal F}(B_+)^{{\cal F}(N_-)}$. We shall prove that $W_h^{Rep}(B_+)$ is isomorphic to ${\cal F}_h(B_+)^{{\cal F}_h(N_-)}$. Let $I\in {\cal F}_h(B_+)^{{\cal F}_h(N_-)}$ be an invariant element. Then $p(I)\in {\cal F}(B_+)^{{\cal F}(N_-)}$, and hence one can find an element $K_0\in W_h^{Rep}(B_+)$ such that $I-K_0=hI_1,~I_1\in {\cal F}_h(B_+)^{{\cal F}_h(N_-)}$. Applying the same procedure to $I_1$ one can find elements $K_1\in W_h^{Rep}(B_+), ~I_2\in {\cal F}_h(B_+)^{{\cal F}_h(N_-)}$ such that $I_1-K_1=hI_2$, i.e. $I-K_0-hK_1=0~(\mbox{mod }h^2)$. We can continue this process. Finally we obtain an infinite sequence of elements $K_i\in W_h^{Rep}(B_+)$ such that $I-\sum_{i=0}^p h^pK_p=0~(\mbox{mod }h^{p+1})$. Since the space ${\cal F}_h(B_+)$ is complete in the $h$--adic topology the series $\sum_{i=0}^\infty h^pK_p\in W_h^{Rep}(B_+)$ converges to $I$. Therefore $I\in W_h^{Rep}(B_+)$, and hence ${\cal F}_h(B_+)^{{\cal F}_h(N_-)}$ is isomorphic to $W_h^{Rep}(B_+)$. We also have the following inclusions: $$ W_h^{Rep}(B_+)\subseteq W_h(B_+)\subseteq {\cal F}_h(B_+)^{{\cal F}_h(N_-)} \cong W_h^{Rep}(B_+). $$ Therefore $W_h^{Rep}(B_+)$ coincides with $W_h(B_+)$. This proves part (ii) of Theorem ${\rm A}_q$ and Theorem ${\rm B}_q$. \chapter*{Acknowledgements} The first words of gratitude are due to Professor Michael Semenov--Tian--Shansky for guidance into the world of Modern Mathematical Physics. This thesis is essentially based on his ideas. I am greatly indebted my advisor in Uppsala Doctor Anton Alekseev for his continuous support, for many insightful discussions and for continuous encouragement. I gratefully acknowledge Prof. Antti Niemi and his research group at the Department of Theoretical Physics at Uppsala University for providing an excellent working atmosphere. The Department of Theoretical Physics in Uppsala University has provided excellent facilities for research, which are gratefully acknowledged. I wish to thank MSc Mats Lilja for his patient help in many and various practical matters during my stay in Uppsala. \pagestyle{plain}
\section{INTRODUCTION} Shortly after the launch of the Rossi X-ray Timing Explorer ({\it RXTE}) in December 1995, observation with {\it RXTE}\ of neutron-star low-mass X-ray binaries (LMXBs) revealed that several sources had a single, highly coherent, high-amplitude brightness oscillation during at least one thermonuclear X-ray burst (for reviews see Strohmayer, Zhang, \& Swank 1997; Strohmayer, Swank, \& Zhang 1998a). The asymptotic frequency of these oscillations in the tails of bursts is so similar in different bursts from a single source, and the oscillation is so coherent in the tail (see, e.g., Strohmayer \& Markwardt 1999), that it is almost certain that this asymptotic frequency is the stellar spin frequency or its first overtone. These burst oscillations therefore provided the first direct evidence for the value of the spin frequencies of these LMXBs, and they corroborate strongly the proposed evolutionary link between LMXBs and millisecond rotation-powered pulsars. In addition, the stability of the frequency in the tails of the bursts has led to their application as promising probes of the binary systems themselves (Strohmayer et al.\ 1998b). The existence of burst oscillations indicates that the emission from the surface, and hence the thermonuclear burning, is not uniform over the entire star. This is in accord with theoretical expectations (Joss 1978; Ruderman 1981; Shara 1982; Livio \& Bath 1982; Fryxell \& Woosley 1982; Nozakura, Ikeuchi, \& Fujimoto 1984; Bildsten 1995), and it suggests that the properties of the burst oscillations, such as the evolution of their frequency or amplitude, may contain valuable information about the propagation of thermonuclear burning over the surface of the neutron star. The lessons learned from study of the thermonuclear propagation in bursts may ultimately further our understanding of thermonuclear propagation in other astrophysical contexts, such as classical novae and Type Ia supernovae. Unlike in novae or Type Ia supernovae, burning in thermonuclear X-ray bursts occurs near the surface and occurs often for a single source, and is therefore relatively easy to observe. The detailed study of burst brightness oscillations therefore has broad importance. Here we describe in detail the frequency behavior of the burst oscillations in five bursts from 4U~1636--536, which is an LMXB with an orbital period of 3.8 hours (see, e.g., van Paradijs et al.\ 1990). This source is of special interest because it produces detectable signals at both the fundamental and the first overtone of the stellar spin frequency (Miller 1999), and because near the beginning of one burst the brightness oscillations reached the highest amplitude ---50\% rms--- so far recorded for oscillations during a thermonuclear burst (Strohmayer et al.\ 1998c). In \S~2 we analyze the light curves of the bursts, and the frequency and amplitude of the brightness oscillations in the four of those five bursts that have strong brightness oscillations for most of the duration of the burst. We find that, despite apparent similarities in the light curves of three of those four bursts, the amplitude and frequency behavior of their brightness oscillations are very different from each other. We also find compelling evidence in one burst, and strong evidence in another burst, for an interval in which the burst oscillation frequency decreases after the peak in the light curve. In \S~3 we focus on the initial portions of the bursts. Analyses of bursts from many sources have shown that the oscillation frequency often changes by a few Hertz over the first few seconds of a burst (see Strohmayer et al.\ 1998a for a review). The change is often a monotonic rise, but there are indications of more complicated behavior in some bursts. It has been pointed out that the magnitude of the frequency change could be explained by a 20--50 meter expansion of the burning layers followed by a slow settling, if the layers conserve angular momentum (see, e.g., Strohmayer et al.\ 1998a), but details have not been worked out. For example, it is not clear how the layers would maintain their coherence throughout the 5--10 complete circuits relative to the body of the star that are implied by the observations. Bildsten (1998) has suggested that the layers may be stabilized by thermal buoyancy or mean molecular weight stratification, but details have not been worked out. In \S~3 we examine in detail the first 0.75 seconds of all five bursts, which was the interval used to construct the candidate waveform for the $\sim$290~Hz oscillation in 4U~1636--536 (Miller 1999). We examine models of the frequency behavior that have increasing complexity: a constant-frequency model; one with a frequency and frequency derivative; a four-parameter model with an initial frequency and frequency derivative followed by a different frequency derivative after a break time; and a five-parameter model with two different frequencies and frequency derivatives separated by a break time. We find that if the same type of frequency model applies to all five of the bursts then the data do not require a model more complicated than the constant-frequency model or, possibly, the model with a single frequency and frequency derivative. Note, however, that this is not inconsistent with the use of the five-parameter model to construct a waveform used in the search for the expected $\sim$290~Hz oscillation (Miller 1999); in such a search, the only goal is to find the best fit to the $\sim$580~Hz oscillations, and the extra parameters need not be justified by a significantly better fit. Finally, in \S~4 we discuss the implications of these results for the current picture of the frequency changes, in which the frequency change occurs because the burning layer is lifted by 20--50 meters from the surface by the radiation flux. We find that the simplest version of this picture has difficulty explaining the observations. \section{OVERVIEW OF THE BURSTS} We used public-domain data from the High Energy Astrophysics Science Archive Research Center. The data were taken in Single Bit Mode, which does not record the energy of photons. We give the starting times of the bursts in Table~1 and the light curves in Figure~\ref{figLightcurves}. In burst~d, the data dropouts are caused by telemetry saturation. \begin{table*}[t] \centering \begin{tabular}{cc} \multicolumn{2}{c}{\bfseries Table~1: Starting Time of Bursts}{\rule[-2mm]{0mm}{6mm}}\\ \tableline \tableline Burst&Date and Time\\ \tableline a&22:39:24.188 UTC on 28 December 1996\\ b&23:54:02.876 UTC on 28 December 1996\\ c&23:26:46.813 UTC on 29 December 1996\\ d&17:36:52.941 UTC on 31 December 1996\\ e&09:57:25.938 UTC on 23 February 1997\\ \tableline \end{tabular} \end{table*} \begin{figure} \psfig{file=lightcurves.ps,height=6.0truein,width=6.0truein} \caption[]{ \label{figLightcurves} Light curves for the bursts. (a)~Burst beginning at 22:39:24 UTC on 28 December 1996. (b)~Burst beginning at 23:54:02 UTC on 28 December 1996. (c)~Burst beginning at 23:26:46 UTC on 29 December 1996. (d)~Burst beginning at 17:36:52 UTC on 31 December 1996. (e)~Burst beginning at 09:57:26 UTC on 23 February 1998. The data gaps in burst~d are caused by telemetry saturation.} \end{figure} Figure~\ref{figDynamic} shows the peaks of the power spectra of the first four bursts, as a function of time. The burst on 23 February 1997 does not have a strong brightness oscillation for most of its duration, and we therefore do not analyze it in the rest of this section. For each burst, the frequency of maximum power in successive nonoverlapping one-second intervals is shown by the solid triangles, and the Leahy et al.\ (1983)-normalized power at this frequency is shown by the solid line. Here we plot only those points with Leahy powers in excess of 10 (chance probability for a single trial less than $6\times 10^{-3}$). The horizontal bars on the frequency points indicate the extent of the interval for which the power density spectrum was calculated. In a few cases, more than one peak exceeds this threshold in a given power density spectrum. We then represent the lower-power peak by an open circle. In burst~a the secondary peak has a Leahy power of 21.2 (single-trial significance $2.5\times 10^{-5}$); in burst~b the secondary peaks have Leahy powers of 44.0 (first interval; significance $2.8\times 10^{-10}$) and 13.5 (second interval; significance $1.2\times 10^{-3}$); and in burst~c the secondary peak has a Leahy power of 12.8 (significance $1.7\times 10^{-3}$). Finally, Figure~\ref{figAmplitudes} shows the rms amplitude of each oscillation, computed for one-second intervals 1/8 second apart. \begin{figure} \psfig{file=dynamic.ps,height=6.0truein,width=6.0truein} \caption[]{ \label{figDynamic} Power spectra as a function of time for the four bursts with strong brightness oscillations. Each solid triangle is at the frequency of maximum power and its $1\sigma$ uncertainties for nonoverlapping 1~second intervals. The frequency of a peak is only plotted if its Leahy power exceeds 10. In bursts 1, 2, and 3 there are intervals in which a second peak exceeds this threshold, and this secondary peak is plotted with an open circle. The power of the secondary peak in burst~(a) is 21.2 (single-trial significance $2.5\times 10^{-5}$); the powers of the secondary peaks in burst~(b) are 44.0 (significance $2.8\times 10^{-10}$) and 13.5 (significance $1.2\times 10^{-3}$); and the power of the peak in burst~(c) is 12.8 (significance $1.7\times 10^{-3}$). The panels are labeled as in Figure~\ref{figLightcurves}. Burst~(e) does not have strong brightness oscillations for most of the burst, and is therefore excluded from this analysis.} \end{figure} \begin{figure} \psfig{file=amp.ps,height=6.0truein,width=6.0truein} \caption[]{ \label{figAmplitudes} Root mean square amplitude of the brightness oscillation for each of the bursts. The amplitudes are calculated for one-second intervals with starting times 1/8 second apart, and the $\pm 1\sigma$ uncertainty bands are shown. The amplitude is only plotted if the Leahy power for the oscillation exceeds 10. The panels are labeled as in Figure~\ref{figLightcurves}.} \end{figure} It is evident from these figures that the frequency behavior can be very complex and can differ greatly from burst to burst. The light curves for bursts~(a), (c), and (d) appear similar to each other, although burst~(d) has a slightly longer decay time than the other two. However, the frequency and amplitude of the brightness oscillations evolve very differently in the three bursts. In burst~(a), there is a strong oscillation near the beginning which disappears for approximately one second, then the oscillation reappears after the peak. The frequency increases continuously, although there is some evidence that in the initial $\sim$0.5 second of the burst the frequency drops (this might help explain the presence of a higher-frequency secondary peak in the power density spectrum). In burst~(c), the brightness oscillation is present for almost the entire time examined. The frequency increases rapidly in the first two to three seconds, then appears to decrease to an asymptotic value. A power density spectrum of a two-second interval starting 1.75 seconds after the beginning of the burst reveals a peak at 581.62$\pm$0.04~Hz. A power density spectrum of a six-second interval starting four seconds after the beginning of the burst has a peak at 581.47$\pm$0.01~Hz. If the latter frequency is the asymptotic frequency of the oscillation, then at a 3$\sigma$ level of certainty it is less than the maximum frequency attained during the burst. The amplitude of the oscillation in the burst tail is high and significant, and there is an abrupt increase in the amplitude 6--8 seconds after the beginning of the burst that is not accompanied by any apparent change in the light curve. In burst~(d) there is a clear {\it decrease} in the frequency of the burst oscillation in the tail of the burst. We explored this further by taking a power density spectrum of a longer interval: five seconds, starting three seconds after the beginning of the burst. We found that, at the 99.99\% confidence level, the frequency change per second during this interval is $-0.54\pm 0.08$~Hz~s$^{-1}$. The best-fit frequency at the beginning of this five-second interval depends on the frequency derivative, and is approximately $\nu_0=581.39{\rm\ Hz}-2({\dot\nu}+0.62~{\rm Hz~s}^{-1}) {\rm\ Hz}$. This means that, relative to a brightness oscillation with a constant frequency equal to the frequency at the beginning of this five-second interval, the observed brightness oscillation has a total phase lag of between $12\pi$ and $16\pi$ radians. The total phase lag is comparable to what is seen in many bursts, except that here the frequency inferred in the tail of the burst is significantly less than the spin frequency inferred from other bursts in this source. There is no sign in this burst that the frequency has reached an asymptotic value. Burst~(b) is the only one with a clearly different light curve. This is a weak burst. The frequency of the brightness oscillation is consistent with what is observed in, at least, bursts~(a) and (c): a rise in the frequency near the beginning of the burst, followed by an approximate leveling off. We note, however, that within the uncertainties the frequency could also reach a maximum and then decline, as appears to be the case for bursts~(c) and (d). \section{BRIGHTNESS OSCILLATIONS AT THE BEGINNING OF THE BURST} Previous analyses have shown that the brightness oscillations in the initial $\sim$second of the bursts are often of particular interest. This is where the highest amplitudes (rms$\sim$50\%; Strohmayer et al.\ 1998c) are reported, and where subharmonics of the strong oscillation have been detected in 4U~1636--536 (Miller 1999) and possibly in the Rapid Burster (Fox \& Lewin 1999). It is therefore important to examine the initial portion more closely to see what hints about the brightness oscillation mechanism can be derived. Before doing so, we need to emphasize an important distinction. If the purpose of the analysis is to characterize the frequency variations of the $\sim$580~Hz oscillation then extra parameters can only be added if the fit to the data is improved sufficiently to justify the additional complexity. The situation is different when the goal is to produce a matched filter for a search for a harmonically related frequency, as in the search for a signal at half of the $\sim$580~Hz dominant brightness oscillation in 4U~1636--536 (Miller 1999). For that purpose, it is not necessary to justify the extra parameters used in the construction of the filter, if no reference is made to the signal for which one is searching. In the case of the search for the $\sim$290~Hz oscillation, a five-parameter matched filter was used for each burst; matched filters with fewer parameters also give a clear signal at $\sim$290~Hz, although with lower significance because the filter does not fit the data as well. A general method to find the best-fit values of parameters and their confidence regions employs a likelihood function. In this approach, we suppose that we have a model in which the countrate as a function of time is predicted to be $s(t)$, from which we can predict the number of counts $s_i$ in one particular bin $i$ of the data, which in this case is 1/8192~s in duration. In general, $s_i$ is not an integer. The actual number of counts observed in bin $i$ is $c_i$, which is an integer. With these definitions, the Poisson likelihood of the full data set given the model $s(t)$ is \begin{equation} {\cal L}=\Pi {s_i^{c_i}\over{c_i!}}e^{-s_i}\; , \end{equation} where the product is over all of the bins of the data. Note that in normal applications of the point likelihood the bin sizes would be so small that a given bin would have either zero or one count, but the fixed bin size of 1/8192~s combined with the high count rates during the bursts (up to $\sim$30,000~c/s; see Figure~\ref{figLightcurves}) means that many of the bins have multiple counts. The likelihood is maximized to determine the best values of the parameters of the model waveform $s(t)$, and approximate confidence contours can be estimated using contours of constant log likelihood: $2\Delta\log{\cal L}=\Delta \chi^2$ (Eadie et al.\ 1971, \S~9.4.3, p. 207). The model waveform $s(t)$ will in general include components related to the relatively slow change in the brightness of the source as well as components related to the high-frequency brightness oscillation. However, the frequency scales are different enough ($\sim$1--5~Hz for the slowly rising component versus $\sim$580~Hz for the fast oscillations) that the fitting of the two components are nearly independent of each other. This means that, when we analyze the behavior of the brightness oscillations, we can simplify by assuming that the burst has a constant average brightness. With this in mind, the model we consider is \begin{equation} s(t)=c_{\rm av}\left(1+A\cos 2\pi[\nu(t)t+\phi_0]\right)\; , \end{equation} where $c_{\rm av}$ is the average countrate, $A$ is the amplitude of the signal (which we assume to be time-independent), $\nu(t)$ is the frequency as a function of time, and $\phi_0$ is the phase of the oscillation at the beginning of the data interval analyzed. In this section we explore four different models for the frequency behavior in the initial 0.75 seconds of each of the four bursts (this time was chosen to conform to the analysis of Miller [1999], which was performed to look for a weaker $\sim$290~Hz oscillation). The four models are: \begin{equation} \nu_1(t)=\nu_0 \end{equation} \begin{equation} \nu_2(t)=\nu_0+{\dot\nu}t \end{equation} \begin{equation} \begin{array}{rl} \nu_3(t) & =\nu_1+{\dot\nu_1}t\;,\;\; t<t_{\rm break}\\ & =\nu_2+{\dot\nu_2}t\;,\;\; t>t_{\rm break} \end{array} \end{equation} where continuity of the frequency is imposed, so that there are four independent parameters, and finally \begin{equation} \begin{array}{rl} \nu_4(t) & =\nu_1+{\dot\nu_1}t\;,\;\; t<t_{\rm break}\\ & =\nu_2+{\dot\nu_2}t\;,\;\; t>t_{\rm break} \end{array} \end{equation} where continuity of the frequency is not imposed, so that there are five independent parameters. For the purposes of this section the most interesting of the parameters of the model waveform $s(t)$ is the frequency, as opposed to the amplitude or the initial phase of the brightness oscillation. If the amplitude $A\ll 1$, as it is in this case, then a tremendous speed-up in the search procedure is possible with the use of the cross-correlation (see, e.g., Helstrom 1960 or Wainstein \& Zubakov 1962 for details of cross-correlation and matched filtering techniques) \begin{equation} H=C\left|\int_{t_0}^{t_0+T}c(t)e^{-i\nu(t)t}dt\right|^2\; , \end{equation} where $t_0$ is the start time of the burst, $T$=0.75~s is the duration of the burst, and $C$ is a normalization constant. In practice this integral is actually calculated as a sum over all of the bins of the data, and $dt$=1/8192~s is the duration of a bin. If $C=2/N_{\rm tot}$, where $N_{\rm tot}$ is the total number of counts in the data set, then $H$ has the same statistical properties as the Leahy power; $H$ is also related to the $Z^2$ statistic used in pulsar period searches (Buccheri et al.\ 1983; see Strohmayer \& Markwardt 1999 for a recent use in the characterization of brightness oscillations during thermonuclear X-ray bursts). To lowest order in the oscillation amplitude $A$ this description is mathematically identical to the likelihood description, but it is much faster to apply because no search need be performed for the amplitude or oscillation phase. It is therefore preferable for low-amplitude oscillations. With this formalism, we can estimate the best values and uncertainty regions for the different frequency models above. The figures in the previous section, which were constructed using a constant-frequency waveform, give this information for the one-parameter, constant-frequency model. \subsection{Two-Parameter Frequency Model} The best-fit values for the two-parameter frequency model are given in Table~2. To estimate uncertainties on these parameters, we performed a Monte Carlo analysis in which we selected $10^6$ random values per burst of $\nu_1$, ${\dot\nu}_1$, ${\dot\nu}_2$, and $t_{\rm break}$, uniformly sampled from, respectively, 576~Hz to 585~Hz; -12~Hz~s$^{-1}$ to 12~Hz~s$^{-1}$; -12~Hz~s$^{-1}$ to 12~Hz~s$^{-1}$; and 0~s to 0.75~s. The quoted uncertainties for single parameters were computed using a Bayesian viewpoint, in which the posterior probability density was calculated throughout the interval and then integrated over the other three parameters to produce a marginalized probability distribution. We have assumed a uniform prior probability density over the whole space searched. This means that the posterior probability density is simply proportional to the likelihood. These confidence regions, which are the smallest regions that encompass 68\% of the probability, are given in Table~3. In some cases the maximum likelihood value of a parameter obtained by extremization in the full two-dimensional parameter space is outside the marginalized 68\% confidence region. This is symptomatic of the fact that the parameters are constrained only weakly by the data. \begin{table}[t] \centering \begin{tabular}{ccc} \multicolumn{3}{c}{\bfseries Table~2: Best-Fit Parameters} {\rule[-2mm]{0mm}{6mm}}\\ \multicolumn{3}{c}{\bfseries for Two-Parameter Model}{\rule[-2mm]{0mm}{6mm}}\\ \tableline \tableline Burst&$\nu_1$ (Hz)&${\dot\nu}_1$ (Hz s$^{-1}$)\\ \tableline a&579.0& 4.0\\ b&580.3& 2.0\\ c&581.0&-3.8\\ d&578.6& 4.2\\ e&581.1&-3.4\\ \tableline \end{tabular} \end{table} \begin{table}[ht] \centering \begin{tabular}{ccc} \multicolumn{3}{c}{\bfseries Table~3: 68\% Confidence Regions} {\rule[-2mm]{0mm}{6mm}}\\ \multicolumn{3}{c}{\bfseries for Two-Parameter Model}{\rule[-2mm]{0mm}{6mm}}\\ \tableline \tableline Burst&$\nu_1$ (Hz)&${\dot\nu}_1$ (Hz s$^{-1}$)\\ \tableline a&579.7--580.1&-8.0--8.0\\ b&580.9--581.2&-8.8--8.0\\ c&579.2--579.7&-8.0--8.0\\ d&578.8--579.9&-8.0--8.0\\ e&579.2--583.6&-8.0--8.0\\ \tableline \end{tabular} \end{table} \subsection{Four-Parameter Frequency Model} The best-fit values for the four-parameter frequency model are given in Table~4. As for the two-parameter model, the uncertainties were estimated by marginalizing over all but the parameter of interest; the confidence regions containing 68\% of the probability are given in Table~5. \begin{table*}[t] \centering \begin{tabular}{ccccc} \multicolumn{5}{c}{\bfseries Table~4: Best-Fit Parameters for Four-Parameter Model}{\rule[-2mm]{0mm}{6mm}}\\ \tableline \tableline Burst&$\nu_1$ (Hz)&${\dot\nu}_1$ (Hz s$^{-1}$) &${\dot\nu}_2$ (Hz s$^{-1}$)&$t_{\rm break}$ (s)\\ \tableline a&581.3&-9.0&11.2&0.28\\ b&579.9& 4.0&0.5&0.32\\ c&579.1& 8.7&-6.7&0.20\\ d&579.7&-2.4&7.9&0.28\\ e&583.2&-12.0&-0.8&0.28\\ \tableline \end{tabular} \end{table*} \begin{table*}[ht] \centering \begin{tabular}{ccccc} \multicolumn{5}{c}{\bfseries Table~5: 68\% Confidence Regions for Four-Parameter Model}{\rule[-2mm]{0mm}{6mm}}\\ \tableline \tableline Burst&$\nu_1$ (Hz)&${\dot\nu}_1$ (Hz s$^{-1}$) &${\dot\nu}_2$ (Hz s$^{-1}$)&$t_{\rm break}$ (s)\\ \tableline a&579.0--581.0&-8.8--2.4&2.4--6.4&0.16--0.42\\ b&578.8--580.6&-0.8--8.0&-4.0--3.2&0.13--0.63\\ c&579.0--581.6&-7.2--5.6&-5.6--4.0&0.13--0.64\\ d&578.0--579.8&-4.8--5.6&-4.0--6.4&0.11--0.64\\ e&579.2--582.0&-8.0--4.0&-5.6--4.8&0.14--0.63\\ \tableline \end{tabular} \end{table*} \subsection{Five-Parameter Frequency Model} The best-fit values for the five-parameter frequency model are given in Table~6. As for the two-parameter model, the uncertainties were estimated by marginalizing over all but the parameter of interest; the confidence regions containing 68\% of the probability are given in Table~7. \begin{table*}[t] \centering \begin{tabular}{cccccc} \multicolumn{6}{c}{\bfseries Table~6: Best-Fit Parameters for Five-Parameter Model}{\rule[-2mm]{0mm}{6mm}}\\ \tableline \tableline Burst&$\nu_1$ (Hz)&$\nu_2$ (Hz)&${\dot\nu}_1$ (Hz s$^{-1}$) &${\dot\nu}_2$ (Hz s$^{-1}$)&t$_{\rm break}$ (s)\\ \tableline a&581.2&578.0&-7.6&6.4&0.34\\ b&579.0&579.6& 8.0&3.2&0.34\\ c&578.6&578.0& 6.8&0.8&0.41\\ d&578.0&582.6&-0.4&-5.2&0.25\\ e&582.0&585.0&-8.0&-0.4&0.41\\ \tableline \end{tabular} \end{table*} \begin{table*}[ht] \centering \begin{tabular}{cccccc} \multicolumn{6}{c}{\bfseries Table~7: 68\% Confidence Regions for Five-Parameter Model}{\rule[-2mm]{0mm}{6mm}}\\ \tableline \tableline Burst&$\nu_1$ (Hz)&$\nu_2$ (Hz)&${\dot\nu}_1$ (Hz s$^{-1}$) &${\dot\nu}_2$ (Hz s$^{-1}$)&t$_{\rm break}$ (s)\\ \tableline a&579.6--581.8&577.4--581.6&-8.8--4.8&0.0--6.4&0.11--0.48\\ b&577.8--580.6&579.4--582.0&-5.6--8.8&-2.4--3.2&0.08--0.52\\ c&577.8--581.2&577.4--581.8&-6.4--8.8&-5.6--1.6&0.09--0.63\\ d&579.2--581.6&577.0--581.6&-9.6--6.4&-2.4--6.4&0.08--0.34\\ e&579.0--582.8&577.6--582.4&-8.8--6.4&-4.8--3.2&0.08--0.56\\ \tableline \end{tabular} \end{table*} \subsection{Summary of Frequency Models} \begin{table*}[t] \centering \begin{tabular}{ccccc} \multicolumn{5}{c}{\bfseries Table~8: Relative Log Likelihoods for Different Models}{\rule[-2mm]{0mm}{6mm}}\\ \tableline \tableline Burst&1-Param&2-Param&4-Param&5-Param\\ \tableline a&0.0&1.3&4.3&4.5\\ b&0.0&1.2&1.6&2.8\\ c&0.0&1.3&2.0&3.1\\ d&0.0&2.2&3.2&5.4\\ e&0.0&0.6&1.0&2.0\\ \tableline \end{tabular} \end{table*} The best-fit parameters and relative log likelihoods are listed in Table~8; as indicated above, $2\Delta\log{\cal L}\approx\Delta\chi^2$. From this table, it is clear that for all but burst four it is not necessary to use the five-parameter fit, and for bursts 2, 3, and 4 it is not necessary to use a model more complicated than the two-parameter model in which the frequency and frequency derivative are constant throughout the first 0.75 seconds. For burst~(d) by itself the five-parameter model is preferred at only the $2\sigma$ level compared to the four-parameter model, and for all five bursts combined the five-parameter model is preferred at less than the $1\sigma$ level relative to the four-parameter model. For all five bursts combined, the four-parameter model is preferred at less than the $1\sigma$ level compared to the two-parameter model, and the two-parameter model is preferred at less than the $2\sigma$ level compared to the one-parameter model. \section{DISCUSSION AND SUMMARY} What can be learned from this detailed characterization of the burst brightness oscillations in 4U~1636--536? The clearest impression left is that there are no simple statements about the frequency behavior that are true for all of the bursts. In two of the bursts, one can make an argument that the oscillation frequency is initially 1--2~Hz below the asymptotic frequency, and then rises. In this interpretation, the asymptotic frequency is extremely close to the spin frequency of the neutron star. This picture can be qualitatively explained by the idea that the burning layer lifts 20--50 meters during the burst and settles down gradually. However, the burst on 31 December 1996 does not follow this pattern. The frequency in the initial second is indeed lower than the maximum value attained, but the significance of this initial signal is low (Leahy power of 10). The maximum is followed by a clear decrease in the frequency over several seconds, with a total phase change equivalent to more than five complete circuits around the star. This happens during a time when the countrate decreases from approximately 2/3 of the maximum to approximately 1/3 of the maximum. The burst on 29 December 1996 has a very strong and significant brightness oscillation in its tail, which appears to level out to a constant frequency. However, near the peak of the light curve for this burst the oscillation frequency is higher than this asymptotic frequency, at a 3$\sigma$ significance level. Such a drop in frequency is not expected in the simplest version of the hypothesis that the frequency changes are caused by the rise of the burning layers. In this model, the highest frequency should be observed when the layers are fully coupled to the core of the star, which is expected to occur when the frequency has reached its asymptotic limit. Another constraint on the hypothesis that the asymptotic frequency equals the spin frequency (after correcting for orbital Doppler shifts) is that the variation in the observed asymptotic frequency must be consistent with the possible modulation due to the binary motion of the neutron star. From binary evolution theory (see, e.g., Lamb \& Melia 1987; Verbunt \& van den Heuvel 1995), an LMXB such as 4U~1636--536 with a 3.8~hr orbital period (van Paradijs et al.\ 1990) that contains a neutron star of mass $M_{\rm NS}$=1.4$M_\odot$ to 2.0$M_\odot$ has a companion star of mass $M_c\approx$0.4$M_\odot$. Assuming that the orbit is approximately circular, the orbital velocity of the neutron star is therefore 90--130~km~s$^{-1}$, implying a maximum frequency modulation of $\Delta\nu/\nu =4.3\times 10^{-4}$, or approximately 0.25~Hz if $\nu$=580~Hz. Therefore, the observed asymptotic frequency cannot be different by more than 0.5~Hz for two different bursts. The analysis of the 31 December 1996 burst reported in \S~2 indicates that eight seconds after the start of the burst the frequency is less than 579.0~Hz. The asymptotic frequency in the burst on 29 December 1996 is 581.43~Hz, so the frequency in the 31 December 1996 burst must rise by 2~Hz to reach a plausible spin frequency. It is difficult to reconcile this frequency behavior with what is expected in the simplest version of the rising burning layer hypothesis. One possibility is that the observed frequency changes are not simply indicative of the spin frequency of the burning layer, but also include a time-dependent change in the phase at which the photons emerge relative to the phase of the burning layer. This would be observationally indistinguishable from a pure frequency change, and would add an extra degree of freedom to the model. Even this, however, is subject to significant observational restrictions. To see this, consider the following observational trends, which have been observed in many bursts from several sources (see, e.g., Strohmayer et al.\ 1998 for a summary). In the remainder of this section we assume that all quantities (e.g., frequencies, times, and phases) are measured at infinity. (1) There are several bursts in which burst oscillations are seen for the entire burst, and do not disappear during the time of peak countrate. (2) Aside from an early phase in which there may be a frequency decrease, the frequency increases smoothly as the burst progresses. (3) The total phase lag of the oscillations compared with a hypothetical oscillation that has a constant frequency equal to the frequency in the burst tail is as much as $10\pi$. The total amount of energy in a burst is $\sim 10^{39}$~ergs. If expansion of a layer and angular momentum conservation are to explain the $\sim$0.3\%--1\% change in the observed angular frequency, then the layer must rise by a distance that is a fraction $\sim$0.2\%--0.5\% of the radius of the star, or 20 to 50 meters. The surface gravity of a neutron star is $\sim 2\times 10^{14}$~cm~s$^{-2}$, so the largest amount of mass that can be lifted to the required 20--50 meter height above the surface is $\sim$1--2$\times 10^{21}$~g. If most of the $\sim 10^{13}$~cm$^2$ surface area of the star is involved, this implies that the greatest column depth which could be lifted to the required height is roughly $10^8$~g~cm$^{-2}$, which is comparable to the expected $10^6-10^8$~g~cm$^{-2}$depth of ignition (see, e.g., Fushiki \& Lamb 1987; Brown \& Bildsten 1998). One may therefore distinguish two scenarios: (1)~the burning layer rotates with the core of the star at a constant spin frequency and the observed frequency shifts are caused by phase shifts induced by radiation transport through more slowly rotating layers, and (2)~the burning layer itself is lifted and rotates more slowly than the core of the star. We now treat these in order. Suppose for simplicity that the burning layer has an infinitesimal vertical extent, that it has some restricted azimuthal extent, and that it all rotates with the same angular frequency $\omega_{\rm burn}(t)$. The energy from this layer propagates upwards through the atmosphere, which in general may be composed of layers with different angular frequencies. Therefore, the phase of emergence of the radiation may differ from the phase of the burning layer at the time of the emission of the radiation. Under the rising burning layer hypothesis, it is expected that the angular frequency of higher layers is less than the angular frequency of lower layers ($d\omega/dh<0$). Hence, there is expected to be a lag $\phi_{\rm lag}>0$ between the phase of emergence and the phase of emission. This phase lag will, in general, have a time-dependence, as the scale height of the atmosphere and the angular frequency of different layers in the atmosphere changes throughout the burst. An observer at infinity will therefore see a net angular frequency of a hot spot that is equal to $\omega_{\rm burn}(t)-{\dot\phi}_{\rm lag}(t)$. Consider first a burning layer that rotates with the stellar core throughout the burst. Then $\omega_{\rm burn}(t)$ =const=$\omega_{\rm spin}$. If neither $\omega(h)$ nor the density or height of the envelope changes with time, then $\phi_{\rm lag}$ is a constant and the observed frequency is just $\omega_{\rm spin}$. Hence, in order to have an apparent frequency shift in this situation, the structure or angular velocity of the envelope must change with time. Now consider an envelope that does change with time. For us to observe a frequency less than $\omega_{\rm spin}$, it is necessary that ${\dot\phi}_{\rm lag}(t)>0$, so the characteristic phase of emergence of the radiation must lag the phase of the source of heat by a greater and greater amount with increasing time (the increase of this phase lag with time must itself decrease with time to produce the observed increase in frequency). But how is this possible? As the envelope settles down, the phase lag should {\it decrease}, because $d\omega(t)/dh<0$. But if the phase lag decreases, the observed frequency should be {\it higher} than the spin frequency. This is not seen in most bursts, and even in the burst on 29 December 1996 where there does appear to be a short period of spindown, the total phase lead implied by the spindown is much smaller than the total phase lag implied by the spinup near the beginning of the burst. Thus, the preceding set of assumptions is inconsistent with the data. This demonstrates that the observed frequency behavior is inconsistent with the source of heat (i.e., the burning layer) rotating at a constant frequency equal to $\omega_{\rm spin}$. Instead, the source of heat must change its frequency during the burst. To analyze this situation, let us now consider a burning layer with a finite thickness, so that the observed photons are a superposition of the photons from many infinitesimal layers such as discussed above. The observed frequency of oscillation is then a superposition of the frequencies due to the infinitesimal layers. Consider two of these infinitesimal slices, labeled 1 and 2, where slice 1 is higher than slice 2. Suppose that these slices are not coupled to each other. Then, by assumption, the angular frequency $\omega_{\rm burn,1}$ of slice 1 is less than the angular frequency $\omega_{\rm burn,2}$ of slice 2. In addition, because the photons from slice 2 have to travel through the same atmospheric layers as the photons from slice 1 in addition to the layers between 2 and 1, the phase lag $\phi_{\rm lag,1}$ of photons from slice 1 is expected to be less than the phase lag $\phi_{\rm lag,2}$ of photons from slice 2. Hence, as the atmospheric scale height decreases, it is expected that $\phi_{\rm lag,2}$ will decrease more rapidly than $\phi_{\rm lag,1}$ does, so that \begin{equation} {\dot\phi}_{\rm lag,2}<{\dot\phi}_{\rm lag,1}<0\; . \end{equation} Therefore, the difference between the angular frequency of the photons from slice 2 and the angular frequency of the photons from slice 1 is \begin{equation} \omega_{\rm burn,2}-\omega_{\rm burn,1}+ {\dot\phi}_{\rm lag,1}-{\dot\phi}_{\rm lag,2}> \omega_{\rm burn,2}-\omega_{\rm burn,1}\; . \end{equation} This means that the phases of emergence of radiation diverge rapidly from each other, which leads quickly to a low amplitude unless the heat source has a small vertical extent. The requirement that the amplitude be significant means that the total azimuthal phase subtended by the emergent radiation has to be much less than $2\pi$. The integrated phase lag relative to the stellar core is often $10\pi$ or larger, hence the average vertical extent of the heat source must be much less than 1/5 of the vertical distance from the original location of the heat source to its location during the burst. An alternative to having the vertical extent of the layer be small is that the burning layer may be tightly coupled to itself, so that its angular frequency is approximately constant over a significant vertical distance. To summarize, several conclusions may be drawn about the standard model for frequency changes during burst oscillations, which we take to be the picture that at least part of the burning layer is lifted and then settles gradually to the surface as the flux drops, producing an observed asymptotic frequency equal to the spin frequency of the neutron star Doppler-shifted by the orbital motion of the neutron star. (1)~The burning region itself (and not just overlaying optically thick layers) must be lifted by 20--50 meters from the surface, (2)~this region must remain decoupled from the rest of the star, presumed to be rotating at the original spin frequency, for several seconds, (3)~to produce the observed coherence of the brightness oscillations during the rise in frequency, the burning layer must either have a vertical extent much smaller than its height above the surface or be strongly coupled to itself to prevent relative azimuthal motion, and (4)~the existence of a frequency greater than the asymptotic frequency (as in the 29 December 1996 burst) implies that something other than differential rotation (e.g., variation in the phase lag) must account for at least part of the observed frequency change. The prolonged decrease in frequency in the tail of the 31 December 1996 burst is not straightforwardly fit into this picture. Despite these difficulties, the high stability (Strohmayer et al.\ 1998) and coherence (Markwardt \& Strohmayer 1999) of the brightness oscillations in the tails of bursts from sources such as 4U~1728--34 argue persuasively that the frequency in the tail of the bursts is close to either the fundamental or the first overtone of the neutron star spin frequency. Moreover, the general picture in which frequency changes are attributed to changes in the height of the emitting layer accounts approximately for the magnitude of the frequency change and explains why the frequency tends to rise near the beginning of the burst. However, in its current form it suffers from apparently serious problems. It is extremely important that there be a detailed investigation of, e.g., the coupling between differentially rotating layers, and that other ideas be explored so that the strengths and weaknesses of the rising layer model are put into sharper focus. \acknowledgements We thank Don Lamb and Fred Lamb for discussions about models of the frequency change, and Don Lamb, Dimitrios Psaltis, and Carlo Graziani for comments on a previous version of this paper. This research has made use of data obtained through the High Energy Astrophysics Science Archive Research Center Online Service, provided by the NASA/Goddard Space Flight Center. This work was supported in part by NASA grant NAG~5-2868, NASA AXAF contract SV~464006, and NASA ATP grant number NRA-98-03-ATP-028.
\section{\ Introduction} Unconventional anisotropic pairing is evidently realized in high temperature superconductors ($HTS$) - see review \cite{KulicRev}, and probably in some heavy Fermion superconductors ($HFS$) - see review \cite{Sauls}. There are good evidences for d-wave pairing in optimally doped $HTS$ oxides \cite {KulicRev} but the type of pairing in HFS is still unclear \cite{Sauls}. The effect of impurities on unconventional pairing is an important tool in analyzing the symmetry of the pairing amplitude, and is the subject of a number of experimental and theoretical works \cite{Pokrovskii}, \cite{Hotta}% . Most of calculations were done assuming an s-wave impurity scattering potential $u_{imp}({\bf p},{\bf p}^{\prime })=const$, and taking either the Born limit ($N(0)u_{imp}\ll 1$) or unitarity limit ($N(0)u_{imp}\gg 1$). Surprisingly, a number of experiments on the optimally $HTS$ oxides have shown that $d-wave$ pairing is quite robust, i.e. not very sensitive to various kinds of impurities and defects. For instance, the decrease of the critical temperature $T_{c}(\rho _{imp})$, with increasing residual resistivity, $\rho _{imp}$, is much smaller than the theory with the $s-wave$ impurity scattering predicts \cite{Pokrovskii}, \cite{Hotta}, \cite{Sun1}. A way out of this experimental and theoretical discrepancy of pair-breaking effects by impurities in $HTS$ oxides was proposed by the authors of Refs. \cite{KuOudo}, \cite{Haran}, who invoked a momentum dependent impurity scattering potential with an appreciable contribution in the d-channel. The microscopic theory in Ref. \cite{KuOudo} accounts for the renormalization of the impurity potential by strong correlations, which gives rise to a pronounced forward scattering peak, while backward scattering is suppressed, as first proposed in Ref. \cite{Kulic1} (see also \cite{KulicRev}). Application of this theory to impurity scattering \cite{Kulic1} shows that in addition to the contribution in the s-channel there is a significant contribution to the Born amplitude from the d-channel of the same magnitude , in particular for low (hole) doping concentration, $\delta <0.2$. As a consequence, the decrease of $T_{c}(\rho _{imp})$ with increasing $\rho _{imp}$ is much slower than the theory with exclusively s-wave impurity scattering predicts \cite{Hotta}. This renormalization effect explains the robustness of d-wave pairing in $HTS$ oxides. One may rise the question whether this robustness also holds far away from $% T_{c}$ and for very strong scattering potential, for instance in the unitarity limit. To answer this question we shall analyze a class of models by calculating the scattering $T$-matrix with an impurity potential that depends on the scattering angles. A related class of problems, which we study in section III, is related to the impurity scattering in the two-band model. Recently, several models for the pairing mechanism in HTS oxides based on two-band and multi-band models \cite{Hofmann}, \cite{Gajic}, \cite{Golubov}, \cite{Combescot} were suggested, and impurity effects studied in Born approximation. Magnetic and non-magnetic interband scattering can lead in this model to a lowering of the critical temperature and also to a relative sign change of the order parameters in different bands \cite{Golubov}. In section III we analyze the changes in the two-band model when going beyond the Born limit. It was shown that in the unitarity limit the Anderson theorem holds. In previous sections we studied a homogeneous superconductor with homogeneously distributed impurities. Selected inhomogeneous problems are studied in section IV, such as the bound states at an impurity, and the pinning energy at an impurity (defect) of singly- and double-quantized vortices. \section{Anisotropic scattering in anisotropic and homogeneous superconductors} In the following we analyze superconducting properties of anisotropic superconductors in the presence of momentum-dependent nonmagnetic impurity scattering by the quasiclassical equations of Eilenberger, Larkin-Ovchinnikov \cite{Eilenberger}, \cite{Larkin} ($ELO$ equations ). For a homogeneous distribution the quasiclassical Green's function matrix, $\hat{% g}({\bf p}_{F},{\bf R},\omega _{n})$, is independent of ${\bf R}$, and the quasiclassical equations read \begin{equation} \lbrack i\omega _{n}\hat{\tau}_{3}-\hat{\Delta}({\bf p}_{F},\omega _{n})-% \hat{\sigma}_{imp}({\bf p}_{F},\omega _{n}),\hat{g}({\bf p}_{F},\omega _{n})]=0 \label{elohom} \end{equation} \begin{equation} \hat{g}^{2}({\bf p}_{F},\omega _{n})=-\hat{1}. \label{norm} \end{equation} We assume weak-coupling superconductivity with $\hat{\Delta}({\bf p}% _{F})(=i\Delta ({\bf p}_{F})\hat{\tau}_{2})$, where $\Delta ({\bf p}_{F})$ is real. The $2\times 2$ matrices$\ \hat{\tau}_{0}\equiv \hat{1}$ and $\hat{% \tau}_{1,2,3}$ are Nambu-Gor'kov matrices. The effect of nonmagnetic impurities is described by the self-energy, $\hat{\sigma}_{imp}$, given in terms of the forward scattering part, $\hat{t}({\bf p}_{F},{\bf p}% _{F}^{\prime },\omega _{n})$, of the T-matrix \cite{Eilenberger}, \cite {Larkin}, \cite{Serene} \begin{equation} \hat{\sigma}_{imp}({\bf p}_{F},\omega _{n})=c\hat{t}({\bf p}_{F},{\bf p}% _{F},\omega _{n}), \label{sigma} \end{equation} where $c(\ll 1)$ is the impurity concentration. For simplicity we assume an isotropic Fermi surface but pairing and impurity scattering are angle-dependent, i.e. $\Delta ({\bf p}_{F})\equiv \Delta (% {\bf s})$, $\hat{g}({\bf p}_{F},\omega _{n})\equiv \hat{g}({\bf s},n)$ and $% \hat{t}({\bf p}_{F},{\bf p}_{F}^{\prime },\omega _{n})\equiv \hat{t}({\bf s,s% }^{\prime }{\bf ,}n)$ where ${\bf s=p}_{F}/p_{F}$.The T-matrix is the solution of the equation \begin{equation} \hat{t}({\bf s,s}^{\prime }{\bf ,}n)=u({\bf s,s}^{\prime })\hat{1}+N(0)\int d% {\bf s}^{\prime \prime }u({\bf s,s}^{\prime \prime })\hat{g}({\bf s}^{\prime \prime },n)\hat{t}({\bf s}^{\prime \prime }{\bf ,s}^{\prime }{\bf ,}n), \label{tmatrix} \end{equation} where for the $2D$ systems, which we consider here, one has $\int d{\bf % s\{..\}}\equiv \int {\bf \{..\}}d\theta /2\pi $. Since $\Delta ({\bf s})$ is real one has $\hat{g}=g_{2}\hat{\tau}_{2}+g_{3}\hat{\tau}_{3}$ and $\hat{t}$ is given by $\hat{t}=t_{0}\hat{\tau}_{0}+t_{1}\hat{\tau}_{1}+t_{2}\hat{\tau}% _{2}+t_{3}\hat{\tau}_{3}$. Because the unperturbed solution has the form ($\omega _{n}=\pi T(2n+1)$) \begin{equation} \hat{g}^{(0)}({\bf s},n)=-\frac{i\omega _{n}\hat{\tau}_{3}-i\Delta _{0}({\bf % s})\hat{\tau}_{2}}{\sqrt{\omega _{n}^{2}+\Delta _{0}^{2}({\bf s})}} \label{g0} \end{equation} then $\hat{g}({\bf s},n)$ is searched for in the form \begin{equation} \hat{g}({\bf s},n)=-\frac{i\tilde{\omega}_{n}({\bf s})\hat{\tau}_{3}-i\tilde{% \Delta}({\bf s,}\omega _{n})\hat{\tau}_{2}}{\sqrt{\tilde{\omega}_{n}^{2}(% {\bf s})+\tilde{\Delta}^{2}({\bf s,}\omega _{n})}}, \label{g} \end{equation} where \begin{equation} \tilde{\omega}_{n}({\bf s})=\omega _{n}({\bf s})+ic_{i}t_{3}({\bf s,s,}n) \label{omega} \end{equation} \begin{equation} \tilde{\Delta}({\bf s,}\omega _{n})=\Delta ({\bf s})-ic_{i}t_{2}({\bf s,s,}% n). \label{delta} \end{equation} The self-consistency equation for $\Delta ({\bf s})$ is given by \begin{equation} \Delta ({\bf s})=N(0)T\sum_{n}\int d{\bf s}^{\prime }V{\bf (s,s}^{\prime }% {\bf )}g_{2}({\bf s}^{\prime },n), \label{selfcons} \end{equation} where the pairing potential $V_{p}{\bf (s,s}^{\prime }{\bf )=}V_{p}Y({\bf s}% )Y({\bf s}^{\prime })$ is assumed in the factorized form with $<Y^{2}({\bf s}% )>_{{\bf s}}=1$. The latter implies that the order parameter has the form $% \Delta ({\bf s})=\Delta \cdot Y({\bf s})$. For convenience we define $\Gamma _{u}(\equiv c\gamma _{u})=c(\pi N(0))^{-1}$ and $v({\bf s,s}^{\prime })\equiv \pi N(0)u({\bf s,s}^{\prime })$. In what follows we consider the effects of anisotropic impurity scattering on the anisotropic pairing where $% <Y({\bf s})>_{{\bf s}}=0$. {\bf 1}. {\bf Anisotropic impurity scattering and nodeless anisotropic pairing} First, we consider the nodeless d-wave like pairing $\Delta ({\bf s})=\Delta \cdot Y({\bf s})$ which is characterized by $\langle Y({\bf s})\rangle _{% {\bf s}}=0$ and $Y^{2}({\bf s})=1$. This means that there is a finite gap everywhere on the Fermi surface, i.e. $\Delta ({\bf s})\neq 0$. It is interesting to mention, that besides the simplicity of this kind of pairing and its adequacy in some qualitative understanding of d-wave pairing it also appears to be a solution of the spin-bag model \cite{SWZ} for $HTS$ oxides. In this model the nodeless $d-wave$ like pairing is due to residual (longitudinal and transverse) spin fluctuations on the antiferromagnetic background, where the $AF$ order is distorted locally by hole doping and the spin-bag is formed around doped holes. The impurity scattering potential is assumed to have the form \begin{equation} v({\bf s,s}^{\prime })=v_{0}+v_{2}Y({\bf s})Y({\bf s}^{\prime }), \label{vsep} \end{equation} i.e. it contains an anisotropic contribution in the same channel as the unconventional pairing. The solution of $Eq.(\ref{tmatrix})$ for $t_{3}$ and $t_{2}$ is searched in the form \begin{equation} t_{3}({\bf s,s}^{\prime })=[\tilde{t}_{30}(n)+\tilde{t}_{32}(n)Y({\bf s})Y(% {\bf s}^{\prime })]g_{3}, \label{t3} \end{equation} \begin{equation} t_{2}({\bf s,s}^{\prime })=\tilde{t}_{2}(n)[g_{2}({\bf s},n)+g_{2}({\bf s}% ^{\prime },n)]. \label{t2} \end{equation} (Note, in this model $g_{2}({\bf s},n)=\tilde{g}_{2}(n)Y({\bf s})$, $% g_{2}^{2}({\bf s},n)=\tilde{g}_{2}^{2}(n)$, $g_{3}({\bf s},n)=g_{3}(n)$ and due to $Eq.(\ref{norm})$ one has $g_{3}^{2}({\bf s},n)+\tilde{g}_{2}^{2}(% {\bf s},n)=-1$.) The solution is given by \begin{equation} \tilde{t}_{30}(n)=\gamma _{u}v_{0}^{2}\frac{1+v_{2}^{2}}{% (1+v_{0}^{2})(1+v_{2}^{2})+(v_{0}-v_{2})^{2}\tilde{g}_{2}^{2}(n)}, \label{t30} \end{equation} \begin{equation} \tilde{t}_{2}(n)=\gamma _{u}v_{0}v_{2}\frac{1+v_{0}v_{2}}{% (1+v_{0}^{2})(1+v_{2}^{2})+(v_{0}-v_{2})^{2}\tilde{g}_{2}^{2}(n)}, \label{t2n} \end{equation} while $\tilde{t}_{32}(n)=\tilde{t}_{30}(n,v_{0}\leftrightarrow v_{2})$. \ Several interesting results comes out in this case. ({\bf a}) {\bf The} {\bf critical temperature} $T_{c}$ In the limit $T\rightarrow T_{c}$ $Eqs.(\ref{selfcons},\ref{t30}-\ref{t2n})$ become \begin{equation} \ln \frac{T_{c}}{T_{c0}}=\Psi (\frac{1}{2})-\Psi (\frac{1}{2}+\frac{\Gamma _{pb}}{2\pi T_{c}}), \label{tc} \end{equation} where the pair-breaking parameter $\Gamma _{pb}$ is given by ($\Gamma _{u}=c\pi /N(0)$) \begin{equation} \Gamma _{pb}=\Gamma _{u}\frac{(v_{0}-v_{2})^{2}}{(1+v_{0}^{2})(1+v_{2}^{2})}. \label{gamapb} \end{equation} Note that $T_{c}$ vanishes for $\Gamma _{pb}^{c}\approx 0.88T_{c0}$ and this pairing is in some respects similar to d-wave pairing. It is apparent from $% Eqs.(\ref{tc}-\ref{gamapb})$ that the pair-breaking effect of impurities is weakened in the presence of momentum-dependent scattering and it is even zero for $v_{0}=v_{2}$. The latter result has been previously derived in the Born approximation \cite{KuOudo}. For $v_{2}\approx v_{0}$ the slope $% dT_{c}/d\rho _{imp}$ can be very small even for appreciable values of $\rho _{imp}\sim \Gamma _{tr}=\Gamma _{u}(\bar{\sigma}_{0}+\bar{\sigma}_{2}),$ because in that case $\Gamma _{pb}\ll \Gamma _{tr}$ as indicate the experimental results of \cite{Sun1}. The parameters $\bar{\sigma}_{i}$ are given by \begin{equation} \bar{\sigma}_{i}=\frac{v_{i}^{2}}{1+v_{i}^{2}}\text{, \ \ }i=0,1,2... \label{sig} \end{equation} The resistivity, $\rho _{imp}$, and the reduction of $T_{c}$ due to impurity scattering, $T_{c}(\rho _{imp})$, depend on the classical transition rate, $% W({\bf s,s}^{\prime })=\Gamma _{u}\mid t_{N}({\bf s,s}^{\prime },n)\mid ^{2}$% in the normal state \cite{Rainer}. This transition rate comprises all the needed information on impurity scattering for either solving the normal state Boltzmann equation to determine $\rho _{imp}$ or the linearized gap equation $Eq.(\ref{selfcons})$ to determine $T_{c}$. For the latter purpose one needs linear (integral) equation for $g_{2}({\bf s},n)$ which reads \begin{equation} \mid \omega _{n}\mid g_{2}({\bf s},n)-\Delta ({\bf s})+\int d{\bf s}^{\prime }W({\bf s,s}^{\prime })[g_{2}({\bf s},n)-g_{2}({\bf s}^{\prime },n)]=0, \label{g2w} \end{equation} where the normal state t-matrix $t_{N}({\bf s,s}^{\prime },n)$ is the solution of the equation \begin{equation} t_{N}({\bf s,s}^{\prime },n)=v({\bf s,s}^{\prime })-i\pi sign(\omega _{n})\int d{\bf s}^{\prime \prime }v({\bf s,s}^{\prime \prime })t_{N}({\bf s}% ^{\prime \prime }{\bf ,s}^{\prime },n). \label{tN} \end{equation} Hence, measurements of $T_{c}(\rho _{imp})$ curve carry not enough informations on the microscopic scattering data, i.e. the scattering T-matrix. More such informations are contained in spectroscopic data on anisotropic superconductors at temperatures $T\ll T_{c}$, such as tunneling data or optical data at about gap frequency. {\bf (b) }The density of states{\bf , }$N(\omega )=N(0)% \mathop{\rm Im}% \int d{\bf s}$ $g_{3}({\bf s,}i\omega _{n}\rightarrow \omega -i\eta )$, depends in the presence of pair-breaking impurities significantly on the values $v_{0}$ and $v_{1}$. It is known \cite{Preosti} that in the case of s-wave scattering only ($v_{2}=0$) one has $N(\omega =0)\neq 0$ for $\Gamma _{u}v_{0}^{2}>\Delta $, and the highest value, $N(\omega =0)=N(0)/[0.5+0.5(1+(2\Delta /\Gamma )^{2})^{1/2}]^{1/2}$, where $\Gamma \equiv \Gamma _{u}\bar{\sigma}_{0}$, is reached in the unitarity limit. On the other hand, one obtains in the limiting case, $v_{0}=v_{2}$, a restoration of the gap, $N(\omega =0)=0$. Despite of the strong scattering limit $N(\omega )$ is BCS-like. {\bf 2. Isotropic impurity scattering and }${\bf d-wave}${\bf \ pairing} Let us study a two-dimensional superconductor with the pairing function $% \Delta ({\bf s})\equiv \Delta (\varphi )(=\Delta \cdot Y_{2}(\varphi ))\sim \cos 2\varphi $ - $d-wave$ pairing. Note this case seems to be more realistic for HTS oxides than the previous one, because the ''$\cos 2\varphi $'' pairing has nodes at the simply connected Fermi surface. We assume that the isotropic impurity potential depends on the transferred scattering angle \begin{equation} v(\varphi ,\varphi ^{\prime })=v_{0}+2v_{1}\cos (\varphi -\varphi ^{\prime })+2v_{2}\cos 2(\varphi -\varphi ^{\prime }), \label{vfi} \end{equation} where $v(\varphi ,\varphi ^{\prime })$ contains the pairing channel ($\sim Y_{2}(\varphi )Y_{2}(\varphi ^{\prime })$) too. (This problem but with $% v_{2}=0$ is studied in \cite{Haas} but there is an inappropriate sign in the $t_{2}$-matrix, which in fact corresponds to a magnetic impurity scattering). From $Eqs.(\ref{t30}-\ref{t2n})$ one obtains the pair-breaking parameter $\Gamma _{pb}$ \begin{equation} \Gamma _{pb}=\Gamma _{u}[\bar{\sigma}_{0}\frac{(1-\alpha )^{2}+\alpha ^{2}(1+v_{0}^{2})}{1+\alpha ^{2}v_{0}^{2}}+\bar{\sigma}_{1}], \label{gamafi} \end{equation} where $\alpha =v_{2}/v_{0}$ and $\bar{\sigma}_{i}$ are given by $Eq.(\ref {sig})$. In order to analyze $T_{c}(\rho _{imp})$ dependence, where the residual impurity resistivity $\rho _{imp}\sim \Gamma _{tr}$, we need the transport scattering $\Gamma _{tr}$ which is in this case given by \begin{equation} \Gamma _{tr}=\Gamma _{u}\{\bar{\sigma}_{0}^{2}+2\bar{\sigma}_{1}^{2}+2\bar{% \sigma}_{2}^{2}-2\bar{\sigma}_{0}[\bar{\sigma}_{1}(1+\frac{1}{v_{0}v_{1}})+% \bar{\sigma}_{2}(1+\frac{1}{v_{0}v_{2}})]\}\text{.} \label{gamatrans} \end{equation} If one wants to interpret depairing effects of impurities and robustness of pairing in HTS oxides in terms of the above results then the experiments \cite{Sun1} imply that the ratio, $\Gamma _{pb}/\Gamma _{tr}$, should be minimum. In the case when $v_{2}\ll v_{1}$ one obtains, $\Gamma _{pb}/\Gamma _{tr}=2$, in both, the Born and unitarity, $v_{0},v_{1}\rightarrow \infty $, limits. For, $v_{1}\ll v_{2},$ $\ $the pair-breaking parameter, $\Gamma _{pb} $, is minimized for $\alpha =1/2$ which gives, $\Gamma _{pb}/\Gamma _{tr}\approx 1/3$, in both limits. This means that the latter case is more appropriate candidate, than the case, $v_{2}\ll v_{1}$, for the qualitative explanation of robustness of d-wave pairing in $HTS$ oxides. \section{\bf Two-band model with nonmagnetic impurities} The interest in two(multi)-band models and in the impurity effects is renewed after the discovery of $HTS$ oxides \cite{Hofmann}, \cite{Gajic}, \cite{Combescot}, \cite{Golubov}, where various kinds of the intra- and inter-band pairing and impurity scattering are considered. By assuming only intra-band pairing $\Delta _{\alpha }$ ($\alpha =1,2$) the effect of the nonmagnetic impurities in the Born approximation is described by the equations ($n$ enumerates Matsubara frequencies) \begin{equation} \tilde{\omega}_{\alpha n}=\omega _{n}+\sum_{\beta }\frac{\tilde{\omega}% _{\beta n}}{2Q_{\beta n}}\gamma _{\alpha \beta } \label{omega2b} \end{equation} \begin{equation} \tilde{\Delta}_{\alpha n}=\Delta _{\alpha }+\sum_{\beta }\frac{\tilde{\Delta}% _{\beta n}}{2Q_{\beta n}}\gamma _{\alpha \beta } \label{delta2b} \end{equation} \begin{equation} \Delta _{\alpha }=\pi T\sum_{\beta ,n}^{-\omega _{D}<\omega _{n}<\omega _{D}}\lambda _{\alpha \beta }\frac{\tilde{\Delta}_{\beta n}}{Q_{\beta n}}, \label{self2b} \end{equation} where $Q_{\alpha n}=\sqrt{\tilde{\omega}_{\alpha n}^{2}+\tilde{\Delta}% _{\alpha n}^{2}}$, $\gamma _{\alpha \beta }=u_{\alpha \beta }^{2}N_{\beta }(0)$ for the nonmagnetic impurity scattering and $\lambda _{\alpha \beta }=V_{\alpha \beta }^{p}N_{\beta }(0)$ are corresponding coupling constants. In Ref. \cite{Golubov} are considered various possibilities for the suppression of the critical temperature, as well as the relative sign of $% \Delta _{1}$ and $\Delta _{2}$, in the Born limit for nonzero values of $% \lambda _{\alpha \beta }$ and $\gamma _{\alpha \beta }$. We note some interesting conclusions obtained in Born approximation which shall be compared with the results obtained in the unitarity limit: $(i)$ the diagonal scattering rate $\gamma _{11}$ and $\gamma _{22}$ disappear from the linearized $Eq.(\ref{self2b})$ for $T_{c}$; $(ii)$ in the case $\lambda _{11}\neq 0,\lambda _{22}=\lambda _{12}=\lambda _{21}=0$ the depression of $% T_{c}$ is given by $\delta T_{c}/T_{c}=-\pi \gamma _{12}/8T_{c}$; $(ii)$ for $\lambda _{11}=\lambda _{22}\neq 0$ and $\lambda _{12}=\lambda _{21}=\lambda _{\perp }<0$ one has $sign(\Delta _{1}/\Delta _{2})=-1$ and $\delta T_{c}/T_{c}=-\pi (\gamma _{12}+\gamma _{21})/8T_{c}$, while the sign of $% \Delta _{1}$ and $\Delta _{2}$ is unchanged by the impurities. The $t$-matrix equation in the two-band model has the form ($\alpha ,\beta ,\gamma =1,2$) ( we consider a rather small impurity concentration and neglect an interband hybridization) \begin{equation} {\bf \hat{t}}(n)={\bf \hat{u}}+\sum_{\gamma }{\bf \hat{u}N}(0){\bf \hat{g}}% (n){\bf \hat{t}}(n), \label{t2b} \end{equation} where ${\bf \hat{t}}(n)=\sum_{i=0}^{3}{\bf t}_{i}{\bf \otimes }\hat{\tau}% _{i} $, ${\bf \hat{g}}(n)={\bf g}_{3}{\bf \otimes }\hat{\tau}_{3}+{\bf g}_{2}% {\bf \otimes }\hat{\tau}_{2}$ and ${\bf \otimes }$ is the direct product of matrices in the band space (bold) and in the Nambu space (hat). ${\bf g}_{2},% {\bf g}_{3}$ and ${\bf N}(0)$ are diagonal matrices in the band space. In the case of nonmagnetic impurities one has ${\bf \hat{u}}^{N}={\bf u}^{N}% {\bf \otimes }\hat{\tau}_{0}$ and since ${\bf g}_{1}=0$ one has \[ {\bf t}_{0}^{N}(n)={\bf u}^{N}+{\bf u}^{N}{\bf N}(0)[{\bf g}_{3}(n){\bf t}% _{3}^{N}(n)+{\bf g}_{3}(n){\bf t}_{3}^{N}(n)] \] \[ {\bf t}_{1}^{N}(n)={\bf u}^{N}{\bf N}(0)[-i{\bf g}_{3}(n){\bf t}_{2}^{N}(n)+i% {\bf g}_{2}(n){\bf t}_{3}^{N}(n)] \] \[ {\bf t}_{2}^{N}(n)={\bf u}^{N}{\bf N}(0)[i{\bf g}_{3}(n){\bf t}_{1}^{N}(n)+% {\bf g}_{2}(n){\bf t}_{0}^{N}(n)] \] \begin{equation} {\bf t}_{3}^{N}(n)={\bf u}^{N}{\bf N}(0)[{\bf g}_{3}(n){\bf t}_{0}^{N}(n)-i% {\bf g}_{2}(n){\bf t}_{1}^{N}(n)] \label{ti2b} \end{equation} Let us consider for simplicity the case when $u_{11}^{N},u_{22}^{N}=0$ but interband scattering, $u_{12}^{N}=u_{21}^{N}=u$ $\neq 0$, and introduce three parameters \begin{equation} \sigma =\frac{\pi ^{2}N_{1}(0)N_{2}(0)u^{2}}{1+\pi ^{2}N_{1}(0)N_{2}(0)u^{2}} \label{sig2b} \end{equation} and \begin{equation} \Gamma _{i}=\frac{c\cdot \sigma }{\pi N_{i}(0)},\text{ \ }i=1,2. \label{Gama2b} \end{equation} After some straightforward calculations one obtains the renormalized frequencies, $\tilde{\omega}_{in}$, and order parameters $\tilde{\Delta}% _{in} $ \begin{equation} \tilde{\omega}_{1n}=\omega _{n}+\Gamma _{1}\frac{(\sigma -1)(\tilde{\omega}% _{1n}^{2}+\tilde{\Delta}_{1n}^{2})\tilde{\omega}_{2n}-\sigma \tilde{\omega}% _{1n}\sqrt{\tilde{\omega}_{1n}^{2}+\tilde{\Delta}_{1n}^{2}}\sqrt{\tilde{% \omega}_{2n}^{2}+\tilde{\Delta}_{2n}^{2}}}{\det 1} \label{omega1} \end{equation} \begin{equation} \tilde{\Delta}_{1n}=\Delta _{1}+\Gamma _{1}\frac{(\sigma -1)(\tilde{\omega}% _{1n}^{2}+\tilde{\Delta}_{1n}^{2})\tilde{\Delta}_{2n}-\sigma \tilde{\Delta}% _{1n}\sqrt{\tilde{\omega}_{1n}^{2}+\tilde{\Delta}_{1n}^{2}}\sqrt{\tilde{% \omega}_{2n}^{2}+\tilde{\Delta}_{2n}^{2}}}{\det 1}, \label{delta1} \end{equation} where \[ \det 1=2(\sigma -1)\sigma \sqrt{\tilde{\omega}_{1n}^{2}+\tilde{\Delta}% _{1n}^{2}}(\tilde{\Delta}_{1n}\tilde{\Delta}_{2n}+\tilde{\omega}_{1n}\tilde{% \omega}_{2n})- \] \begin{equation} -[2(\sigma -1)\sigma +1](\tilde{\omega}_{1n}^{2}+\tilde{\Delta}_{1n}^{2})% \sqrt{\tilde{\omega}_{2n}^{2}+\tilde{\Delta}_{2n}^{2}}. \label{det1} \end{equation} The solution for the second band is obtained from $Eqs.(\ref{omega1}-\ref {det1})$ by replacing $1\Longleftrightarrow 2$. In the Born limit one gets \begin{equation} \tilde{\omega}_{1n}=\omega _{n}+\Gamma _{1}\frac{\sigma \tilde{\omega}_{2n}}{% \sqrt{\tilde{\omega}_{2n}^{2}+\tilde{\Delta}_{2n}^{2}}} \label{omegab} \end{equation} \begin{equation} \tilde{\Delta}_{1n}=\Delta _{1}+\Gamma _{1}\frac{\sigma \tilde{\Delta}_{2n}}{% \sqrt{\tilde{\omega}_{2n}^{2}+\tilde{\Delta}_{2n}^{2}}}, \label{deltab} \end{equation} i.e. the interband scattering mixes both bands. In the unitarity limit $% \sigma \rightarrow 1$ ($u\rightarrow \infty $) the bands are decoupled, i.e. \begin{equation} \tilde{\omega}_{\alpha n}=\omega _{n}+\Gamma _{\alpha }\frac{\tilde{\omega}% _{\alpha n}}{\sqrt{\tilde{\omega}_{\alpha n}^{2}+\tilde{\Delta}_{\alpha n}^{2}}} \label{omegauni} \end{equation} \begin{equation} \tilde{\Delta}_{\alpha n}=\Delta _{\alpha n}+\Gamma _{\alpha }\frac{\tilde{% \Delta}_{\alpha n}}{\sqrt{\tilde{\omega}_{\alpha n}^{2}+\tilde{\Delta}% _{\alpha n}^{2}}}. \label{deltauni} \end{equation} So, in this case the Anderson theorem is restored, i,e, the thermodynamic properties are impurity independent. The latter result can be generalize to the case \begin{equation} {\bf u}^{N}=\left( \begin{array}{cc} \alpha u & u \\ u & u_{22} \end{array} \right) . \label{un} \end{equation} For $u\rightarrow \infty $ but $\alpha $ and $u_{22}$ finite, $\alpha $ and $% u_{22}$ drop out from equations and the bands are decoupled with $\tilde{% \omega}_{\alpha n}$ and $\tilde{\Delta}_{\alpha n}$ given by $Eqs.(\ref {omegauni}-\ref{deltauni})$. At $T_{c}$ one has \[ \tilde{\omega}_{1n}=\omega _{n}+\Gamma _{1}sign(\omega _{n}) \] \begin{equation} \tilde{\Delta}_{1n}=\Delta _{1}+\Gamma _{1}(\frac{\tilde{\Delta}_{1n}}{\mid \tilde{\omega}_{1n}\mid }+\frac{(1-\sigma )\tilde{\Delta}_{2n}}{\mid \tilde{% \omega}_{2n}\mid }). \label{omdel} \end{equation} From $Eq.(\ref{omdel})$ it is seen that in the unitarity limit, $\sigma \rightarrow 1$, the renormalized order parameters are decoupled and $T_{c}$ is unrenormalized. For $\sigma <1$ it can be easily shown that $T_{c}$ is reduced. \section{Small anisotropic defect in anisotropic superconductors} In what follows we consider the effect of a single impurity (small defect) with small scattering length $a$, which is supposed to be much smaller than the superconducting coherence length, $\mid a\mid \ll \xi _{0}$. Hence, the impurity can be considered as a localized perturbation, but with negligible renormalization of $\hat{\Delta}({\bf p}_{F},{\bf R})$, giving rise to the quasiclassic equations \cite{Thuneberg}, \cite{Thuneberg2} \[ \lbrack (i\omega _{n}+e{\bf v}_{F}\cdot {\bf A(R)})\hat{\tau}_{3}-\hat{\Delta% }({\bf p}_{F},{\bf R}),\delta \hat{g}({\bf p}_{F},{\bf R},\omega _{n})]+i% {\bf v}_{F}\nabla _{{\bf R}}\delta \hat{g}({\bf p}_{F},{\bf R},\omega _{n})= \] \begin{equation} =[\hat{t}({\bf p}_{F},{\bf p}_{F},\omega _{n}),\hat{g}_{imt}({\bf p}_{F},% {\bf R},\omega _{n})]\delta ({\bf R-R}_{imp}). \label{eloinhom} \end{equation} Here, $\delta \hat{g}({\bf p}_{F},{\bf R},\omega _{n})=\hat{g}({\bf p}_{F},% {\bf R},\omega _{n})-\hat{g}_{imt}({\bf p}_{F},{\bf R},\omega _{n})$. The extra term proportional to $\delta ({\bf R-R}_{imp})$ describes a jump in $% \hat{g}({\bf p}_{F},{\bf R},\omega _{n})$ at the site ${\bf R}_{imp}$ of the impurity (defect), while the intermediate Green's function $\hat{g}_{imt}(% {\bf p}_{F},{\bf R},\omega _{n})$ describes the quasiclassic motion in the absence of impurity (defect) and it is the solution of $Eq.(\ref{eloinhom})$ by putting the right-side to zero. $\hat{g}_{imt}({\bf p}_{F},{\bf R},\omega _{n})$ is normalized according to $Eq.(\ref{norm})$. The t-matrix is the solution of $Eq.(\ref{tmatrix})$ where $\hat{g}({\bf p}_{F},\omega _{n})$ is replaced by $\hat{g}_{imt}({\bf p}_{F},{\bf R=R}_{imp},\omega _{n})$. The change of the superconducting free-energy in the presence of a single impurity (defect) is given by \cite{Thuneberg}, \cite{Thuneberg2} \begin{equation} \delta F({\bf R}_{imp})=N(0)T\sum_{n}\int_{0}^{1}d\lambda \int \frac{d^{2}% \hat{k}_{F}}{4\pi }\int d^{3}RTr[\delta \hat{g}({\bf p}_{F},{\bf R},\omega _{n})\hat{\Delta}_{b}({\bf p}_{F},{\bf R})], \label{deltafg} \end{equation} where $\hat{\Delta}_{b}({\bf p}_{F},{\bf R})$ and the vector potential ${\bf % A}_{b}({\bf R})$ are calculated in the absence of the impurity. The Green's function, $\delta \hat{g}({\bf p}_{F},{\bf R},\omega _{n})$, must be evaluated for an order parameter $\hat{\Delta}({\bf p}_{F},{\bf R})=\lambda \hat{\Delta}_{b}({\bf p}_{F},{\bf R})$. In the following we study the consequences of anisotropic impurity scattering for three selected examples of inhomogeneous anisotropic superconductors. {\bf 1}. {\bf Bound states due to the anisotropic impurity} Let us consider the local change of superconductivity in the presence of a single anisotropic impurity with the potential $v({\bf s,s}^{\prime })$ given by $Eq.(\ref{vsep})$ and analyze the impurity-induced quasiparticle bound state{\bf \ }and the change in the free-energy $\delta F({\bf R}% _{imp}) ${\bf . }By assuming that{\bf \ }$2\pi \bar{\sigma}_{i}\ll E_{F}/\Delta _{0}$ , where $i=0,2$ and $\bar{\sigma}% _{i}=v_{i}^{2}/(1+v_{i}^{2})$, the t-matrix is given by the same expression as $Eqs.(\ref{t30}-\ref{t2n})$, but with $\tilde{g}_{2}(n)$ is replaced by $% g_{2}^{(0)}(n)$. The bound state energy $\omega _{B,anis}<\Delta _{0}$, which is due to the pair-breaking impurity effects, can be obtained as a pole of the $t$-matrix which gives \begin{equation} \omega _{B,anis}=\Delta _{0}\sqrt{1-\bar{\sigma}_{pb},} \label{bound} \end{equation} where \begin{equation} \bar{\sigma}_{pb}=\bar{\sigma}_{0}\bar{\sigma}_{2}\frac{(v_{0}-v_{2})^{2}}{% v_{0}^{2}v_{2}^{2}}. \label{sigpb} \end{equation} In the unitarity limit for both channels, i.e. $v_{0}\gg 1$, $v_{2}\gg 1$ but $v_{2}/v_{0}$ finite, one has $\omega _{B,anis}\rightarrow \Delta _{0}$ contrary to the unitarity limit for the s-wave scattering ($v_{0}\gg 1$, $% v_{2}=0$) where $\omega _{B,iso}\rightarrow 0$. However, the zero-energy bound state $\omega _{B,anis}\rightarrow 0$ appears when $v_{0}v_{2}=-1$, i.e. if one channel is in the unitarity limit the other one must be in the Born limit. Due to the bound state there is a change (increase) of the free-energy $% \delta F({\bf R}_{imp})\equiv \delta F_{imp}$. By solving $Eq.(\ref{eloinhom}% )$ with $\hat{g}_{imt}({\bf p}_{F},{\bf R},\omega _{n})$ given by $Eq.(\ref {g0})$ and $\hat{t}$ given by $Eqs.\ref{t3}-\ref{t2n})$ one gets $\delta F_{imp}$ from $Eq.(\ref{deltafg})$ \[ \delta F_{imp}=T\sum_{n}\int_{0}^{1}d\lambda \bar{\sigma}_{pb}\frac{\lambda \Delta _{0}^{2}\omega _{n}^{2}}{[\omega _{n}^{2}+\lambda ^{2}\Delta _{0}^{2}][\omega _{n}^{2}+\omega _{B,anis}^{2}]}= \] \begin{equation} =2T\ln \frac{\cosh (\Delta _{0}/2T)}{\cosh [(1-\sigma _{pb})^{1/2}\Delta _{0}/2T]}, \label{deltafimp} \end{equation} where $\bar{\sigma}_{pb}$ is given in $Eq.(\ref{sigpb})$. It is seen that there is a loss in the condensation energy,$\delta F({\bf R}_{imp})>0$, which is related to the pair-breaking effect of impurity. For $v_{0}=v_{2}$ such an impurity does not affect superconductivity and $\delta F({\bf R}% _{imp})=0$. The obtained results tell us that in for angle-dependent impurity scattering even a strong impurity potential may have very weak effect on $T_{c}$, the bound state, and the free-energy of anisotropic and unconventional pairing. In that case the anisotropic pairing is robust in the presence of impurities. {\bf 2. Pinning of single-vortex by a small anisotropic defect} Because in $HTS$ oxides strong correlations give rise to strong momentum-dependent charge scattering processes it is interesting to analyze the elementary-flux-pinning potential of a small defect by using the approach of Thuneberg et al., \cite{Thuneberg}, \cite{Thuneberg2}, who showed that in s-wave superconductors the pinning energy of a small defect ($% a\ll \xi _{0}$) is dominated by scattering processes at the defect. It is proportional to the product of the scattering cross section and coherence length ($\propto a^{2}\xi _{0}$), instead of (naively believed) $a^{3}$. The case of anisotropic pairing with s-wave impurities and near $T_{c}$ was recently studied in \cite{Friesen}. In what follows we study the effect of scattering anisotropy on the pinning energy of a small defect in an anisotropic superconductor at any temperature below $T_{c}$. We use the model potential given in $Eq.(\ref{vsep})$ and assume that the vortex is placed at the defect. In order to calculate the elementary flux-pinning energy one has to solve the quasiclassical equations for various ballistic trajectories with ${\bf R}$-dependent the vector potential, ${\bf A(R)}$, and order parameter, \begin{equation} \Delta _{b}({\bf p}_{F},{\bf R})=\mid \Delta ({\bf p}_{F},{\bf R})\mid e^{i\theta }Y(\theta ) \label{deltaR} \end{equation} In the gauge where $\theta $ is the angle with respect to the X-axis then $% {\bf A(R)}$ has no radial component. The solution of $Eq.(\ref{eloinhom})$ requires for a realistic vortex numerical calculations. For a qualitative discussion we will adopt a simplified vortex model \cite{Thuneberg}, \cite {Thuneberg2} which neglects the suppression of the order parameter in the vortex core and sets $\mid \Delta _{b}({\bf p}_{F},{\bf R})\mid =\Delta _{0}(% {\bf p}_{F})$, i.e. independent of ${\bf R}$. Hence, the order parameter along a trajectory passing through the vortex center has constant magnitude but its phase changes abruptly by $\pi $ when going through the vortex core. This ''zero-core model'' gives the right order of magnitude of the pinning energy, $\delta F_{pin}({\bf R}_{imp})$, when compared with the numerical calculations \cite{Thuneberg2}. In order to calculate $\delta F_{pin}({\bf R}% _{imp})$ two quantities are needed: $\hat{g}_{imt}({\bf p}_{F},{\bf R=R}% _{imp},\omega _{n})\equiv \hat{g}_{v}({\bf p}_{F},{\bf R=R}_{imp},\omega _{n})$ in the presence of the zero-core vortex and the impurity $t$-matrix calculated with the Green's function, $\hat{g}_{v}({\bf p}_{F},{\bf R=R}% _{imp},\omega _{n})$, of the zero-core model. The solution is straightforward \cite{Thuneberg2} and gives \begin{equation} \hat{g}_{v}({\bf p}_{F},{\bf R=R}_{imp},\omega _{n})=\frac{1}{\omega _{n}}% [(-\Delta _{2}\hat{\tau}_{1}+\Delta _{1}\hat{\tau}_{2})Y(\theta )+(-i\alpha _{n})\hat{\tau}_{3}], \label{gv} \end{equation} and \begin{equation} \hat{t}({\bf p}_{F},{\bf p}_{F},\omega _{n})=t_{3}\hat{\tau}_{3}=-i\gamma _{u}\alpha _{n}\omega _{n}[\frac{\bar{\sigma}_{0}}{\omega _{n}^{2}+\bar{% \sigma}_{0}\Delta _{0}^{2}}+\frac{\bar{\sigma}_{2}}{\omega _{n}^{2}+\bar{% \sigma}_{2}\Delta _{0}^{2}}]\hat{\tau}_{3}. \label{tv} \end{equation} Here, $\alpha _{n}=\sqrt{\omega _{n}^{2}+\Delta _{0}^{2}}$. $Eq.(\ref {eloinhom})$ can be solved by the Fourier (or Laplace) transform which gives the expression for the pinning free-energy \begin{equation} \delta F_{pin}=\delta F_{pin}^{(stiff)}(\bar{\sigma}_{0},\bar{\sigma}% _{2})+\delta F_{pin}^{(pb)}(\sigma _{pb}) \label{deltafpin} \end{equation} \begin{equation} \delta F_{pin}^{(stiff)}=-2T\ln \{\cosh \frac{\sqrt{\bar{\sigma}_{0}}\Delta _{0}}{2T}\cdot \cosh \frac{\sqrt{\bar{\sigma}_{2}}\Delta _{0}}{2T}\}, \label{stif} \end{equation} \begin{equation} \delta F_{pin}^{(imp)}=-2T\ln \frac{\cosh (\Delta _{0}/2T)}{\cosh [(1-\bar{% \sigma}_{pb})^{1/2}\Delta _{0}/2T]}. \label{pinimp} \end{equation} $Eqs.(\ref{deltafpin}-\ref{pinimp})$ imply that, $\delta F_{pin}<0$, and the vortex is attracted (pinned) by the defect. A comparison of $Eq.(\ref {deltafpin})$ with the corresponding results for $s-wave$ superconductors with an s-wave scattering potential shows, that in the former case two additional terms are present. The first one, depending on $\bar{\sigma}_{2}$% , appears also in s-wave superconductors with anisotropic scattering accounted for. In fact $\delta F_{pin}^{(stiff)}$ describes the reduction of the superconducting stiffness in the presence of impurities. For instance near $T_{c}$ $Eq.(\ref{stif})$ gives \begin{equation} \delta F_{pin}^{(stiff)}=-(\bar{\sigma}_{0}+\bar{\sigma}_{2})\cdot \frac{% \Delta _{0}^{2}(T)}{4T_{c}}\approx -7.6\frac{\bar{\sigma}_{0}+\bar{\sigma}% _{2}}{v_{F}^{2}}\xi _{0}^{2}\cdot T_{c}\Delta _{0}^{2}(T), \label{stiff} \end{equation} and $\delta F_{pin}^{(stiff)}$ is proportional to the total scattering amplitude $\bar{\sigma}_{0}+\bar{\sigma}_{2}$. For vortex far away from the impurity there is loss in the condensation energy $\delta F_{pin}^{(pb)}(% \bar{\sigma}_{pb})$ due to pair-breaking effect of the impurity, i.e. $% \delta F_{pin}^{(pb)}(\bar{\sigma}_{pb})=-\delta F_{imp}(\bar{\sigma}_{pb})$ where $\delta F_{imp}(\bar{\sigma}_{pb})$ is given by $Eq.(\ref{deltafimp})$% . Therefore this part enters in $Eq.(\ref{deltafpin})$ with the negative sign, thus increasing the pinning energy when vortex is sitting on the defect and stabilizing it additionally. Near $T_{c}$ one has \begin{equation} \delta F_{pin}^{(pb)}(\sigma _{eff})=-\bar{\sigma}_{pb}\cdot \frac{\Delta _{0}^{2}(T)}{4T_{c}}, \label{cond} \end{equation} For the s-wave scattering only, $v_{2}=0$, one has $\bar{\sigma}_{2}=0$, $% \bar{\sigma}_{pb}=\bar{\sigma}_{0}$, and the pair-breaking effect is maximal while the condensation energy is gained maximally for vortex sitting on the defect. However, for, $v_{2}=v_{0}$ the pair-breaking of impurity is absent $% \bar{\sigma}_{pb}=0$ and $\delta F_{pin}^{(pb)}=0$, i.e. in this case the pinning by the small defect is similar to that in s-wave superconductors. The physical picture of the vortex pinning by small defect given above is based on the known results based on the microscopic derivation of the Ginzburg-Landau equations in the presence of impurities. An explanation based on the quasiclassical approach is given in \cite{Thuneberg}, \cite {Thuneberg2} and we briefly discuss it in order to develop an intuition for the case of a double-vortex pinning, which is studied below. Because the order parameter changes its phase by $\pi $ along the trajectories across the vortex core it leads to the phase change of $\hat{g}_{v}({\bf p}_{F},% {\bf R},\omega _{n})(\equiv \hat{g}_{imt}({\bf p}_{F},{\bf R},\omega _{n}))$ on the distance $\xi _{0}$, thus causing a cost in the condensation energy, i.e. the maximal increase of the free-energy. Note, the function $\hat{g}% _{v}({\bf p}_{F},{\bf R},\omega _{n})$ describes the quasiclassical motion of particles (or pairs) along trajectories across the vortex core where the maximal phase change ($\pi $) occurs. In the presence of defect the motion of particles is described by the function $\hat{g}({\bf p}_{F},{\bf R}% ,\omega _{n})$ which contains scattering of particles to new directions where the phase change (mismatch) is less than $\pi $ and it costs less condensation energy. Therefore the vortex is attracted to the defect because scattering helps superconductivity to sustain abrupt changes in the order parameter. The latter explains the contribution $\delta F_{pin}^{(stiff)}$, while in the anisotropic superconductors, due to pair-breaking effects of impurities, there is a gain of condensation energy $-\delta F_{imp}(\sigma _{eff})$ for vortex sitting on the defect. {\bf 3. Pinning of double-vortex by small defect} We extend the calculations in \cite{Thuneberg2} to the pinning of multiply-quantized vortices on small defects. First, a s-wave superconductor is considered and we put the question - is it possible to pin the double-flux-vortex ($\Phi =2\Phi _{0}$) by the small defect, which is for simplicity characterized by the parameter $\bar{\sigma}_{0}$ for s-wave scattering only? The ''zero-core model'' is assumed again. In that case the order parameter can be parametrized in the form \begin{equation} \Delta _{b}({\bf p}_{F},{\bf R})=\mid \Delta ({\bf p}_{F},{\bf R})\mid e^{2i\theta }. \label{delta2R} \end{equation} For particle motion across the double-vortex core the order parameter does not change phase and in that case the solutions for $\hat{g}_{imt}({\bf p}% _{F},{\bf R=R}_{imp},\omega _{n})(\equiv \hat{g}_{2v}({\bf p}_{F},{\bf R=R}% _{imp},\omega _{n}))$ and for $\hat{t}({\bf p}_{F},{\bf p}_{F},\omega _{n})$ are given by \begin{equation} \hat{g}_{2v}({\bf p}_{F},{\bf R=R}_{imp},\omega _{n})=\frac{i}{\alpha _{n}}% [\Delta _{1}\hat{\tau}_{1}+\Delta _{2}\hat{\tau}_{2}+(-\omega _{n})\hat{\tau}% _{3}] \label{g2v} \end{equation} and \begin{equation} \hat{t}({\bf p}_{F},{\bf p}_{F},\omega _{n})=t_{3}\hat{\tau}_{3}=-i\omega _{n}\alpha _{n}\gamma _{u}\frac{\bar{\sigma}_{0}}{\omega _{n}^{2}+\tilde{% \sigma}_{0}\Delta _{0}^{2}}\hat{\tau}_{3}, \label{t2v} \end{equation} where $\tilde{\sigma}_{0}=\bar{\sigma}_{0}/v_{0}^{2}$. Then by solving $Eq.(% \ref{eloinhom})$ and by using $Eq.(\ref{deltafg})$ and this solutions one obtains the pinning energy of the double-vortex within the zero-core model \begin{equation} \delta F_{2v,pin}=2T\sum_{n}\int_{0}^{1}d\lambda \frac{\lambda \Delta _{0}^{2}\omega _{n}^{2}\bar{\sigma}_{0}}{\alpha _{n}^{2}[\omega _{n}^{2}+% \tilde{\sigma}_{0}\lambda ^{2}\Delta _{0}^{2}]}>0. \label{f2vpi} \end{equation} The main conclusion coming out from $Eq.(\ref{f2vpi})$ is that because $% \delta F_{pin}>0$ the double-vortex in s-wave superconductors is repelled from the defect - i.e. the zero-core double-vortex can not be pinned. Contrary to the single-vortex, where the defect scatters particles to new directions where the phase change is smaller, in the case of double-vortex the particles are scattered to directions where the phase change is larger. However, it might be that the above obtained results are an artefact of the ''zero-core model'', where there is no suppression of the superconducting order due to the vortex core, and numerical calculations are required for a realistic double-vortex structure \cite{Endres}. In the case of an unconventional pairing, like that in Section $II.1$, the order parameter is given by $\Delta _{b}({\bf p}_{F},{\bf R})=\mid \Delta (% {\bf p}_{F},{\bf R})\mid \exp (2i\theta )Y(\theta )$ and the t-matrix contains also terms $t_{1},t_{2}\neq 0$ leading to a decrease of the jump $[% \hat{t}({\bf p}_{F},{\bf p}_{F},\omega _{n}),\hat{g}_{imt}({\bf p}_{F},{\bf R% },\omega _{n})]$ in $Eq.(\ref{eloinhom})$. In this case the pinning energy contains the additional term (gain in energy) due to the pair-breaking effect of the impurity, $-\delta F_{imp}$, i.e. $\delta F_{pin}=\delta F_{2v,pin}-\delta F_{imp}$. Since $\delta F_{2v,pin}$ is less positive (repulsive) than for a s-wave superconductor in $Eq.(\ref{f2vpi})$, and because $-\delta F_{imp}<0$ one can happen that $\delta F_{pin}<0$\ and even the zero-core double-vortex can be pinned by the defect. A realistic calculation of $\delta F_{pin}$ for anisotropic superconductors with anisotropic scattering of single and double-vortex will be discussed elsewhere \cite{Endres}. In conclusion the anisotropic impurity scattering gives rise to new qualitative effects in unconventional and anisotropic superconductors, where for instance it ''screens'' the strength of the scattering in some quantities (like $T_{c}$ - robustness of pairing, bound states, pinning, etc.) even in the unitarity limit. It seems that this situation is partly realized in $HTS$ oxides where the d-wave pairing is robust in the presence of even very strong impurity scattering. In two-band models nonmagnetic impurities do not affect thermodynamic properties of s-wave superconductors in the unitarity limit for the interband scattering, contrary to the Born limit, i.e. in this case the Anderson theorem is restored in the unitarity limit. {\bf Acknowledgments} M. L. K. thanks Dierk Rainer for valuable discussions, suggestions and for careful reading and correcting the manuscript. M. L. K. acknowledges gratefully the support of the Deutsche Forschungsgemeinschaft through the Forschergruppe ''Transportph\"{a}nomene in Supraleitern und Suprafluiden''. O.V. D. thanks Nils Schopohl for valuable discussions and support.
\section{Introduction}\label{sec:one} Experiments at the high energy frontier require a good understanding of jets, their distribution in phase space with respect to each other and to the leptons and photons produced in mixed electroweak-QCD processes. The internal structure of jets is equally important. Internal jet structure is intimately related to the number of jets reconstructed by different jet-defining algorithms. It also is an issue in reconstructing the kinematic properties of single jets, such as their transverse momentum or direction, and the invariant mass of two- or three-jet systems. Both aspects are important when searching for signs of new physics in hadron collider experiments. One measure of internal jet structure is the transverse energy flow inside jets. This internal jet shape is defined as the fraction of a jet's transverse energy, $E_T$, which is deposited inside a sub-cone of radius $r<R$, where $R$ is the cone size of the jet in the $\eta$-$\phi$ plane. Such jet shape measurements have been performed in the past at $p\bar p$ colliders~\cite{ua1,tevatron} and also in photo-production and deep-inelastic-scattering (DIS) events at HERA~\cite{h1,zeusgamma,zeusdis}, and have been compared to theoretical calculations~\cite{eks,kramer,giele,seymour}. Comparisons of the shape of gluon-rich jets at hadron colliders with the quark-dominated jets produced in $ep$ and $e^+e^-$ collisions~\cite{opal} confirm the broader structure of gluon jets, which is expected because the larger color charge of gluons as compared to quarks leads to enhanced collinear radiation. A major obstacle for a precise comparison of jet shape data to perturbative QCD predictions is the fact that, at tree level, QCD jets are single, massless partons. Hence, a nontrivial jet shape only arises at higher order, where a jet may contain two or more partons. Two partons inside a jet, which appear in typical NLO cross section calculations, produce jet shapes at lowest order only. Jets with three partons first appear in two-loop calculations and thus a determination of jet shapes at true NLO, in a given physical process, is extremely demanding theoretically. Photo-production of jets~\cite{kramer} or dijet production at the Tevatron~\cite{eks} is a case in point: the kinematics of the event requires at least two hard jets in the final state, with balancing transverse momentum. A NLO jet shape, with jets consisting of three partons, thus requires four-parton final states at tree level and one-loop corrections to all $2\to 3$ processes involving quarks and gluons. These corrections are only now becoming available~\cite{nlo3jet}. In this letter we perform a full NLO jet shape calculation in a kinematically simpler situation: DIS at HERA. At sufficiently large $Q^2$, the scattered electron or positron in a DIS event provides the transverse momentum which is required to balance a high-$E_T$ jet. A three-parton final state then suffices to generate jets containing three colored partons. The soft and collinear divergences, which are generated by integrating the three-parton contributions over the entire phase space, are canceled by one-loop corrections to two-parton final states and two-loop corrections to one-parton contributions. However, a single parton cannot produce internal jet structure and thus all true two-loop effects can be neglected when determining differential jet shapes for up to three partons. The full one-loop QCD corrections for two- and three-parton final states in DIS are implemented in the MEPJET Monte Carlo program~\cite{mepjet}. Consequently, it is possible to extract full NLO jet shapes for the current jet in DIS with MEPJET. In the following, we analyze NLO corrections to the differential jet shape $\rho(r,R,E_{Tj},\eta_j)$ for events with a single jet of transverse energy $E_{Tj}$ and pseudo-rapidity $\eta_j$. The differential jet shape is defined as \begin{equation} \label{eq:rho} \rho(r,R,E_{Tj},\eta_j) = {1\over d^2\sigma_{NLO}/dE_{Tj}\, d\eta_j} \sum_{n}\; \int dr_n\; {E_{Tn} \over E_{Tj}} \delta(r-r_n) {d^3 \sigma \over dE_{Tj}\, d\eta_j\, dr_n}\; . \end{equation} Here the sum runs over all partons, $n$, belonging to the jet, which have a separation $r_n = \sqrt{(\eta_n-\eta_j)^2+(\phi_n-\phi_j)^2}=r<R$ from the axis of the jet in the legoplot. In practice, both the data and the Monte Carlo calculation replace the differential distributions by integrals over finite bin sizes in $E_{Tj}$, $\eta_j$ and $r$. Since we want to compare our NLO results with the ZEUS data, we follow the ZEUS event selection~\cite{zeusdis} and study current jets in neutral current (NC) DIS events with \begin{equation} \label{eq:cutszeus} E_e>10\;{\rm GeV}\;, \qquad Q^2>100\;{\rm GeV}^2\; , \qquad E_{Tj}>14\;{\rm GeV}\;, \qquad -1<\eta_j<2\; , \end{equation} where $E_e$ is the energy of the scattered electron in the laboratory frame. The default jet cone size is $R=1$ and cone slices of width $\Delta r = 0.1$ are considered. Because of the modest jet $E_T$, $b$-quarks are unlikely to appear inside a current jet. Production of massive $\bar bb$ pairs via photon-gluon fusion is small also, for the cuts considered below. Thus it is appropriate to perform the calculations in a fixed 4-flavor scheme, neglecting any $b$-quark contributions. Matching 4-flavor parton distributions are provided by the CTEQ4F4 set~\cite{cteq4f4}; for consistency with the parton distribution functions we use two-loop expressions for the running strong coupling constant, with $\Lambda^{(4)}_{\overline {\rm MS}}=292\;{\rm MeV}$. For the theoretical jet shapes, the factors in the denominator of Eq.~(\ref{eq:rho}) are determined using NLO DIS cross sections at ${\cal O}(\alpha_s)$, within a given set of selection cuts. Therefore, they correspond to 1-jet exclusive cross sections for the 1-jet cuts described below. For the ZEUS cuts, 2-jet events also enter with single weight only. The numerator is determined at full ${\cal O}(\alpha_s^2)$ for the NLO results and at ${\cal O}(\alpha_s)$ for the LO calculations. The default version of the MEPJET2.2 program is written for the calculation of NLO dijet cross sections in DIS, i.e. allowing for one soft or collinear parton in the final state. For the calculation of jet shapes at NLO, up to two soft and/or collinear partons must be generated. We have modified the MEPJET phase space generator to cover this enlarged phase space region for two- and three-parton final states. The matrix elements needed for the calculation of NLO jet shapes are identical to the ones used for the NLO dijet cross section and have been tested previously~\cite{mepjet,graudenz}. The cancellation of collinear and infrared singularities employs the phase space slicing method of Giele, Glover, and Kosower~\cite{ggk}, which removes soft and/or collinear regions of the three-parton phase space, where some pair of final and/or initial parton momenta satisfies $2p_i\cdot p_j < s_{min}$. Contributions from these regions are approximated by the appropriate asymptotic expressions and added analytically to the two-parton contributions, where they cancel the soft and collinear singularities of the virtual contributions. The final result must be independent of the choice of $s_{min}$, for values of this soft cutoff sufficiently small to make the asymptotic approximations valid. Numerical $s_{min}$ independence is a powerful test for the correct implementation of observables and of the phase space generator, as well as for the infrared stability of the jet clustering algorithms~\cite{nlo3jet}. Results of this $s_{min}$ test are shown in Fig.~\ref{fig:smin} for the differential jet shapes at $r=0.15$, 0.45, 0.65, and 0.95 in the successive combination algorithm of Ellis and Soper~\cite{es} (see below). Within Monte Carlo statistical errors of about 1-5\%, results are independent of $s_{min}$ below $s_{min}\approx 0.01\;{\rm GeV}^2$. We have checked that similar results hold for our implementation of the PUCELL jet algorithm employed by ZEUS~\cite{zeusgamma}. We will use $s_{min}=0.01\;{\rm GeV}^2$ in the following. \begin{figure}[t] \vspace*{0.5in} \begin{picture}(0,0)(0,0) \special{psfile=fig1.ps voffset=-300 hoffset=590 hscale=75 vscale=75 angle=90} \end{picture} \vspace{7.0cm} \caption{$s_{min}$ dependence of jet shapes $\rho(r)$ in four representative $r$-bins. The Ellis-Soper $k_T$ algorithm~\protect\cite{es} has been used, within the cuts of Eqs.~(\protect\ref{eq:cutszeus},\protect\ref{eq:cutsopt}) and requiring that only one hard jet, of $E_{Tj}>5$~GeV is seen in the pseudo-rapidity range $-2.5<\eta_j<2.5$. Errors represent Monte Carlo statistics only. Note that results in three of the four bins have been rescaled by the factors given in parentheses. } \vspace*{0.2in} \label{fig:smin} \end{figure} In our NLO calculation of jet shapes, the current jet contains up to three partons, which provides a much more detailed simulation of internal jet structure than is possible in a LO analysis. The extra detail suggests a direct comparison of NLO theory with data, with the $E_T$ flow of hadrons in the data replaced by parton $E_T$. Previous LO analyses modeled hadronization effects by introducing a phenomenological parameter, $R_{sep}$, which controlled the clustering of partons with legoplot separations between $R$ and $2R$\cite{rsep}. Our analysis implements the full experimental jet algorithms at the parton level, without introducing extra, tunable parameters. We have studied two algorithms in detail, the successive combination algorithm of Ellis and Soper~\cite{es}, called $k_T$ algorithm in the following, and the PUCELL algorithm used by the ZEUS Collaboration~\cite{zeusgamma,zeusdis}, for default cone sizes of $R=1$. The $k_T$ algorithm successively combines pairs of nearest partons/protojets to new protojets, up to a distance $R$ in the legoplot. Initially all partons are classified as protojets of transverse energy $E_T=p_T$. The algorithm then compares the $E_T$'s of protojets, via $d_i=E_{Ti}R$, with the $E_T$ weighted distances, $d_{ij}= {\rm min}(E_{Ti},E_{Tj}) \sqrt{(\eta_i-\eta_j)^2+(\phi_i-\phi_j)^2}$, of pairs of protojets in the legoplot. The pair of protojets with the smallest $d_{ij}$ is recombined if $d_{ij}$ is smaller than all the $d_i$ (which implies that their distance in the legoplot is smaller than $R$). Otherwise the protojet with the smallest $E_T$, and hence the smallest $d_i$, is called a jet and eliminated from the list of protojets. This process is iterated until all protojets have been assigned to jets. Note that some of these jets may be eliminated by subsequent selection cuts, as in Eq.~(\ref{eq:cutszeus}). Some freedom exists in the assignment of jet momenta, i.e. in the recombination scheme. For the $k_T$ algorithm we use the $E$-scheme, i.e. the jet four-momentum is the sum of the four-momenta of the partons belonging to the jet and the jet $E_T$ is the sum of the parton $p_T$'s. Apart from the definition of recombined momenta, our $k_T$ algorithm is identical to the ones described in Refs.~\cite{seymour,es}. The PUCELL algorithm is an iterative fixed cone algorithm. In a first step all partons with $p_T>300$~MeV are considered as seeds for the formation of pre-clusters. Starting with the highest $p_T$ seed, a pre-cluster is formed, containing this seed and all seeds within a cone of radius $R$ around it. This procedure is iterated with the remaining seeds, which do not yet belong to a pre-cluster, in order of decreasing $p_T$. All the seeds belonging to a given pre-cluster are then merged, according to the Snowmass convention~\cite{snowmass}, i.e., the legoplot variables of the pre-clusters are defined as the $E_T$-weighted averages over the seeds, \begin{eqnarray} \label{eq:snowmass1} E_{T,p.c.} & = & \sum_i E_{Ti} \;,\\ \label{eq:snowmass2} \eta_{p.c.}& = & {1\over E_{T,p.c.}} \sum_i E_{Ti}\;\eta_i \;,\\ \label{eq:snowmass3} \phi_{p.c.}& = & {1\over E_{T,p.c.}} \sum_i E_{Ti}\;\phi_i \;. \end{eqnarray} In a second step, all partons are considered, even if they fall below the seed threshold of 300~MeV. A new cluster axis is calculated as in (\ref{eq:snowmass1}-\ref{eq:snowmass3}), in terms of all partons inside a cone of radius $R$ of the old (pre)cluster axis. This second step is iterated until the contents of all clusters stabilize. A third step deals with overlapping clusters, i.e. with clusters which share partons. If the energy of the common partons is more than 75\% of the energy of the lower-energy cluster, the two clusters are merged to a single jet. Otherwise two separate jets are formed and common partons are assigned to the nearest jet. We can now compare the jet shape predictions for these jet algorithms at both leading and next-to-leading order. One way to assess the improvements from a NLO calculation is to study the scale dependence of observables. The process at hand is inclusive DIS which basically has a single scale only, $Q$, the momentum transfer carried by the virtual photon. $Q$ is the obvious choice for the factorization scale and we fix $\mu_f^2=Q^2$ throughout. On the other hand, we are investigating the inner structure of jets, and the average jet mass or the average transverse momentum of partons, with respect to the jet axis, appear as reasonable scale choices\footnote{The mass of the jet in an individual event appears as an integration variable in the determination of $\rho(r)$ and is not a physical scale of the observable.}. These scales are proportional to the intrinsic scale of the entire process, $Q$, and become different from zero only at ${\cal O}(\alpha_s)$, thus suggesting $\alpha_s Q^2$ as a scale choice. This leads us to investigate variations of the renormalization scale $\mu_r^2=\xi \alpha_s(Q^2)Q^2$ with the scale factor $\xi$. Searching for minimal sensitivity~\cite{stevenson} of $\rho(r)$, a ``correct'' scale $\mu_r=Q$ would appear as a flat $\xi$-dependence near $\xi=1/\alpha_s$. Thus, for all practical purposes, our choice is general enough. For the PUCELL algorithm and two representative bins in $r$, the scale variation of the differential jet shape, $\rho(r)$, is shown in Fig.~\ref{fig:scale.PUC}. \begin{figure}[t] \vspace*{0.5in} \begin{picture}(0,0)(0,0) \special{psfile=fig2a.ps voffset=-280 hoffset=460 hscale=65 vscale=65 angle=90} \special{psfile=fig2b.ps voffset=-280 hoffset=680 hscale=65 vscale=65 angle=90} \end{picture} \vspace{7.0cm} \caption{Renormalization scale dependence of the differential jet shape $\rho(r)$ at (a) $r=0.15$ and (b) $r=0.45$, using the PUCELL algorithm. \Blue{The dash-dotted and dashed lines are the LO and NLO results for the acceptance cuts of Eq.~(\protect\ref{eq:cutszeus})}. \Red{Also shown are LO (dotted line) and NLO (solid line) results for events with one single jet only of $E_T>5$~GeV (see Eqs.~(\protect\ref{eq:jetveto},\protect\ref{eq:cutsopt}))}. } \vspace*{0.2in} \label{fig:scale.PUC} \end{figure} The dash-dotted and dashed lines show the LO and NLO results for the generic ZEUS acceptance cuts of Eq.~(\ref{eq:cutszeus}). The $E_T$ flow is somewhat higher at NLO, i.e. jets are broader. However, the renormalization scale dependence is almost as large at NLO as for the LO case. This disappointing result can be traced to a large contribution from events with an additional low $E_T$ jet. In DIS events with two jets, at ${\cal O}(\alpha_s^2)$, at best one jet can contain two partons, and, hence, the jet shape is modeled at LO only. In order to enhance the contribution from jets with the maximal number of partons, which then are truly modeled at NLO, events which contain additional jets should be eliminated. We achieve this goal by vetoing events containing any additional jets with \begin{equation} \label{eq:jetveto} E_{Tj}>5\;{\rm GeV}\;, \qquad -2.5<\eta_j<2.5\; . \end{equation} These veto cuts are chosen such that they can be easily implemented experimentally. A further reduction of events with hadronic activity beyond the current jet is achieved by requiring the scattered lepton and the observed hard jet to be back-to-back in azimuth. Allowing for finite detector resolution we require \begin{equation} \label{eq:cutsopt} \Delta\phi_{ej}=|\phi_e-\phi_j|>3\;. \end{equation} Also shown in Fig.~\ref{fig:scale.PUC} is the scale dependence of $\rho(r)$ for events with the jet veto cut of Eq.~(\ref{eq:jetveto}) and the back-to-back cut of Eq.~(\ref{eq:cutsopt}). In the following we call these restrictions ``1-jet cuts''. For these 1-jet events the scale independence of $\rho(r)$ is significantly improved at NLO, out to distances of $r\approx 0.6$. The factors in the denominator of Eq.~(\ref{eq:rho}) are very stable against scale variations for inclusive DIS events, i.e. within the ZEUS cuts of Eq.~(\ref{eq:cutszeus}). The 1-jet cuts, however, lead to a sizable reduction of $d^2\sigma_{NLO}/dE_{Tj}\, d\eta_j$ in the denominator of Eq.~(\ref{eq:rho}), corresponding to a subtraction of two-parton final states with dijet-type kinematics. This subtraction term is modeled at leading order only and is strongly scale dependent. However, it can be calculated at ${\cal O} (\alpha_s^2)$ also, which results in a more reliable determination. For the 1-jet cuts we find minimal scale sensitivity~\cite{stevenson} near $\mu_r^2= \alpha_s(Q^2)Q^2$, and at this value the ${\cal O} (\alpha_s)$ and ${\cal O} (\alpha_s^2)$ results for the subtraction terms virtually agree. We therefore use $\mu_r^2= \alpha_s(Q^2)Q^2$ in the denominator, which increases the normalization of the resulting jet shapes by 10-15\% as compared to the choice $\mu_r=Q$. The renormalization scale dependences of jet shapes in the PUCELL and the $k_T$ algorithm are compared in Fig.~\ref{fig:scale.opt}, using the 1-jet cuts. Results are shown for four representative $r$-bins, at LO (dash-dotted and dotted lines) and at NLO (dashed and solid lines). The LO curves are virtually identical for the two algorithms. This is to be expected, since the criterion for merging two partons is the same at LO, namely their separation in the legoplot must be less than $R=1$. Small differences are either statistical or due to the different recombination schemes for parton momenta. At NLO, the PUCELL and the $k_T$ algorithm produce quite similar results for the central part of the jet (at small $r$). In this region, sensitivity to the choice of renormalization scale is minimal near $\xi=1$ which confirms our basic choice of $\mu_r^2=\alpha_s Q^2$. At distances from the jet axis beyond $r\approx 0.5$, the iterative nature of the PUCELL algorithm is more likely to gather a third parton into the jet, making it somewhat broader on average than jets reconstructed by the $k_T$ algorithm. Since this broadening effect requires three partons and is therefore only modeled at tree level in our NLO calculation, the NLO enhancement of $\rho(r)$ at large $r$ shows a pronounced scale dependence, which is significantly stronger in the PUCELL than in the $k_T$ algorithm. \begin{figure}[t] \vspace*{0.5in} \begin{picture}(0,0)(0,0) \special{psfile=fig3a.ps voffset=-190 hoffset=457 hscale=60 vscale=55 angle=90} \special{psfile=fig3b.ps voffset=-190 hoffset=657 hscale=60 vscale=55 angle=90} \special{psfile=fig3c.ps voffset=-380 hoffset=450 hscale=60 vscale=55 angle=90} \special{psfile=fig3d.ps voffset=-380 hoffset=650 hscale=60 vscale=55 angle=90} \end{picture} \vspace{11.0cm} \caption{Renormalization scale dependence of the differential jet shape $\rho(r)$ at (a) $r=0.15$, (b) $r=0.55$, (c) $r=0.75$, and (d) $r=0.95$, for events satisfying the 1-jet cuts of Eqs.~(\protect\ref{eq:jetveto},\protect\ref{eq:cutsopt}). Results are shown for the \Blue{PUCELL algorithm at LO (dash-dotted line) and NLO (dashed line)} and for the \Red{$k_T$ algorithm at LO (dotted line) and NLO (solid line)}. } \label{fig:scale.opt} \end{figure} Differential jet shapes in DIS, as a function of $E_T$ and pseudo-rapidity of the jet, have been measured by ZEUS~\cite{zeusdis}. In Fig.~\ref{fig:zeus}(a) we compare our LO and NLO QCD predictions with the ZEUS measurements for events with at least one hard jet. This jet is identified with the PUCELL algorithm and must lie in the interval 14~GeV~$<E_T<$~21~GeV and $-1<\eta<2$. The agreement between data and theory in the $0.1<r<1$ range is improved significantly at NLO. Calculation of the jet shape at $r=0$ is currently not possible, since it would require the inclusion of two-loop contributions and the resummation of multiple soft and collinear emission. As discussed previously, the scale dependence of the NLO results is quite large within the ZEUS acceptance cuts. The ZEUS data (and our NLO simulations) show, on the other hand, that the jet shapes depend very little on the jet pseudo-rapidity, and modestly on jet $E_T$. Imposing the 1-jet cuts of Eqs.~(\protect\ref{eq:jetveto},\protect\ref{eq:cutsopt}) does not significantly change $\rho(r)$ for the default scale choice of $\mu_r^2=\alpha_s Q^2$, but it reduces the scale uncertainty. Included in Fig.~\ref{fig:zeus} are the NLO results using the 1-jet cuts. The NLO QCD predictions for $\rho(r)$ agree very well with the data, up to values around $r\approx 0.8$ for the 1-jet cuts and $r\approx 0.6$ for the generic ZEUS cuts. This corresponds to the regions where we have found a small scale dependence. In other words, the differences between data and NLO QCD are consistent with higher order QCD effects. Note that the agreement between theory and data is much worse at LO. \begin{figure}[t] \vspace*{0.5in} \begin{picture}(0,0)(0,0) \special{psfile=fig4a.eps voffset=-210 hoffset=-10 hscale=50 vscale=50 angle=0} \special{psfile=fig4b.eps voffset=-210 hoffset=230 hscale=50 vscale=50 angle=0} \end{picture} \vspace{7.0cm} \caption{Comparison of ZEUS jet shape data~\protect\cite{zeusdis} with QCD predictions for DIS jets reconstructed by the PUCELL algorithm. Jet cuts are: $-1<\eta<2$ and (a,b) 14~GeV~$<E_T<$~21~GeV, (b) 37~GeV~$<E_T<$~45~GeV. ZEUS data (circles and squares) are compared in the lower $E_T$ range (a) with \Red{LO (lower band)} and \Blue{NLO (dashed line)} QCD predictions. The upper band in (a) and the two lines in (b) represent NLO jet shapes within the 1-jet cuts of Eqs.~(\protect\ref{eq:jetveto},\protect\ref{eq:cutsopt}). The width of the bands corresponds to varying the renormalization scale between $\mu_r^2=\alpha_s Q^2/4$ and $\mu_r^2=4\alpha_s Q^2$. } \vspace*{0.2in} \label{fig:zeus} \end{figure} Similar agreement between true NLO QCD and ZEUS data is found for different jet $E_T$ and $\eta$ ranges. One example is shown in Fig.~\ref{fig:zeus}(b), where jet shapes for jets with 14~GeV~$<E_T<$~21~GeV and 37~GeV~$<E_T<$~45~GeV are compared. The data and the NLO QCD predictions clearly show that higher $E_T$ jets are narrower. It should be noted that the excellent agreement between data and theory, at the 10\% level, is obtained only when applying the minimal sensitivity criteria~\cite{stevenson} to pick the renormalization scale. A na\"{\i}ve choice, like $\mu_r=Q$ (roughly corresponding to $\xi\approx 10$ in Fig.~\ref{fig:scale.opt}) would lead to substantially larger deviations from the experimental results. We have performed a first analysis of true NLO jet shapes in DIS. We find excellent agreement between NLO QCD and data. This agreement is achieved without introducing extra phenomenological parameters describing hadronization effects, like $R_{sep}$. For the PUCELL and the $k_T$ algorithm we have shown that the scale uncertainty of the QCD predictions for jet shapes is substantially reduced at NLO. A precise comparison of experiment with theory is facilitated by cuts which suppress multi-jet events. Our analysis can be easily extended to other jet algorithms and momentum recombination schemes and allows one to investigate questions like the infrared safety of jet algorithms or the reconstruction of kinematical variables from jets at full NLO. Their effect on the matching of theory and data for $ep$ collisions at HERA and more generally at hadron colliders can now be investigated at the one-loop level. \acknowledgements This research was supported in part by the University of Wisconsin Research Committee with funds granted by the Wisconsin Alumni Research Foundation and in part by the U.~S.~Department of Energy under Contract No.~DE-FG02-95ER40896.
\section{Introduction} The covariances of algebraic structures are of central importance for the physical and mathematical theories constructed by using these algebraic structures. The investigation of quantum group covariant structures gives rise to the introduction of braided groups (a self contained review can be found in \cite{marev}) which are obtained from quantum groups by a covariantization process called transmutation . Quantum groups of function algebra type are not invariant under their own transformation, i.e., the quantum algebra $A(R)$ is not a comodule algebra under the adjoint coaction. However, it is possible to obtain an $A(R)$-comodule algebra called braided algebra $B(R)$ obtained by covariantizing the algebra of the coacted copy keeping the coalgebra unchanged. This covariantization process deforms the notion of tensor product and leads to the generalization of the usual Hopf algebra axioms called braided Hopf algebra axioms \cite{majid1}. Thus the braided Hopf algebras can be used to generalize the supersymmetric structures via the generalization of super tensor product and super Hopf algebras \cite{dunne} or to introduce oscillators with braid statistics \cite{ali}. In this work, we investigate the covariance properties of the algebra satisfied by the entries of the braided matrix $BSL_q(2)$. In other words, we investigate if there is any Hopf algebra other the quantum algebra the coaction of which makes the braided algebra a comodule algebra. Since the coacting and the coacted algebras have the same coalgebra structure in transmutation theory, we first investigate the general braided Hopf algebra structure. We find that there is one more braided Hopf algebra other the one given in the literature. We also find that the nonbraided Hopf algebra whose coproduct is the same as the braided one is a two parameter deformed Hopf algebra. This algebra, like the quantum algebra, is not a comodule algebra under the adjoint coaction. For a certain value of one of the deformation parameters it turns out that the braided algebra becomes a comodule algebra under the coaction of this one parameter nonbraided algebra. We explicitly construct the transmutation of the noncovariant (nonbraided) algebra to the covariant (braided) one. \section{Prelimineries and Review} The right coaction of a Hopf algebra $A$ on $H$ is a linear map $\beta :H\rightarrow H\otimes A$, i.e., \be\label{eq:coaction9} \beta (h)=\Sigma h^{(i)}\otimes a^{(i)},\quad h^{(i)}\in H,\quad a^{(i)}\in A \ee \noindent satisfying \be \label{eq:beta} (\beta \otimes id)\circ \beta =(id\otimes \Delta )\circ \beta \quad (id\otimes \epsilon )\circ \beta =id. \ee \noindent The algebra $H$ is a right $A$-comodule algebra if the map $\beta$ is an algebra homomorphism such that \be \beta (h\cdot g)=\beta (h)\cdot \beta (g) \quad \forall h,g\in H. \ee \noindent The consistency of the algebra homomorphism requires that $\beta (1_H)=1_H\otimes 1_A $. The right adjoint coaction of a Hopf algebra on itself is defined by \be\label{eq:adaction} \beta (h)=\Sigma h_{(2)}\otimes S(h_{(1)})\cdot h_{(3)} \ee \noindent where $h_{(1)},h_{(2)},h_{(3)}$ are given by \be \Delta ^{2}(h)= \Sigma h_{(1)}\otimes h_{(2)}\otimes h_{(3)}. \ee \noindent The quantum algebra ($A(R)$) of the quantum matrix $SL_q(2)$ is generated by $a,b,c,d$ and $1$ satisfying the relations \begin{eqnarray}\label{eq:quantumgrouprelations} a\cdot b &=& q^{-1}b\cdot a, \nonumber \\ a\cdot c &=& q^{-1}c\cdot a,\nonumber \\ b\cdot d &=& q^{-1}d\cdot b,\nonumber \\ c\cdot d &=& q^{-1}d\cdot c, \\ b\cdot c &=& c\cdot b, \nonumber \\ a\cdot d-d\cdot a &=& (q^{-1}-q)b\cdot c \nonumber \\ a\cdot d-q^{-1}b\cdot c &=& =1 \nonumber \end{eqnarray} \noindent the Hopf structure of which is given by the coproducts \begin{eqnarray}\label{eq:cop2} \Delta (a) &=& a\otimes a+b\otimes c, \nonumber \\ \Delta (b) &=& a\otimes b+b\otimes d, \nonumber \\ \Delta (c) &=& c\otimes a+d\otimes c, \\ \Delta (d) &=& c\otimes b+d\otimes d \nonumber \end{eqnarray} \noindent the counits \be \epsilon (a)=1,\ \ \epsilon (b)=0,\ \ \epsilon (c)=0,\ \ \epsilon (d)=1, \ee \noindent and by the antipodes \be S(a)=d,\ \ S(b)=-qb,\ \ S(c)=-q^{-1}c,\ \ S(d)=a. \ee \noindent The $*$-algebra structure with real $q$ is given by \be a^{*}=d,\ \ b^{*}=-qb,\ \ c^{*}=-q^{-1}c,\ \ d^{*}=a. \ee \noindent The adjoint coaction of $A(R)$ on itself is then calculated to give \begin{eqnarray}\label{eq:coaction1} \beta (a) &=& a\otimes d\cdot a+b\otimes d\cdot c+c\otimes(-qb\cdot a)+d\otimes(-qb\cdot c) \nonumber \\ \beta (b) &=& a\otimes d\cdot b+b\otimes d\cdot d +c\otimes (-qb\cdot c)+d\otimes (-qb\cdot d)\nonumber \\ \beta (c) &=& a\otimes (-q^{-1}c\cdot a)+b\otimes (-q^{-1}c\cdot c)+c\otimes a\cdot a+d\otimes a\cdot c \\ \beta (d) &=& a\otimes (-q^{-1}c\cdot b)+b\otimes(-q^{-1}c\cdot d)+c\otimes a\cdot b+d\otimes a\cdot d. \nonumber \end{eqnarray} \noindent It can easily be seen that these mappings do not define an algebra homomorphism. Hence $A(R)$ itself is not an $A(R)$-comodule algebra. However, the multiplication ($\cdot $) of the coacted copy (first elements in the tensor product) is replaced by a multiplication ($\underline{\cdot } $) such that \begin{eqnarray}\label{eq:braidedmultiplication} a\underline{\cdot }a &=& a\cdot a\nonumber \\ a\underline{\cdot }b &=& a\cdot b\nonumber \\ a\underline{\cdot }c &=& qc\cdot a\nonumber \\ a\underline{\cdot }d &=& a\cdot d+(q-q^{-1})c\cdot b\nonumber \\ b\underline{\cdot }a &=& q^2a\cdot b\nonumber \\ b\underline{\cdot }b &=& qb\cdot b\nonumber \\ b\underline{\cdot }c &=& q^{-1}b\cdot c+(1-q^{-2})(d-a)\cdot a\nonumber \\ b\underline{\cdot }d &=& qb\cdot d-(1-q^{-2})a\cdot b\\ c\underline{\cdot }a &=& q^{-1}c\cdot a\nonumber \\ c\underline{\cdot }b &=& q^{-1}c\cdot b\nonumber \\ c\underline{\cdot }c &=& qc\cdot c\nonumber \\ c\underline{\cdot }d &=& qc\cdot d\nonumber \\ d\underline{\cdot }a &=& d\cdot a\nonumber \\ d\underline{\cdot }b &=& d\cdot b\nonumber \\ d\underline{\cdot }c &=& d\cdot c-q^{-1}(1-q^{-2})c\cdot a\nonumber \\ d\underline{\cdot }d &=& d\cdot d-q^{-1}(1-q^{-2})c\cdot b.\nonumber \end{eqnarray} \noindent Then the algebra ($B(R)$) satisfied by $a,b,c,d$ and $1$ with this new multiplication ($\underline{\cdot }$) \begin{eqnarray}\label{eq:braidedgrouprelations} b\underline{\cdot }a&=& q^2a\underline{\cdot }b, \nonumber \\ c\underline{\cdot }a&=& q^{-2}a\underline{\cdot }c, \nonumber \\ a\underline{\cdot }d&=& d\underline{\cdot }a, \\ b\underline{\cdot }c&=& c\underline{\cdot }b\ +\ (1-q^{-2})a\underline{\cdot }(d-a), \nonumber \\ d\underline{\cdot }b&=& b\underline{\cdot }d\ +\ (1-q^{-2})a\underline{\cdot }b, \nonumber \\ c\underline{\cdot }d&=& d\underline{\cdot }c\ +\ (1-q^{-2})c\underline{\cdot }a, \nonumber \\ a\underline{\cdot }d-q^2c\underline{\cdot }b &=& 1. \nonumber \end{eqnarray} \noindent is an $A(R)$-comodule algebra under the coaction (\ref{eq:coaction1}). This covariantization process is called transmutation. The transformation under a noncommutative algebra ($A(R)$ for instance) deforms the notion of tensor product because the transformed algebras are no longer independent. The deformed tensor product is called the braided tensor product and denoted by $ \underline{\otimes}$. The coalgebra of the transmuted algebra $B(R)$ is of the same form as the original algebra $A(R)$, i.e., \begin{eqnarray}\label{eq:cop1} \underline{\Delta} (a)&=&a\underline{\otimes} a+b\underline{\otimes} c, \nonumber \\ \underline{\Delta} (b)&=&a\underline{\otimes} b+b\underline{\otimes} d, \nonumber \\ \underline{\Delta} (c)&=&c\underline{\otimes} a+d\underline{\otimes} c, \\ \underline{\Delta} (d)&=&c\underline{\otimes} b+d\underline{\otimes} d \nonumber \\ \underline{\epsilon} (a)&=&\underline{\epsilon} (d)=1,\ \underline{\epsilon} (b)=\underline{\epsilon} (c)=0. \nonumber \end{eqnarray} \noindent The antipodes of the generators of $B(R)$ are given by \be \label{eq:ant3} \underline{S}(a)=q^{2}d+(1-q^{2})a,\ \ \underline{S}(b)=-q^{2}b,\ \ \underline{S}(c)=-q^{2}c,\ \ \underline{S}(d)=a \ee \noindent The coproducts define an algebra homomorphism in the braided tensor product space with the braided tensor product algebra such that \be (x\underline{\otimes }y)\underline{\cdot }(w\underline{\otimes} z)=x\psi(y\underline{\otimes} z)d\quad x,y,w,z \in B. \ee \noindent where $\psi $ is a generalization of the permutation map in boson algebras. The braiding relations define a nontrivial statistics between copies of algebras. The braiding relations for the braided algebra (\ref{eq:braidedgrouprelations}) with the coalgebra (\ref{eq:cop1}) are given by \begin{eqnarray}\label{eq:braiding1} \psi (a\underline{\otimes} a)&=&a\underline{\otimes} a+(1-q^2)b\underline{\otimes} c \nonumber\\ \psi (a\underline{\otimes} b)&=&b\underline{\otimes} a \nonumber\\ \psi (a\underline{\otimes} c)&=&c\underline{\otimes} a+(1-q^2)(d-a)\underline{\otimes} c \nonumber\\ \psi (a\underline{\otimes} d)&=&d\underline{\otimes} a+(1-q^{-2})b\underline{\otimes} c \nonumber\\ \psi (b\underline{\otimes} a)&=&a\underline{\otimes} b+(1-q^2)b\underline{\otimes} (d-a) \nonumber\\ \psi (b\underline{\otimes} b)&=&q^2b\underline{\otimes} b \nonumber\\ \psi (b\underline{\otimes} c)&=&q^{-2}c\underline{\otimes} b+(1+q^2)(1-q^2)^2b\underline{\otimes} c-(1-q^{-2})(d-a)\underline{\otimes} (d-a) \nonumber\\ \psi (b\underline{\otimes} d)&=&d\underline{\otimes} b+(1-q^{-2})b\underline{\otimes} (d-a) \nonumber\\ \psi (c\underline{\otimes} a)&=&a\underline{\otimes} c \nonumber\\ \psi (c\underline{\otimes} b)&=&q^{-2}b\underline{\otimes} c \\ \psi (c\underline{\otimes} c)&=&q^2c\underline{\otimes} c \nonumber\\ \psi (c\underline{\otimes} d)&=&d\underline{\otimes} c \nonumber\\ \psi (d\underline{\otimes} a)&=&a\underline{\otimes} d+(1-q^{-2})b\underline{\otimes} c \nonumber\\ \psi (d\underline{\otimes} b)&=&b\underline{\otimes} d \nonumber\\ \psi (d\underline{\otimes} c)&=&c\underline{\otimes} d+(1-q^{-2})(d-a)\underline{\otimes} c \nonumber\\ \psi (d\underline{\otimes} d)&=&d\underline{\otimes} d-q^{-2}(1-q^{-2})b\underline{\otimes} c. \nonumber\\ \end{eqnarray} \noindent The central element $q^{-1}a+qd$ in quantum algebra (which is the quantum trace in the matrix algebra) is not only central in the braided algebra but also bosonic in the sense that \be \psi((q^{-1}a+qd)\underline{\otimes}x)=x\underline{\otimes}(q^{-1}a+qd), \ \ \psi(x\underline{\otimes}(q^{-1}a+qd))=(q^{-1}a+qd)\underline{\otimes}x\ \ \forall x\in B(R). \ee \noindent The algebras in the braided category satisfy the braided Hopf algebra axioms \cite{majid1} \begin{eqnarray} \label{eq:bhopf} m\circ (id\otimes m)&=&m\circ (m\otimes id) \nonumber \\ m\circ (id\otimes \eta )&=&m\circ (\eta \otimes id) = id \nonumber \\ (id\otimes \Delta )\circ \Delta &=&(\Delta \otimes id)\circ \Delta \nonumber \\ (\epsilon \otimes id)\circ \Delta &=&(id\otimes \epsilon )\circ \Delta = id \nonumber \\ m\circ (id\otimes S)\circ \Delta &=&m\circ (S\otimes id)\circ \Delta =\eta \circ \epsilon \nonumber \\ \psi \circ (m \otimes id) &=& (id \otimes m)\circ (\psi \otimes id)\circ (id \otimes \psi) \nonumber \\ \psi \circ (id \otimes m) &=& (m \otimes id)\circ (id \otimes \psi)\circ (\psi \otimes id)\nonumber \\ (id \otimes \Delta)\circ \psi &=& (\psi \otimes id)\circ (id \otimes \psi )\circ (\Delta \otimes id) \nonumber \\ (\Delta \otimes id)\circ \psi &=& (id \otimes \psi )(\psi \otimes id)\circ (id \otimes \Delta) \nonumber \\ \Delta \circ m &=& (m \otimes m)(id\otimes \psi \otimes id)\circ (\Delta \otimes \Delta ) \\ S\circ m &=& m\circ \psi \circ (S\otimes S)\nonumber \\ \Delta \circ S &=& (S\otimes S)\circ \psi \circ \Delta \nonumber \\ \epsilon \circ m &=& \epsilon \otimes \epsilon \nonumber \\ (\psi \otimes id)\circ (id \otimes \psi)\circ (\psi \otimes id)&=&(id \otimes \psi)\circ (\psi \otimes id)\circ (id \otimes \psi) \nonumber \end{eqnarray} \noindent and the $\underline{*}$ algebras in the same category with real deformation parameter also satisfy \cite{majid2} \begin{eqnarray}\label{eq:*1} \Delta \circ * &=& \pi \circ (* \otimes *)\circ \Delta\nonumber \\ S\circ * &=& *\circ S \\ (x\otimes y)^{*} &=& y^{*}\otimes x^{*},\ \forall x,y\in B \nonumber \end{eqnarray} \noindent where we omit the underlining for the sake of clarity and all mappings are in the braided sense in the axioms (\ref{eq:bhopf})-(\ref{eq:*1}). Note that in the $\psi\rightarrow \pi$ limit the axioms (\ref{eq:bhopf}) reduce to the usual Hopf algebra axioms. It is also possible to define more general Hopf algebras where the counit map is no longer an algebra homomorphism \cite{durdevic}. The involutions for (\ref{eq:braidedgrouprelations}) \be \label{eq:star1} a^{\underline{*}}=a,\ \ b^{\underline{*}}=c,\ \ c^{\underline{*}}=d,\ \ d^{\underline{*}}=d \ee \noindent complete the braided Hopf $\underline{*}$-algebra structure. \section{A New Braided Hopf Algebra Solution} Expressing the braided algebra in terms of a central and bosonic element \be \label{eq:p} p\equiv q^{-2}a+d \ee \noindent and three other generators $a,b,c$ makes a lot simplification in the calculations. The algebra (\ref{eq:braidedgrouprelations}) can equivalently be expressed as \begin{eqnarray}\label{eq:ba1} b\underline{\cdot }a &=& q^2a\underline{\cdot }b \nonumber \\ a\underline{\cdot }c &=& q^2c\underline{\cdot }a \\ b\underline{\cdot }c &=& c\underline{\cdot }b-(1-q^{-4})a^2+(1-q^{-2})p\underline{\cdot }a \nonumber \\ q^{-2}a\underline{\cdot }a+a\underline{\cdot }p-q^2c\underline{\cdot } b &=& 1 \nonumber \end{eqnarray} \noindent The $\underline{*}$-structure (\ref{eq:star1}) implies \be a^{\underline{*}}=a,\ \ b^{\underline{*}}=c,\ \ c^{\underline{*}}=b,\ \ p^{\underline{*}}=p. \ee \noindent We write the general forms of the coproducts \begin{eqnarray}\label{eq:gbha} \underline{\Delta }(a) &=& A_1a\underline{\otimes } a+A_2b\underline{\otimes } c +A_3c\underline{\otimes } b+A_4p\underline{\otimes } a+A_5a\underline{\otimes } p +A_61\underline{\otimes } a \nonumber \\ & &+A_7a\underline{\otimes } 1+A_81\underline{\otimes } 1+A_9p\underline{\otimes } p+A_{10}1\underline{\otimes } p+A_{11}p\underline{\otimes } 1, \nonumber \\ \underline{\Delta }(b) &=& B_1a\underline{\otimes } b+B_2b\underline{\otimes } a+B_3b\underline{\otimes } p+B_4p\underline{\otimes } b+B_51\underline{\otimes } b+B_6b\underline{\otimes } 1, \\ \underline{\Delta }(c) &=& B_1c\underline{\otimes } a+B_2a\underline{\otimes } c+B_3p\underline{\otimes } c+B_4c\underline{\otimes } p+B_5c \underline{\otimes } 1+B_61\underline{\otimes } c, \nonumber \\ \underline{\Delta }(p) &=& C_1a\underline{\otimes } a+C_2b\underline{\otimes } c+C_3c \underline{\otimes } b+C_4p\underline{\otimes } a+C_5a\underline{\otimes } p \nonumber \\ & & +C_61\underline{\otimes } a +C_7a\underline{\otimes } 1+C_81\underline{\otimes } 1+C_9p\underline{\otimes } p+C_{10}1\underline{\otimes } p+C_{11}p\underline{\otimes } 1 \nonumber \end{eqnarray} \noindent the counits \begin{eqnarray} \underline{\epsilon }(a) &=& e_1, \nonumber \\ \underline{\epsilon }(b) &=& \underline{\epsilon }(c)=e_2, \\ \underline{\epsilon }(p) &=& e_3 \nonumber \end{eqnarray} \noindent the antipodes \begin{eqnarray} \underline{S}(a) &=& k_1a+k_2b+k_3c+k_4p+k_5, \nonumber \\ \underline{S}(b) &=& m_1a+m_2b+m_3c+m_4p+m_5, \\ \underline{S}(c) &=& m_1a+m_2c+m_3b+m_4p+m_5, \nonumber \\ \underline{S}(p) &=& n_1a+n_2b+n_3c+n_4p+n_5 \nonumber \end{eqnarray} \noindent the braidings \begin{eqnarray}\label{eq:babraidings} \psi (a\underline{\otimes } a) &=& g_1a\underline{\otimes } a+g_2b\underline{\otimes } c +g_3c\underline{\otimes } b+g_4p\underline{\otimes } a+g_5a\underline{\otimes } p \nonumber \\ & &+g_61\underline{\otimes } a+g_7a\underline{\otimes } 1+g_81\underline{\otimes } 1+g_9p\underline{\otimes } p+g_{10}1\underline{\otimes } p+g_{11}p\underline{\otimes } 1, \nonumber \\ \psi (a\underline{\otimes } b) &=& d_1b\underline{\otimes } a+d_2a\underline{\otimes } b+d_3p\underline{\otimes } b+d_4b\underline{\otimes } p+d_51\underline{\otimes } b+d_6b\underline{\otimes } 1, \nonumber \\ \psi (a\underline{\otimes } c) &=& f_1c\underline{\otimes } a+f_2a\underline{\otimes } c+f_3p\underline{\otimes } c+f_4c \underline{\otimes } p+f_51\underline{\otimes } c+f_6c\underline{\otimes } 1, \nonumber \\ \psi (b\underline{\otimes } b) &=& z_1b\underline{\otimes } b, \\ \psi (c\underline{\otimes } b) &=& c_1a\underline{\otimes } a+c_2b\underline{\otimes } c +c_3c\underline{\otimes } b+c_4p\underline{\otimes } a+c_5a\underline{\otimes } p \nonumber \\ & & +c_61\underline{\otimes } a+c_7a\underline{\otimes } 1+c_81\underline{\otimes } 1+c_9p\underline{\otimes } p+c_{10}1\underline{\otimes } p+c_{11}p\underline{\otimes } 1, \nonumber \\ \psi (b\underline{\otimes } c) &=& a_1a\underline{\otimes } a+a_2b\underline{\otimes } c +a_3c\underline{\otimes } b+a_4p\underline{\otimes } a+a_5a\underline{\otimes } p \nonumber \\ & & +a_61\underline{\otimes } a+a_7a\underline{\otimes } 1+a_81\underline{\otimes } 1 +a_9p\underline{\otimes } p+a_{10}1\underline{\otimes } p+a_{11}p\underline{\otimes } 1 \nonumber \end{eqnarray} \noindent and their $\underline{*}$-involutions to find the solutions for the braided Hopf algebra structure. The symbols with a subscript are the parameters to be determined. For the bosonic trace , the braiding of the central element $p$ is trivial. We substitute these general forms into the braided Hopf algebra axioms (\ref{eq:bhopf}) and solve the equations by using the computer programming Maple V. We find that that there are only two solutions. For the first solution the coproducts \begin{eqnarray}\label{eq:firsttype} \underline{\Delta }(a) &=& a\underline{\otimes } a+b\underline{\otimes } c \nonumber \\ \underline{\Delta }(b) &=& a\underline{\otimes } b-q^{-2}b\underline{\otimes } a+b\underline{\otimes } p \\ \underline{\Delta }(c) &=& c\underline{\otimes } a-q^{-2}a\underline{\otimes } c+p\underline{\otimes } c \nonumber \\ \underline{\Delta }(p) &=& (q^{-2}+q^{-4})a\underline{\otimes } a+q^{-2}b\underline{\otimes } c +c\underline{\otimes } b-q^{-2}p\underline{\otimes } a-q^{-2}a\underline{\otimes } p+p\underline{\otimes } p \nonumber \end{eqnarray} \noindent the counits \begin{eqnarray} \underline{\epsilon }(a) &=& 1, \nonumber \\ \underline{\epsilon }(b) &=& 0 \\ \underline{\epsilon }(c) &=& 0, \nonumber \\ \underline{\epsilon }(p) &=& 1+q^{-2} \nonumber \end{eqnarray} \noindent the antipodes \begin{eqnarray} \underline{S}(a) &=& q^2(p-a), \nonumber \\ \underline{S}(b) &=& -q^2b, \\ \underline{S}(c) &=& -q^2c, \nonumber \\ \underline{S}(p) &=& p \nonumber \end{eqnarray} \noindent together with the braidings \begin{eqnarray}\label{eq:firstbraidings} \psi (a\underline{\otimes } a) &=& a\underline{\otimes } a+(1-q^2)b\underline{\otimes } c, \nonumber \\ \psi (a\underline{\otimes } b) &=& b\underline{\otimes } a, \nonumber \\ \psi (a\underline{\otimes } c) &=& c\underline{\otimes } a+(q^2-q^{-2})a\underline{\otimes } c+(1-q^2)p\underline{\otimes } c, \nonumber \\ \psi (b\underline{\otimes } b) &=& q^2b\underline{\otimes } b, \\ \psi (c\underline{\otimes } b) &=& q^{-2}b\underline{\otimes } c, \nonumber \\ \psi (b\underline{\otimes } c) &=& (-1-q^{-2}+q^{-4}+q^{-6})a\underline{\otimes } a+(q^2-1-q^{-2}+q^{-4})b\underline{\otimes } c \nonumber \\ & & +q^{-2}c\underline{\otimes } b+ (1-q^{-4})a\underline{\otimes } p+(1-q^{-4})p\underline{\otimes } a+(q^{-2}-1)p\underline{\otimes } p \nonumber \end{eqnarray} \noindent and their $\underline{*}$-involutions define a braided Hopf $\underline{*}$-algebra. For the second solution the coproducts \begin{eqnarray}\label{eq:secondtype} \underline{\Delta }(a) &=& a\underline{\otimes } a+q^4c\underline{\otimes } b, \nonumber \\ \underline{\Delta }(b) &=& -q^2a\underline{\otimes } b+b\underline{\otimes } a+q^2p\underline{\otimes } b, \\ \underline{\Delta }(c) &=& -q^2c\underline{\otimes } a+a\underline{\otimes } c+q^2c\underline{\otimes } p, \nonumber \\ \underline{\Delta }(p) &=& (1+q^2)a\underline{\otimes } a+q^2b\underline{\otimes } c+q^4c\underline{\otimes } b -q^2p\underline{\otimes } a-q^2a\underline{\otimes } p+q^2p\underline{\otimes } p \nonumber \end{eqnarray} \noindent the counits \begin{eqnarray} \underline{\epsilon }(a) &=& 1, \nonumber \\ \underline{\epsilon }(b) &=& 0, \\ \underline{\epsilon }(c) &=& 0, \nonumber \\ \underline{\epsilon }(p) &=& 1+q^{-2} \nonumber \end{eqnarray} \noindent and the antipodes \begin{eqnarray} \underline{S}(a) &=& -q^{-2}a+p, \nonumber \\ \underline{S}(b) &=& -q^{-2}b, \\ \underline{S}(c) &=& -q^{-2}c, \nonumber \\ \underline{S}(p) &=& p \nonumber \end{eqnarray} \noindent together with the braidings \begin{eqnarray} \psi (a\underline{\otimes } a) &=& a\underline{\otimes } a+(q^4-q^2)c\underline{\otimes } b, \nonumber \\ \psi (a\underline{\otimes } b) &=& b\underline{\otimes } a+(q^{-2}-q^2)a\underline{\otimes } b+(q^2-1)p\underline{\otimes } b, \nonumber \\ \psi (a\underline{\otimes } c) &=& c\underline{\otimes } a \nonumber \\ \psi (b\underline{\otimes } b) &=& q^{-2}b\underline{\otimes } b, \\ \psi (b\underline{\otimes } c) &=& q^2c\underline{\otimes } b, \nonumber \\ \psi (c\underline{\otimes } b) &=& (q^2+1-q^{-2}-q^{-4})a\underline{\otimes } a+q^2b\underline{\otimes } c+(q^4-q^2+q^{-2}-1)c\underline{\otimes } b \nonumber \\ & & +(q^{-2}-q^2)a\underline{\otimes } p+(q^{-2}-q^2)p\underline{\otimes } a+(q^2-1)p\underline{\otimes } p \nonumber \end{eqnarray} \noindent and their $\underline{*}$-involutions makes $B(R)$ a braided Hopf $\underline{*}$-algebra. Note that in the $q\rightarrow 1$ limit, not only the algebra becomes commutative but also the braided tensor product becomes the ordinary tensor product, and we obtain the bosonic statistics for both braided Hopf algebra solutions. When we express the solutions in terms of the original generators of the algebra, i.e., in terms of $a,b,c,d$ by using (\ref{eq:p}) we see that that the first solution is completely equivalent to the braided Hopf algebra given in the literature which we give in (\ref{eq:cop1})-(\ref{eq:braiding1}). But the second solution with the coproducts \begin{eqnarray}\label{eq:cop3} \underline{\Delta }(a) &=& a\underline{\otimes } a+q^4c\underline{\otimes } b, \nonumber \\ \underline{\Delta }(b) &=& (1-q^2)a\underline{\otimes } b+b\underline{\otimes } a+q^2d\underline{\otimes } b, \nonumber \\ \underline{\Delta }(c) &=& (1-q^2)c\underline{\otimes } a+a\underline{\otimes } c+q^2c\underline{\otimes } d, \\ \underline{\Delta }(d) &=& (q^2-1)a\underline{\otimes } a+(q^4-q^2)c\underline{\otimes } b+q^2b\underline{\otimes } c +(1-q^2)a\underline{\otimes } d+(1-q^2)d\underline{\otimes } a+q^2d\underline{\otimes } d \nonumber \end{eqnarray} \noindent counits \be \label{eq:counit3} \underline{\epsilon }(a)=\underline{\epsilon }(d)=1,\ \ \ \underline{\epsilon }(b)= \underline{\epsilon }(c)=0, \ee \noindent antipodes \be \underline{S}(a)=d,\ \ \underline{S}(b)=-q^{-2}b,\ \ \underline{S}(c)=-q^{-2},\ \ \underline{S}(d)=q^2d+(1-q^2)a \ee \noindent and braidings \begin{eqnarray}\label{eq:bra2} \psi (a\underline{\otimes } a)&=&a\underline{\otimes } a+(q^4-q^2)c\underline{\otimes } b, \nonumber \\ \psi (a\underline{\otimes } b)&=&b\underline{\otimes } a+ (1-q^2)a\underline{\otimes } b+(q^2-1)d\underline{\otimes } b, \nonumber \\ \psi (a\underline{\otimes } c)&=&c\underline{\otimes } a , \nonumber \\ \psi (a\underline{\otimes } d)&=&d\underline{\otimes } a+(1-q^2)c\underline{\otimes } b , \nonumber \\ \psi (b\underline{\otimes } a)&=&a\underline{\otimes } b , \nonumber \\ \psi (b\underline{\otimes } b)&=&q^{-2}b\underline{\otimes } b, \nonumber \\ \psi (b\underline{\otimes } c)&=&q^2c\underline{\otimes } b, \nonumber \\ \psi (b\underline{\otimes } d)&=&d\underline{\otimes } b, \\ \psi (c\underline{\otimes } a)&=&a\underline{\otimes } c +(1-q^2)c\underline{\otimes } a+(q^2-1)c\underline{\otimes } d, \nonumber \\ \psi (c\underline{\otimes } b)&=&(q^2-1)a\underline{\otimes } a+q^2b\underline{\otimes } c+(q^4-q^2+q^{-2}-1)c\underline{\otimes } b \nonumber \\ & &+(1-q^2)a\underline{\otimes } d+(1-q^2)d\underline{\otimes } a+(q^2-1)d\underline{\otimes } d, \nonumber \\ \psi (c\underline{\otimes } c)&=&q^{-2}c\underline{\otimes } c, \nonumber \\ \psi (c\underline{\otimes } d)&=&d\underline{\otimes } c+(1-q^{-2})c\underline{\otimes } a+(q^{-2}-1)c\underline{\otimes } d, \nonumber \\ \psi (d\underline{\otimes } a)&=&a\underline{\otimes } d+(1-q^2)c\underline{\otimes } b, \nonumber \\ \psi (d\underline{\otimes } b)&=&b\underline{\otimes } d+(1-q^{-2})a\underline{\otimes } b+(q^{-2}-1)d\underline{\otimes } b, \nonumber \\ \psi (d\underline{\otimes } c)&=&c\underline{\otimes } d, \nonumber \\ \psi (d\underline{\otimes } d)&=&d\underline{\otimes } d+(1-q^{-2})c\underline{\otimes } b \nonumber \end{eqnarray} \noindent defines a different braided Hopf algebra. \section{A New Algebra and its Transmutation} We find that the nonbraided Hopf algebra generated by the generators $a,b,c,d$ and $1$ whose coalgebra is of the form (\ref{eq:cop3})-(\ref{eq:counit3}) is a two parameter ($q,r$) deformed Hopf algebra. The algebra part is found to satisfy \begin{eqnarray} a\cdot b &=& rb\cdot a, \nonumber \\ a\cdot c &=& rc\cdot a, \nonumber \\ b\cdot c &=& c\cdot b, \\ b\cdot d &=& rd\cdot b+(q^{-2}-1)(r^2-1)b\cdot a \nonumber \\ c\cdot d &=& rd\cdot c+(q^{-2}-1)(r^2-1)c\cdot a, \nonumber \\ a\cdot d-d\cdot a &=& (r-r^{-1})q^2b\cdot c, \nonumber \\ q^2d\cdot a+(1-q^2)a\cdot a-r^{-1}q^4c\cdot b &=& 1.\nonumber \end{eqnarray} \noindent Finally the antipodes \be S(a)=(1-q^2)a+q^2d,\ S(b)=-rb,\ S(c)=-r^{-1}c,\ S(d)=(2-q^2)a+(q^2-1)d \ee \noindent completes Hopf algebra. For the $*$-algebra structure we find \be a^{*}=(1-q^2)a+q^2d,\ \ b^{*}=-r^{-1}c,\ \ c^{*}=-rb,\ \ d^{*}=(2-q^2)a+(q^2-1)d. \ee The adjoint coaction of this Hopf algebra on itself calculated by using (\ref{eq:adaction}) \begin{eqnarray}\label{eq:coaction} \beta (a) &=& a\otimes [(1-q^2)a\cdot a+q^2d\cdot a-r^{-1}q^4(1-q^2)c\cdot b]+b\otimes [-r^{-1}q^4c\cdot a] \nonumber \\ & & +c\otimes [q^4(1-q^2)a\cdot b+q^6d\cdot b]+d\otimes [-r^{-1}q^6c\cdot b],\nonumber \\ \beta (b) &=& a\otimes [-q^2a\cdot b]+b\otimes [a^2]+c\otimes [-q^4rb\cdot b]+d\otimes [q^2a\cdot b] \\ \beta (c) &=& a\otimes [q^4d\cdot c+q^2(1-q^2)a\cdot c]+c\otimes [(1-q^2)^2a\cdot a+q^2(1-q^2)a\cdot d+q^2(1-q^2)d\cdot a+q^4d\cdot d]\nonumber \\ & & b\otimes [-q^4r^{-1}c\cdot c]+d\otimes [q^2(q^2-1)r^{-1}c\cdot a+q^4r^{-1}c\cdot d]\nonumber \\ \beta (d) &=& a\otimes [(q^6-q^4)r^{-1}c\cdot b-q^4rc\cdot b]+c\otimes [(q^6-q^4-q^4r^2)d\cdot b+(q^4-q^2)(r^2-q^2+1)a\cdot b]\nonumber \\ & & b\otimes [(q^2(1-q^2)r^{-1}+q^2r)c\cdot a]+d\otimes [(1-q^2)a\cdot a+q^2a\cdot d-(q^6-q^4)r^{-1}c\cdot b]\nonumber \end{eqnarray} \noindent does not define an algebra homomorphism. However when the coacted copy satisfies the braided algebra (\ref{eq:braidedgrouprelations}), the transformations (\ref{eq:coaction}) define an algebra homomorphism for $r=q$. It can be shown that the braiding of the transformation in the braided tensor product space is equal to the transformation of the braiding on the same space, i.e., \be \psi (\beta (x\underline{\otimes }y))=\beta (\psi (x\underline{\otimes }y))\ \ \ \forall x,y\in B(R) \ee \noindent where the braidings are given by (\ref{eq:bra2}). It can also be shown that the braided algebra is a comodule algebra under the adjoint coaction for both of the braided Hopf algebra solutions. The transmuted multiplication ($\underline{\cdot }$) in terms of the multiplication ($\cdot$) of the coacting nonbraided algebra \begin{eqnarray} a\underline{\cdot } a &=&a\cdot a \nonumber \\ a\underline{\cdot } b &=&q^{-1}b\cdot a \nonumber \\ a\underline{\cdot } c &=&a\cdot c \nonumber \\ a\underline{\cdot } d &=&a\cdot d+(q-q^3)b\cdot c \nonumber \\ b\underline{\cdot } a &=&qb\cdot a \nonumber \\ b\underline{\cdot } b &=&q^{-1}b\cdot b \nonumber \\ b\underline{\cdot } c &=&qb\cdot c \\ b\underline{\cdot } d &=&q^{-1}b\cdot d+q(1-q^{-2})^2b\cdot a \nonumber \\ c\underline{\cdot } a &=&q^{-2}a\cdot c \nonumber \\ c\underline{\cdot } b &=&qc\cdot b+(q^{-2}-1)(d-a)\cdot a \nonumber \\ c\underline{\cdot } c &=&q^{-1}c\cdot c \nonumber \\ c\underline{\cdot } d &=&q^{-1}c\cdot d+(1-q^{-2})a\cdot c \nonumber \\ d\underline{\cdot } a &=&d\cdot a \nonumber \\ d\underline{\cdot } b &=&d\cdot b+(q^{-2}-q^{-4})a\cdot b \nonumber \\ d\underline{\cdot } c &=&d\cdot c \nonumber \\ d\underline{\cdot } d &=&d\cdot d+(q-q^{-1})b\cdot c \nonumber \end{eqnarray} \noindent completes the transmutation process.