content
stringlengths
1
15.9M
\section{Introduction} \label{intro} Over the past several years it has become widely accepted that the global star formation rate --- at least as seen in the optical and ultraviolet light --- had a strong peak at around $z = 1$\ and then fell very steeply at lower redshifts (e.g.\ Pozzetti, Madau \& Dickinson 1998 and references therein). This would mean that most of the integrated total star formation producing optically visible light would have occurred at roughly half the present age of the Universe. However, the argument for the steep fall-off at $z < 1$\ relies on a single important analysis by \markcite{lilly96}Lilly et al.\ (1996) based on the Canada France Redshift Survey (CFRS) sample. This sample has two weaknesses when it is used to determine the evolution of rest-frame UV light. The first of these is that it is a red ($I$\ band) selected sample with $V$\ and $I$\ photometry, primarily, and only partial $B$\ and $K$\ coverage. At low redshifts this requires a very substantial extrapolation across the $4000~{\rm \AA}$\ break to obtain a $2800~{\rm \AA}$\ rest-frame luminosity. The second weakness is that the CFRS is slightly too shallow for this problem ($I_{AB} = 22.5$, or $I \sim 22.1$\ in the Kron-Cousins system). At the highest redshifts, near $z = 1$, the sample does not probe deep enough in the luminosity function to allow a reliable extrapolation to a total luminosity density. In this paper we utilize a large, extremely deep, and highly complete spectroscopic redshift survey of galaxies observed in seven colors ($U^{\prime}$\ [$3400\pm 150~{\rm \AA}$], $B$, $V$, $R$, $I$, $J$, $HK^{\prime}$; the limits in the blue and red are $B_{AB} = 24.75$\ and $I_{AB} = 23.5$, respectively) in the Hawaii Survey Fields and the Hubble Deep Field to investigate the rest-frame UV luminosity density evolution from $z = 1$\ to the present. With this sample we are able to avoid the above problems encountered in the CFRS study. First, the availability of both ultraviolet and blue data means that we can {\it select\/} objects based on their rest-frame ultraviolet magnitudes at all redshifts and hence avoid the serious problem of selecting at redder wavelengths and extrapolating to obtain the UV colors. Second, the substantial additional depth of our sample allows us to probe to the flat segments of the luminosity function, thereby giving a more accurate determination of the luminosity density function. It also enables us to extend the results to redshifts beyond $z = 1$. This uniform sample selection strongly minimises possible errors due to dust in the galaxies, unless there is substantial evolution of galaxy dust properties with redshift. Thus, the shape of the present UV luminosity density evolution should be an accurate representation of the shape of the star formation history to within the one residual uncertainty of differential dust evolution. We find that the $z < 1$\ UV luminosity density falls with a shallower slope --- roughly $(1 + z)^{1.5}$\ for $q_0 =0.5$\ --- than reported by Lilly et al.\ (1996). Our low redshift UV luminosity density value is considerably higher than that found by Lilly et al.\ and agrees with the local UV luminosity density recently determined by \markcite{treyer}Treyer et al.\ (1998). Consquently, much of the integrated total of optically visible stars is still being assembled at the present time. The outline of the paper is as follows. In \S 2 we provide a brief, self-contained analysis of the distribution of light with redshift in the Hubble Deep Field. If this small field is representative, then the analysis can be used to demonstrate the primary conclusion of the paper, namely that the fall-off in the rest-frame ultraviolet luminosity density below $z = 1$\ is relatively shallow. We then proceed to the main analysis. We discuss the data in \S 3; the photometry and the statistical and systematic errors are presented in an Appendix for textual clarity. In \S 4 we construct the rest-frame ultraviolet luminosity functions and the UV luminosity density as a function of redshift. We then present comparisons with the results of other analyses. In order to be fully self-consistent, we specifically restrict ourselves to the evolution of the ultraviolet luminosity density and do not attempt to compare with studies of emission-line evolution. In \S 5 we briefly summarize our conclusions. \section{Preliminary Digression} \label{prelim} The Hubble Deep Field (HDF; \markcite{will96}Williams et al.\ 1996) provides our deepest view of the faint galaxy populations. The color information in the HDF has been extensively used to estimate the high-$z$\ star formation history of optically-selected galaxies (e.g.\ \markcite{mad97}Madau et al.\ 1997; \markcite{con}Connolly et al.\ 1997; \markcite{sawicki}Sawicki, Lin \& Yee 1997). As large numbers of spectroscopic redshifts have accumulated for HDF galaxies (\markcite{steid}Steidel et al.\ 1996; \markcite{cohen}Cohen et al.\ 1996; \markcite{low}Lowenthal et al.\ 1997; \markcite{phillips}Phillips et al.\ 1997; \markcite{song97}Songaila 1997; \markcite{barg}Barger et al.\ 1999), it is now possible to use these spectroscopic data to examine the history of star formation at $z < 1$\ and to show directly that, within the HDF itself, a substantial fraction of the star formation is occurring at low redshift. A similar conclusion is reached by \markcite{pasc98}Pascarelle et al.\ (1998) based on a very different photometric redshift analysis. The sky surface brightness of the integrated galaxy light computed from the \markcite{will96}Williams et al.\ objects in the deeper areas of the WFC chips --- an area of $4.9~{\rm arcmin}^2$ --- is shown by the solid line in Figure~\ref{fig:1}. There are 96 objects in this region of the HDF that now have spectroscopic identifications. These identified objects, which are generally the brighter galaxies, contain most of the optical and roughly half of the $3000~{\rm \AA}$\ light. We have subdivided the contributions by redshift, showing the light from known $z < 1$\ galaxies as the solid boxes and that from known $z < 0.5$\ galaxies as solid diamonds. The dashed line, which shows the combined light from all known $z > 1$\ galaxies and {\it all\/} unidentified or unobserved objects, represents an extreme upper bound to the fraction of light arising from galaxies at $z > 1$. Since some (possibly substantial) fraction of the fainter unidentified galaxies will lie at $z < 1$, the ratio of the $z > 1$\ light to that at lower redshift is overestimated here but can already be used to make the argument that a large fraction of star formation is local. As was pointed out in \markcite{cowie88}Cowie (1988), \markcite{song90}Songaila et al.\ (1990), and \markcite{mad97}Madau et al.\ (1997), the integrated massive star formation, which can be characterized by the local metal density, is directly proportional to the observed-frame ultraviolet sky brightness, independent of cosmology, through the relation \begin{equation} \rho_{\rm metals} = {{4\pi\,S_{\nu}} \over {\varepsilon\,c}} \end{equation} Estimates of the normalizing factor $\varepsilon$\ can be found in \markcite{song90}Songaila et al.\ (1990) and \markcite{mad97}Madau et al.\ (1997). Applying this relation to the $3000~{\rm \AA}$\ light in Figure~\ref{fig:1}, we see immediately that comparable amounts of metals have been produced by $z < 0.5$\ galaxies and by $z > 1$\ galaxies, since the ultraviolet sky brightness of the known $z < 0.5$\ galaxies is already 54\% of the maximum possible $z > 1$\ light. However, in this simple form, the argument depends onthe assumption that the light from star-forming galaxies is roughly flat in $f_{\nu}$. A more conservative approach is to compare the known $z < 0.5$\ $3000~{\rm \AA}$\ light with the possible $z > 1$\ $6020~{\rm \AA}$\ light. This comparison at equivalent rest-frame wavelengths allows for the shapes of the spectral energy distributions (SEDs) and for dust extinction. It also takes into account the disappearance of very high redshift galaxies from the $3000~{\rm \AA}$\ band. The ratio $R_{SF} \equiv {\rm light}(z < 0.5)/{\rm light}(z > 1)$\ from Figure~\ref{fig:1} is then 26\%. This represents a very extreme lower bound on the ratio of the integrated star formation at $z < 0.5$\ to that at $z > 1$. The implications for the evolution of the low-$z$\ SFR [$\dot{\rho}_{\ast}(z)$] can be shown with the simple model \begin{equation} \begin{array}{rcll} \dot{\rho}_{\ast}(z) &=& A\,(1 + z)^{\alpha} & ,\quad z < 1 \nonumber\\ &=& 2^{\alpha}\, A & ,\quad z > 1 \end{array} \end{equation} that can be integrated analytically. For $q_0 = 0.5$\ $R_{SF} = $18\% for $\alpha = 4$, 30\% for $\alpha = 3$\ and 48\% for $\alpha = 2$. Thus, even comparing the {\it candidate\/} $z > 1$\ visual light with the {\it known\/} $z < 0.5$\ ultraviolet light requires $\alpha \ll 3.2$. Provided only that the small HDF is a representative field, even the very conservative and extremely robust version of the argument requires a shallower slope than has been inferred from the CFRS data ({\it viz}., $3.9 \pm 0.75$). The remainder of the paper verifies this conclusion using our wide-field ultraviolet and optically-selected samples. \section{Data Sets} \label{data} The primary data sets used in the present study are two $6^{\prime} \times 2\,'\mskip-7.2mu.\, 5$\ areas crossing the HDF and the Hawaii Survey Field SSA22 (\markcite{lcg}Lilly, Cowie \& Gardner 1991). In each case the field has been imaged in 7 colors --- $U^{\prime}$ ($3400 \pm 150~{\rm \AA}$), $B$, $V$, $R$, $I$, $J$\ and $HK^{\prime}$ --- where $R$\ and $I$\ are Kron-Cousins and the $HK^{\prime}$\ ($1.9 \pm 0.4~\mu$) filter is described in \markcite{wains}Wainscoat \& Cowie (in preparation). The images were obtained on the Keck~II telescope using LRIS (\markcite{oke}Oke et al.\ 1995) and on the University of Hawaii 2.2m and the Canada-France-Hawaii 3.6m telescopes using QUIRC (\markcite{hodapp}Hodapp et al.\ 1996), ORBIT, and the UH8K CCD Mosaic Camera built by Metzger, Luppino, and Miyazaki. The fields, whose sizes were set by the usable LRIS area, are fully covered in all colors. All magnitudes were measured in $3^{\prime\prime}$\ diameter apertures and corrected to total magnitudes following the procedures in \markcite{cowie94}Cowie et al.\ (1994). A further two similarly sized areas covering the Hawaii Survey Fields SSA4 (\markcite{cowie94}Cowie et al.\ 1994) and SSA13 (\markcite{lcg}Lilly, Cowie \& Gardner 1991) with $B,V,I,HK^{\prime}$ data were used to augment the SSA22 and HDF $B$, $I$, and $HK^{\prime}$-selected samples. The fields, areas, magnitude limits, and numbers of spectroscopically-identified galaxies, stars, and unknowns (unobserved or unidentified objects) for each of the color-selected samples are given in Table~\ref{tbl:1}. A more detailed discussion and analysis of the photometric and spectroscopic samples used in this paper can be found in Barger et al.\ (1999). \section{Rest-Frame Ultraviolet Luminosity Density Evolution} \label{uvlum} \subsection{Construction of the luminosity functions} For redshifts, $z$, for which the rest-frame UV wavelength, $\lambda_U$, maps directly to the wavelength of one of the observed color bands, the absolute rest-frame UV magnitude, $M_U$, in the $AB$\ system ($M_U = 0$\ corresponds to $\log_{10}\,F_{\nu} = -48.57$) is given by \begin{equation} M_U = m_{AB} + 2.5\,\log_{10}(1+z) - 5\,\log_{10}(d_L(z)/10\,{\rm pc}) \end{equation} where $m_{AB}$\ is the observed magnitude at the redshifted wavelength $\lambda_U(1+z)$\ in the $AB$\ system and $d_L(z)$\ is the luminosity distance. For redshifts such that $\lambda_U(1+z)$\ differs by a small amount from this central wavelength, we can apply the same formula but with the addition of a small differential $k$\ correction $dk(z)$, defined as \begin{equation} dk(z) = -2.5\,\log_{10}\,{{f_{\nu}\,(\lambda_U(1+z))} \over {f_{\nu}\,(\lambda_U(1+z_C))} } \end{equation} where $f_{\nu}$\ is the SED of the galaxy, and $z_C$\ is the redshift corresponding to the center of the band. Because of the frequent, regularly-spaced color sampling, we can simply construct $dk(z)$\ by interpolation from the neighboring color bands. The value of $dk(z)$\ is generally small (less than 0.2~mag) for the redshift intervals which are used. The $2500~{\rm \AA}$\ rest-frame absolute magnitudes, computed in the appropriate redshift intervals from the $U^{\prime}$, $B$, $V$\ and $I$\ samples, are shown versus redshift in Figure~\ref{fig:2}. In the overlapping redshift intervals the absolute magnitudes compare well. The rise in the maximum absolute UV rest-frame luminosity with redshift mapped by the upper envelope of the distribution is partly a reflection of the larger volume sampled at high redshift but also partly reflects the known evolution that causes the maximum UV luminosity to increase with redshift (\markcite{lilly96}Lilly et al.\ 1996; \markcite{cowie94}Cowie et al.\ 1996); we will discuss this in more detail below. We adopt the traditional $V_{max}$\ method of \markcite{felten}Felten (1977) for constructing the luminosity functions. The number density of galaxies in the redshift range [$z_1,z_2$] with magnitude $M$\ is given by \begin{equation} \phi(M)\,dM \propto \sum\, {{1} \over {V_{max}(M)}} \end{equation} where the sum is over all galaxies with magnitude $M \pm dM/2$. $V_{max}(M)$\ is the maximum total volume in all the samples where galaxies with absolute magnitude $M$ are observable in the appropriate apparent magnitude range and lie in the redshift range. (A very complete description of the procedure may be found in \markcite{ellis96}Ellis et al.\ 1996). The ultraviolet luminosity density, $\ell$, is then \begin{equation} \ell = 4.4 \times 10^{20}\, \sum {{10^{-0.4M}} \over {V_{Max}(M)}}\quad {\rm ergs\ s^{-1}\ Hz^{-1}} \end{equation} where the sum is over all the observed objects. We use Schechter function fits with $\alpha = -1$\ to extend $\ell$\ from the brighter absolute magnitudes observed to $M = -16$, which we then use as our total luminosity density. In the computation of the luminosity functions, we assign errors based on the Poisson distribution corresponding to the number of galaxies contributing to the absolute magnitude bin. However, a larger potential source of error is the missing or unidentified galaxies in each sample. Following the standard procedure, we compute the luminosity function after assuming that these objects follow the redshift distribution of the identified objects, but because of systematic effects, this assumption may not be valid. We have therefore also computed the LFs with all, and also with none, of the missing objects allocated into the redshift bin, which should provide extremal estimates of this error. In the case where we allocated the missing objects into the redshift bin, the allocated redshifts were randomly distributed uniformly within the redshift interval. We are able to use this very robust procedure because of the extremely high completeness of the spectroscopic samples. \subsection{UV luminosity functions versus rest wavelength} A key issue is the optimum wavelength at which to compute the rest-frame LFs and the UV light densities. Previous efforts, such as those by the CFRS, have computed the light densities at a relatively long wavelength of $2800~{\rm \AA}$, which is optimal for a red sample and minimises dust extinction but may be a poorer measure of massive star formation rates than shorter wavelengths. The local UV sample of \markcite{treyer}Treyer et al.\ (1998), which provides an invaluable low redshift comparison, is at $2000~{\rm \AA}$. However, our own sample gives the best combination of wide redshift range and largest galaxy sample at a rest wavelength around $2500~{\rm \AA}$. In the following discussion we shall use all three of these rest wavelengths for specific comparisons and analyses. Before proceeding, however, we use the various color samples to show that at $z = 1$\ the luminosity functions and light densities are only weakly sensitive to the wavelength choice. At $z = 1$\ the $U^{\prime}$, $B$\ and $V$\ samples correspond to rest-wavelengths of $1700~{\rm \AA}$, $2250~{\rm \AA}$\ and $2750~{\rm \AA}$. In Figure~\ref{fig:3} we compare the luminosity functions constructed at each of these rest wavelengths in the redshift interval $0.7 < z < 1.3$. It is apparent that the luminosity functions are quite similar, with only a slight fading of $M_U$\ with decreasing wavelength. The corresponding luminosity densities for galaxies with luminosities greater than $M_U = -17.5$\ (squares), or extrapolated to $M = -16$\ assuming an $\alpha = -1$\ Schechter function (dashed line), are shown in Figure~\ref{fig:4}. The total UV luminosity density above $M = -16$\ in this redshift interval is fit by a power law \begin{equation} \ell = 2.6\times 10^{26}\,h_{65}\,\left( {{\lambda} \over {3000~\rm\AA}}\right)^{1.1}\,\quad {\rm ergs\ Mpc^{-3}\ Hz^{-1}} \end{equation} where $h_{65}$\ is the Hubble constant in units of $65~{\rm km\ s^{-1}\ Mpc^{-1}}$. We shall generally avoid comparing light densities at different wavelengths. Where necessary, however, we use the weak wavelength dependence of equation\ (7), assuming, somewhat arbitrarily, that this is valid for other redshifts. \subsection{The redshift evolution of the $2500~{\rm \AA}$\ rest-frame luminosity density} We first construct the $2500~{\rm \AA}$\ rest-frame luminosity function in three redshift bands: $z = 0.20 - 0.50$\ using the $U^{\prime}$-selected sample (Figure~\ref{fig:5}), $z = 0.6 - 1.0$\ using the $B$-selected sample (Figure~\ref{fig:6}), and $z = 1.0 - 1.5$\ using the $V$-selected sample (Figure~\ref{fig:7}). In each case the solid curve is the incompleteness-corrected luminosity function with $\pm 1~\sigma$\ Poisson error bars on the symbols; the dotted curve is the luminosity function obtained from only the observed objects (the minimal function); and the dashed curve is the luminosity function when all the unidentified objects are taken to lie within the redshift interval (the maximal function). We next compute the luminosity density for galaxies more luminous than $M_U = -16$. This is a directly determined quantity in the lowest redshift interval, but for the higher redshift intervals it requires an extrapolation which was made with an $\alpha = -1$\ Schechter function fit. For $z = 0.6 - 1.0$ the observed light density for $M_{AB} \le -17.5$\ is $1.7\times 10^{26}\,h_{65}\ {\rm ergs\ s^{-1}\ Hz^{-1}\ Mpc^{-3}}$, and the extrapolated light density is 21\% higher. For $z = 1.0 - 1.5$\ the observed light density for $M_{AB} \le -18.25$\ is $1.4\times 10^{26}\,h_{65}\ {\rm ergs\ s^{-1}\ Hz^{-1}\ Mpc^{-3}}$, and the extrapolated light density is 34\% higher. The incompleteness-corrected $M_{AB} \le -16$\ light density as a function of redshift is shown in Figure~\ref{fig:8} as the filled squares; the redshift range is shown as the thin horizontal lines. The statistical and systematic errors (see Appendix) are shown as the thicker portions of the error bar. We have performed similar calculations for both the minimal and maximal functions. These represent the extremal range of the possible light density at a given redshift and are shown as the open squares joined by the vertical thin lines. The solid line shows a simple $(1 + z)^{1.5}$\ evolution law which would result in equal amounts of star formation per unit redshift interval and would constitute an acceptable fit to the data. The dashed line shows a redshift evolution of $(1 + z)^4$\ matched to the second data point; this is clearly too steep, even allowing for the maximum incompleteness correction. The shape and normalization of the rest-frame UV luminosity function is of course a consequence of a complex mix of the evolution of the star formation rates in galaxies of various types and masses and the decrease in $M_{U\ast}$\ with declining redshift, which most likely reflects a preferential drop in the star formation rates of the most massive galaxies (\markcite{cowie96}Cowie et al.\ 1996). Nevertheless, it is interesting to compare the functions to determine if there is evidence for a change of shape (Figure~\ref{fig:9}). For $q_0 = 0.5$\ the functions can be quite well overlayed with a simple shift in the absolute magnitude (pure luminosity evolution). Though there is a hint that the higher redshift functions might be steeper, there is no statistically significant difference in the shapes as a function of redshift. The absolute magnitude shifts used in constructing Figure~\ref{fig:9} (+0.65~mag at $z = 0.6 - 1.0$\ and +0.75~mag at $z = 1.0 - 1.5$, both relative to $z = 0.2 - 0.5$) can also be used to check the $2500~{\rm \AA}$\ rest-frame luminosity density evolution and would imply a ratio of 1 : 1.8 : 2.0 in these redshift bins; the best-fit power law corresponds to $\ell \propto (1 + z)^{1.4}$, which is quite similar to the estimates given above. \subsection{Comparison with other samples} The most directly comparable sample is the low-redshift, rest-frame ultraviolet ($2000~{\rm \AA}$) selected sample of Treyer et al.\markcite{treyer} (1998). To make a direct comparison with this sample we have computed the $2000~{\rm \AA}$\ luminosity functions at $z = 0.5 - 0.9$\ using the $U^{\prime}$-selected sample and at $z = 1.0 - 1.5$\ using the $B$-selected sample. We compare these with the luminosity function determined by Treyer et al.\ in Figures~\ref{fig:10} and \ref{fig:11}. The solid line with the symbols shows the presently-determined luminosity function. The dashed line shows the Treyer et al.\ function for ${\rm H}_0 = 65~{\rm km\ s^{-1}\ Mpc^{-1}}$\ and converting the magnitude system used by Treyer et al.\ to $AB$\ magnitudes with a 2.29~mag offset. Here we have used a renormalization rather than a magnitude offset to overlay the functions. For $z = 0.5 - 0.9$\ the normalization of the Treyer et al.\ function is multiplied by 1.7. For $z = 1.0 - 1.5$\ the normalization is multiplied by 2.8. The corresponding luminosity density changes would be an increase of $\ell \propto (1+z)^{1.3}$. The measured $2000~{\rm \AA}$\ luminosity densities in the Treyer et al.\ sample are compared with the incompleteness-corrected $2000~{\rm \AA}$\ luminosity densities from this work in Figure~\ref{fig:12}, where once again we have shown the maximal and minimal luminosity densities with open squares. As was the case for the $2500~{\rm \AA}$\ data, the combined data sets exhibit a slow evolution, as shown by the $(1 + z)^{1.5}$\ solid line. A steeper dependence, such as the $(1 + z)^4$\ evolution of the dashed line, is radically inconsistent with the data. Because of the steep rise in the Treyer et al.\ luminosity function at the faintest magnitudes, it is best fit by an $\alpha = -1.6$\ power law rather than the $\alpha = -1.0$\ power law (characteristic of the optical luminosity functions) which we have adopted in extrapolating the luminosity functions to the fainter magnitudes. In order to investigate the dependence on $\alpha$\ we have fitted the $2000~{\rm \AA}$\ luminosity functions at $z = 0.5 - 0.9$\ and $z = 1 - 1.5$\ with Schechter functions with indices $\alpha = -1.0$\ and $\alpha = -1.5$. The fitted parameters and the luminosity densities above $-16$\ are summarized in Table~\ref{tbl:2} where they are compared with the values derived by Treyer et al. Increasing the index from $-1.0$\ to $-1.5$\ has the effect of preferentially increasing the luminosity density at the higher redshifts where the extrapolation is larger. However, the effect is not large. Using the $-1.5$\ derived luminosity density would increase the slope to $(1+z)^{1.7}$\ from $(1+z)^{1.3}$\ for the $\alpha = -1.0$\ case. We next compare the present sample with the \markcite{lilly96}Lilly et al.\ (1996) analysis of the CFRS data at a rest-frame wavelength of $2800~{\rm \AA}$. In order to make the most direct comparison possible, we followed the Lilly et al.\ redshift intervals of [0.2,0.5], [0.5,0.75], [0.75,1], constructing the luminosity functions with the $U^{\prime}$\ sample, the $B$\ sample, and the $V$\ sample, respectively. The luminosity densities constructed from these samples are shown in Figure~\ref{fig:13}, where the filled squares are again the incompleteness-corrected luminosity density, the thicker portions of the error bars show the systematic and statistical errors, and the open squares show the minimal and maximal light densities. The open diamonds show the Lilly et al.\ analysis; their three higher redshift points are based on the CFRS data and the lowest redshift point is based on the data of \markcite{love}Loveday et al.\ (1992). The lowest redshift point is already contradicted by Treyer et al.'s (1998) analysis, shown as the filled diamond, corrected upwards by a factor of 1.5 for the wavelength difference between $2000~{\rm \AA}$\ and $2800~{\rm \AA}$. The Treyer et al.\ point is also comparable with the [0.2,0.5] Lilly point. The present data exhibit a much shallower slope than the Lilly et al.\ data in the $z = 0.2 - 1$\ range. This results primarily from the [0.75,1] point being about a factor of two lower than the highest redshift Lilly point. The two intermediate redshift points agree well with the CFRS analysis. We also compare with the analyses of \markcite{con}Connolly et al.\ (1997) and \markcite{sawicki}Sawicki et al.\ (1997), shown as open triangles and inverted open triangles, respectively, in Figure~\ref{fig:13}. These two analyses are based on photometrically estimated redshifts in the small HDF proper and are in broad general agreement with each other and with the work of \markcite{pasc98}Pascarelle et al.\ (1998). (We do not include a direct comparison to the latter work here as it is computed at $1500~{\rm \AA}$.) Following the discussion in the Appendix, we have made a small 20\% downward correction to allow for the slightly higher counts in the relevant magnitude range in the HDF proper versus the wide surrounding field. Both the [0.5,1] and [1.0,1.5] points from these analyses are higher than our incompleteness-corrected estimate but are consistent within the statistical and incompleteness errors. Our best estimate value may indeed be slightly low here, since there may be preferential incompleteness in these redshift ranges. However, there is also relatively little validation of photometrically estimated redshifts in the range $z = 1$\ to 2 where the observed spectral range ($\lambda = 3000~{\rm \AA}$\ to $22,000~{\rm \AA}$) is quite featureless for blue galaxies. Thus, the high redshift Connolly et al.\ and Sawicki et al.\ points could be overestimated if lower redshift blue irregulars have been assigned to the wrong bin. Since our maximal incompleteness-corrected estimate is not substantially different from the photometrically estimated values, the discussion of \S 5 does not depend on the resolution of this issue. \subsection{The relative evolution of the rest-frame UV, blue and red light} As has been known for several years now (Lilly et al.\markcite{lilly96} 1996; Cowie et al.\markcite{cowie96} 1996) the rest-frame red luminosity has an extremely weak evolution with redshift out to $z = 1$. This slow evolution can be clearly seen in Figure~\ref{fig:14}, in which we compare the rest-frame $8000~{\rm \AA}$\ luminosity at $z = 0.2 - 0.5$, obtained using the $I$-selected sample, with that at $z = 0.7 - 1.3$, computed using the $K$-selected sample. The two functions can be brought into full consistency with a very small (0.2~mag) pure luminosity shift corresponding to a decrease of only 20\% in the red light from $z = 0.8$\ to $z = 0.35$, thereby confirming with the present sample that the downward evolution of the UV light density by a factor of two between $z = 1$\ and $z = 0.35$\ does take place against an essentially invariant shape and normalization for the red light density. \section{Summary} We have investigated the evolution of the universal ultraviolet luminosity density from $z=1$ to the present using a magnitude-limited sample selected on the basis of rest-frame ultraviolet colors at all redshifts from an extremely deep and highly complete spectroscopic redshift survey. Our uniform selection procedure avoids the serious problem of selecting the sample at redder wavelengths and then extrapolating to obtain ultraviolet colors, and it strongly minimises possible errors due to dust in the galaxies, unless there is substantial evolution of galaxy dust properties with redshift. The depth of the current sample is also sufficiently deep to probe to the flat segments of the luminosity function and to extend the results to redshifts beyond $z=1$, thereby enabling us to make an accurate determination of the luminosity density function. We find that our incompleteness-corrected rest-frame 2500$~{\rm \AA}$ luminosity densities for $M_U\le -16$ as a function of redshift are well fit by an $l\propto (1+z)^{1.5}$ evolution law for $q_0 = 0.5$. (The slope would be shallower for open geometries.) The decline in $M_U$ with decreasing redshift most likely reflects a preferential drop in the star formation rates of the most massive galaxies. A direct comparison of the measured low-redshift 2000$~{\rm \AA}$ luminosity density of Treyer et al.\ (1998) with incompleteness-corrected 2000$~{\rm \AA}$ luminosity densities computed from our sample also shows that the slow evolution law of $l\propto (1+z)^{1.5}$\ provides a good fit. When we compare the Lilly et al.\ (1996) CFRS analysis with our sample analyzed at rest-frame 2800$~{\rm \AA}$, we find that our sample gives a much shallower slope in the $z=0.2-1$ range. This most probably arises in part due to the relatively large extrapolation at low redshifts Lilly et al.\ needed to make to go from a primarily $V$ and $I$-based data sample to ultraviolet colors. The two lowest-redshift Lilly et al.\ points are also in disagreement with the Treyer et al.\ analysis. Interestingly, however, the present UV luminosity density analysis leads to a closer agreement with star formation rates found in an H$\alpha$ analysis of the CFRS data by Tresse \& Maddox (1998). We can summarise our results in the form of the widely-discussed plot of UV luminosity density versus redshift, our version of which is shown for the $2500~{\rm \AA}$\ rest frame in Figure~\ref{fig:8} and at $2000~{\rm \AA}$\ in Figure~\ref{fig:12}. This plot is rather different from the now conventional `Madau' form which has a strong rise to $z = 1$. Rather, we see a fairly slow decline in $\dot\rho_{UV}$\ from $z = 1.5$\ to $z = 0$, which is reasonably well described by a simple $\dot\rho_{UV} = 7\times 10^{25}\,(1 + z)^{1.5}~{\rm ergs\ s^{-1}\ Mpc^{-3}\ Hz^{-1}}$\ relation. This result has fairly profound philosophical implications in that the integrated star formation is continuing to rise smoothly at the present time and the bulk of the star formation has occurred at recent times. (This result holds irrespective of cosmological geometry.) Put succinctly, as seen in the optical, now may be the epoch of galaxy formation and not $z = 1$.
\subsection*{R\'ef\'erences bibliographiques\markboth{\sc R\'ef\'erences bibliographiques}{\sc R\'ef\'erences bibliographiques}} \list {[\arabic{enumi}]}{\settowidth\labelwidth{[#1]}\leftmargin\labelwidth \advance\leftmargin\labelsep \usecounter{enumi}} \def\hskip .11em plus .33em minus -.07em{\hskip .11em plus .33em minus -.07em} \sloppy \sfcode`\.=1000\relax} \newtheorem{thm}{Theorem} \newtheorem{prop}[thm]{Proposition} \newtheorem{fthm}{Th\'eor\`eme} \newtheorem{fdefn}[fthm]{D\'efinition} \newtheorem{fex}[fthm]{Exemple} \newcommand{\nolinebreak\hfill{$\Box$}\par\vspace{0.5\parskip}}{\nolinebreak\hfill{$\Box$}\par\vspace{0.5\parskip}} \newcommand{\nolinebreak\hfill{$\triangle$}\par\vspace{0.5\parskip}}{\nolinebreak\hfill{$\triangle$}\par\vspace{0.5\parskip}} \newcommand{\nolinebreak\hfill{$\Diamond$}\par\vspace{0.5\parskip}}{\nolinebreak\hfill{$\Diamond$}\par\vspace{0.5\parskip}} \def\tilde{U}{\tilde{U}} \def\tilde{V}{\tilde{V}} \def\tilde{f}{\tilde{f}} \def\Delta{\Delta} \def\Gamma{\Gamma} \def\Gamma_{\alpha}{\Gamma_{\alpha}} \def\Delta_{\alpha}{\Delta_{\alpha}} \def\tilde{\omega}{\tilde{\omega}} \def\ot_{\alpha}{\tilde{\omega}_{\alpha}} \def\ot_{\beta}{\tilde{\omega}_{\beta}} \def\tilde{\tau}_{\alpha}{\tilde{\tau}_{\alpha}} \def\phi_{\alpha}{\phi_{\alpha}} \def\phi_{\beta}{\phi_{\beta}} \defU_{\alpha}{U_{\alpha}} \def\ut_{\alpha}{\tilde{U}_{\alpha}} \defV_{\alpha}{V_{\alpha}} \def\vt_{\alpha}{\tilde{V}_{\alpha}} \defU_{\beta}{U_{\beta}} \defg_{\alpha\beta}{g_{\alpha\beta}} \def\mbox{\bbb{C}}{\mbox{\bbb{C}}} \def\mbox{\bbb{Q}}{\mbox{\bbb{Q}}} \def\mbox{\bbb{R}}{\mbox{\bbb{R}}} \def\mbox{\bbb{Z}}{\mbox{\bbb{Z}}} \def\R^k{\mbox{\bbb{R}}^k} \def|z|^2{|z|^2} \def|w|^2{|w|^2} \def|z_1|^2{|z_1|^2} \def|z_2|^2{|z_2|^2} \defe^{2\pi i\theta}{e^{2\pi i\theta}} \defe^{2\pi i\theta\frac{s}{t}}{e^{2\pi i\theta\frac{s}{t}}} \def\eones{e^{\frac{2\pi i\theta}{s}}} \def\mbox{\frak d}{\mbox{\frak d}} \def\mbox{\frak n}{\mbox{\frak n}} \def\mbox{\frak t}{\mbox{\frak t}} \def\d^*{\mbox{\frak d}^*} \def\n^*{\mbox{\frak n}^*} \def\t^*{\mbox{\frak t}^*} \def\mbox{X}{\mbox{X}} \markright{\sc Sur une g\'en\'eralisation de la notion de $V$-vari\'et\'e} \begin{document} \noindent {\large{\sc Sur une g\'en\'eralisation de la notion de $V$-vari\'et\'e}}\\ \noindent{\bf Elisa PRATO}\\ \noindent \small{Laboratoire Dieudonn\'e, Universit\'e de Nice, Parc Valrose, 06108 Nice Cedex 2, France\\ Courriel: {\tt <EMAIL>}\\ Web: {\tt http://www-math.unice.fr/\~{}elisa/home.html}}\\ \noindent {\bf R\'esum\'e.} {\small Nous consid\'erons un espace topologique qui est localement isomorphe au quotient de $\R^k$ par l'action d'un groupe discret et nous l'appelons {\em quasi-vari\'et\'e} de dimension $k$. Les quasi-vari\'et\'es g\'en\'eralisent les vari\'et\'es et les $V$-vari\'et\'es et repr\'esentent le cadre naturel pour la r\'eduction symplectique par rapport \`a l'action induite d'un sous-groupe de Lie, compact ou non, d'un tore. Nous d\'efinissons les {\em quasi-tores}, les actions hamiltoniennes de quasi-tores et l'application moment sur une quasi-vari\'et\'e symplectique, et nous montrons que tout polytope convexe simple, rationnel ou non, est l'image de l'application moment pour l'action d'un quasi-tore sur une quasi-vari\'et\'e.}\\ \noindent {\sc On a generalization of the notion of orbifold}\\ \noindent {\bf Abstract.} {\small We consider a topological space which is locally isomorphic to the quotient of $\R^k$ by the action of a discrete group and we call it {\em quasifold} of dimension $k$. Quasifolds generalize manifolds and orbifolds and represent the natural framework for performing symplectic reduction with respect to the induced action of any Lie subgroup, compact or not, of a torus. We define {\em quasitori}, Hamiltonian actions of quasitori and the moment mapping for symplectic quasifolds, and we show that every simple convex polytope, rational or not, is the image of the moment mapping for the action of a quasitorus on a quasifold.}\\ \subsection*{Introduction} Soit $M$ une vari\'et\'e symplectique compacte et connexe et soit $T$ un tore agissant sur $M$ de fa\c{c}on hamiltonienne. Alors, d'apr\`es le th\'eor\`eme de convexit\'e d'Atiyah et Guillemin-Sternberg \cite{a,gs}, l'image de l'application moment correspondante est un polytope convexe {\em rationnel}. Si l'action de $T$ est effective et $\dim{M}=2\dim{T}$ alors, d'\`apres un th\'eor\`eme de Delzant \cite{d}, le polytope image est un polytope rationnel simple satisfaisant \`a une certaine condition d'int\'egralit\'e, et ce polytope d\'etermine la vari\'et\'e \`a symplectomorphisme \'equivariant pr\`es. Au cours de la d\'emonstration de ce th\'eor\`eme, Delzant donne une construction explicite de la vari\'et\'e \`a partir du polytope; cette construction se base sur la technique de r\'eduction symplectique. Les m\'ethodes d'Atiyah, Guillemin-Sternberg et Delzant ont succesivement \'et\'es adapt\'ees par Lerman-Tolman \cite{lt} au cas des $V$-vari\'et\'es, mais les polytopes image, obtenus dans ce cas, sont toujours des polytopes rationnels. Il est d'ailleurs tr\`es naturel de se demander si tout polytope simple, {\em rationnel ou non}, est l'image de l'application moment pour un espace symplectique ad\'equat. Pour pouvoir r\'epondre \`a cette question nous consid\'erons des espaces topologiques qui g\'en\'eralisent les vari\'et\'es et les $V$-vari\'et\'es, mais qui ne sont plus forcement des espaces s\'epar\'es. Ces espaces, que nous appelons {\em quasi-vari\'et\'es}, permettent l'utilisation de la technique de r\'eduction symplectique sous des hypoth\`eses assez faibles; par cons\'equent, ils donnent le moyen d'\'etendre la construction de Delzant au cas des polytopes non rationnels. L'unicit\'e dans le th\'eor\`eme de Delzant n'est plus valable dans ce contexte (voir exemple~\ref{quasisphere2}). Dans le cas o\`u $\Delta$ est rationnel cette construction nous donne une famille de $V$-vari\'et\'es, en accord avec \cite{lt}; si $\Delta$ v\'erifie aussi la condition d'int\'egralit\'e de Delzant, on trouve parmi cette famille une vraie vari\'et\'e, en accord avec \cite{d}. Remarquons que l'espace des orbites de l'action d'un groupe discret sur une vari\'et\'e a \'et\'e etudi\'e par Connes \cite[chapitre II]{c} dans le cadre de la g\'eom\'etrie non commutative. Les quasi-vari\'et\'es sont aussi reli\'ees \`a la g\'eom\'etrie des quasi-cristaux \cite{se}. Nous renvoyons le lecteur \`a \cite{p1} pour les preuves des r\'esultats \'enonc\'es, ainsi que pour de nombreux exemples; dans une suite \`a \cite{p1} on trouvera un traitement plus detaill\'e des propri\'et\'es de convexit\'e de l'application moment. Finalement dans un travail en collaboration avec Battaglia \cite{bp} nous introduisons des structures complexes et des structures de K\"ahler sur les quasi-vari\'et\'es, et nous montrons que les espaces du th\'eor\`eme~\ref{toutsimple} peuvent \^etre consid\'er\'es comme la g\'en\'eralisation naturelle des vari\'et\'es toriques qu'on associe \`a tout polytope convexe simple qui est rationnel. \subsection*{D\'efinitions et r\'esultats} Commen\c{c}ons par donner la d\'efinition de mod\`ele de quasi-vari\'et\'e et de diff\'eomorphisme entre mod\`eles. \begin{fdefn}[Mod\`ele]{\rm Soit $\tilde{U}$ une vari\'et\'e connexe de dimension $k$ et soit $\Gamma$ un groupe discret agissant diff\'erentiablement sur $\tilde{U}$ de fa\c{c}on que l'ensemble des points o\`u l'action est libre, soit connexe et dense. Un {\em mod\`ele} de dimension $k$, $\tilde{U}/\Gamma$, est l'espace des orbites de l'action de $\Gamma$ sur $\tilde{U}$, muni de la topologie quotient.}\nolinebreak\hfill{$\triangle$}\par\vspace{0.5\parskip}\end{fdefn} On peut toujours supposer, en passant \'eventuellement au rev\^{e}tement universel, que $\tilde{U}$ soit simplement connexe, ce que nous ferons par la suite. \begin{fdefn}[Diff\'eomorphisme de mod\`eles]{\rm Soient $p\,\colon\tilde{U}\rightarrow\tilde{U}/\Gamma$ et $q\,\colon\tilde{V}\rightarrow\tilde{V}/\Delta$ deux mod\`eles. On dira qu'une application $f\,\colon \tilde{U}/\Gamma\longrightarrow \tilde{V}/\Delta$ est un {\em diff\'eomorphisme de mod\`eles} s'il existe un diff\'eomorphisme $\tilde{f}\,\colon\tilde{U}\longrightarrow\tilde{V}$ et un isomorphisme $F\,\colon\Gamma\rightarrow\Delta$ tels que $q\circ \tilde{f}=f\circ p$ et $\tilde{f}(\gamma\cdot-)=F(\gamma)\cdot\tilde{f}(-)$, $\gamma\in\Gamma$. Nous dirons alors que $\tilde{f}$ est un {\em relev\'e} de $f$. }\nolinebreak\hfill{$\triangle$}\par\vspace{0.5\parskip}\end{fdefn} On montre aisement que deux relev\'es d'un diff\'eomorphisme $f$ co\"{\i}ncident \`a moins d'une multiplication \`a droite par un \'el\'ement de $\Gamma$ (ou \`a gauche par un \'el\'ement de $\Delta$). Nous pouvons maintenant d\'efinir une quasi-vari\'et\'e \`a partir des deux d\'efinitions pr\'ec\'edentes. \begin{fdefn}[Quasi-vari\'et\'e] {\rm Soit $M$ un espace topologique. Un {\em atlas de quasi-vari\'et\'e} de dimension $k$ sur $M$ est une collection d'ouverts ${\cal A}= \{\; U_{\alpha}\,|\,\alpha\in A\;\}$, dits {\em cartes}, ayant les propri\'et\'es suivantes: \begin{enumerate} \item les $U_{\alpha}$ recouvrent $M$; \item pour tout $\alpha\in A$ il existe un mod\`ele $\ut_{\alpha}/\Gamma_{\alpha}$, o\`u $\ut_{\alpha}$ est un sous-ensemble ouvert, connexe et simplement connexe de $\R^k$, et un hom\'eomorphisme $\phi_{\alpha}\,\colon\ut_{\alpha}/\Gamma_{\alpha}\longrightarrowU_{\alpha}$; \item pour tous $\alpha, \beta \in A$ tels que $U_{\alpha}\capU_{\beta}\neq\emptyset$ l'application $$g_{\alpha\beta}=\phi_{\beta}^{-1}\circ\phi_{\alpha}\,\colon\phi_{\alpha}^{-1}(U_{\alpha}\capU_{\beta}) \longrightarrow\phi_{\beta}^{-1}(U_{\alpha}\capU_{\beta})$$ est un diff\'eomorphisme de mod\`eles. Nous dirons alors que $g_{\alpha\beta}$ est un {\em changement de cartes}. \end{enumerate} Deux atlas sur $M$ sont \'equivalents si l'union des leurs cartes est encore un atlas de $M$. Une {\em structure de quasi-vari\'et\'e} sur $M$ est une classe d'\'equivalence d'atlas; nous dirons alors qu'un espace $M$ avec une structure de quasi-vari\'et\'e est une {\em quasi-vari\'et\'e}. }\nolinebreak\hfill{$\triangle$}\par\vspace{0.5\parskip}\end{fdefn} Si les groupes $\Gamma_{\alpha}$ sont finis on retrouve ici la d\'efinition classique de $V$-vari\'et\'e, s'ils sont triviaux on retrouve la d\'efinition de vraie vari\'et\'e. \begin{fex}[Quasi-sph\`ere]\label{quasisphere}{\rm Soient $s,t$ deux nombres r\'eels positifs tels que $s/t\notin\mbox{\bbb{Q}}$. Consid\'erons $\mbox{\bbb{C}}^2$ avec sa forme symplectique canonique $\omega_0=\frac{1}{2\pi i}(dz_1\wedge d\bar{z}_1 + dz_2\wedge d\bar{z}_2)$ et l'action de $\mbox{\bbb{R}}$: $(\theta, (z_1,z_2)) = ( e^{2\pi i\theta} z_1, e^{2\pi i\theta\frac{s}{t}} z_2)$ d'application moment $\Psi(z_1,z_2)=|z_1|^2 +\frac{s}{t}|z_2|^2-s$. Consid\'erons le niveau r\'egulier $\Psi ^{-1} (0)$; alors l'espace des orbites $M=\Psi^{-1}(0)/\mbox{\bbb{R}}$ est une quasi-vari\'et\'e de dimension $2$. }\nolinebreak\hfill{$\Diamond$}\par\vspace{0.5\parskip}\end{fex} Soit $M$ une quasi-vari\'et\'e. Pour d\'efinir une g\'eom\'etrie sur $M$ nous proc\'edons de la fa\c{c}on suivante: d'une mani\`ere g\'en\'erale, un objet g\'eom\'etrique sur une quasi-vari\'et\'e est la donn\'ee d'un objet g\'eom\'etrique sur tout ouvert $\ut_{\alpha}$, invariant par rapport \`a l'action du groupe discret $\Gamma_{\alpha}$ et satisfaisant \`a des conditions de compatibilit\'e. Par exemple, une forme diff\'erentielle de degr\'e $h$, $\omega$, sur $M$ est la donn\'ee, pour tout $\alpha\in{\cal A}$, d'une forme diff\'erentielle $\Gamma_{\alpha}$-invariante de degr\'e $h$, $\ot_{\alpha}$, sur $\ut_{\alpha}$ qui se transforme de mani\`ere compatible aux changements de cartes $g_{\alpha\beta}$; une $2$-forme est dite symplectique si toute $\ot_{\alpha}$ est symplectique (ferm\'ee et non-d\'eg\'en\'er\'ee). Nous renvoyons le lecteur \`a l'article \cite{p1} pour les d\'efinitions de: champ de vecteurs, application diff\'erentiable, diff\'eomorphisme, image r\'eciproque d'une forme, image directe d'un champ de vecteurs, diff\'erentiel et produit int\'erieur. L'analogue du tore dans la g\'eom\'etrie des quasi-vari\'et\'es s'appelle quasi-tore. Soit $\mbox{\frak d}$ un espace vectoriel de dimension $n$. \begin{fdefn}[Quasi-r\'eseau, quasi-tore, quasi-alg\`ebre de Lie]{\rm Un {\em quasi-r\'eseau}, $Q$, dans $\mbox{\frak d}$ est le $\mbox{\bbb{Z}}$-module engendr\'e par un ensemble de vecteurs $X_1,\ldots,X_d$ qui engendrent $\mbox{\frak d}$. Nous appelons {\em quasi-tore} de dimension $n$ le groupe quotient $D=\mbox{\frak d}/Q$. La {\em quasi-alg\`ebre} de Lie du quasi-tore $D$ est l'espace vectoriel $\mbox{\frak d}$.} \nolinebreak\hfill{$\triangle$}\par\vspace{0.5\parskip}\end{fdefn} Un quasi-tore est evidemment une quasi-vari\'et\'e et les operations de groupe sont diff\'erentiables. Si $Q$ est un vrai r\'eseau (par exemple si $d=n$) on retrouve dans la d\'efinition pr\'ec\'edente le tore et son alg\`ebre de Lie. Les quasi-tores de dimension un ont \'et\'es \'etudi\'es par Donato, Iglesias et Lachaud \cite{di,i, il}; Iglesias introduit \`a cette occasion la terminologie {\em tores irrationnels}. \begin{fex}[Quasi-cercle]\label{quasicercle}{\rm Consid\'erons dans $\mbox{\bbb{R}}$ le quasi-r\'eseau $Q=s\mbox{\bbb{Z}}+t\mbox{\bbb{Z}}$, $s/t\notin\mbox{\bbb{Q}}$. Alors $D^1=\mbox{\frak d}/Q$ est un quasi-tore de dimension $1$. \nolinebreak\hfill{$\Diamond$}\par\vspace{0.5\parskip}}\end{fex} Les quasi-tores apparaissent de fa\c{c}on tr\`es naturelle, comme le montre la proposition suivante. \begin{prop} Soit $T$ un tore et soit $N$ un sous-groupe de Lie de $T$. Alors $T/N$ est un quasi-tore de dimension $n=\dim{T}-\dim{N}$. \nolinebreak\hfill{$\Box$}\par\vspace{0.5\parskip}\end{prop} Par exemple le quotient du tore de dimension $2$ par une droite de pente irrationnelle $s/t\notin\mbox{\bbb{Q}}$ est le quasi-tore de l'exemple~\ref{quasicercle}. \begin{fdefn}[Action diff\'erentiable]{\rm Une {\em action diff\'erentiable} d'un quasi-tore $D$ sur une quasi-vari\'et\'e $M$ est une application diff\'erentiable $\tau\,\colon\, D\times M \longrightarrow M$ telle que $\tau(d_1\cdot d_2,m)=\tau(d_1,\tau(d_2,m))$, et $\tau(1_{D},m)=m$ pout tout $d_1, d_2 \in D$ et $m\in M$.} \nolinebreak\hfill{$\triangle$}\par\vspace{0.5\parskip}\end{fdefn} Etant donn\'e une action diff\'erentiable d'un quasi-tore sur une quasi-vari\'et\'e il est toujours possible d'associer \`a tout $X\in\mbox{\frak d}$, comme dans le cas des vari\'et\'es, un champ de vecteurs sur $M$, dit {\em champ fondamental de l'action}, not\'e $\mbox{X}_M$. \begin{fdefn}[Action hamiltonienne, application moment]{\rm Une action diff\'erentiable d'un quasi-tore $D$ sur une quasi-vari\'et\'e $M$ est dite {\em hamiltonienne} si elle respecte la forme symplectique et s'il existe une application diff\'erentiable $D$-invariante $\Phi\,\colon\, M\longrightarrow\d^*$, dite {\em application moment}, ayant la propri\'et\'e que $\imath(\mbox{X}_M)\omega=d<\Phi,X>$, pour tout $X\in\mbox{\frak d}$. \nolinebreak\hfill{$\triangle$}\par\vspace{0.5\parskip}}\end{fdefn} \begin{fex}\label{quasisphere2}{\rm Consid\'erons la quasi-sph\`ere de l'exemple~\ref{quasisphere} et le quasi-cercle de l'exemple~\ref{quasicercle}. L'application $\tau([\theta],[z:w])=[\eones z:w]$ d\'efinit une action hamiltonienne de $D^1$ sur $M$, d'application moment $\Phi([z:w])=\frac{|z|^2}{s}=1-\frac{|w|^2}{t}$. Remarquons que $\Phi(M)=[0,1]$ (indep\'endamment de $s$ et $t$) comme pour la rotation de la sph\`ere de $\mbox{\bbb{R}}^3$ autour de l'axe $Oz$.}\nolinebreak\hfill{$\Diamond$}\par\vspace{0.5\parskip}\end{fex} Pour obtenir beaucoup d'autres exemples d'actions hamiltoniennes de quasi-tores sur des quasi-vari\'et\'es il suffit de prendre la r\'eduction symplectique par rapport \`a un sous-groupe de Lie d'un vrai tore agissant sur une vraie vari\'et\'e symplectique. \begin{fthm}[R\'eduction] Soit $T$ un tore d'alg\`ebre de Lie $\mbox{\frak t}$, soit $T\times X \longrightarrow X$ une action hamiltonienne de $T$ sur une vari\'et\'e symplectique $X$ et supposons que l'application moment $J\,\colon X \longrightarrow\t^*$ soit propre. Consid\'erons l'action induite d'un sous-groupe de Lie $N$ de $T$ et supposons que $0$ soit une valeur r\'eguli\`ere de l'application moment correspondante: $\psi=i^*\circ J\,\colon X \longrightarrow \n^*$ ($\mbox{\frak n}$ d\'enote l'alg\`ebre de Lie de $N$ et $i$ l'inclusion dans $\mbox{\frak t}$.) Alors $M=\psi^{-1}(0)/N$ est une quasi-vari\'et\'e symplectique de dimension $\dim{X}-2\dim{N}$ et l'action induite de $T/N$ sur $M$ est hamiltonienne. \nolinebreak\hfill{$\Box$}\par\vspace{0.5\parskip} \end{fthm} Ce dernier th\'eor\`eme permet d'\'etendre la construction de Delzant et donc de montrer que tout polytope simple $\Delta\subset\d^*$ est l'image par l'application moment d'une quasi-vari\'et\'e. Rappelons qu'un polytope convexe est dit simple si de chaque sommet sont issues exactement $n$ ar\^etes (ici nous supposons par simplicit\'e $\dim{\Delta}=n$). Nous avons alors: \begin{fthm}\label{toutsimple} Pour tout polytope convexe simple $\Delta\subset\d^*$ il existe un quasi-tore de dimension $n$ et de quasi-alg\`ebre de Lie $\mbox{\frak d}$, $D$, une quasi-vari\'et\'e symplectique compacte de dimension $2n$, $M$, et une action hamiltonienne effective de $D$ sur $M$ telle que l'image de l'application moment correspondante soit $\Delta$. \nolinebreak\hfill{$\Box$}\par\vspace{0.5\parskip}\end{fthm}
\section{Introduction} Physical mechanisms underlying many cardiac arrhythmias, in particular the transition from ventricular tachycardia (VT) to ventricular fibrillation (VF), are not fully understood. The ventricular tissue is known, both experimentally and theoretically, to support long-living spiral excitations, and it is thought that a breakup of such a spiral could give rise to a turbulent, chaotic activity commonly associated with VF. (Spirals are reviewed in books \cite{books}.) A considerable effort is now being directed towards understanding of these defect-mediated transitions to turbulence within mathematical models of ventricular tissue. The currently popular approach (reviewed in Ref. \cite{Panfilov}) considers a spiral in a patch (or slab) of ventricular tissue; the patch is taken in isolation from any pacemaking source. One then follows numerically the time evolution of that initial spiral. In the real beating heart, however, the ventricles are not isolated from other regions, and the heart, viewed as a whole, supports a (more or less) periodic autonomous activity---the heartbeat itself. In this case, any defect should be properly viewed as a disturbance of the normal heartbeat, rather than a structure in isolated tissue. In this paper we present some general results on the evolution of initial disturbances in autonomously active media and discuss their possible applications to cardiac arrhythmias. In particular, we identify a simple mechanism of defect-induced transition to turbulence in discrete (lattice) systems. We also find that the more rigid is the system in maintaining locally the undisturbed form of activity, the more easily the transition to turbulence occurs. This observation can potentially identify a useful therapeutic target. The assumed lattice structure need not (though it may) be related to the mechanical structure of the medium. The size of the lattice spacing in our models simply represents the smallest spatial scale on which the rhythmic activity can be desynchronized: a region smaller than that scale will necessarily fire as one. Discrete models of fibrillation have a long history, cf. the 1964 model of Moe {\em et al.} \cite{Moe&al}. (Unlike these authors, though, we do not introduce any frozen inhomogeneity in the parameters of the medium, apart from the lattice structure itself.) In addition, the importance of a discrete (granular) structure of the medium has been emphasized in theoretical studies of {\em de}fibrillation \cite{defib}. We introduce an interaction of an excitable region (like the ventricles) with a pacemaking region using the following simplified (not anatomical) model. We consider a three-dimensional (3d) slab of simulated medium whose extent in the $z$ direction is limited by the planes $z=0$ and $z=L_{z}$. The properties of the medium change in the $z$ direction: the region near $z=0$ is spontaneously oscillatory and represents the pacemaking region; the region at larger $z$ is merely excitable and represents the ventricular tissue. The $z$ direction will be also called longitudinal, and the other two directions, $x$ and $y$, will be called transverse. The medium supports a spontaneous rhythmic activity, in which an infinite train of pulses propagates from small to large $z$. This steady activity is independent of $x$ and $y$ and is supposed to model the heart's normal rhythm, in which pulses propagate from the inner surface of the ventricles out. The goal of our study was to see what happens if at some instant the spontaneous rhythmic activity is disturbed in a spatially nonuniform fashion, and then the system is left to itself. We approach this question in two steps. First, we consider the case when the initial disturbance is very smooth, i.e. almost uniform across the medium; in particular, it captures no topological defects. In this case, we expect that {\em locally} the activity rapidly relaxes close to its undisturbed form. The state can then be described using a single field $\tau(x,y,z;t)$, which measures the space- and time-dependent delay (or advance) in activity among the local regions. This field is a phase variable: it is defined modulo the period $T$ of the steady rhythm. For these smooth perturbations, we expect that the dynamics of $\tau$ at large times will be universal: it will be described by an equation whose form (although not the precise values of the coefficients) does not depend on the details of electrophysiology or on the microstructure of the medium. In particular, this large-time dynamics does not ``see'' the granular structure of the medium. The form of the equation depends on the symmetries of the medium at large scales and can be obtained by keeping terms of the lowest order in space and time derivatives consistent with the symmetries. For simplicity, we will assume that at large scales the properties of the medium are invariant under translations and rotations in the $x$--$y$ plane and that $\tau$ does not depend on $z$, i.e. the disturbance is effectively two-dimensional (2d). (Recall that $z$ is the direction of propagation of the normal rhythm.) In this case, the equation describing the large-time dynamics has the form \begin{equation} \partial_{t} \theta = a \nabla_{2}^{2} \theta + c (\nabla_{2}\theta)^{2} \; , \label{lwl} \end{equation} where the phase $\theta(x,y;t)$ is related to $\tau$ via \begin{equation} \theta(x,y;t)=2\pi \tau(x,y;t)/T \; , \label{theta0} \end{equation} and $a$ and $c$ are coefficients; $\nabla_{2}$ is the 2d gradient: $\nabla_{2}=(\partial_{x}, \partial_{y})$. We define a smooth disturbance by the condition \begin{equation} |\nabla_2 \theta| \ll 2\pi/L \; , \label{smo} \end{equation} where $L=\max\{L_x, L_y\}$ is the transverse size of the medium. Under this condition, the second term in on the right-hand side of (\ref{lwl}) is much smaller than the first. We keep it nonetheless, because it is the leading term that breaks the $\theta\to -\theta$ symmetry. As we will see, terms breaking this symmetry play an important role in evolution of non-smooth disturbances, such as topological defects. So, it is essential to establish that the coefficient $c$ is indeed nonzero. For smooth disturbances, though, the second term is unimportant, and eq. (\ref{lwl}) shows that when $a>0$ a smooth initial disturbance relaxes back to the uniform steady rhythm ($\theta=\rm const$). The relaxation process is ordinary diffusion. It is important to provide a derivation of (\ref{lwl}) from an electrophysiological model. In particular, that would supply certain values for the yet unknown coefficients $a$ and $c$. In Sect. 2 we show how $\theta$ (or $\tau$) can be defined within such a model. The smaller are gradients of $\theta$, the slower it evolves. One might think that, given an electrophysiological model, it should be easy to separate away the slow dynamics and obtain, quite generally, a closed equation for $\theta$. This task, however, turns out to be far from straightforward, and as of this writing we have not been able to obtain a general derivation of (\ref{lwl}); in Sect. 2 we illustrate the nature of the difficulty. To establish that the coefficient $c$ is indeed nonzero, we then have resorted to the following argument. The simple electrophysiological model that we consider can be driven, by a choice of the parameters, to a critical (bifurcation) point, at which the autonomous rhythmic activity is extinguished. Near the critical point, the system can be described by a complex Ginzburg-Landau (CGL) model of a complex order parameter whose phase is our time-delay field $\theta$. For a smooth, almost uniform, perturbation, the CGL description reduces to an equation for $\theta$ alone, and that has the precise form (\ref{lwl}), with definite values of $a$ and $c$. In particular, we find that $a>0$ and $c\neq 0$. As we move away from the critical point and towards the form of activity representative of the normal heartbeat, the CGL description ceases to be valid. But as it is difficult to imagine how $c$ would now suddenly become identically zero, we assume that the large-time dynamics of $\theta$ is still described by (\ref{lwl}) with a nonzero $c$. We also assume that $a>0$, so that the uniform state is stable. The electrophysiological model that we use is reviewed in Sect. 3, and the CGL description is derived in Sect. 4. The second step of our program is promoting the above description of smooth perturbations to a description including not-so-smooth perturbations, in particular, topological defects. The latter description will not be universal. The lack of universality means (by definition) that the description, and the type of the resulting dynamics, depend on the microstructure of the medium. Because no activity can be fine-grained indefinitely, it is natural to assume a granular, or lattice, structure. In Sect. 5, we construct lattice models and study their dynamics. In Sect. 6 we summarize our results. \section{Description of smooth disturbances} In this section we want to show how the slow variable $\theta$, or equivalently $\tau$, can be defined within the context of an electrophysiological model. This variable evolves arbitrarily slow in the limit of arbitrarily small gradients; it should not be confused with ``slow'' recovery variables of electrophysiology. Our definition of $\tau$ works for any medium supporting an autonomous periodic activity that is stable with respect to smooth, almost uniform, perturbations. For definiteness, we consider here an electrophysiological equation of the form \begin{equation} \epsilon \ddot{g} - \nabla^{2} \dot{g} - b \nabla^{2} g - F(g,\dot{g};z) = 0 \; . \label{eqm} \end{equation} Overhead dots denote time derivatives, $\nabla$ is the 3d gradient, and $\epsilon$ and $b$ are parameters. The change in properties of the medium in the $z$ direction is described by the function $F$, which explicitly depends on $z$. Eq. (\ref{eqm}) obtains, for instance, when a medium described by the two-variable FitzHugh-Nagumo (FHN) model \cite{FHN} is placed in an external static electric field (we will show that below). In that case, $g$ is the deviation of the recovery variable of the FHN model from the static solution. We consider cases when eq. (\ref{eqm}) (or, more precisely, a suitable boundary problem based on it) has a periodic in time solution of the form \begin{equation} g(\r,t) = \phi(z, t) \; . \label{phi} \end{equation} For example, this solution may describe a train of pulses propagating in the $z$ direction. The periodicity means that $\phi(z,t+T)=\phi(z,t)$ for some period $T$. Notice that, because of the translational invariance of (\ref{eqm}) in time, $\phi(z, t-\tau)$ is also a solution of (\ref{eqm}), for any real $\tau$ (albeit with different initial conditions). We now consider a smooth (in space) perturbation of the periodic activity described by (\ref{phi}) and assume that a sufficiently smooth perturbation relaxes back to the periodic state. After the relaxation has been under way for a while, we expect that deviations of $g$ from $\phi$ are already small---except perhaps in the softest mode, associated with the time translation. We thus seek a solution to (\ref{eqm}) of the form \begin{equation} g(\r,t) = \phi(z, t- \tau(\r,t)) + \chi(\r,t) \; , \label{sol} \end{equation} where $\tau(\r,t)$ is a slowly changing (on the scale of the period $T$) function of time: $\dot{\tau} \ll \tau/T$. In the limit $\dot{\tau} \to 0$, we should return to the solution (\ref{phi}) merely shifted in time, so in this limit $\chi$ should vanish. Thus, when $\dot{\tau}$ is small, $\chi$ is also small, although not necessarily slowly changing. Because of the periodicity of $\phi$ in time, $\tau(\r,t)$ is a phase variable: at each spatial point, it is defined modulo the period $T$. The condition that the perturbation be smooth reduces this ambiguity to a common shift by $T$ in the entire space. Note that separation of a perturbation into $\tau$ and $\chi$ is not completely defined by (\ref{sol}): a time-dependent variation in $\tau$ can be absorbed by a variation in $\chi$. This ambiguity can be fixed by an additional condition---for instance, by requiring that $\chi$ is orthogonal to $\dot{\phi}$ with respect to a certain inner product. Eq. (\ref{sol}) together with the additional condition will then provide a complete definition of the slow variable $\tau$. Now, let us illustrate the nature of the difficulty that arises when one tries to derive a closed equation for $\tau$ from eq. (\ref{eqm}). We substitute (\ref{sol}) into (\ref{eqm}) and expand the right-hand side to the leading order in small quantities---the function $\chi$ and the derivatives of $\tau$. The dependence on $\chi$ will be contained in an expression of the form ${\hat M}(\phi)\chi$, where ${\hat M}$ is a linear operator, which acts on $\chi$ and depends on $\phi(z, t- \tau(\r,t))$. Because of the translational invariance of (\ref{eqm}) in time, the operator ${\hat M}(\phi)$ almost annihilates $\dot{\phi}(z, t- \tau(\r,t))$: \begin{equation} {\hat M}(\phi)\dot{\phi} \approx 0 \; ; \label{annih} \end{equation} the approximate equality means an equality up to terms of order of the small quantity $\partial_{t} \tau$. If the operator ${\hat M}(\phi)$ were Hermitean with respect to an inner product of the form \begin{equation} \langle \chi_1 , \chi_2 \rangle = \int_0^{L_z} dz \int_0^T dt w(z,t) \chi_1(z,t) \chi_2(z,t) \; , \label{inpro} \end{equation} for some fixed weight $w(z,t)$, then taking the inner product of (\ref{eqm}) with $\dot{\phi}$ would, to the leading order, project away $\chi$ and produce a closed equation for $\tau$. In the case of eq. (\ref{eqm}), however, the explicit form of the operator ${\hat M}$ is \begin{equation} {\hat M}(\phi)\chi = \left( \epsilon\partial_t^2 - \nabla^2 \partial_t - b \nabla^2 - \dpar{F}{\phi} -\dpar{F}{\dot{\phi}} \partial_t \right) \chi \; , \label{M} \end{equation} where $F$ is $F(\phi, \dot{\phi}; z)$. This operator is clearly not Hermitean with respect to (\ref{inpro}) with $w=1$, and indeed we have not found any weight that would render it Hermitean. Thus, we were unable to directly separate the slow dynamics of $\tau$ from the fast dynamics of $\chi$. While it seems intuitively clear that the slow dynamics will be described by an equation of the form (\ref{lwl}), to establish that the coefficients $a$ and $c$ are indeed both nonzero, we had to resort to an indirect method, which we describe below. \section{A model of the heartbeat} In this section, we describe in some detail the pacemaking mechanism with which we model the heartbeat. This simple model, based on the two-variable FitzHugh-Nagumo (FHN) kinetics, will be sufficient for our argument justifying (\ref{eqm}) with nonzero $a$ and $c$. Consider a slab of medium described by a FitzHugh-Nagumo model, \begin{eqnarray} \epsilon\dpar{E}{t} & = & \nabla^{2} E + f(E) - G \; , \label{eqE} \\ \dpar{G}{t} & = & E - b G \; , \label{eqG} \end{eqnarray} placed in a static uniform external electric field, such as the field of a parallel capacitor. Here $E$ is the transmembrane voltage, $G$ is the recovery variable, $\epsilon>0$ and $b>0$ are parameters, and $\nabla$ is the 3d gradient. The direction of the external field is our longitudinal, or $z$, direction, and the slab extends in that direction from $z=0$ to $z=L_z$. The boundary conditions corresponding to this arrangement are \begin{equation} \partial E/\partial z(0) = \partial E/\partial z(L_z) = -{\cal F} \; , \label{bc} \end{equation} where ${\cal F}$ is a positive constant---the magnitude of the external field. The boundary problem (\ref{eqE})--(\ref{bc}) has a static solution, $E_0(z)$, $G_0(z)$. Deviations from the static solution are $e(\r,t)=E(\r,t)-E_{0}(z)$ and $g(\r,t)=G(\r,t)-G_0(z)$. Excluding the variable $e$ with the help of (\ref{eqG}), we obtain an equation of the form (\ref{eqm}) with \begin{equation} F(g,\dot{g};z) = f(E_0+ b g+\dot{g}) - f(E_0) - g - \epsilon b\dot{g} \; . \label{F} \end{equation} The explicit dependence of $F$ on $z$ appears through the $z$ dependence of $E_0$. For a range of ${\cal F}$ the static solution to (\ref{eqE})--(\ref{bc}) is unstable, for various choices of $f(E)$, with respect to arbitrarily small fluctuations of $E$ and $G$, and the instability gives rise to an unending time-dependent activity \cite{RK}. This will be our pacemaking mechanism. The corresponding linear stability analysis introduces a number of useful definitions, so we briefly go over it here. Expanding eqs. (\ref{eqE})--(\ref{eqG}) to the first order in $e$ and $g$, we obtain \begin{equation} \left(\begin{array}{c} \partial e/\partial t \\ \partial g/\partial t \end{array} \right)= \left(\begin{array}{cc} \frac{1}{\epsilon} \left( \nabla_{2}^{2} +\dpar{^{2}}{z^{2}} + f'[E_{0}(z)] \right) & -{1\over\epsilon} \\ 1 & -b \end{array} \right) \left(\begin{array}{c} e \\ g \end{array} \right) \; . \label{eqdel} \end{equation} This equation should be supplemented by the boundary conditions \begin{equation} \dpar{e}{z}(0)=\dpar{e}{z}(L_{z})=0 \; . \label{bou1} \end{equation} Consider eigenfunctions $\psi_{n}(z)$, $n\geq 0$, of the $z$-dependent operator in (\ref{eqdel}), \begin{equation} \left(-\dpar{^{2}}{z^{2}} -f'[E_{0}(z)] \right) \psi_{n}(z)=\lambda_{n}\psi_{n}(z) \; , \label{ope} \end{equation} with the boundary conditions \begin{equation} \dpar{\psi_{n}}{z}(0)=\dpar{\psi_{n}}{z}(L_{z})=0 \; . \label{bou2} \end{equation} We assume that the eigenfunctions $\psi_{n}$ are real and form a complete orthonormal system on $L_{2}[0, L_{z}]$. The fields $e$ and $g$ can be expanded in the complete orthonormal system $\{\psi_{n}\}$: \begin{eqnarray} e(\r,t) & = & \sum_{n=0}^{\infty} u_{n}(\r_{2}, t) \psi_{n}(z) \label{expe} \; , \\ g(\r,t) & = & \sum_{n=0}^{\infty} v_{n}(\r_{2}, t) \psi_{n}(z) \; ; \label{expg} \end{eqnarray} here $\r_{2}$ is the two-dimensional coordinate: $\r_{2}=(x,y)$. Eq. (\ref{eqdel}) then reduces to the following second-order in time linear equation \begin{equation} \ddot{v}_{n}+\left( b+\frac{\lambda_{n}-\nabla_{2}^{2}}{\epsilon} \right) \dot{v}_{n} + {1\over \epsilon}\left(1+b[\lambda_{n}-\nabla_{2}^{2}] \right) v_{n} = 0 \; . \label{so} \end{equation} Eq. (\ref{so}) describes a collection of independent oscillators, one for each value of the integer $n\geq 0$ and of the 2d wave number $\k$. These oscillators have frequencies squared equal to $\omega_{n}^{2}+ b k^{2}/\epsilon$ and friction coefficients equal to $\gamma_{n}+k^{2}/\epsilon$, where \begin{eqnarray} \omega_{n}^{2}& = & (1+b\lambda_{n})/\epsilon \; , \label{ome} \\ \gamma_{n}& = & b+\lambda_{n}/\epsilon \; . \label{gam} \end{eqnarray} Assuming that the boundary conditions in the $x$--$y$ plane allow for the $\k=0$ mode, we conclude that the necessary and sufficient condition for instability is that \begin{equation} \lambda_{n}<\max\{-\epsilon b, -1/b\} \label{cond} \end{equation} for at least one of the eigenvalues $\lambda_n$. This condition corresponds to there being a negative $\omega_{n}^{2}$ or a negative $\gamma_{n}$, or both. The parameter $\epsilon$ sets the ratio of time scales characterizing changes in the voltage $E$ and in the recovery variable $G$ and is typically small. When $\epsilon <1/b^{2}$, the condition (\ref{cond}) becomes \begin{equation} \lambda_{n}<-\epsilon b \; , \label{cond1} \end{equation} or equivalently $\gamma_{n} < 0$, where $\gamma_{n}$ is the friction (\ref{gam}). The question that we now address is whether the condition (\ref{cond1}) is ever satisfied for physiologically relevant values of the parameters. We choose $\epsilon=0.06$, $b=0.7$, and $f(E)=6.75E(E-0.25)(1-E)$, as recommended in Ref. \cite{Starmer&al} for ventricular tissue with ``normal'' Na and K conductances. The only other parameter (besides ${\cal F}$) that we need to choose is $L_z$, the thickness of the slab in the $z$ direction. This represents the thickness of the ventricles in our simplified model. We have done numerical simulations with $L_z=3.2$. For lengths, Ref. \cite{Starmer&al} recommends scaling by a factor of 0.5 cm. A somewhat smaller scaling factor of 0.2 cm is obtained if we equate the characteristic (``Debye'') length $\xi=0.57$, at which a weak static field gets screened inside the medium, to a realistic value of 1 mm. With either scaling, though, $L_z=3.2$ corresponds to a physical length of order 1 cm. To find out if the instability occurs for a given value of ${\cal F}$, one can numerically solve the boundary problem (\ref{ope})--(\ref{bou2}) and check the condition (\ref{cond}). Alternatively, one can numerically integrate the time-dependent problem (\ref{eqE})--(\ref{bc}) with initial conditions corresponding to small fluctuations near the static solution. This second approach also allows one to find the form of the time-dependent attractor emerging as the instability is cutoff by nonlinear effects, so we have adopted it. For the purposes of this section, it is sufficient to consider initial fluctuations that are independent of $x$ and $y$. Using numerical integrations of (\ref{eqE})--(\ref{bc}) with such initial conditions and with the above values of the parameters, we have found that the static solution is stable as long as ${\cal F} \leq {\cal F}_1\approx 0.4$. The value ${\cal F}_{1}$ is the lower critical value, at which the static solution first becomes unstable as ${\cal F}$ is increased. The instability persists as long as ${\cal F}_{1}< {\cal F}<{\cal F}_{2}$ but disappears when ${\cal F}$ reaches the upper critical value ${\cal F}_{2}\approx 1$. The form of the time-dependent attractor, which develops from small initial fluctuations near the static solution, is qualitatively different for values of ${\cal F}$ that are close to the upper critical field as compared to those elsewhere in the instability window. These two different forms correspond to propagating versus nonpropagating activity \cite{RK}. In the range ${\cal F}_1 < {\cal F} < {\cal F}_p$, where ${\cal F}_p$ is somewhat smaller than ${\cal F}_{2}$, the attractor is an unending train of pulses propagating in the positive $z$ direction. In our model, this corresponds to the normal heartbeat. On the other hand, when ${\cal F}_p < {\cal F} < {\cal F}_2$, the development of the instability is cut off by nonlinear effects when the deviation from the static solution is too small to generate a full-fledged pulse. In this case, the entire attractor lies in the proximity of the static solution. As ${\cal F}$ approaches ${\cal F}_{2}$, the activity is extinguished {\em gradually}: the closer is ${\cal F}$ to ${\cal F}_{2}$, the smaller is the deviation from the static solution. This gradual disappearance of activity is reminiscent of a second-order phase transition. \section{The CGL description} Near the upper critical field, which from now on we will call the {\em critical point}, the fields $e(\r,t)=E(\r,t)-E_{0}(z)$ and $g(\r,t)=G(\r,t)-G_{0}(z)$ are small ($E_{0}$ and $G_{0}$ denote the static solution). Expanding the system (\ref{eqE})--(\ref{eqG}) in $e$ and $g$ so as to retain the leading nonlinearities, we obtain \begin{eqnarray} \epsilon\dpar{e}{t} &=& \nabla^{2} e + f'(E_{0}) e + {1\over 2} f''(E_{0}) e^{2} + {1\over 6} f'''(E_{0}) e^{3} - g \; , \label{hoe} \\ \dpar{g}{t} &=& e- b g \; . \label{hog} \end{eqnarray} As it turns out, the effect of the $e^{2}$ term is relatively suppressed and is of the same order as the effect of the $e^{3}$ term. So, we kept both types of terms in eq. (\ref{hoe}). Substituting the expansions (\ref{expe})--(\ref{expg}) into (\ref{hoe})--(\ref{hog}), we obtain \begin{eqnarray} \epsilon\dpar{u_{n}}{t} &=& (\nabla_{2}^{2}-\lambda_{n}) u_{n} - v_{n} - \alpha_{nmm'} u_{m} u_{m'} - \beta_{nmm'm''} u_{m} u_{m'} u_{m''} \; , \label{equ} \\ \dpar{v_{n}}{t} & = & u_{n}- b v_{n} \; ; \label{eqv} \end{eqnarray} repeated indices are summed over. Here $\nabla_{2}$ is the 2d gradient: $\nabla_{2}=(\partial_{x}, \partial_{y})$, $\lambda_{n}$ is the eigenvalue of the Schr\"{o}dinger problem (\ref{ope})--(\ref{bou2}), and $\alpha$ and $\beta$ are defined as \begin{eqnarray} \alpha_{nmm'} & \equiv & -{1\over 2} \int_{0}^{L_{z}} dz f''(E_{0}) \psi_{n} \psi_{m} \psi_{m'} \; , \label{alpha} \\ \beta_{nmm'm''}& \equiv & -{1\over 6} \int_{0}^{L_{z}} dz f'''(E_{0}) \psi_{n} \psi_{m} \psi_{m'} \psi_{m''} \; . \label{beta} \end{eqnarray} We stay closely enough to the critical point, so that on that side of it where the static solution is unstable there will be only one $\lambda_{n}$ satisfying the instability condition (\ref{cond}). That will be $\lambda_{0}$. In what follows we only consider cases when $\epsilon < 1/b^{2}$. Then, the instability condition takes the form \begin{equation} \gamma_{0} < 0 \; , \label{cond2} \end{equation} where $\gamma_{0}= b + \lambda_{0}/\epsilon$ is the friction coefficient (\ref{gam}) for $n=0$. The closer the system is to the critical point, the smaller is $|\gamma_{0}|$. We make it small enough, so that the frequency squared (\ref{ome}) with $n=0$ (and hence with all $n>0$ as well) is positive and much larger than $\gamma_{0}^{2}$: \begin{equation} \omega_{0}^{2} = 1/\epsilon - b^{2} + b\gamma_{0} \gg \gamma_{0}^{2} \; . \label{ome1} \end{equation} The large positive $\omega_{0}$ sets the time scale of rapid oscillations of $u_{n}$ and $v_{n}$. We now want to show that when the system is sufficiently close to the critical point its dynamics on time scales of order of and larger than $|\gamma_{0}|^{-1}$ is described by a 2d complex Ginzburg-Landau (CGL) model. The field $\Psi(\r_{2},t)$ of this CGL model is defined via the expansion \begin{equation} v_{0}(\r_{2},t) = \left( {\Psi \over b-i\omega_{0}}{\rm e}^{-i\omega_{0}t} +{A_{0} \Psi^{2}\over b-2i\omega_{0}} {\rm e}^{-2i\omega_{0}t} + {\rm c.c.} \right) + {C_{0} \over b} \Psi^{\dagger} \Psi + \ldots \; , \label{psi} \end{equation} where the omitted terms are higher harmonics, proportional to the third and higher powers of $\exp(\pm i\omega_{0}t)$; c.c. means complex conjugate. The coefficients $A_{0}$ and $C_{0}$ are in principle series in $\Psi^{\dagger} \Psi$, but near the critical point $\Psi$ is small, and to the leading order $A_{0}$ and $C_{0}$ can be regarded as constants, which will be determined later. The definition (\ref{psi}) separates away the rapid oscillations with frequency $\omega_{0}$ and its multiples and, in this sense, is analogous to a transition to the nonrelativistic limit in field theory. The CGL description is obtained by substituting (\ref{psi}) into eqs. (\ref{equ})--(\ref{eqv}), expanding to the third order in $\Psi$, and finally retaining only terms that contain $\exp(\pm i\omega_{0}t)$ in powers 0, 1, and 2. One can verify that terms omitted in (\ref{psi}) will not contribute to the resulting equation. For instance, terms proportional to $\exp(\pm 3i\omega_{0}t)$ are of order $\Psi^{3}$; to convert them into terms of lower order in $\exp(\pm i\omega_{0}t)$ one will need to multiply them by at least one power of $\Psi$ or $\Psi^{\dagger}$, which will make them of the fourth order in $\Psi$. The CGL description allows us to consider disturbances of the uniform activity that satisfy the conditions \begin{equation} \dot{\Psi} = O(\Psi^{3}) \; , ~~~~~\nabla_{2}^{2}\Psi = O(\Psi^{3}) \; . \label{smo1} \end{equation} These are less restrictive than the smoothness condition (\ref{smo}), which now takes the form \begin{equation} |\nabla_{2} \Psi|/|\Psi| \ll 2\pi/L \; . \label{smo2} \end{equation} In particular, unlike (\ref{smo1}), the condition (\ref{smo2}) explicitly prohibits topological defects, which are centered at zeroes of $|\Psi|$. Under the more restrictive condition (\ref{smo2}), the CGL dynamics reduces, at sufficiently large times, to dynamics of the phase of $\Psi$ alone. To the third order in $\Psi$, $u_{0}$ is obtained from (\ref{eqv}) and (\ref{psi}) as \begin{equation} u_{0}(\r_{2},t) = C_{0} \Psi^{\dagger} \Psi + \left( \Psi{\rm e}^{-i\omega_{0}t}+ A_{0} \Psi^{2}{\rm e}^{-2i\omega_{0}t} + {\dot{\Psi} \over b-i\omega_{0} } {\rm e}^{-i\omega_{0}t} + {\rm c.c.} \right) + \ldots \; , \label{u0} \end{equation} where dots again denote higher harmonics. As will be checked a posteriori, $v_{n}$ and $u_{n}$ with $n>0$ are of order $\Psi^{2}$. In this approximation, eqs. (\ref{equ})--(\ref{eqv}) with $n=0$ become \begin{eqnarray} \epsilon\dpar{u_{0}}{t} &=& (\nabla_{2}^{2}-\lambda_{0}) u_{0} - v_{0} - \alpha_{000} u_{0}^{2} - 2\alpha_{00\nu} u_{0} u_{\nu} - \beta_{0000} u_{0}^{3} \; , \label{equ0} \\ \dpar{v_{0}}{t} & = & u_{0}- b v_{0} \; ,\label{eqv0} \end{eqnarray} where $\nu>0$, while for $n=\nu>0$ they become \begin{eqnarray} \epsilon\dpar{u_{\nu}}{t} &=& -\lambda_{\nu} u_{\nu} - v_{\nu} - \alpha_{\nu 00} u_{0}^{2} \; , \label{equnu} \\ \dpar{v_{\nu}}{t} & = & u_{\nu}- b v_{\nu} \; . \label{eqvnu} \end{eqnarray} We see that in this approximation the modes with $n=\nu>0$ are damped linear oscillators driven by the external force proportional to $u_{0}^{2}$. For the purpose of calculating $u_{\nu}$, it is sufficient to take $u_{0}^{2}$ computed to the second order in $\Psi$: \begin{equation} u_{0}^{2} = 2 \Psi^{\dagger}\Psi +\left( \Psi^{2} {\rm e}^{-2i\omega_{0} t} +{\rm c.c.} \right) + O(\Psi^{3}) \; . \label{force} \end{equation} Then, the solution for $u_{\nu}$ at large times is \begin{equation} u_{\nu} = A_{\nu}\Psi^{2}{\rm e}^{-2i\omega_{0} t} + A_{\nu}^{*} (\Psi^{\dagger})^{2}{\rm e}^{2i\omega_{0} t} + C_{\nu} \Psi^{\dagger} \Psi + O(\Psi^{3}) \; , \label{unu} \end{equation} where \begin{eqnarray} A_{\nu} & = & - \alpha_{\nu 00} \left(\lambda_{\nu} - 2i\epsilon \omega_{0} + {1\over b -2i\omega_{0}} \right)^{-1} \; , \label{A} \\ C_{\nu} & = & - 2 \alpha_{\nu 00} \left(\lambda_{\nu} + 1/b \right)^{-1} \; . \label{C} \end{eqnarray} Substituting this expression for $u_{\nu}$ into eq. (\ref{equ0}) for $u_{0}$ we see that the only effect of the modes with $n>0$ is a local (in space and time) renormalization of the dynamics of the $n=0$ mode. To complete our derivation of the CGL description, we now turn to eq. (\ref{equ0}) and compose separate equations for different powers of $\exp(-i\omega_{0}t)$. The equations for the zeroth and second powers give expressions for $C_{0}$ and $A_{0}$ that are of the same form as (\ref{A})--(\ref{C}) but with $\nu$ everywhere replaced by 0. The equation for the first power then gives the CGL equation \begin{equation} \dot{\Psi} = D\nabla_{2}^{2} \Psi - {1\over 2} \gamma_{0} \Psi - s \Psi^{2} \Psi^{\dagger} \; , \label{cgl} \end{equation} where the complex diffusion coefficient is \begin{equation} D = {1\over 2\epsilon} \left( 1 + {ib\over \omega_{0}} \right) \; , \label{D} \end{equation} and the complex coupling constant is \begin{equation} s = D \left( -2 \sum_{n=0}^{\infty} \alpha_{00n}^{2} \left( {2b\over \epsilon \omega_{n}^{2}} +\left(\lambda_{n} - 2i\epsilon \omega_{0} + {1\over b -2i\omega_{0}} \right)^{-1} \right) + 3\beta_{0000} \right) \; . \label{s} \end{equation} Recall that the condition of instability of the static solution is $\gamma_{0} < 0$, and near the critical point $|\gamma_{0}|$ is small. Spatially uniform activity near the critical point (for $\gamma_{0}<0$) is described by the following solution of (\ref{cgl}): \begin{equation} \Psi_{0}(t) = \rho_{0} \exp(-i s_{I} \rho_{0}^{2} t) \; , \label{psi0} \end{equation} where $\rho_{0} = (|\gamma_{0}|/2s_{R})^{1/2}$; $s_{R}$ and $s_{I}$ are the real and imaginary parts of $s$. Of course, this solution exists only when $s_{R}> 0$. For a smooth perturbation of this uniform activity (which, in particular, contains no topological defects), we can define the modulus $\rho(\r_{2}, t)$ and the phase $\theta(\r_{2},t)$ via \begin{equation} \Psi(\r_{2},t) = \rho(\r_{2}, t) \exp(-i s_{I} \rho_{0}^{2} t + \theta(\r_{2}, t) ) \; . \label{theta} \end{equation} Substituting this into eq. (\ref{psi}) shows that $\theta$ measures the phase shifts in periodic activity among local regions, so it is precisely the variable that we defined in Sect. 2. As the modulus $\rho$ relaxes close to $\rho\approx \rho_{0}$ everywhere in the 2d space, eq. (\ref{cgl}) reduces to an equation for the phase $\theta$ alone. That equation is of the form (\ref{lwl}), with $a={\rm Re} D$, and $c=-{\rm Im} D$. \section{Construction of lattice models} As we move away from the critical point and towards the form of activity that is more representative of the normal heartbeat, the CGL description ceases to be valid. Nevertheless, we expect that eq. (\ref{lwl}) will still apply for sufficiently smooth perturbations. That is because $\theta$ is the only variable that can change arbitrarily slowly (for arbitrarily small gradients), and the two terms on the right-hand side of (\ref{lwl}) are the only two terms of the lowest (second) order in gradients that are consistent with the symmetries of our model and the assumption that $\theta$ does not depend on $z$. Moreover, we now have a reason to believe that both coefficients $a$ and $c$ will be nonzero: we have seen that they were both nonzero near the critical point, and it is hard to imagine how either of them would vanish identically when we move away. So, we consider eq. (\ref{lwl}) to be reasonably well justified. The next step is to build upon (\ref{lwl}) to construct models that would apply to not-so-smooth perturbations of the normal rhythm, in particular, to those containing topological defects. As we consider perturbations of progressively smaller spatial scales, there are two effects that lead to deviations from (\ref{lwl}). On the one hand, the granular (lattice) structure of the medium becomes important; on the other hand, the local form of activity deviates from its unperturbed form, so that other variables besides $\theta$ come into play. We have found that the resulting dynamics depends crucially on which of these two effects becomes important first, i.e. at larger spatial scales. In what follows, we contrast the corresponding two types of the dynamics. Finding out which one is realized in a specific medium will require a detailed electrophysiological model. The required model will have to include the details of the granular structure, so it cannot be a simple continuum model of the type we used to justify eq. (\ref{lwl}). First, consider the case when the local activity is very {\em rigid} in maintaining its form. That means that each grain---or lattice site---still carries on essentially the undisturbed activity, so the field $\theta$ remains the only requisite variable. In this case, the dynamics is described by a model of classical lattice $XY$ spins. For definiteness, we consider here a model on a square lattice, with interactions restricted to the nearest neighbors (NN). (Similar results were obtained for a model that includes interactions of next-to-nearest neighbors.) We take the model equation in the form \begin{equation} \partial_{t} \theta_{i} = h^{-2} \sum_{j\in {\rm NN}(i)} \left[ a \sin(\theta_{j} - \theta_{i}) + c (1 - \cos(\theta_{j} - \theta_{i})) \right] \; . \label{eqlat} \end{equation} The index $i$ labels the sites of a 2d square lattice, and $h$ is the lattice spacing. Matching to the long-wave limit (\ref{lwl}) identifies $a$ and $c$ in (\ref{eqlat}) with those in (\ref{lwl}). Near the critical point, $c/a= - b/\omega_{0}$, which is proportional to the small $\sqrt{\epsilon}$. Away from the critical point, however, there is no reason to expect $|c/a|$ to be small, and we need to explore the dynamics of the model for diverse values of this ratio. We assume that $a>0$ and set $a=1$ by a rescaling of time. When $c=0$, eq. (\ref{eqlat}) becomes the usual diffusive $XY$ model. This model has stable topological defects---vortices and antivortices. A nonzero $c$ gives these defects a rotation (clockwise or counterclockwise, depending on the sign of $c$), so vortices and antivortices become spirals. By numerically integrating (\ref{eqlat}), we have found that for small values of $|c|$ these spirals are stable---or at least no instability could be detected during finite times of our computer runs. As $|c|$ is increased, the spirals become more tightly wound and at a sufficiently large $|c|$ they become unstable. Formation of a tightly wound but still stable spiral is illustrated by Figs. \ref{fig:init}, \ref{fig:spiral}. Fig. \ref{fig:init} shows an initial state, containing a single vortex, and Fig. \ref{fig:spiral} shows the spiral that develops from that initial state for $a=1$ and $c=-0.5$. The values of $\theta$ at a given time are represented as directions of lattice spins, as measured clockwise from 12 noon \cite{dl}. These results were obtained via Euler's explicit time-stepping scheme on a $33\times 33$ lattice with side length $L=10$ and discretized Neumann boundary conditions. For picture clarity, only a $22\times 22$ square is shown. Evolution of an unstable defect is illustrated by Fig. \ref{fig:bubble}. This picture was obtained for $a=1$ and $c=-2$ on the same lattice and with the same initial condition as Fig. \ref{fig:spiral}. The center of the defect now serves as a nuclei of a new phase, a featureless turbulent state. A bubble of the new phase originates at the center of the defect and rapidly grows, eating up the ``normal'' phase, until the new phase occupies the entire volume. As far as we can tell, the resulting turbulent state is persistent. Fig. \ref{fig:bubble} shows the bubble during its growth. This growth is indeed so rapid that the initial vortex does not have time to fully develop into a spiral, although some fragments of spiral structure can be seen near the wall of the bubble. A patch of the turbulent state is seen inside the bubble, away from the wall. When the turbulent state occupies the entire volume, it remains disordered: directions of the spins are uncorrelated beyond a few lattice spacings. In addition, spins in the turbulent state rapidly change their directions with time. Next, we consider a case when the local activity is {\em flexible}, i.e. it readily changes its form in response to a short-scale perturbation. For instance, we can supply the lattice spins with a variable length by making $\theta$ the phase of a complex field $\Phi=|\Phi| \exp(i\theta)$. This introduces an additional degree of freedom associated with $|\Phi|$. As an illustration, consider $\Phi$ that obeys a complex Ginzburg-Landau (CGL) equation: \begin{equation} \dpar{\Phi}{t} = D\nabla^{2}\Phi + r \Phi (1- |\Phi|^{2} ) \; , \label{cgl1} \end{equation} where $D=a- ic$; for simplicity we take the coupling $r$ to be real: $r> 0$. We can now discretize eq. (\ref{cgl1}) on a 2d square lattice of spacing $h$ and vary the parameter $r$ in relation to $h^{-2}$. At large $r$, the modulus $|\Phi|$ freezes out at $|\Phi| \approx 1$, and we obtain a lattice model of $\theta$ alone, in the spirit (although not necessarily of the exact form) of eq. (\ref{eqlat}). At small $r$, the natural size of a defect's core will be set by $(|D|/r)^{1/2}$, rather than the lattice spacing, so we expect that the discretization will be irrelevant, and the dynamics will approach that of the continuum 2d CGL model. This latter model has spiral solutions that are at least core-stable in a certain range of its parameters \cite{cgl}. Numerically integrating discretized eq. (\ref{cgl1}), we have found that by varying $r$, for a fixed $c/a$, one can interpolate between the unstable spirals of a lattice model with fixed-length spins and the stable spirals of the continuum CGL model. \section{Conclusion} In this paper we tried to implement consistently the idea that a disturbance in the normal heartbeat can be viewed as a collection of ``clocks'', each of which measures the local phase of the activity. In conjunction with the view that the heart has a granular (or lattice) structure, this idea leads to a description of the heart via lattice models of classical spins. Our main results are as follows. (i) Assuming that sufficiently smooth (almost uniform across the medium) disturbances of the normal rhythm relax back to it, one can write down a universal description of this relaxation process. Universality means that the form of the equation is independent of details of microscopics. For a simplified model of the heartbeat, and disturbances depending only on the transverse (with respect to the direction of pulse propagation) coordinates, the universal description is eq. (\ref{lwl}). Although we have not derived this equation in the general case, we have justified it by presenting a derivation near a critical (bifurcation) point. (ii) For not-so-smooth disturbances, including topological defects, dynamics begins to depend on the assumed lattice structure and the details of electrophysiology. In particular, we have found that it depends strongly on how rigid the local activity is in maintaining its form. When the activity is very rigid (fixed length spins), the system, for a range of the parameter space, is prone to a defect-induced instability, which leads to a disordered, turbulent state. We expect that the local rigidity of the medium (in the above sense) will depend on its longitudinal size (the thickness of the ventricles) and on the electrophysiological parameters, such as Na and K conductances. Since, according to our results, the local rigidity plays such an important role in the transition to turbulence (fibrillation), its dependence on the parameters may serve to identify useful therapeutic targets.
\section{Introduction} The heavy vector mesons $B^{*}$ and $D^{*}$ (of spin-parity $1^-$) decay via spin-flip electromagnetic or strong interactions to the well studied pseudoscalar ground states $B$ and $D.$ The decays of $D^{*}$ are known to proceed either as a strong transition \mbox{$D^{*} \to \, D\, \pi\,$} with a final pion with momentum of about $40\, {\rm MeV}$ or as an electromagnetic one \mbox{$D^{*} \to \, D\, \gamma\,$}, with a final photon with momentum of about $140 {\rm MeV}.$ The situation is different for $B^{*}$ which has a mass of $5324.9 \pm 1.8 $ MeV; since the mass difference $M_{B^{*}}-M_B$ is only $45.8$ MeV, there is no strong $B^{*}$ decay and the radiative process \mbox{$B^{*} \to \, B\, \gamma\,$} is the dominant decay mode for $B^{*}.$ In the present paper we study another possible electromagnetic decay, the two-photon decay processes \mbox{$B^{*} \to \, B\, \gamma\,\gamma\,$} and \mbox{$D^{*} \to \, D\, \gamma\,\gamma\,$} which were not considered previously in the literature. In addition to the intrinsic interest in these novel modes, we point out that their study could provide information on the strong couplings $g_{B^{*}B\pi\,}$ and $g_{D^{*}D\pi\,}$ and on the electromagnetic ones $g_{B^*B\gamma},g_{D^*D\gamma}.$ The strong couplings are directly related to the basic strong coupling $g$ of the effective heavy meson chiral Lagrangian, which describes the interactions of heavy mesons with low-momentum pions. There is a major dissimilarity between the possibility of measuring the couplings in the charm and beauty sectors, as a result of the different mass difference between the respective vector and pseudoscalar mesons. In the charm sector, the experimentally measured branching ratios of the ${D^*}^0,{D^*}^+$ decays into the allowed $D\pi$ and $D\gamma$ modes lead to relations between $g_{D^{*}D\pi}$ and $g_{D^{*}D\gamma}.$ Henceforth, the \mbox{$D^{*} \to \, D\, \gamma\,\gamma\,$} decay under study here is expressible in terms of the strong coupling only and would provide a convenient tool for its measurement. On the other hand, in the beauty sector where the \mbox{$B^{*} \to \, B\, \pi\,$} decay is forbidden by phase space, the $g_{B^{*}B\pi}$ coupling is not directly accessible. And although the \mbox{$B^{*} \to \, B\, \gamma\,$} decay is experimentally detected, the direct measurement of its strength is an unlikely proposition at present, in view of the smallness of the expected value of its decay width. However, as we show in this paper, the two photon decay \mbox{$B^{*} \to \, B\, \gamma\,\gamma\,$} branching ratio turns out to be a function of $g_{B^{*}B\pi}$ and $g_{B^{*}B\gamma},$ which opens the possibility for their determination, especially in particularly favorable regions of the $[g_{B^{*}B\pi}, g_{B^{*}B\gamma}]$ parameter space. Our analysis singles out the neutral modes \mbox{${B^{*}}^{0} ({D^{*}}^{0})\to \, B^0 (D^0)\, \gamma\,\gamma\,$} as the more relevant ones in relation with the determination of the the couplings under consideration. In addition to the direct radiative transition which is the sole contribution in neutral decays, and on which we concentrate in this paper, there is also the two-photon decay arising from bremsstrahlung in the charged \mbox{${B^{*}}^{+} ({D^{*}}^{+})\to \, B^+ (D^+)\, \gamma\,$} channel. Since this radiation overwhelms the direct mode, as we will show, one has to resort to the investigation of the neutral modes if one aims for a cleaner determination of the strong couplings. In section II we review the present status of the main decays of $B^{*}$ and $D^{*},$ with which the rare two photon decays must be compared. In section III we present the theoretical framework of our approach. Section IV contains the explicit treatment of the decay amplitudes. In the last section we summarize our predictions and we discuss certain features of the calculation. \section{The experimental and theoretical status of $B^{*}$ and $D^{*}$ principal decays} The vector mesons $B^{*}$ were firstly observed \cite{cusb} by the CUSB collaboration at the Cornell Electron Storage Ring (CESR) by detecting the photon signal from the radiative decay \mbox{$B^{*} \to \, B\, \gamma\,$}. This signal of $46$ MeV photons was confirmed in improved CUSB-II measurements, with the vector mesons produced at CESR at the $\Upsilon(5S)$ resonance \cite{cusb2}. Recently the process \mbox{$B^{*} \to \, B\, \gamma\,$} has been observed also at LEP with the various detectors \cite{opal} in a sample of over 4 milion hadronic $Z^{0}$ decays. The rate of the $B^{*}$-meson production relative to that of $B$ mesons is found generally to be consistent with the expectation from spin counting. It is expected that this production rate will be maintained in future $B$ experiments at hadron machines, like BTeV at Fermilab and LHC at CERN, where samples of the order of $10^{10}$ $B'$s are expected. Then, a fairly high sensitivity can be achieved in the study of $B^{*}$ decays and measurements of $B^{*}$ branching ratios of the order $10^{-7} - 10^{-8}$ could be accessible. In view of the small mass difference $\Delta M(B^{*}-B)=45.78\pm 0.35$ MeV \cite{PDG}, which forbids strong $B^{*}$ decays,the electromagnetic transition \mbox{$B^{*} \to \, B\, \gamma\,$} appears as the main decay of $B^{*}.$ This decay has been studied in a variety of theoretical models, including quark models \cite{Eichten,Godfrey,Ivanov}, the chiral bag model \cite{Singer} followed by effective chiral Lagrangian approaches for heavy and light mesons \cite{Cho,Amundson,Colangelo}, potential models \cite{Bank,DeFazio} and QCD sum rules \cite{Dosch,Aliev}. The predictions of these calculations span a range of nearly one order of magnitude for the expected decay widths, between $\Gamma({B^{*}}^0 ({B^{*}}^+)\rightarrow B^0(B^+) \gamma)=0.04(0.10)$ KeV \cite{Dosch} and $\Gamma({B^{*}}^0 ({B^{*}}^+)\rightarrow B^0(B^+) \gamma)=0.28(0.62)$ KeV \cite{Singer}, with most of the calculations \cite{Godfrey,Ivanov,Cho,Aliev} giving values closer to the larger values of Ref. \cite{Singer}. The $D^{*}$ meson was discovered more than twenty years ago \cite{Peruzzi} and has been subsequently studied in several experiments at different accelerators e.g. \cite{JADE,HRS,CLEO,ARGUS,CLEO1,ACCMOR}. The ${D^{*}}^+$ ($ M=2010.0\pm 0.5 \,{\rm MeV}$) and ${D^{*}}^0$ ($ M=2006.7\pm 0.5 \,{\rm MeV}$) have relatively little phase space for strong decay into $ D + \pi.$ The current PDG averages \cite{PDG} for the measured branching ratios of the observed decays are ${\rm Br}({D^{*}}^+ \rightarrow D^+\pi^0):{\rm Br}({D^{*}}^+ \rightarrow D^0\pi^+):{\rm Br}({D^{*}}^+ \rightarrow D^+\gamma)=(30.6\pm2.5)\%:(68.3\pm1.4)\%: (1.1+2.1-0.7)\% $ and ${\rm Br}({D^{*}}^0 \rightarrow D^0\pi^0):{\rm Br}({D^{*}}^0 \rightarrow D^0\gamma)=(61.9\pm2.9)\%:(38.1\pm2.9)\%.$ The most recent experiment on $ {D^{*}}^+$ decays \cite{CLEO1} gives more accurate branching ratios for the three decay channels as follow $(30.7\pm0.7)\%:(67.6\pm0.9)\%: (1.7\pm 0.6)\% .$ Although there are, by now, good data on the branching ratios, there is still no absolute measurement of any of the partial decay widths. The tightest upper limit has been established by the ACCMOR Collaboration at CERN \cite{ACCMOR} from the measurement of 127 ${D^{*}}^+$ events using a high-resolution silicon vertex detector, to be $\Gamma({D^{*}}^+)<131$ KeV. The other closest limit, obtained by the HRS collaboration \cite{HRS}, gives upper limits of 1.1 MeV and 2.1 MeV for the total decay widths of the charged and neutral ${D^{*}}'$ s. The decays of ${D^{*}}$ have also been treated extensively in a plethora of theoretical models. Many of the papers we mentioned concerning the ${B^{*}}$ decay \cite{Eichten,Godfrey,Ivanov,Singer,Cho,Amundson,Colangelo,Bank,DeFazio,Dosch,Aliev} discuss also the ${D^{*}}$ decays. In addition, we want to mention the early approaches \cite{Thens,Pham} with effective Lagrangian including symmetry-breaking, a relativistic quark model \cite{Donnel}, the study of ${D^{*}}$ decays using the chiral-bag model which contains pion exchange effects (pion loops) \cite{Miller}, the use of QCD sum rules \cite{Belyaev}, a chiral model with $ M_{c}\rightarrow \infty $ \cite{Troytskaya} and the comprehensive analysis of Kamal and Xu \cite{Kamal}. Recently \cite{Stewart}, the strength of the various decay channels of $D^*$ has been extracted from an analysis of the experimental branching ratios by the use of the chiral perturbation theory. In the ${D^{*}} $ case, the theoretical calculations again span an order of magnitude range for the prediction of the absolute decay widths, from a small width of $\Gamma({D^{*}}^0)\simeq (3-10)$ KeV \cite{Dosch,Troytskaya} to $\Gamma({D^{*}}^0)\simeq (60-120)$ KeV \cite{Cho,Pham,Donnel}, including fairly large uncertainties. It should be emphasized at this point that the chiral bag calculation ($\chi$) has offered the best estimation for the branching ratios \cite{Miller} ${\rm Br}^{\chi}({D^{*}}^+ \rightarrow D^+\pi^0):{\rm Br}^{\chi}({D^{*}}^+ \rightarrow D^0\pi^+):{\rm Br}^{\chi}({D^{*}}^+ \rightarrow D^+\gamma)=31.2\%:67.5\%:1.3\%$ and ${\rm Br}^{\chi}({D^{*}}^0 \rightarrow D^0\pi^0):{\rm Br}^{\chi}({D^{*}}^0 \rightarrow D^0\gamma)=64.3\%:35.7\%.$ The recent experiments have confirmed \cite{PDG,CLEO1} these relative ratios and dispersed the puzzling features which prevailed previously concerning the radiative branching ratio and the relative ratios of ${D^{*}}^+$ strong channels (see, e.g. Ref.\cite{Kamal}). The prediction of the chiral bag model \cite{Miller} is $\Gamma({D^{*}}^+\rightarrow all) =79$ KeV, $\Gamma({D^{*}}^0\rightarrow all)= 59.4$ KeV. Several of the other calculations result in fairly similar values \cite{Eichten,Godfrey,Kamal} as well as predicting for $\Gamma({D^{*}}^0)$ a value approximately $25\%$ smaller than for $\Gamma({D^{*}}^+).$ There are also calculations in which these widths are nearly equal \cite{Colangelo,Dosch} or, on the contrary, calculations giving $\Gamma({D^{*}}^+)$ to be at least twice larger than $\Gamma({D^{*}}^0)$ \cite{Aliev,Belyaev,Troytskaya}. The experimental and theoretical survey we presented here is obviously of direct relevance to our calculation as the absolute values of $B^{*}$ and $D^{*}$ widths will affect the probability of the \mbox{$B^{*} \to \, B\, \gamma\,\gamma\,$} and \mbox{$D^{*} \to \, D\, \gamma\,\gamma\,$} future detection. \section{A Model for Two-Photon Transition.} Substantial progress has been made in recent years in the treatment of the interactions of heavy mesons containing a single heavy quark with low momentum pions, by the use of an effective Lagrangian \cite{Wise,Burdmann,Yan}, the so-called ``Heavy Meson Chiral Lagrangian" (HM$\chi$ L),which embodies two principal symmetries of Quantum Chromodynamics (for a comprehensive review of this theoretical framework and its applications, see \cite{Casalbuoni}). At the leading order in an $1/M_{H}$ expansion ($M_{H}$ is the mass of the heavy meson) and the chiral limit for the light quarks ($m_{l}\rightarrow 0,l=u,d,s$), the Lagrangian carries flavor and spin symmetry in the heavy meson sector, as well as $SU(3)_L\otimes SU(3)_R$ chiral invariance in the light meson one. We adopt this framework for the calculation of the processes we study here, namely the $B^*(D^*)\rightarrow B(D) \gamma \gamma $ decays, and we shall use it to display the possible usefulness of these transitions for the determination of the $g_{B^{*}B\pi\,}, g_{D^{*}D\pi\,}$ and $g_{B^{*}B\gamma}$ couplings. The heavy vector ($B^*$ or $D^*$) and pseudoscalar ($B$ or $D$) mesons are represented by a $4\times 4$ Dirac matrix $H$, with one spinor index for the heavy quark and the second one for the light degree of freedom. \begin{equation} H=\frac{1+\not v}{2}\left[B^*_\mu \gamma^\mu - B\gamma_5\right],\,\,\,\bar{H} =\gamma_0 H^\dagger \gamma_0, \end{equation} where $v$ is the meson velocity, $v^\mu B^*_\mu =0$ and $B_\mu^*,B$ are the respective annihilation operators of the meson fields. We shall usually refer henceforth to \mbox{$B^{*} \to \, B\, \gamma\,\gamma\,$} with the understanding that the same treatment holds for \mbox{$D^{*} \to \, D\, \gamma\,\gamma\,$}. However we shall specify the two channels separately when numerical or other specific features make it necessary. The relevant interaction term of the ${\rm HM}\chi{\rm L},$ representing the coupling of heavy mesons to an odd number of pions, is given by \cite{Wise,Burdmann} \begin{equation} \label{lchir} {\cal L}^{\rm int}_{{\rm HM}\chi L} = g {\rm Tr} \left( \bar{H}_a \gamma_\mu \gamma_5 {\cal A}^\mu_{ab}H_b \right) \end{equation} where the axial current ${\cal A}^\mu$ is \begin{equation} {\cal A}^\mu =\frac{i}{2}\left(\xi^\dagger \partial^\mu\xi-\xi\partial^\mu\xi^\dagger\right) \end{equation} and $\xi={\rm exp}(i{\cal M}/f),$ with ${\cal M}$ being the usual $3\times 3$ matrix describing the octet of pseudoscalar Nambu-Goldstone bosons. The axial coupling constant $g$ is one of the basic parameters of HM$\chi$L, which is of direct import to our problem. $a,b$ denote light quark flavours ($a,b$=1,2,3) and f is the pion decay constant, $f=132$ MeV. Expanding the axial current and using the first term ${\cal A}^\mu=-(1/f)\partial^\mu {\cal M}+....$ one obtains the effective Lagrangian representing $B^*B$-pion and $B^*B^*$-pion interactions, which are the relevant ones in our problem. Thus, \begin{equation} \label{leff} {\cal L}^1_{\rm eff}= \left[-\frac{2g}{f}B^*_\mu \partial^{\mu}{\cal M} B^{\dagger} +h.c.\right]+\frac{2 g i}{f}\epsilon_{\alpha\beta\mu\nu}{B^*}^{\beta} \partial^{\mu}{\cal M} {{B^*}^{\dagger}}^\alpha v^\nu . \end{equation} The dimensionless ${B^*}B\pi$ coupling is defined as \cite{Casalbuoni} \begin{equation} \label{defgb} \langle \pi(q)\bar{B}(v_1)|{B^*}(v_2,\epsilon_2) \rangle = g_{B^{*} B\pi}(q^2) q_\mu \epsilon_{2}^{\mu} \end{equation} where $\epsilon^\mu$ is the polarization vector of $B^{*},$ with the physical coupling given by the limit $q^2\rightarrow m_\pi^2.$ We use the same normalization convention as in \cite{Casalbuoni}. Throughout this work, we assume that the variation of $g_{B^{*} B\pi}$ with $q^2$ in the region of our treatment can be safely neglected. Likewise, we define \begin{equation} \langle \pi(q)\bar{B^*}(v_1,\epsilon_1)|B^*(v_2,\epsilon_2) \rangle = g_{B^{*} B^* \pi}(q^2) \epsilon_{\alpha\beta\mu\nu}\epsilon_1^\alpha \epsilon_2^\beta q^\mu v_1^\nu, \end{equation} with the same remarks as above. We also note that the isospin symmetry requires \begin{eqnarray} \label{gcopb} g_{B^* B\pi} & \equiv & g_{{B^*}^+ B^0\pi^+} = -\sqrt{2}g_{{B^*}^+ B^+\pi^0} = \sqrt{2}g_{{B^*}^0 B^0\pi^0} \nonumber \\ &=& -g_{{B^*}^0 B^+\pi^-} \end{eqnarray} and the $g_{B^* B\pi} $ so defined is the commonly used in the literature. The (\ref{gcopb}) relation holds similarly for $g_{B^* B^*\pi},g_{D^* D\pi}$ and $g_{D^* D^*\pi}$ couplings. Now, from (\ref{lchir})-(\ref{leff}), using the definition (\ref{defgb}) one has \begin{eqnarray} \label{gchir} g_{B^* B\pi} &=&\frac{2 M_B}{f}g=g_{B^* B^*\pi} \nonumber \\ g_{D^* D\pi} &=&\frac{2 M_D}{f}g=g_{D^* D^*\pi}. \end{eqnarray} Note that, in deriving (\ref{gchir}) one assumes $B,B^*(D,D^*)$ mass degeneracy. In order to calculate the \mbox{$B^{*} \to \, B\, \gamma\,\gamma\,$} decay width we use the interaction Lagrangian (\ref{leff}) to the leading order in chiral perturbation theory, which is an appropriate tool here, in view of the smallness of $M_{B^*}-M_B.$ The calculation of the radiative processes $B^*(D^*)\rightarrow B(D)\gamma\gamma$ obviously requires the incorporation of the electromagnetic interaction in our Lagrangian (\ref{leff}), which is performed \cite{Stewart,Cheng2} by the usual procedure of gauging the Lagrangian with the $U(1)$ photon field. This leads to the replacement of the derivative operators in the Lagrangian by covariant derivatives containing the photon field, explicitly exhibited in \cite{Cheng2}. Nevertheless, the new Lagrangian still does not provide for couplings to induce the observed \mbox{$B^{*} \to \, B\, \gamma\,$}, \mbox{$D^{*} \to \, D\, \gamma\,$} magnetic dipole transitions. This necessitates the introduction of an additional term in the Lagrangian, a contact gauge invariant interaction proportional to the electromagnetic field $F_{\mu\nu},$ which is given by \cite{Stewart,Cheng2} \begin{equation} \label{lem} {\cal L}= \frac{e\mu}{4} {\rm Tr} \left( \bar{H}_{a} \sigma_{\mu\nu} F_{\mu\nu}H_{b}\delta_{ab} \right), \end{equation} where $\mu$ is the strength of this anomalous magnetic dipole interaction, having mass dimension $[1/M].$ Additional terms arising from an $1/M_{H}$ expansion exist \cite{Stewart,Cheng2}, however several of them, including the radiation of the heavy quark can be absorbed in equation (\ref{lem}) by redefining $\mu.$ In the present paper, we shall consider $\mu$ as an effective coupling, representing the strength of the $B^*(D^*)B(D)\gamma$ transition. Expanding $H$ in terms of the components, one obtains for the additional electromagnetic interaction \begin{equation} \label{leffem} {\cal L}^2_{\rm eff}=-e\mu F^{\mu\nu} \left[i{B^*_\mu}^{\dagger}B^*_\nu +\epsilon_{\alpha\beta\mu\nu}v^\alpha (B^{\dagger}B^*_\beta+h.c.)\right] \end{equation} which exhibits $B^*B\gamma$ and $B^*B^*\gamma$ couplings with equal strength, as given by the heavy quark symmetry. From (\ref{leffem}) we obtain the respective vertices, which are \begin{eqnarray} \langle \gamma(k,\epsilon) \bar{B^*}(v_{1},\epsilon_{1}) |B^*(v_{2},\epsilon_{2})\rangle&=& e \mu M_{B^*}(\epsilon_{1}.k\epsilon.\epsilon_{2}- \epsilon_{2}.k\epsilon.\epsilon_{1}) \\ \langle \gamma(k,\epsilon) \bar{B}(v_{1}) |B^*(v_{2},\epsilon_{2})\rangle &=& -i e M_{B^*}\mu \epsilon_{\mu\nu\alpha\beta}{\epsilon}^{\mu} k^{\nu}v_{2}^{\alpha}\epsilon_2^{\beta}. \end{eqnarray} The propagator of the heavy vector meson is given by $-i(g^{\mu\nu}-v^\mu v^\nu)/2\left[(v.k)-\Delta/4\right],$ where $\Delta = M_{B^*}-M_B$ and $v,k$ are the velocity and the residual momentum. The propagator of the heavy pseudoscalar meson is $i/2\left[(v.k)+3\Delta/4\right]$ \cite{Casalbuoni}. Now, considering Lagrangians (\ref{leff}) and (\ref{leffem}), as well as the axial anomaly responsible for the $\pi^0\rightarrow \gamma\gamma,$ we classify the diagrams contributing to \mbox{$B^{*} \to \, B\, \gamma\,\gamma\,$} in the leading order of chiral perturbation theory as follows: There is the diagram ${B^*}^0\rightarrow B^0 ``\pi"\rightarrow B^0 \gamma \gamma$ (Fig.1), via a virtual pion (the somewhat different situation in \mbox{$D^{*} \to \, D\, \gamma\,\gamma\,$} will be analyzed in the last section). This diagram contains the known strength of the pion axial anomaly. Then, there is the loop graph ${B^*}^0\rightarrow ({B^*}^{+} \pi^{-})\rightarrow B^0 \gamma \gamma$ (Fig.2), with the photons radiated from the virtual charged pion in the loop, with additional graphs of the same class, as specified in the next section. The first graph is proportional to the $g_{B^* B\pi}$ coupling, while the loop graph contains the $g_{B^* B\pi} g_{B^* B^*\pi}$ product. In addition we have the tree level diagram with two insertions of the magnetic operator defined in (\ref{leffem}),leading to ${B^*}^0\rightarrow {B^*}^0\gamma\rightarrow B^0\gamma \gamma$ (Fig.3). This graph depends only on the $\mu$ magnetic moment. Finally we have a class of one loop diagrams which involve both the magnetic moment and the strong coupling, which is exhibited in Fig.(4-6). We did not include the contribution of diagrams containing three heavy meson propagators which is negligible. Needless to say, the determination of $g_{B^* B\pi}$ would be simpler, should the first two graphs dominate. However, this is not true for $D^*$ decay, while it can be true for the $B^*$ decay for an opportune range of parameters as we discuss in the next section. We remark at this point that corrections to (\ref{lchir}) which arise from higher terms in the $1/M_H$ expansion as well as in chiral breaking have also been investigated \cite{Stewart,Casalbuoni,Cheng,Boyd}. A comprehensive inclusion of these corrections in the calculation of the two-photon decays of heavy vector mesons is beyond our scope in this first treatment of these processes. Nevertheless, we note that we shall use physical masses for the degenerate doublet of heavy mesons both in the loop propagators and in the decay calculations, moreover the chiral loops included are themselves of order $1/M_{H}.$ The terms we include are the leading ones in chiral perturbation theory, and are of the same order in an $1/N_c$ expansion; moreover to this order there are no counterterms \cite{Leibovich}. It is appropriate to mention now that the $g_{B^* B\pi},g_{D^* D\pi}$ couplings were estimated in recent years by the use of a variety of theoretical techniques, like QCD sum rules \cite{Dosch,Belyaev}, soft pion approximation \cite{Cola} and other methods \cite{Cho,Amundson,Stewart,Cheng,Boyd,Manohar}. Generally, the values of $g$ obtained in these works are in the range $g=0.25-0.7,$ significantly smaller than the quark model result of $g=1$ or of modified quark models \cite{Pham,Isgur,Nussinov} which brought this value slightly below one. The most recent determinations of $g$ include an analysis \cite{Casalbuoni} of various theoretical approaches which leads to a ``best estimate" of $g=0.38,$ a recent lattice determination giving $0.42(4)(8)$ \cite{UKQCD} and the analysis of Stewart \cite{Stewart} which incorporates symmetry breaking terms in the Lagrangian and obtains $g=0.27^{+0.9}_{-0.4}.$ Finally, the experimental limit $\Gamma({D^*}^+)<131$ KeV \cite{ACCMOR} puts an upper limit of $g<0.71,$ using $\Gamma({D^*}^+\rightarrow D^0\pi^+ +D^+\pi^0)= (g^2/4\pi f^2) |\vec{p}_\pi|^3.$ The existing theoretical estimates we mentioned, give $0.04 {\rm KeV} <\Gamma(\mbox{${B^{*}}^{0} \to \, B^0\, \gamma\,$})< 1 {\rm KeV}$ and $0.10 {\rm KeV} <\Gamma(\mbox{${B^{*}}^{+} \to \, B^+\, \gamma\,$})< 1 {\rm KeV},$ where we allowed for a slightly higher upper limit. We redefine the magnetic coupling in eq.(\ref{leffem}) to a dimensionless quantity $\bar{\mu}=M_{B^*}\mu=g_{{B^*}^0B^0\gamma},$ and $\bar{\mu}_{+}=M_{B^*}\mu_{+}=g_{{B^*}^+B^+\gamma},$ then the above limits give the following ranges, $2.2<\bar{\mu}<11.0,$ and $3.5<\bar{\mu}_+<11.0.$ \section{The decay amplitudes.} We present now the explicit expressions of the decay amplitudes, which in our approach, to leading order in chiral perturbation theory, consist of the contribution of the anomaly graph (Fig. 1), the tree level graph (Fig.3) and the loop graphs (Figs. 2,4-6). In presenting the differential decay distribution, we use the following variables: \begin{eqnarray} s&=&(p-p^\prime)^2=(k_1+k_2)^2\nonumber \\ t&=&(p-k_1)^2=(p^\prime+k_2)^2\nonumber \\ u&=&(p-k_2)^2=(p^\prime+k_1)^2 \end{eqnarray} with \begin{equation} t+s+u=M_{B^*}^2+M_{B}^2, \end{equation} where $k_1, k_2$ are the four momenta of the two photons and $p,p^\prime$ are the four-momenta of the decaying $B^*$ and the final $B$ respectively. The allowed ranges for $s$ and $t$ are \begin{eqnarray} 0\leq s&\leq& (M_{B^*}-M_{B})^2, \,\,\,\,\,t_{-}\leq t\leq t_{+}\nonumber \\ t_{\pm} &=& \frac{1}{2}\left[(M_{B^*}^2+M_{B}^2-s)\pm \sqrt{(M_{B^*}^2+M_{B}^2-s)^2-4M_{B^*}^2M_{B}^2}\right]. \end{eqnarray} The amplitudes are given for \mbox{${B^{*}}^{0} \to \, B^0\, \gamma\,\gamma\,$} and we shall remark on the changes appearing in \mbox{${D^{*}}^{0}\to \, D^0\, \gamma\,\gamma\,$} whenever required. The amplitude from the anomaly graph, mediated by a pion is: \begin{eqnarray} \label{Ampan} {\rm A}_{\rm anomaly}(\mbox{${B^{*}}^{0} \to \, B^0\, \gamma\,\gamma\,$})=\frac{\alpha g_{B^*B\pi}}{\sqrt{2}\pi f} \epsilon^\mu_{B^*} \epsilon_{1}^\lambda \epsilon_{2}^\gamma \frac{1}{s -m_\pi^2} \epsilon_{\lambda\gamma\tau\rho}k_1^\tau k_2^\rho (k_1+k_2)_\mu, \end{eqnarray} where $\epsilon_{B^*}, \epsilon_1, \epsilon_2$ are the polarization vectors of the heavy vector meson ${B^*}$ and the two photons respectively. There are additional contributions from $\eta,\eta^{\prime}$ which are not specified in (\ref{Ampan}). As we shall describe in the next section, their contribution is rather small and we may safely neglect them at this stage. For the tree level graph we find \begin{eqnarray} \label{Atree} {\rm A}_{\rm tree}(\mbox{${B^{*}}^{0} \to \, B^0\, \gamma\,\gamma\,$})&=& \frac{4\pi \alpha \bar{\mu}^2}{ M^{*}_{B}} \epsilon_{\gamma\delta\alpha\beta}{\epsilon_{B^*}}_ {\sigma} {p^{\prime}}^\alpha \nonumber \\ && \left[\frac{ \epsilon_{2}^\gamma k_2^{\delta}(\epsilon_{1}^{\sigma} k_{1}^{\beta}-\epsilon_{1}^{\beta} k_{1}^{\sigma})}{t-M_{B^*}^2}+ \frac{ \epsilon_{1}^\gamma k_1^{\delta}(\epsilon_{2}^{\sigma} k_{2}^{\beta}-\epsilon_{2}^{\beta} k_{2}^{\sigma})}{u-M_{B^*}^2}\right] \end{eqnarray} The loop contribution which depends only on the strong coupling is given by a sum of several diagrams. In addition to that explicitly shown in Fig.2 there are diagrams with one photon radiated by the virtual pion in the loop and the other emitted from the $B^* B^* \pi,$ $B^* B \pi$ vertices or both photons emitted from these vertices, or both photons emitted from the loop by the $\pi\pi\gamma\gamma$ vertex. In the limit of photon momenta small compared to the pion mass, which we find to be a suitable approximation, the class of diagrams of Fig.2 gives \begin{eqnarray} \label{Amploop1} {\rm A}^{1}_{\rm loop}(\mbox{${B^{*}}^{0} \to \, B^0\, \gamma\,\gamma\,$}) &=& \frac{\alpha g_{B^* B\pi} g_{B^* B^*\pi}}{8\pi M_{B^*}} \epsilon_{\mu\eta\alpha\beta} \epsilon^\mu_{B^*} v^{\beta} \epsilon_{1}^\lambda \epsilon_{2}^\gamma (k_1+k_2)^\eta \nonumber \\ & & \times \left[3\,\left( g^\alpha_\gamma v_\lambda + g^\alpha_\lambda v_\gamma \right)+\frac{1}{9 m_{\pi}}\left( g^\alpha_\gamma {k_2}_\lambda + g^\alpha_\lambda {k_1}_\gamma \right)\right]. \end{eqnarray} Finally other loop contributions to the \mbox{$B^{*} \to \, B\, \gamma\,\gamma\,$} decay come from diagrams where both the strong coupling and the magnetic one are involved. There are diagrams with one photon radiated by the virtual pion in the loop and the other emitted from the ingoing $B^*$ particle through the $B^*B^*(B)\gamma$ vertices (Fig.4), or from the outgoing particle $B^*$ which becomes $B$ through the $B^*B\gamma$ vertex (Fig.5), or from the $B$ in the loop which becomes $B^*$ through the $B^*B\gamma$ vertex (Fig.6). The amplitude corresponding to Fig.4 with a $B^*B^*\gamma$ vertex is: \begin{eqnarray} \label{Amploop2} {\rm A}^{2}_{\rm loop}(\mbox{${B^{*}}^{0} \to \, B^0\, \gamma\,\gamma\,$}) &=&\frac{11\alpha (M_{B^*}-M_B)g_{B^* B^*\pi}g_{B^* B\pi}\bar{\mu}}{16\pi M^2_{B^*}}\epsilon_{\alpha\sigma\gamma\delta}\nonumber \\ && \left[(k_1^\alpha(\epsilon_{B^*}.\epsilon_1)- \epsilon_1^\alpha(\epsilon_{B^*}.k_1)) \frac{(p-k_1)^\delta\epsilon_2^\gamma k_2^\sigma}{t-M^2_{B^*}} \right. \nonumber \\ &+& \left. (k_2^\alpha(\epsilon_{B^*}.\epsilon_2)- \epsilon_2^\alpha(\epsilon_{B^*}.k_2)) \frac{(p-k_2)^\delta\epsilon_1^\gamma k_1^\sigma}{u-M^2_{B^*}} \right] \end{eqnarray} The amplitude corresponding to Fig.4 where a $B^*B\gamma$ vertex replaces the $B^*B^*\gamma$ appearing in Fig.4, is \begin{eqnarray} \label{Amploop3} {\rm A}^{3}_{\rm loop}(\mbox{${B^{*}}^{0} \to \, B^0\, \gamma\,\gamma\,$}) &=&\frac{9\alpha m_{\pi}^2 g_{B^* B\pi}^2\bar{\mu}}{4\pi M_{B^*}}\epsilon_{\mu\nu\alpha\beta}\epsilon^{\nu}_{B^*} v^{\alpha}v_{\gamma}\nonumber \\ && \left[\frac{\epsilon_1^{\mu}k1^{\beta}\epsilon_2^{\gamma}} {t-M^2_{B^*}}+ \frac{\epsilon_2^{\mu}k2^{\beta}\epsilon_1^{\gamma}} {u-M^2_{B^*}} \right] \end{eqnarray} The amplitude corresponding to Fig.5 with $B$ in the loop is: \begin{eqnarray} \label{Amploop4} {\rm A}^{4}_{\rm loop}(\mbox{${B^{*}}^{0} \to \, B^0\, \gamma\,\gamma\,$}) &=& \frac{3 \alpha m_{\pi}^2 g_{B^* B\pi}^2\bar{\mu}}{8 \pi M_{B^*}^2} \epsilon_{\mu\nu\alpha\rho} \epsilon^\sigma_{B^*} (g_{\gamma}^{\rho}v_{\sigma}+g_{\sigma}^{\rho}v_\gamma) \nonumber \\ && \left[ \frac{(p-k_1)^\alpha\epsilon_2^\mu k_2^\nu\epsilon_1^{\gamma}} {t-M^2_{B^*}}+ \frac{(p-k_2)^\alpha\epsilon_1^\mu k_1^\nu\epsilon_2^{\gamma}} {u-M^2_{B^*}} \right] \end{eqnarray} and the one with $B^*$ in the loop is \begin{eqnarray} \label{Amploop5} {\rm A}^{5}_{\rm loop}(\mbox{${B^{*}}^{0} \to \, B^0\, \gamma\,\gamma\,$}) &=& \frac{3 \alpha m_{\pi}^2 g_{B^* B^*\pi}^2\bar{\mu}}{16 \pi M_{B^*}^3} \epsilon_{\alpha\beta\gamma\delta} \epsilon^\beta_{B^*}v^{\delta} \epsilon_{\eta\alpha\xi\rho}\epsilon_{\mu\nu\sigma\eta} (g^{\gamma\rho}v_{\tau}+g^{\gamma}_{\tau}v^\rho) \nonumber \\ && \left[ \frac{(p-k_1)^\xi(p-k_1)^\sigma\epsilon_1^\tau \epsilon_2^\mu k_2^\nu}{t-M^2_{B^*}}+ \frac{(p-k_2)^\xi(p-k_2)^\sigma\epsilon_1^\mu \epsilon_2^\tau k_1^\nu}{u-M^2_{B^*}} \right] \end{eqnarray} Finally the amplitude corresponding to Fig.6 is: \begin{eqnarray} \label{Amploop6} {\rm A}^{6}_{\rm loop}(\mbox{${B^{*}}^{0} \to \, B^0\, \gamma\,\gamma\,$}) &=& \frac{11 \alpha g_{B^* B\pi}^2 \bar{\mu}_{+}}{24 \pi M_{B^*}^3} \epsilon_{\mu\nu\alpha\beta} \epsilon^\sigma_{B^*} \left[\frac{1}{3}(k_1+k_2)_\sigma \epsilon_1^\alpha\epsilon_2^\mu k_1^\nu k_2^\beta \right. \nonumber \\ &+& \left. \frac{1}{2}p^\alpha\left[\epsilon_2^\mu k_2^\beta ({\epsilon_1}_\sigma k_1^\nu+\epsilon_1^\nu {k_1}_\sigma)+ \epsilon_1^\mu k_1^\beta ({\epsilon_2}_\sigma k_2^\nu+\epsilon_2^\nu {k_2}_\sigma)\right]\right]. \end{eqnarray} In the $B^*$ decay, the pions are the sole contributions in the loop, while in the \mbox{$D^{*} \to \, D\, \gamma\,\gamma\,$} calculation we include both pions and kaons. Let us call A the sum of all the amplitudes: \begin{equation} {\rm A}={\rm A}_{\rm anomaly}+{\rm A}_{\rm tree}+ {\rm A}_{\rm loops} \end{equation} where ${\rm A}_{\rm loops}$ is the sum of all the amplitudes $A^{i}_{\rm loops},\,i=1,...,6$ which come from the loops. The square of the above amplitude, when averaged over the initial spin and summed over the final spins is: \begin{equation} |\bar{A}|^2=\frac{1}{2}\sum_{\rm spins} |A|^2, \end{equation} where we have included the factor $\frac{1}{2}$ in order to take into account two identical particles in the final state. The differential decay rate of photon energy is obtained by integrating the following expression over the variable $t,$ \begin{equation} \label{dgamma} \frac{d\Gamma}{ds}=\frac{1}{(2\pi)^3}\frac{1}{32 M_{B*}^3} \int_{t_{-}}^{t_{+}} |\bar{A}|^2 dt. \end{equation} There is a major difference in the anomaly contribution of the $B^*$ and $D^*$ decays. Since the $\pi^0$ appears in the physical region in the \mbox{${D^{*}}^{0} \to \, D^0\, \pi^0\,$} decay we have to isolate the on-shell $\pi^0$ decay in the \mbox{${D^{*}}^{0}\to \, D^0\, \gamma\,\gamma\,$} mode. Hence, for the $D^*$ case we limit ourselves in the integration of $ d\Gamma/ds $ to a region which goes from $s =0$ up to $20$ MeV away from the pion mass. Using now the physical masses of $ M_{B^*}, M_{B}, M_{D^*} M_{D}$ \cite{PDG} and the eq.(\ref{gchir}) we obtain for the decay rates of \mbox{$B^{*} \to \, B\, \gamma\,\gamma\,$}, the following expression: \begin{eqnarray} \label{gbsgg} \Gamma(\mbox{$B^{*} \to \, B\, \gamma\,\gamma\,$}) &=& 3.40\times 10^{-16} g^2 + 1.53\times 10^{-12} g^4 + 4.81\times 10^{-17} g^3 \bar{\mu} \nonumber \\ &+& 1.53\times 10^{-13} g^4 \bar{\mu}+ 4.71\times 10^{-17} g \bar{\mu}^2 + 9.81\times 10^{-14} g^4 \bar{\mu}^2 \nonumber \\ &+& 2.65\times 10^{-16} g^2 \bar{\mu}^3 + 1.38\times 10^{-16} \bar{\mu}^4+ 2.67\times 10^{-16} g^3 \bar{\mu}_{+} \nonumber \\ &+& 2.90\times 10^{-17}g^4 \bar{\mu}_{+} + 8.94\times 10^{-20} g^4 \bar{\mu} \bar{\mu}_{+} \nonumber \\ &+& 9.11\times 10^{-20} g^2 \bar{\mu}^2 \bar{\mu}_{+}+ 7.21\times 10^{-17} g^4 \bar{\mu}_{+}^2. \end{eqnarray} As we can see this is a function of both the strong coupling $g$ and the magnetic dipole strength $\bar{\mu},\bar{\mu}_{+}$ which represent the effective $g_{{B^{*}}^0B^0\gamma}$ and $g_{{B^{*}}^+B^+\gamma}$ couplings. For the \mbox{$D^{*} \to \, D\, \gamma\,\gamma\,$} we have: \begin{eqnarray} \label{gdsgg} \Gamma(\mbox{$D^{*} \to \, D\, \gamma\,\gamma\,$}) &=& 2.52\times 10^{-11} g^2 + 5.85\times 10^{-10} g^4 + 1.79\times 10^{-12} g^3 \bar{\mu} \nonumber \\ &+& 4.43\times 10^{-11} g^4 \bar{\mu} + 1.30\times 10^{-12} g \bar{\mu}^2 + 3.50\times 10^{-11} g^4 \bar{\mu}^2 \nonumber \\ &+& 2.42\times 10^{-13} g^2 \bar{\mu}^3 + 2.18\times 10^{-12} \bar{\mu}^4 + 1.01\times 10^{-11} g^3 \bar{\mu}_{+} \nonumber \\ &+& 2.95\times 10^{-13} g^4 \bar{\mu}_{+} + 2.11\times 10^{-14} g^4 \bar{\mu} \bar{\mu}_{+} \nonumber \\ &+& 1.70\times 10^{-14} g^2 \bar{\mu}^2 \bar{\mu}_{+} + 2.05\times 10^{-12} g^4 \bar{\mu}_{+}^2. \end{eqnarray} In this case we can relate the magnetic coupling to the strong coupling using the existing experimental informations on $\Gamma(\mbox{${D^{*}}^{0} \to \, D^0\, \pi^0\,$}): \Gamma(\mbox{${D^{*}}^{0} \to \, D^0\, \gamma\,$})$ of $(61.9\pm 2.9)\%:(38.1\pm 2.9)\%,$ which gives rise to the relation: $ \bar{\mu} \simeq 6.6 g,$ and $\Gamma(\mbox{${D^{*}}^{+} \to \, D^0\, \pi^+\,$}): \Gamma(\mbox{${D^{*}}^{+} \to \, D^+\, \gamma\,$})$ of $(67.6\pm 0.9)\%:(1.7\pm 0.6)\%,$ which gives rise to the relation: $ \bar{\mu}_{+} \simeq 1.7 g.$ Then we can write the decay width solely as a function of $g,$ which is a crucial step in the engagement of this decay as a tool for measuring $g.$ \begin{eqnarray} \label{pos} \Gamma(\mbox{$D^{*} \to \, D\, \gamma\,\gamma\,$})&=& 2.52\times 10^{-11} g^2 + 5.66\times 10^{-11} g^3 \nonumber \\ &+& 4.76\times 10^{-9} g^4 + 3.64\times 10^{-10} g^5 + 1.53\times 10^{-9} g^6. \end{eqnarray} We used in eqs.(\ref{gbsgg})-(\ref{gdsgg}) the same notation of $\mu,\mu_+$ for the magnetic moment strength in both the charmed and beauty sectors although they are probably not equal for the physical processes. However since in the charm case, the magnetic coupling has been related to the strong one, the $\mu, \mu_+$ will denote in the rest of the paper the strength of the ${B^*}^{0,+}\rightarrow B^{0,+}\gamma$ transitions. The experimentally measured branching ratios of \mbox{$D^{*} \to \, D\, \pi\,$}, \mbox{$D^{*} \to \, D\, \gamma\,$} lead to relations between $\bar{\mu}, \bar{\mu}_+$ and $g$ modulo an unknown phase. We have allowed also for the possibilities of negative relative signs among $\bar{\mu}, \bar{\mu}_+$ and $g$ and as it turns out, this affects only slightly the numerical picture, due to the fact that the main contribution is given by quadratic terms. In eq.(\ref{neg}) we give for comparison the expression obtained for $\bar{\mu}=-6.6 g, \bar{\mu}_+=-1.7 g.$ \begin{eqnarray} \label{neg} \Gamma(\mbox{$D^{*} \to \, D\, \gamma\,\gamma\,$})_{\bar{\mu}={\rm neg}}&=& 2.52\times 10^{-11} g^2 + 5.66\times 10^{-11} g^3 \nonumber \\ &+& 4.70\times 10^{-9} g^4 - 3.64\times 10^{-10} g^5 + 1.53\times 10^{-9} g^6. \end{eqnarray} This ambiguity will be further discussed in the next section. The difference between the rates of $B^*$ and $D^*$ is mainly due to the different phase space. In discussing the two photon radiative decays, we shall refer in the next section to the following quantities \begin{equation} \label{BRb} {\rm BR}(\mbox{$B^{*} \to \, B\, \gamma\,\gamma\,$}) = \frac{\Gamma(\mbox{${B^{*}}^{0} \to \, B^0\, \gamma\,\gamma\,$})}{\Gamma({B^*}^0)}= \frac{\Gamma(\mbox{${B^{*}}^{0} \to \, B^0\, \gamma\,\gamma\,$})}{\Gamma(\mbox{${B^{*}}^{0} \to \, B^0\, \gamma\,$})} \end{equation} \begin{equation} \label{BRd} {\rm BR}(\mbox{$D^{*} \to \, D\, \gamma\,\gamma\,$}) = \frac{\Gamma(\mbox{${D^{*}}^{0}\to \, D^0\, \gamma\,\gamma\,$})}{\Gamma({D^*}^0)}= \frac{\Gamma(\mbox{${D^{*}}^{0}\to \, D^0\, \gamma\,\gamma\,$})}{\Gamma(\mbox{${D^{*}}^{0} \to \, D^0\, \gamma\,$})+\Gamma(\mbox{${D^{*}}^{0} \to \, D^0\, \pi^0\,$})}. \end{equation} \section{Discussion and Conclusions.} The formalism we have presented refers to the decays of the neutral heavy vector mesons ${B^*}^0,{D^*}^0,$ as it will the numerical analysis of our results, which will be given below. For the charged decays, \mbox{${B^{*}}^{+} \to \, B^+\, \gamma\,\gamma\,$} and \mbox{${D^{*}}^{+}\to \, D^+\, \gamma\,\gamma\,$} one has to consider also the bremsstrahlung emission which appears in diagrams of Fig.3 and Fig.4 and additional ones.The bremsstrahlung radiation comes from the initial or the final charged particles. To give an idea of this effect we have calculated the part of the bremsstrahlung amplitude which is due to radiation from the final $B^+$ particle in the amplitude ${B^*}^+\rightarrow (B^+\gamma)\rightarrow B^+\gamma\gamma.$ It is \begin{equation} \label{Ampbrem} A^{B^+}_{\rm brem}= 4\pi\alpha g_{B^* B\gamma} \epsilon^\mu_{B^*} \epsilon_{1}^\lambda \epsilon_{2}^\gamma p^\alpha \left[ \frac{ \epsilon_{\lambda \mu \alpha\beta} k_1^\beta p^\prime_\gamma}{(t-M_B^2)} + \frac{\epsilon_{\gamma \mu \alpha\rho} k_2^\rho p^\prime_\lambda}{(u-M_B^2)} \right]. \end{equation} This amplitude (and the ones given by ${B^*}^+$ radiation) have to be added in order to get the full amplitude for the \mbox{${B^{*}}^{+} \to \, B^+\, \gamma\,\gamma\,$} decay. An estimate of the bremsstrahlung decay width, from (\ref{Ampbrem}) only, using for the unknown $g_{B^*B\gamma}$ vertex a value leading to $\Gamma(\mbox{${B^{*}}^{+} \to \, B^+\, \gamma\,$})=0.14$ KeV \cite{Casalbuoni} leads to a decay width of $\sim 10^{-9}$ GeV for $k_1,k_2\geq 10$ MeV, considerably larger than (\ref{gbsgg}). The use of the charged \mbox{${B^{*}}^{+} \to \, B^+\, \gamma\,\gamma\,$} thus involves a different type of analysis in view of the relative size of the different components of $|A(\mbox{${B^{*}}^{+} \to \, B^+\, \gamma\,\gamma\,$})|^2$ and is less useful for a determination of $g.$ A similar situation is encountered for \mbox{${D^{*}}^{+}\to \, D^+\, \gamma\,\gamma\,$}. Thus, we concentrate here on the ``safer'' neutral decays and we relegate the discussion of the charged decays to a separate work, in which we consider the usefulness of $\Gamma(\mbox{${B^{*}}^{+} \to \, B^+\, \gamma\,\gamma\,$})$ for the determination of $g_{{B^*}^+B^+\gamma} $ \cite{Dafne}. We proceed now to analyse the results on the two decays separately and we start with \mbox{${D^{*}}^{0}\to \, D^0\, \gamma\,\gamma\,$} transition for which the rate (\ref{pos}) was obtained. The many theoretical estimates for $g$ we mentioned in Sec.3 are spread over the range $0.25<g<1$ (we also remind the reader that the experimental result \cite{ACCMOR} on the upper limit of $\Gamma({D^*}^+\rightarrow all)$ can be interpreted as $g<0.71$). Using this range and, eqs.(\ref{pos}),(\ref{neg}) we can establish the expectation \begin{equation} \Gamma(\mbox{${D^{*}}^{0}\to \, D^0\, \gamma\,\gamma\,$})\simeq(0.022-6.73){\rm eV}. \end{equation} The most promising feature of the present analysis arises when we use the existing experimental informations on $\Gamma(\mbox{${D^{*}}^{0} \to \, D^0\, \pi^0\,$}):\Gamma(\mbox{${D^{*}}^{0} \to \, D^0\, \gamma\,$})$ of $(61.9\pm 2.9)\%:(38.1\pm2.9)\%$ to transform eq.(\ref{BRd}) into a ratio of $\Gamma(\mbox{${D^{*}}^{0}\to \, D^0\, \gamma\,\gamma\,$})$ to the total ${D^*}^0$ width which becomes proportional to $g^2.$ Using \begin{equation} \Gamma(\mbox{${D^{*}}^{0} \to \, D^0\, \pi^0\,$})=\frac{1}{12\pi}\frac{g^2}{f^2} |\vec{p}_\pi|^3=1.25\times 10^{-4} g^2 {\rm GeV} \end{equation} and the $(61.9\pm 2.9)\%:(38.1\pm 2.9)\%$ relative branching ratio, we arrive at \begin{equation} \label{gamma} \Gamma({D^*}^0\rightarrow all)= (2.02\pm 0.12)\times 10^{-4} g^2 {\rm GeV}. \end{equation} Thus from (\ref{gamma}) with (\ref{pos}) we can obtain a branching ratio {\it which depends on} $g^2$ {\it only:} \begin{equation} \label{gd} {\rm Br}(\mbox{${D^{*}}^{0}\to \, D^0\, \gamma\,\gamma\,$})= \frac{(0.025+ 0.057 g + 4.76 g^2 + 0.36 g^3 +1.53 g^4)\times 10^{-9}g^2} {2.02\times 10^{-4} g^2}. \end{equation} With our model for \mbox{${D^{*}}^{0}\to \, D^0\, \gamma\,\gamma\,$}, the measurement of this ratio will thus constitute a measurement of the $g$ coupling. Using again the accepted expectation of $0.25\leq g\leq 1,$ we predict \begin{equation} \label{bdsgg} {\rm Br}(\mbox{$D^{*} \to \, D\, \gamma\,\gamma\,$})=\frac{\Gamma(\mbox{${D^{*}}^{0}\to \, D^0\, \gamma\,\gamma\,$})}{\Gamma(D^*)}= (0.16-3.3)\times 10^{-5}. \end{equation} A few remarks are in order. Firstly, the sign question The observed branching ratios do not afford to estabilish experimentally the sign of $g/g_{{D^*}^0D^0\gamma}.$ On the other hand, there is theoretical support from the analysis of Stewart \cite{Stewart} on the positive sign of this ratio. However even if we assume opposite sign for various pairs of the couplings, we found that the changes are rather small, and this is explicitly exhibited in Table 2, and included in Eq.(\ref{bdsgg}). The differential distribution in the $s$-variable can also be used to learn about the value of $g,$ due to the fact that the different contributions depend on different powers of $g.$ Finally we remark that the contributions from the diagrams exhibited in Fig.4 are rather small, as a result of two heavy propagators. The main contributions are those of the anomaly and of the graphs of Fig.(2) and (3). In Figs.(7) and (8) we present the differential distributions in $s$ for $g=0.7$ and $g=0.25.$ In the latter, the contributions containing a higher power of $g$ are diminished and the effect of the anomaly becomes visible in the higher end of the spectrum. Turning now to the \mbox{$B^{*} \to \, B\, \gamma\,\gamma\,$}, we have a rather different situation. Firstly, there is only one major decay of $B^*,$ namely \mbox{$B^{*} \to \, B\, \gamma\,$}, which precludes an analysis like in $D^*$ decays. The branching ratio (\ref{BRb}) depends on three parameters, $g_{B^*B\pi}$ (or $g$), $g_{{B^*}^0B^0\gamma}$ (or $\mu$) and $g_{{B^*}^+B^+\gamma}$ (or $\mu^+$). At this point, we rely on the theoretical estimates presented in Section 2 and contain our analysis to the regions given by existing calculations. Now, an inspection of eq.(\ref{gdsgg}) shows that $\mu_+,$ which appears only in diagram of Fig.6 has a very little effect on the rate, whether $\mu,\mu_+$ are at the lowest or at the highest end of their value, for any value of $g.$ Hence, we continue our analysis in the parameter space of $[g,\mu]$ only. In table 2 we present the values of ${\rm Br}(\mbox{$B^{*} \to \, B\, \gamma\,\gamma\,$})$ for different values of $g$ and the two extreme values of $\mu,$ corresponding to $\Gamma(\mbox{$B^{*} \to \, B\, \gamma\,$})=40 {\rm eV}$ and $1{\rm KeV}.$ Again, assuming that relative negative signs are possible we give in the last column the Br for $\mu=-2.2.$ Clearly, a branching ratio in the $10^{-7}-10^{-6}$ range will not allow to pinpoint accurate values for the two couplings. Nevertheless if the branching ratio turns out to be in the $10^{-5}$ range, it can only be caused by large values of $g,$ say $g>0.6.$ We wish also to mention an additional scenario: the $g$ coupling will probably be measured directly in $D^*$ decays, or indirectly from ${\rm Br}(\mbox{${D^{*}}^{0}\to \, D^0\, \gamma\,\gamma\,$})$ or other methods. With this knowledge, ${\rm Br}(\mbox{${B^{*}}^{0} \to \, B^0\, \gamma\,\gamma\,$})$ becomes a function of $g_{{B^*}^0B^0\gamma}$ only and it could provide the desirable measurement of this coupling. This is a very interesting issue, since as pointed out already some time ago \cite{Miller}, there is no other possibility of measuring the width of the \mbox{$B^{*} \to \, B\, \gamma\,$} decay with presently known techniques. In Figs.(9),(10) and (11) we give the differential distribution of $d\Gamma(\mbox{$B^{*} \to \, B\, \gamma\,\gamma\,$})/ds$ for $g=0.5$ and three different values of $\mu.$ Clearly, once $g$ is known one may use accurate differential distributions to distinguish between different $\mu$ values. At this point we also wish to make some remarks on the similar decays of the strange heavy vector mesons, ${B_s^{*}}^0 \to \, B^0_s\, \gamma\,\gamma$ and ${D_s{*}}^+ \to \, D^+_s\, \gamma\,\gamma$ which were not mentioned so far. In both these cases the pion anomaly is further suppressed, since ${B_s^{*}}^0 \to \, B^0_s\, \pi^0, {D_s^{*}}^+ \to \, D^+_s\, \pi^0 $ can proceed only by isospin violation, e.g. via $\eta-\pi^0$ mixing. On the other hand, both decays can proceed via chiral loops with charged $K$-mesons in the loops, ${B^{*}}^0_s\to(K^+ {B^*}^-) \to \, B^0_s\, \gamma\,\gamma$ and ${D^{*}}^+_s \to(K^+ {D^*}^0)\to \, D^+_s\, \gamma\,\gamma.$ However, one must add that for ${D^{*}}^+_s \to \, D^+_s\, \gamma\,\gamma$ there is the complication of the bremstrahlung and we shall disregard it here. We calculated therefore only the ${B^*}^0_s \to B^0_s \gamma \gamma$ decay, and the situation is quite similar to that encountered in ${B^*}^0$ decay; therefore we do not repeat this analysis here. Before concluding we comment on a few points which were neglected in our treatment. \begin{itemize} \item[1] We calculated also the contribution to the anomaly term of a virtual $\eta$ exchange for the \mbox{$D^{*} \to \, D\, \gamma\,\gamma\,$} decay. The inclusion of $\eta$ modifies our result in eq.(\ref{gdsgg}) by a factor of $\left(1+\frac{g_{D^* D\eta}}{10 g_{D^*D\pi}}\right).$ Since $g_{D^* D\eta}$ and $ g_{D^*D\pi}$ are expected to be comparable, this is a small effect. \item[2] We neglected the off-shell $q^2$-dependence of the anomaly which could have some effect, especially in the $D^*$-decay. This should be included in a more detailed treatment. \end{itemize} To summarize, we have used the Heavy Meson Chiral Lagrangian to present the first treatment of the rare \mbox{${B^{*}}^{0} \to \, B^0\, \gamma\,\gamma\,$}, \mbox{${D^{*}}^{0}\to \, D^0\, \gamma\,\gamma\,$} decays. The decay rates depend on the strong $g_{B^{*}B\pi\,},g_{D^{*}D\pi\,}$ couplings and on the strength of the magnetic dipole transitions $g_{B^*B\gamma},g_{D^*D\gamma}.$ The strong couplings are expressed in the chiral Lagrangian by the strong axial coupling $g.$ We have shown that Br(\mbox{${D^{*}}^{0}\to \, D^0\, \gamma\,\gamma\,$}) can be given as a function of $g$ only, and as such, it would provide an appropriate tool for its measurement. For the conventionally envisaged range $0.25<g<1$ we calculated $1.6\times 10^{-6}<{\rm Br}(\mbox{${D^{*}}^{0}\to \, D^0\, \gamma\,\gamma\,$})<3.3\times 10^{-5}.$ On the other hand, Br(\mbox{${B^{*}}^{0} \to \, B^0\, \gamma\,\gamma\,$}) is a function of $g_{B^{*}B\pi\,},g_{{B^*}^0B^0\gamma}$ and $g_{{B^*}^+B^+\gamma}.$ The latter coupling has little effect on the branching ratio. Nevertheless, one cannot determine specific values for the first two couplings from the measured branching ratio, unless $g$ is in the higher part of its expected range, say $0.6-1.$ The differential $d\Gamma/ds$ distributions in both cases can be used as additional help for extracting the values of the coupling constants. If the value of $g_{B^{*}B\pi\,}$ will turn out to be in the ``measurable" range, it will be of great interest to check the $HM\chi L$ relation $g_{B^{*}B\pi\,}/g_{D^{*}D\pi\,}=M_B/M_D.$ \section{Acknowledgement} We are in debt to Professor Gad Eilam for helpful remarks and stimulating discussions. We also acknowledge discussions with Drs. Simon Robins, Yoram Rozen and Shlomit Terem on the feasibility of the relevant detection experiments. Special thanks are extended to the referee, who pointed out to us a serious omission in the original version. The research of P.S. has been supported in part by the Fund for Promotion of Research at the Technion.
\section{Introduction} The exact gauge {\it low-energy effective action} (LEEA) in the Coulomb branch of N=2 supersymmetric Yang-Mills theory in {\it four} spacetime dimensions (4d), which includes both perturbative (one-loop) and nonperturbative (instanton) quantum corrections, was determined by Seiberg and Witten \cite{sw}. The main tools of their construction were the general constraints implied by N=2 extended supersymmetry, the known anomaly structure, and electric-magnetic duality. The N=2 off-shell supersymmetry implies the unique `Ansatz' for the N=2 (abelian) vector multiplet LEEA, in terms of a holomorphic function ${\cal F}(W)$ of the N=2 (restricted chiral) superfield strength $W$. The chiral anomaly determines the perturbative (logarithmic) contribution to the function ${\cal F}''(W)$. Nonperturbative consistency and duality unambiguously fix the instanton corrections to ${\cal F}(W)$, which are related to the BPS monopoles representing nonperturbative degrees of freedom and belonging to hypermultiplets. The exact Seiberg-Witten ($SU(2)$-based) solution can be encoded in terms of an elliptic curve $\Sigma_{\rm SW}$, by integrating certain abelian differential $\lambda_{\rm SW}$ over the torus periods. Since a generic 4d, N=2 gauge field theory has both N=2 vector multiplets {\it and} hypermultiplets, the latter may also have their own N=2 supersymmetric LEEA \cite{rev2}. The hypermultiplet LEEA is highly constrained by N=2 extended supersymmetry and its automorphisms too, so that its exact form can also be determined. For instance, in three spacetime dimensions, Seiberg and Witten \cite{sw3} used the so-called {\it c-map} \cite{cfg}, relating the special K\"ahler geometry of the N=2 vector multiplet moduli space to the hyper-K\"ahler geometry of the hypermultiplet moduli space. They further argued that the {\it Atiyah-Hitchin} (AH) metric \cite{ati} is the only regular exact solution.~\footnote{An abelian gauge vector is dual to a scalar in three dimensions.} This proposal was later confirmed by 3d instanton calculations \cite{dkmtv}, which also discovered a one-parameter family of possible hyper-K\"ahler metrics generalizing the AH metric. We propose the most general `Ansatz' for the 4d hypermultiplet LEEA, which is compatible with all unbroken symmetries and has two parameters. It allows us to reformulate the solution in the very transparent geometrical way. The earlier approaches \cite{sw3,dkmtv} are formulated only in 3d on the gauge field theory side, and they are not manifestly supersymmetric, which may cast some doubt on their ultimate consistency. Our main purpose in this Letter is a derivation of the exact hypermultiplet LEEA directly in {\it four} spacetime dimensions, in the manifestly N=2 supersymmetric way. We confine ourselves to a {\it single} hypermultiplet for simplicity. We fully exploit the restrictions implied by {\it off-shell} N=2 supersymmetry and its $SU(2)_R$ automorphisms, by making both of them manifest in HSS. Converting the HSS action into N=2 {\it projective superspace} (PSS) allows us to calculate the effective hyper-K\"ahler metric. The solution can be put into the Seiberg-Witten form after introducing the auxiliary elliptic curve associated with an $O(4)$ projective multiplet in N=2 PSS. \section{N=2 supersymmetry and hyper-K\"ahler metrics} The scalar kinetic part of the hypermultiplet LEEA is of the second order in spacetime derivatives, so that it has the form of a {\it non-linear sigma-model} (NLSM). By N=2 supersymmetry in 4d, the metric of this 4d NLSM has to be {\it hyper-K\"ahler} \cite{hklr}. Making N=2 supersymmetry manifest (i.e off-shell) also makes manifest the hyper-K\"ahler nature of the hypermultiplet LEEA. In the {\it harmonic superspace} (HSS) approach \cite{gikos}, both an N=2 vector multiplet and a hypermultiplet can be introduced off-shell on equal footing. For example, the Fayet-Sohnius hypermultiplet is described by an unconstrained complex analytic superfield $q^+$ of $U(1)$ charge $(+1)$, whereas an N=2 vector multiplet is described by an unconstrained analytic superfield $V^{++}$ of $U(1)$ charge $(+2)$. The general N=2 NLSM Lagrangian in HSS reads \cite{gios} $$ {\cal L}^{(+4)}= -\sbar{q}{}^+D^{++}q^+ - {\cal K}^{(+4)}(\sbar{q}{}^{+},q^+;u^{\pm})~,\eqno(2.1)$$ where ${\cal K}^{+(4)}$ is called a hyper-K\"ahler potential, while the harmonic covariant derivative $D^{++}$ includes central charges. The N=2 central charge $Z$ can be treated as the abelian N=2 vector superfield background whose N=2 gauge superfield strength is given by $\VEV{W}=Z$ \cite{ikz}. The role of the analytic function ${\cal K}^{+(4)}$ in the hypermultiplet LEEA (2.1) is similar to the role of the holomorphic Seiberg-Witten potential ${\cal F}$ in the N=2 gauge LEEA. Because of manifest N=2 supersymmetry by construction, the equations of motion for the HSS action (2.1) determine (at least, in principle) the component hyper-K\"ahler NLSM metric in terms of a single analytic function ${\cal K}^{(+4)}$. An explicit form of this relation is, however, not known in general. A crucial simplification arises when the $SU(2)_R$ automorphisms of N=2 supersymmetry are also preserved, which implies that the hyper-K\"ahler potential is independent upon harmonics. Since $SU(2)_R$ is known to be non-anomalous \cite{sw,sw3}, the most general `Ansatz' for the hypermultiplet LEEA takes the form of a real quartic polynomial, $$ {\cal K}^{(+4)}= \fracm{\lambda}{2}(\sbar{q}{}^+)^2(q^+)^2+\left[ \gamma \sbar{(q^+)}{}^4 + \beta\sbar{(q^+)}{}^3 q^+ +{\rm h.c.}\right]~,\eqno(2.2)$$ with one real $(\lambda)$ and two complex $(\beta,\gamma)$ parameters. The $Sp(1)_{\rm PG}$ internal symmetry of the free hypermultipet action leaves the form of the quartic (2.2) invariant but not the coefficients. Hence, $Sp(1)_{\rm PG}$ can be used to reduce the number of coupling constants in the family of hyper-K\"ahler metrics associated with the hyper-K\"ahler potential (2.2) from five to two. Equations (2.1) and (2.2) also imply the conservation law \cite{gios} $$ D^{++}{\cal K}^{(+4)}=0 \eqno(2.3)$$ on the equations of motion, $D^{++}\sbar{q}{}^+ =\pa{\cal K}^{(+4)}/\pa q^+$ and $D^{++}q^+ =-\pa{\cal K}^{(+4)}/\pa\sbar{q}{}^+$. \section{Perturbative hypermultiplet LEEA} The manifestly N=2 supersymmetric HSS description of the hypermultiplet LEEA allows us to exploit the constraints imposed by unbroken N=2 supersymmetry and its automorphism symmetry in the very efficient and transparent way. For example, as regards a perturbation theory in 4d, N=2 supersymmetric QED (or in the Coulomb branch of N=2 supersymmetric $SU(2)$ Yang-Mills theory \cite{sw}), the unbroken symmetry is given by $$ SU(2)_{R,~{\rm global}}\times U(1)_{\rm local}~.\eqno(3.1)$$ The {\it unique} hypermultiplet self-interaction consistent with N=2 supersymmetry and the internal symmetry (3.1) in HSS is described by the hyper-K\"ahler potential $$ {\cal K}^{(+4)}_{\rm TN}=\fracmm{\lambda}{2}\left(\sbar{q}{}^+q^+\right)^2~, \eqno(3.2)$$ just because it is the only function of $U(1)$ charge $(+4)$ that is independent upon harmonics and invariant under $U(1)_{\rm local}\,$. \begin{figure} \vglue.1in \makebox{ \epsfxsize=4in \epsfbox{box.eps} } \end{figure} The coupling constant $\lambda$ in eq.~(3.2) (in the one-loop approximation) is determined by the HSS graph shown in Fig.~1. The analytic propagator (wave lines in Fig.~1) of the N=2 vector superfield (in N=2 Feynman gauge) is (see \cite{ikz} for details) $$ i\VEV{ V^{++}(1)V^{++}(2)}=\fracmm{1}{\Box_1}(D_1^+)^4\delta^{12}({\cal Z}_1-{\cal Z}_2) \delta^{(-2,2)}(u_1,u_2)~.\eqno(3.3)$$ The hypermultiplet analytic propagator (solid lines in Fig.~1) with non-vanishing central charges (in the pseudo-real notation) reads \cite{ikz} $$ i\VEV{q^+(1)q^+(2)}=\fracmm{-1}{\Box_1+m^2} \fracmm{(D^+_1)^4(D^+_2)^4}{(u^+_1u_2^+)^3}e^{\tau_3[v(2)-v(1)]}\delta^{12} ({\cal Z}_1-{\cal Z}_2)~,\eqno(3.4)$$ where $m^2=\abs{Z}^2$ is the hypermultiplet (bare) BPS mass and $$ iv =-Z(\bar{\theta}^+\bar{\theta}^-)-\bar{Z}(\theta^+\theta^-)~.\eqno(3.5)$$ In the one-loop approximation one finds \cite{ikz} the predicted form (3.2) with $$ \lambda=\fracmm{g^4}{\pi^2}\left[ \fracmm{1}{m^2}\ln\left( 1+\fracmm{m^2}{\Lambda^2} \right) -\fracmm{1}{\Lambda^2+m^2}\right] \eqno(3.6)$$ in terms of the gauge coupling constant $g$, the BPS mass $m^2$ and the IR-cutoff $\Lambda$. Note that $\lambda\neq 0$ only when $Z\neq 0$. The dependence of $\lambda$ upon the IR-cutoff is expected to disappear after summing up all contributions from higher loops. To understand the hyper-K\"ahler geometry associated with the hyper-K\"ahler potential (3.2), it is convenient to rewrite the HSS action into PSS, by going partially on-shell, in terms of an N=2 {\it tensor} multiplet. Unlike the Fayet-Sohnius hypermultiplet, the N=2 tensor superfield $L^{(ij)}$ has a finite number of auxiliary fields. In the standard N=2 superspace $({\cal Z})$ the $L^{(ij)}$ is defined by the off-shell constraints $$ D\low{\alpha}{}^{(i}L^{ik)}=\bar{D}_{\dt{\alpha}}{}^{(i}L^{jk)}=0~,\eqno(3.7)$$ and the reality condition $$ \Bar{L^{ij}}=\varepsilon_{ik}\varepsilon_{jl}L^{kl}~.\eqno(3.8)$$ It is not difficult to verify that eq.~(3.7) implies $$ \de_{\alpha}G\equiv (D^1_{\alpha}+\xi D^2_{\alpha})G=0~,\qquad \Delta_{\dt{\alpha}}G\equiv (\bar{D}^1_{\dt{\alpha}}+\xi\bar{D}^2_{\dt{\alpha}})G=0~,\eqno(3.9)$$ for {\it any} function $G(Q^2_2(\xi),\xi)$ depending upon $$ Q^2_2(\xi)\equiv \xi_i\xi_j L^{ij}(Z)~,\quad \xi_i\equiv (1,\xi)~, \eqno(3.10)$$ where the $CP(1)$ inhomogeneous coordinate $\xi$ has been introduced. It follows from eq.~(3.9) that one can construct N=2 invariant actions by integrating the potential $G(Q^2_2(\xi),\xi)$ over the rest of the N=2 superspace coordinates \cite{hklr}, $$ S = \int d^4x\fracmm{1}{2\pi i}\oint_C \fracmm{d\xi}{(1+\xi^2)^4}\tilde{\de}^2 \tilde{\Delta}^2 G(Q^2_2(\xi),\xi) +{\rm h.c.}~,\eqno(3.11)$$ where the new Grassmann superspace derivatives, $$ \tilde{\de}_{\alpha}=\xi D^1_{\alpha}-D^2_{\alpha}~,\quad \tilde{\Delta}_{\dt{\alpha}}=\xi \bar{D}^1_{\dt{\alpha}}-\bar{D}^2_{\dt{\alpha}}~,\eqno(3.12)$$ `orthogonal' to those of eq.~(3.9), have been introduced. After being reduced to 4d, N=1 superspace, eqs.~(3.10) and (3.11) take the form $$ \left.Q^2_2(\xi)\right| = \Phi + \xi H -\xi^2 \bar{\Phi}~,\eqno(3.13)$$ and $$ S = \int d^4x d^4\theta\fracmm{1}{2\pi i}\oint_C \fracmm{d\xi}{\xi^2} G(\left.Q^2_2(\xi)\right|,\xi)+{\rm h.c.}~,\eqno(3.14)$$ in terms of the N=1 chiral superfield $\Phi$ and the N=1 real linear superfield $H$. The N=2 tensor multipet constraints (3.7) and (3.8) read in HSS as $$ D^{++}L^{++}=0 \quad {\rm and} \quad \sbar{L}{}^{++}=L^{++}~,\eqno(3.15)$$ respectively, where $L^{++}=u^+_iu^+_jL^{ij}({\cal Z})$. Let's substitute (we temporarily set $\lambda=1$) $$ {\cal K}^{(+4)}_{\rm TN}= \fracm{1}{2}(\sbar{q}{}^+q^+)^2=-2(L^{++})^2~,\quad {\rm or,~ equivalently,}\quad \sbar{q}{}^+q^+=2iL^{++}~,\eqno(3.16)$$ which is certainly allowed because of eq.~(2.3). The constraints (3.7) can be incorporated off-shell by using extra real analytic superfield $\omega$ as the Lagrange multiplier. Changing the variables from $(\sbar{q}{}^+,q^+)$ to $(L^{++},\omega)$ amounts to an N=2 duality transformation in HSS. The explicit solution to eq.~(3.16) reads $$ q^+=-i\left(2u^+_1+if^{++}u^-_1\right)e^{-i\omega/2}~, \quad \sbar{q}{}^+=i\left( 2u^+_2-if^{++}u^-_2\right)e^{i\omega/2}~,\eqno(3.17)$$ where the function $f^{++}$ is given by \cite{gio} $$ f^{++}(L,u) = \fracmm{2(L^{++}-2iu_1^+u_2^+)}{1+\sqrt{1-4u^+_1u^+_2 u^-_1u^-_2 -2iL^{++}u_1^-u_2^-}}~.\eqno(3.18)$$ It is straightforward to rewrite the free (massless) HSS action (2.7) in terms of the new variables. It results in the {\it improved} (i.e. N=2 superconformally invariant) N=2 tensor multiplet action \cite{gio} $$S\low{\rm impr.} = {\fracmm12} \int d\zeta^{(-4)}du (f^{++})^2~.\eqno(3.19)$$ The action dual to the NLSM action defined by eqs.~(2.1) and (3.2) is thus given by a sum of the non-improved (quadratic) and improved (non-polynomial) HSS actions for the N=2 tensor multiplet \cite{hklr,gio}, $$ S_{\rm TN}[L;\omega]= S\low{\rm impr.} + {\fracmm12} \int d\zeta^{(-4)}du \left[(L^{++})^2 +\omega D^{++}L^{++}\right]~.\eqno(3.20)$$ The equivalent PSS action is given by eq.~(3.11) with $$ \oint G = M\oint_{C_0} \fracmm{(Q^2_2)^2}{2\xi} + \oint_{C_r} Q^2_2(\ln Q^2_2-1)~,\eqno(3.21)$$ where we have restored the dependence upon $\lambda$ by setting $M={\fracmm12}\lambda^{-1/2}$. The contour $C_0$ in complex $\xi$-plane goes around the origin, whereas the contour $C_r$ encircles the roots of the quadratic equation \cite{hklr} $$ Q^2_2(\xi)=0~. \eqno(3.22)$$ The hyper-K\"ahler metric of the N=2 NLSM defined by eqs.~(3.21) and (3.22) is equivalent to the Taub-NUT metric with the mass parameter $M={\fracmm12}\lambda^{-1/2}$ \cite{hklr,gios}. The $SU(2)_R$ transformations act in PSS in the form of {\it projective} (fractional) transformations \cite{oldket} $$ \xi'=\fracmm{\bar{a}\xi-\bar{b}}{a+b\xi}~,\qquad \abs{a}^2+\abs{b}^2=1~,\eqno(3.23)$$ while a generic PSS action (3.11) is not invariant under these transformations. Nevertheless, eq.~(3.11) with some non-trivial contour $C_r$ is going to be invariant under the transformations (3.23) provided that $$ G({Q_2^2}'(\xi'),\xi')=\fracmm{1}{(a+b\xi)^2}G(Q_2^2(\xi),\xi)~,\quad {\rm where}\quad {Q_2^2}'(\xi')=\fracmm{1}{(a+b\xi)^2}Q_2^2(\xi)~.\eqno(3.24)$$ Eq.~(3.24) implies that the invariant PSS potential $G(Q^2_2)$ should be `almost' linear in $Q^2_2$, like in the second term of the action (3.21). The transition $u_i\to\xi_i=(1,\xi)$ describes the holomorphic projection of HSS on PSS, where the analytic superfield $L^{++}(\zeta,u)$ is replaced by the holomorphic (with respect to $\xi$) section $Q^2_2(L,\xi)$ of the line bundle $O(2)$ whose fiber is parametrized by constrained superfields. The equation (3.13) defines the Riemann sphere in ${\bf C}^2$ parametrized by $(Q_2,\xi)$. \section{Exact hypermultiplet LEEA and $O(4)$ bundle} In an abelian quantum field theory there are no instantons, so that the one-loop results of sect.~3 are, in fact, {\it exact} in that case. If, however, the underlying N=2 gauge field theory has a non-abelian gauge group whose rank is more than one, one expects nonperturbative contributions to the LEEA of a single (magnetically charged) hypermultiplet from instantons and anti-instantons \cite{hwit}. It may happen, e.g., in the Higgs branch where the gauge symmetry is completely broken. Given the most general $SU(2)_R$-invariant hyper-K\"ahler potential (2.2), let's make a substitution $$ {\cal K}^{(+4)}(q,\sbar{q})\equiv\fracm{\lambda}{2}(\sbar{q}{}^+)^2(q^+)^2+\left[ \gamma\sbar{(q^+)}{}^4 + \beta\sbar{(q^+)}{}^3q^+ +{\rm h.c.}\right] =L^{++++}(\zeta,u)~,\eqno(4.1)$$ where the real analytic superfield $L^{++++}$ satisfies the conservation law (2.3), $$ D^{++}L^{++++}=0~.\eqno(4.2)$$ Eq.~(4.2) can be recognized as the {\it off-shell} N=2 (standard) superspace constraints $$ D\low{\alpha}{}^{(i}L^{jklm)}=\bar{D}_{\dt{\alpha}}{}^{(i}L^{jklm)}=0~,\eqno(4.3)$$ where $L^{++++}=u^+_iu^+_ju^+_ku^+_lL^{ijkl}({\cal Z})$, while eq.~(4.1) implies the reality condition $$ \Bar{L^{ijkl}}=\varepsilon_{im}\varepsilon_{jn}\varepsilon_{kp}\varepsilon_{lq}L^{mnpq}~,\eqno(4.4)$$ defining together the $O(4)$ projective multiplet \cite{oldket}. Unlike the $O(2)$ tensor supermultiplet (sect.~3), the $O(4)$ supermultiplet does not have a conserved vector (or a gauge antisymmetric tensor) amongst its field components. The N=2 invariant PSS action construction (3.11) in terms of a PSS potential $G(Q^2_{4}(\xi),\xi)$ equally applies to the projective $O(4)$ supermultiplets,~\footnote{In fact, it also applies to any projective $O(2k)$ multiplets satisfying the off-shell N=2 superspace \newline ${~~~~~}$ constraints generalizing those of eqs.~(4.3) and (4.4) with $k>2$ \cite{oldket}.} while $L^{ijkl}$ should enter the action via the argument \cite{oldket} $$ Q^2_{4}(\xi)=\xi_{i}\xi_{j}\xi_{k}\xi_lL^{ijkl}({\cal Z})~,\quad \xi_i=(1,\xi)~. \eqno(4.5)$$ The N=1 superspace projections of the N=2 superfield (4.5) and the N=2 invariant PSS action are given by $$ \left.Q^2_4(\xi)\right|=\Phi + \xi H +\xi^2V-\xi^3\bar{H} +\xi^4\bar{\Phi}~, \eqno(4.6)$$ and $$ S = \int d^4x d^4\theta\fracmm{1}{2\pi i}\oint_C \fracmm{d\xi}{\xi^2} G (\left.Q^2_4(\xi)\right|,\xi)+{\rm h.c.}~,\eqno(4.7)$$ respectively, in terms of the N=1 chiral superfield $\Phi$, the N=1 complex linear superfield $H$, and the N=1 general (unconstrained) real superfield $V$ \cite{hklr,oldket}. The N=1 superfield $V$ enters the action (4.7) as the Lagrange multiplier, whose elimination via its `equation of motion' implies the algebraic constraint \cite{oldket} $$ {\rm Re}\,\oint \fracmm{\pa G}{\pa Q^2_4}=0~.\eqno(4.8)$$ Eq.~(4.8) reduces the number of independent N=2 NLSM physical real scalars from five to four, which is consistent with the well-known fact that the real dimension of any hyper-K\"ahler manifold is a multiple of four \cite{hklr}. After solving the constraint (4.8), the complex linear N=1 superfield $H$ can be traded for yet another N=1 chiral superfield $\Psi$, by the use of the N=1 superfield Legendre transform that results in the N=1 superspace K\"ahler potential $K(\Phi,\Psi,\bar{\Phi},\bar{\Psi})$ associated with the N=2 supersymmetric NLSM of eq.~(4.7) \cite{hklr}. The most straightforward procedure of calculating the dependence $q(L)$, as well as performing an explicit N=2 transformation of the unconstrained HSS action into the PSS action in terms of the constrained N=2 superfield defined by eq.~(4.1), use roots of the quartic polynomial. Remarkably, the N=2 PSS action in question can be fixed without calculating the roots in the manifestly N=2 supersymmetric approach. It is the $SU(2)_R$ invariance that is powerful enough to fix the PSS action equivalent to the HSS action of eqs.~(2.1) and (2.2) ({\it cf.} ref.~\cite{gio}). The one real and two complex constants, $(\lambda,\beta,\gamma)$, respectively, parametrizing the hyper-K\"ahler potential (2.2), are naturally united into the $SU(2)$ 5-plet $c^{ijkl}$ subject to the reality condition (4.4). After extracting a constant piece out of $q^+$, say, $q^+_a=u^+_a + \tilde{q}^+_a$ and $u_a=(1,\xi)$, and collecting all constant pieces on the left-hand-side of eq.~(4.1), we can identify their sum with a constant piece $c^{++++}=c^{ijkl}u^+_iu^+_ju^+_ku^+_l$ of $L^{++++}$ on the right-hand-side of eq.~(4.1), representing the constant vacuum expectation values of the N=1 superfield components of $L^{++++}$ defined by eq.~(4.6), i.e. $$ \lambda=\VEV{V}~,\quad \beta=\VEV{H}~,\quad\gamma=\VEV{\Phi}~.\eqno(4.9)$$ The $SU(2)_R$ transformations in PSS are the projective transformations (3.24), so that the PSS potential $G$ of the `improved' $O(4)$ multiplet action having the form (3.11) must be proportional to $Q_4\equiv\sqrt{Q^2_4}$ because of the relations $$G({Q_4^2}'(\xi'),\xi')=\fracmm{1}{(a+b\xi)^2}G(Q_4^2(\xi),\xi) \quad{\rm and}\quad {Q^2_4}'(\xi')=\fracmm{1}{(a+b\xi)^4}Q^2_4(\xi)~.\eqno(4.10)$$ The most general non-trivial contour $C_r$ in complex $\xi$-plane, whose definition is compatible with the projective $SU(2)$ symmetry, is the one encircling the roots of the quartic ({\it cf.} sect.~3), $$\left.Q^2_4(\xi)\right|=p + \xi q + \xi^2 r -\xi^3\bar{q}+\xi^4\bar{p}~, \eqno(4.11)$$ with one real $(r)$ and two complex $(p,q)$ additional parameters belonging to yet another 5-plet of $SU(2)$. The projective $SU(2)$ invariance of the PSS action defined by eqs.~(4.7) and (4.11) can be used to reduce the number of independent parameters in the corresponding family of hyper-K\"ahler metrics from five to two, which is consistent with the HSS predictions of sect.~2.~\footnote{The generalization of eq.~(3.22) similarly to eq.~(4.11) is `empty' since the quadratic polynomial \newline ${~~~~~}$ $c^2_2(\xi)=p+\xi r-\xi^2\bar{p}$ can always be removed by an $SU(2)$ transformation.} We didn't attempt to establish an explicit relation between the HSS coefficients $(\lambda,\gamma,\beta)$ and the PSS coefficients $(r,q,p)$. The most natural (non-trivial) contour $C_r$ surrounds the roots of the equation $$ Q^2_4(\xi)=0~, \eqno(4.12)$$ and it leads to the only non-singular hyper-K\"ahler NLSM metric (sect.~5). The $SU(2)$-invariant PSS action, equivalent to the one defined by eqs.~(2.1) and (4.1), is therefore given by $$ \fracmm{1}{2\pi i}\oint G =-\fracmm{1}{2\pi i}\oint_{C_0} \fracmm{Q^2_4}{\xi} + \oint_{C_r} Q_4~.\eqno(4.13)$$ The constraint (4.8) in the case (4.12) takes the form $$ \oint_{C_r} \fracmm{d\xi}{\sqrt{Q^2_4}}=1~.\eqno(4.14)$$ The component form of the metric associated with eqs.~(4.13) and (4.14) was found in ref.~\cite{iro}. Because of the reality condition (4.4), the quartic (4.12) has two pairs of roots $(\rho,-1/\bar{\rho})$ related by an $SL(2,{\bf Z})$ transformation and satisfying the defining relation $$ Q^2_4(\xi)=c(\xi-\rho_1)(\bar{\rho}_1\xi+1)(\xi-\rho_2)(\bar{\rho}_2\xi+1)~.\eqno(4.15)$$ The branch cuts of the root in eq.~(4.14) can be chosen to run from $\rho_1$ to $-1/\bar{\rho}_2$ and from $\rho_2$ to $-1/\bar{\rho}_1$. The contour integration in eq.~(4.14) can thus be reduced to the complete elliptic integral (in the Legendre normal form) over the branch cut \cite{iro}, $$ \fracmm{4}{\sqrt{c(1+\abs{\rho_1}^2)(1+\abs{\rho_2}^2)}} \int^1_0\fracmm{d\xi}{\sqrt{(1-\xi^2)(1-k^2\xi^2)}}=1~,\eqno(4.16)$$ with the modulus $$ k^2= \fracmm{(1+\rho_1\bar{\rho}_2)(1+\rho_2\bar{\rho}_1)}{(1+\abs{\rho_1}^2) (1+\abs{\rho_2}^2)}~.\eqno(4.17)$$ The constraint (4.16) can be explicitly solved in terms of the complete elliptic integrals, $$ K(k)=\int^{\pi/2}_0 \,\fracmm{d\gamma}{\sqrt{1-k^2\sin^2\gamma}} \quad {\rm and} \quad E(k) =\int^{\pi/2}_0 d\gamma\,\sqrt{1-k^2\sin^2\gamma}~~,\eqno(4.18)$$ of the first and second kind, respectively, by using the following parametrization \cite{iro}: $$ \eqalign{ \Phi ~=~& 2e^{2i\varphi} \left[ \cos(2\psi) (1+\cos^2\vartheta)\right. \cr & \left.+ 2i\sin(2\psi)\cos\vartheta + (2k^2-1)\sin^2\vartheta \right] K^2(k)~,\cr H ~=~& 8e^{i\varphi}\sin\vartheta \left[ \sin(2\psi) \right. \cr & \left.- i\cos(2\psi)\cos\vartheta + i(2k^2-1)\cos\vartheta \right] K^2(k)~,\cr V ~=~& 4\left[ -3\cos(2\psi)\sin^2\vartheta +(2k^2-1)(1-3\cos^2\vartheta)\right] K^2(k)~,\cr}\eqno(4.19)$$ in terms of the Euler `angles' $(\vartheta,\psi,\varphi)$ and the modulus $k$ representing the independent (superfield) coordinates of the N=2 NLSM under consideration. Applying the generalized Legendre transform \cite{iro} to the function (4.13) with respect to $H$ gives rise to the {\it Atiyah-Hitchin} (AH) metric \cite{ati} $$ ds^2_{\rm AH}=\fracmm{1}{4}A^2B^2C^2\left(\fracmm{dk}{k{k'}^2K^2}\right)^2 + A^2(k)\sigma_1^2 + B^2(k)\sigma^2_2 + C^2(k)\sigma^2_3~,\eqno(4.20)$$ whose coefficient functions satisfy the relations \cite{ati} $$ \eqalign{ AB ~=~ & -K(k)\left[E(k)-K(k)\right]~,\cr BC~=~ & -K(k)\left[E(k)-{k'}^2K(k)\right]~,\cr AC~=~ & -K(k)E(k)~.\cr}\eqno(4.21)$$ while $\sigma_i$ stand for the $SO(3)$-invariant one-forms $$ \eqalign{ \sigma_1 ~=~ & +\frac{1}{2}\left(\sin\psi d\vartheta - \sin\vartheta \cos\psi d \varphi\right)~,\cr \sigma_2 ~=~ & -\frac{1}{2}\left(\cos\psi d\vartheta + \sin\vartheta \sin\psi d \varphi\right)~,\cr \sigma_3 ~=~ & +\frac{1}{2}\left(d\psi + \cos\vartheta d\varphi\right)~,\cr}\eqno(4.22)$$ and $k'$ is known as the complementary modulus, ${k'}^2=1-k^2$. In the limit $k\to 1$ (or, equivalently, $k'\to 0$), one has an asymptotic expansion $$K(k)\approx -\log\, k'\left[ 1 +\fracmm{(k')^2}{4}\right] +\ldots \eqno(4.23)$$ Eq.~(4.23) suggests us to make a redefinition $$ k'=\sqrt{1-k^2}\approx 4\exp\left(\fracmm{1}{\gamma}\right)~,\eqno(4.24)$$ and describe the same limit at $\gamma\to 0^-$. After substituting eq.~(4.23) into eq.~(4.21) one finds that the AH metric becomes {\it exponentially close} to the Taub-NUT metric in the form (4.20) subject to the additional relations: $$ A^2\approx B^2\approx \fracmm{1+\gamma}{\gamma^2}~~,\quad C^2\approx \fracmm{1}{1+\gamma}~~.\eqno(4.25)$$ The extra $U(1)$ symmetry of the Taub-NUT metric is the direct consequence of the relation $A^2=B^2$ arising from the AH metric in the asymptotic limit described by eq.~(4.25). The vicinity of $k'\approx 0^+$ describes the region of the hypermultiplet moduli space where quantum perturbation theory applies, with the exponentially small AH corrections to the Taub-NUT metric being interpreted as the one-instanton and anti-instanton contributions to the hypermultiplet LEEA \cite{sw3,dkmtv}. From the N=2 PSS viewpoint, the transition from the perturbative hypermultiplet LEEA to the nonperturbative one corresponds to the transition from the $O(2)$ holomorphic line bundle associated with the standard N=2 tensor supermultiplet to the $O(4)$ holomorphic line bundle associated with the $O(4)$ N=2 supermultiplet. The two holomorphic bundles are topologically different: with respect to the standard covering of $CP(1)$ by two open affine sets, the $O(2)$ bundle has transition functions $\xi^{-1}$, whereas the $O(4)$ bundle has transition functions $\xi^{-2}$. The variable $Q$ is the coordinate of the corresponding fiber over $CP(1)$. \section{Atiyah-Hitchin metric and elliptic curve} The quadratic dependence of $Q^2_2$ on $\xi$ in eqs.~(3.10) and (3.13) allows us to globally interpret it as a holomorphic (of degree 2) section of PSS, fibered by the superfields $(\Phi,H)$ and topologically equivalent to a complex line (or Riemann sphere of genus $0$). Similarly, the quartic dependence of $Q^2_4$ on $\xi$ in eqs.~(4.5) and (4.6) allows us to globally interpret it as a holomorphic (of degree 4) section of PSS, fibered by the superfields $(\Phi,H,V)$ and topologically equivalent to an {\it elliptic curve} $\Sigma_{\rm hyper.}$ (or a torus of genus $1$). The non-perturbative hypermultiplet LEEA can therefore be encoded in terms of the genus-one Riemann surface $\Sigma_{\rm hyper.}$ in close analogy to the exact N=2 gauge LEEA in terms of the elliptic curve $\Sigma_{\rm SW}$ of Seiberg and Witten \cite{sw}. The classical twistor construction of hyper-K\"ahler metrics \cite{ati} is known to be closely related to the {\it Hurtubise} elliptic curve $\Sigma_{\rm H}$ \cite{hur}. This curve can be identified with $\Sigma_{\rm hyper.}$ that carries the same information and whose defining eq.~(4.6) can be put into the Hurtubise form, $$ \tilde{Q}^2_4(\tilde{\xi})= K^2(k)\tilde{\xi}\left[ kk'(\tilde{\xi}^2-1)+(k^2-{k'}^2)\tilde{\xi} \right]~,\eqno(5.1)$$ by a projective $SU(2)$ transformation. In its turn, eq.~(5.1) is simply related (by a linear transformation) to another standard (Weierstrass) form, $y^2=4x^3-g_2x-g_3$. Therefore, in accordance with ref.~\cite{ati}, the real period $\omega$ of $\Sigma_{\rm H}$ is $$\omega\equiv 4k_1~,\quad {\rm where}\quad 4k_1^2=kk'K^2(k)~,\eqno(5.2)$$ whereas the complex period matrix of $\Sigma_{\rm H}$ is given by $$\tau=\fracmm{iK(k')}{K(k)}~.\eqno(5.3)$$ At generic values of the AH modulus $k$, $0<k<1$, the roots of the Weierstrass form are all different from each other, while they all lie on the real axis, say, at $e_3<e_2<e_1<\infty=(e_4)$. Accordingly, the branch cuts are running from $e_3$ to $e_2$ and from $e_1$ to $\infty$. The $C_r$ integration contour in the PSS formulation of the exact hypermultiplet LEEA in eq.~(4.13) can now be interpreted as the contour integral {\it over the non-contractible $\alpha$-cycle of the elliptic curve} $\Sigma_{\rm H}$ \cite{bakas}, again in the very similar way as the Seiberg-Witten solution to the $SU(2)$-based N=2 gauge LEEA is written down in terms of the abelian differential $\lambda_{\rm SW}$ integrated over the periods of $\Sigma_{\rm SW}$ \cite{sw}. The most general (non-trivial) integration contour $C_r$ in eq.~(4.13) is given by a linear combination of the non-contractible $\alpha$ and $\beta$ cycles of $\Sigma_{\rm H}$, while an integration over $\beta$ is known to lead to a singularity \cite{ati}. This simple observation implies that the AH metric is the only regular solution. The perturbative (Taub-NUT) limit $k\to 1$ corresponds to the situation when $e_2\to e_1$, so that the $\beta$-cycle of $\Sigma_{\rm H}$ degenerates. The curve (5.1) then asymptotically approaches a complex line, $\tilde{Q}_4\sim \pm K\tilde{\xi}$. Another limit, $k\to 0$, leads to a (coordinate) bolt-type singularity of the AH metric in the standard parameterization (4.20) \cite{ati}. In the context of monopole physics, this corresponds to the coincidence limit of two centered monopoles. In the context of the hypermultiplet LEEA, $k\to 0$ implies $e_2\to e_3$, so that the $\alpha$-cycle of $\Sigma_{\rm H}$ degenerates, as well as the whole hypermultiplet action associated with eq.~(4.13). The two limits, $k\to 1$ and $k\to 0$, are related by the modular transformation exchanging $k$ with $k'$, and $\alpha$-cycle with $\beta$-cycle \cite{bakas}. The non-perturbative corrections to the hypermultiplet LEEA are therefore dictated by the hidden (in 4d) elliptic curve parametrizing the exact solution. \section*{Acknowledgements} I am grateful to Ioannis Bakas, Jim Gates and Olaf Lechtenfeld for useful discussions.
\section{INTRODUCTION} The Hall effect in the mixed state of high-temperature superconductors (HTS) is one of the most interesting and controversial problems related to vortex dynamics. Many experiments have shown that the Hall anomaly occurs not only in HTS, such as YBa$_{2}$Cu$_{3}$O$_{7-\delta}$ (YBCO)\cite {Hagen1990,Chien1991,Rice1992}, Bi$_{2}$Sr$_{2}$CaCu$_{2}$O$_{8}$ (Bi-2212) \cite{Samoilov1993}, and Tl$_{2}$Ba$_{2}$CaCu$_{2}$O$_{8}$ (Tl-2212)\cite {Hagen1991,Samoilov1994}, but also in conventional superconductors, for example, in the thin-film and the single-crystalline forms of Nb\cite {Hagen1990,Noto1976,Usui1968}, V\cite{Usui1968}, and In-Pb alloys\cite {Weijsenfeld1968}. So far, two kinds of sign reversals have been observed. The first is a simple, single sign reversal as observed in YBCO\cite{Rice1992,Samoilov1995} and La$_{2-x}$Sr$_x$CuO$_4$(LSCO)\cite{Matsuda1995}, and the second is a double sign reversal from positive to negative and then to positive again as the temperature decreases. The distinction between the two is that the single sign reversal is observed for the case of relatively low anisotropy while the double sign reversal is observed in the case of relatively high anisotropy, such as Bi-\cite{Samoilov1993}, Tl-\cite{Hagen1991,Samoilov1995}% , and Hg-based compounds\cite{WNKang1997,WNKang1999}. The anisotropy ratio is known to be on the order of $10^{4}$ for Tl-2212, $10^{2}\sim 10^{3}$ for LSCO, and $10\sim 10^{2}$ for YBCO, and the ratio for Hg-based superconductors\cite{YCKim1995} is between the values for Tl-2212 and YBCO. Recently, even a third sign reversal was observed in the low-temperature region for heavy-ion-irradiated Hg-based compounds\cite{WNKang1999c}. This observation is quite meaningful because this multiple sign reversal was predicted by Kopnin and depended on the behaviors of the density of states and of the gap of the superconductor\cite{Kopnin1996}. This multiple sign reversal is possible if there are localized or almost localized energy states in the superconducting state. Just after the detection of the Hall effect in Nb crystals\cite{Reed1965}, Bardeen and Stephen\cite{Bardeen1965} derived the flux-flow resistivity and the Hall resistivity due to vortex motion. However, in their theory, the sign of the Hall resistivity is always positive. Quite a few other theories, based on the flux-backflow\cite{Hagen1990}, two-band\cite{Hirsch1991}, or induced-pinning\cite{Wang1994} phenomena, have also been developed to explain the Hall effect in mixed states. However, the origin of the Hall anomaly is still not well understood. In this paper, we report the magnetic-field dependence of the Hall conductivity in the mixed state of HgBa$_{2}$CaCu$_{2}$O$_{6+\delta}$ (Hg-1212) and HgBa$_{2}$Ca$_{2}$Cu$_{3}$O$_{8+\delta}$ (Hg-1223) thin films. As expected, a double sign reversal is observed in these highly anisotropic superconductors. The measured Hall conductivities in the mixed states are better fitted by the form $\sigma_{xy}=C_{1}/H+C_{2}+C_{3}H$, which is different from the case of the less anisotropic YBCO superconductor\cite {Ginsberg1995} where $C_{2}$ is negligible. In this Hg-based superconductor, $C_{1}$ and $C_{2}$ depend strongly on the temperature, but that is not the case for $C_{3}$. $C_{1}$ scales as $\sim (1-t)^{n}$, which is partially understood from the temperature dependence of the gap and the coupling constant\cite{Kopnin1996}, but this understanding is not rigorous. Here, we claim that $C_{2}$, which appears only for highly anisotropic materials, scales as $C_{2}\sim (1-t)^{n^{\prime}}$. The critical exponent $n^{\prime}$ is $2.0\pm 0.2$ for Hg-1223 and $3.2\pm 0.1$ for Hg-1212. We observe, for the first time to the best of our knowledge, this scaling behavior of $C_{2}$ for Hg-based superconductors. A similar behavior was previously observed in LSCO\cite{Matsuda1995}, but was not analyzed. $C_{3}$ for these Hg-based thin films weakly depends on the temperature, which is a different behavior than those observed for YBCO and LSCO. \section{THEORETICAL BACKGROUD} Kopnin {\it et al.}\cite{Kopnin1993} and Dorsey\cite{Dorsey1992b} obtained the Hall conductivity by using the time-dependent Ginzburg-Landau (TDGL) theory in which the relaxation time of the order parameter was taken to be complex. Kopnin {\it et al.}\cite{Kopnin1993} claimed that the negative Hall effect was very much related to the energy derivative of the density of states. According to their theory, the Hall conductivity can be expressed by two contributions. The first contribution due to the vortex motion is proportional to $1/H$ and is dominant in the low-field region. The second contribution, which originates from quasiparticles, is proportional to $H$. An analysis of the Hall conductivity, based on the TDGL theory, for the YBCO single crystal was reported by Ginsberg and Manson\cite{Ginsberg1995}. In their paper, the Hall conductivity $\sigma_{xy}(H)$ was well explained by the sum of $H$- and $1/H$-dependent parts. However, the behavior of $% \sigma_{xy}(H)$ varies on a case-by-case basis. For example, the $H$% -dependent part for YBCO is replaced by a field-independent part for Tl-2212 \cite{Samoilov1994}. In the case of LSCO\cite{Matsuda1995}, the Hall conductivity is expressed as the sum of three terms : a $1/H$-dependent term, an $H$-dependent term, and an $H$-independent term. The temperature dependence of the coefficient of each component in the Hall conductivity was also investigated\cite{Matsuda1995,Ginsberg1995}. The coefficient of the $1/H $ term varies as $(1-t)^{n}$, where $t=T/T_{c}$ is the reduced temperature. $% n$ is observed to be 2 for YBCO and $2\sim 3$ for LSCO. Recently, Kopnin {\it et al.}\cite{Kopnin1996,Kopnin1995} calculated the Hall conductivity based on the kinetic equations and the TDGL theory. Their approach included an additional force due to the kinetic effects of changing the quasiparticle densities in the normal core and in the superconducting state. The total Hall conductivity is given by \begin{equation} \sigma_{xy}(H)=\sigma_{H}^{(L)}+\sigma_{H}^{(D)}+\sigma_{H}^{(A)} , \label{sigma1} \end{equation} where $\sigma_{H}^{(L)}$ comes from localized excitations in the vortex cores, $\sigma_{H}^{(D)}$ is from delocalized quasiparticles above the gap, and $\sigma_{H}^{(A)}$ is from the additional force due to the kinetic effects of charge imbalance relaxation. In the vicinity of $T_{c}$, this term can be expressed as \begin{equation} \sigma_{H}^{(A)} \sim \frac{1}{H \lambda} \left( \frac{d\nu}{d\zeta} \right) \Delta^{2} , \label{sigmaA} \end{equation} where ${d\nu}/{d\zeta}$ is the energy derivative of the density of states at the Fermi surface, $\lambda$ is the coupling constant, and $\Delta$ is the superconducting energy gap. $\sigma^{(L)}_{H}$ and $\sigma^{(A)}_{H}$ depend on $1/H$, while $\sigma^{(D)}_{H}$ is proportional to $H$. One important thing we have to notice is that first two terms on the right-hand side of Eq. (\ref{sigma1}) are always positive. For the dirty case, $\sigma^{(L)}_{H} $ is very small; hence, it can be neglected near $T_{c}$. A sign reversal can occur when $\sigma_{H}^{(A)}$ dominates over $\sigma_{H}^{(D)}$. \section{EXPERIMENTALS} High-quality Hg-1212 and Hg-1223 thin films were grown by using the pulsed laser deposition and post annealing method. The details are reported elsewhere\cite{WNKang1998,WNKang1999b}. The onset-transition temperatures, $% T_c$, are 127 K for Hg-1212 and 132 K for Hg-1223. The sizes of the specimens were 3 mm $\times$ 10 mm $\times$ 1 $\mu$m. A 20-T superconducting magnet system (Oxford Inc.) was used for the dc magnetic fields, and a two-channel nanovoltmeter (HP34420A) was used to measure the Hall resistivity ($\rho_{xy}$) and the longitudinal resistivity ($\rho_{xx}$) by using the standard dc five-probe method. The external magnetic field was applied parallel to the $c$ axis of the thin films, and the transport current density was $200\sim 250$ A/cm$^{2}$. Both the Hall resistivity and the longitudinal resistivity showed Ohmic behavior, i. e., corresponding to the flux-flow region, at the current used in this study. \section{RESULTS and DISCUSSION} We measured the longitudinal resistivities and the Hall resistivities of Hg-1212 and Hg-1223 thin films in the magnetic field region 0 T $\leq H \leq$ 18 T, and the results for Hg-1212 are shown in Fig. 1 while those for Hg-1223 are shown in Fig. 2. Compared to most of previous experiments performed at lower fields, we extended the magnetic field up to 18 T. The motivation for doing this was to check whether the previous analysis of the field dependence of the Hall conductivity based on the TDGL theory was valid even at this high field. Figures 1(a) and 2(a) show the field dependences of $\rho_{xx}(H)$ for various temperatures. $\rho_{xx}(H,T)$ increases monotonically with increasing temperature. In Figures 1(b) and 2(b), $% \rho_{xy}(H)$ is plotted and has a nearly linear dependence on the field in the high-field region. The sign of the Hall resistivity in the low-field region near the transition temperature becomes negative, which is opposite to the positive sign of the Hall resistivity for the normal state. The range of the field in which sign reversal is observed for Hg-1223 is narrower than that for Hg-1212. The insets of Figs. 1 and 2 show detailed representations of the low-field region. The Hall conductivity is typically defined as $\sigma_{xy}\simeq \rho_{xy}/\rho_{xx}^{2}$ by assuming $\rho_{xx}\gg \rho_{xy}$. In Fig. 3, the field dependences of the Hall conductivities of Hg-1212 (Fig. 3(a)) and Hg-1223 (Fig. 3(b)) are shown for various temperatures. Based on the theoretical prediction of Kopnin {\it et al.}\cite {Kopnin1996,Kopnin1993,Kopnin1995}, we analyze $\sigma_{xy}(H)$ by using \begin{equation} \sigma_{xy}(H)= \frac{C_{1}}{H}+C_{2}+C_{3}H , \label{sigma2} \end{equation} which is plotted with solid lines in Fig. 3. Compared to YBCO, the component $C_{2}$ is added for better fitting. The data are well fitted in the region of 115 K $\leq T \leq 125$ K for Hg-1212 and 125 K $\leq T \leq 130$ K for Hg-1223. In this figure, the downward curves, as approaching zero field, show a sign reversal, but the upward curves do not. If the curve is downward, $C_{1}/H$ is negative, but if the curve is upward, it is positive. The temperature dependences of $C_{1}$ and $C_{2}$ for Hg-1212 and Hg-1223 are shown in Fig. 4. Experiment shows that $C_{1}$ scales with temperature near $T_{c}$ as \begin{equation} C_{1}\sim (1-t)^{n} , \label{c1} \end{equation} where $n$ is $2.3\pm 0.2$ for Hg-1212 and $1.8\pm 0.3$ for Hg-1223, as shown in the Table I. The scaling form of $C_{1}$ can be partially understood from the temperature dependences of the gap and of the coupling constant in Eq. (% \ref{sigmaA}) based on the theoretical prediction by Kopnin. However, this has not yet been proven rigorously. Scaling of $C_{1}$ has also been reported for YBCO\cite{Ginsberg1995} and LSCO\cite{Matsuda1995} with $n$ values of 2 for YBCO and $2\sim 3$ for LSCO. These are not significantly different from those for Hg-1212 and Hg-1223. Compared to several other critical exponents, such as the magnetization scaling and the irreversibility lines, $n$ does not critically depend on the anisotropy. As shown in Fig. 4(b), $C_{2}$ steeply increases with decreasing temperature. Therefore, we can extract the following scaling form for $C_{2}$ near $T_{c}$ : \begin{equation} C_{2}\sim (1-t)^{n^{\prime}} , \label{c2} \end{equation} where $n^{\prime}$ is $3.2\pm 0.1$ for Hg-1212 and $2.0\pm 0.2$ for Hg-1223. Differently from the case of YBCO, in which $C_{2}=0$, we find that $C_{2}$ is not negligible for Hg-based superconductors. $C_{2}$ seems to be associated with the anisotropy ratio of the material, because the $C_{2}$ part of the resistivity becomes more explicit for the highly anisotropic superconductors, such as LSCO\cite{Matsuda1995} and Tl-1212\cite {Samoilov1994}. Specifically, a similar tendency to that shown in Fig. 4(b) for $C_{2}$ was observed in data previously reported for LSCO, but the temperature-scaling behavior was not determined. Not much information is reported for Tl-1212; however, the Hall conductivity is well fitted by $% \sigma_{xy}=C_{1}/H+C_{2}$. As explained before, the origins of the $1/H$ and $H$ dependences can be explained by the TDGL theory\cite {Kopnin1993,Dorsey1992b} or the microscopic theory\cite {Kopnin1996,Kopnin1995}, but neither of them can explain the scaling behavior of $C_{2}$. The coefficient $C_{3}$ of the term linear in $H$ shows a weak temperature dependence in both Hg-1212 and Hg-1223. This is different from the cases of underdoped and slightly overdoped LSCO\cite{Matsuda1995}, in which $C_{3}$ decreases as the temperature decreases. On the other hand, $C_{3}$ in YBCO decreases linearly with temperature. In order to investigate the effect of anisotropy on the coefficients $C_{1}$% , $C_{2}$, and $C_{3}$ and on the powers $n$ and $n^{\prime}$, we summarize our results along with previous results for HTS in Table I. As the anisotropy increases, the absolute value of $C_{1}$ and $C_{3}$ decrease while $C_{2}$ is increases. These values are evaluated at $t \simeq 0.92$. According to this tendency, $C_{2}$ is very small for the case of low anisotropy whereas $C_{3}$ is very small for the highly anisotropic case. As a result, in Eq. (\ref{sigma2}), the second term is negligible for YBCO, and third term is negligible for Tl-2212. In case of Hg-base superconductors, however, since the anisotropy ratio ranges between that of YBCO and that of Tl-2212, all terms in Eq. (\ref{sigma2}) are required, just as in the case of LSCO\cite{Matsuda1995}. The Hall conductivities measured up to very high magnetic fields (0 T $\leq H\leq 18$ T) for Hg-based superconductors are still well described by Eq. (\ref{sigma2}), but the temperature dependences of these coefficients have not yet been explained theoretically. \section{SUMMARY} We investigate the Hall effects for Hg-1212 and Hg-1223 thin films as functions of the magnetic field up to 18 T. The Hall conductivity in the mixed state is expressed well by $\sigma_{xy}(H) = C_{1}/H + C_{2} + C_{3}H$% . The coefficient $C_{1}$ scales with temperature as $(1-t)^{n}$ with $% n\simeq 2.3$ and 1.8 for Hg-1212 and Hg-1223, respectively; these values of $% n$ are comparable to the values observed for YBCO and LSCO. We find that $% C_{2}$ is more important for highly anisotropic compounds. $C_{2}$ is observed to follow the same scaling form, but with exponent $n^{\prime}\sim 3.2$ and 2.0 for Hg-1212 and Hg-1223, respectively. These scaling behaviors of $C_{1}$ and $C_{2}$ have not yet been explained theoretically. \acknowledgments This work is supported by the Ministry of Science and Technology of Korea through the Creative Research Initiative Program.
\section{\@startsection {section}{1}% {\z@}{5ex plus .2ex minus .4ex}% {1.5ex plus.4ex minus .1ex}% {\centering\ifpr@pstyle\else\ifx\undefined\reset@font\else% \reset@font\fi\large\fi\bf}} \def\subsection{\@startsection{subsection}% {2}{\z@}{3.25ex plus .4ex minus .4ex}% {1ex plus .2ex} {\centering\ifpr@pstyle\else\ifx\undefined\reset@font\else% \reset@font\fi\normalsize\fi\bf}} } \bigpage \newfont{\fourteencp}{cmcsc10 scaled\magstep2} \newfont{\titlefont}{cmbx10 scaled\magstep2} \newfont{\authorfont}{cmcsc10 scaled\magstep1} \newfont{\fourteenmib}{cmmib10 scaled\magstep2} \skewchar\fourteenmib='177 \newfont{\elevenmib}{cmmib10 scaled\magstephalf} \skewchar\elevenmib='177 \newfont{\ninemib}{cmmib9} \skewchar\ninemib='177 \makeatletter \newcommand\nonsequentialeqnum{ \nons@qeqtrue \@addtoreset{equation}{section} \def\arabic{section}.\arabic{equation}}{\arabic{section}.\arabic{equation}}} \newif\ifp@bblock \p@bblocktrue \newcommand\nopubblock{\p@bblockfalse} \newcommand\topspace{\hrule height 0pt depth 0pt \vskip} \newcommand\p@bblock{\begingroup \tabskip=\hsize minus \hsize \baselineskip=1.5\ht\strutbox \topspace-2\baselineskip \halign to\hsize{\strut ##\hfil\tabskip=0pt\crcr \the\Pubnum\crcr\the\date\crcr}\endgroup} \newcommand\YUKAWAmark{\hbox{ \ifpr@pstyle\ninemib\else\elevenmib\fi Yukawa\hskip1mm Institute\hskip1mm Kyoto \hfill}} \newtoks\date \newtoks\Pubnum \let\pubnum=\Pubnum \Pubnum={} \date={\today} \newcommand{\vspace{12pt plus .5fil minus 2pt}}{\vspace{12pt plus .5fil minus 2pt}} \def\@authoraddress{} \def\@title{} \def\title#1{\gdef\@title{\vspace{12pt plus .5fil minus 2pt} \begin{center}{\titlefont #1}\end{center}\par}} \def\@author#1{\vspace{12pt plus .5fil minus 2pt}\par\begin{center}{\authorfont #1} \end{center} \nobreak} \def\author#1{\expandafter\def\expandafter\@authoraddress\expandafter {\@authoraddress{\@author{#1}}}} \def\andauthor#1{\expandafter\def\expandafter\@authoraddress\expandafter {\@authoraddress{\vspace{12pt plus .5fil minus 2pt}\centerline{and}\@author{#1}}}} \def\authors#1{\expandafter\def\expandafter\@authoraddress\expandafter {\@authoraddress{\vspace{12pt plus .5fil minus 2pt}\noindent #1}}} \def\@address#1{\par\begin{center}{\sl #1}\end{center}\par} \def\address#1{\expandafter\def\expandafter\@authoraddress\expandafter {\@authoraddress{\@address{#1}}}} \def\andaddress#1{\expandafter\def\expandafter% \@authoraddress\expandafter {\@authoraddress{\par\centerline{\sl and}\@address{#1}}}} \renewcommand{\thanks}[1]{\footnote{#1}} \def\maketitle{\par \begingroup \def\fnsymbol{footnote}{\fnsymbol{footnote}} \thispagestyle{empty} \baselineskip=\paperbaselineskip \@maketitle \endgroup \setcounter{footnote}{0} \let\maketitle\relax \let\@maketitle\relax \let\@thanks\relax \let\@title\relax \let\@title\relax \let\@authoraddress\relax \let\thanks\relax} \def\@maketitle{% \ifpr@pstyle\vspace{-1.0cm}\else\vspace{-1.7cm}\fi \YUKAWAmark\vskip0.6cm \ifp@bblock\p@bblock \else\hrule height 0pt \relax \fi \@title \@authoraddress } \makeatother \begin{document} \pubnum{YITP-99-13} \date{February 1999} \title{Scaling Analysis of Fluctuating Strength Function} \author{Hirokazu Aiba} \address{Koka Women's College, 38 Kadono-cho Nishikyogoku, Ukyo-ku, 615 Kyoto, Japan} \andauthor{Masayuki Matsuo} \address{Yukawa Institute for Theoretical Physics, Kyoto University, 601-01 Kyoto, Japan} \maketitle \vfill \centerline{\fourteencp Abstract} \vspace{8pt} We propose a new method to analyze fluctuations in the strength function phenomena in highly excited nuclei. Extending the method of multifractal analysis to the cases where the strength fluctuations do not obey power scaling laws, we introduce a new measure of fluctuation, called the local scaling dimension, which characterizes scaling behavior of the strength fluctuation as a function of energy bin width subdividing the strength function. We discuss properties of the new measure by applying it to a model system which simulates the doorway damping mechanism of giant resonances. It is found that the local scaling dimension characterizes well fluctuations and their energy scales of fine structures in the strength function associated with the damped collective motions. \vspace{30mm} \section{Introduction} \label{sec:intro} Collective or single-particle modes of excitation which have high excitation energies usually become damped motions due to coupling with background compound states with high level density. Strengths of exciting such modes form a broad peak (when plotted as a function of the excitation energy of a nucleus) which are spread over an interval of excitation energy. A typical example is the giant resonances which have a damping width of MeV order while the resonance energy is several to a few tens MeV depending on quantum numbers and nuclear masses \cite{bertsch,woude}. Strength distribution plotted as a function of the excitation energy is called strength function. The strength function usually shows a smooth profile such as Lorentzian shape, but it also exhibits fluctuations called fine structures when the strength function is studied in high resolution experiments \cite{woude,hansen,kuhner,kilgus,kamer}. An important origin of the damping of the giant resonances in heavy nuclei is the doorway coupling \cite{bertsch}. The collectivity of the excitation mode is described well in terms of a coherent superposition of 1p-1h excitations, as often prescribed by the RPA models. On the other hand, the damping of the mode is caused by (besides the particle escaping whose effect is small in heavy nuclei) coupling to other kinds of excitation modes such as 2p-2h excitations or 1p-1h plus low-lying vibrational modes, which are called doorway states. Although the doorway states couple further with more complicated states, e.g. many-particle many-hole states and have their own spreading width, the total damping width of the giant resonances is explained reasonably well by the doorway coupling \cite{bertsch}. It is argued that the doorway states influence also the fine structure of the strength distribution\cite{kilgus,kamer}. However, relation between the doorway states and the fine structures remains to be understood since it would depend on the spreading of doorway states into more complicated states, which is not very well known. If we look at the strength function with the finest energy scale dissolving individual energy eigenstates by assuming small particle escaping, another mechanism of the strength fluctuation is expected to emerge. Because of the chaotic nature of the compound states at the high excitation energy, strengths associated with individual eigenstates would show statistical fluctuation following the Porter-Thomas distribution \cite{porter,brody}. When the particle escaping width is larger than mean level spacing so that individual level peaks overlap, the Ericson fluctuation emerges\ \cite{brody,ericson,bohigas-weidenmuller}. However it is natural to expect that these statistical behaviors based on the random matrix theories \cite{brody,dyson,mehta,bohigas} govern the strength fluctuation only at small energy scales. One should rather expect that the strength function associated with the damping of giant resonances involves different kinds of fluctuations coexisting at different energy scales. (Note also there exists an argument that experimental fine structures of particle decay strength distribution under a certain situation resembles large fluctuations in simulated spectra generated solely on the basis of the statistical Porter-Thomas fluctuation, known as the pandemonium \cite{hansen,pandemonium}. Although this is a another possible origin of the fine structures, the property of the statistical fluctuation should depend on the level density of compound states under discussion). Thus it is important to analyze the fluctuations of strength function in connection with the energy scales of the fluctuation. In this paper, we propose an new method to perform such an fluctuation analysis of the strength functions. With this method we characterize the strength fluctuation with focus on its scaling behavior by taking a similar approach widely adopted to analyze fluctuation phenomena having self-similar multifractal structure\cite{maccauly1,maccauly2,feder}. The multifractal fluctuations accompany a characteristic scaling property obeying the power laws, and they are found in various kinds of chaotic classical dynamical systems \cite{maccauly1,maccauly2,feder}. Furthermore multifractal fluctuations in quantum wave functions are found recently for electrons in disordered matter\cite{aoki,castellani,janssen}. The idea is also applied to the strength function fluctuations \cite{aiba,gorski}, and an approximate power scaling is found in a shell model calculation of the giant dipole resonances \cite{gorski}. In contrast to these multifractal analyses, however, we consider in this paper more general situations where the fluctuation does not necessarily follow the scaling laws, and we rather try to characterize the scaling behavior which may be dependent on the energy scales due to different coexisting fluctuation mechanisms. For this purpose, we introduce a new measure of fluctuation by extending the generalized fractal dimension \cite{hentschel-procaccia,halsey} used in the standard multifractal analysis. The new measure, which we call the local scaling dimension, characterizes the scaling property as a function of the energy scale of fluctuation. With use of simple examples, we will demonstrate in this paper how the local scaling dimension carries information on the dynamics of the damping process. For the analysis of the strength fluctuations, the autocorrelation function has been used widely as a measure of fluctuation in various contexts. It was applied for example to analyze the Ericson fluctuation in compound nuclear reactions to derive the particle decay width\ \cite{brody,ericson}, and also to the strength of giant resonances and the decay particle spectra \cite{hansen,kilgus}. The autocorrelation function and its Fourier transform is also used to characterize energy level fluctuations in the random matrix theories and chaotic quantal systems \cite{leviander,wilkie,lombardi,alhassid}. We will show that the present approach and the autocorrelation analysis are related to each other and that the local scaling analysis can be used to reveal the energy scales involved in the strength fluctuation. The present method treats in a single framework strength fluctuations in wide range of the energy scale, covering the statistical fluctuations at the finest scale around the level spacing order as well as those associated with the doorway states at the larger scale. In the present paper, therefore, we use the word fine structure to denote all kinds of strength fluctuations including the doorway fluctuations, which are sometimes called the intermediate structures in the literature. The paper is organized as follows. In section\ \ref{sec:ldim}, we describe the formalism of the fluctuation analysis and introduce the local scaling dimension. In addition, relation to the autocorrelation function is clarified. In section\ \ref{sec:numerical}, employing a schematic model which simulates the doorway damping mechanism of giant resonance, we discuss in detail how the local scaling dimension characterizes fluctuations in the model strength function. In particular, an emphasis is put on the relation between the profile of the local scaling dimension and the energy scales involved in the doorway damping process. Finally, section\ \ref{sec:conclusion} is devoted to conclusions. \section{Scaling Analysis of Strength Function} \label{sec:ldim} \subsection{Local Scaling Dimension} \label{sec:localscaling} The strength function is expressed as \cite{Bohr-Mottelson} \begin{equation} S(E)=\sum_i S_i\delta(E-E_i+E_0), \label{defstr} \end{equation} for exciting the nucleus with excitation energy $E$ by a probe which excite a mode under consideration if we neglect the coupling to continuum states of escaping particles and experimental resolutions for simplicity. Here $E_i$ and $E_0$ are the energy of discrete levels and the ground state energy of the whole nuclear many-body Hamiltonian, respectively, and $S_i$ denotes the strength of exciting the $i$-th energy level. Let us assume strengths are normalized as $\sum_i S_i=1$. By smoothing the strength function with respect to the excitation energy $E$ with a sufficiently large smoothing energy width, or by putting the strengths into energy bins with large bin width, a smooth profile (e.g. a Lorentzian shape) will show up. If small smoothing width or bin width is used, resultant distribution will exhibit fluctuation associated with fine structures of the original strength function. Thus the fluctuation of the strength function generally depends on the energy scale with which one measures. To evaluate the scale dependent fluctuation, we apply the method described below. It is an extension of the scaling analysis widely used to characterize multifractal structures\ \cite{maccauly1,maccauly2,feder}. Let us consider binned distribution of the strength function $S(E)$ by dividing whole energy interval under consideration into $L$ bins with length $\epsilon$. Strength contained in $n$--th bin is denoted by $p_n$, \begin{equation} p_n\equiv\sum_{i\in n{\rm -th~ bin}}S_i. \label{defp} \end{equation} To characterize the fluctuation of the binned strength distribution $\{p_n\}$, moments of the distribution are introduced. This leads to the so-called partition function $\chi_m(\epsilon)$ defined by \begin{equation} \chi_m(\epsilon)\equiv \sum_{n=1}^L p_n^m \\ =L\langle p_n^m\rangle. \label{partition} \end{equation} For completeness of the definition, we need to introduce an average of $\chi_m(\epsilon)$ with respect to the bin boundary. This is because there is still arbitrariness in Eq.(\ref{partition}) concerning how the boundaries of bins are chosen even if the bin width $\epsilon$ is fixed. To overcome this uncertainty, we make an average over different choices of the bin boundaries. When the average is done with sufficiently large numbers of choices so that the boundaries cover densely the energy axis, the averaged partition function becomes independent of the boundaries. It is of special interest if the partition function has a scaling property. If we consider a trivial case where the strength is uniformly distributed with no fluctuation, i.e., $p_n=const=1/L\propto \epsilon$ for all $L$ and $\epsilon$, then an power low scaling $\chi_m(\epsilon)\propto\epsilon^{(m-1)}$ holds for $\chi_m(\epsilon)$ with respect to the bin width $\epsilon$. When the fluctuation is present, the partition function generally deviates from this limit. An extreme case of large fluctuation is the situation where the strength is concentrated in a single energy level. In this case the partition function becomes constant $\chi_m(\epsilon)=1$ for any value of $\epsilon$ since the binned strength is then $p_n=1$ for only one bin, and $p_{n'}=0$ for the others. In other words the partition function $\chi_m(\epsilon)$ scales with zero power. If the fluctuation shows a fractal structure, i.e., it has a self-similar structure against the change of the scale, the partition function shows a power scaling with power different from the trivial cases, like \begin{equation} \chi_m(\epsilon)\propto\epsilon^{D_m^{\rm fractal}(m-1)}, \label{chifrac} \end{equation} or equivalently a linear scaling between $\log\chi_m(\epsilon)$ and $\log\epsilon$, \begin{equation} \log\chi_m(\epsilon)\approx (m-1){D_m^{\rm fractal}\log\epsilon}. \label{chifraclog} \end{equation} The scaling coefficient $D_m^{\rm fractal}$ called the generalized fractal dimension characterizes the multifractal\ \cite{hentschel-procaccia,halsey}. The trivial case of uniform distribution corresponds to $D_m^{\rm fractal}=1$, implying a one dimensional uniform object in the $E$ axis. The case of the single sharp peak corresponds to $D_m^{\rm fractal}=0$ since it is just a dot (zero dimensional object). It should be noted however that the partition function does not follow the power scaling law in general. The fluctuation of the binned strength may change when we change the bin width. This will happen when the strength function contains fluctuations having specific energy scales. For the case of the damping of giant resonance, fluctuation may reflect the doorway states such as 2p-2h configurations, for which their spreading width could be one of the energy scales involved. Also, when we look at very small energy scale comparable with average level spacing, strength fluctuation associated with individual energy levels may emerge. Therefore, it is more useful to analyze ``how the partition function scales for different energy scales'' than to seek the power scaling property. For this purpose, we introduce an extension of the generalized fractal dimension by defining the scaling coefficients as local linear coefficients of $\log\chi_m(\epsilon)$ vs. $\log\epsilon$ at each energy scale $\epsilon$, or equivalently defining by \begin{equation} D_m(\epsilon)\equiv {1\over m-1} \frac{\partial\log\chi_m(\epsilon)} {\partial\log\epsilon}. \label{scaledim} \end{equation} We call this quantity the local scaling dimension since it is defined at each energy scale $\epsilon$ and this is a function of $\epsilon$. Note also that if the local scaling dimension $D_m(\epsilon)$ is constant over a long interval of the energy scale $\epsilon$, it implies presence of the multifractal structure of the strength fluctuation. In this case the local scaling dimension reduces to the standard generalized fractal dimension $D_m^{\rm fractal}$ defined by Eq.\ (\ref{chifraclog}). In actual calculation of the local scaling dimension, we define it by means of finite difference under the change of a factor 2, \begin{equation} D_m(\sqrt{2}\epsilon) = {1\over m-1}\frac{ \log\chi_m(2\epsilon)- \log\chi_m(\epsilon)} {\log 2}, \label{approscaledim} \end{equation} rather than the derivative in Eq.\ (\ref{scaledim}). Using the finite difference, the calculation is very simple for all the moments. The local scaling dimension $D_2(\epsilon)$ defined by the derivative has an exact form (\ref{d2explicit}) expressed in terms of the individual energy levels $E_i$ and associated strengths ${S_i}$, as given in Appendix\ \ref{appa}. However, exact calculation with use of Eqs. (\ref{scaledim}) and (\ref{inichitheta2}) is more complicated than the calculation using the finite difference\ (\ref{approscaledim}) except for the second moment local scaling dimension $D_2(\epsilon)$. Note also that the finite difference gives an effective smoothing. We will discuss difference between the two definitions in Sec.\ \ref{sec:results}. \subsection{Schematic Examples} \label{sec:exampes} Let us illustrate how the local scaling dimension carries information on the fluctuation of strength function by using some schematic examples. \subsubsection{Lorentzian} \label{sec:lorentzian} As the first example, we consider the Lorentzian distribution \begin{equation} S(E)= {1\over 2\pi}\frac{\Gamma}{E^2 + \Gamma^2/4}. \end{equation} This represents a typical peak profile associated with many kinds of damping processes. The Lorentzian distribution shows a smooth variation of the strength centered around the peak. Discreteness of the energy levels is neglected. Note that the Lorentzian distribution has one characteristic energy scale which is the FWHM of the distribution, $\Gamma$. The local scaling dimension $D_m(\epsilon)$ associated with the Lorentzian is shown in Fig.\ref{figloren} as a function of the binning energy scale $\epsilon$. When the bin width $\epsilon$ is smaller than $\Gamma$, the local scaling dimension $D_m(\epsilon)$ takes the value close to 1, while $D_m(\epsilon) \sim 0$ for $\epsilon$ much larger than $\Gamma$. For $\epsilon$ around $\Gamma$, $D_m(\epsilon)$ decreases sharply with increasing $\epsilon$. This behavior arises because continuous and smooth aspect of the Lorentzian distribution is dominantly measured by the scaling analysis using small bin width $\epsilon \ll \Gamma$, and the local scaling dimension goes to the limit of uniform distribution $D_m(\epsilon)=1$. With large bin width $\epsilon \gg \Gamma$, most of strength is concentrated in a single bin, thus the local scaling dimension reaches to the sharp peak limit $D_m(\epsilon)= 0$. With the bin width $\epsilon$ being the same order of the width $\Gamma$ of the Lorentzian, a transition between the two limits occurs. Thus the profile of $D_m(\epsilon)$ relates to the characteristic energy scale $\Gamma$ of the Lorentzian. In fact the value of $\epsilon$ where $D_m(\epsilon)$ decreases most steeply corresponds well to $\Gamma$ as seen in Fig. \ref{figloren}. \subsubsection{GOE strength fluctuation} \label{sec:GOE} If we distinguish the strengths associated with individual energy levels of highly excited nucleus, the strengths fluctuate from level to level on top of a smooth profile of strength distribution. The level-by-level fluctuation originates from the chaotic nature of excited nuclei and is known to follow generic statistical rules described by the random matrix theory of the Gaussian orthogonal ensemble (GOE) \cite{brody,dyson,mehta,bohigas}. Let us characterize the GOE strength fluctuation by means of the scaling analysis. When the strength fluctuation is described by the GOE, the strengths $S_i$ for individual states fluctuate independently and their statistics follows the Porter-Thomas distribution \cite{porter,brody} at large $N_{\rm tot}$ limit ($N_{\rm tot}$ being the dimension of the random matrix). The distribution of energy levels $E_i$ follows a semi-circle level density in average, and the position of the individual energy levels fluctuates locally around the average. We treat here only the strength fluctuation by neglecting effects both of the average level distribution and of local energy level fluctuations. Namely we assume the uniform distribution of the energy levels. As discussed in Appendix\ \ref{appb}, influence of the local energy level fluctuation on the local scaling dimension is small due to the spectral rigidity of the GOE energy levels. In this treatment, the energy levels are chosen to be equally spaced: $E_i = id~ (i=1,2,...,N_{\rm tot})$, where $d$ is the level spacing. The local scaling dimension is then given by \begin{equation} D_m(l)={1\over m-1}\sum_{i=1}^{m-1}{l\over l+2i}, \label{dimgoederriv} \end{equation} where $l\equiv\epsilon/d$ (the bin width measured in the unit of the level spacing $d$) is used as the scaling parameter. We give its derivation in Appendix\ \ref{appb}. An example of the GOE strength function is shown in Fig.\ \ref{figgoe}(a). The strengths of individual states fluctuate strongly, i.e., there exist levels which have as large strength as factor tens of the average value of strength. This is because the Porter-Thomas distribution is a distribution with large skewness. Because of the strength fluctuation, the binned strengths also fluctuate when the bin width $\epsilon$ is not very large compared with the level spacing $d$. The fluctuation of the binned strength $\{p_n\}$ decreases as $\epsilon$ increases since the binning operates as an averaging over levels included in bins. The local scaling dimension $D_m(l)$, Eq.\ (\ref{dimgoederriv}), is plotted in Fig.\ \ref{figgoe}(b), which shows how fluctuation of the binned strength distribution changes as the energy scale (bin width) $\epsilon$ changes. When we use the bins whose width is comparable with the level spacing ($l \lesssim 10$), the binned distribution shows many spiky peaks caused by large $S_i$ components. This leads to small values of $D_m(\epsilon)$. The fluctuation of the binned distribution is still large for the bin width of order of ten level spacings ($l \sim 10^1$), where $D_m(\epsilon)$ deviate significantly from the uniform limit $D_m(\epsilon)=1$. It is washed out mostly only with large bin width of order of $100d$, where $D_m(\epsilon) \approx 1$ is realized. It is also noted that the local scaling dimension $D_m(\epsilon)$ for the higher moment (larger $m$) takes significantly smaller values compared with the second moment ($D_2(\epsilon)$) especially for the small bin widths $l \lesssim 10^1$. Since taking the higher moment emphasizes large components of the binned distribution $\{p_n\}$, the local scaling dimension $D_m(\epsilon)$ for higher $m$ reflects scaling property of those components. Small value of $D_m(\epsilon)$ for large $m$ indicates that the large components are distributed sparsely along the $E$ axis, reflecting the spiky distribution of the GOE strength. We evaluated also the local scaling dimension $D_m(\epsilon)$ based on the finite difference definition, both by using the analytic evaluation Eq.\ (\ref{dimgoedif}) and by using numerical evaluation adopting 60 realizations of GOE. They give essentially the same curves as the those plotted in Fig.\ \ref{figgoe}(b) based on Eq.\ (\ref{dimgoederriv}), and the difference is almost invisible in this plotting scale. When we include the energy level fluctuation, the local scaling dimensions is affected for small bin width $\epsilon$ comparable to the level spacing. However, the effect is not very large and becomes negligible for $\epsilon > 10d$, as discussed in Appendix\ \ref{appb}. \subsubsection{Poisson fluctuation} As the third example, we consider the Poisson fluctuation. If we consider experimental spectra where events are counted in channels which corresponds to excitation energy bins, there always exists fluctuation in the spectra that originates from the counting statistics. Even if the strength distribution is completely uniform, counts in bins fluctuate statistically obeying the Poisson distribution. Relating the normalized `strength' $p_n$ to the counts $\{ r_n \}$ in bins by $p_n= r_n/N$ ($N$ being the total counts), the partition function $\chi_m$ is expressed in terms of the $m$-th moment of the Poisson distribution. Using the average number of counts $l=\langle r_n \rangle$ as the scale of the bin width, the local scaling dimension $D_m$ is given by \begin{eqnarray} D_m(l)&= &{1\over(m-1)}\frac{d\log(\langle r_n^m\rangle/l)}{d\log l} \nonumber \\ &= &{l \over l+1},~ {l^2+{3\over2}l \over l^2 + 3l+1},~ {l^3+4l^2+{7\over3}l \over l^3+6l^2+ 7l+1},~ {l^4+{15\over2}l^3+{25\over2}l^2+{15\over4}l \over l^4+10l^3+25l^2+ 15l+1} \ \ (m=2,3,4,5). \label{dmpoisson} \end{eqnarray} The local scaling dimension takes the value significantly smaller than one for small $l$ ($l \lesssim 10$). It monotonically increases with increasing $l$ and approaches to one. The behavior is similar to that of the GOE strength fluctuation, Fig.\ref{figgoe}(b), but the values of $D_m$ for the Poisson fluctuation is closer to $D_m=1$ than the GOE fluctuation since the fluctuation associated with the Poisson distribution is smaller than the GOE fluctuation. Expression\ (\ref{dmpoisson}) indicates that effects of the counting statistics diminish for $l$ larger than several tens for which the uniform limit $D_m \approx 1$ is almost achieved. We neglect the counting statistics in the other part of this paper. \subsection{Relation to Autocorrelation Analysis} One often uses the autocorrelation function to characterize fluctuations of the strength function. The autocorrelation function for the strength function, defined by \begin{eqnarray} C_2(\epsilon) &\equiv &\int S(E)S(E+\epsilon)dE,\\ \label{def1c2} &=&\sum_{i,j}S_iS_j\delta(\epsilon-E_j+E_i), \label{def2c2} \end{eqnarray} quantifies the fluctuation correlation as a function of the displacement energy $\epsilon$. Dependence of $C_2(\epsilon)$ on the displacement energy $\epsilon$ reveals the characteristic energy scale involved in the strength fluctuations. Since the local scaling dimension has a similar property as discussed above, one may expect relation between the autocorrelation function and the local scaling dimension. In fact, we can prove that $C_2(\epsilon)$ is related to the partition function $\chi_2(\epsilon)$ and the local scaling dimension $D_2(\epsilon)$ for the second moment. To this end, we introduce an integral function of $C_2(\epsilon)$ by \begin{equation} B_2(\epsilon)\equiv \int_{-\epsilon}^\epsilon C_2(\epsilon')d\epsilon' \\ =\sum_{ij}S_iS_j\theta(\epsilon-|E_i-E_j|). \label{b2theta} \end{equation} Comparing Eq.\ (\ref{b2theta}) with the closed form expression Eq.\ (\ref{inichi2theta}) for the partition function, we find an equation \begin{equation} (1+\epsilon{d\over d\epsilon})\chi_2(\epsilon) =B_2(\epsilon). \label{derivchi2b2} \end{equation} Thus, the partition function is expressed as \begin{equation} \chi_2(\epsilon)={1\over\epsilon}\int_0^\epsilon B_2(\epsilon')d\epsilon', \label{chi2b2} \end{equation} in terms of the integral of $B_2(\epsilon)$. Using this expression for Eq.\ (\ref{scaledim}), the local scaling dimension $D_2(\epsilon)$ is written as \begin{equation} D_2(\epsilon)=\frac{B_2(\epsilon)-\chi_2(\epsilon)}{ \chi_2(\epsilon)}, \label{dmbm} \end{equation} which is expressed in terms of the integral $B_2(\epsilon)$ and the double integral $\chi_2(\epsilon)$ of the autocorrelation function $C_2(\epsilon)$. The close connection between the autocorrelation function $C_2(\epsilon)$ and the local scaling dimension $D_2(\epsilon)$ can be seen easily for the case of the Lorentzian distribution. For the Lorentzian distribution, the autocorrelation function $C_2(\epsilon)={1\over\pi}{\Gamma\over\Gamma^2 + \epsilon^2}$ decreases monotonically with increasing $ \epsilon$ and its value changes most steeply around $\epsilon \sim \Gamma$, where $\Gamma$ is the FWHM of the Lorentzian strength distribution. On the other hand, the local scaling dimension $D_2(\epsilon)$ is given with use of Eq.\ (\ref{dmbm}) by \begin{equation} D_2(\epsilon) = \frac{{\Gamma\over 2\epsilon}\log((\epsilon/\Gamma)^2+1)} {\tan^{-1}{\epsilon\over\Gamma} - {\Gamma\over 2\epsilon}\log((\epsilon/\Gamma)^2+1)}. \end{equation} This function also decreases monotonically from $D_2=1$ at $\epsilon=0$ with increasing $\epsilon$. The point of its steepest slope, $\epsilon=0.80\Gamma$, (or $\epsilon=1.8\Gamma$ when plotted as a function of $\log\epsilon$) corresponds well to the FWHM of the Lorentzian. This correspondence holds also for the local scaling dimensions $D_m$ of higher moments as seen in the previous subsection. There are two figures of merit for the use of the scaling analysis in comparison with the autocorrelation analysis. Firstly, the local scaling dimension $D_m(\epsilon)$ has smooth $\epsilon$ dependence while the autocorrelation function $C_2(\epsilon)$ is a sum of the delta functions (if we do not include any smoothing procedure). This difference can be seen in Eq. (\ref{dmbm}) where the $D_2(\epsilon)$ is expressed in terms of the integral functions of $C_2(\epsilon)$. Secondly, the local scaling dimensions $D_m(\epsilon)$ for $m>2$ carry information on the higher moments of strength fluctuation. \section{Analysis of Doorway Damping Model} \label{sec:numerical} In the following we shall discuss fluctuations of the strength function which is associated with damping phenomena of a specific state, e.g., a collective vibrational state, embedded in the background states in the highly excited nuclei. Keeping in mind the damping mechanism of the giant resonances, we herewith adopt a model system in which the damping of a collective state takes place through coupling to doorway states which consist of only a part of the background states. The strength function of the collective state exhibits not only spreading of the strength common for all damping phenomena but also fine structures reflecting the presence of the doorway states. We shall discuss in detail how the scaling analysis describes characteristics of this kind of strength fluctuations. \subsection{Model} The Hamiltonian of the model is given by \begin{equation} H=\omega_c|c\rangle\langle c|+V_{\rm doorway}+H_{\rm bg}, \label{hamiltonian} \end{equation} where $|c\rangle$ denotes the collective state, whose energy is $\omega_c$. The third term represents the Hamiltonian for the background states, which include the doorway states as well as the other background states. The second term $V_{\rm doorway}$ represents the coupling of the collective state $|c\rangle$ to the doorway states. Here we keep in mind the picture that shell model many-particle many-hole excitations form basis states $\{|\mu\rangle\}$ of the background states and the interaction among the basis states causes configuration mixing. As a simplified treatment, we assume that $H_{\rm bg}$ has diagonal term representing the unperturbed energies of the basis states, and for the interaction among the basis states we employ a GOE Hamiltonian. Namely, \begin{equation} H_{\rm bg}=\sum_\mu\omega_\mu|\mu\rangle\langle \mu|+H_{\rm GOE}, \label{bghamil} \end{equation} \begin{equation} H_{\rm GOE}=\sum_{\mu} v^{\rm d}_{\mu\mu}|\mu\rangle\langle\mu|+ \sum_{\mu > \nu}v^{\rm nd}_{\mu\nu}(|\mu\rangle\langle \nu|+h.c.). \label{goehamil} \end{equation} The basis energy $\omega_\mu$ is given by an equidistant model, \begin{equation} \omega_\mu=(-{N_{\rm bg}\over 2}+\mu)d~~~(\mu=1,\cdots,N_{\rm bg}), \label{bgenergy} \end{equation} where $d$ is the level spacing of the background states and $N_{\rm bg}$ represents the total number of backgrounds. Matrix elements $v^{\rm d}_{\mu\mu}$ and $v^{\rm nd}_{\mu\nu}$ ($\mu\neq \nu$) for the GOE Hamiltonian are independent Gaussian random variables with the zero mean and the variance satisfying $\langle (v^{\rm d}_{\mu\mu})^2\rangle = 2 \langle (v^{\rm nd}_{\mu\nu})^2\rangle$. Because of the interaction, the background states have their own spreading width, denoted $\gamma$ hereafter, which can be estimated as, \begin{equation} \gamma=2\pi{\langle (v^{\rm nd}_{\mu\nu})^2\rangle \over d}. \label{bggamma} \end{equation} The background Hamiltonian\ (\ref{bghamil}) is equivalent to the model adopted in ref.\ \cite{guhr} and the same as the Wigner ensemble random matrix model\ \cite{wigner}. The second term in Eq.\ (\ref{hamiltonian}) representing the coupling between the collective state and doorway states is given by \begin{equation} V_{\rm doorway} = \mathop{{\sum}'}_{\mu={\rm doorway}} V_\mu^{\rm door} (|c\rangle\langle \mu|+h.c.). \label{doorwaycouple} \end{equation} Here the interaction is present only for a part of the background $\mu$ states which are supposed to be the doorway states of the damping, and the summation in Eq.(\ref{doorwaycouple}) runs only over the doorway states. The doorway $\mu$ states are selected in every $L$ $\mu$-states (i.e., $\mu$ states with $\mu=L,2L,3L,...$). Accordingly, the level spacing of doorway states is $D=Ld$. The coupling matrix elements $V^{\rm door}_\mu$ are random variables taken from a Gaussian distribution with zero mean. By diagonalizing the Hamiltonian with use of the basis consisting of all the $\mu$ states and the collective state $|c\rangle$, we obtain for each realization of random variables the energy eigenstates $|i\rangle$, their energy $E_i$ and associated strength $S_i=|\langle c|i \rangle|^2$ of the collective state. The strength function of the collective state \begin{equation} S(E)=\sum_i S_i\delta(E-E_i) \label{strfun} \end{equation} is thus obtained. Since the collective-doorway couplings are uniform with respect to the energy of the background states, the strength function exhibits as a global profile the Lorentzian distribution. In fact, when we average the strength distribution over many realizations of the Hamiltonian and make smoothing with respect to the energy $E$, the averaged strength function $\tilde{S}(E)$ becomes very close to the Lorentzian whose FWHM is given by the golden rule estimate \begin{equation} \Gamma=2\pi{\langle (V^{\rm door}_\mu)^2\rangle \over D}. \label{gamma} \end{equation} In this paper we do not discuss the global spreading profile since it is trivial in this model and it often has simple profile such as Lorentzian also in the experimental data. Instead we deal only with fine structures and fluctuations of the strength function. It is then useful to remove the global smooth profile by normalizing the strength $S_i$ as \begin{equation} \bar{S}_i ={\cal N}{S_i\tilde{\rho}(E_i) \over \tilde{S}(E_i)}. \label{unfold} \end{equation} Here, $\tilde{S}(E)$ and $\tilde{\rho}(E)$ are obtained by averaging the calculated strength function, Eq.(\ref{strfun}), and the level density $\rho(E)=\sum_i\delta(E-E_i)$ over realizations and by smoothing with respect to $E$ with use of the Gaussian weighting and the Laguerre polynomials as adopted in the Strutinsky method\ \cite{ring-schuck}. $\cal{N}$ is an overall normalization factor to guarantee $\sum_i\bar{S}_i=1$. As we discussed in connection with the GOE strength fluctuation (Sec.\ \ref{sec:GOE} and Appendix\ \ref{appb}), effect of the energy level fluctuation on the fluctuation of the strength function is small except for very small energy scale less than about ten times of the level spacing $d$. Furthermore the level fluctuations within a small energy interval is described by the GOE model as far as $\gamma$ is not small. Thus we neglect the fluctuation of the energy levels in the following analysis. In other words, we use the level order $i$ as the scale of the excitation energy of the levels instead of the energy $E_i$ itself. This is equivalent to assume that the energy levels are equidistantly distributed as $E_i = id$. In the following, we treat unfolded strength function $\bar{S}(E)=\sum_i\bar{S}_i\delta(E-E_i)$ thus obtained. The number of background states is fixed to $N_{\rm bg}=8191$ so that the size of the Hamiltonian matrix has the dimension of 8192. The unperturbed energy of the collective state is placed at the energy center of the spectrum by putting $\omega_c=0$, and we set the FWHM of the global strength distribution $\Gamma=2000d$ so that the most of the strengths are located dominantly in the central region of the total spectra. For the interaction matrix elements in $H_{\rm bg}$, or equivalently, the spreading width $\gamma$ of the background states, we use several different values $\gamma=64d$, $128d$, $256d$, and $512d$ to see the dependence on $\gamma$. The spacing $D$ between the doorway states is also varied as $D=64d$, $128d$, and $256d$, corresponding to the number of doorways 128, 64, 32, respectively. By generating random numbers for the Gaussian variables in the Hamiltonian, a realization of the Hamiltonian is obtained. The spectra and the strength distribution are calculated for each realization. We perform an average over many realizations of the Hamiltonian to obtain the features which is independent of specific realizations. For this purpose we generate repeatedly the random numbers in $V_{\rm doorway}$, but fix the coefficients in $H_{\rm bg}~ (H_{\rm GOE})$ since the results do not depend very much on realizations of $H_{\rm GOE}$ because of sufficiently large dimension $N_{\rm bg}$. For the ensemble average, we perform averaging over 60 realizations, which is enough to obtain almost exact results for the GOE fluctuations (Sec.\ \ref{sec:GOE}). The partition function is calculated for the bin widths of powers of 2 of the level spacing $d$, i.e., $\epsilon=ld$, $l = 1,2,4,8,16,...,4096$. Since the number of levels is finite, we need to take care of the edges of the strength distribution. The bins placed at either edge of the spectra have smaller bin width than those in the other part of the spectra, and this would affect the partition function when the bin width is not very small compared to the whole spectrum width. To avoid the edge effect, we assumed a periodic continuation of the spectra by imposing $\bar{S}_{i+N_{\rm tot}}=\bar{S}_i$, where $N_{\rm tot}$ is the total number of levels $N_{\rm tot}=N_{\rm bg}+1$. For the calculation of the partition function, we performed averaging over 16 bin boundary positions by sliding the bin boundaries by $ l/16$ (we average over $l$ boundary positions for $l<16$), which provides sufficient averaging. An example of the strength function is shown in Fig.\ \ref{figvarstr}(a). Qualitatively the distribution looks a Lorentzian distribution as a gross profile. However, the distribution has large fluctuation around the Lorentzian shape. In addition to apparent level-to-level fluctuations, one can recognize clustering of strengths. The latter appears to reflect the distribution of the collective-doorway couplings as seen from the comparison with Fig.\ \ref{figvarstr}(b), which shows the strength distribution which is obtained by assuming pure doorway states, i.e., by neglecting the interaction between the doorway states and the other background states. Note however that it is very hard to identify individual doorway states in the strength distribution of Fig.\ \ref{figvarstr}(a) since in this example the spreading width of doorway states, $\gamma$, is larger than the spacing between the doorways, $D$. The strengths associated with individual doorways overlap with each other. Fig.\ \ref{figvarstr}(c) shows the distribution of the normalized strength (corresponding to Fig.\ \ref{figvarstr}(a)) for which we apply the scaling analysis. Fig.\ \ref{fignns}(a) shows the nearest neighbor level spacing distribution (NND) for $D=128d$ and $\gamma=256d$. The NND follows the Wigner distribution \cite{brody} perfectly, indicating that the fluctuation of the energy level at a short energy interval obeys the GOE random matrix limit. By using the same Hamiltonian as our background Hamiltonian $H_{\rm bg}$, Guhr et al.\cite{guhr} showed that the GOE limit of the NND is reached as the spreading width of the background states $\gamma$ becomes larger than several tens of the level spacing $d$. Fig.\ \ref{fignns}(b) shows the statistical distribution of strength $\bar{S}_i$ of individual level for $D=128d$ and $\gamma=256d$. Here the histogram of distribution of square root strength $\bar{S}_i^{1/2}$ is plotted together with the Porter-Thomas distribution drawn by the dashed curve. Note that the Porter-Thomas distribution $P(S)=\exp(-S/2)/\sqrt{2\pi S}$ has a Gaussian shape in this plotting. Deviation from the Porter-Thomas distribution is clearly seen. The excess of large and small strengths compared to the Porter-Thomas distribution indicates that the strength fluctuation is larger than the GOE limit. The deviation is caused by the clustering of the strength distribution, which originates from the fine structures of the strength functions associated with doorway states of the damping of the collective state. \subsection{Scaling Analysis} \label{sec:results} \subsubsection{Characteristics of $D_m(\epsilon)$} Let us analyze the strength fluctuation by applying the scaling analysis which we introduced in the previous section. Figures\ \ref{figpartition} and\ \ref{figmodelfra} display the partition functions $\chi_m(\epsilon)$ and the local scaling dimensions $D_m(\epsilon)$, respectively, for the strength functions calculated with four different values of $\gamma$ and fixed value of $D=128d$. From the partition function $\chi_m(\epsilon)$, one sees that the fluctuation does not follow the power scaling law (or linear relation in the log-log plot). Correspondingly, the local scaling dimension $D_m(\epsilon)$ varies as a function of $\epsilon$. The local scaling dimensions in the four cases show common features. Namely, at small values of $\epsilon$, $D_m(\epsilon)$ monotonically increases with $\epsilon$. As $\epsilon$ increases further, $D_m(\epsilon)$ starts to decrease, but it turns to increase again. At the largest value of $\epsilon$ comparable to the whole spectrum range, $D_m(\epsilon)$ converges to the uniform limit $D_m(\epsilon)=1$. Comparing with the local scaling dimension $D_m(\epsilon)$ for the GOE strength fluctuation indicated with dashed curve in the figure, one sees that the behavior of $D_m(\epsilon)$ at small $\epsilon$ follows the GOE limit whereas it deviates from the GOE limit as $\epsilon$ increases. It is seen that the energy scale $\epsilon$ where $D_m(\epsilon)$ starts to deviate from the GOE limit, which we denote $\epsilon^{*}$, has strong correlation to the value of $\gamma$. As we discussed in terms of the level spacing distribution, the GOE limit is realized as far as the small energy interval is concerned. However, mixing between the background states (including the doorway states) takes place only within a finite energy interval, that is, the spreading width $\gamma$ of the background states. Therefore, the fluctuations far below the energy scale $\gamma$ will obey the GOE limit, while those comparable to or larger than $\gamma$ may deviate from the GOE limit. In fact, one observes in Fig.\ \ref{figmodelfra} that the energy scale $\epsilon^{*}$ where $D_m(\epsilon)$ starts to deviate from the GOE limit is related approximately to $\gamma$ as $\epsilon^{*} \sim 1/5 \gamma$. Figure\ \ref{figmodelfra} also shows clearly that the decrease of $D_m(\epsilon)$ at the intermediate energy scale $\epsilon$ depends strongly on the parameter $\gamma$. In fact, the value of $\epsilon$, denoted $\epsilon^{**}$, where $D_m(\epsilon)$ decreases most sharply corresponds well to the value of $\gamma$, indicated by the arrow in the figure, and it moves as one changes $\gamma$. Namely $\epsilon^{**} \sim \gamma$ is observed for all panels in the figure. The close connection between $\gamma$ and the decrease of $D_m(\epsilon)$ can be explained as follows. The presence of the doorway states causes the fine structure or the clustering of the strength distribution. When the spreading width $\gamma$ of the doorway states is small, the doorway states form many peaks (clusters) where width of individual peaks is given by $\gamma$. Using the bin width $\epsilon$ comparable with the spreading width $\gamma$, the scaling analysis uncovers the profile of the peaks associated with the doorways. Since the profile will look the Lorentzian shape of width $\gamma$, the local scaling dimension $D_m(\epsilon)$ decreases sharply around $\epsilon \sim \gamma$ as we found for the smooth Lorentzian distribution (Sec.\ref{sec:lorentzian}). When $\gamma$ is larger than the spacing $D$ of the doorway states, the peaks overlap with each other and one cannot trace the individual doorway states. However, the fluctuations of the strength distribution still reflects the spreading width $\gamma$ of doorway states since the fluctuations associated with the doorway states are not completely smeared out. Thus correspondence between the spreading width $\gamma$ of doorway states and the point of steepest decrease in $D_m(\epsilon)$ holds also for the cases of $\gamma \gtrsim D$. This situation corresponds to Figs.\ \ref{figmodelfra}(c) and (d). The increase of $D_m(\epsilon)$ at the largest energy scales arises from another features of the doorway states, that is, the fluctuation in the collective-doorway coupling. If we consider only the doorway states and neglect their coupling to other backgrounds, the strengths are concentrated only on the doorway states, as shown in Fig.\ \ref{figvarstr}(b). In this limit the strengths of the doorways follow a GOE strength fluctuation, since the collective-doorway coupling matrix elements is chosen as Gaussian random variables. It should be noted that the energy scale associated with this GOE is not the level spacing $d$ of the whole spectrum, but the spacing $D$ between the doorways, which is much larger than $d$. The dotted curve in Fig.\ \ref{figmodelfra} represents the GOE strength fluctuation characterized by the level spacing $D$ of the doorway states. When the bin width $\epsilon$ is taken larger than the spreading width $\gamma$ of the doorway states, the fluctuations at smaller energy scale are smeared out, and only the fluctuation of the collective-doorway coupling remains. In all cases in Fig.\ \ref{figmodelfra}, the local scaling dimensions follow this curve at the largest energy scales, indicating that the strength fluctuation in this domain reflects the structure of the collective-doorway coupling. One may also notice that the dip of $D_m(\epsilon)$ around $\epsilon \sim \gamma$ becomes small and $D_m(\epsilon)$ approaches to the GOE limit as one increase the value of $\gamma$. It is also noted that for small value of $\gamma$, the local scaling dimension $D_m(\epsilon)$ for the higher moments $m$ takes significantly smaller value around the intermediate $\epsilon$ than $D_m(\epsilon)$ for the lower moments (eg. $D_5$ vs. $D_2$), while, for large $\gamma$, $D_m(\epsilon)$ for all moments takes similar value. The higher moments are sensitive to components of the large strength $p_n$. Therefore, small value of $D_m(\epsilon)$ for higher moments implies that the strength distribution is very spiky. In other words the strengths are clustered and form large peaks. On the contrary, similar values of $D_m(\epsilon)$ close to 1, obtained for large $\gamma$ values, implies relatively uniform distribution of strengths. The above two features characterize the way how the fine structures (or clustering) associated with the doorway states diminishes as $\gamma$ increases. In Fig.\ \ref{figotherdfra}, we show dependence on the parameter $D$, the spacing of the doorway states, with the spreading width $\gamma$ being kept constant. It is seen that the values of local scaling dimension $D_m(\epsilon)$ at larger energy scale $\epsilon$ depends on the spacing $D$ of doorway states. When $D$ is small (e.g. Fig.\ \ref{figotherdfra}(a), where $D=64d$), $D_m(\epsilon)$ takes larger value close to the GOE limit than the case of larger $D$. This implies that when the strength of each doorway states overlaps significantly with each other, namely, the spacing $D$ of doorway states is much smaller than spreading width $\gamma$, the fluctuation associated with the doorways tends to be smeared. On the other hand, when $\gamma \lesssim D$, the fluctuation associated with the doorways stands up clearly, then $D_m(\epsilon)$ exhibits large deviation from the GOE limit. However, it is more important to note that the spacing $D$ of the doorways does not affect very much the behavior of $D_m(\epsilon)$ at smaller energy scale. In both cases of Fig.\ \ref{figotherdfra}, the values $\epsilon^{*}$ and $\epsilon^{**}$ where $D_m(\epsilon)$ deviates from the GOE and decreases most sharply, respectively, have relations to $\gamma$, $\epsilon^{*}\sim 1/5\gamma$ and $\epsilon^{**}\sim\gamma$, as was observed in Fig.\ref{figmodelfra}, indicating that these features hold in spite of the change of $D$ values. In Fig.\ \ref{figmodelfra} we also plot with the dashed-dotted curve $D_2(\epsilon)$ which is evaluated by means of the exact expression Eq.\ (\ref{d2explicit}) based on the derivative definition Eq.\ (\ref{scaledim}). It is seen that difference between the exact evaluation and the evaluation Eq.\ (\ref{approscaledim}) based on the finite difference is almost indistinguishable. This justifies the use of the finite difference evaluation of the local scaling dimension. \subsubsection{$D_m(\epsilon)$ as a measure of doorway spreading width} The close relation between the spreading width $\gamma$ and the profile of the local scaling dimension may be used as a tool to estimate the spreading width $\gamma$ of the doorway states. The number of doorway states which are involved in the damped collective states such as the giant resonances is not small. In such cases, it is difficult to evaluate the spreading width by distinguishing individual doorway states from the whole strength distribution since the strengths associated with each doorway states may overlap with each other. On the other hand, since the scaling analysis discussed in this paper does not assume separation of the peaks, it is applicable in such overlapping cases. Indeed, the numerical analysis presented above indicates that the characteristic behavior of $D_m(\epsilon)$, namely the decrease around the energy scale comparable with $\gamma$, shows up even in cases where the spreading width $\gamma$ is larger than the spacing $D$ between the doorway states. Thus one may estimate the spreading width $\gamma$ by locating the point $\epsilon^{**}$ where $D_m(\epsilon)$ decreases most steeply. It is instructive to compare with the autocorrelation analysis, which is often used to analyze the strength fluctuations. As discussed in the previous section, there is the relation between the autocorrelation analysis and the scaling analysis, in particular, between the autocorrelation function $C_2(\epsilon)$ and the local scaling dimension $D_2(\epsilon)$ for the second moment. Figure\ \ref{figc2} shows the autocorrelation function $C_2(\epsilon)$ for $D=128d$ and $\gamma=256d$, which is also calculated for the normalized strength distribution (e.g. Fig.\ \ref{figvarstr}(c)) and averaged over 60 realizations as done for the scaling analysis. The autocorrelation function $C_2(\epsilon)$ exhibits the following property. It has a broad peak centered around $\epsilon = 0$ which is built upon a flat `background'. The broad peak contains the information of the fine structure of the strength fluctuations which arises from the doorway states. The width of the broad peak corresponds well to the spreading width $\gamma$ of the doorway states. It is helpful to remind that the autocorrelation function used to analyze the Ericson fluctuation in cross sections of highly excited compound nuclei, where resonance overlaps with neighboring ones. In that case, the width of the autocorrelation function is related to the particle decay width of individual states. In our case, on the other hand, the width of the autocorrelation function is related to the spreading width of the doorway states. Comparing Fig.\ \ref{figmodelfra} and Fig. \ref{figc2}, one sees close relationship between the local scaling dimension $D_2(\epsilon)$ and the autocorrelation function $C_2(\epsilon)$, both of which decrease sharply at the energy scale corresponding to $\gamma$. \subsubsection{Analysis without ensemble average} The results presented above are obtained by averaging over sixty different strength functions which are calculated for every realization of the random Hamiltonian. This is possible for the model systems. However, there will be a situation where one has only a single experimental data for a given nucleus so that one cannot perform ensemble average. When one applies the scaling analysis to a single given spectrum, the result will contain both the property which is common for the ensemble which the data belongs to and the behavior which is specific to the given realization of spectra. The latter can be regarded as fluctuation associated with different realizations. The fluctuation indicates an uncertainty when we infer the ensemble averaged properties from analysis of a single spectrum. To evaluate this uncertainty in the local scaling dimension $D_m(\epsilon)$, we calculate the local scaling dimensions $D_m(\epsilon)$ for each of 60 realizations of the model Hamiltonian, and evaluate the average and the standard deviation in $D_m(\epsilon)$. The results are plotted in Fig.\ \ref{figflucd2}. We also plot in the same figure an example of $D_m(\epsilon)$ for an arbitrarily chosen realization. It is seen that the local scaling dimension calculated for a single realization shows $\epsilon$ dependence which is qualitatively same as the ensemble averaged $D_m(\epsilon)$. The standard deviation from the average is smaller than the typical size of $\epsilon$ dependence of $D_m(\epsilon)$ for the cases shown in the figure. It is seen that the deviation increases as $m$ increases, namely, as the higher moments are concerned. The standard deviation in $D_m(\epsilon)$ is estimated approximately as $\sigma(D_m) \approx \sqrt{3}m/\sqrt{N_{\rm tot}}$ by assuming independent Porter-Thomas fluctuation for the individual strengths $S_i$. Thus, the local scaling dimensions $D_m(\epsilon)$ calculated for a single data are not affected very much by the fluctuation due to the realizations and keep information on the background mechanisms such as the spreading width $\gamma$ of the doorway states as far as a sufficient number of levels are included in the analysis and the low order moments are concerned. \section{Conclusions} \label{sec:conclusion} We proposed a new method to analyze fluctuation properties of the strength functions by means of the scaling analysis similar to the multifractal analysis. In order to treat the non-scaling fluctuations, we introduced the local scaling dimension $D_m(\epsilon)$ defined as a function of a scaling parameter $\epsilon$, which is the energy bin width subdividing the strength function. This quantity reduces to the generalized fractal dimension if a self-similar structure exists in the strength fluctuation. The local scaling dimension also has a close connection with the autocorrelation function of the strength function. Strength functions associated with damped excited modes such as the giant resonances may exhibit various fluctuations and fine structures arising from different mechanisms including the doorway structures of damping and the Porter-Thomas or GOE strength fluctuations associated with individual levels. Employing a model which mimics the doorway damping of the giant resonances, we applied the method to analyze the strength fluctuations associated with the doorway structures. The close relation between the profile of the local scaling dimension and the fluctuations embedded in the strength function is studied in detail. We showed that the behavior of the local scaling dimension $D_m(\epsilon)$ as a function of the binning energy scale $\epsilon$ manifests clearly both the GOE fluctuation at the small energy scale and the properties associated with the doorway states. For example, the binning energy scale $\epsilon^{**}$ where the local scaling dimension $D_m(\epsilon)$ decreases sharply corresponds to the spreading width $\gamma$ of the doorway states. This demonstrates that the local scaling dimension can be an useful measure to analyze the fluctuations and fine structures of the strength function associated with the damping modes in highly excited nuclei. It is in principle possible to apply the scaling analysis discussed in the paper to the experimental spectra and also to strength functions in realistic descriptions of the giant resonances in order to study the mechanism responsible for the strength fluctuations, such as the spreading width of the doorway states. The method itself is, on the other hand, quite general and not restricted to the giant resonances, and applicable to any strength fluctuation phenomenon in nuclei as well as in the other fields of quantum physics. \acknowledgments The authors acknowledge helpful discussions with Kenichi Matsuyanagi, Toru Suzuki, and Shoujirou Mizutori. The numerical calculations are performed at the Yukawa Institute Computer Facility. The work is supported by the Grant-in-Aid for Scientific Research from the Ministry of Education, Science and Culture (No. 10640267).
\section{Introduction} In \cite{BC}, we proposed a model for quantized discrete general relativity with a Euclidean signature. The model was constructed by combining the structure of a certain tensor category and of a full subcategory of it with the combinatorics of a triangulated 4-manifold. The category was the representations of ${\rm U}_q{\rm so}(4)$, for $q$ a $4n$-th root of unity and the subcategory the representations which are called balanced or simple. The main technical tool was the introduction of spin networks for these representations, which we called relativistic spin networks. The adjective relativistic is used because the relativistic spin networks are related to four-dimensional geometry whereas the original ${\rm SU}(2)$ spin networks of Penrose \cite P are related to three-dimensional geometry. The name anticipated the development of relativistic spin networks for the physically realistic case of the Lorentz group, ${\rm SO}(3,1)$. The wider context for the model is explained in \cite{BZ, BZ3}. It is closely related to topologically invariant models \cite{CKY, W3D, TV, PR, BQG}. The purpose of this paper is to supply the analogous concepts to our previous work for the Lorentz group and its $q$-deformation, ${\rm U}_q{\rm SL}(2,{\mathbb C})$. The geometrical description of the relativistic spin network evaluation for ${\rm SO}(3,1)$ is developed, in which hyperboloids in Minkowski space replace the role played by the sphere $S^3$ for the ${\rm SO}(4)$ case \cite {BC, B, FK}. We develop a correspondence between relativistic spin networks and the Lorentzian geometry of simplicies. One of the predictions we make is that the area of a timelike surface is quantized but the area of a spacelike surface can take any value in a continuous range. After removing a trivial infinite factor of the volume of the hyperbolic space, we obtain a expression for the evaluation of the relativistic spin network associated to a 4-simplex as a multiple integral on hyperbolic space. We conjecture that this integral is absolutely convergent. The study of the $q$-deformation of the Lorentzian case is not so far advanced, but we give some observations on this and the possible implications for Lorentzian signature state sum models. Before explaining our results, we give a brief review of the Euclidean case. \subsection {Review of the Euclidean model} The steps of the construction of the model in \cite{BC} were as follows: \smallskip \begin{enumerate} \item Describe the geometry of a discrete Riemannian triangulated 4-manifold by assigning bivectors to the 2-simplices satisfying appropriate constraints. \item Identify the bivectors with Lie algebra elements. \item Quantize the bivectors by replacing the Lie algebra with a sum over its representation category. \item Implement the constraints that the bivectors are simple bivectors by passage to a subcategory. \item Switch to the representations of a quantum group with $q$ a root of unity in order to create a finite model. \item Determine a quantum state for each tetrahedron. This is a morphism intertwining the four representations which are associated to the four boundary faces of the tetrahedron. \item Connect the representations and morphisms around the boundary of each 4-simplex into a closed diagram called a relativistic spin network. The diagram consists of one tetravalent vertex at the centre of each tetrahedon with edges corresponding the four faces of the tetrahedron. Edges corresponding to the common face of two adjacent tetrahedra are then joined. This gives a closed graph embedded in the boundary of the 4-simplex, $S^3$. The evaluation of this relativistic spin network determines the amplitude for a single 4-simplex. \item Multiply the amplitudes for each 4-simplex in the space-time manifold together, and then sum over the representation labels introduced to give a discrete version of a path integral. \end{enumerate} We introduced two variants of the model, given by different relativistic spin networks for the 4-simplex. One of them is based on a \fifteenj\ from representation theory (a complex number which is a function of 15 irreducible representations). The other is based on a new function of 10 irreducible representations which one might call the \tenj\footnote{despite the fact that it is not one of the traditional 3NJ--symbols}. Firstly we describe the model based on the \tenj. There is a uniquely determined morphism which satisfies all the constraints for the tetrahedron, which is taken to be a 4-valent vertex for the relativistic spin network. The vertex was introduced in \cite{BC} and generalised to other valencies in \cite Y. For the $q=1$ case, the vertex was proved to be the unique one satisfying the constraints for the geometry of a tetrahedron in \cite R, and its detailed properties were investigated in \cite{Barb2,BB}. This vertex can be regarded as the quantum state determined by quantizing the four bivectors determined by the four faces of a classical geometrical tetrahedron. Starting with this 4-valent vertex one obtains an amplitude for the 4-simplex which depends on the 10 representation labels at the 10 triangles. The irreducible representations in question are indexed by the integers, and the geometrical interpretation of these is that they are the areas of these 10 triangles in suitable units. The asymptotic properties of this amplitude are related to the Einstein-Hilbert Lagrangian in \cite {B},\cite{BWA}, and the classical solutions for a simplicial manifold are determined in \cite{BRW}. We refer to the expression formed from the tetravalent diagram on a 4-simplex as a \tenj, and the model constructed from them as the 10J model. For the 15J model one takes the relativistic spin network $$\DoubleY{}{}{}{}{b}$$ obtained by composing two trivalent vertices as the quantum state for the tetrahedron. This depends on the additional choice of a simple/balanced representation $b$ for the centre edge. The closed relativistic spin network for the 4-simplex is a trivalent graph called a \fifteenj. The asymptotic properties of the \fifteenj\ are related to the Einstein-Hilbert Lagrangian in \cite{CYA}. The principal difference to the 10J model is that the amplitude for the 4-simplex depends on 15 irreducible representations, not 10. For each tetrahedron, one has a representation on each triangular face, together with a fifth representation associated to the tetrahedron. To evaluate the \fifteenj\ one requires a choice of splitting the four faces into two pairs. The geometrical picture of the 15J model in terms of bivectors is a little more complicated than the 10J case, as there are now five bivectors for a tetrahedron instead of four. Four of the bivectors are now associated to the faces of a combinatorial tetrahedron, but they no longer satisfy the geometrical constraints for the bivectors of a tetrahedron. The constraints that these satisfy are that the bivectors for one pair of faces and the fifth bivector lie in a hyperplane $H_1\subset{\mathbb R}^4$, and the bivectors for the other pair of faces and the fifth bivector lie in a second hyperplane $H_2$. This is clear by carrying out the analogous geometrical analysis to that in \cite{B}; each vertex of a closed relativistic spin network is associated a unit vector $n\in{\mathbb R}^4$ and an edge is associated the bivector $* (n_1\wedge n_2)$. The fact that there are two hyperplanes for the tetrahedron and not one is not accidental; the variable $b$ is canonically conjugate to the angle between the two hyperplanes, so that if $b$ is specified precisely then the angle is necessarily indeterminate, due to the uncertainty principle. The trade-off between these two versions of the model is that introducing the extra representation on the tetrahedron increases the coupling between neighbouring 4-simplexes in a state sum model but destabilises the geometry of the 4-simplex. At present the best we can say is that the interpretation of the model is unclear at this point, but the clarity is neither better nor worse in its Lorentzian version. \subsection{Results in Lorentzian signature} The purpose of the present paper is to propose analogous models for the Lorentzian signature. As we will explain below, this essentially amounts to replacing ${\rm U}_q{\rm so}(4)$ with the noncompact quantum group ${\rm U}_q\sll(2,{\mathbb C})_{\mathbb R}$, (known in the literature as the Quantum Lorentz Algebra) whose representations have been studied in \cite{PUSZ, BR}. Some new phenomena occur due to the fact that the Lorentz group is not compact. For example, it seems no longer to be necessary to set the quantization parameter $q$ equal to a root of unity, instead, we conjecture that any real number will do. Unfortunately, the representation categories of noncompact Lie groups or quantum groups are more complex objects than the categories which have appeared in constructions of TQFTs or Euclidean general relativity. In order to formulate categorical state sums for these new categories, we will eventually (although not here), find it necessary to formulate the axioms for a new class of tensor categories to be called ``measured categories''. The new class of categories is a generalization of the theory of unitary representations of noncompact groups due to Mackey \cite{MA} and of the notion of fabrics due to Moussouris\cite{MOU}. For the purposes of this paper, the difference is that the irreducible representations are not discrete, but form a measure space. This leads to a natural generalization of a state sum which we call a state integral. It is not quite possible to give explicit formulae for our model with the present stage of knowledge of the representation theory of the Quantum Lorentz Algebra. The progress we have made is as follows \begin{enumerate}\item There is a natural definition of the \tenj\ for ${\rm SO}(3,1)$ which we conjecture to be finite. \item The integration measure for the state integral is a finite measure for the $q$-deformed case, the QLA. \end{enumerate} In the case of 1.~we present some arguments for this conjecture. Obviously one would like to conjecture that the \tenj s remain finite for the QLA. However we do not yet have the correct definition of the $q$-deformation. Also, the question of whether there is a good definition of the \fifteenj\ for ${\rm SO}(3,1)$ and for the QLA is not yet clear. The following is an outline of the rest of this paper: chapter 2 gives a review of the representation theory of ${\rm SO}(3,1)$. Chapter 3 explains the geometry of relativistic bivectors and its relationship to representation theory and quantization. Chapter 4 deals with the question of the vertex for the theory. Chapter 5 discusses the state sum model for the theory in the divergent case of the classical Lorentz group, giving details of the definition of the Lorentzian relativistic spin network evaluation. Chapter 6 recalls the basic facts about the representation theory of the Quantum Lorentz Algebra and proposes a $q$-deformed model with integrals with respect to a finite measure. Finally, chapter 7 proposes some natural directions for the investigation of the model. \section{The representation theory of ${\rm SO}(3,1)$} The purpose of this brief chapter is to acquaint the reader with the necessary facts about the representation theory of the Lorentz group and its Lie algebra, the Lorentz algebra \cite{6,7,8}. The history of the subject is actually rather strange. It was originally studied by Dirac, who thought that representations of the Lorentz group, which he called ``extensors'' \cite{Dirac} might play a role in physics. In the subsequent development of quantum field theory, the representations of the Poincar\'e group which were important were not the ones which came from the Lorentz group, and physical interest in the subject disappeared (at least until the present). On the other hand, the work of Gelfand and Naimark \cite{6,7,8} was extremely influential in mathematics. In this paper we are only interested in the unitary irreducible representations of the Lorentz algebra called the principal series. We use unitary representations in order to have a probability interpretation. Existing state sum models use the Plancherel measure in the summation over representations. In the Plancherel measure for the Lorentz group the irreducibles which are not in the principal series have measure zero and can therefore be neglected. The principal series representations are labelled by two parameters, one a half integer $k$, the other a continuous real parameter $p$; we can denote the representations $R(k, p)$. The two parameters are naturally combined into a single complex number, $w=k+ip$. There are no isomorphisms among them except that $R(k, p)= R(-k,-p)$. Each is irreducible, infinite dimensional and isomorphic to its dual. Any unitary representation of the Lorentz algebra can be written as a direct integral of irreducibles. The regular representation can be written as a direct integral involving only the principal series. Representations of the principal series are classified as fermions or bosons accordingly as $k$ is half integral or integral. Both determine representations of the covering group of the Lorentz group, which is the group ${\rm SL}(2,{\mathbb C})$ considered as a real Lie group. The bosonic representations give representations of the Lorentz group itself. The tensor product of any two members of the principal series is a direct integral of all the bosonic members if both of the representations are bosons or both fermions, and a direct integral of all the fermions otherwise. The Lorentz algebra can be thought of as $\sll(2,{\mathbb C})$ considered as a real Lie algebra, or as ${\rm so}(3,{\mathbb C})$ considered as a real Lie algebra. This corresponds to writing a general element of the Lorentz algebra as a sum of a rotation $J$ and a boost $K$. Then $J+iK$ is the corresponding element of ${\rm so}(3,{\mathbb C})$. There are two invariant inner products on this Lie algebra, $$\langle L,L\rangle ={\frac12} L_{ab}L^{ab} =J^2-K^2$$ and $$\langle L,*L\rangle ={\frac14} L_{ab}L^{cd}\epsilon^{abcd}=2J\cdot K$$ The corresponding Casimir elements in the Lie algebra have eigenvalues $$C_1=k^2-p^2-1$$ $$C_2=2kp$$ These are the real and imaginary parts of $w^2-1$. \section {Quantizing simple bivectors} The space of bivectors over ${\mathbb R}^4$ is a six dimensional real vector space. If we equip ${\mathbb R}^4$ with a Euclidean metric, then the bivectors have a natural identification with the Lie algebra ${\rm so}(4)$, which was necessary in the construction in \cite{BC}. As explained in \cite{B, BB}, the correct isomorphism is to take $b=*L$, or $$ b^{ab}={\frac12}\epsilon^{abcd}{L^e}_d g_{ec}$$ with $g_{ec}$ the Euclidean metric. If take $g$ to be a Lorentzian pseudometric on ${\mathbb R}^4$ (Minkowski space) instead of the Euclidean metric, then the bivectors are naturally identified with ${\rm so}(3,1)$ instead of ${\rm so}(4)$. Although changing the signature of the Lorentzian psuedometric would introduce a minus sign into the above identification, all bilinear conditions on the bivectors are unaffected. The bivector associated to a triangle is the wedge of two edges, a {\it simple} bivector. The condition that the bivector is simple is $$\langle b,*b\rangle=0,$$ the same equation as for the Euclidean case. In addition, we would like to differentiate spacelike, null and timelike simple bivectors, which correspond to planes in Minkowski space with an induced metric which is Euclidean, degenerate or Minkowskian. These are determined by the sign of $\langle b,b\rangle$. This is positive for spacelike bivectors, zero for null bivectors and negative for timelike bivectors. However the Hodge $*$ interchanges timelike and spacelike bivectors, so the corresponding Lie algebra element $L$ has $\langle L,L\rangle$ negative for spacelike bivectors, zero for null bivectors and positive for timelike bivectors. The idea behind the construction of this paper, as well as the model in \cite{BC}, is that having transformed all the variables for discretized GR into the form of constrained angular momenta (equals bivectors thought of as infinitesimal rotations) we quantize them the same way we quantize ordinary angular momenta, by replacing them with representations of the appropriate Lie algebra. Perhaps this is slightly obscured by the fact that in three dimensions bivectors are Hodge dual to vectors, so nonrelativistic angular momentum is written as a vector operator. In this program, it is now desirable to find an expression of the constraint that a bivector be simple translated into the category of unitary representations of the Lorentz algebra. The condition that the bivector is simple, $\langle b,*b\rangle=0$, translates into the vanishing of the corresponding Casimir $C_2=2kp$. Thus either $k=0$ or $p=0$. The quantization of $\langle L,L\rangle$ is then $C_1+1=k^2-p^2$. If $p=0$, then $\langle L,L\rangle$ is positive, and so the Lie algebra element $L$ is spacelike. This means that the bivector $b=*L$ is timelike. This leads us to propose the subcategory of representations $R(k,0)$ as the quantization of the timelike bivectors, and the subcategory generated by the $R(0,p)$ representations as a quantization of the spacelike ones. Since the Hodge $*$ interchanges timelike and spacelike bivectors, the space of simple timelike Lie algebra elements with a given square $\langle L,L\rangle$ is topologically the same as the corresponding space of spacelike elements. However the Poisson structures are different. Indeed, the spacelike simple Lie algebra elements include the subalgebra ${\rm su}(2)$, for which there is a quantization condition that the symplectic form is integral. This leads to a discrete series of representations. However the corresponding cohomology class for the timelike elements vanishes, so there is no quantization condition in this case and the parameter $p$ is arbitrary. There is another physical motivation for labelling the faces with the Hodge duals of their geometric bivectors. The approach we are exploring makes contact with the spin foam proposal of Rovelli and Reisenberger \cite{RR}. In that picture states for quantum gravity are described by embedded spin networks in 3d space. Their evolution is modelled by the world sheets of evolving spin networks which can change in time by developing vertices. On closer study, it seems that the kind of 2-complex they investigate is dual to a triangulation labelled as a term in our state sum. Although their approach has self-dual ${\rm su}(2)$ connections, the natural extension to ${\rm so}(4)$ connections would contain bivector operators which are canonically conjugate to the connection variables. As explained in \cite{BB}, the Poisson structure which comes in here is the one determined by the Einstein action, and this identifies (dual) Lie algebra elements with bivectors using the Hodge $*$. This contact with the kinematics of what is superficially a very different approach to quantizing gravity strengthens the specific proposal we have for the form of our state sum, which was suggested by the facts of the representation theory of the Lorentz group. \section{Vertices} In this section we consider the possible forms for the vertices for relativistic spin networks by considering the case $q=1$, in which the constructions can be described geometrically. Naimark \cite7 has shown that between any triple of representations of ${\rm SL}(2,{\mathbb C})$ satisfying the parity condition there is a unique (up to scalar multiple) intertwining operator. This means that, apart from normalisations, the trivalent vertex for the relativistic spin networks is defined. To obtain a picture for the tetrahedron, we consider the possible 4-valent vertices. In the case of ${\rm SO}(4)$, the requirement that each tetrahedron lies in a hyperplane (3-plane) translated into the constraint that the sum of any pair of bivectors on two faces of a tetrahedron add up to form another simple bivector. This is still true in the Lorentzian case. However it is not true that if the bivectors on the faces are spacelike then the sums of pairs will necessarily be spacelike as well. Likewise if the faces are timelike then the sum of any pair of bivectors need not be timelike. These facts are mirrored in the representation theory: tensor products of two even-spin $R(k,0)$ representations will contain copies of the $R(0,p)$ representations and vice versa. This will need to be taken into account below, when we decide exactly how to define a state sum model. Our preferred representations $R(k,0)$ and $R(0,p)$ can be realised in the spaces of square integrable functions on the hyperboloids in Minkowski space, ${\mathbb R}^4$ with inner product $x\cdot x=(x^0)^2-(x^1)^2-(x^2)^2-(x^3)^2$, using the Gelfand-Graev transform \cite{GG, GGV}. We consider the cases $Q_1$ given by $x\cdot x=1$, $x^0>0$, the positive null cone $Q_0$ given by $x\cdot x=0$, $x^0>0$ and the de-Sitter space $Q_{-1}$, given by $x\cdot x=-1$. In \cite{BC}, we noted that the simple/balanced representations of ${\rm SO}(4)$ could be realised in the space of functions on $S^3$. The relativistic spin network formalism in terms of functions on $S^3$ was developed in \cite{B, FK}. This can be directly generalised to the present case in a number of ways by replacing $S^3$ with either one of the hyperboloids $Q_1$, $Q_0$ or $Q_{-1}$. The representations of ${\rm SO}(3,1)$ have a Fourier decomposition as the following direct integrals and direct sums: $$ L^2(Q_1)\cong\bigoplus_p R(0,p) \dd \mu_p$$ $$ L^2(Q_0)\cong2\bigoplus_p R(0,p) \dd \mu_p$$ $$ L^2(Q_{-1})\cong\left(2\bigoplus_p R(0,p)\dd \mu_p\right)\bigoplus\left(\bigoplus_k R(k,0)\right). $$ The notation $\oplus \dd \mu_p$ indicates a direct integral, using the measure $\dd \mu_p=p^2\dd p$, which is the Plancherel measure restricted to $k=0$. An element of the direct integral is a vector function of $p$, $v_p\in R(0,p)$, with square $\int\|v_p\|^2\dd \mu_p$. The 2 indicates that the following term appears as a summand twice. This decomposition can be understood in the following way. The Casimir operator $C_1$ is the Laplacian on $Q_1$. So the irreducible representation $R(0,p)$ can be considered the space of solutions to the eigenvalue equation for the Laplacian with eigenvalue $-1-p^2$. The Casimir $C_1$ is also the wave operator on $Q_{-1}$ with eigenvalues either $-1-p^2$ or $k^2$. In this case one has solutions of the wave equation associated to both the time orientations, somewhat analogous to positive energy and negative energy solutions of the wave equation in Minkowski space. This gives the two copies of $R(0,p)$ in the Fourier decomposition of $L^2(Q_{-1})$\cite{V}. In addition there are the `tachyons', giving the $R(k,0)$. For the hyperboloid $Q_0$, one has a degenerate version of these formulae. One would like to have an intuitive understanding of these Fourier decompositions in terms of the quantization of bivectors. Bivectors in Minkowski space determine geodesics on the hyperboloids. For example, a spacelike simple bivector is Hodge dual to an oriented timelike plane through the origin in Minkowski space, which intersects $Q_1$ in an oriented geodesic. Following Mukunda \cite{M}, quantum mechanics on $Q_1$ has as a classical limit the motion of free particles on $Q_1$. These move along geodesics given by the Hamiltonian $H=P^2$ on the phase space $T^*Q$, where $P$ is the coordinate for the cotangent space. In fact, the constraint surface $\{H={\rm constant}\}$ decomposes into the orbit space under this flow, namely the space of geodesics on $Q_1$, or the space of simple spacelike bivectors in Minkowski space. Each timelike plane intersects $Q_0$ and $Q_{-1}$ in {\it two} timelike or null geodesics each, one future-directed and one past-directed. This explains the multiplicity 2 of the $R(0,p)$ in the Fourier decompositions of these spaces. Likewise, a timelike simple bivector is Hodge dual to a spacelike plane through the origin in Minkowski space, which intersects $Q_{-1}$ in exactly one spacelike geodesic. These planes do not intersect $Q_0$ or $Q_1$. Thus the $R(k,0)$ representations occur only in the Fourier decomposition of $Q_{-1}$. In this way, we understand the multiplicities in the above Fourier decompositions. The Gelfand-Graev transform gives precise formulae for the decompositions. The case of the three-dimensional hyperbolic space $Q_1$ is particularly important in the following, so we develop the formulae here. The other cases have analogous formulae. In general the representations $R(k,p)$ can be realised in the space of square-integrable sections of a line bundle over ${\mathbb C} P^1$. However for the $R(0,p)$ one can simplify this to give the representation in the space of homogeneous functions of degree $-1+ip$ on the light cone in Minkowski space.\footnote{Gelfand and coauthors use the parameter $\rho=2p$.} This means an element $f$ of this space satisfies $f(\lambda \xi)=\lambda^{-1+ip}f(\xi)$ for null vectors $\xi$ and for any real number $\lambda\ne0$. The inner product on this space is determined by the integral \begin{equation}\label{innerproduct} \int_\Gamma \bar f_1(\xi) f_2(\xi) \dd\xi \end{equation} over the two-sphere $\Gamma$ given by the null vectors satisfying $\xi^0=1$. The measure is the standard rotationally-invariant measure $\dd\xi$, normalised to total volume $1$. The Gelfand-Graev transform gives the function $$ \hat f(x)=\int_\Gamma f(\xi) (x\cdot\xi)^{-1-ip}\dd\xi$$ defined on $Q_1$. Our proposal for a $k$-valent relativistic spin network vertex for ${\rm SO}(3,1)$ is given by a map $R(0,p_1)\otimes\ldots\otimes R(0,p_k)\to{\mathbb C}$ by the formula $$ f_1\otimes f_2\otimes \ldots f_k\mapsto {\frac 1{2\pi^2}}\int_Q \hat f_1(x) \hat f_2(x) \dots \hat f_k(x)\dd x.$$ In this formula, $Q$ is one of the hyperboloids $Q_{\pm1}$, $Q_0$, and $\hat f$ denotes one of the representations of $f$ as a function on the hyperboloid, as given for example for the case of $Q_1$ by the preceding formula. The measure $\dd x$ is the standard Riemannian (or pseudo-Riemannian) volume measure on $Q$. The map is not defined on every element of the tensor product, and the important point is to understand it as a generalised function of the $p_i$ variables as well as the $\xi_i$. Alternatively the vertex can be defined by the formula for the kernel of the integral. For the case of $Q_1$ this is $$V(\xi_1,\xi_2,\ldots,\xi_k)={\frac1{2\pi^2}}\int_{Q_1} (x\cdot\xi_1)^{-1-ip_1}(x\cdot\xi_2)^{-1-ip_2}\ldots(x\cdot\xi_k)^{-1-ip_k} \dd x$$ Each of the representations for the different hyperboloids gives rise to a particular formula for the vertex (of arbitrary valence) for the relativistic spin networks. Using $Q_1$, there is a single vertex for relativistic spin networks labelled with $R(0,p)$. The uniqueness is due to the fact that each irreducible appears only once in the decomposition of $L^2(Q_1)$. Remarkably, this also implies a decomposition of the four-valent vertex \begin{equation}\label{yy} \vcenter{\FourX{a}{b}{c}{d}} =\int_p \dd \mu_p \vcenter{\DoubleY{a}{b}{c}{d}{p} } \end{equation} with the intermediate edge labelled with the $R(0,p)$. This formula is obtained by applying the Fourier decomposition to the intermediate product function $\hat f_1(x)\hat f_2(x)$. As this is also a function on $Q_1$, its decomposition only involves the $R(0,p)$. Precise formulae and a proof of this decomposition are given below in section \ref{evaluation}. This formula is a direct analogue of the corresponding formula for the Euclidean case which formed the original definition of the four-valent relativistic spin network vertex in \cite{BC}. The analogy is given in section \ref{comparison}. Our interpretation of this formula is of a tetrahedron which lies in a spacelike hypersurface. In this situation, the sum of the two bivectors for two faces is always again a simple spacelike bivector. This is reflected in the fact that the decomposition formula only requires intermediate representations of the form $R(0,p)$. Using $Q_{-1}$, there are a number of vertex formulae given by analogous integrals. If the representation on a free end is $R(0,p)$, then one has to specify whether the representation is to be realised in $L^2(Q_{-1})$ as the future- or past-directed solutions of the wave equation. These two transforms are given by $$ \hat f(x)=\int_\Gamma f(\xi)G^\pm(x,\xi)\dd\xi$$ with $$G^+(x,\xi)= \left\{ \matrix(x\cdot\xi)^{-1-ip}& x\cdot\xi>0\\ 0& x\cdot\xi<0\endmatrix \right. $$ and $G^-(x,\xi)=G^+(-x,\xi)$\footnote{This differs from the integral transform given in \cite{GGV}, which is valid for functions which are even under inversion.} Also, it is possible to put the $R(k,0)$ on the free ends, in only one way, using the formulae in \cite{GGV}. Our interpretation of these 4-valent vertices is that they represent tetrahedra which lie in a timelike (Minkowski signature) hypersurface in Minkowski space. Accordingly the faces of such a tetrahedron can be timelike, either future-pointing or past pointing, or spacelike. These possibilities correspond to the different possible vertex formulae. There is also a decomposition formula analogous to (\ref{yy}) for this case, which entails the use of two copies of $R(0,p)$ (the choice of $G^+$ or $G^-$) and the use of the $R(k,0)$ representations. This corresponds to the fact that in a Minkowski signature tetrahedron the sum of the bivectors for a pair of faces can be spacelike (future or past oriented) or timelike. Finally, the analogous formulae for $Q_0$ give null tetrahedra. \section{The state sum model} We shall now choose a geometric form for the model. Let us assume that the classical geometry is a triangulation of our manifold into 4-simplices all of whose boundary tetrahedra are spacelike. This means that all of the bivectors on the 2-simplices and all of the sums of bivectors of 2-simplices in the boundary of a common tetrahedron must be simple and spacelike. As discussed in the previous section, this is consistent with a categorical calculus in which the representations $R(0,p)$ only are used everywhere. This is the simplest form for a model, though we note that it may be interesting to investigate the other possibilities. For the 10J version of the model, a simplicial manifold is labelled with a value of $p$ on each triangle, and the weight for this state is the product of the symbols for each 4-simplex. In the 15J version of our model there is additionally a value of $p$ on each tetrahedron, and we use the product of the \fifteenj s. The model is formally the integral with the measure $\dd \mu_p$ on each variable in the simplicial manifold. Of course, a model constructed from representations of $\sll(2,{\mathbb C})$ would involve either an integral over all values of $p$ (or for other forms of the model a sum over all values of $k$). This would make it divergent. In \cite{BC} we solved the analogous problem by passing to a quantum group. We shall see in section \ref{qla} how this seems to work out in the Lorentzian case. In this section we state the form of the model for $q=1$. \subsection{Relativistic spin network evaluation}\label{evaluation} Firstly there is the question of how to evaluate the categorical diagrams we are to associate with the 4-simplices. Given a function $h\in L^2(Q_1)$ the $p$-th irreducible component is given by $$ f_p(\xi)={\frac1{2\pi^2}}\int_{Q_1}h(x)(x\cdot \xi)^{-1+ip}\dd x$$ giving a homogeneous function on the light cone, $\xi\cdot\xi=0$. The inverse transform is $$ h(x)=\int_0^\infty\dd\mu_p\int_\Gamma f_p(\xi) (x\cdot\xi)^{-1-ip}\dd\xi.$$ Composing these two transforms gives the delta function $\delta(x,y)$ on $Q_1$. However if one does the same calculation but does not integrate over $p$, the result is a projection operator onto the $p$-th irreducible component $h_p$ of the function $h$ on $Q_1$. The kernel of this operator can be explicitly evaluated. The projection is given by $$ h_p(x)={\frac1{2\pi^2}}\int_{Q_1}K_p(x,y)h(y)\dd y$$ where \begin{equation}\label{zonal} K_p(x,y)=\int_\Gamma (x\cdot\xi)^{-1-ip}(y\cdot\xi)^{-1+ip}\dd\xi =\frac{\sin pr}{p\sinh r}, \end{equation} $r$ being the hyperbolic distance (boost parameter) between $x$ and $y$. For a fixed $x$, the function of $y$ is called the zonal spherical function \cite{V}. It is the solution of the Helmholtz equation which is spherically symmetric about $x$. One can check explicitly that the delta function is regained by integrating over $p$ \begin{equation}\label{delta} \int_0^\infty K_p(x,y) \dd\mu_p=2\pi^2\delta(x,y) \end{equation} For a fixed $x$, $K_p(x,y)$ gives a continuous function of $y$ which is absolutely bounded by the value $1$ at $x$ and decays exponentially at infinity, but is not square integrable. The normalisation is in fact $$ {\frac1{2\pi^2}}\int_{Q_1} K_p(x,y) K_{p'}(y,z) \dd y = K_p(x,z)\frac{\delta(p-p')}{p^2},$$ expressing that fact that $K$ is a projection to the $p$-th Fourier component on $Q_1$. Here it is assumed that $p, p'>0$. Now we can give the rules for evaluating a relativistic spin network. The network is a graph with each edge labelled by a parameter $p$. The graph is allowed to have a boundary. This means that the edges do not have to end in vertices of the graph. The ends of the edges which do not meet vertices are called free ends, and the set of all these is the boundary of the graph. If there are no free ends the graph is called a closed graph. The naive evaluation associates one variable $\xi$ to each edge and is given by taking the product of one kernel $V(\xi_1,\ldots,\xi_k)$ for each vertex and integrating along the interior edges (those without free ends) using the inner product (\ref{innerproduct}) in the $\xi$ variables.\footnote{The hermitian inner product (\ref{innerproduct}) is written as a bilinear product by replacing $p$ with $-p$ in one of the two factors.} For the interior edges this procedure gives a factor of $K_p(x_1,x_2)$ for each edge. This is because performing the transform from functions on $Q_1$ to $R(0,p)$ and back to the space of functions on $Q_1$ at the next vertex is the same as inserting $K_p(x,y)$ for the edge. This can be illustrated by giving the promised proof of (\ref{yy}). This uses first equation (\ref{delta}), then (\ref{zonal}). \begin{multline*} \vcenter{\FourX{p_1}{p_3}{p_2}{p_4}} =V(\xi_1,\xi_2,\xi_3,\xi_4)\\ =\int_{Q_1} \frac {\dd x}{2\pi^2} \int_{Q_1}\frac {\dd y}{2\pi^2} \int_0^\infty \dd\mu_p\; K_p(x,y) (x\cdot\xi_1)^{-1-ip_1}(x\cdot\xi_2)^{-1-ip_2} (y\cdot\xi_3)^{-1-ip_3} (y\cdot\xi_4)^{-1-ip_4} \\ =\int_0^\infty \dd\mu_p \int_\Gamma \dd\xi\; V(\xi_1,\xi_2,\xi)V(\xi,\xi_3,\xi_4)\\ =\int_0^\infty \dd\mu_p \vcenter{\DoubleY {p_1}{p_3}{p_2}{p_4}p} \end{multline*} \subsection{The regularised evaluation for closed networks} For closed networks the naive evaluation formula can be re-expressed entirely in terms of the kernels $K$. In this section we give the definition of the evaluation for closed networks using these kernels as we shall use it in the rest of this paper. It will be necessary to modify the naive evaluation formula a little to obtain the actual definition of the evaluation which we are to use. Each vertex is associated a variable $x\in Q_1$. Each edge is associated the factor $K_p(x_1,x_2)$, the variables being the ones associated to the vertices at either end of the edge. The naive evaluation is then \begin{equation}\label{naive} \int_{{Q_1}^n} \prod K_{p(ij)}(x_i,x_j) \dd x_1 \dd x_2 \ldots \dd x_n. \end{equation} The evaluation of closed networks is a problem. Since the irreducible representations are infinite dimensional, we cannot trace on them. This is because the trace of the identity operator is infinite. In our integral definition this is reflected in the fact that the integral is invariant under $SO(3,1)$ and the orbits are not compact, so one has infinite factors for the volume of each orbit. There are two essentially equivalent ways of dealing with this. The definition we adopt is simply to remove the integration over the variable in $Q_1$ at one of the vertices in \eqref{naive}. Using the Lorentz invariance, it does not matter which variable is chosen for this. We take the variable $x_1$ to be fixed. \begin{defn} The regularised evaluation for closed networks is \begin{equation}\label{closed} \int_{{Q_1}^{n-1}} \prod K_{p(ij)}(x_i,x_j) \dd x_2 \ldots \dd x_n. \end{equation} \end{defn} Due to the Lorentz invariance, this integral is independent of the chosen value for $x_1$. This definition is similar to the suggestion in \cite D that we can open up any closed network by cutting one edge to give a network with two free ends and regard the network as an intertwiner from one irreducible to another. This gives a multiple of the identity operator. This multiple is then the evaluation of the original closed network. The evaluation is illustrated by several very simple cases. Firstly, a graph given by a single loop with one vertex on it $$\monogon p$$ has the evaluation $$ K_p(x,x)=1.$$ A graph with two vertices on a loop $$\bigon ab$$ has the evaluation $$ {\frac1{2\pi^2}} \int_{Q_1} K_a(x,y) K_{b}(y,x)\dd y =\frac{\delta(a-b)}{a^2}.$$ This elementary example illustrates very clearly that the relativistic spin network evaluation may be distributional in the spin parameters $a,b$. This means that for particular values of these parameters (here $a=b$) the evaluation may not have a finite value. Also we note that in the simpler case of a closed network with two vertices joined with just one edge, $$\unigon a$$ the evaluation is a divergent integral for all values of the spin parameter. $${\frac1{2\pi^2}} \int_{Q_1} K_a(x,y)\dd y =\infty.$$ The theta symbol with two vertices and three edges $$\thetagraph abc$$ has the evaluation \begin{multline*}{\frac1{2\pi^2}} \int_{Q_1} K_a(x,y) K_{b}(x,y) K_c(x,y)\dd y\\ =\frac2{\pi abc}\int_0^\infty \frac{\sin ar\sin br\sin cr}{\sinh r}\dd r\\ =\frac1{4 abc}\bigl(f(b+c-a)+f(c+a-b)+f(a+b-c)-f(a+b+c)\bigr)\end{multline*} where $$ f(k)=\frac2\pi\int_0^\infty\frac{\sin kr}{\sinh r}\dd r=\tanh(\frac\pi2 k).$$ For large values of $|k|$, $f(k)$ tends to $\pm1$. As a consequence, when the Euclidean triangle inequalities for $a$,$b$,$c$ are satisfied the value of the theta symbol approximates $1/(2abc)$. When the inequalities are violated exactly one of the four $f$ terms becomes negative, giving approximately zero for the evaluation of the theta symbol in this limit. In fact the value dies away exponentially fast as $a$,$b$,$c$ increase away from the critical values of a degenerate triangle, characteristic of quantum effects in a classically forbidden region of configurations. In terms of bivectors this agrees with the interpretation developed above. The simple spacelike bivectors correspond to planes in Minkowski space which lie in a common spacelike hypersurface in Minkowski space. Whenever these three bivectors add to zero then their magnitudes satisfy the Euclidean triangle inequalities. A similar analysis can be carried out for the graph with two vertices and four connecting edges. $$\fourtheta abcd$$ This graph is important because it gives a magnitude for the vertex for the tetrahedron. For this graph the evaluation is \begin{multline*}\frac1{4 abcd}\Bigl(g(b+c+d-a)+g(c+d+a-b)+g(d+a+b-c)+g(a+b+c-d)\\ -g(b+c-a-d)-g(c+a-b-d)-g(a+b-c-d)-g(a+b+c+d) \Bigr)\end{multline*} where $$ g(k)=\frac k2 \coth( \frac \pi2 k).$$ For large $k$, the function $g$ is asymptotically $|k|/2$. If one of four spins is bigger than the sum of the other three, say $d>a+b+c$, then the amplitude is asymptotically zero, just as for the theta symbol. However this condition occurs precisely when it is impossible to form a spacelike tetrahedron in Minkowski space with these numbers as the areas of the faces. This confirms our interpretation of the four-valent vertex as the quantization of a spacelike tetrahedron. For more complicated graphs it is a non-trivial task to determine whether our expression for the evaluation is finite. In the special case of a \tenj, we give arguments that the integral is finite and determines a function of the $p$. We do not at this point know a finite expression for a \fifteenj, or even a \sixj. If this is not just an artifact of our knowledge, it may be that only the 10J model survives in Lorentzian signature. The evaluation of relativistic spin networks is full of interesting open questions. \subsection{Finite \tenj s For ${\rm SL}(2,{\mathbb C})$}\label{finite} The \tenj\ is the evaluation of the relativistic spin network based on the complete graph on 5 vertices, all of whose edges are labelled with $R(0,p)$ representations of ${\rm SL}(2,{\mathbb C})$. \begin{conj} The regularised evaluation of the \tenj\ labelled with $R(0,p)$ representations is finite. \end{conj} {\bf Evidence for the conjecture:} This integral is similar in form to a Feynman integral on hyperbolic space. The kernel $K(x,y)$ is bounded as $x\to y$, so only infrared divergences need be considered. In other words, the only divergence possible would be at infinity in hyperbolic space. We therefore need to use two facts: the radial growth of the area of a large sphere in $Q_1$, and the asymptotic behavior at infinity of the kernel $K$. Using the fact that the area of the sphere is asymptotically $e^{2r}$, while the $K$ is asymptotically $e^{-r}$ it is possible to make an estimate for the integral expression (\ref{closed}). A delicate analysis shows that the expression is in fact absolutely convergent, even if the sine terms in $K$ are omitted and $K$ is approximated by $e^{-r}$ for large $r$. Thus the behaviour of the integral (\ref{closed}) as all variables go to infinity separately is dominated by the integral of a positive function, and apart from the trivial factors of $1/p$, is independent of the values of the spin labels $p$ defining the representations on the edges. In order to show that the integral is actually finite we need to consider also the regions where some vertices go to infinity and some do not. This corresponds to subdivergences in the standard Feynmanological language. We believe they do not appear in our case, but further analysis is needed. It is interesting to note that the analogous analysis for a \sixj\ gives an indeterminate expression and it is not clear if the corresponding integral converges. It would be interesting to look for a regularisation procedure to give finite values for all diagrams in this category. \subsection{Comparision with the Euclidean case}\label{comparison} In our original paper on relativistic spin networks, we used the representation theory of ${\rm SO}(4)$ rather than ${\rm SO}(3,1)$. We call this the Euclidean case. In this, we used irreducible unitary representations, which are finite dimensional. For these representations the spin network evaluations always exist as there is no difficulty taking the trace of a finite dimensional matrix. At first sight, the formalism for the infinite dimensional representations of the Lorentz group looks completely different. However the formulae we obtain have a strong similarity, and can say that the two evaluations are analogous but not the same. The irreducible representations that are used in the Euclidean case, the simple/balanced ones, are labelled by the spin $n$, an integer\footnote{This parameter is often given as a half-integer $j=n/2$.}. The following considerations show that the Lorentzian relativistic spin network evaluation can be considered as an extension of the Euclidean evaluation to the case where $n$ is a complex number of the form $-1+ip$, for $p\in{\mathbb R}$. In \cite{B} the relativistic spin network evaluation for ${\rm SO}(4)$ was written in terms of integrals over $S^3$. This space is analogous to the hyperboloid $Q_1$ in the present work. Each edge is labelled with an integer $n$, and the kernel for each edge in \cite{B} is $$K_E=\frac {\sin(n+1)\theta}{\sin\theta},$$ where $\theta$ is the distance between two points on $S^3$. On substituting $r=i\theta$ and $p=-i(n+1)$, the kernel $K$ for the Lorentz group becomes exactly $1/(n+1)$ times the formula for the Euclidean kernel $K_E$. This suggests that a better comparison between the magnitudes of Lorentzian and the Euclidean evaluations is to divide the later by a factor of $n+1$ for each edge. Indeed this is confirmed by the example of the loop with one vertex. In \cite{B} the evaluation of the loop is $(-1)^n(n+1)$, whereas here it is $1$. With the adjustment, the evaluations agree up to a phase factor. In general understanding the comparison of the phase factors would require a deeper investigation. For the theta symbol, the Euclidean evaluation gives $1$ if the three integer spin labels satisfy the triangle inequalities for a Euclidean triangle and sum to an even integer, and zero otherwise. For the Lorentzian evaluation, as computed above, the spin labels $a,b,c$ are continuous parameters and, after making the adjustment of multiplying by $abc$ (equivalent in magnitude to dividing the Euclidean formula by factors of $n+1$), the evaluation interpolates smoothly between asymptotic values of $1/2$ far into the interior of the region where the triangle inequalities are satisfied and $0$ far into the region where they are violated. The integration measure is analogous too, as the factor of $1/(2\pi^2)$ which has been included with each integration on the unit hyperboloid is the volume of the unit three-sphere; so the analogous measure on $S^3$ is normalised to total volume $1$. This is the measure used in \cite B. The decomposition formula (\ref{yy}) is directly analogous once one makes the correction of a factor of $n+1$ for the middle edge in the Euclidean evaluation. After this correction, the Euclidean formula would be to sum over the spins on the middle edge with weight $(n+1)^2$ (again ignoring phase factors). This is directly analogous to the measure $\dd \mu_p=p^2\dd p$ on substituting $n+1=ip$. \section{Passing to the Quantum Lorentz Algebra: a finite model?}\label{qla} The process of passing to the representation category of a quantum group should not be thought of merely as a clever regularization scheme for this family of models. Quantum groups fit very firmly into the program of noncommutative geometry. A quantum group is a noncommutative space with a symmetry structure like a Lie group. In fact, the noncompact quantum group which is referred to in the literature as the Quantum Lorentz Algebra (QLA) has a good $C^*$ algebra version \cite{BR,PW}. The set of irreducible representations of a $C^*$ algebra is the noncommutative analog of the set of points of a space. Thus, the construction of the model in this paper can be viewed as an exploration of a noncommutative version of general relativity. The approach of this paper allows us to interpret the discoveries about the representations of the QLA as a species of quantum geometry. In terms of classical physics, the $q$ deformation can also be interpreted as the introduction of a cosmological constant \cite{Kod,BZ2}. The representation theory of the QLA \cite{BR} is not as well understood as that of the classical Lorentz algebra. There are two possible forms of the QLA, one with the deformation parameter real and one where it is a complex phase. It seems that only the real case has been studied in the literature, so we shall attempt to use it in our construction. As in the classical case, the irreducible unitary representations are infinite dimensional and classified by two parameters, one discrete and one continuous. The first difference is that the continuous parameter is only allowed to take values in a bounded set of finite measure, which depends on the discrete parameter. The Plancherel measure (rather subtly defined) is given in \cite{BR} as $$\frac {h}{2\pi} (\cosh(2hk)-\cos(2hp))\dd p,$$ where $q=e^h$. This measure is defined over $p\in [-\pi/h, \pi /h]$ and $k\in Z$. However, since the values $(k,p)$ and $(-k,-p)$ are equivalent, the measure is integrated over one half of this region, a fundamental domain for this identification. For $k=0$, the measure reduces to $$\frac {h}{2 \pi}\left(1-\cos(2hp)\right)\dd p$$ over the interval $[0, \pi /h]$, which is a finite measure. Now we can see that the effect of passing to the QLA is to make the Plancherel measure restricted to the quantum version of the spacelike simple bivectors a finite measure. On the other hand, the sum corresponding to the timelike simple bivectors is not truncated. This is consistent with our choice of model integrating over the representations of the form $R(0,p)$, thinking of them as the Hodge duals of spacelike bivectors. (At present we are actually constrained to do this because the version of the QLA with a complex phase for $q$ is not studied. One could easily conjecture that passing to a root of unity would give a truncation in $k$. This is a subject for further study). The rest of the situation with respect to the representations of the QLA has not yet been fully clarified. The \threej s (trivalent vertices) have been defined only for a finite dimensional representation paired to a unitary one. On the other hand, the universal R matrix is known, and Buffenoir and Roche have announced a program to find the missing \threej s \cite{BR}. \section{Prospects} Clearly, the first item on the agenda in pursuing this model will be a careful study of the 3J, 6J 10J and \fifteenj s of the QLA. Once the definitions of these symbols are clear, a number of natural questions will arise. In the first place it will be interesting to see if asymptotic formulae for the relevant $q$-deformed symbols can be found as in the compact case. This will allow us to see if the argument recovering the Einstein Hilbert Lagrangian in the classical limit for the Euclidean signature case can be extended to Lorentzian signature, as it does in 2+1 dimensions \cite D. Secondly, it will be necessary to study the state integral carefully to find out if the integrand contains any singularities which are not integrable. Beyond this, there are many interesting questions we would like to study about the model. One would like to see whether small disturbances in initial conditions propagate causally. The model differs from other quantum gravity models in that it allows a continuous spectrum for the area of a spacelike surface. It might be interesting to see if this can generate any interesting predictions for black hole spectra. Farther down the line, we would like very much to know what the model does as we refine the triangulation on which it is based. Good behavior might tell us how to use the model to construct an actual theory of quantum general relativity.
\section{The Problem of the X/$\gamma$-Ray Low Luminosity of the Galactic Nucleus } Increasing evidences in support of the presence of a massive (2.5~10$^{\rm 6}$~M$_{\odot}$) Black Hole (BH) at the dynamical center of our Galaxy have been collected in the past years. Proper and radial motions of stars in the central parsec of the Galaxy, obtained with high resolution near-infrared observations \cite{ref1}, show indeed the presence of a dark mass with density $>$~10$^{\rm 12}$~M$_{\odot}$~pc$^{\rm -3}$ located within $<$~0.01~pc from Sgr~A$^{\rm *}$, and governing the dynamics of the mass of the region. The compact synchrotron radiosource Sgr~A$^{\rm *}$, which coincides with the Galaxy dynamical center and shows very low proper motion, is therefore considered the visible counterpart of the Galactic Nucleus (GN) BH, and its spectrum has been studied at all wavelengths (for recent reviews on the Galactic Center see \cite{ref2} \cite{ref3}). However, unlike stellar-mass BHs in binary systems and super massive BHs in AGNs, the Galactic Nucleus is found in the infrared and X-ray domains extremely faint and underluminous. In particular the results of the Galactic Center SIGMA/GRANAT Survey \cite{ref4} \cite{ref5} \cite{ref6} coupled to Rosat \cite{ref7}, ART-P \cite{ref8} and ASCA \cite{ref9} observations have shown that Sgr~A$^{\rm *}$ total X-ray luminosity is well below 10$^{\rm 37}$~erg~s$^{\rm -1}$, i.e. $<$~10$^{\rm -7}$ times the Eddington Luminosity for such a BH. This is rather intriguing since stellar winds of the close IRS~16 star cluster provide enough matter to power the accreting BH. Due to non uniformities in the winds the matter is accreted with substantial angular momentum and so the flow is certainly not in form of pure spherical free fall. In standard thin accretion disks around BH about 10$\%$ of the energy provided by the accreted matter ($\dot M$c$^{\rm 2}$) is expected to be radiated between infrared and X-ray frequencies. Estimates of accretion rate for Sgr~A$^{\rm *}$ are in the range $(6-200)~10^{\rm -6}$~M$_{\odot}$~yr$^{\rm -1}$ \cite{ref10} \cite{ref11} and luminosities around $10^{\rm 40}-10^{\rm 42}$~erg~s$^{\rm -1 }$ are therefore expected. New models of accretion flow have been recently proposed, in which very low radiation efficiency is obtained, even for non-spherical infall, by assuming that most of the energy is advected into the BH rather then being radiated. These ``advection dominated accretion flow'' models (ADAF) seem to be able to interpret the whole Sgr~A$^{\rm *}$ spectrum from radio to $\gamma$-rays, and in particular to explain the observed low X-ray flux \cite{ref12} \cite{ref11} \cite{ref13}. Sgr~A$^{\rm *}$ is in fact considered the test case for this set of models and observations in the X-ray domain of this object are crucial to establish the validity of the model assumptions and/or to constrain their parameters. \begin{figure} \centering \psfig{file=15_AGOLDWURM1_1.ps, width=12cm} \caption{ Images of the Galactic Center in galactic coordinates obtained from the 1990-1997 SIGMA Survey in 4 energy bands: 30-40~keV up-left, 40-75~keV up-right, 75-150~keV bottom-left, 150-300~keV bottom-right. Contours are in units of standard deviations, starting from 4.5~$\sigma$ with logarithmic steps of 1.4. Crosses indicate positions of SIGMA point sources detected within the central degree in different periods \cite{ref6} while the black square shows the position of Sgr~A$^{\rm *}$. }\label{fig1} \end{figure} \section{The 1990-1997 SIGMA/GRANAT Survey of the Galactic Center: 30-300~keV Upper Limits on Sgr~A$^{\rm *}$} The 30-1300~keV SIGMA telescope \cite{ref14} on the GRANAT satellite, observed the Galactic Center between March 1990 and October 1997 about twice a year, for a total of 9.2~10$^{\rm 6}$~s effective time. The telescope provided an unprecedented angular resolution ($15'$) at these energies and, for the quoted observing time, a typical 1$\sigma$ flux error of $<$ 2-3 mCrab (1 mCrab is about 8 $\times$ 10$^{-12}$ erg cm$^{-2}$ s$^{-1}$ in the 40-80 keV band). Analysis of a subset of these data already provided the most precise hard X-ray images of the Galactic Nucleus \cite{ref4} \cite{ref5} and proved that the Sgr~A$^{\rm *}$ luminosity in the 40-150 keV band is lower then 10$^{\rm 36}$~erg~s$^{\rm -1}$. These results were important because, though it was known from the Einstein Observatory data \cite{ref15}, that the GN was not bright in soft X-rays, it was still possible that Sgr~A$^{\rm *}$ could have, as any respectable BH in low state, a bright hard tail extending to $>$~100 keV and possibly also a component of 511~keV line emission. For example the close source 1E~1740.7-2942, weak at $<$~4~keV was later observed, in particular by SIGMA, to be bright in soft $\gamma$-rays, was found associated to radio jets and even to display transient events of 500~keV emission (see references in \cite{ref4}). Indeed 1E~1740.7-2942 is now recognized to be a good BH candidate at only 40$'$ from the GN and to be responsible for most of the hard X-ray/$\gamma$-ray activity observed by early low-angular-resolution experiments from the direction of the Galactic Center and previously associated to the GN \cite{ref16}. Recently Goldoni et al. (1999) \cite{ref6}, have re-analyzed the full set of SIGMA data to improve the upper limits estimation. In this analysis they included models with both point-source and diffuse emission to fit sky images. Reconstructed images of the central 4$^\circ$$\times$4$^\circ$ region in different energy bands are presented in Fig.~1. Four variable point-sources are contributing to the emission from the central 1$^\circ$ circle (1E~1740.7-2942, A~1742-294, GROJ~1744-28, GRS~1743-290). The last one is only $11'$ from Sgr~A$^{\rm *}$ and, if considered a single source, cannot be associated to Sgr~A$^{\rm *}$. However the presence of some weak contribution from the Galactic Nucleus cannot be excluded and therefore 2 sets of limits were derived for the Sgr~A$^{\rm *}$ hard X-ray flux: one for which all GRS~1743-290 emission is attributed to one single source (case A) and another obtained by including a source at the Sgr~A$^{\rm *}$ position (fixed parameter) in the fitting procedure (case B). Results are reported in Table~1, for the four energy bands of Fig.~1. At high energy no emission at all is detected apart from 1E~1740.7-2942 and values in the two colums are identical. \begin{table} \begin{center} \begin{tabular}{ccc} \multicolumn{3}{c}{}\\ \multicolumn{3}{c}{\large \bf SIGMA/GRANAT Upper Limits on Sgr~A$^{\rm *}$}\\ \multicolumn{3}{c}{}\\ \hline \noalign{\smallskip} {\rm Energy~range}&{\rm Integrated~luminosity~(A)}&{\rm Integrated~luminosity~(B)}\\ {\rm keV} & {\rm erg~s$^{\rm -1}$} & {\rm erg~s$^{\rm -1}$} \\ \noalign{\smallskip} \hline \noalign{\smallskip} $30-40~{\rm keV}$ & $2.6 \times 10^{35}$ & $4.5 \times 10^{35}$\\ \hline \noalign{\smallskip} $40-75~{\rm keV}$ & $2.0 \times 10^{35}$ & $3.4 \times 10^{35}$\\ \hline \noalign{\smallskip} $75-150~{\rm keV}$ & $2.0 \times 10^{35}$ & $2.4 \times 10^{35}$\\ \hline \noalign{\smallskip} $150-300~{\rm keV}$ & $5.2 \times 10^{35}$ & $5.2 \times 10^{35}$\\ \hline \noalign{\smallskip} \end{tabular} \vspace{3mm} \caption{SIGMA/GRANAT 2$\sigma$ upper limits for Sgr~A$^{\rm *}$ luminosity} \label{Table1} \end{center} \end{table} The most stringent upper limits of Table~1 (A column) have been reported in Fig.~2 in units of E~L$_{\rm E }$ for a distance of 8.5 kpc, and compared with the ADAF model predicted spectrum of Sgr~A$^{\rm *}$. The reported model is the ``best model'' as defined in \cite{ref11}, and refers to a BH mass of 2.5~10$^{\rm 6}$~M$_{\odot}$, a mass accretion in Eddington units ${\dot m}$~=~1.3~10$^{\rm -4}$ (i.e. 7~10$^{\rm -6}$~M$_{\odot}$~yr$^{\rm -1}$), a viscosity parameter $\alpha$~=~0.3, an equipartition parameter $\beta$~=~0.5 (exact equipartition between gas pressure and magnetic pressure), and a fraction of viscous heat converted in electron heat of $\delta$~=~0.001. Comparison with X-ray results, i.e. luminosities measured by Rosat (0.8-2.5~keV) and ASCA (2-10~keV), is also shown in Fig.~2. ASCA value is actually reported as upper limit, following Narayan et al. 1998 \cite{ref11}, because the observed flux was rather associated to the X-ray burster A~1742-289, located only at 1$'$ from Sgr~A$^{\rm *}$ \cite{ref9}. Note however that this interpretation is controversial and the ASCA measure may well contain some contribution from Sgr~A$^{\rm *}$ \cite{ref17}. \begin{figure} \centering \psfig{file=15_AGOLDWURM1_2.ps, width=12cm} \caption{ Sgr~A$^{\rm *}$ 30-300~keV upper limits in units of E~L$_{\rm E }$ (d=8.5 kpc), obtained from the 1990-1997 SIGMA/GRANAT Galactic Center Survey, compared to the predicted ADAF spectrum (for M$_{BH}$=2.5~10$^{\rm 6}$~M$_{\odot}$ and ${\dot m}$=1.3 10$^{\rm -4}$) \cite{ref11} (blue dashed-dotted line), along with the SIGMA sensitivity (red dashed line) scaled to the survey time of 9~10$^{\rm 6}$~s. Rosat Sgr~A$^{\rm *}$ flux and the ASCA upper limit are also reported. IBIS (black full lines) and JEM-X (green dotted line) broad-band sensitivities for the Galactic Nucleus and the INTEGRAL Galactic Center Deep Exposure time of 4~10$^{\rm 6}$~s are also reported (JEM-X sensitivity at 20~keV is lower then IBIS one due to smaller FOV). A typical thermal spectrum (kT$\approx$10~keV) of a neutron star LMXB (1E~1743.1-2864, \cite{ref8}) (light-blue dashed-double-dotted line) is also shown for comparison. }\label{fig2} \end{figure} \section{The IBIS/INTEGRAL Galactic Center Deep Survey} The Imager on Board the Integral Satellite (IBIS) is one of the two main instruments of INTEGRAL, the ESA $\gamma$-ray mission to be launched in 2001. It provides, thanks to its coded-aperture imaging system composed by a tungsten mask and two pixellated detector layers, fine imaging (12$'$ FWHM resolution), good spectral resolution ($<$~8~$\%$ at 100~keV and $\approx$~6~$\%$ at 1~MeV) and good sensitivity over a wide (20~keV-10~MeV) energy range and a very wide field of view (29$^\circ$$\times$29$^\circ$ at 0 sensitivity) \cite{ref18}. The INTEGRAL Core Program includes the so called Galactic Center Deep Exposure (GCDE) program, the deep observation of a central Milky Way region 60$^\circ$ wide in longitude and 20$^\circ$ in latitude. The GCDE will consist of a grid of 31$ \times $11 pointings separated by 2$^\circ$. The net GCDE observing time is 4.8 10$^{\rm 6}$~s per year, of which about 84~$\%$ performed in pointing mode \cite{ref19}. Considering the GCDE scan in pointing mode and the IBIS sensitivity over its FOV we estimated that the Galactic Nucleus will be actually observed by IBIS for an effective (on-axis equivalent) time of $\approx$~4~10$^{\rm 6}$~s in 3 years (see Fig.~3 for the IBIS/Sensitivity of the INTEGRAL GCDE scan). Adding slew and Galactic Plane Survey times the total net 3-years IBIS Core Program exposure on the GN will increase to 5~10$^{\rm 6}$~s. IBIS broad-band sensitivity is shown in Fig.~2 for the pointing GCDE time on the E~L$_{\rm E}$~vs.~E plot to compare it with the ADAF model spectrum for Sgr~A$^{\rm *}$ \cite{ref11}. Two curves for IBIS are represented for two extreme estimates of the in-flight background. In Fig.~2 is reported also the expected GCDE sensitivity for the INTEGRAL X-Ray monitor (JEM-X) which provides images in the range 3-60~keV but in a smaller field of view. This picture demonstrates that IBIS sensitivity estimated for the GCDE will allow either to detect the high energy emission predicted by ADAF from Sgr~A$^{\rm *}$ or to set tighter constraints on the model parameters. \begin{figure} \centering \psfig{file=15_AGOLDWURM1_3.ps, width=9cm} \caption{ IBIS Sensitivity for the INTEGRAL CGDE Survey, as fraction of on-axis IBIS sensitivity for the same exposure time. Contours are between 0.06 and 4.8 with steps of 0.06. Crosses indicate pointing scan of the GCDE \cite{ref19}. Maximum value (at l=0$^{\circ}$ b=0$^{\circ}$) is 0.51. }\label{fig3} \end{figure} \section{ Simulations of the Galactic Nucleus IBIS/INTEGRAL Observations} ASCA, ART-P and Rosat results have shown that at low energies lots of point sources and also diffuse emission are present in the area and can make identifications difficult. In particular activity of the X-ray burster A1742-289, only 1$'$ away from Sgr~A$^{\rm *}$, would not be easily separated by the GN one even by X-ray instruments like JEM-X. On the other hands results at higher energies, i.e. around 80~keV, should be rather ideal to reveal emission from the GN, since at these energies, due to their softer spectra, most of the Neutron Star binaries will be significantly fainter than the expected emission from the GN BH (see NS binary typical spectrum in Fig.~2). Even at energies $>$~50~keV, however, imaging will be crucial since SIGMA showed the presence of several high-energy sources located within 2$^\circ$ from the Galactic Nucleus (see Fig.~1). To prove the imaging capabilities of the IBIS telescope we performed simulations of a deep IBIS observation of the Galactic Center with the ISGRI low energy (20-700 keV) $\gamma$-ray detector layer of the telescope \cite{ref20}. Basic simulation procedure was described in \cite{ref21} \cite{ref22}, we used results of the SIGMA survey for source fluxes \cite{ref4} and the predicted Sgr~A$^{\rm *}$ flux from ADAF spectrum. Sky image reconstruction procedures are based on standard cross-correlation techniques and iterative analysis and removal of sources as described e.g. in \cite{ref23}. One simplified assumption was that the observation was performed as a single pointing rather then a sum of pointings along a grid as it is actually expected. This should not influence the results since the scanning will rather help to remove background unknown systematic effects, not presently included in the simulation. Work is in progress to include more realistic operational conditions. Fig.~4 shows some of the results of the simulations. These are zooms of the central part of reconstructed sky images obtained by an iterative decoding and point-source cleaning algorithm applied to the simulated detector images. A cluster of 4 high energy sources in the central 2$^\circ$ circle appears in the images (Fig.~4, left). Fine image analysis allow to position them and to remove their contribution (Fig.~4, right). Sgr~A$^{\rm *}$ is then detected at the expected position and signal to noise ratio of $\approx$~6~$\sigma$, corresponding to L$_{\rm 50-140~keV}$~$\approx$~3.5~10$^{34}$~erg~s$^{-1}$, the level predicted by the ADAF model. Including diffuse emission (possibly present at $<50$~keV energies \cite{ref9}) showed that its presence can make detection of faint sources more complicated and somemore refined procedures should be employed to model the composite diffuse plus point-sources emission. However diffuse emission will not influence the search for the GN emission at high energies. \begin{figure} \centering \psfig{figure=15_AGOLDWURM1_4.ps,width=12cm} \caption{ The central 3$^\circ$$\times$3$^\circ$ of the deconvolved and cleaned sky images in standard deviations (1 pixel $\sim~5'$), from a simulation of the IBIS/ISGRI GCDE observation (4~10$^{\rm 6}$~s) in the 50-140 keV band (logarithm scale from 1 to 200~$\sigma$ left, linear scale from 1 to 5~$\sigma$ right). Simulation included 11 sources in the IBIS FOV with fluxes from 4 to 70 mCrabs \cite{ref4}, a background of 150~cts~s$^{\rm -1}$ resulting in 1~$\sigma$ imaging error of 0.11 mCrab and a 0.6~mCrab point-source in Sgr~A$^{\rm *}$ (L$_{50-140~keV}$ $\approx$~3.5~10$^{34}$~erg~s$^{-1}$). After removal of the 4 brightest sources Sgr~A$^{\rm *}$ is detected at the simulated position and expected S/N level of $\approx$~6~$\sigma$. } \label{fig4} \end{figure} \section{Conclusions: XMM Core Program Observations of Sgr~A$^{\rm *}$} We have presented here the best available hard X-ray upper limits on Sgr~A$^{\rm *}$, obtained by the deep Galactic Center SIGMA/GRANAT survey. As shown above (Fig.~2), they do not constrain the present ADAF models, invoked to resolve the apparent contradiction between the presence of a massive black hole at the Galactic Center and its lack of activity in the X-ray domain. However we note that the ADAF model critically depends on mass accretion since L~$\propto$~${\dot m}^{\rm 2}$. The assumed value of accretion rate in the model (${\dot m}$=$1.3~10^{\rm -4}$ \cite{ref11}, actually determined by Rosat flux) is close to the lower limit of the estimated range of ${\dot m}$ (i.e. $(1-30)~10^{\rm -4}$, \cite{ref11} \cite{ref10}). Even a factor 3 higher in ${\dot m}$ would make the model not consistent with our limits, and also not compatible with other spectral data of Sgr~A$^{\rm *}$. New generation of telescopes aboard the future X-ray (AXAF/Chandra, XMM) and gamma-ray (INTEGRAL) missions, will allow to deeply search and study the GN high energy emission and in this way to test present models for massive BH accretion. We have shown with simulations that IBIS/INTEGRAL will be able to disentangle the hard X-ray emission of the Milky Way central square degree and to detect ADAF emission from Sgr~A$^{\rm *}$ or to set appropriate upper limits over the band 50-140~keV, by using 3 year data of the INTEGRAL/GCDE core program. In case of detection, the reconstructed flux will allow to test ADAF spectra and the radiation processes expected in such a massive BH. On the other hands the set of upper limits will imply an $\dot M$ at least a factor 1.4 lower then the present value of ADAF model, making the problem of low accretion rate even more difficult. The ADIOS models \cite{ref24}, in which advection is coupled to inflows/outflows of matter and which allow even lower emission then ADAF models for the same mass supply rate, may have then to be invoked. Alternatively different hypothesis on magnetic field will have to be included, e.g. allowing for lower values of the $\beta$ equipartition parameter in order to steepen the spectrum and reduce contribution at high energies \cite{ref11}. In this context hard X-ray results will be even more valuable if coupled to high resolution results at lower energies which could constrain the mass accretion rate and allow to study the spectral shape. XMM observations of the Galactic Center are planned in the Core Program of the first year of the mission operations. More than 5~10$^{\rm 4}$~s of the XMM Galactic Center scan will be devoted to observe the central half-degree of the Galaxy with the EPIC cameras. Such exposure time will provide sensitivities of the order of 10$^{32}$~erg~s$^{-1}$ ($5~\sigma$ for 8.5 kpc distance) in the energy range $2-10$~keV, with 6$''$ angular resolution (FWHM) and good spectral resolution. We expect that these observations will allow to resolve the confusion over the soft X-ray sources associated to Sgr~A$^{\rm *}$ \cite{ref17} and to obtain spectral data to test radiative properties of the closest supermassive accreting black hole. \section*{REFERENCES}
\section{Introduction} Turbulence is a ubiquitous feature of the interstellar medium although its precise nature is poorly understood and its role in the formation of stars is unclear. Theories of isolated star formation generally assert that gravitational collapse occurs onto a thermally supported core (e.g., Shu, Adams, \& Lizano 1987) and motions are quasi-static until very late times (Basu \& Mouschovias 1994; Ciolek \& Mouschovias 1995; Li 1998). Observations of infall at small scales ($\sim 0.01$~pc) in the isolated starless core, L1544, in Taurus are not in contradiction with such theories if it is indeed sufficiently close to forming a star (Williams et al. 1999), although large scale motions ($\simgt 0.1$~pc) do appear to require an alternative explanation (Tafalla et al. 1998), such as turbulent dissipation (Myers \& Lazarian 1998). The Serpens molecular cloud (d=310~pc; de Lara, Chavarria-K, \& Lopez-Molina 1991) has more embedded YSOs and is more turbulent than the Taurus cloud. In the northwest region of this cloud lies a cluster of Class 0 sources (Hurt \& Barsony 1996) which has been the subject of numerous studies in the millimeter and sub-millimeter regime (Casali, Eiroa, \& Duncan 1993; McMullin et al. 1994; White et al. 1995; Hurt, Barsony, \& Wootten 1996; Wolf-Chase et al. 1998; Testi \& Sargent 1998). Since the dense cores around Class 0 sources often show the spectral signature of inward motion (Mardones et al. 1997), we embarked on a study of the dense gas dynamics in this young cluster forming region in order to compare with more isolated star forming sites such as in Taurus. In this {\it Letter}, we present observations of a previously unrecognized core adjacent to the Class 0 source S68N that appears to be starless, contracting, and is highly turbulent. We compare its properties with its neighbor, S68N, deduce a chemical timescale for its formation that suggests that it is very young, and determine an average infall speed by spectral line modeling. This core, which we designate S68NW, demonstrates that turbulent motions in the ISM cannot be ignored in the formation of individual stars in clusters. \section{Observations} \label{sec:obs} A well tested method of diagnosing inward motions onto a star forming region is to search for self-absorbed lines where emission at low velocities is brighter than at high velocities (Leung \& Brown 1977; Walker et al. 1986; Zhou et al. 1993). We observed CS(2--1) and \mbox{N$_2$H$^+$}(1--0) toward the cluster of Class 0 sources since both lines are reasonably bright hence quick to map, both are strongly excited by gas of density $\mbox{$n_{\rm H_2}$}\sim 10^5$~\cc, and generally CS is optically thick and \mbox{N$_2$H$^+$}\ is optically thin at the resolution of these observations. Singledish maps were made at the Five College Radio Astronomy Observatory\footnotemark \footnotetext{FCRAO is supported in part by the National Science Foundation under grant AST9420159 and is operated with permission of the Metropolitan District Commission, Commonwealth of Massachusetts} (FCRAO) 14~m telescope in December 1996 using the QUARRY 15 beam array receiver and the FAAS backend consisting of 15 autocorrelation spectrometers with 1024 channels set to an effective resolution of 24~kHz (0.06~km/s). The observations were taken in frequency switching mode and, after folding, 3rd order baselines were subtracted. The pointing and focus were checked every 3 hours on nearby SiO maser sources. The FWHM of the telescope beam is $50''$, and a map covering $6'\times 8'$ was made at Nyquist ($25''$) spacing. Observations were subsequently made with the 10 antenna Berkeley-Illinois-Maryland array\footnotemark \footnotetext{Operated by the University of California at Berkeley, the University of Illinois, and the University of Maryland, with support from the National Science Foundation} (BIMA) for two 8 hour tracks in each line during April 1997 (CS) and October/November 1997 (\mbox{N$_2$H$^+$}). A two field mosaic was made with phase center $\alpha(2000)=18^{\rm h}29^{\rm m}47\hbox{$.\!\!^{\rm s}$} 5, ~\delta(2000)=01^\circ 15'51\farcs 4$ and a second slightly overlapping pointing at $\Delta\alpha=33\farcs 0, \Delta\delta=-91\farcs 0$. Amplitude and phase were calibrated using 4 minute observations of 1751+096 (4.4~Jy) interleaved with each 22 minute integration on source. The correlator was configured with two sets of 256 channels at a bandwidth of 12.5~MHz (0.15~\kms\ per channel) in each sideband and a total continuum bandwidth of 800~MHz. The flexible correlator setup allowed us to observe CH$_3$OH($2_1-1_1$) in addition to CS(2--1), and \mbox{C$^{34}$S}(2--1) along with \mbox{N$_2$H$^+$}(1--0): the methanol line was found to map the outflow associated with S68N (Wolf-Chase et al. 1998) but the \mbox{C$^{34}$S}\ line was detected only marginally. The data were calibrated and maps produced using standard procedures in the MIRIAD package. Since the emission is extended, analysis of the spectra must correct for the spatial filtering properties of the interferometer. To allow for this, we combined the FCRAO and BIMA data using the task IMMERGE. Maps were compared within the region of visibility overlap (6~m to 14~m) and small pointing corrections made to the FCRAO data ($< 6''$, about a tenth of the beam) which was then scaled using a gain of 43.7~Jy~K\e.\footnotemark \footnotetext{Information regarding aperture efficiency measurements on the FCRAO 14~m telescope can be found on the World Wide Web at http://donald.phast.umass.edu/$\sim$fcrao/library/techmemos/gain96.html} The resolution of the resulting maps was $10\farcs 0\times 7\farcs 8$ at p.a. $-72^\circ$ for CS which was observed twice in the compact C configuration and $8\farcs 5\times 4\farcs 6$ at p.a. $+2^\circ$ for \mbox{N$_2$H$^+$}\ which was observed once in C configuration and once in the wider B configuration. \section{Analysis} \label{sec:analysis} \subsection{S68N and S68NW} Analysis of the large scale maps is deferred to a later paper. Here, we restrict attention to a remarkable region, $\sim 1\farcm 5\times 2'$, around the S68N protostar. This source was discovered from earlier 3-element BIMA CS observations by McMullin et al. (1994) but we re-observed it with the 10-element array to obtain greater sensitivity and resolution. It was originally undetected in the 3~mm continuum but is readily apparent in the new data at an integrated flux level of 12.3~mJy, in agreement with OVRO observations by Testi \& Sargent (1998). S68N has also been detected at shorter wavelengths and its spectrum was fit by a modified blackbody with dust temperature 20~K and luminosity $5~L_\odot$ by Wolf-Chase et al. (1998). Maps of the integrated intensity of CS(2--1) and \mbox{N$_2$H$^+$}(1--0) around S68N are displayed in Fig.~1. The position of the 3~mm continuum peak, indicated by the star, lies at the center of the \mbox{N$_2$H$^+$}\ emission but is offset by $7"$ from the CS core. However, there is both high velocity emission from the outflow and self-absorption present in the CS spectra which may skew the map of integrated intensity relative to the distribution of dense gas around the star. Equally striking in the CS map, however, is the presence of a compact core, hereafter S68NW, that lies $\sim 50''$ west of S68N. It is also present in the map of \mbox{N$_2$H$^+$}\ integrated intensity but is not nearly so prominent. It was not detected in the continuum to a $3\sigma$ sensitivity of 3.3~mJy~beam\e, nor is it apparent in the slightly higher sensitivity OVRO Testi \& Sargent observations. It is also undetectable in maps at 1~mm (Casali et al. 1993; Tafalla \& Mardones, private communication), at 12, 25, 60, and 100~$\mu$m in the Hurt \& Barsony IRAS HIRES maps, in the near-infrared ($2~\mu$m; Eiroa \& Casali 1992) or in the Digital Sky Survey. These observations constrain the luminosity of any embedded object in S68NW to be less than $0.5~L_\odot$. \subsection{Abundance differences} Fig.~1 suggests a difference in the chemistry between the star forming S68N and starless S68NW core. We have estimated the abundance of \mbox{N$_2$H$^+$}\ in the two cores by comparing the mass of \mbox{N$_2$H$^+$}\ derived from the integrated emission with the virial mass derived from the size and linewidth. Given the compact appearance of the cores, the assumption of virialization is unlikely to be greatly in error. We define the boundaries of each core as the FWHM contour of the \mbox{N$_2$H$^+$}\ maps and calculate sizes, linewidths, and integrated emission within these limits. Core properties are listed in Table~1. The inferred virial \mbox{N$_2$H$^+$}\ abundance is much lower in S68NW than S68N. This is due to a combination of a smaller size, greater linewidth, and lower integrated intensity in S68NW, but the relatively low emission is the dominant factor. In the following section, infall model fits do not find such an extreme abundance difference between the two but nevertheless confirms that the \mbox{N$_2$H$^+$}\ abundance in S68NW is unusually low, $\sim 1\times 10^{-10}$, compared to $\sim 4\times 10^{-10}$ in other dense cores (Womack, Ziurys, \& Wyckoff 1992; Ungerechts et al. 1997). A potential explanation that suits the starless nature of S68NW is chemical evolution: Bergin et al. (1997) show that whereas CS forms very quickly in a dense core, it takes $\simgt 10^5$~yr to form substantial amounts of \mbox{N$_2$H$^+$}. Observations of other time-sensitive molecular species such as HC$_3$N offer a test of this hypothesis. \subsection{Spectral line modeling} The majority of the CS spectra in this region are double-peaked and the magnitude of the dips between the peaks tends to increase closer to the core centers. Average spectra within the FWHM contour of \mbox{N$_2$H$^+$}\ emission for each core are displayed in Fig.~2. Unlike the case of L1544 (Williams et al. 1999), the \mbox{N$_2$H$^+$}\ spectra are not self-absorbed and we use these data to determine the velocity and linewidth of the cores. For each core, the central dip in the CS spectrum lines up with the \mbox{N$_2$H$^+$}\ velocity indicating that the CS emission is self-absorbed. The S68N spectrum shows prominent outflow wings but it is quite symmetric in marked contrast to S68NW for which the lower velocity (blue) peak is much brighter than the higher velocity (red) peak. For a radially decreasing excitation gradient, such as would exist for a centrally condensed core at constant kinetic temperature, this indicates that the outer self-absorbing gas is red-shifted (i.e., infalling) relative to the inner emitting region. To estimate the speed of the infalling gas, we have fit the spectra using a simple two layer model consisting of two isothermal layers, the near side (to the observer) at low density and the far side at high density. This model resembles those discussed by Myers et al. (1996) and Williams et al. (1999): emission from the rear layer is absorbed by the lower excitation front layer with the location of the absorption dependent on the relative velocity between the front and rear layers (i.e., the infall speed). Observations are used to constrain the models as much as possible: gaussian fits to the isolated \mbox{N$_2$H$^+$}\ hyperfine component are used to set the systemic velocity and linewidth of the core, the line-of-sight widths of each layer are set equal to the measured radius of the cores, and the \mbox{N$_2$H$^+$}\ abundances are constrained to vary only within a factor of two of the virial estimates derived in the previous section. The free parameters are the densities of each layer, their common kinetic temperature, the molecular abundances, and the infall speed of the front layer onto the rear layer. In addition, a low optical depth component between the two layers was added to the S68N model to allow for the outflow, and a gaussian component $\sim 2$~\kms\ from line center was used in the S68NW model to fit excess emission at low velocities. The model spectra are shown in relation to the observations in Fig.~2 and model parameters listed in Table~2. The rear layer densities are very similar, approximately equal to the critical density of the two transitions, and the foreground layer densities are comparable to each other and similar to the density of $^{13}$CO emitting gas. The kinetic temperatures are the same for both cores and equal to the dust temperature of S68N as determined from the spectral energy distribution by Wolf-Chase et al. (1998). The CS abundances are also the same and consistent with observations of Orion by Ungerechts et al. (1997) but the fits require a smaller difference in \mbox{N$_2$H$^+$}\ abundance than the virial estimates derived in the previous section (note, however, that the \mbox{N$_2$H$^+$}\ abundance of S68NW is still very low). The parameter that is most different between the two cores is the infall speed. The low infall speed in S68N is implied by the near symmetry of the line profile and is little affected by the addition of the outflow component which is also quite symmetrical. The inferred infall speed for S68NW, however, is high because of the large blue-red asymmetry but its precise value is very sensitive to the strength of an additional gaussian component added at low velocities. This extra component appears to be physically unrelated to S68NW; it lacks a red counterpart and peaks in emission $\sim 2$~\kms\ from the \mbox{N$_2$H$^+$}\ line. Channel maps suggest that it is a second CS core, slightly offset along the line of sight. Its contribution to the integrated CS intensity within one linewidth of the central velocity of \mbox{N$_2$H$^+$}\ is less than 20\% at its peak and is generally much less in other spectra. The addition of this extra component reduces the blue-red ratio required in the infall model resulting in a smaller infall speed: in the absence of this component, the inferred infall speed is $\simgt 0.5$~\kms. Therefore, we believe that the value listed in Table~2, $0.34$~\kms, is a {\em lower limit} which implies that the S68NW core is contracting supersonically ($v_{\rm in}/\sigma_{\rm thermal}({\rm H}_2)>1$) although not necessarily super-Alfv\'{e}nically ($v_{\rm in}/\sigma_{\rm non-thermal}\simgt 0.5$). The implied mass infall rate of the front layer onto the rear layer is $1\times 10^{-7}~M_\odot$~yr\e. \section{Discussion} The data presented here indicate that S68NW is a turbulent core in the process of contraction and increasing its mass substantially. Furthermore its proximity to the S68N core and embedded protostar suggests that S68NW may soon form a low-mass star, as the next part of the sequence of low-mass star formation events which have occurred in the Serpens complex over the last Myr. If so, the formation of stars and cores in Serpens may substantially overlap in time, in contrast to the idea that star formation in clusters is coeval, such as in response to a single triggering event (e.g. Zinnecker, McCaughrean \& Wilking 1993). Instead these observations imply that the cloud is forming a core while its already formed cores are still forming stars. If so, core formation and star formation may have relatively similar timescales, each shorter than the overall cluster formation timescale. The instantaneous collapse timescale, $t_{\rm coll}=6200~{\rm AU}/0.34~{\rm km~s^{-1}}\simeq 10^5~{\rm yr}$ is approximately equal to the free-fall time for gas of density $n({\rm H_2})\sim 10^5$~cm\eee. Such dynamic motions are achieved in models of ambipolar diffusion only at very late times, $\sim 10^7$~yr (Basu \& Mouschovias 1994; Ciolek \& Mouschovias 1995; Li 1998), which appears to be inconsistent with the low \mbox{N$_2$H$^+$}\ abundance. In addition, the supersonic infall speed requires either low ionization levels, $x_e<10^{-8}$ at $n({\rm H_2})=3\times 10^4$~cm\eee\ (Basu \& Mouschovias 1995a), significantly less than measured in cores in Taurus (Williams et al. 1998) and Orion (Bergin et al. 1999), or a weak magnetic field, $B\simlt 10~\mu$G at $n({\rm H_2})=5\times 10^3$~cm\eee\ (Basu \& Mouschovias 1995b) which would imply a smaller Alfv\'{e}n speed, $\sim 0.2$~\kms, than the observed linewidth. Such a weak field at these relatively high densities may also conflict with HI Zeeman measurements of similar size fields at much lower gas densities in Ophiuchus (Goodman \& Heiles 1994). Finally, ambipolar diffusion models do not predict the large size scales over which infall occurs: asymmetric, self-absorbed CS line profiles extend well beyond the FWHM \mbox{N$_2$H$^+$}\ contour, indicating detectable inward motions over $\simgt 0.1$~pc. Similarly large infall zones have also been observed in the isolated core L1544 (Tafalla et al. 1998) and in the cluster forming regions, L1251B and NGC1333--IRAS4 (Mardones 1998). A possible explanation is that the collapse front propagates outward at the non-thermal, rather than the thermal, sound speed (e.g. Myers \& Fuller 1993) in which case the ratio of infall speed to effective sound speed remains comfortably within the bounds of ambipolar diffusion models. However, this requires that the non-thermal motions be maintained in the core over the ambipolar diffusion timescale, $t_{\rm AD}\simeq 10^6$~yr, but Nakano (1998) shows that the timescale for turbulent dissipation is approximately the same as the free fall time and much less than $t_{\rm AD}$ in the absence of any internal driving sources. Indeed, it may be the decay of the non-thermal motions, with a corresponding loss of pressure support, that drives the fast inward motions in S68NW (Myers \& Lazarian 1998) over the large observed size scales. Although we do not observe a decrease in the \mbox{N$_2$H$^+$}\ velocity dispersion toward S68NW, this does not preclude the existence of a small less turbulent central core. Observing such an object remains a challenge for the future. Note that in this scenario, the timescale of the next stage of star formation would be that of ambipolar diffusion if the core were magnetically subcritical, or would be dynamical if the core were magnetically supercritical. It will be useful to determine the incidence of cores like S68NW in other star-forming regions, and to study such cores in lines sensitive to a wide range of gas density. \acknowledgments This research was partially supported by NASA Origins of Solar Systems Program, grant NAGW-3401. JPW thanks the Radio Astronomy Laboratory at the University of California at Berkeley for support during the writing of the manuscript and Chris McKee and Frank Shu for informative discussions. Conversations with Shantanu Basu, Glenn Ciolek, and Zhi-Yun Li are also gratefully acknowledged.
\section{Introduction} Simple synchrotron models for the afterglow emission of cosmological Gamma-Ray Burst (GRB) sources imply an ambient gas density, $\sim 1~{\rm cm^{-3}}$, which is characteristic of the interstellar medium of galaxies (e.g., Waxman 1997a,b; Wijers \& Galama 1998). Indeed, direct imaging of the neighborhood of well-localized GRBs revealed faint host galaxies at cosmological distances in many cases (Bloom et al. 1999 and references therein). The detection of spectral signatures that can be associated with the GRB environment is of great interest both for distance measurements and for learning about the environment itself in which GRBs occur. A knowledge of the GRBs birthplace can help to constrain the validity of a given model for their formation. At the same time, it is also of fundamental interest to know how GRBs themselves affect their environments. The most popular models for GRB formation currently involve either the collapse of a single massive star (the so-called ``hypernova scenario'' [Woosley 1993, Paczy\'nski 1998; MacFadyen \& Woosley 1998]), or the coalescence of two compact objects such as two neutron stars or a neutron star and a black hole (Eichler et al. 1989; Narayan et al. 1992; Ruffert \& Janka 1998). These models could be constrained by knowing the GRB environment. Massive stars have very short lives, thus they will explode in star-forming regions, which are typically characterized by very dense environments. On the other hand, most merging neutron stars would be very old and would have typically traveled far from their birthplace. The scenario of compact merger progenitors could thus be suggested by medium to low density environments. We have previously shown (Perna \& Loeb 1998) that the X-ray and UV components of the afterglow radiation create an ionized bubble of radius $\sim 100$ pc $n_1^{-1/3}$ in the surrounding galaxy, where $n_1$ is the ambient density in units of $1~{\rm cm}^{-3}$. On a short timescale, as long as the afterglow radiation is still effective to ionize, the gradual ionization of the medium can produce time dependent absorption (Perna \& Loeb 1998; Meszaros \& Rees 1998) and emission lines (Ghisellini et al. 1998; B\"ottcher et al. 1998). In this paper we compute the emission spectrum which results as the ionized gas slowly cools and recombines. Cooling times are typically very long, $t_{\rm cool}\sim 10^5(T/10^5{\rm K})/(n_e/1~{\rm cm}^{-3})$ yr, at a temperature $T$ and an electron density, $n_e$. If GRBs occur in galaxies, then their rate is estimated to be $\sim (10^{6-7} f_b~{\rm yr})^{-1}$ (Wijers et al. 1997), where $f_b\leq 1$ is the unknown beaming factor (covering fraction) of the $\gamma$-ray emission. This implies that in every galaxy there is a non-negligible probability of finding an ionized GRB remnant at any given time. The identification of these remnants in nearby galaxies will allow a much closer study of the sites where GRBs occurred and will provide an estimate for the energy output and occurrence frequency of the events (Loeb \& Perna 1998). The hydrodynamic impact of a GRB blast wave on its environment lasts longer than the radiative ionization effect. It takes tens of millions of years for the GRB blast wave to slow down to a velocity of $\sim 10~{\rm km~s^{-1}}$, at which point it may be erased by interstellar turbulence. Hence, old GRB remnants should consist of a large-size ($\sim {\rm kpc}$), expanding, cold HI shell, similar to the HI supershells which were identified for two decades in nearby galaxies (Loeb \& Perna 1998, Efremov, Elmegreen, \& Hodge 1998; and references therein). However, it appears difficult to distinguish the old hydrodynamic remnants produced by GRBs from those produced by the accumulated effect of more conventional energy sources, such as multiple supernovae, stellar winds from OB associations, or impact from high-velocity clouds. In some cases, a class of these possibilities is disfavored. For example, recent deep CCD imaging of the HI holes in the Holmberg II galaxy (Rhode et al. 1999), did not reveal the anticipated optical emission from a normal stellar population in several of these holes, in conflict with the multiple supernova or stellar wind interpretations. Nevertheless, even in this case, the old age of the hydrodynamic remnants does not allow for a unique identification of GRBs as their energy sources. On the other hand, since the energy release in GRB remnants is impulsive, it should be easier to distinguish them from conventional sources by identifying their unique spectral signatures at a sufficiently early time when they are {\it young} and radiant. As we show later, the emission from young GRB remnants with ages of $\la 10^4~{\rm years}$ is affected mostly by the radiative ionization effect of the early GRB afterglow on its surrounding interstellar medium, since it takes much more than $10^4~{\rm years}$ for the non-relativistic blast wave to traverse the photoionized region. The impulsive energy release of hard ionizing radiation is unique to GRB sources and could distinguish young GRB remnants from the remnants of multiple supernovae. The goal of this paper is to identify the spectral signatures that are peculiar to GRB remnants and that can be distinguished from those due to the remnants of other explosive events, such as supernovae. In \S 2 we present the computational scheme adopted for this problem. In \S 3 we show our numerical results, and analyze the particular spectral signatures of young GRB remnants. In \S 4 we consider the expected effect of variations in the input parameters used in our calculations. Finally, \S 5 summarizes our main conclusions. \section{Model Assumptions and Computational Scheme} We consider a GRB source which turns on at time $t=0$ and illuminates a stationary ambient medium of uniform density $n$, with a time dependent luminosity per unit frequency, $L_\nu(t)$. The release in the surrounding medium of a large amount of ionizing radiation is a distinctive feature of GRBs and their afterglows, as opposed to supernova explosions, where any impulsive electromagnetic release would not escape promptly, but would be degraded by adiabatic expansion of the envelope before it could leak out. In our work, we limit our analysis to the effects of the afterglow photoionizing radiation on the medium. The blast wave lags behind the ionizing front, and until the time it reaches larger radii, from which most of the absorption and reemission comes, it is not expected to greatly affect the ionization state of the medium and the resulting luminosity. We will discuss this point in greater detail in \S3. Afterglows are most naturally explained by models in which the bursts are produced by relativistically expanding fireballs (Paczy\'nski \& Rhoads 1993; Meszaros \& Rees 1997; Vietri 1997a; Waxman 1997a,b; Wijers, Rees, \& Meszaros 1997; Vietri 1997b; Sari 1997). On encountering an external medium, the relativistic shell which emitted the initial GRB decelerates and converts its bulk kinetic energy to synchrotron radiation, giving rise to the afterglow. The combined radio and optical data imply that the fireball energy is $\sim 10^{51-52}$ erg. In the simplest unbeamed synchrotron model (e.g., Waxman 1997a,b), the time and frequency dependence of the afterglow luminosity is given by \begin{equation} L_\nu(t)=L_{\nu_{\rm m}}\left(\frac{\nu}{\nu_{\rm m}(t)}\right)^{-\alpha}\;, \label{eq:lum} \end{equation} where, \begin{equation} \nu_m(t)=1.7\times 10^{16} \left({\xi_e\over 0.2}\right)^2 \left({\xi_B\over 0.1}\right)^{1/2} E_{52}^{1/2}t_{\rm hr}^{-3/2}\;{\rm Hz}\;. \label{eq:num} \end{equation} Here $\xi_B$ and $\xi_e$ are the fractions of the equipartition energy in magnetic field and accelerated electrons, $E=10^{52} E_{52}~{\rm erg}$ is the fireball energy, $t_{\rm hr}\equiv (t/{\rm hr})$, and \begin{equation} L_{\nu_{\rm m}}= 8.65\times 10^{29} \sqrt{n_1}\left({\xi_{\rm B}\over 0.1}\right)E_{52}\; {\rm {erg~s^{-1}~Hz^{-1}}}, \label{eq:lumm} \end{equation} where $n_1$ is the ambient proton density in units of $1~{\rm cm^{-3}}$. The spectral index $\alpha$ is chosen to have the values $\alpha_1=-1/3$ for $\nu\le\nu_m$ and $\alpha_2=0.7$ for $\nu>\nu_m$, so as to match the temporal decay slope observed for GRB 970228 (Fruchter et al. 1998) and GRB 970508 (Galama et al. 1998). We consider a uniform medium which is initially neutral and in thermodynamic equilibrium, with a temperature $T\sim 10^4$ K, and include all the most important astrophysical elements, that is H, He, C, N, O, Ne, Mg, Si, S, Ar, Ca, Fe, Ni. Their abundances are taken from Anders \& Grevesse (1989). We consider a region surrounding the GRB site of size $R$ and medium density $n$, and we split it up into a radial grid with steps $\Delta r$. In propagating from a point at position $r$ to another point at position $r+\Delta r$, the afterglow flux is reduced according to \begin{equation} F_\nu(r+\Delta r,t+\Delta t) = F_\nu(r,t)\exp [-\Delta \tau_\nu(r,t)] \frac{r^2}{(r+\Delta r)^2}\;, \label{eq:flux} \end{equation} where $F_{\nu}$ is in units of ${\rm {erg~cm^{-2}~s^{-1}~Hz^{-1}}}$. We denote the local number densities of the ions of the various elements by $n_a^j(r,t)$, where the superscript $a$ characterizes the element and the subscript $j$ characterizes the ionization state. The optical depth due to photoabsorption within the distance $\Delta r$ is then given by \begin{equation} \Delta \tau_\nu(r,t)= \Delta r \sum_{a,j}n^a_j(r,t)\sigma^a_j(\nu)\;. \label{eq:tau} \end{equation} The photoionization cross sections are taken from Reilman \& Manson (1979). The abundances of the ions of the elements are determined by solving the system of equations \begin{equation} \frac{dn^a_j(r,t)}{dt}=q_{j-2}n^a_{j-2}+ q_{j-1}n^a_{j-1} + c_{j-1}n_{j-1}^a n_e -(q_j+c_j n_e +\alpha_j n_e)n_j^a +\alpha_{j+1}n_{j+1}^a n_e\;. \label{eq:dndt} \end{equation} The $q_j$ and $c_j$ are respectively the photoionization and collisional ionization coefficients of ion $j$, while $\alpha_j$ is the recombination coefficient. Note that $q_{j-2}$ refers to inner shell photoionization followed by Auger ionization. The collisional ionization rates are calculated according to Younger (1981). We compute the terms due to photoionization by integrating $F_\nu\sigma_\nu$ numerically. The recombination rates are given by the sum of the radiative and dielectronic recombination rates. The radiative recombination process is the inverse of photoionization, so the rates to the ground states are computed from the photoionization cross section with the help of the detailed balance relation. Hydrogenic rates are used for radiative recombination to excited levels. The dielectronic recombination rates are taken from Burgess (1966) with modifications to take more recent calculations into account. Most important is the reduction due to autoionization to excited states (Jacobs et al. 1977), with an appropriate treatment of the weakening of this effect at higher $Z$ (Smith et al. 1985). Since we are dealing with a non-equilibrium situation, the ionization fractions are calculated within the program. The emissivity of the medium, $E_\nu(r,t)$, is calculated by using an $X$-ray emission code developed by Raymond (1979), and modified with updated atomic rates, as in Cox \& Raymond (1979). This code computes the spectrum of radiation emitted by a hot, optically thin plasma. The basic processes which produce the continuum radiation are bremsstrahlung, recombination and two-photon continuum. Permitted-line radiation and the most important forbidden lines are also included, as well as the recombination-line radiation from H-- and He--like ions. Photoionization heating and radiative cooling are calculated within the same code, and used to update the temperature of the plasma as a function of position and time. Compton heating and cooling of the electrons by the radiation is also taken into account, as well as the secondary effect of the radiation emitted by the gas on the gas itself. This effect is especially important during the late phase of cooling. We start the simulation ($t=0$) at a position $R_{\rm min}\ll R$, and let the afterglow flux propagate and evolve according to equation~(\ref{eq:flux}), while calculating, at each position $r_i\le ct$ of the grid, the abundances of all the ions of each element, the temperature of the plasma, and the local emissivities $E_\nu(t,r_i)$. Let $t_{\rm obs}$ be the observer time, such that the radiation detected at $t_{\rm obs}=0$ corresponds to that emitted at $t=0$ in the source frame. Then a photon emitted at position $r$ at an angle $\theta$ with the line of sight will be detected by the observer at a time $t_{\rm obs}$ if it is emitted in the source frame at a time $t=t_{\rm obs}+{r\cos\theta}/{c}$. The total emitted radiation that reaches the observer at time $t_{\rm obs}$ is given by \begin{eqnarray} E_\nu^{\rm tot}(t_{\rm obs})&=& 2\pi\int_0^{R_{\rm max}}dr r^2\int_{-1}^{1} d\cos\theta \;E_\nu\left(r,t_{\rm obs}+\frac{r\cos\theta}{c}\right)\nonumber \\ &=& 2\pi c\int_0^{R_{\rm max}}dr r\int_{t_{\rm obs}-\frac{r}{c}}^{{t_{\rm obs} +\frac{r}{c}}} dt \; E_\nu(r,t)\;. \label{eq:emtot} \end{eqnarray} \section{Spectral Signatures of GRB Remnants} Figure 1 depicts the temperature profile of a GRB remnant at several times. In Figure 1a we consider the situation where a GRB of energy $E=10^{52}$ ergs occurs in a typical interstellar medium, for which we assume the density $n=1~{\rm cm^{-3}}$. Figure 1b shows the case of a burst of the same energy occurring in a dense cloud of density $n=10^2~{\rm cm^{-3}}$ and size $R=10$ pc. As the afterglow flux is proportional to $\sqrt{n}$ [cf. Eq.~ (\ref{eq:lumm})], the gas is heated to a higher temperature close to the source than in the lower density case. However it cools much faster than a lower density gas heated to the same temperature, because $t_{\rm cool}\propto n^{-1}$. In both figures, the bold line shows the ionized hydrogen fraction H$^+$/H$^0$ at the times immediately following the passage of the afterglow radiation through the various shells. Figure 2a and 2b show the emission spectrum above 13 eV at several times during cooling for the same set of parameters used in Figure 1. This ionizing flux is important for the luminosities and intensity ratios of the optical lines after the gas cools to around $10^4$ K. The time behavior of some of the most important lines in the observable regions of the spectrum is shown separately in Figure 3a and 3b, again with the same set of parameters as in Figures 1 and 2. Notice how the emission from the remnant is very weak in the first tens of years, and rapidly rises when $t_{\rm obs}\ga 300$ yr (particularly noticeable in panels (a) and (b) of Figure 3a). This is a result of the fact that the emission to the observer starts to come from the entire volume of the remnant only when $t_{\rm obs}$ becomes comparable to the light crossing time $R/c$. Figures 2 and 3 show that the energy from a remnant in a typical interstellar medium is mostly reemitted in the optical, UV and soft $X$-ray band. This is to be contrasted with the emission from a young supernova remnant, where the gas, heated by the shock to temperatures $\ga 10^7$K, produces a strong emission in harder regions of the $X$-ray band. In the high density case, the lifetime of the emission lines is shorter, [see Figure 3b], due to the rapid cooling of the dense gas. Here, for the high density case, we considered a cloud of size 10 pc (corresponding to a column density of $3\times 10^{21}$ cm$^{-2}$). Much higher column densities are not typically inferred in GRBs. In any event, for a burst which occurs in a bigger dense region, leaving a larger fraction of its energy in the surrounding medium, luminosities up to about two orders of magnitude higher than the ones shown could be observed in its remnant. Figures 4a and 4b show the ratios between some strong optical lines as a function of time. Some important diagnostic plots commonly used to distinguish among various excitation mechanisms (Baldwin, Phillips, \& Terlevich 1981; Baum, Heckman \& van Brugel 1992; Dopita \& Sutherland 1995) are shown in Figures 5a and 5b. The emission--line ratios exhibited by the nebulae reflect the mechanism by which the gas is ionized and the chemical abundances and physical conditions in the line-emitting gas. If the gas is purely photoionized, the ionization state of the gas is determined primarily by the ionization parameter, defined by $U=Q(H)/4\pi r^2 n_e c$, where $Q(H)$ is the number of ionizing photons per second emitted by the source, and to a lesser extent is affected by the shape of the ionizing flux. A remarkable feature of our diagnostic plots is the generally high value of the ratio between the [O III] $\lambda$5007 line and H$_\beta$. Numerical simulations (Shull \& McKee 1979) show that such high ratios (i.e. [O III] $\lambda$5007/H$_\beta \ge 5$) cannot be produced in shocks but are produced by photoionization models in which the ionization parameter is relatively high (i.e. $\sim 5\times 10^{-3}$ for [O III]/H$_\beta\simeq 10$; this ratio increases with ionization parameter). In our case, the ionization parameter is $\gg 1$ close to the source, and is higher than $\sim 10^{-3}$ for most of the ionized gas. At early times, because of time--delay effects, the bulk of the emission comes from the region close to the source, with very high values of the ionization parameter, and this leads to correspondingly high values of [O III] $\lambda$5007/H$_\beta$, not typically found in regions excited by other mechanisms. Note however that high values of [O III] $\lambda$5007/H$_\beta$ are occasionally observed in supernova remnant shocks, but only during a brief period of incomplete cooling (e.g. Raymond et al. 1998), or in oxygen rich supernova remnants (e.g. Morse et al. 1996). The ratio between [O III] $\lambda$5007 and [O III] $\lambda$4363 is a diagnostic of the temperature of the emitting plasma. Its increase with time is a signature of the fact that the gas is cooling. The temperature indicated is generally far higher than is observed in steady-state photoionized plasmas such as H II regions. It is even higher than is common in supernova remnants (Raymond et al. 1998) for much of the GRB cooling time. A third commonly used line ratio diagnostic for nebulae is the [S II]$\lambda$ 6717,27/H$_\alpha$ ratio, which is small in H II regions and planetary nebulae, and $\ga 0.4$ in most shocks. A GRB remnant shows the signatures of photoionization for most of its cooling time. A density-diagnostic line is the [O II] $\lambda$3727,29. An increase in the electron density generally leads to a weakening of this line. This can be seen in our case by comparing panels (d) of Figures 4a and 4b. Perhaps the most unusual feature of the optical emission is the high ratio of He II $\lambda$4686 to H$_\beta$. While this ratio is high for only a short time, the He II emission is extremely weak in H II regions and seldom exceeds 0.1 in supernova remnants. The strenght here results from the existence of a huge volume of gas at $10^5$ K or more, and the faintness of the Balmer lines at these temperatures. In our simulation of the impact of a GRB on the external medium, we have considered only the effects of photoionization. As a matter of fact, a shock front lags behind, and we need to estimate how it affects the GRB signatures that we discussed. Needless to say, the photoionization model is only valid until the blast wave produced by the GRB event reaches the photoionized material. As long as the shocked gas is very hot, however, it will have little effect on the optical spectrum. The shock compresses the gas, thereby increasing its emissivity, but it also heats the gas. This tends to increase the energy emitted, but to decrease the number of photons produced. The blast wave will strongly affect the optical spectrum when: (a) it has swept up a substantial fraction of the photoionized gas or, (b) when the blast wave becomes radiative, producing strong ionizing radiation and strong optical emission from the cooling gas. The former case occurs when the blast wave reaches 2/3 the radius of the photoionized region (thus reducing the volume of the optically emitting volume by 30\%). The latter occurs when the shock slows to about 300 $\rm km~s^{-1}$, depending upon the explosion energy only as the 1/11 power (Cox 1972). For an explosion with an energy $E_{52}\times 10^{52}$ ergs in a uniform medium of density $n_1$ cm$^{-3}$, the late phase of the blast wave evolution is described by the Sedov (1959) solution: $R\approx (19 {\rm pc})(E_{52}/n_1)^{1/5}t_4^{2/5}$, where $t_4$ is the time from the explosion in units of $10^4$ yr. The corresponding velocity of the wave is $v\approx $(750 km s$^{-1}$)$(E_{52}/n_1)^{1/5}t_4^{-3/5}$. For an explosion with $10^{52}$ ergs in a medium of density $n_1$, the blast wave will reach a distance of about 50 pc after $t\approx 10^6$ yr, while the shock reaches a velocity of 300 km sec$^{-1}$ after $t\approx 4.6\times 10^4$ yr. At that time, the shock has traveled a distance of about 35 pc. Let us consider the effect of the emission from the shock on a particularly important line, such as [O III] $\lambda$5007. From the simulations of Hartigan, Raymond \& Hartmann (1987), we see that the flux in the 5007 line from a shock at a velocity of 300 km sec$^{-1}$ in a medium of density 1 cm$^{-3}$ is $\sim 10^{-4}$ ergs cm$^{-2}$ s$^{-1}$. At a distance of 100 pc, this flux is $\sim 10^{-5}$ ergs cm$^{-2}$ s$^{-1}$ and it has to be compared with the flux from the same line due to the photoionized gas. This is on the order of $10^{38}/(4\pi r^2)\sim 10^{-4}$ ergs cm$^{-2}$ s$^{-1}$. Thus, the contribution of the shock to the emission when condition (b) is satisfied is only a few percent. The same analysis for the higher density case shows, instead, that after a time $\sim 10^4$ yr the optical spectrum becomes dominated by the emission from the shock. \section{Discussion} The model that we assumed for our GRB has a typical energy of $10^{52}$ ergs which is released isotropically, and the afterglow is produced in the standard fireball model. However, this might not always be the real scenario. A prompt optical-UV flash was detected for GRB990123 (Akerlof et al. 1999). If such a flash (coincident with the GRB and lasting for less than a minute) is generic in GRBs and carries as much energy as the gamma-ray emission (i.e. much more than the optical-UV afterglow emission), then GRBs might ionize a larger region than we previously considered. The photoionization signatures would be even stronger, and the luminosities higher. If, on the other hand, the optical-UV afterglow is beamed (and thus its energy lower than commonly estimated) then GRBs have a weaker effect on their environment and in this case it would be more difficult to distinguish them from other photoionized nebulae. A situation of non-steady state caused by photoionization, in a region where there is no evidence of a nearby photoionizing source, is however more generally typical of a GRB remnant. Unless the progenitor of a GRB is a massive star, one does not generally expect to find GRB remnants in star forming regions. On the other hand, photoionized nebulae are generally found around OB associations. Note that recently Wang (1999) has reported observations of $X$-ray emitting regions in M101 which did not show any evidence for OB associations, and has made the hypothesis that they could be associated with GRB remnants. Other complications could arise from a non-homogeneous medium. If the medium has dense clumps in it, then these will absorb more flux than the surrounding region; they will be more luminous but will cool faster. Depending on the pressure gradient at its boundary, a dense clump may expand (and cool adiabatically), or suffer additional compression. Small clumps are more likely to be heated to a comparable temperature with respect to the surrounding medium, and thus they will expand, due to the higher pressure caused by their higher density. On the other hand, a large dense clump will absorb a considerable amount of flux, and thus it will show a much steeper temperature gradient with respect to the lower-density surrounding medium. In this case the clump might expand in one direction and be compressed in another. The time-dependence of the luminosity as a function of time in these cases would show more complicated patterns. Modelling all these secondary effects is beyond our scope, especially because the real conditions of the medium are unknown, and the possible--early production of the afterglow and its degree of beaming are as yet a subject of debate. Also, it is still far from clear whether there is a unique scenario for all the bursts, or if instead there can be substantial differences from one burst to another. Our final discussion needs to address the issue of how many of these GRB remnants are detectable with current instruments. To this purpose, let us make some rough estimates of the possibility of detection of a strong line, say the [O III]$\lambda$5007, for example. Its luminosity is $\ga 10^{37}$ ergs s$^{-1}$ for a time $t_L\approx 4\times 10^4$ yr. The corresponding flux of photons at a distance $d_{\rm Mpc}\times$ Mpc is $F_{\rm signal}= 2\times 10^{-2} d_{\rm Mpc}^{-2}~{\rm cm^{-2}~s^{-1}}$. The number of background sky photons for a slit of 10 ${\rm \AA}$ around the 5007\AA~wavelength in observations from the ground, is $F_{\rm noise}\approx 10^{-5}$ cm$^{-2}$ s$^{-1}$ arcsec$^{-2}$(Roach \& Gordon 1973). For a telescope with a diameter $D=10$~m, a spectroscopic detection efficiency of $\epsilon=0.1$, and an integration time of $t_{\rm int}=10$ hr, the signal-to-noise ratio S/N=$N_{\rm signal}/\sqrt{N_{\rm noise}+N_{\rm signal}}$~obtains a value of $\sim 3.4 \times 10^5 d^{-2}_{\rm Mpc} (1+2\times 10^3 d_{\rm Mpc}^{-2})^{-1/2}$, where $N=\epsilon F (\pi D^2/4) t_{\rm int}$ is the number of photons detected if a 1'' resolution element is assumed. A signal-to-noise ratio S/N$\ge 10$ will thus correspond to a maximum detection distance of $d_{\rm max}\approx 200$ Mpc. This distance is an order of magnitude larger than the distance to the Virgo cluster of galaxies ($d_{\rm Virgo}\approx 16$Mpc) and is comparable to the distance of Coma cluster ($d_{\rm Coma}\approx 10^2$Mpc). Let us hence estimate the number of remnants that could be detected. For a population of GRBs that follow the star formation history, Wijers et al. (1997) estimated the local rate per galaxy to be $\Gamma_{\rm GRB}=2.5\times 10^{-8}{\rm yr}^{-1}$. Multiplying this number by $t_L$ yields a probability of about $10^{-3}$ of finding a young remnant per galaxy. The number of galaxies in the Virgo cluster up to a magnitude B=19 is about 2500 and the total number of bright ($L_\star$) galaxies out to a distance of 200 Mpc is $\sim 10^5$. Thus a few GRB remnants could be easily detected in the Virgo cluster, and about a hundred are detectable out to the limiting distance of 200 Mpc. We note that the above estimate is sensitive to the redshift distribution of GRBs; for example, if GRBs constitute a non-evolving population, the estimated GRB rate per galaxy is about 40 times higher (Fenimore \& Bloom 1995). If the $\gamma$--ray emission from GRBs is beamed to within a fraction $f_b$ of their sky, then the number of remnants in Virgo would {\it increase} as $\Gamma_{\rm GRB}\propto f_b^{-1}$ while the maximum number of remnants out to the limiting distance would {\it decrease} as $\Gamma_{\rm GRB} d_{\rm max}^3\propto f_b^{1/2}$. Hence, for a relatively modest beaming factor of $f_b\la 0.1$, there should be more remnants observable in Virgo than elsewhere. Moreover, Virgo remnants are easier to detect because they can be resolved, while distant remnants cannot be resolved and their flux could be easily dominated by contaminating light from their host galaxy. Based on these considerations, we conclude that an effective observational search should focus on identifying young GRB remnants in the Virgo cluster. A photo-ionized remnant of radius $\sim 100~{\rm pc}$ at a distance of $20~{\rm Mpc}$ occupies an angular diameter of $2^{\prime\prime}$ on the sky and could therefore be resolved. The strong emission lines from such a remnant can be detected with a signal-to-noise ratio of S/N$\approx 100$ after one hour of integration on the Keck telescope. Because of the temperature decrease at outer radii, we expect such a remnant to be center-filled in narrow-band imaging of high ionization lines (such as [O III] at 5007 \AA~or He II at 4686 \AA), and limb-brightened for low-ionization lines (such as [S II] at 6717 \AA). The non-relativistic blast wave does not reach the outer edge of young remnants, and might be visible in a deep exposure of a high resolution image. There might also be synchrotron emission in the radio band from the accelerated electrons in this shock. More interestingly, GRB remnants are expected to show ionization cones if the early UV afterglow emission from GRBs is beamed. The hydrodynamic spreading of the photo-ionized gas is negligible, as the gas expands at the sound speed of $\sim 10^2~{\rm km~s^{-1}}$ and can traverse only a distance of $\la 5~{\rm pc}~(t_L/5\times 10^4~{\rm yr})$ during the lifetime of the remnant. This distance is at least an order of magnitude smaller than the remnant radius, and so lateral expansion could smooth only extreme beaming factors of $f_b\la (0.1^2/4\pi)= 10^{-3}$. \section{Conclusions} We have computed the emission spectrum which results from the cooling and recombination of an interstellar medium whose equilibrium state has been altered by a GRB and its subsequent afterglow emission. We have identified some generic signatures which are quite likely to bear the footprints of a GRB, and whose close study in nearby galaxies can in turn give us direct information on the sites where GRBs typically occur, and, maybe, lead us to the discovery of the remnant (if there is one) of the object which triggered the initial burst. The $X$-ray emission is very weak compared to the UV and optical. This property could help separate GRBs from sources which provide a more steady energy supply, such as multiple supernovae or stellar winds; the latter type of sources tend to fill their remnants with hot X-ray emitting gas. We have found that the [O III]$\lambda$5007 to $H_\beta$ line ratio is indicative of the high values of the ionization parameter in GRB remnants (see Figures 4 and 5). Detection of this and similar generic lines (cf. Figure 3) at a signal-to-noise ratio S/N$=100$ is feasible for remnants in the Virgo cluster after an hour of integration time with the Keck 10 meter telescope. Narrow-band imaging of such remnants could resolve the shock from the photoionized region inside these remnants, and should reveal ionization cones if the early (prompt or afterglow) UV emission from GRBs is beamed. \acknowledgments We thank John Huchra for useful discussions. AL and RP were supported in part by NASA grants NAG 5-7039 and NAG 5-7768. JR was supported in part by NASA grant NAG 5-2845.
\section{Introduction.} The Kerr metric plays a very prominent r\^{o}le in Einstein's theory of general relativity. Obtained by R.Kerr in 1963 \cite{Kerr}, it was the first explicitly known stationary rotating (i.e. non-static) asymptotically flat vacuum spacetime. Although many other stationary and axially symmetric asymptotically flat vacuum solutions are known at present, the Kerr metric remains, somehow, the simplest one. More importantly, the Kerr spacetime has a very special status among stationary vacuum solutions due to the black hole uniqueness theorem, which states, roughly speaking, that the exterior geometry of a stationary, asymptotically flat, vacuum black hole must be Kerr. Despite its importance, the reasons why this geometry plays such a privileged r\^{o}le remain somewhat obscure. In comparison, the Schwarzschild metric has nice uniqueness properties; in addition to Birkhoff's theorem, this metric is well-known to be the only static and asymptotically flat vacuum solution such that the induced metric of the hypersurfaces orthogonal to the static Killing vector are conformally flat. This leads one to consider whether there exists any property of the Kerr metric which singles it out among the class of stationary, asymptotically flat vacuum metrics. Some characterizations are already known. The first such was found by B. Carter \cite{Carter1} who analyzed stationary and axisymmetric electrovacuum spacetimes such that the Hamilton-Jacobi and Schr\"odinger equations are separable in some adapted coordinates and found a finite parameter family of solutions containing the Kerr metric. However, this analysis relies on coordinate dependent conditions, which is not appropriate for characterizing a spacetime. Furthermore, axial symmetry is assumed from the outset. A second characterization was obtained by Z. Perj\'{e}s \cite{Perjes1}, who restricted the analysis to the strictly stationary subclass (i.e. to spacetimes with a Killing vector which is timelike everywhere). Working on the manifold of trajectories (i.e. the set of orbits of the isometry group, which is assumed to be a manifold), Perj\'es showed that the Kerr metric can be found uniquely by demanding the existence of ``geodesic eigenrays'' of the Killing vector. The third and, probably the most interesting of the known characterizations of Kerr is due to W. Simon \cite{Simon}, who defined a tensor (now called Simon tensor) on the manifold of trajectories which is identically zero for the Kerr metric. Conversely, if the Simon tensor vanishes in an asymptotically flat, vacuum spacetime, then there exists an open neighbourhood of infinity which can be isometrically embedded into the Kerr spacetime. Moreover, the Simon tensor is equivalent to the Cotton tensor (which vanishes in a three-dimensional Riemannian manifold if and only if the metric is locally conformally flat) when the Killing vector is hypersurface orthogonal, so that one of the known characterizations of Schwarzschild is recovered. Despite the clear interest of this characterization it still suffers from two drawbacks. The first one is that the theorem establishes the isometry with Kerr only in some neighbourhood of infinity. Even though any strictly stationary vacuum metric is analytic \cite{mzHaagen}, this does not ensure that the isometry near infinity extends everywhere, because analytic extensions of manifolds need not be unique. However, this is probably not a serious objection and could presumably be fixed by exploiting the interesting results by Z.Perj\'es in \cite{Perjes2}, who found all the possible local forms of the metric in any strictly stationary spacetime with vanishing Simon tensor. In particular, Perj\'es found that the most general metric with those properties depends only on a few parameters, thus showing that the asymptotically flatness condition in the characterization of Kerr is only necessary in order to fix the value of some constants. The second objection is more serious and will be the object of this paper. Since the whole construction of the Simon tensor is done on the manifold of trajectories, the characterization in terms of the Simon tensor only works in the strictly stationary region, i.e. outside any ergosphere. However, for most practical purposes (in particular, for the black hole uniqueness theorems) the relevant region of Kerr is the domain of outer communication, which contains a non-empty ergoregion. Therefore the characterization in terms of objects defined in the manifold of trajectories is not sufficient. Thus, it is necessary to find a property of Kerr that distinguishes this metric irrespectively of the norm of the Killing vector. In order to treat this problem, we should look for a characterization involving only spacetime objects. Due to the relevance of the Simon tensor, the obvious strategy is to try and obtain a spacetime version of the Simon tensor in the region where the Killing vector is timelike and analyze whether or not this spacetime tensor remains regular at points where the Killing vector becomes null. Provided this is the case, it would remain to show that the extended Simon tensor vanishes everywhere for Kerr and prove also that the converse holds, i.e. that an asymptotically flat, vacuum spacetime with vanishing spacetime Simon tensor is locally isometric to Kerr (we cannot expect a global isometry to exist without further global conditions). This is the problem we want to solve in this paper. More specifically, in section 2 we recall the definition of the Simon tensor on the manifold of trajectories and write down the main result proven in \cite{Simon}. In section 3, we obtain the spacetime counterpart of the Simon tensor in the strictly stationary region and show that it remains regular at points where the Killing vector becomes null. This requires some analysis of the algebraic properties of the Simon tensor. Then, we define a tensor on any spacetime with a Killing vector (with no assumptions on its norm) such that it coincides with the Simon tensor when projected down into the manifold of trajectories in the strictly stationary region, and we obtain the consequences of the vanishing of this tensor in terms of the Weyl tensor. In section 4, we prove that an asymptotically flat vacuum spacetime with a Killing field which is timelike near infinity and such that the corresponding spacetime Simon tensor vanishes, is locally isometric to Kerr at every point, thus extending the results in \cite{Simon}. It is worth emphasizing that obtaining a spacetime characterization of Kerr is interesting not only in order to understand better what is so special about Kerr, but also for more practical purposes. The problem we have in mind is the black hole uniqueness theorem. The existing proofs for this theorem require rather strong hypotheses on the spacetime, like connectedness of the black hole event horizon (see Weinstein \cite{We1}, \cite{We2} for interesting progress in the non-connected case), non-existence of closed timelike curves in the domain of outer communication (see Carter \cite{Carter3} for a discussion) and analyticity of the metric and the event horizon (see Chru\'sciel \cite{Ch1}, \cite{Ch3}). This last point is crucial in order to apply the so-called Hawking rigidity theorem \cite{HE}, which ensures the existence of a second Killing field when the black hole is rotating. It is reasonable to believe that the theorem still holds when some (or perhaps all) of these hypotheses are significantly relaxed. However, trying to prove those results is a hard problem, and it is conceivable that a more detailed knowledge of the Kerr geometry, and in particular having at hand a spacetime characterization of Kerr can prove helpful for attacking those questions. In this respect, let us mention that the characterization of Schwarzschild in terms of the conformal flatness of the hypersurfaces orthogonal to the static Killing vector is an essential tool in all known proofs of the black hole uniqueness theorems in the non-rotating case. \section{The Simon tensor on the manifold of trajectories} Let us start by fixing our definitions and conventions. Throughout this paper a $C^n$ spacetime denotes a paracompact, Hausdorff, connected, $C^{n+1}$ four-dimensional manifold endowed with a $C^n$ Lorentzian metric $g$ of signature $(-1,1,1,1)$. Smooth means $C^{\infty}$. We will also assume that spacetimes are orientable and time-orientable. The Levi-Civita covariant derivative of $g$ is denoted by $\nabla$, the volume form is $\eta_{\alpha\beta\gamma\mu}$ and our sign conventions of the Riemann and Ricci tensors follow \cite{KSMH}. Throughout this section $({\cal M},g)$ denotes a $C^2$ spacetime satisfying Einstein's vacuum field equations $R_{\alpha\beta}=0$ and admitting a Killing vector field $\vec{\xi}$ which is timelike everywhere. The construction of the Simon tensor is as follows. First, an equivalence relation $\approx$ is introduced in ${\cal M}$ so that two points $p,q \in {\cal M}$ are equivalent if and only if they belong to the same integral line of $\vec{\xi}$. If the spacetime satisfies the chronology condition (see \cite{HE} for the definition) and the Killing field is complete, then the quotient set ${\cal N} = {\cal M} /\approx$ is a differentiable manifold \cite{mantraj} such that the canonical projection $\pi : {\cal M} \rightarrow {\cal N}$ is differentiable. Then, there exists a one-to-one correspondence between tensors on ${\cal N}$ and tensors on ${\cal M}$ which are completely orthogonal to $\vec{\xi}$ (i.e. orthogonal to $\vec{\xi}$ with respect to any index) and with vanishing Lie derivative along $\vec{\xi}$. Then, the so-called norm and twist of the Killing field are defined by $\lambda= - \xi^{\alpha} \xi_{\alpha}$ and $\omega_{\alpha} = \eta_{\alpha\beta\gamma\delta} \xi^{\beta} \nabla^{\gamma} \xi^{\delta}$. Since $\lambda >0$ in ${\cal M}$, the tensors $h_{\alpha\beta} = g_{\alpha\beta} - \lambda^{-1} \xi_{\alpha}\xi_{\beta}$ and $\gamma_{\alpha\beta} = \lambda h_{\alpha\beta}$ are well-defined on ${\cal M}$. The corresponding tensors on ${\cal N}$ are denoted by $\lambda$, $\omega_{i}$, $h_{ij}$ and $\gamma_{ij}$ respectively (tensors on ${\cal N}$ carry Latin indices). Both $h_{ij}$ and $\gamma_{ij}$ are symmetric, non-degenerate and positive definite. Denoting by $D$ the Levi-Civita covariant derivative of $({\cal N}, \gamma )$ and introducing the complex one form \begin{eqnarray} \sigma_{j} = D_{j} \lambda - i ~ \omega_{j}, \label{sigmai} \end{eqnarray} the Simon tensor associated with $\vec{\xi}$ is defined\footnote{This definition differs by a non-zero factor from the original one in \cite{Simon}.} via \cite{Simon} \begin{eqnarray*} S_{ijk} = 2 \left (\sigma_{\left [ k \right.} D_{\left . j \right ]} \sigma_{i} - u_{\left [ k \right .} \gamma_{\left . j \right ] i } \right ), \hspace{1cm} u_k = \gamma^{ij} \sigma_{\left [k\right.} D_{\left . j \right ]} \sigma_i. \end{eqnarray*} For later convenience, let us introduce a tensor $\hat{Z}_{ij} = D_{j} \sigma_i$ (which is symmetric by virtue of the vacuum field equations) so that the Simon tensor reads \begin{eqnarray*} S_{ijk} = \hat{Z}_{ij} \sigma_k - \hat{Z}_{ik} \sigma_j - \frac{1}{2} \gamma_{ij} \left (\hat{Z} \sigma_k - \sigma^{l} \hat{Z}_{lk} \right ) + \frac{1}{2} \gamma_{ik} \left (\hat{Z} \sigma_j - \sigma^{l} \hat{Z}_{lj} \right), \end{eqnarray*} where $\hat{Z} = \gamma^{ij} \hat{Z}_{ij}$ and indices are raised with $\gamma^{ij}$. One of the main results proven in \cite{Simon} is (the concept of asymptotically flat end is defined later in section 4) \begin{theorem}{(Simon 1984)} Let $\left ({\cal M},g \right)$ be a $C^2$ vacuum spacetime with a timelike Killing field $\vec{\xi}$. Assume that the spacetime contains an asymptotically flat end ${\cal M}^{\infty}$. If the Simon tensor associated with $\vec{\xi}$ vanishes identically, then there exists an asymptotically flat open submanifold ${\cal M}_1 \subset {\cal M}^{\infty}$ which is isometrically diffeomorphic to an open submanifold of the Kerr spacetime. \end{theorem} The key idea of the proof is to show that the set of multipole moments defined in the asymptotic end ${\cal M}^{\infty}$ of $({\cal M},g)$ coincides with the set of multipole moments of the Kerr spacetime (for a certain mass and angular momentum). Then, a previous result by Beig and Simon \cite{BS} stating that two asymptotically flat spacetimes with the same set of multipole moments must be isometric in a neighbourhood of infinity completes the proof. \section{The spacetime Simon tensor} In this section we will obtain a tensor defined on any spacetime with a Killing vector, which coincides with the Simon tensor when projected down into to the manifold of trajectories (whenever this exists) and such that it vanishes identically for the Kerr spacetime. Throughout this section $\left ({\cal M},g \right )$ denotes a $C^2$ vacuum spacetime admitting a non-trivial Killing field $\vec{\xi}$. We do not require that the orbits of this Killing vector are complete (i.e. the Killing vector $\vec{\xi}$ need not generate a one parameter isometry group). A key object in our analysis is the exact two-form ${\bm F}_{\alpha \beta} = \nabla_{\alpha} \xi_{\beta}$ (we use boldface characters to denote two-forms). A complex two-form $\cm{B}$ satisfying $\cm{B}^{\star} = - i \cm{B}$, where $\star$ is the Hodge dual operator, is called self-dual (calligraphic letters will be used to denote them). Given any real two-form $\bm B_{\alpha\beta}$, its self-dual part $\cm{B}_{\alpha\beta}$, is defined by $\cm{B}_{\alpha\beta} = \bm{B}_{\alpha\beta} + i \, \, \bm{B}^{\star}_{\alpha \beta}$. In particular, \begin{eqnarray} \cm{F}_{\alpha\beta} = \bm{F}_{\alpha\beta} + i \bm{F}^{\star}_{\alpha\beta} \label{Fdual} \end{eqnarray} is called {\it Killing form} throughout this paper. Since most of the calculations in this section involve two-forms, let us recall some well-known identities (see e.g. \cite{Israel}). Let $\bm{X}$ and $\bm{Y}$ be arbitrary two-forms on ${\cal M}$, then \begin{eqnarray*} \bm X_{\mu\sigma} \bm Y_{\nu}^{\ssigma} - {\bm X^{\star}}_{\mu\sigma} {\bm Y^{\star}}_{\nu}^{\ssigma} = \frac{1}{2} g_{\mu\nu} \bm X_{\alpha\beta} \bm Y^{\alpha\beta} , \quad \bm X_{\mu\sigma} {\bm X^{\star}}_{\nu}^{\ssigma} =\frac{1}{4} g_{\mu\nu} \bm X_{\alpha\beta} {\bm X^{\star}}^{\alpha\beta}, \end{eqnarray*} which specialized to arbitrary self-dual two-forms, $\bm{\cal X}$, $\bm{\cal Y}$, give \begin{eqnarray} \bm{\cal X}_{\mu\sigma} \bm{\cal Y}_{\nu}^{\ssigma} + \bm{\cal Y}_{\mu\sigma} \bm{\cal X}_{\nu}^{\ssigma} = \frac{1}{2} g_{\mu\nu} \bm{\cal X}_{\alpha\beta} \bm{\cal Y}^{\alpha\beta}, \quad {\bm{\cal X}}_{\mu\sigma} \bm{\cal X}_{\nu}^{\ssigma} = \frac{1}{4} g_{\mu\nu} \bm{\cal X}_{\alpha\beta} \bm{\cal X}^{\alpha\beta}. \label{dual} \end{eqnarray} For any self-dual two-form $\bm{\cal Z}$ we have \begin{eqnarray} \bm{\cal Z}_{\mu\sigma} \bm Z_{\nu}^{\ssigma} - \bm{\cal Z}_{\nu\sigma} \bm Z_{\mu}^{\ssigma} =0, \label{orto} \end{eqnarray} where $\bm{Z}$ denotes the real part of $\bm{\cal Z}$. From the equation $\nabla_{\alpha} \nabla_{\beta} \xi_{\mu} = \xi^{\sigma} \, C_{\sigma\alpha\beta\mu}$, which follows from the Killing equations in vacuum, we obtain \begin{eqnarray} \nabla_{\alpha} \F_{\beta\gamma} = \xi^{\sigma} \C_{\sigma\alpha\beta\gamma}, \label{df} \end{eqnarray} where $\C_{\alpha\beta\gamma\delta}$ is the so-called right self-dual Weyl tensor, defined as $\C_{\alpha\beta\gamma\delta} = C_{\alpha\beta\gamma\delta} + \frac{i}{2} \eta_{\gamma\delta\rho\sigma} C_{\alpha\beta}^{\,\,\,\,\,\,\,\, \rho\sigma}$. This tensor shares all the symmetries with the Weyl tensor, i.e. it is a symmetric double two-form, with vanishing trace, which satisfies the Bianchi identity $\C_{\alpha \left [\beta\gamma\delta \right ] } =0$. It follows from (\ref{df}) that $\F$ satisfies Maxwell's equations $d \F = 0$. The Ernst one-form is defined by \begin{eqnarray} \sigma_{\mu} \equiv 2 \xi^{\alpha} \F_{\alpha\mu} = \nabla_{\mu} \lambda - i \omega_{\mu}, \label{Ernst} \end{eqnarray} which has as an immediate consequence that \begin{eqnarray} \sigma_{\mu} \sigma^{\mu} = -\lambda \F_{\alpha\beta} \F^{\alpha\beta}. \label{s2} \end{eqnarray} The Ernst one-form $\sigma_{\alpha}$ is closed, $\nabla_{\left [\mu\right .} \sigma_{\left . \nu \right ]} = 2 \F_{\alpha \left [ \nu \right .} \bm{F}_{\left . \mu \right ]}^{\,\,\,\,\alpha} + \xi^{\alpha} \nabla_{\alpha} \F_{\mu\nu} = 0, $ where $d\F = 0$ was used in the first equality and (\ref{orto}) and (\ref{df}) in the second one. Let us assume for the moment that the set ${\cal M}^{+}= \left \{ p \in {\cal M} ; \lambda|_p > 0 \right \}$ is non-empty and that ${\cal N}^{+}= {\cal M}^{+} / \approx$ is a manifold, so that the Simon tensor can be defined along the lines described in the previous section. We want to translate the Simon tensor into a tensor in ${\cal M}^{+}$ and analyze whether it can be extended to all of ${\cal M}$. The spacetime counterpart of $\sigma_i$ in (\ref{sigmai}) is the Ernst one-form $\sigma_{\mu}$. The symmetric tensor $\hat{Z}_{ij}$ will also have a spacetime counterpart in ${\cal M}^{+}$, which we denote by $\hat{Z}_{\mu\nu}$. As we shall see below, $\hat{Z}_{\mu\nu}$ is singular on the boundary of ${\cal M}^{+}$ (if non-empty). In order to see that the Simon tensor remains nevertheless regular, we need to analyze the structure of the Simon tensor. To that end, let us consider an arbitrary point $p \in {\cal M}^{+}$ and define the following linear map between two-index symmetric, covariant tensors at $p$ which are completely orthogonal to $\vec{\xi}$ and three-index covariant tensors at $p$ \begin{eqnarray} U(Z)_{\alpha\beta\gamma} = Z_{\alpha\beta} \sigma_{\gamma} - Z_{\alpha\gamma} \sigma_{\beta} - \frac{1}{2} h_{\alpha\beta} \left ( Z \sigma_{\gamma} - \sigma^{\mu} Z_{\mu\gamma} \right )+ \frac{1}{2} h_{\alpha\gamma} \left ( Z \sigma_{\beta} - \sigma^{\mu} Z_{\mu\beta} \right ), \label{defU} \end{eqnarray} where $Z = g^{\mu\nu} Z_{\mu\nu}$. The algebraic properties of this map are summarized in the following lemma, which is proven by straightforward calculation. \begin{lemma} \label{proper} At any point $p \in {\cal M}^{+}$ and for any symmetric tensor $Z_{\alpha\beta} |_p$ completely orthogonal to $\vec{\xi}$, the tensor $U_{\alpha\beta\gamma} = \left . U(Z)_{\alpha\beta\gamma} \right |_p$ defined by (\ref{defU}) satisfies \begin{eqnarray*} 1) ~ U_{\alpha\beta\gamma} \mbox{ is completely orthogonal to } \vec{\xi}, \hspace{9mm} 2) ~ U_{\alpha\beta\gamma} = - U_{\alpha\gamma\beta}, \hspace{9mm} 3) ~ U^{\alpha}_{\,\,\,\,\,\alpha\beta} = 0, \\ 4) ~ U_{\left [\alpha\beta\gamma\right ]} = 0, \hspace{9mm} 5) ~ \sigma^{\alpha} \sigma^{\beta} \left ( U_{\alpha\beta\gamma} \sigma_{\delta} - U_{\alpha\beta\delta} \sigma_{\gamma} \right ) + \sigma^{\mu} \sigma_{\mu} \sigma^{\alpha} U_{\alpha\gamma \delta} = 0. \hspace{9mm} \end{eqnarray*} \end{lemma} Properties $2)$, $3)$ and $4)$ are standard for Cotton-like tensors. Property $5)$ is specific for the Simon tensor. A trivial calculation shows that the map $U(Z)$ satisfies \begin{eqnarray} U \left ( Z_{\alpha\beta} + s_1 h_{\alpha\beta} + s_2 \sigma_{\alpha} \sigma_{\beta} \right ) = U \left ( Z_{\alpha\beta} \right ) \label{kern1} \end{eqnarray} where $s_1$ and $s_2$ are arbitrary complex constants. This indicates that the kernel of $U(Z)$ is at least two dimensional. The following lemma shows that at points where $\F_{\alpha\beta} \F^{\alpha\beta} \neq 0$ the kernel of $U(Z)$ is indeed two-dimensional and that $U(Z)$ is surjective onto the set of tensors satisfying properties $1)$ to $5)$ above. \begin{lemma} Let $U_{\alpha\beta\gamma}$ be a tensor at $p \in {\cal M}^{+}$ satisfying properties $1)$ to $5)$ in lemma \ref{proper} and assume that $\F_{\alpha\beta} \F^{\alpha\beta} |_p \neq 0$. Then $\sigma_{\mu} \sigma^{\mu} |_p \neq 0$ and the general solution of the algebraic equation $U(Z^{0} )_{\alpha\beta\gamma} = U_{\alpha\beta\gamma}$ is given by \begin{eqnarray} Z^{0}_{\alpha\beta} = \left . \frac{\sigma^{\gamma} \left ( U_{\alpha\beta\gamma} +U_{\beta\alpha\gamma} \right ) }{2\sigma_{\mu} \sigma^{\mu}} + \frac{3 \sigma^{\nu} \sigma^{\gamma} \left ( U_{\nu\alpha\gamma}\sigma_{\beta} +U_{\nu\beta\gamma}\sigma_{\alpha} \right )} {2\left (\sigma^{\mu}\sigma^{\mu} \right )^2} + s_1 h_{\alpha\beta} + s_2 \sigma_{\alpha} \sigma_{\beta} \right |_p \label{aleqsol} \end{eqnarray} where $s_1$ and $s_2$ are arbitrary constants. \label{inv} \end{lemma} {\it Proof.} Since $\lambda \F_{\alpha\beta} \F^{\alpha\beta} |_p \neq 0$, equation (\ref{s2}) shows $\sigma_{\mu} \sigma^{\mu} |_p \neq 0$. Making use of property (\ref{kern1}) the problem can be restricted to the one of finding the general solution of \begin{eqnarray} \left . {\tilde{Z}}_{\alpha\beta} \sigma_{\gamma} - {\tilde{Z}}_{\alpha\gamma} \sigma_{\beta} + \frac{1}{2} h_{\alpha\beta} \sigma^{\mu} {\tilde{Z}}_{\mu\gamma} - \frac{1}{2} h_{\alpha\gamma} \sigma^{\mu} {\tilde{Z}}_{\mu\beta} \right |_p= U_{\alpha\beta\gamma}, \label{aleq} \end{eqnarray} for tensors ${\tilde{Z}}_{\alpha\beta}$ satisfying ${\tilde Z}= {\tilde{Z}}_{\alpha\beta} \sigma^{\alpha} \sigma^{\beta} |_p =0$. Contracting (\ref{aleq}) with $\sigma^{\gamma}$ and $\sigma^{\alpha}$ alternatively, it is easy to see that the only possible solution of this equation is given by the right-hand side of (\ref{aleqsol}) with $s_1 = s_2=0$. It remains to show that (\ref{aleq}) is fulfilled. This can be proven by introducing three mutually orthogonal, unit complex vectors $e^i_{\mu}$, $i=1,2,3$, at $p$ satisfying $\xi^{\mu} e^i_{\mu} |_p = 0$ and such that $\sigma_{\mu} |_p = (\sigma_{\beta} \sigma^{\beta})^{1/2} |_p e^3_{\mu}$. Expanding $U_{\alpha\beta\gamma}$ in terms of these vectors it is easy to obtain the most general form allowed by the algebraic properties 1)-5) in lemma \ref{proper}. Expanding also ${\tilde Z}_{\alpha\beta}$ in terms of this basis, it is a matter of simple calculation to check that (\ref{aleq}) holds identically. $\hfill \Box$ Let us return to the tensor $\hat{Z}_{\alpha\beta}$. The following lemma is key to show that the Simon tensor can be extended to all of ${\cal M}$. \begin{lemma} $\hat{Z}_{\nu\mu}$, i.e. the spacetime counterpart of $D_m \sigma_n$ on ${\cal M}^{+}$, can be written as \begin{eqnarray*} \hat{Z}_{\nu\mu} =2 \xi^{\alpha} \xi^{\beta} \C_{\alpha\mu\beta\nu} - \frac{1}{2} \bm{{\cal F}^2} h_{\mu\nu} - \frac{1}{2 \lambda} \sigma_{\mu} \sigma_{\nu}, \end{eqnarray*} where $\bm{{\cal F}^2} \equiv \F_{\alpha\beta} \F^{\alpha\beta}$. \label{imp} \end{lemma} {\it Proof.} We shall start with the tensor $\hat{Z}_{mn}$, defined on ${\cal N}^{+}$ and find its spacetime counterpart. Since the two metrics $h_{ij}$ and $\gamma_{ij}$ are conformally related, it follows that \begin{eqnarray*} D_{m} \sigma_l = D^{h}_{m} \sigma_l + \frac{1}{2 \lambda } \left (h_{lm} \sigma^{k} D_k \lambda - \sigma_l D_m \lambda - \sigma_m D_l \lambda \right ), \end{eqnarray*} where $D^h$ denotes the Levi-Civita covariant derivative of $h$, and indices are raised with $h^{ij}$. The spacetime counterpart of the right-hand side can be written down by recalling that each covariant derivative on ${\cal N}^{+}$ must be transformed into a covariant derivative on ${\cal M}^{+}$ and then projected with $h^{\,\,\alpha}_{\beta} = \delta^{\,\,\alpha}_{\beta} + \lambda^{-1} \xi_{\beta} \xi^{\alpha}$. Hence \begin{eqnarray*} D_{m} \sigma_n \quad \longleftrightarrow \quad \hat{Z}_{\mu\nu}= h_{\mu}^{\,\,\mu^\prime} h_{\nu}^{\,\,\nu^\prime} \nabla_{\mu^\prime} \sigma_{\nu^\prime} + \frac{1}{2 \lambda } \left (h_{\mu\nu} \sigma^{\delta} \nabla_{\delta} \lambda - \sigma_{\mu} \nabla_{\nu} \lambda - \sigma_{\nu} \nabla_{\mu} \lambda \right ), \end{eqnarray*} where the arrow stands for the one-to-one correspondence between tensors on ${\cal N}^{+}$ and tensors on ${\cal M}^{+}$. Using (\ref{df}) and (\ref{Ernst}), we immediately obtain \begin{eqnarray*} h_{\mu}^{\,\,\mu^\prime} h_{\nu}^{\,\,\nu^\prime} \nabla_{\mu^\prime} \sigma_{\nu^\prime} = 2 h_{\mu}^{\,\,\mu^\prime} h_{\nu}^{\,\,\nu^\prime} \bm{F}_{\mu^\prime}^{\,\,\,\alpha} \F_{\alpha\nu \prime} + 2 \xi^{\alpha} \xi^{\beta} \C_{\alpha\mu\beta\nu}. \end{eqnarray*} Thus, proving the lemma amounts to showing that \begin{eqnarray} 2 h_{\mu}^{\,\,\mu^\prime} h_{\nu}^{\,\,\nu^\prime} \bm{F}_{\mu^\prime}^{\,\,\,\alpha} \F_{\alpha\nu^\prime} + \frac{1}{2 \lambda } \left (h_{\mu\nu} \sigma^{\delta} \nabla_{\delta} \lambda - \sigma_{\mu} \nabla_{\nu} \lambda - \sigma_{\nu} \nabla_{\mu} \lambda \right ) + \frac{1}{2} \bm{{\cal F}^2} h_{\mu\nu} + \frac{1}{2 \lambda} \sigma_{\mu} \sigma_{\nu}= 0 \label{bigeq} \end{eqnarray} holds on ${\cal M}^{+}$. The imaginary part of this equation is an immediate consequence of the imaginary part of (\ref{s2}) which reads $\omega^{\alpha} \nabla_{\alpha} \lambda = \lambda \bm F_{\alpha\beta} {\bm F^{\star}}^{\alpha\beta}$. Regarding the real part of (\ref{bigeq}), the relation follows from the remarkable identity (here comma stands for partial differentiation) \begin{eqnarray} \omega_{\mu} \omega_{\nu} + \lambda_{,\mu} \lambda_{,\nu} = \hspace{11cm} \nonumber \\ \hspace{15mm} 2 \left ( \lambda g_{\mu\nu} + \xi_{\mu} \xi_{\nu} \right ) \nabla^{\alpha} \xi^{\beta} \nabla_{\alpha} \xi_{\beta} + g_{\mu\nu} \lambda_{,\alpha} \lambda^{, \alpha} - 4 \lambda \nabla_{\nu} \xi_{\alpha} \nabla_{\mu} \xi^{\alpha} - 4 \lambda^{, \alpha} \xi_{\left ( \nu \right .} \nabla_{\left . \mu \right )} \xi_{\alpha} \label{Ident} \end{eqnarray} which holds for any Killing vector. This identity can be proven, after a somewhat long but trivial calculation, by expanding $\eta_{\alpha\beta\gamma\delta} \,\eta_{\mu\nu\rho\sigma}$, appearing in $\omega_{\mu}\omega_{\nu}$ on the left-hand side, in terms of products of $g_{\alpha\beta}$. $\hfill \Box$ This lemma shows that, although $\hat{Z}_{\alpha\beta}$ becomes singular at points where $\lambda$ goes to zero, the diverging part belongs to the kernel of the map $U(\hat{Z})$. Therefore, the spacetime counterpart of the Simon tensor, which is $S=U(\hat{Z})$, remains regular at the boundary of ${\cal M}^{+}$. In other words, let us define the symmetric, trace-free tensor \begin{eqnarray} Y_{\mu\nu} = 2 \xi^{\alpha} \xi^{\beta} \C_{\alpha\mu\beta\nu} \label{Y} \end{eqnarray} which satisfies, due to (\ref{dual}), $\sigma^{\alpha} Y_{\alpha\beta} = - \lambda \xi^{\mu} \C_{\mu\beta\gamma\delta} \F^{\gamma\delta}$. Using lemmas \ref{inv} and \ref{imp}, we find that the spacetime counterpart of the Simon tensor at any point $p \in {\cal M}^{+}$ reads \begin{eqnarray*} S_{\alpha\beta\nu} = U \left (Y \right )_{\alpha\beta\nu} = Y_{\alpha\beta} \sigma_{\nu} - Y_{\alpha\nu} \sigma_{\beta} - \frac{1}{2} \gamma_{\alpha\beta} \xi^{\mu} \C_{\mu\nu\rho\delta} \F^{\rho\delta} +\frac{1}{2} \gamma_{\alpha\nu} \xi^{\mu} \C_{\mu\beta\rho\delta} \F^{\rho\delta}. \end{eqnarray*} All the objects in this expression are well-defined in ${\cal M}$. Moreover, this definition makes sense irrespectively of whether the vacuum field equations hold or not. Thus, let us put forward the following definition \begin{definition} Let $({\cal M},g)$ be a $C^2$ spacetime with a Killing field $\vec{\xi}$. Construct $\F_{\mu\nu}$ and $\sigma_{\mu}$ through expressions (\ref{Fdual}) and (\ref{Ernst}). The spacetime Simon tensor with respect to $\vec{\xi}$ is defined by \begin{eqnarray*} S_{\alpha\beta\nu} = U \left (Y \right )_{\alpha\beta\nu} = 2 \xi^{\mu} \xi^{\rho} \C_{\mu\alpha\rho\beta} \sigma_{\nu} - 2 \xi^{\mu} \xi^{\rho} \C_{\mu\alpha\rho\nu} \sigma_{\beta} - \frac{1}{2} \gamma_{\alpha\beta} \xi^{\mu} \C_{\mu\nu\rho\delta} \F^{\rho\delta} +\frac{1}{2} \gamma_{\alpha\nu} \xi^{\mu} \C_{\mu\beta\rho\delta} \F^{\rho\delta}, \end{eqnarray*} where $\gamma_{\alpha\beta} = \lambda g_{\mu\nu} + \xi_{\alpha} \xi_{\beta}$ and $\lambda = - \xi^{\alpha} \xi_{\beta}$. \end{definition} Let us now consider the implications of the vanishing of the spacetime Simon tensor. So, let $(V,g)$ be a spacetime with a Killing field $\vec{\xi}$ such that the corresponding spacetime Simon tensor vanishes everywhere. Consider a point $p \in V$ where $\vec{\xi}$ has non-zero norm and $\bm{{\cal F}^2}$ is non-zero. From lemma \ref{inv} and the fact that $Y_{\alpha\beta}$ is trace-free it follows that \begin{eqnarray} \left .Y_{\alpha\beta} \right |_p = \left .2 \xi^{\mu} \xi^{\nu} \C_{\mu\alpha\nu\beta} \right |_p = \left .\frac{Q(p)}{2} \left ( \sigma_{\alpha}\sigma_{\beta} + \frac{1}{3} \gamma_{\alpha\beta} \bm{{\cal F}^2} \right ) \right |_p \label{elec} \end{eqnarray} where $Q(p)$ is an arbitrary complex constant. $\C_{\alpha\beta \gamma\delta}$ is a double symmetric self-dual two-form which is well-known to be uniquely characterized by its electric and magnetic parts (see e.g. \cite{KSMH}). Thus we have \begin{eqnarray} \left . \C_{\alpha\beta\gamma\delta} \right |_p = Q(p) \left. \left (\F_{\alpha\beta} \F_{\gamma\delta} - \frac{1}{3} {\cm I}_{\alpha\beta\gamma\delta} \bm{{\cal F}^2} \right ) \right |_p \label{weylSi} \end{eqnarray} where ${\cm I}_{\alpha\beta\gamma\delta} \equiv (g_{\alpha\gamma}\, g_{\beta\delta}-g_{\alpha\delta}\,g_{\beta\gamma} + i \,\eta_{\alpha\beta\gamma\delta})/4$ is the metric in the space of self-dual two-forms. The equality above holds because, by virtue of (\ref{elec}), the electric and magnetic parts of both sides coincide. The following lemma will be important to prove the characterization of Kerr. \begin{lemma} Let $(V,g)$ be a $C^3$ vacuum, not locally flat, spacetime with a Killing vector $\vec{\xi}$ which is non-null on a dense subset of $V$ and such that the corresponding spacetime Simon tensor vanishes everywhere. Assume also that $\bm{{\cal F}^2} \neq 0$ everywhere. Then, the Ernst one-form is exact,(i.e. it exists a function $\sigma$ such that $\sigma_{\alpha} = \nabla_{\alpha} \sigma$) and the Weyl tensor and $\bm{{\cal F}^2}$ take the form \begin{eqnarray*} \C_{\alpha\beta\gamma\delta} = \frac{-6}{c - \sigma} \left (\F_{\alpha\beta} \F_{\gamma\delta} - \frac{1}{3} {\cm I}_{\alpha\beta\gamma\delta} \bm{{\cal F}^2} \right ), \hspace{1cm} \bm{{\cal F}^2} = A \left (c -\sigma \right )^4 \end{eqnarray*} where $c$ and $A\neq 0$ are complex constants. \label{formWeyl} \end{lemma} {\it Proof.} A trivial consequence (\ref{df}) and (\ref{weylSi}) is \begin{eqnarray} \nabla_{\alpha} \bm{{\cal F}^2} = \frac{2}{3} Q \bm{{\cal F}^2} \sigma_{\alpha}. \label{eqF2} \end{eqnarray} Since the spacetime is $C^3$ we can use the second Bianchi identities which, in terms of the right self-dual Weyl tensor, read $\nabla_{\alpha} \C^{\alpha}_{\,\,\,\,\beta\gamma\delta}=0$. A straightforward calculation shows \begin{eqnarray} 0 = 4 \xi^{\beta}\xi^{\gamma} \nabla_{\alpha} \C^{\alpha}_{\,\,\,\,\beta\gamma\delta}= \frac{\lambda}{9} \bm{{\cal F}^2} Q^2 \sigma_{\delta} - \frac{\bm{{\cal F}^2}}{3} \left ( \lambda \nabla_{\delta} Q + \xi_{\delta} \xi^{\alpha} \nabla_{\alpha} Q \right ) - \sigma_{\delta} \sigma^{\alpha} \nabla_{\alpha} Q, \label{expre} \end{eqnarray} which, after contraction with $\sigma^{\delta}$ yields $\sigma^{\alpha} \nabla_{\alpha} Q = \frac{1}{6} \lambda \bm{{\cal F}^2} Q^2$, where we used that $\lambda$ is zero at most on a set with empty interior and $\nabla_{\alpha} Q$ is continuous. Inserting this expression back into (\ref{expre}) we find $ \lambda \left (\nabla_{\delta} Q + \frac{1}{6} Q^2 \sigma_{\delta} \right ) - \xi_{\delta} \xi^{\alpha} \nabla_{\alpha} Q = 0$. From $\C_{\alpha\beta\gamma\delta} \F^{\alpha\beta} \F^{\gamma\delta} = 2/3 \left ( \bm{{\cal F}^2} \right )^2 Q$ we have $\xi^{\alpha} \nabla_{\alpha} Q =0$ and, therefore, \begin{eqnarray} \nabla_{\alpha} Q + \frac{1}{6} Q^2 \sigma_{\alpha} = 0. \label{EqQ} \end{eqnarray} The spacetime is not locally flat by assumption, so $Q$ is not identically zero. Let us define the set $K = \{p \in V ; Q(p)\neq 0 \}$ and suppose first that $K \neq V$. Take a point $q \in\partial K$ (i.e. in the topological boundary of $K$). Recall that the Ernst one-form is closed on all of $V$ and therefore there exists a sufficiently small open neighbourhood $V_q$ of $q$ where the Ernst one-form is exact, i.e. it exists a function $\sigma |_{V_q}$ such that $\sigma |_{V_q} = \nabla_{\alpha} \sigma |_{V_q}$. On $K \cap V_q$, equation (\ref{EqQ}) can be integrated to give $\sigma - \frac{6}{Q} = c$ where $c$ is a complex constant. Since $\sigma$ is well-defined in $V_q$ and in particular bounded at $q$ it follows that $Q$ cannot vanish on $q$, against the assumption. Hence $Q$ is non-zero everywhere. Then, equation (\ref{EqQ}) shows that $\sigma_{\alpha}$ is exact, i.e. $\sigma_{\alpha} = \nabla_{\alpha} \sigma$. The connectedness of $V$ gives $Q = -6/(c - \sigma)$ everywhere. The integration of (\ref{eqF2}) completes the proof. $\hfill \Box$. \section{The main Theorem} A straightforward calculation shows that the spacetime Simon tensor $S_{\alpha\beta\mu}$ with respect to the asymptotically timelike Killing vector vanishes identically in the Kerr metric. The aim of this section is to prove that the converse is also true in the asymptotically flat case. More precisely, we prove the following theorem \begin{theorem} Let $(V,g)$ be a smooth spacetime with the following properties \begin{enumerate} \item The metric $g$ satisfies the Einstein vacuum field equations, \item $(V,g)$ admits a smooth Killing field $\vec{\xi}$ such that the spacetime Simon tensor associated to $\vec{\xi}$ vanishes everywhere, \item $(V,g)$ contains a stationary asymptotically flat four-end $V^{\infty}$, $\vec{\xi}$ tends to a time translation at infinity in $V^{\infty}$ and the Komar mass of $\vec{\xi}$ in $V^{\infty}$ is non-zero. \end{enumerate} Then $(V,g)$ is locally isometric to a Kerr spacetime. \label{Main} \end{theorem} {\bf Remark 1}. By stationary asymptotically flat four-end we understand an open submanifold $V^{\infty} \subset V$ diffeomorphic to $I\times \left (\Bbb{R}^3 \setminus B(R) \right )$, ($I \in \Bbb{R}$ is an open interval and $B(R)$ is a closed ball of radius $R$), such that, in the local coordinates defined by the diffeomorphism, the metric satisfies \begin{eqnarray*} \left |g_{\mu\nu} - \eta_{\mu\nu} \right | + \left |r \partial_i g_{\mu\nu} \right | \leq C r^{-\alpha}, \hspace{1cm} \partial_t g_{\mu\nu} =0 \end{eqnarray*} where $C,\alpha$ are positive constants, $r = \sqrt{ \sum (x^i)^2}$ and $\eta_{\mu\nu}$ is the Minkowski metric. Usually, the definition of asymptotically flat four-end requires $I = \Bbb{R}$, but this is not necessary for our purposes. A result by Kennefick and \'{O} Murchadha \cite{KO} (see also Proposition 1.9 in \cite{Ch3}) shows that the Einstein field equations and the existence of a timelike Killing vector force $\alpha \geq 1$. It is then well-known (see e.g. Beig and Simon \cite{BS1}) that the metric can be brought into the asymptotic form \begin{eqnarray} g_{00} = -1 + \frac{2M}{r} + O(r^{-2}), \hspace{6mm} g_{0i} = - \epsilon_{ijk} \frac{4 S^j x^k}{r^3} + O(r^{-3}), \hspace{6mm} g_{ij} = \delta_{ij} + O(r^{-1}), \label{AF} \end{eqnarray} where $M$ is the Komar mass \cite{Ko} of $\vec{\xi}$ in the asymptotically flat end $V^{\infty}$ and $\epsilon_{ijl}$ is the alternating Levi-Civita symbol. It is worth noticing that assumption 3 in the theorem is used only in order to prove lemma \ref{exWeyl} below, which fixes the value of the two arbitrary constants appearing in lemma \ref{formWeyl}. Hence, assumption 3 in the theorem could be replaced by any hypothesis under which lemma \ref{exWeyl} still holds. {\bf Remark 2}. By Kerr spacetime we mean, as usual, the maximal analytic extension of the Kerr metric, as described by Boyer and Lindquist \cite{BY} and Carter \cite{Carter2}. An element of the Kerr family will be denoted by $(V_{M,a},g_{M,a})$, where $M$ denotes the Komar mass and $a$ the specific angular momentum. In particular, $(V_{M,0},g_{M,0})$ is the Kruskal extension of the Schwarzschild spacetime. {\bf Remark 3}. As lemma \ref{formWeyl} above shows, the vanishing of the Simon tensor at points where $\bm{{\cal F}^2} \neq 0 $ and $\lambda \neq 0$ implies that the Weyl tensor takes the form (\ref{weylSi}), which in particular shows that the Petrov type of the Weyl tensor is D. The general solution for vacuum spacetimes of Petrov type D was found by Kinnersley in \cite{Kinner}. Hence, we could in principle use his results in order to prove the theorem and indeed we share with his method the fact that we employ the Newman-Penrose formalism to prove the theorem. However, we do not assume analyticity of the metric as is usually done in the field of exact solutions, where the main motivation is obtaining explicit metrics satisfying Einstein field equations. This requires extra care with the choice of tetrads and coordinate systems and it is more convenient to produce a self-contained proof. In particular, we try to use invariantly defined quantities and avoid changing the null tetrad in order to make some spin coefficients zero. By doing this, the proof becomes geometrically more transparent and some insight is gained into the path leading from the vanishing of the Simon tensor up to the Kerr metric. {\bf Remark 4}. The theorem states that for any point $p \in (V,g)$, there exists an open neighbourhood $U_p$ of $p$ which is isometrically diffeomorphic to an open submanifold of $V_{M,a}$. Since the characterization of Kerr in terms of the spacetime Simon tensor is local and the only global requirement we make is the existence of an asymptotically flat end (which, as pointed out in Remark 1, is only used in order to fix the value of two constants), we should non expect in principle that the local isometry extends to an isometric embedding of $(V,g)$ into $(V_{M,a},g_{M,a})$. There can be topological obstructions for this global embedding to exist. Analyzing this question in detail would require classifying the spacetimes which are locally isometric to Kerr and which are asymptotically flat (in the sense above, or perhaps under the stronger requirement $I=\Bbb{R}$). It would be necessary, among other things, to determine the discrete isometry groups of $(V_{M,a},g_{M,a})$ such that the quotient metric still has an asymptotically flat four-end. This is not an easy problem because the global structure of the Kerr spacetime is not particularly simple. In the context of black hole uniqueness theorems one is mainly interested in the domain of outer communication. In this case, it is probably easy to show, after assuming $I = \Bbb{R}$ in the definition of asymptotically flat four-end, that the domain of outer communication of $(V,g)$ is diffeomorphic to the domain of outer communication of $(V_{M,a},g_{M,a})$. However, the analysis of this problem will be relevant only if the characterization of Kerr in terms of the spacetime Simon tensor proves useful for extending the black hole uniqueness theorems to the non-analytic case, and we will not consider this question any further here. We should emphasize, however, that theorem \ref{Main} is ``semi-local'' because the existence of the local isometry is shown {\it everywhere}. \hspace{3mm} Throughout this section $(V,g)$ denotes a spacetime fulfilling the requirements of theorem \ref{Main}. Let us normalize the Killing vector and choose the integration constant in the twist potential $\omega$ so that $\sigma \, \rightarrow 1$ at infinity in $V^{\infty}$. A simple calculation using the asymptotic form of the metric (\ref{AF}) in $V^{\infty}$ gives $\bm{{\cal F}^2} = - 4 M^2/r^4 + O (r^{-5})$. Since by assumption $M\neq 0$ we have that the open submanifold $\hat{V}_{f} = \{ p \in V ; \bm{{\cal F}^2} |_p \neq 0 \}$ is non-empty. Possibly after redefining $V^{\infty}$ as an appropriate asymptotically flat open submanifold of $V^{\infty}$, we can assume that $V^{\infty} \subset \hat{V_f}$. Let us define the spacetime $(V_{f}, g_f)$ as the connected component of $\hat{V}_{f}$ containing the asymptotically flat region $V^{\infty}$, with the induced metric. We have the following lemma \begin{lemma} Let $(V,g)$ satisfy the hypotheses of the theorem \ref{Main} and $(V_f,g_f)$ be defined as above. Then, the Weyl tensor and $\bm{{\cal F}^2}$ in $(V_{f},g_{f})$ take the form \begin{eqnarray*} \C_{\alpha\beta\gamma\delta} = \frac{ - 6 }{1 - \sigma} \left (\F_{\alpha\beta} \F_{\gamma\delta} - \frac{1}{3} {\cm I}_{\alpha\beta\gamma\delta} \bm{{\cal F}^2} \right ), \hspace{1cm} \bm{{\cal F}^2} = - \frac{1}{4M^2}\left (1 - \sigma \right )^4. \end{eqnarray*} \label{exWeyl} \end{lemma} {\it Proof.} In order to apply lemma \ref{formWeyl} we must show that the set of points where $\vec{\xi}$ has non-zero norm is dense in $V_{f}$. We use the equation \begin{eqnarray} \nabla^{\mu} \nabla_{\mu} \sigma = - \bm{{\cal F}^2} = - 2 \left ( {\bm F}_{\alpha\beta} {\bm F}^{\alpha\beta} + i \, {\bm F^\star}_{\alpha\beta} {\bm F}^{\alpha\beta} \right ), \label{lapsigma} \end{eqnarray} which follows from $\sigma_{\mu} = 2 \xi^{\alpha} \F_{\alpha\mu}$ and $\nabla^{\alpha} \F_{\alpha\beta}=0$. Assume there exists an open set $U \subset V_{f}$ where $\vec{\xi}$ has zero norm. The real part of (\ref{lapsigma}) implies ${\bm F}_{\alpha\beta} {\bm F}^{\alpha\beta}|_{U} = 0$ and identity (\ref{Ident}) provides $\omega_{\mu} |_{U}=0$. Using (\ref{lapsigma}) again we find $\bm{{\cal F}^2} |_U = 0$, which is impossible in $V_f$. Thus, lemma \ref{formWeyl} can be applied. The asymptotic form of the metric (\ref{AF}) implies $\sigma = 1 - 4M/r + O(r^{-2}) $ in $V^{\infty}$, which combined with the asymptotic behaviour of $\bm{{\cal F}^2}$ gives $c =1$ and $A= -1/(4M^2)$. $\hfill \Box$ It is convenient to define a function $P|_{V_f} = (1-\sigma)^{-1} |_{V_f}$. Equations (\ref{s2}) and (\ref{lapsigma}) read, in terms of $P$, \begin{eqnarray} \nabla_{\mu} P \nabla^{\mu} P = \frac{\lambda}{4M^2}, \hspace{1cm} \nabla^{\mu} \nabla_{\mu} P = \frac{1}{4 M^2} \frac{2 \ov{P} -1 }{P \ov{P}}, \label{P2} \end{eqnarray} where the bar denotes complex conjugate. Define the real functions $y$ and $z$ by $P= y + i \, z$. Notice that $P$ is nowhere zero on $V_f$ and thus $y$ and $z$ cannot vanish simultaneously in $V_f$. As we shall see, the scalar functions $y$ and $z$ are very closely related with the radial coordinate $r$ and the angular coordinate $\theta$ in the Boyer-Lindquist coordinates of the Kerr metric. They have an intrinsic definition in terms of the Ernst potential and will be essential for proving the existence of the local isometry with the Kerr spacetime. Since $\F_{\alpha\beta}$ is self-dual and $\bm{{\cal F}^2} |_{V_{f}} = -1/(4M^ 2 P^4)$, there exist two real, smooth, non-zero, null vector fields $\vec{l}_{\pm}$ satisfying $\left( \vec{l}_{+}, \vec{l}_{-} \right ) = -1$ such that \begin{eqnarray*} \F_{\alpha\beta} = \frac{1}{4 M P^2} \left ( -l_{+\alpha} l_{-\beta}+l_{+\beta} l_{-\alpha} - i ~ \eta_{\alpha\beta \gamma\delta} l^{\gamma}_{+} l^{\delta}_{-} \right ), \end{eqnarray*} ($\vec{l}_{\pm}$ are the eigenvectors $\F_{\alpha\beta}$, i.e. $\F_{\alpha\beta} l_{\pm}^{\alpha} \propto l_{\pm\beta}$). The contraction of this equation with $\xi^{\alpha}$ gives, after splitting into the real and imaginary parts, \begin{eqnarray} \nabla_{\beta} y = \frac{1}{2M} \left [ - \left (\vec{\xi},\vec{l}_{+} \right ) l_{-\beta} +\left (\vec{\xi},\vec{l}_{-} \right ) l_{+\beta} \right ] \hspace{1cm} \nabla_{\beta} z = - \frac{1} {2M} \,\, \eta_{\alpha\beta\gamma\delta} \xi^{\alpha} l_{+}^{\gamma} l_{-}^{\delta}, \label{split} \end{eqnarray} where the parentheses denote the scalar product with respect to the metric $g_f$. Moreover ${\cal L}_{\vec{\xi}} \F_{\alpha\beta}=0$ provides, after using $(\vec{l}_{+},\vec{l}_{-})=-1$, the relation \begin{eqnarray} \left [\vec{\xi}, \vec{l}_{\pm} \right ] = \pm \, C_1 \vec{l}_{\pm} \hspace{5mm} \Longrightarrow \hspace{5mm} \xi^{\alpha} \nabla_{\alpha} \left (\vec{\xi},\vec{l}_{\pm} \right )= \pm \, C_1 \left (\vec{\xi}, \vec{l}_{\pm} \right ) \label{commutator} \end{eqnarray} for some function $C_1$. Contracting $\sigma_{\beta} = 2 \xi^{\alpha} \F_{\alpha\beta}$ with $\F_{\mu}^{\,\,\,\,\beta}$ and using (\ref{split}) we get \begin{eqnarray} \xi_{\beta} = - \left (\vec{\xi},\vec{l}_{+} \right ) l_{-\beta} - \left (\vec{\xi},\vec{l}_{-} \right ) l_{+\beta} - 2M \, \eta_{\beta\mu\gamma\delta} \nabla^{\mu} z \, l_{+}^{\gamma} l_{-}^{\delta}. \label{Killing} \end{eqnarray} \begin{proposition} The norms of $\nabla_{\alpha} z$ and $\nabla_{\alpha} y$ are \begin{eqnarray} \left . \nabla_{\alpha} z \nabla^{\alpha} z \right |_{V_{f}} = \left . \frac{B - z^2}{4 M^2 \left (y^2 + z^2 \right )} \right |_{V_{f}} \hspace{1cm} \left . \nabla_{\alpha}y \nabla^{\alpha} y \right |_{V_{f}} =\left . \frac{y^2 - y + B}{4 M^2 \left (y^2 + z^2 \right )} \right |_{V_{f}} \end{eqnarray} where $B$ is a non-negative constant. Moreover $z^2 |_{V_{f}} \leq B$. \label{normyz} \end{proposition} {\it Proof.} The real part of the first equation in (\ref{P2}) reads \begin{eqnarray} \nabla_{\alpha} y \nabla^{\alpha} y - \nabla_{\alpha} z \nabla^{\alpha} z = \frac{1}{4 M^2} \left (1- \frac{y}{y^2+ z^2} \right ) \label{reldydz} \end{eqnarray} so that we only need to prove $4 M^2 \nabla_{\alpha} z \nabla^{\alpha} z = (B- z^2)/(y^2+ z^2)$. Let us define a function $H = 4 M^2 P \ov{P} \nabla_{\alpha} z \nabla^{\alpha} z$. Take an arbitrary point $p \in V_f$ and consider a sufficiently small open neighbourhood $U_p \subset V_f$ of $p$ so that there exists a smooth complex vector field $\vec{m}|_{U_p}$ such that $\{\vec{l}_{+}, \vec{l}_{-}, \vec{m}, \vec{\ov{m}} \}$ is a null tetrad in $U_p$ with positive orientation, ( i.e. satisfying $\eta_{\alpha\beta\gamma\delta} l_{+}^{\alpha} l_{-}^{\beta} m^{\gamma} \ov{m}^{\delta} = - i$). We shall use the Newman-Penrose (NP) notation to denote the Ricci rotation coefficients associated with this null basis. We shall follow the conventions in \cite{KSMH} except that we use $\vec{l}_{+}$ and $\vec{l}_{-}$ instead of $\vec{k}$ and $\vec{l}$. Lemma \ref{exWeyl} shows that $\vec{l}_{\pm} $ are principal null directions of the Weyl tensor and, from the Goldberg-Sachs theorem \cite{GS}, they define geodesic and shearfree null congruences, which in NP notation means $\kappa = \sigma = \nu = \lambda = 0$.\footnote{The spin coefficient $\sigma$ has nothing to do with the Ernst potential $\sigma$ we have been using throughout. This is the only place where it occurs and therefore no confusion should arise.} The subset of NP equations we shall require are \begin{eqnarray} D \tau = \rho \left (\tau + \ov{\pi} \right ) + \tau \left (\epsilon - \ov{\epsilon} \right ), \hspace{4mm} & & \hspace{4mm} \delta \rho = \rho \left (\ov{\alpha} + \beta \right ) + \tau \left ( \rho - \ov{\rho} \right ), \label{seteqs1} \\ \Delta \pi = - \mu \left (\pi + \ov{\tau} \right ) + \pi \left (\ov{\gamma} - \gamma \right ), \hspace{4mm} & & \hspace{4mm} \ov{\delta} \mu = - \pi \left (\mu - \ov{\mu} \right ) - \mu \left ( \alpha + \ov{\beta} \right ). \label{seteqs2} \end{eqnarray} For later use, let us also quote the commutators between the tetrad vectors, which read \begin{eqnarray} \Delta \, D - D \, \Delta & = & \left (\gamma + \ov{\gamma} \right ) D + \left (\epsilon + \ov{\epsilon} \right ) \Delta - \left (\tau + \ov{\pi} \right) \ov{\delta} - \left (\ov{\tau} + \pi \right ) \delta \label{DeltaD}, \\ \delta \,D - D \, \delta & =& \left (\ov{\alpha} + \beta - \ov{\pi} \right ) D - \left ( \ov{\rho} + \epsilon - \ov{\epsilon} \right ) \delta \label{deltaD}, \\ \delta \, \Delta - \Delta \, \delta & =& \left (\tau - \ov{\alpha} - \beta \right ) \Delta + \left (\mu - \gamma + \ov{\gamma} \right ) \delta \label{deltaDelta}, \\ \ov{\delta} \, \delta - \delta \, \ov{\delta} & =& \left (\ov{\mu} - \mu \right ) D + \left ( \ov{\rho} - \rho \right ) \Delta - \left (\ov{\alpha} - \beta \right ) \ov{\delta} - \left (\ov{\beta} - \alpha \right ) \delta. \label{deltadeltabar} \end{eqnarray} From lemma \ref{exWeyl} the only non-zero component of the Weyl spinor in this tetrad is $\Psi_2= - 1/(8m^2 P^3)$. The Bianchi identities in the NP formalism are $D P = - \rho P$, $\Delta P = \mu P$, $\delta P = - \tau P$ and $\ov{\delta} P = \pi P$. Thus, (\ref{split}) becomes \begin{eqnarray} \left . \nabla_{\beta} y \right |_{U_p} = P \left (\rho l_{-\beta} - \mu l_{+\beta} \right ), \hspace{1cm} \left . \nabla_{\beta} z \right |_{U_p} = - i \,\, \pi P m_{\beta} + i \,\, \tau P \ov{m}_{\beta} \label{Nabyz} \end{eqnarray} which implies $\rho P = \ov{\rho P }$, $\mu P = \ov{\mu P } $ and $\tau P = \ov{\pi P } $. In terms of the spin coefficients, the function $H$ reads $H = 8 M^2 P^2 \ov{P}^2 \pi \ov{\pi} \geq 0$ and equation (\ref{reldydz}) becomes $H = 8 M^2 P^3 \ov{P} \rho \mu + y - y^2 - z^2$. A straightforward calculation using equations (\ref{seteqs1})-(\ref{seteqs2}) gives $D H = 0,$ $\Delta H = 0$, $\delta H = -2 i\,\ov{\pi} \ov{P} z$. Equation (\ref{Nabyz}) implies $D z = \Delta z =0$ and $\delta z = i \,\ov{\pi} \ov{P}$, from which the constancy of $H+ z^2$ in $U_p$ follows trivially. Since this holds for any $p \in V_f$ and $V_f$ is connected, we obtain $H |_{V_f} = B - z^2 |_{V_f}\geq 0$ where $B$ is a non-negative constant. $\hfill \Box$ Let us now define two open sets $V_{\pm} = \{ p \in V_{f} \,\, ; \,\, (\vec{\xi}, \vec{l}_{\pm}) |_p \neq 0 \}$ and introduce the null vector fields \begin{eqnarray*} \left . \vec{s}_{\pm} \,\, \right |_{V_{\pm}} = \left . \frac{2M}{ \left (\vec{\xi},\vec{l}_{\pm} \right )} \,\, \vec{l}_{\pm} \right |_{V_{\pm}} , \hspace{1cm} \left . \vec{k}_{\pm} \,\, \right |_{V_{\pm}} = \left . \frac{ \left (\vec{\xi},\vec{l}_{\pm} \right )}{2M} \,\, \vec{l}_{\mp} \right |_{V_{\pm}} \end{eqnarray*} which satisfy $\left . \left (\vec{s}_{\pm},\vec{\xi} \right ) \right |_{V_{\pm}}=2M$, $\left .\left (\vec{s}_{\pm},\vec{k}_{\pm} \right ) \right |_{V_{\pm}}=-1 $. In terms of $\vec{s}_{\pm}$ and $\vec{k}_{\pm}$, equation (\ref{split}) becomes \begin{eqnarray} \left. \nabla_{\alpha} y \right |_{V_{\pm}}= \left . \mp \left ( k_{\pm\alpha} + W s_{\pm\alpha} \right ) \right |_{V_{\pm}}= \left . \mp \left ( k_{\pm\alpha} + \frac{-y^2+y-B}{2M^2(y^2+z^2)} s_{\pm\alpha} \right ) \right |_{V_{\pm}}. \label{newdy} \end{eqnarray} where we have defined \begin{eqnarray} \label{defW} \left . W \right |_{V_f} \equiv \left . - \frac{\left (\vec{\xi}, \vec{l}_{+} \right ) \cdot \left (\vec{\xi}, \vec{l}_{-} \right )}{4M^2} \right |_{V_f} = \left . \frac{-y^2 + y - B}{8M^2 (y^2+ z^2)} \right |_{V_f}, \end{eqnarray} the equality being a consequence of proposition \ref{normyz}. The case in which the Killing vector $\vec{\xi}$ is hypersurface orthogonal requires a separate treatment. This case is characterized by $\nabla_{\mu} z=0$ in some open set ${\cal U}$. Then, the imaginary part of the second equation in (\ref{P2}) implies $z |_{\cal U}=0$. Proposition \ref{normyz} shows $B=0$ and hence that $z=0$ everywhere. \begin{proposition} Assume that the Killing vector $\vec{\xi}$ of $(V_f,g_f)$ is hypersurface orthogonal. Then, at any point $p \in V_{+} \cup V_{-}$ there exists a neighbourhood $U_p$ of $p$ which can be isometrically embedded into the Kruskal-Schwarzschild spacetime. \label{Schwar} \end{proposition} {\it Proof.} For definiteness, we shall give the proof only for $p \in V_{+}$. The proof for $V_{-}$ requires only a few sign changes and will be omitted. So, take a point $p \in V_{+}$ and a sufficiently small open neighbourhood $U_p \subset V_{+}$ of $p$. We shall use the same notation as in the proof of proposition \ref{normyz}. Since $P=y$ is real, so are $\mu$ and $\rho$. Furthermore $z=0$ implies $\pi = \tau = 0$. The commutators (\ref{DeltaD}) and (\ref{deltadeltabar}) show that the two-planes spanned by $\{\vec{l}_{+},\vec{l}_{-} \}$ and by $\{\vec{m},\vec{\ov{m}} \}$ are surface-forming. Denote by $\{S^{\iota}_{l} \}$ and $\{S^{\iota}_{m} \}$ ($\iota$ is an index labeling each surface) the corresponding families of surfaces. Thus, $U_p$ (or some open subset thereof containing $p$) is foliated by two families of mutually orthogonal two-surfaces. We want to show that the induced metric in $S^{\iota}_{m}$ is that of a two-sphere. To that end, we use the Gauss equation to obtain the curvature scalar of the induced metric in $S^{\iota}_{m}$. Recall that the second fundamental form ${\cal K}$ of a surface is defined by ${\cal K}(\vec{X},\vec{Y})=\left (\nabla_{\vec{X}} \vec{Y} \right )^{\bot}$, where $\vec{X}, \vec{Y}$ are tangent vectors to the surface and $\bot$ denotes the orthogonal projection to the surface. A simple calculation using the definition of the NP coefficients and (\ref{split}) gives \begin{eqnarray} \left . {\cal K} (\vec{X}, \vec{Y} )^{\beta} \right |_{S^{\alpha}_{m}} = \left . - \frac{\nabla^{\beta} y}{y} \, \, g_f \left (\vec{X}, \vec{Y} \right ) \right |_{S^{\iota}_{m}}. \label{secform} \end{eqnarray} Using the Gauss equation (see e.g. \cite{KN}) combined with lemma \ref{exWeyl}, we obtain $\hat{R}^{\iota} = 1/(2M^2 y^2)|_{S^{\iota}_{m}}$, where $\hat{R}^{\iota}$ is the Ricci scalar of the surface $S^{\iota}_{m}$. Since $\delta y = 0$ (i.e. $y$ is constant on $S^{\iota}_{m}$) it follows that $S^{\iota}_{m}$ has positive constant curvature and therefore its metric is locally the round metric of a two-sphere of radius $r=2My$. As $\nabla_{\alpha} y$ is nowhere zero on $V_{+} \cup V_{-}$, there exist three functions $x^0, x^2, x^3$ defined on $U_p$ so that $\{x^0,y,x^2,x^3\}$ is a coordinate system on $U_p$ adapted to the foliation, i.e such that $S^{\iota}_{m}$ is defined by $x^0 = \mbox{const}$, $y=\mbox{const}$ and $S^{\iota}_{l}$ is defined by $x^A= \mbox{const}, \: A=2,3$. Expression (\ref{Killing}) shows that $\vec{\xi}$ is non-zero on $V_{+} \cup V_{-}$ and tangent to $S^{\iota}_{l}$, hence $\vec{\xi}\, |_{U_p}= \xi^0 \partial_{x^0} |_{U_p}$ for some non-zero function $\xi^0$. The Killing equations imply $\partial_{x^B} \xi^0 |_{U_p} = 0$. Regarding $\vec{s}_{+}$, (\ref{newdy}) implies $\vec{s}_{+}|_{U_p} = \partial_y + s^0 \partial_{x^0} |_{U_p}$ for some function $s^0$. The commutator (\ref{deltaD}) shows $\partial_{x^B} s^0 = 0$. Using $[\vec{\xi},\vec{s}_{+} ]=0$ it is easy to see that there exists a function $u |_{U_p}$ such that $\{u,y,x^2,x^3\}$ is a coordinate system in $U_p$ in which $\vec{\xi}= \partial_{u} |_{U_p}$ and $\vec{s}_{+}=\partial_{y} |_{U_p}$. It remains to determine $g_{AB} \equiv (\partial_{x^A}, \partial_{x^B} ) |_{U_p}$. Using equation (\ref{secform}) we easily obtain $g_{AB} = 4 M^2 y^2 g^{0}_{AB} (x^C)$, for some functions $g^0_{AB}$ independent of $u$ and $y$. Imposing that the induced metric of $S^{\iota}_{m}$ is the metric of a sphere of radius $r=2My$ we obtain, after defining $r = 2 M y$, that \begin{eqnarray*} \left . ds^2 \right |_{U_p} = \left . \left [ \left (-1 + \frac{2M}{r} \right ) du^2 + 2 du \, dr + r^2 \left (d\theta^2 + \sin^2 \theta d \phi^2 \right ) \right ] \right |_{U_p}, \end{eqnarray*} which proves that $U_p$ can be isometrically embedded into the Kruskal-Schwarzschild spacetime. $\mbox{$\hfill \Box$}$ We now return to the general case (i.e. without imposing staticity). Let us define the vector field \begin{eqnarray} \left .\vec{\eta} \, \right |_{V_{f}} \equiv \left . \frac{1}{2M} \left (B + y^2 \right) \vec{\xi} + \frac{\left (y^2 + z^2 \right)}{2M} \left [ \left (\vec{\xi}, \vec{l}_{-} \right ) \vec{l}_{+} + \left (\vec{\xi}, \vec{l}_{+} \right ) \vec{l}_{-} \right ] \right |_{V_{f}}, \label{eta} \end{eqnarray} which, in $V_{\pm}$, can be rewritten as \begin{eqnarray*} \left . \vec{\eta} \, \right |_{V_{\pm}} = \left . \frac{1}{2M} \left (B + y^2 \right) \vec{\xi} + \left (y^2 + z^2 \right) \vec{k}_{\pm} + \frac{y^2-y+B}{8M^2} \vec{s}_{\pm} \,\,\, \right |_{V_{\pm}}. \end{eqnarray*} Let us still define another open set $V_1 = \{ p \in V_{f} \, \, ; \,\, z^2 |p < B \}$ and a vector field $b^{\alpha} |_{V_1} = 4M^2 (y^2 + z^2 )(B - z^2)^{-1} \nabla^{\alpha} z |_{V_1}$. \begin{proposition} Assume that the Killing vector $\vec{\xi}$ is not hypersurface orthogonal. Then $\{ \vec{\xi}, \vec{s}_{\pm}, \vec{b}, \vec{\eta}\}$ define a holonomic basis of vector fields on $V_{\pm} \cap V_1$. Furthermore, at any point $p$ in $(V_{+} \cup V_{-} ) \cap V_1$ there exists a neighbourhood $U_p$ of $p$ that can be isometrically embedded into the Kerr spacetime. \label{basis} \end{proposition} {\it Proof.} The proof will again be given only for $V_{+} \cap V_1$. The proof in $V_{-}\cap V_1$ is essentially the same. First, we must prove that $\{ \vec{\xi}, \vec{s}_{+}, \vec{b}, \vec{\eta}\}$ commute and are linearly independent at any point on $V_{+} \cap V_1$. An immediate consequence of (\ref{commutator}) is $[\vec{\xi}, \vec{s}_{+} ] |_{V_{+}} = \vec{0}$, $[\vec{\xi}, \vec{k}_{+} ] |_{V_{+}} = \vec{0}$ and hence $[\vec{\xi}, \vec{\eta} \, ] |_{V_{+}} = \vec{0}$. $[\vec{\xi}, \vec{b} \, ] |_{V_1} = \vec{0}$ follows easily from the Killing equations. Let us take an arbitrary point $p \in V_{+} \cap V_1$ and a sufficiently small neighbourhood $U_p \subset V_{+} \cap V_1$ of $p$ where there exists a complex null field $\vec{m}$ such that $\{\vec{s}_{+},\vec{k}_{+},\vec{m}, \vec{\ov{m}} \}$ is a positively oriented null tetrad on $U_p$. All the relations obtained in the proof of proposition \ref{normyz} are still valid when all NP spin coefficients referred to this new null tetrad. Moreover, relation (\ref{newdy}) implies $DP=1$ and $\Delta P = W$ and therefore $\rho = -1/P$, $\mu= W/P$, which can be used to simplify equations (\ref{seteqs1}) and commutators (\ref{DeltaD})-(\ref{deltadeltabar}). In particular, the second equation in (\ref{seteqs1}) gives $\ov{\alpha} + \beta = \ov{\pi}$ which simplifies the calculations considerably. Using (\ref{Nabyz}), $\vec{b}$ takes the form $ b^{\alpha} |_{U_p} = 4 \, i \, M^2 \left (y^2 + z^2 \right )(B - z^2)^{-1} ( \ov{\pi} \ov{P} \ov{m}^{\beta} - \pi P m^{\beta} )$. $[\vec{s}_{+}, \vec{b} \, ]_{U_p} = [\vec{\eta}, \vec{b} \, ]_{U_p} = \vec{0}$ follow after a simple calculation using (\ref{seteqs1}), (\ref{deltaD}) and (\ref{deltaDelta}). In order to prove $[\vec{s}_{+}, \vec{\eta} \,]|_{U_p}= \vec{0}$ we need to evaluate $[\vec{s}_{+},\vec{k}_{+}]$. This is done by rewriting (\ref{Killing}) in the null tetrad, $\vec{\xi} \, |_{U_{p}}= 2 M \left ( W \vec{s}_{+} - \vec{k}_{+} - P \pi \vec{m} - \ov{\pi} \ov{P} \vec{\ov{m}} \right )|_{U_{p}}$. Imposing $[\vec{\xi}, \vec{k}_{+} ] =\vec{0}$, we easily obtain \begin{eqnarray*} \left . \left [\vec{s}_{+} ,\vec{k}_{+} \right ] \right |_{V_{+}}= \left . - \frac{2y}{y^2+z^2} \vec{k}_{+} - \frac{y}{M \left (y^2 + z^2 \right)} \vec{\xi} + \frac{1-2y}{8M^2 \left (y^2+ z^2 \right )} \vec{s}_{+} \right |_{V_{+}}, \end{eqnarray*} from which $[\vec{s}_{+},\vec{\eta} ]|_{V_{+}}=\vec{0}$ is proven by simple calculation. In order to show that $\vec{\xi}$, $\vec{s}_{+}$, $\vec{\eta}$, $\vec{b}$ are linearly independent, we evaluate their scalar products on $V_{+} \cap V_1$, which give \begin{eqnarray} \left (\vec{\xi}, \vec{\xi} \,\right ) = -1 + \frac{y}{y^2+z^2}, \quad \left (\vec{s_{+}}, \vec{\xi} \,\right ) = 2 M, \quad \left (\vec{\xi}, \vec{b} \,\right ) =0, \quad \left (\vec{\xi}, \vec{\eta}\, \right ) = \frac{y \left (B- z^2 \right )}{2 M \left (y^2+z^2 \right )}, \nonumber \\ \left (\vec{s}_{+}, \vec{s}_{+} \right ) =0, \quad \left (\vec{s}_{+}, \vec{b} \,\right ) =0, \quad \left (\vec{s}_{+}, \vec{\eta} \,\right ) = B - z^2, \quad \left (\vec{b}, \vec{b} \,\right ) = \frac{4M^2 \left (y^2 + z^2 \right)}{B - z^2 }, \nonumber \\ \left (\vec{b}, \vec{\eta} \,\right )=0, \quad \left (\vec{\eta}, \vec{\eta} \,\right ) = \frac{B-z^2}{4 M^2 \left (y^2+z^2 \right )} \left [ z^2 \left (y^2-y+B \right ) + y^4+B y^2+By \right ]. \label{scalar} \end{eqnarray} The determinant of the matrix formed by these scalar products (in the obvious order) is $-4M^2 (y^2+z^2) \neq 0$ which implies the linear independence of the vectors. Let us define a positive constant $a$ by $B = a^2/(4M^2)$ and introduce, as before, a function $r$ by $r = 2 M y |_{V_{f}}$. Since $z^2 |_{V_1} < B$, we can also define a function $\theta$ on $V_1$ by $z = a \cos \theta /(2M) |_{V_1}$. The holonomicity of $\{ \vec{\xi}, \vec{s}_{\pm}, \vec{b}, \vec{\eta}\}$ implies that there exist two functions $u|_{U_p}$ and $\phi |_{U_p}$ such that $\{u,r,\theta,\phi \}$ is a coordinate system in $U_p$ satisfying \begin{eqnarray*} \left . \vec{\xi}\, \right |_{U_p} =\frac{\partial}{\partial u}, \quad \left . \vec{s}_{+} \right |_{U_p} = 2 M \frac{\partial}{\partial r}, \quad \left . \vec{b} \,\right |_{U_p} = - \frac{2 M}{a \sin \theta} \frac{\partial}{\partial \theta}, \quad \left . \vec{\eta} \,\right |_{U_p} = \frac{a}{8M^3} \frac{\partial}{\partial \phi}, \end{eqnarray*} which, after using (\ref{scalar}), yields \begin{eqnarray*} \left . ds^2 \right |_{U_p} = \left (-1 + \frac{2Mr}{r^2 + a^2 \cos^2 \theta} \right ) du^2 + 2 du \, dr + \frac{4 M a r \sin^2 \theta}{r^2 + a^2 \cos^2 \theta} du \, d\phi + 2 a \sin^2 \theta dr \, d \phi \\ \left . + \left (r^2 + a^2 \cos^2 \theta \right ) d\theta^2 + \frac{\sin^2 \theta \left [ \left (r^2 + a^2 \right )^2 - a^2 \sin^2 \theta \left (r^2 - 2Mr +a^2 \right) \right ] }{r^2 + a^2 \cos^2 \theta} d\phi^2 \right |_{U_p}. \end{eqnarray*} This is the form of the Kerr metric with mass $M$ and specific angular momentum $a$ in the so-called Kerr coordinates. This shows that $U_p$ is isometrically diffeomorphic to an open submanifold of the Kerr spacetime. $\hfill \Box$ We have shown before that $\nabla_{\alpha} z |_U=0$ for an open set $U \subset V_f$ implies $z |_{V_{f}}=0$ and hence that $\vec{\xi}$ is hypersurface orthogonal everywhere. This shows that $V_1$ is either dense in $V_f$ (when $\vec{\xi}$ is not hypersurface orthogonal) or $V_1$ is empty (when $\vec{\xi}$ is hypersurface orthogonal). The next lemma gives some properties of the complements of $V_1$ and $V_{+}\cup V_{-}$. \begin{lemma} The vector field $\vec{\eta}$ defined above is a Killing vector field in $V_f$. The complement of $V_1$ in $V_f$ is $V_f \setminus V_1= \{ p \in V_f; \vec{\eta} \, |_p=\vec{0}\}$. Moreover, \begin{itemize} \item If $B=0$ then $V_{f} \setminus (V_{+} \cup V_{-} ) = \{p \in V_f ; \vec{\xi} \, |_p= \vec{0}\}$ \item If $0 < B \leq 1/4$ then $V_{f} \setminus (V_{+} \cup V_{-} ) = \{p \in V_f ; (\vec{\eta} - y_{+} \vec{\xi}\, ) \,|_p =\vec{0} \mbox{ or } (\vec{\eta} - y_{-} \vec{\xi}\,) \,|_p = \vec{0} \}$, where $y_{\pm} = (1 \pm (1-4B)^{1/2})/4M$. \item If $B>1/4$ then $V_{f} \setminus (V_{+} \cup V_{-})$ is empty. \end{itemize} \label{fixedpoints} \end{lemma} {\it Proof.} In the particular case $B=0$ we have $z|_{V_f}=0$ and (\ref{eta}) implies $\vec{\eta}|_{V_f}=\vec{0}$, so that the first two claims of the lemma become obvious. When $B\neq 0$, proposition \ref{basis} shows that $\vec{\eta}$ is a Killing vector on the non-empty set $(V_{+} \cup V_{-}) \cap V_1$, which is linearly independent of $\vec{\xi}$. In the non-hypersurface orthogonal case $V_1$, is dense in $V_f$ and hence $\vec{\eta}$ is a Killing vector in $\ov{V_{+} \cup V_{-}}$ (where the bar over a set denotes its closure). Let us show that $\ov{V_{+} \cup V_{-}} = V_f$. Assume that the open set $U_1 \equiv V_f \setminus (\ov{V_{+} \cup V_{-}})$ is non-empty. From the definition of $V_{\pm}$, we have $(\vec{\xi}, \vec{l_{\pm}} )|_{U_1} = 0$ and (\ref{split}) implies $\nabla_{\alpha} y |_{U_1}= 0$. The real part of the second equation in (\ref{P2}) implies then $y |_{U_1} = 1/2$ and proposition \ref{normyz} fixes $B= 1/4$. The vector field $\vec{\eta} - 1/(4M) \vec{\xi}$ is a non-zero Killing field in $\ov{V_{+} \cup V_{-}}$ which vanishes identically on the open set $U_1$, but this is impossible due to well-known properties of Killing vectors. Thus, $\vec{\eta}$ is a Killing field everywhere. The second part of the lemma is an obvious consequence of the definition of $\vec{\eta}$ (\ref{eta}). Regarding the statements for $V_f \setminus (V_{+} \cup V_{-})$, the case $B=0$ is trivial from equation (\ref{Killing}), while the last two statements follow easily from (\ref{defW}) and (\ref{Killing}), after noticing that the vanishing of $(\vec{\xi},\vec{l}_{+})$ implies that of $W$. $\hfill \Box$ We can now prove the main theorem. {\it Proof of the Theorem.} Propositions \ref{Schwar} and \ref{basis} show that the local isometry exists in a neighbourhood of any point in $(V_{+} \cup V_{-})$ in the hypersurface orthogonal case and of any point in $(V_{+} \cup V_{-}) \cap V_1$ in the non-hypersurface orthogonal case. It remains to show that the local isometry exists also in the complement of those sets in $V_f$ which, from lemma \ref{fixedpoints} correspond to the fixed points of certain Killing vectors in $V_f$. Define the open set $W_1= V_{+} \cup V_{-}$ in the hypersurface orthogonal case and $W_1= (V_{+} \cup V_{-}) \cap V_1$ in the non-hypersurface orthogonal one. Take an arbitrary point $q \not\in W_1$ and a neighbourhood $U_q \subset V_f$ of $q$. Since the set of fixed points of a Killing vector is either an isolated point or a smooth, two-dimensional, totally geodesic surface \cite{K}, the set $\tilde{U}_q = U_q \cap W_1$ is connected. Furthermore, any point in $\tilde{U}_q$ has a neighbourhood which can be isometrically embedded in the Kerr spacetime. Combining this two facts and choosing $U_q$ small enough, it is easy to see that $\tilde{U}_q$ can be isometrically embedded into the Kerr spacetime. This map can be extended to the closure of $\tilde{U}_q$ by continuity. Since $\tilde{U}_q$ is dense in $U_q$, this defines a map $\Psi_q \, : \, U_q \,\, \rightarrow (V_{M,a},g_{M,a})$. It is now a simple matter to show, using the continuity of the metric, that $\Psi_q$ is an isometric embedding. In order to complete the proof we need to show that $V_{f} = V$. Assume, on the contrary, that $V_{f}$ is a proper subset of $V$ and take a point $q \in \partial V_{f}$. From continuity of $\bm{{\cal F}^2}$ we must have $\lambda |_q =1$, hence $\vec{\xi}$ is timelike in some neighbourhood of $q$. Consider a smooth curve $\gamma_p (s)$ defined on some interval $s_0 < s \leq 1$ such that $\gamma_p (s) \in V_f$, $\forall s \in (s_0,1)$, $\gamma_p(1)=p$ and such that the tangent vector $\dot{\gamma} (s)$ is orthogonal to $\vec{\xi}$ and of unit length (the existence of such a curve is easy to establish). Define the real function $Y(s) = (y \circ \gamma_p) (s)$. Since $\sigma \rightarrow 1$ when we approach $p$ and $z$ remains bounded, we must have $Y(s) \rightarrow \infty$ when $s \rightarrow 1$. Using $4 M \nabla_{\alpha} y \nabla^{\alpha} y |_{Y(s)} \rightarrow 1$ when $s \rightarrow 1$, we can assume that $\nabla_{\alpha} y |_{Y(s)}$ is spacelike for $s \in (s_0,1)$. Then \begin{eqnarray*} \left (\frac{dY}{ds} (s) \right )^2 = \left ( \nabla_{\alpha} y |_{Y(s)} \dot{\gamma}^{\alpha}_p(s) \right )^2 \leq \nabla_{\alpha} y \nabla^{\alpha} y |_{Y(s)} \end{eqnarray*} where we have used the fact that $\{ \nabla_{\alpha} y |_{\gamma_p(s)}, \dot{\gamma}_p(s) \}$ define a spacelike two-plane and we have applied the Schwarz inequality. Hence $\frac{dY}{ds}$ stays bounded which contradicts $Y \rightarrow \infty$ when $s \rightarrow 1$. This completes the proof of the theorem. $\hfill \Box$. Let us remark that the existence of a local isometry between two spacetimes $(M_1,g_1)$ and $(M_2,g_2)$ on a dense subset of $M_1$ does not imply, in general, the existence of the local isometry everywhere on $M_1$. It is easy to construct counterexamples when the manifolds are not analytic. It is very likely that this cannot happen when the target manifold is analytic and inextendible. In any case, this result is not needed here because the local isometry is ensured on an open, dense and {\it connected} set, which makes the extension of the local isometry to all of $M_1$ simple. \section{Discussion} The characterization of the Kerr metric given by W.Simon \cite{Simon} involved a three-index tensor in the quotient manifold which did not have a clear geometrical interpretation (except in the hypersurface orthogonal case, when it is equivalent to the Cotton tensor). In this paper we have translated the Simon tensor into the spacetime and have extended it everywhere. By doing this, we can also obtain a geometrical interpretation for the Simon tensor and hence a nice geometrical property which characterizes the Kerr metric. Any spacetime with a non-trivial Killing field $\vec{\xi}$ has a privileged two-form ${\bm F}_{\alpha\beta}= \nabla_{\alpha} \xi_{\beta}$, which can be algebraically classified by analyzing the eigenvalue problem associated with its self-dual two-form $\F_{\alpha\beta}$. At points where $\F_{\alpha\beta}$ is so-called regular (i.e. $\bm{{\cal F}^2} \neq 0$), there exist two principal null directions defined as the eigenvectors of $\F_{\alpha\beta}$ (i.e. $\F_{\alpha\beta} l^{\beta} \propto l_{\alpha}$. At points where $\F_{\alpha\beta}$ is singular (i.e $\bm{{\cal F}^2} =0$ with $\F_{\alpha\beta} \neq 0$) there exists one principal null direction. Similarly, the Petrov classification of the Weyl tensor is the algebraic classification of the endomorphism of the space of self-dual two-forms defined by \begin{eqnarray} \C \left ({\cm X} \right )_{\alpha\beta} = \C_{\alpha\beta}^{\,\,\,\,\,\,\,\,\mu\nu} {\cm X}_{\mu\nu}. \label{endom} \end{eqnarray} There exist four principal null directions of the Weyl tensor (which degenerate when the Weyl tensor is algebraically special). A repeated principal null direction of the Weyl tensor is a non-zero null vector field $\vec{l}$ satisfying \begin{eqnarray*} \C_{\alpha\beta\gamma\delta} l^{\beta} l^{\delta} = F l_{\alpha} l_{\gamma} \end{eqnarray*} for some function $F$. In section 3 we have shown that, at points where $\vec{\xi}$ is non-null and $\F_{\alpha\beta}$ is regular, the vanishing of the Simon tensor is equivalent to the Weyl tensor taking the form (\ref{weylSi}). Let us state without proof that this also holds at points $\bm{{\cal F}^2} =0$ provided $\F_{\alpha\beta}\neq 0$. The Petrov classification of (\ref{weylSi}) is very easy. In this case, the eigenvalues of the endomorphism (\ref{endom}) are $2/3 Q(p) \bm{{\cal F}^2}|_p$ (with eigenvector $\F_{\alpha\beta}|_p$) and $-1/3 Q(p) \bm{{\cal F}^2} |_p$ (the eigenspace of which is the set of self-dual two forms ${\cm X}$ satisfying ${\cm X}_{\alpha\beta} \F^{\alpha\beta} |_p = 0$). Hence, the Weyl tensor (\ref{weylSi}) is of Petrov type D whenever $Q \bm{{\cal F}^2} |_p \neq 0$, type N when $\bm{{\cal F}^2} |_p =0, Q(p) \neq 0$ and Type 0 when $Q(p)=0$. More importantly, the principal null directions of the Weyl tensor coincide exactly with the principal null directions of $\F_{\alpha\beta}$ (whenever $Q(p) \neq 0$). At points where $\F_{\alpha\beta}$ is regular, each principal null direction of $\F_{\alpha\beta}$ is a double principal null direction of the Weyl tensor and at points where $\F_{\alpha\beta}$ is singular, the principal null direction of $\F_{\alpha\beta}$ is a quadruple principal null direction of the Weyl tensor. Hence, the Kerr metric is characterized (among asymptotically flat vacuum solutions with a Killing field which tends to a time translation at infinity) by the fact that the principal null directions of the Weyl tensor coincide with the principal null directions of the Killing form $\F_{\alpha\beta}$. This also shows why the Kerr metric enjoys such a privileged position among asymptotically flat stationary vacuum solutions. The geometrical simplicity of this characterization also indicates that the Kerr metric is, in that sense, the simplest possible asymptotically flat stationary vacuum metric. A systematic study of the Killing form $\F_{\alpha\beta}$ for electrovacuum spacetimes admitting a Killing field has been carried out recently by F.Fayos and C.F.Sopuerta \cite{PS}, who have found a way to determine the principal null directions of $\F_{\alpha\beta}$ and of the electromagnetic field. Several conditions on the principal null directions can then be imposed, thus restricting the possible geometries. For instance, the condition that the principal null directions of $\F_{\alpha\beta}$ are aligned with the principal null directions of the electromagnetic field can be imposed. Similarly, conditions like some (or all) principal null directions of $\F_{\alpha\beta}$ being shear-free and/or geodesic can also be studied. A particular interesting case would be studying which electrovacuum spacetimes admitting a Killing vector field are allowed such that the principal null directions of $\F_{\alpha\beta}$ and those of the electromagnetic field are geodesic and shear-free. From the Goldberg-Sachs theorem, these null directions must then be repeated principal null directions of the Weyl tensor, so that the Petrov type is D, N or 0 (depending on the degree of degeneracy). A straightforward calculation shows that the Kerr-Newman metric has Petrov type D with the two repeated principal null directions aligned with the principal null directions of $\F_{\alpha\beta}$ (as in the Kerr metric) {\it and} also aligned with the principal null directions of the electromagnetic field. It is a matter for future research to analyze whether the converse is also true (after assuming asymptotic flatness) in order to obtain a characterization of the Kerr-Newman spacetime. \section*{Acknowledgements} I am very grateful to Bobby Beig and Walter Simon for helpful discussions and for a careful reading of the manuscript. I would also like to thank Jos\'e M.M. Senovilla for comments on a previous version. I wish to thank the European Union for financial support under the Marie Curie fellowship ERBFMBICT972520.
\section{Introduction} The Sine-Gordon (SG) model in two-dimensional Minkowski space-time is described by the action\footnote{ We have rescaled $\beta \to \sqrt{4\pi} \beta $ in comparison with the usual convention, so that the free fermion (FF) point occurs at $\beta^2 = 1$.} \begin{equation}} \def\bn{\beq\label{SinGor} S_{\rm SG}=\frac{1}{4\pi} \int dt\,dx\ \sk{ \frac{1}{2} \frac{\partial^2\Phi(x,t)} {\partial t^2} -\frac{1}{2} \frac{\partial^2\Phi(x,t)} {\partial x^2} +\frac{m^2}{\beta^2}\ \sk{\cos(\beta\Phi(x,t))-1} }. \end{equation}} \def\ed{\eeq The quantum SG theory is perhaps the most fundamental of the integrable quantum field theories in two dimensions, and thus plays an important role in the development of new methods. The $S$-matrix of soliton-antisoliton scattering was obtained in \cite{ZZ}. This $S$-matrix (see \r{S-mat}) depends on so called renormalized coupling constant $\xi$ and the relation of this parameter to the SG coupling constant $\beta$ is \begin{equation}} \def\bn{\beq\label{ren-coup} \xi=\frac{\beta^2}{2-\beta^2}\ . \end{equation}} \def\ed{\eeq The quantum SG model is a superrenormalizable theory for the real values of the coupling constant $0<\beta^2<2$ which corresponds to the restriction to the real positive values of $\xi$, $0<\xi<\infty$. The regime $1<\xi<\infty$ is the breatherless one, where solitons and antisolitons do not form bound states. One can see that modulo the overall scalar factor and for appropriate choice of the multiplicative spectral parameter $z=e^{-\theta/\xi}$, where $\theta$ is a rapidity of the particles, the soliton-antisoliton $S$-matrix can be written in the form \begin{equation}} \def\bn{\beq\label{s-mul} S(\theta,\xi)=\rho(\theta,\xi) \left(\begin{array}{cccc} {zq-z^{-1}q^{-1}}&0&0&0\\ 0&{z-z^{-1}}&{q-q^{-1}}&0\\ 0&{q-q^{-1}}&{z-z^{-1}}&0\\ 0&0&0&{zq-z^{-1}q^{-1}} \end{array}\right),\quad q=\exp\sk{\pi i\frac{\xi+1}{\xi}} \end{equation}} \def\ed{\eeq which signifies in particular a quantum group symmetry of the Hilbert space of states of the model with respect to the finite dimensional quantum group $U_q(sl_2)$ \cite{RS,AL}. The SG model was also one of the first continuous integrable models where the quantum inverse scattering method (QISM) was tested. It was shown in the paper \cite{FST} that the quantum monodromy matrices ${\cal T}(u)$ satisfy the commutation relation \begin{equation}} \def\bn{\beq\label{YBt} R(u_1-u_2,\xi){\cal T}_1(u_1){\cal T}_2(u_2)={\cal T}_2(u_2){\cal T}_1(u_1)R(u_1-u_2,\xi)\ , \end{equation}} \def\ed{\eeq where the $R$-matrix has the same structure as in \r{s-mul} (see \r{RR-mat} for the exact formula) in terms of additive an spectral parameter $u$ $\sk{z=e^{-u/(\xi+1)}}$, but with deformation parameter replaced by \begin{equation}} \def\bn{\beq\label{q'-def} q'=\exp\sk{\pi i\frac{\xi}{\xi+1}}. \end{equation}} \def\ed{\eeq The equation \r{YBt} implies that \begin{equation}} \def\bn{\beq\label{trace} [{\rm tr}\,{\cal T}(u_1),{\rm tr}\,{\cal T}(u_2)]=0 \end{equation}} \def\ed{\eeq and signifies that after proper expansion of the quantity ${\rm tr}\,{\cal T}(u)$ with respect to the spectral parameter $u$ it generates the local integrals of motion and \r{trace} shows that they are in involution. Note that even at the FF point where $\beta^2=1$ the $R$-matrix in the commutation relation of monodromy matrices \r{YBt} is nontrivial since $q'=i$; this can be traced to the fact that the monodromy matrix is constructed from the fields $\exp(i\Phi /2)$, which are non-local in terms of the fermions since the fermion bilinear is $\exp(i\Phi )$. As we see, the SG model naturally contains two quantum group symmetries, with different deformation parameters related by the duality transformation \r{dual-tr}. An attempt to explain this phenomena was made in \cite{Lu1} in the framework of the bosonization technique in massive integrable field theories. This approach was generalized then for the lattice integrable models \cite{FIJKMY}. Following the ideas presented in these papers a screening current algebra was proposed in \cite{KLP}. The specific coalgebraic properties of this infinite-dimensional algebra allowed to reconstruct the bosonization approach of \cite{Lu1} from algebraical analysis of the representation theory of the screening current algebra. Essential progress toward understanding quantum integrable models in the infinite volume limit was made in the framework of Baxter's corner transfer matrix (CTM) method \cite{B}. It was observed that the CTM of some lattice integrable models in the infinite volume limit has equidistant spectrum bounded from below and so can be described by the infinite set of oscillators. This fact allows one to develop a new approach to quantum integrable models on the lattice. This was done by the Kyoto group for the XXZ model in the anti-ferroelectric regime \cite{JM}. The model was completely solved, namely, the correlation functions of local operators and form-factors of local operators were calculated explicitly, using infinite-dimensional representations of quantum affine algebra $U_{q}(\widehat{sl}_2)$ with real parameter of deformation satisfying $-1<q<0$. One of the main ideas of the construction is to divide the total Hilbert space of the model, which is identified in the infinite volume limit with an infinite product of two-dimensional spaces where local operators act, \begin{equation}} \def\bn{\beq\label{hilXXZ} \CH_{\rm XXZ}\approx \cdots\Bbb{C}^2\otimes\Bbb{C}^2\otimes\Bbb{C}^2\otimes\Bbb{C}^2\otimes\Bbb{C}^2\otimes\Bbb{C}^2\cdots \end{equation}} \def\ed{\eeq into two semi-infinite products of these spaces \begin{equation}} \def\bn{\beq\label{hilXXZ1} \CH_{\rm XXZ}\approx\bigr(\cdots\Bbb{C}^2\otimes\Bbb{C}^2\otimes\Bbb{C}^2\bigl)\otimes \bigr(\Bbb{C}^2\otimes\Bbb{C}^2\otimes\Bbb{C}^2\cdots\bigl)\approx \CH^*_{\rm CTM}\otimes\CH_{\rm CTM}={\rm End}\sk{\CH_{\rm CTM}} \end{equation}} \def\ed{\eeq which are denoted by $\CH_{\rm CTM}$ and where the corner transfer matrix acts naturally. Each of these semi-infinite products is identified with level 1 and level $-1$ integrable modules of $U_{q}(\widehat{sl}_2)$, where operators of the algebra act naturally. The decomposition \r{hilXXZ1} led in particular to the identification of the states in the Hilbert space $\CH_{XXZ}$ with the operators acting in $\CH_{CTM}$. The space ${\rm End}\sk{\CH_{\rm CTM}}$ is equipped with a natural scalar product $(A,B)={\rm Tr}\,_{\CH_{\rm CTM}}AB$ and the vacuum vector in $\CH_{\rm XXZ}$ is defined as $(-q)^ {H_{\rm CTM}}$, where $H_{\rm CTM}$ is a corner transfer matrix hamiltonian. The representation theory of the quantum affine algebra $U_{q}(\widehat{sl}_2)$ provides certain operators which intertwine its action in $\CH_{\rm CTM}$ (type I and type II intertwining operators). Type II intertwining operators are used for the construction of the basis of asymptotic states in ${\rm End}\sk{\CH_{\rm CTM}}$, and type I operators are used for the construction of the transfer matrix and the local hamiltonian in this picture. Moreover, the adjoint action of the elements of the quantum affine algebra in ${\rm End}\sk{\CH_{\rm CTM}}$ describe a level 0 $U_{q}(\widehat{sl}_2)$ symmetry of the model. As a consequence, the form-factors of the local operators and correlation functions of their product are presented in a form of certain multiple integrals, which come as a trace over $\CH_{\rm CTM}$ of certain products of the intertwining operators. In the continuous integrable models an approach to implement Baxter's CTM method was developed in the papers \cite{Lu1,BrL} and was based on the method of the angular quantization. The total Hilbert space of the continuous quantum integrable model in infinite volume was supposed to be embedded into a tensor product \begin{equation}} \def\bn{\beq\label{decomp} \CH \hookrightarrow \CH_L \otimes \CH_R\ , \end{equation}} \def\ed{\eeq where $\CH_L$ ($\CH_R$) are the Hilbert spaces of the quantization in the left (right) wedge. The right Rindler wedge (RRW) in two-dimensional Minkowski space-time is \begin{equation}} \def\bn{\beq\label{rrw} (x^0)^2-(x^1)^2<0,\quad x^1 >0\ , \end{equation}} \def\ed{\eeq where $x^0$ is a time and $x^1$ is coordinate, while the left Rindler wedge (LRW) \begin{equation}} \def\bn{\beq\label{lrw} (x^0)^2-(x^1)^2<0,\quad x^1 <0\ . \end{equation}} \def\ed{\eeq Let us fix the parametrization of space-time coordinates in RRW \begin{equation}} \def\bn{\beq\label{para} x^0=r\,{\rm sh}\,\alpha, \quad x^1=r\,{\rm ch}\,\alpha,\quad r\geq0,\quad \alpha\in\Bbb{R}\ . \end{equation}} \def\ed{\eeq With this parametrization, the coordinates $x^0, x^1$ cover the RRW, since $x^1 >0$. The LRW is formally obtained by the rotation $\alpha \to \alpha -i\pi$ or by applying the operator $e^{\pi K}$ where $K$ is Lorentz boost generator $K = -i \partial_\alpha$. The space $\CH_L$ can be identified with the dual to $\CH_R$ and so the states in the total Hilbert space can be realized as the operators in $\CH_R$. It was suggested in \cite{Lu1} to realize $\CH_R$ for the SG model as a Fock space with a natural action of the operators satisfying the commutation relations of Zamolodchikov-Faddeev (ZF) algebra. Further, in \cite{KLP} these operators were identified with intertwining operators of the scaled elliptic algebra ${\cal A}(\widehat{sl}_2)$ which can be observed in the bosonization picture \cite{Lu1} by the presentation using screening currents. One of the main arguments in favor of these mathematical constructions was the coincidence of form factors of certain local operators in SG theory with trace calculations in $\CH_R$. In this paper, we try to develop the method of the angular quantization in two directions. First, we analyze the SG model in RRW at the free fermion point, where the canonical quantization can be done explicitly. We see here that the usual conserved charges \cite{Le} diverge and the only chance to get a rich algebra of symmetries is to use a certain analytical continuation of the conserved charges, or equivalently, the scattering data. In this case the bosonization \cite{Lu1} naturally appears. We see further that in order to close the algebra, we are forced to use the currents with dual monodromy properties and the algebra of (nonlocal) conserved currents which we find here coincide with specialization of the scaling elliptic algebra ${\cal A}(\widehat{sl}_2)$ proposed in \cite{KLP} and specialized to the free fermion point ($\xi=1$). Second, we go into further details of the description of the continuous SG model analogous to the group-theoretical description of the space of states in the XXZ model \cite{JM}. We show that starting from level one representation of the scaling elliptic algebra ${\cal A}(\widehat{sl}_2)$ we can correctly define the vacuum, the asymptotic states and operators which act on the space of the asymptotic states, namely, the transfer matrix, the hamiltonian, the local integrals of motion. Contrary to the lattice case they are given now via coefficients of the asymptotic expansion of the family of commuting operators. We define the adjoint action of the algebra ${\cal A}(\widehat{sl}_2)$ on the space of states and show that known symmetries of this space related to the conserved nonlocal currents \cite{L,BL} and formulated in terms of quantum affine algebras at level zero, can be obtained from this adjoint action by the asymptotical expansion. Let us roughly explain this description. The total Hilbert space $\CH$ is supposed to be divided as in \r{decomp}. The spaces $\CH_R$ and $\CH_L$ are level 1 and level $-1$ highest weight modules over the algebra ${\cal A}(\widehat{sl}_2)$ so the states in $\CH$ can be identified with some operators in $\CH_R$. In particular, the physical vacuum state $|\mbox{vac}\rangle_{\rm ph}$ is identified with boost operator \begin{equation}} \def\bn{\beq\label{vacph} |\mbox{vac}\rangle_{\rm ph}=e^{\pi K}=e^{-i\pi\partial_\alpha}, \end{equation}} \def\ed{\eeq where $\alpha$ is angular time in RRW and the states $|\theta_1,\ldots,\theta_n\rangle_{\varepsilon_1,\ldots,\varepsilon_n}$ are identified with the product \begin{equation}} \def\bn{\beq\label{stat} |\theta_1,\ldots,\theta_n\rangle_{\varepsilon_1,\ldots,\varepsilon_n}= Z^*_{\varepsilon_1}(\theta_1)\ldots Z^*_{\varepsilon_n}(\theta_n)e^{\pi K}\ , \end{equation}} \def\ed{\eeq where $Z^*_{\varepsilon}(\theta)$ are certain `twisted' intertwining operators of the screening currents algebra ${\cal A}(\widehat{sl}_2)$, which also act in $\CH_R$. The adjoint action of the algebra ${\cal A}(\widehat{sl}_2)$ is not standard because this algebra is not a Hopf algebra. Indeed, the commutation and comultiplication relations of the algebra ${\cal A}(\widehat{sl}_2)$ in terms of $L$-operators can be written in the form \begin{equation}} \def\bn{\beq\label{scr-com} \CR(u_1-u_2,\xi+c)L_1(u_1,\xi)L_2(u_2,\xi)=L_2(u_2,\xi)L_1(u_1,\xi) \CR(u_1-u_2,\xi) \end{equation}} \def\ed{\eeq \begin{equation}} \def\bn{\beq\label{scr-coal} \Delta^{\rm op} L(u,\xi)=L(u- i\pi c^{(2)}/4,\xi+c^{(2)}) \ {\dot\otimes}\ L(u+ i\pi c^{(1)}/4,\xi) \nonumber} \def\gamma{\gamma \end{equation}} \def\ed{\eeq where $\CR(u,\xi)$ means $R$-matrix defined by \r{R-mat} and $c$ is a central element of the algebra ${\cal A}(\widehat{sl}_2)$. Note that $R$-matrices in the left and right hand sides of \r{scr-com} differ by the central element of the algebra, which signifies that the algebra under consideration is not coassociative. This algebra is not a usual Hopf algebra. Nevertheless, a coalgebraic structure of this algebra was used in \cite{KLP} to construct the intertwining operators for highest weight modules over this algebra at the value of the central element $c=1$. There are also some indications that this screening current algebra is a quasi-Hopf algebra \cite{Drquasi} (see papers \cite{JKOS1-2} on the lattice variants of this algebra). The adjoint action has the different form on the subspaces $\CH_i\in\CH$, $i=0,1$ of even and odd number of particles and includes the involution of the algebra ${\cal A}(\widehat{sl}_2)$ \begin{equation}} \def\bn{\beq \iota \sk{L(u)}= \sigma_z L(u) \sigma_z\ . \label{L-oper-rel} \end{equation}} \def\ed{\eeq For the state $X_i\in\CH_i$, $i=0,1$ it is defined as follows \begin{equation}} \def\bn{\beq {\rm Ad}_{L(u;\xi)}\cdot X_{k}= \iota\sk{L^{-1}(u+i\pi c/4;\xi)}\,X_{k}\,\iota^{k+1} \sk{L(u-i\pi +i\pi c/4;\xi)}\ . \label{adj} \end{equation}} \def\ed{\eeq We prove that so defined adjoint action realizes the level zero representation of the algebra ${\cal A}(\widehat{sl}_2)$ onto the space of states $\CH$, such that $n$-particle states compose $n$-fold tensor products of two-dimensional representations. The quantum affine symmetry of the Hilbert space $\CH$ found in \cite{BL} can be realized via the asymptotical expansion of the adjoint action of the currents of the algebra ${\cal A}(\widehat{sl}_2)$. The paper is organized as follows. In the second section we consider the canonical quantization of the SG model at FF point and its specialization to right Rindler wedge. Then we construct the nonlocal integrals of motion and develop bosonization of all the objects in terms of these integrals of motion. The third section is devoted to the description of the screening currents algebra for general value of SG coupling constant satisfying $1<\beta<\sqrt{2}$. In the last section we develop the angular quantization scheme in full aspect; for example, we construct the monodromy matrix on the total Hilbert space and investigate some of its properties. \setcounter{equation}{0} \section{Canonical quantization} \subsection{Sine-Gordon model at free fermion point} It is well known \cite{Col} that the SG model model with the action \r{SinGor} is equivalent on the quantum level to the massive Thirring model defined by the action \begin{equation}} \def\bn{\beq\label{Thir} S_{\rm Th}=\int dx^0dx^1\ \left[{1\over 2}\left(\overline{\Psi}(x) i\gamma^\mu\df \mu \Psi(x)- \df \mu\overline{\Psi}(x) i\gamma^\mu \Psi(x)\right)-m\overline{\Psi}(x)\Psi(x) -\frac{g}{2}\sk{\overline{\Psi}(x)\gamma^\mu\Psi(x)}^2 \right] \end{equation}} \def\ed{\eeq where $g=\frac{\pi(1-\xi)}{2\xi}$. The equivalence is established by the following bosonization rules \begin{equation}} \def\bn{\beq\label{S-T} \frac{\beta}{2\pi}\varepsilon^{\mu\nu}\partial_{\nu}\Phi(x^0,x^1)= \overline{\Psi}(x^0,x^1)\gamma^\mu\Psi(x^0,x^1),\quad \cos(i\beta\Phi(x^0,x^1))= \overline{\Psi}(x^0,x^1)\Psi(x^0,x^1) \end{equation}} \def\ed{\eeq where $\varepsilon^{\mu\nu}$ is antisymmetric tensor normalized $\varepsilon^{01}=1$. At the FF point ($\xi=1$) the interaction in the Thirring model vanishes and its lagrangian becomes a lagrangian of free massive Dirac fermions \begin{equation}} \def\bn{\beq {\cal L}(x^0,x^1)= \left[{1\over 2}\left(\overline{\Psi}(x) i\gamma^\mu\df \mu \Psi(x)- \df \mu\overline{\Psi}(x) i\gamma^\mu \Psi(x)\right)-m\overline{\Psi}(x)\Psi(x)\right], \end{equation}} \def\ed{\eeq where $\overline{\Psi}(x)=\Psi^\dag(x)\gamma^0$ is a Dirac conjugated spinor. We fix the $\gamma$-matrices to be $$ \gamma^0=\left(\begin{array}{cc} 0& -i\\ i&0\end{array}\right), \quad \gamma^1=\left(\begin{array}{cc} 0& i\\ i&0\end{array}\right). $$ The complete set of solutions to the corresponding linear equation of motion can be chosen in the form \begin{equation}} \def\bn{\beq \Psi_\theta(x^0,x^1)=\sqrt{m\over 2}\spin {e^{\theta/ 2}}{ie^{-\theta/2}} e^{-im{\rm ch}(\theta)x^0+im{\rm sh}(\theta)x^1} \label{ps} \end{equation}} \def\ed{\eeq and satisfies the completeness relation with respect to the scalar product \begin{equation}} \def\bn{\beq (\Psi_{\theta},\Psi_{\theta^\prime})= \int_{-\infty}^\infty dx^1 \overline{\Psi}_\theta(x^0,x^1)\gamma^0\Psi_{\theta^\prime}(x^0,x^1)=\delta(\theta-\theta^\prime). \label{nops} \end{equation}} \def\ed{\eeq Note that the solutions \r{ps} are anti-periodic with respect to the shift $\theta\to\theta+2\pi i$. The completeness relation \r{nops} allows one to quantize the Dirac field $\Psi(x)\equiv\Psi_+(x)$ and its hermitian conjugate $\Psi^\dag_+(x)\equiv\Psi_-(x)$ \begin{eqnarray} \Psi_+(x^0,x^1)=\int_{-\infty}^\infty d\theta\left[ c(\theta)\Psi_\theta(x) +d^{\dag}(\theta)\Psi^*_\theta(x)\right], \label{ts1+} \end{eqnarray} \begin{eqnarray} \Psi_-(x^0,x^1)=\int_{-\infty}^\infty d\theta\left[ d(\theta)\Psi_\theta(x) +c^{\dag}(\theta)\Psi^*_\theta(x)\right], \label{ts1-} \end{eqnarray} by imposing equal time anticommutation relations \begin{equation}} \def\bn{\beq\label{commm} \{\psi_+(x^0,x),\psi_-(x^0,x')\}=2\pi\delta(x-x'),\quad \{\overline\psi_+(x^0,x),\overline\psi_-(x^0,x')\}=2\pi\delta(x-x'), \end{equation}} \def\ed{\eeq where $\psi_\pm(x^0,x^1)$ and $\overline\psi_\pm(x^0,x^1)$ are components of spinor $\Psi_\pm(x^0,x^1)$. One can verify now that \r{commm} and the normalization condition \r{nops} imply the standard anticommutation relations \begin{equation}} \def\bn{\beq \{c(\theta),c^\dag(\theta')\}=\delta(\theta-\theta'), \quad \{d(\theta),d^\dag(\theta')\}=\delta(\theta-\theta') \label{A+A-} \end{equation}} \def\ed{\eeq and all others are trivial. The Hilbert space $\CH$ of this model is defined by application of the creation operators $c^\dag(\theta)$ and $d^\dag(\theta)$ to the `physical' vacuum vector $|\mbox{vac}\rangle_{\rm ph}$ annihilated by the operators $c(\theta)$ and $d(\theta)$. The integrals of motion are constructed from the conserved currents: $$ \dif{J_y}{\overline y}-\dif{J_{\overline y}}{y}=0 $$ and are given by the contour integral \begin{equation}} \def\bn{\beq\label{charges} Q^J=\int \sk{dy\ J_y + d\overline y\ J_{\overline y}}\ . \end{equation}} \def\ed{\eeq The charge $Q^J$ is conserved along the evolution which is `orthogonal' to the contour in the definition of $Q^J$. In the standard quantization picture this contour is chosen to be equal time line $x^0=\hbox{const}$ in the space-time. The charges which are conserved along the evolution with respect to the time $x^0$ can be obtained from the currents \begin{eqnarray} J_y^\pm&=&(D \psi_\pm) \psi_\pm,\quad J_{\overline y}^\pm=(D\overline\psi_\pm) \overline\psi_\pm,\label{iii1}\\ J_y^0&=&(D\psi_- )\psi_+,\quad J_{\overline y}^0=(D\overline\psi_- ) \overline\psi_+,\label{iii2} \end{eqnarray} or equivalently from \begin{eqnarray} \widetilde J_y^\pm&=&\psi_\pm (D \psi_\pm),\quad \widetilde J_{\overline y}^\pm=\overline\psi_\pm (D \overline\psi_\pm),\label{ii1}\\ \widetilde J_y^0&=&\psi_- (D \psi_+),\quad \widetilde J_{\overline y}^0=\overline\psi_- (D \overline\psi_+),\label{ii2} \end{eqnarray} for the operator $D=\partial_y^n$ or $\partial_{\overline y}^n$, $n\geq0$. Let us denote the neutral charges which correspond to the first operator as $I_n$ and to the second as $\overline I_n$. The charged conserved quantities we denote by $I^\pm_n$ and $\overline I^\pm_n$ respectively. They have the explicit expressions in terms of the operators acting in total Hilbert space: \begin{equation}} \def\bn{\beq\label{int-n} I_n=m^n\int_{-\infty}^\infty d\theta\ e^{n\theta}\sk{c^\dag(\theta)c(\theta)+ d^\dag(\theta)d(\theta)},\quad \overline I_n=m^n\int_{-\infty}^\infty d\theta\ e^{-n\theta}\sk{c^\dag(\theta)c(\theta)+ d^\dag(\theta)d(\theta)} \end{equation}} \def\ed{\eeq and similar formulas for $I^\pm_n$ and $\overline I^\pm_n$. The Hamiltonian $H$ which describes the evolution of the quantum fields \r{ts1+} and \r{ts1-} with respect to the time $x^0$ is given by the sum $(I_1+\overline I_1)/2$ and has eigenvalue $m{\rm ch}\,\theta$ on the one-particle states generated by $c^\dag(\theta)$ and $d^\dag(\theta)$ from physical vacuum $|\mbox{vac}\rangle_{\rm ph}$. However there is no a direct way to quantize the SG field $\Phi(x^0,x^1)$ at the FF point using the quantization of the Dirac fermion fields $\Psi(x^0,x^1)$. In particular, it is difficult to construct the realization of the commutation relations \r{YBt} directly in the infinite volume limit using the canonical anticommutation relations \r{commm} and without referring to the lattice regularization. On the other hand the canonical quantization of the free massive Dirac fermions in RRW allows one to construct the operators which are building blocks of the angular quantization method. This will be done in the next subsections with the main goal being to demonstrate the nonabelian symmetry algebra which appears in the angular quantization approach to the SG model. \subsection{Free fermions in Rindler wedge} Let us solve the equation of motion for free massive Dirac field in RRW using the parametrization \r{para}. The solution to the Dirac equation of motion normalized with respect to the scalar product \begin{eqnarray} (\Psi,\Psi')=\int_{0}^\infty dr\ \sk{e^{-\alpha}\overline \psi\overline \psi'+e^{\alpha} \psi\psi'}\ ,\quad \Psi=\sk{\begin{array}{c}\overline \psi\\ \psi \end{array}} \label{scalar} \end{eqnarray} is given in terms of MacDonald func\-tions \begin{equation}} \def\bn{\beq \Psi_\nu(r,\alpha)={\sqrt{m}\over\sqrt{\pi} \Gamma(i\nu+1/2)} \spin{{\bf e}^{(2i\nu+1/2)\alpha} K_{i\nu+1/2}(mr)} {{\bf e}^{(2i\nu-1/2)\alpha} K_{i\nu-1/2}(mr)},\quad \nu\in\Bbb{R}\ , \label{sol} \end{equation}} \def\ed{\eeq and has exponentially decreasing asymptotics in RRW when $r\to\infty$. On the other hand we observe that rotation of angular time $\alpha$ by $2\pi i$ which corresponds to the path around origin in euclidean plane multiplies the solution \r{sol} by the factor $-e^{-2\pi\nu}$. This signifies the fact that the space of functions used in canonical and angular quantization are completely different. Nevertheless, the completeness relation \begin{equation}} \def\bn{\beq (\Psi_\nu,\Psi_{\nu'})=\delta(\nu+\nu') \label{orth} \end{equation}} \def\ed{\eeq allows one to quantize the Dirac fields in RRW \begin{equation}} \def\bn{\beq \Psi_\pm(r,\alpha)=\spin{\overline\psi_\pm(r,\alpha)}{\psi_\pm(r,\alpha)}=\int_{-\infty}^\infty d\nu\ b_\pm(\nu)\Psi_\nu(r,\alpha) \label{rw+} \end{equation}} \def\ed{\eeq by imposing the equal `time' ($\alpha=\hbox{const}$) anticommutation relations \begin{equation}} \def\bn{\beq \{\psi_+(r,\alpha),\psi_-(r',\alpha)\}=-{\bf e}^{-\alpha}\delta(r-r'),\quad \{\overline\psi_+(r,\alpha),\overline\psi_-(r',\alpha)\}=-{\bf e}^{\alpha}\delta(r-r') \label{coa} \end{equation}} \def\ed{\eeq which are equivalent to \begin{equation}} \def\bn{\beq \{b_\pm(\nu),b_\pm(\nu')\}=0,\quad \{b_+(\nu),b_-(\nu')\}=\delta(\nu+\nu')\ . \label{stand} \end{equation}} \def\ed{\eeq Rindler fermionic Fock space $\CH_R^f$ is defined by the vacuum state $|\mbox{vac}\rangle_f$ which satisfies \begin{equation}} \def\bn{\beq\label{Fock} b_\pm(\nu)|\mbox{vac}\rangle_f=0,\quad \nu>0\ . \end{equation}} \def\ed{\eeq The left vacuum vector $_f\langle\mbox{vac}|$ is correspondingly defined: \begin{equation}} \def\bn{\beq\label{lFock} _f\langle\mbox{vac}| b_\pm(\nu)=0,\quad \nu<0\ . \end{equation}} \def\ed{\eeq \subsection{Scattering transform} For the quantum fermionic fields \r{rw+} we introduce the scattering transform \cite{Lu-pri} \begin{equation}} \def\bn{\beq\label{sc-tr} \Psi(r,\alpha)\to\Lambda_\pm(\theta,\alpha)= \frac{\sqrt{m}}{2\sqrt\pi}\ \int_{0}^\infty dr\ e^{-mr{\rm ch}\theta}\sk{ \overline\psi_\pm(r,\alpha)e^{(\theta-\alpha)/2}+\psi_\pm(r,\alpha)e^{(\alpha-\theta)/2}} \ , \end{equation}} \def\ed{\eeq where $\theta\in\Bbb{C}$ is spectral parameter. Using the free fermion equation of motion in RRW we can verify that the dependence of the operators $\Lambda_\pm$ on the angular time $\alpha$ reduces to a simple shift of the spectral parameter \begin{equation}} \def\bn{\beq\label{stpr} \Lambda_\pm(\theta,\alpha)=\Lambda_\pm(\theta+\alpha)\ , \end{equation}} \def\ed{\eeq where $\Lambda_\pm(\theta)$ is the value of the scattering transform at the initial time, say $\alpha=0$. It is clear that the scattering transform \r{sc-tr} is not defined for all values of the spectral parameter $\theta$. For example, if the solutions $\overline\psi_\pm(r,\alpha)$ and $\psi_\pm(r,\alpha)$ have the constant asymptotics when $r\to\infty$ then the integral in \r{sc-tr} is convergent if $|{\rm Im}\,\,\theta|<\pi/2$, which follows from the inequality ${\rm Re}\,\,{\rm ch}\,\theta>0$. However, the solutions \r{sol} of the Dirac equation in RRW have exponentially decreasing asymptotics. Using the fact that the leading term of the asymptotic of the MacDonald function $K_x(z)$ when $z\to\infty$ does not depend on the index $x$ and is proportional to $z^{-1/2}e^{-z}$ we find that the inequality mentioned above is replaced by the more weak inequality \begin{equation}} \def\bn{\beq\label{ineq} {\rm Re}\,\,{\rm ch}\,\theta>-1\ . \end{equation}} \def\ed{\eeq The solution of \r{ineq} defines a larger domain of existence of the scattering transform than specified above, namely \begin{equation}} \def\bn{\beq\label{dom1} |{\rm Im}\,\,\theta|<\pi/2+\epsilon,\quad\hbox{where}\quad \epsilon=\pi/2-\arccos\sk{\sk{{\rm ch}\,{\rm Re}\,\,\theta}^{-1}} \end{equation}} \def\ed{\eeq so the domain of the possible values of the spectral parameter is a strip whose width depends on the value of ${\rm Re}\,\,\theta$. An important consequence of this observation is the fact that the points ${\rm Im}\,\,\theta=\pm\pi/2$ are {\it always} in the domain of existence of the scattering transform. This leads to the fact that the vacuum expectation value of the product $\Lambda_\pm(\theta) \Lambda_\mp(\theta')$ is a well defined meromorphic function in the domain $|{\rm Im}\,\,(\theta-\theta')|\leq \pi+\varepsilon$ for some positive number $\varepsilon$. Using the expression of the scattering data operators $\Lambda_\pm(\theta)$ in terms of the fermionic operators $b_\pm(\nu)$ \begin{equation}} \def\bn{\beq\label{zz5} \Lambda_\pm(\theta)= \frac{1}{2\pi}\int_{-\infty}^\infty d\nu\ b_\pm(\nu)\Gamma\sk{\frac{1}{2} -i\nu}e^{i\nu\theta} \end{equation}} \def\ed{\eeq we can calculate this function explicitly: \begin{equation}} \def\bn{\beq\label{vev-sc} \langle\Lambda_\pm(\theta)\Lambda_\mp(\theta')\rangle= \frac{1}{4\pi^2}\beta\sk{\frac{\theta-\theta'+\pi i}{2\pi i}}, \end{equation}} \def\ed{\eeq where the $\beta$-function is $\beta(x)= \partial_x\ln\sk{\left.\Gamma\sk{\frac{x+1}{2}} \right/\Gamma\sk{\frac{x}{2}}}$. Since the function \r{vev-sc} has the poles only in the points $\theta=\theta'-\pi i(2k+1)$, $k=0,1,\ldots$ the domain of the `existence' of this function can be extended to ${\rm Im}\,(\theta-\theta')>-3\pi$ with a simple pole at the point $\theta=\theta'-\pi i$. The function $\langle\Lambda_\pm(\theta)\Lambda_\mp(\theta')\rangle$ is given by the meromorphic function \r{vev-sc} in this domain. An immediate consequence of this fact is the anticommutation relation \begin{equation}} \def\bn{\beq\label{alLam} \{\Lambda_+(\theta),\Lambda_-(\theta')\}=\frac{\pi}{{\rm ch}((\theta-\theta')/2)}, \quad |{\rm Im}\,(\theta-\theta')|<3\pi\ . \end{equation}} \def\ed{\eeq The scattering transform \r{sc-tr} describes an evolution of the initial data (the quantum fields $\Psi_\pm(r,0)$ at initial value of the angular time $\alpha=0$) with respect to this angular time. Since this evolution reduces to a simple shift in the spectral parameter one can easily restore the quantum fields $\Psi_\pm(r,\alpha)$ at arbitrary time $\alpha$ by solving inverse scattering problem, restoring quantum fields $\Psi_\pm(r,\alpha)$ from the operators $\Lambda_\pm(\theta)$. This can be done using the operators ${\cal Z}_\pm(\theta)$ related to the operators $\Lambda_\pm(\theta)$ by the integral transform \begin{equation}} \def\bn{\beq\label{ZL7} \Lambda_\pm(\theta)=\frac{1}{4\pi}\int_{-\infty}^\infty d\theta'\ \frac{{\cal Z}_\pm(\theta')} {{\rm ch}((\theta-\theta')/2)} \end{equation}} \def\ed{\eeq which can be inverted as follows \begin{equation}} \def\bn{\beq\label{LZ7} {\cal Z}_\pm(\theta)=\Lambda(\theta+\pi i)+\Lambda(\theta-\pi i)\ . \end{equation}} \def\ed{\eeq The inverse scattering problem has a solution in terms of the operators ${\cal Z}_\pm(\theta)$: \begin{equation}} \def\bn{\beq\label{psi-Z} \Psi_\pm(r,\alpha)={\sqrt{m}\over2\sqrt{\pi}}\int_{-\infty}^\infty d\theta\ {\cal Z}_\pm(\theta+\alpha)e^{-mr{\rm ch}\,\theta} \spin{e^{(\theta+\alpha)/2}}{e^{-(\theta+\alpha)/2}}. \end{equation}} \def\ed{\eeq One can observe that substitution of \r{psi-Z} into \r{sc-tr} leads to the integral transform \r{ZL7}. Formula \r{LZ7} allows one to calculate the vacuum expectation values of the different operators \begin{eqnarray} \langle \Lambda_\pm(\theta_1) {\cal Z}_\mp(\theta_2)\rangle &=& \langle {\cal Z}_\pm(\theta_1) \Lambda_\mp(\theta_2)\rangle= -\frac{1}{2\pi i}\ \frac{1}{\theta_1-\theta_2}, \quad {\rm Im}\,(\theta_1-\theta_2)>-2\pi\ ,\nonumber} \def\gamma{\gamma\\ \langle {\cal Z}_\pm(\theta_1) {\cal Z}_\mp(\theta_2)\rangle &=& -\frac{1}{2\pi i} \left[ \frac{1}{\theta_1-\theta_2-i\pi}+\frac{1}{\theta_1-\theta_2+i\pi} \right],\quad {\rm Im}\,(\theta_1-\theta_2)>-\pi\ . \label{vev} \end{eqnarray} These formulas allow to verify that the canonical anticommutation relations \r{coa} follow from the solution of the inverse scattering problem \r{psi-Z}. The formula \r{vev} demonstrates also that the operators ${\cal Z}_\pm(\theta)$ anticommute for real values of the spectral parameter $\theta$. We would like to remark here that the operators ${\cal Z}_\pm(\theta)$ being expressed in terms of the fermionic operators $b_\pm(\nu)$ \begin{equation}} \def\bn{\beq\label{ZZ6} {\cal Z}_\pm(\theta)= \int_{-\infty}^\infty d\nu\ \frac{b_\pm(\nu)}{\Gamma\sk{\frac{1}{2} +i\nu}}e^{i\nu\theta} \end{equation}} \def\ed{\eeq should not be understood literally, but rather as a certain normal ordering expression, where the normal ordering is dictated by the prescription \r{vev} on the domains of analyticity of the products of the scattering data operators. The naive use of the vacuum expectation value $$ \langle b_\pm(\nu)b_\mp(\nu')\rangle=\delta(\nu+\nu')\Theta(\nu) $$ where $\Theta(\nu)$ is the step-function, in order to calculate \r{vev} does not allow to find the domain where the vacuum expectation value $\langle {\cal Z}_\pm(\theta_1) {\cal Z}_\mp(\theta_2)\rangle$ is defined since this information is encoded in the analytical properties of the scattering transform. In order to obtain \r{coa} from \r{vev} it is convenient to introduce the operators $Z_\pm^*(\theta)$ and $Z_\pm(\theta)$ as a shift by $\pm\pi i/2$ of the operators ${\cal Z}_\pm(\theta)$ using the freedom to move the contour in the integral representation \r{psi-Z}: \begin{equation}} \def\bn{\beq\label{ZF-oper} Z^*_\pm(\theta)={\cal Z}_\pm(\theta-\pi i/2),\quad Z_\pm(\theta)={\cal Z}_\mp(\theta+\pi i/2)\ . \end{equation}} \def\ed{\eeq Then the fermion fields \r{psi-Z} can be rewritten in the two equivalent form: \begin{eqnarray} \Psi_\pm(r,\alpha)&=&{\sqrt{m}\over2\sqrt{\pi}}\int_{-\infty}^\infty d\theta\ Z_\mp(\theta+\alpha)e^{-imr{\rm sh}(\theta)} \spin{e^{\pi i/4}e^{(\theta+\alpha)/2}}{e^{-\pi i/4}e^{-(\theta+\alpha)/2}}\nonumber} \def\gamma{\gamma\\ &=&{\sqrt{m}\over2\sqrt{\pi}}\int_{-\infty}^\infty d\theta\ Z^*_\pm(\theta+\alpha)e^{imr{\rm sh}(\theta)} \spin{e^{-\pi i/4}e^{(\theta+\alpha)/2}}{e^{\pi i/4}e^{-(\theta+\alpha)/2}}.\nonumber} \def\gamma{\gamma \end{eqnarray} The operators $Z_\pm(\theta)$ or $Z^*_\pm(\theta)$ acting in angular fermionic or bosonic Hilbert spaces can be associated with the states in total Hilbert space $\CH$ of the model. In the next subsections we will identify them with intertwining and dual intertwining operators for the screening current algebra. The fact that the pole at the point $\theta_1=\theta_2+i\pi$ does not produce the restriction on the domain of the analyticity can be seen in the general situation, because the origin of this pole is the pinching of the contour in the integral representation of the function $\langle {\cal Z}_\pm(\theta_1) {\cal Z}_\mp(\theta_2)\rangle$ when $\theta_1\to\theta_2+i\pi$. \subsection{Integrals of motion} The operator which describes the evolution with respect to the angular time $\alpha$ is the Lorentz boost operator $K$. In terms of the fields: $$ K= \frac{i}{2}\int_{0}^\infty dr\ r\ \Bigl[e^{-\alpha} \sk{\overline\psi_-\sk{\partial_r\overline\psi_+}-\sk{\partial_r\overline\psi_-}\overline\psi_+}- e^{\alpha} \sk{\psi_-\sk{\partial_r\psi_+}-\sk{\partial_r\psi_-}\psi_+} +2m \overline\psi_-\psi_+ -2m \psi_-\overline\psi_+ \Bigr]\ . $$ Using canonical anticommutation relations \r{coa} one can find the action of this operator on the components of the Dirac spinor: \begin{eqnarray} {[}K, \overline\psi_\pm(r,\alpha){]}&=& i\frac{\partial\overline\psi_\pm(r,\alpha)}{\partial\alpha}- \frac{i}{2}\overline\psi_\pm(r,\alpha) ,\nonumber} \def\gamma{\gamma\\ {[}K, \psi_\pm(r,\alpha){]}&=& i\frac{\partial\psi_\pm(r,\alpha)}{\partial\alpha}+ \frac{i}{2}\psi_\pm(r,\alpha)\ . \label{b-a} \end{eqnarray} Note that the boost operator acts differently on the different components of spinors $\Psi_\pm(r,\alpha)$. The formulas \r{b-a} can be rewritten as follows \begin{eqnarray} e^{-i\eta K}\overline\psi_\pm(r,\alpha)e^{i\eta K}&=& e^{-\eta/2}\overline\psi_\pm(r,\alpha+\eta),\nonumber} \def\gamma{\gamma\\ e^{-i\eta K}\psi_\pm(r,\alpha)e^{i\eta K}&=& e^{\eta/2}\psi_\pm(r,\alpha+\eta)\ .\label{b-ae} \end{eqnarray} After rotation of the fermions around the origin by $\eta=2\pi i$ both of the equations \r{b-ae} become \begin{equation}} \def\bn{\beq\label{b-at} e^{2\pi K}\Psi_\pm(r,\alpha)e^{-2\pi K}=-\Psi_\pm(r,\alpha+2\pi i)\ . \end{equation}} \def\ed{\eeq In terms of Rindler fermions $b_\pm(\nu)$ the boost operator has a form \begin{equation}} \def\bn{\beq\label{boost} K=\int_{-\infty}^\infty d\nu\ \nu\ {:}b_-(\nu)b_+(-\nu){:}\equiv \int_{-\infty}^\infty d\nu\ \nu\ {:}b_+(\nu)b_-(-\nu){:} \end{equation}} \def\ed{\eeq and yields the value of continuous mode $\nu$ \begin{eqnarray} {[}K,b_\pm(\nu){]}=\nu b_\pm(\nu),\quad \forall\ \nu\in\Bbb{R}\ . \label{b-prop} \end{eqnarray} A second important operator is the operator of topological charge which can be written in terms of the fields as follows \begin{equation}} \def\bn{\beq\label{Q} Q=-\int_{0}^\infty dr\ \sk{e^\alpha\psi_-(r,\alpha)\psi_+(r,\alpha) +e^{-\alpha}\overline\psi_-(r,\alpha)\overline\psi_+(r,\alpha) }. \end{equation}} \def\ed{\eeq In terms of the fermionic modes it has the form \begin{equation}} \def\bn{\beq Q =-\int_{-\infty}^\infty d\nu\ {:}b_-(\nu)b_+(-\nu){:}= \int_{-\infty}^\infty d\nu\ {:}b_+(\nu)b_-(-\nu){:} \label{fer-ch} \end{equation}} \def\ed{\eeq and is normalized in such a way that the charges of the Rindler fermions $b_\pm(\nu)$ correspond to their indexes \begin{equation}} \def\bn{\beq {[}Q,b_\pm(\nu){]}=\pm b_\pm(\nu),\quad \forall \nu\in\Bbb{R}. \label{ch-prop} \end{equation}} \def\ed{\eeq In Rindler's parametrization the contours in the definition of the conserved charges \r{charges} are the straight rays $\alpha=\hbox{const}$. So in RRW we have \begin{equation}} \def\bn{\beq\label{i1} Q^J={1\over2}\int_0^\infty dr\left({\bf e}^{\alpha}J_y + {\bf e}^{-\alpha}J_{\overline y}\right). \end{equation}} \def\ed{\eeq Let us consider the charges given by this formula and compute them for the conserved currents $J^0$ \r{iii2} and $\widetilde J^0$ \r{ii2}. We obtain the result \begin{equation}} \def\bn{\beq\label{xxxx} I_n=(-1)^n\widetilde I_n= \int_{-\infty}^\infty d\theta\ e^{n\theta} {:}{\cal Z}_-(\theta)\Lambda_+(\theta){:}. \end{equation}} \def\ed{\eeq Using formulas \r{zz5} and \r{ZZ6} we can observe that arbitrary non-vanishing matrix elements of the integrals $I_n$ or $\widetilde I_n$ in the Fermionic Fock space are divergent. To avoid this divergence we consider the analytical continuation of discrete index $n\to -i\lambda$ to the imaginary axis, where $\lambda\in\Bbb{R}$. In this case the charges given by the currents \r{ii1}, \r{ii2} and \r{iii1}, \r{iii2} will produce well defined quantities. In contrast to \r{xxxx} the charges corresponding to the currents $J^0$ and $\widetilde J^0$ do not coincide. We denote the ones corresponding to neutral currents \r{ii2} and \r{iii2} as $a_\lambda$, $\widetilde a_\lambda$. In terms of the scattering data operators or in Rindler fermions they have the form \begin{eqnarray} a_\lambda&=&\int_{-\infty}^\infty d\theta \ e^{i\lambda\theta} {:} {\cal Z}_-(\theta)\Lambda_+(\theta){:}= \int_{-\infty}^\infty d\nu\ \frac {\Gamma\sk{\frac{1}{2}-i\nu}} {\Gamma\sk{\frac{1}{2}+i(\lambda-\nu)}} {:}b_-(\lambda-\nu)b_+(\nu){:}\ , \label{ii6}\\ \widetilde a_\lambda&=&\int_{-\infty}^\infty d\theta \ e^{i\lambda\theta} {:}\Lambda_-(\theta){\cal Z}_+(\theta){:}= \int_{-\infty}^\infty d\nu\ \frac {\Gamma\sk{\frac{1}{2}-i(\lambda-\nu)}} {\Gamma\sk{\frac{1}{2}+i\nu}} {:}b_-(\lambda-\nu)b_+(\nu){:}\ , \label{i6} \end{eqnarray} where normal ordering is defined with respect to fermionic vacuum vectors. By comparing the formulas \r{i6} and \r{ii6} we conclude that the conserved charges $a_\lambda$ and $\widetilde a_\lambda$ are related to each other by some complicated integral transform. This integral transform can be described algebraically by extending the algebra of the operators $a_\lambda$ and $\widetilde a_\lambda$. This will be demonstrated in the Appendix \ref{app1}. Using \r{stand} we see that \begin{equation}} \def\bn{\beq\label{i15} [a_\lambda,a_\mu]=[\widetilde a_\lambda,\widetilde a_\mu]=\lambda\delta(\lambda+\mu). \end{equation}} \def\ed{\eeq and this Heisenberg type commutation relation allows us to use these operators for the bosonization. \subsection{The bosonization and the screening currents} Using \r{stand} we can obtain the commutation relations: \begin{eqnarray} \label{ii16} [a_\lambda,{\cal Z}_-(\theta)]&=&e^{-i\lambda\theta} {\cal Z}_-(\theta),\qquad [a_\lambda,\Lambda_+(\theta)]=-e^{-i\lambda\theta} \Lambda_+(\theta)\\ \label{i16} [\widetilde a_\lambda,\Lambda_-(\theta)]&=&e^{-i\lambda\theta} \Lambda_-(\theta),\qquad [\widetilde a_\lambda,{\cal Z}_+(\theta)]=-e^{-i\lambda\theta} {\cal Z}_+(\theta) \end{eqnarray} These formulas together with \r{i15} allow to bosonize the operators $\Lambda_+(\theta)$ and ${\cal Z}_-(\theta)$ in terms of a free bosonic field constructed from continuous bosons $a_\lambda$ and the operators $\Lambda_-(\theta)$ and ${\cal Z}_+(\theta)$ from the analogous free field constructed from the bosons $\widetilde a_\lambda$. This bosonization should conserve all the properties of the Hilbert space $\CH_R^f$ and the action of the operators $\Lambda_\pm(\theta)$, ${\cal Z}_\pm(\theta)$ on it. It is clear that it is impossible to do this using only bosonic modes $a_\lambda$ because they carry the charge 0, while the Hilbert space $\CH_R^f$ is naturally graded with respect to topological charge operator: \begin{equation}} \def\bn{\beq\label{HF-gr} \CH_R^f=\bigoplus_{n\in\Bbb{Z}}\CH_{R,n}^f,\quad \CH_{R,n}^f=\left\{\left. x\in \CH_{R}^f\ \right|\ Qx=nx\right\} \end{equation}} \def\ed{\eeq Because of the formulas \r{zz5} and \r{ZZ6} the operators $\Lambda_\pm(\theta)$ and ${\cal Z}_\pm(\theta)$ change the topological charge \begin{equation}} \def\bn{\beq\label{tcch} \Lambda_\pm(\theta),\ {\cal Z}_\pm(\theta)\ : \ \CH^f_{R,n}\to \CH^f_{R,n\pm1}\ . \end{equation}} \def\ed{\eeq Note that in our normalization the topological charge operator $Q$ coincides with the operators $-a_0\equiv -\widetilde a_0$. To conserve these properties of $\CH_{R}^f$ in the bosonization picture we introduce a pair of zero mode operators ${{\cal Q}}$ and ${{\cal P}}$ which satisfy the commutation relations \begin{equation}} \def\bn{\beq\label{zero-mo} [{{\cal P}},{{\cal Q}}]=i \end{equation}} \def\ed{\eeq and bosonic vacuum vectors $\bvac{n}$, $n\in\Bbb{Z}$ which are annihilated by all nonnegative bosonic modes and are eigenstates of the operator ${{\cal P}}$ \begin{equation}} \def\bn{\beq\label{bFock} a_\lambda\bvac{n}=0,\qquad \hbox{$\lambda\geq 0$},\qquad {{\cal P}}\bvac{n}=n\bvac{n}\ . \end{equation}} \def\ed{\eeq We identify $ \CH^f_{R,n}$ with bosonic space $ \CH^b_{R,n}$ generated from bosonic vacuum vector $\bvac{n}$ \begin{equation}} \def\bn{\beq\label{bFock1} \int_{-\infty}^0 f_n(\lambda_n) a_{\lambda_n} d\lambda_n\ldots \int_{-\infty}^0 f_1(\lambda_1) a_{\lambda_1} d\lambda_1\ \bvac{n}\ , \end{equation}} \def\ed{\eeq where the functions $f_i(\lambda)$ are analytical functions in a neighborhood of ${\Bbb{R}_-}$ except $\lambda=0$, where they can have a simple pole. Then due to \r{i16} the operators $\Lambda_-(\theta)$ and ${\cal Z}_+(\theta)$ can be bosonized as follows: \begin{equation}} \def\bn{\beq \label{i17} \Lambda_+(\theta)=\exp\sk{i{{\cal Q}}+ \int_{-\infty}^\infty\hspace{-20.5pt}-\hspace{11.5pt} \frac{d\lambda}{\lambda}\ a_\lambda e^{i\lambda\theta}},\quad {\cal Z}_-(\theta)=\exp\sk{-i{{\cal Q}}- \int_{-\infty}^\infty\hspace{-20.5pt}-\hspace{11.5pt} \frac{d\lambda}{\lambda}\ a_\lambda e^{i\lambda\theta}}. \end{equation}} \def\ed{\eeq The integral under the exponent is understood as principal value integral to exclude the singularity at zero: \begin{equation}} \def\bn{\beq\label{vpint} \int_{-\infty}^\infty\hspace{-20.5pt}-\hspace{11.5pt} f(\lambda)\, d\lambda=\lim{\epsilon\to+0}\sk{\int_{-\infty}^{-\epsilon} f(\lambda)\, d\lambda + \int_{\epsilon}^\infty f(\lambda)\, d\lambda}. \end{equation}} \def\ed{\eeq We define the products of the operators like \r{i17} to be $\zeta$-function regularized \cite{JM1,JMK} \begin{eqnarray} &\exp\left(\int_{-\infty}^\infty d\lambda\ g_1(\lambda)\,a_\lambda\right)\ \cdot\ \exp\left(\int_{-\infty}^\infty d\mu\ g_2(\mu)\,a_\mu\right) = \nonumber} \def\gamma{\gamma\\ &\quad =\frac{e^{\gamma}}{2\pi} \exp\left(\int_{\widetilde C}{d\lambda\,\ln(-\lambda)\over2\pi i}\ c(\lambda)g_1(\lambda)g_2(-\lambda) \right) \exp\left(\int_{-\infty}^\infty d\lambda\ (g_1(\lambda)+g_2(\lambda))\,a_\lambda\right)\ . \label{normal} \end{eqnarray} where $c(\lambda)=\lambda$ and $\gamma$ is Euler constant and the contour $\widetilde C$ is shown in Fig.~1. \bigskip \unitlength 1.00mm \linethickness{0.4pt} \begin{picture}(121.00,20.00) \put(17.00,15.00){\makebox(0,0)[cc]{0}} \put(20.00,15.00){\makebox(0,0)[cc]{$\bullet$}} \put(132.00,15.00){\makebox(0,0)[cc]{$+\infty$}} \put(20.00,15.00){\line(1,0){100.33}} \put(40.00,10.00){\line(1,0){80.33}} \put(120.00,20.00){\line(-1,0){100.00}} \put(30.00,5.00){\makebox(0,0)[cc]{Fig.~1.}} \put(121.00,10.00){\vector(1,0){0.2}} \put(100.00,10.00){\line(1,0){21.00}} \put(20.00,10.00){\line(1,0){22.00}} \put(20.67,10.00){\line(1,0){22.00}} \put(20.17,15.00){\oval(15.00,10.00)[l]} \end{picture} \smallskip Naturally there is an alternative way to bosonize the fermionic Fock space $\CH_R^f$ using modes $\widetilde a_\lambda$ and introducing the corresponding zero mode operators $\widetilde{{\cal Q}}$, $\widetilde{{\cal P}}$, the bosonic vacuum vectors and bosonic Fock spaces. It is clear that these vacuum vectors are not a priori the same as for bosons $a_\lambda$ because of the complicated commutation relations between the bosons $a_\lambda$ and $\widetilde a_\lambda$ (see Appendix \ref{app1}). This alternative bosonization looks as follows \begin{equation}} \def\bn{\beq \label{ii17} \Lambda_-(\theta)=\exp\sk{-i\widetilde{{\cal Q}}- \int_{-\infty}^\infty\hspace{-20.5pt}-\hspace{11.5pt} \frac{d\lambda}{\lambda}\ \widetilde a_\lambda e^{i\lambda\theta}},\quad {\cal Z}_+(\theta)=\exp\sk{i\widetilde {{\cal Q}}+ \int_{-\infty}^\infty\hspace{-20.5pt}-\hspace{11.5pt} \frac{d\lambda}{\lambda}\ \widetilde a_\lambda e^{i\lambda\theta}}\ . \end{equation}} \def\ed{\eeq We also define the charged currents in the momentum space which correspond to the charged conserved currents \r{ii1} and \r{iii1} by the requirement that they relate the components of the operators ${\cal Z}_\pm(\theta)$. Let \begin{equation}} \def\bn{\beq\label{I22} E(\theta)= {:}\Lambda_+(\theta+\pi i/2) \Lambda_+(\theta-\pi i/2){:},\qquad \widetilde F(\theta)= {:}\Lambda_-(\theta+\pi i/2) \Lambda_-(\theta-\pi i/2){:} \end{equation}} \def\ed{\eeq be the operator valued currents in the momentum space. Using vacuum expectation values and Wick theorem we can prove the formulas \begin{eqnarray} {\cal Z}_+(\theta)&=&-i\int_{C_1}du\ e^{u-\theta}{\cal Z}_-(\theta)E(u)+ i\int_{C_2}du\ e^{u-\theta}E(u){\cal Z}_-(\theta)= \Lambda_+(\theta+\pi i)+\Lambda_+(\theta-\pi i)\ ,\label{Z+Z-}\\ {\cal Z}_-(\theta)&=&-i\int_{C_1}du\ e^{u-\theta}{\cal Z}_+(\theta)\widetilde F(u)+ i\int_{C_2}du\ e^{u-\theta}\widetilde F(u){\cal Z}_+(\theta)= \Lambda_-(\theta+\pi i)+\Lambda_-(\theta-\pi i)\ ,\label{Z-Z+} \end{eqnarray} where the contour $C_1$ goes from $-\infty$ to $+\infty$ and is above all the poles in the operator product expansion ${\cal Z}_-(\theta)E(u)$ and ${\cal Z}_+(\theta)\widetilde F(u)$ and the contour $C_2$ is also from $-\infty$ to $+\infty$ and below all the poles in the OPE $E(u){\cal Z}_-(\theta)$ and $\widetilde F(u){\cal Z}_+(\theta)$. Let us prove \r{Z+Z-}. It follows from the OPE \begin{eqnarray} {\cal Z}_-(\theta)E(u)&=&{:}{\cal Z}_-(\theta)E(u){:}- \frac{1}{2\pi i}\left[ \frac{\Lambda_+\sk{u-\frac{\pi i}{2}}}{\theta-u-\frac{\pi i}{2}} -\frac{\Lambda_+\sk{u+\frac{\pi i}{2}}}{\theta-u+\frac{\pi i}{2}} \right],\nonumber} \def\gamma{\gamma\\ E(u){\cal Z}_-(\theta)&=&{:}{\cal Z}_-(\theta)E(u){:}- \frac{1}{2\pi i}\left[ \frac{\Lambda_+\sk{u+\frac{\pi i}{2}}}{u-\theta-\frac{\pi i}{2}} -\frac{\Lambda_+\sk{u-\frac{\pi i}{2}}}{u-\theta+\frac{\pi i}{2}} \right].\nonumber} \def\gamma{\gamma \end{eqnarray} Now the formula \r{Z+Z-} follows from the trivial calculation of the integrals. The second relation \r{Z-Z+} is proved analogously. Using formulas \r{i17} and \r{ii17} we can write down the bosonized expressions for the screening currents \begin{eqnarray} E(u)&=&\exp\sk{2i{{\cal Q}}+2\int_{-\infty}^\infty\hspace{-20.5pt}-\hspace{11.5pt}{d\lambda\over\lambda}\ e^{iu\lambda}{\rm ch}\sk{\pi\lambda/2}\ a_\lambda}\ ,\label{E-scr}\\ \widetilde F(u)&=&\exp\sk{-2i\widetilde{{\cal Q}}-2\int_{-\infty}^\infty\hspace{-20.5pt}-\hspace{11.5pt}{d\lambda\over\lambda}\ e^{iu\lambda}{\rm ch}\sk{\pi\lambda/2}\ \widetilde a_\lambda}\ ,\label{F-scr} \end{eqnarray} in terms of the bosons $a_\lambda$ for the current $E(u)$ and $\widetilde a_\lambda$ for $\widetilde F(u)$. Because the operators $a_\lambda$ and $\widetilde a_\lambda$ do not form a closed algebra the screening currents $E(u)$ and $\widetilde F(u)$ also do not form the closed algebra. In the next subsection we will define another pair of the screening currents $\widetilde E(u)$ and $F(u)$ such that the pairs $E(u)$, $F(u)$ or $\widetilde E(u)$, $\widetilde F(u)$ do form the closed algebra both isomorphic to the screening current algebra introduced in \cite{Lu1,KLP}. \subsection{Another pair of screening currents and quantum Jost functions} The commutation relations \r{YBt} have a smooth classical limit when $\xi\to0$ and correspondingly $q'\to1$. In this limit these commutation relations become Poisson brackets for the elements of monodromy matrices for classical SG model \cite{FTbook}. The elements $Z'_{\varepsilon}(\alpha)$ of the `monodromy' matrix associated with the half-line were introduced in \cite{Lu1}. It was shown in \cite{Lu2} that they are the quantum analogs of classical Jost functions. These classical objects can be written explicitly as path-ordered exponents of SG connections in RRW and satisfy the Poisson bracket relation \begin{equation}} \def\bn{\beq\label{PBrel} \{{\cal Z}'_{\varepsilon_1}(\alpha_1),{\cal Z}'_{\varepsilon_2}(\alpha_2)\}=r_{\varepsilon_1\varepsilon_2}^{\varepsilon'_1\varepsilon'_2} (\alpha_1-\alpha_2) {\cal Z}'_{\varepsilon'_2}(\alpha_2){\cal Z}'_{\varepsilon'_1}(\alpha_1)\ , \end{equation}} \def\ed{\eeq where $r_{\varepsilon_1\varepsilon_2}^{\varepsilon'_1\varepsilon'_2} (\alpha) $ is a classical trigonometric $r$-matrix \begin{equation}} \def\bn{\beq\label{r-clas} r_{\varepsilon_1\varepsilon_2}^{\varepsilon'_1\varepsilon'_2} (\alpha) =\lim{\xi\to0}\frac{R_{\varepsilon_1\varepsilon_2}^{\varepsilon'_1\varepsilon'_2} (\alpha,\xi)-1}{\pi i\xi}=\frac{1}{2} \sk{\begin{array}{cccc} {\rm th}(\alpha/2) & & & \\ & -{\rm cth}(\alpha/2) & 2{\rm sh}^{-1}(\alpha) & \\ & 2{\rm sh}^{-1}(\alpha) & -{\rm cth}(\alpha/2) & \\ & & & {\rm th}(\alpha/2) \end{array}} \end{equation}} \def\ed{\eeq obtained from the $R$-matrices \r{RR-mat}. Note, that the scalar factor of the $R$-matrix \r{RR-mat} also contribute to the classical $r$-matrix \r{r-clas}. Unfortunately, a way to obtain the quantum analog of the relations \r{PBrel} \begin{equation}} \def\bn{\beq\label{PB-q} {\cal Z}'_{\varepsilon_1}(\alpha_1){\cal Z}'_{\varepsilon_2}(\alpha_2)=R_{\varepsilon_1\varepsilon_2}^{\varepsilon'_1\varepsilon'_2} (\alpha_1-\alpha_2,\xi) {\cal Z}'_{\varepsilon'_2}(\alpha_2){\cal Z}'_{\varepsilon'_1}(\alpha_1) \end{equation}} \def\ed{\eeq starting from SG Lagrangian is not known. Nevertheless, one can formulate the properties of these operators which allows to reconstruct them uniquely. These properties follow from the interpretation of the operators ${\cal Z}_\pm(\theta)$ as the operators in $\CH_R$ which correspond to the states in total Hilbert space $\CH$ of the model. Since the operators ${\cal Z}'_\pm(\alpha)$ are related to the quantum Jost functions and to the integrals of motion it is natural to require their commutativity with the operators ${\cal Z}_\varepsilon(\theta)$ up to the phase \begin{equation}} \def\bn{\beq\label{Z-Z'} {\cal Z}_\varepsilon(\theta){\cal Z}'_\nu(\alpha)=\varepsilon\nu\ \phi(\theta-\alpha)\ {\cal Z}'_\nu(\alpha){\cal Z}_\varepsilon(\theta) \end{equation}} \def\ed{\eeq where $\phi(\theta)$ is a yet unknown function (see \r{ZFI-II} below). Because of the relations \r{Z+Z-} and \r{Z-Z+} this requirement is equivalent to the following \begin{equation}} \def\bn{\beq\label{EF-Z'} E(u){\cal Z}'_-(\alpha)=-{\cal Z}'_-(\alpha) E(u),\quad \widetilde F(u){\cal Z}'_+(\alpha)=-{\cal Z}'_+(\alpha)\widetilde F(u) \end{equation}} \def\ed{\eeq These anticommutation relations can be satisfied by the following bosonization of the operators $Z'_\nu(\alpha)$ \begin{eqnarray} {\cal Z}'_-(\alpha)=\exp\sk{i{{\cal Q}}/2 +\int_{-\infty}^\infty \frac{d\lambda}{\lambda}\ \frac{a_\lambda}{2{\rm ch}\,\pi\lambda/2}\ e^{i\lambda\alpha}}, \label{Z'+bos}\\ {\cal Z}'_+(\alpha)=\exp\sk{-i\widetilde {{\cal Q}}/2 -\int_{-\infty}^\infty \frac{d\lambda}{\lambda}\ \frac{\widetilde a_\lambda}{2{\rm ch}\,\pi\lambda/2}\ e^{i\lambda\alpha}}. \label{Z'-bos} \end{eqnarray} Using this bosonization and also the rule of the normal ordering \r{normal} we can observe that the operators ${\cal Z}'_\pm(\alpha)$ are related to the scattering data operators $\Lambda_\pm(\theta)$ as follows \begin{equation}} \def\bn{\beq\label{rel1} \Lambda_{-\varepsilon}(\alpha)=g^{-1} {\cal Z}'_\varepsilon\sk{\alpha-\frac{i\pi}{2}} {\cal Z}'_\varepsilon\sk{\alpha+\frac{i\pi}{2}}={:}{\cal Z}'_\varepsilon\sk{\alpha-\frac{i\pi}{2}} {\cal Z}'_\varepsilon\sk{\alpha+\frac{i\pi}{2}}{:}. \end{equation}} \def\ed{\eeq The normalization constant is given in terms of double $\Gamma$-functions (see Appendix \ref{app2} for the definition of these functions). Note that formula \r{rel1} allows to identify the scattering data operators $\Lambda_\pm(\theta)$ with the generating functions of the local operators introduced in the paper \cite{Lu1}. From the formulas \r{Z'+bos} and \r{i17} we can easily find the function $\phi(\theta)={\rm ctg}\sk{\frac{\pi}{4}+\frac{\theta}{2i}}$. The formulas in \r{rel1} are equalities in different bosonic Fock spaces generated by the operators $a_\lambda$ and $\widetilde a_\lambda$ respectively. We can translate them in equal bosonic spaces introducing another pair of screening currents $F(u)$ and $\widetilde E(u)$ \begin{equation}} \def\bn{\beq\label{scr2} F(u)\equiv {\cal Z}_-(u),\quad \widetilde E(u)\equiv {\cal Z}_+(u). \end{equation}} \def\ed{\eeq These relations are given by the integral transforms \begin{eqnarray} {\cal Z}_+'(\alpha)&=&2e^{-2\gamma}\pi^{3/2}\int_{C'}du\ e^{\frac{u-z}{2}} \left[e^{\frac{\pi i}{4}} {\cal Z}'_-(\alpha)F(u)+ e^{-\frac{\pi i}{4}}F(u) {\cal Z}'_-(\alpha) \right]\nonumber} \def\gamma{\gamma\\ &=&\frac{e^{-3\gamma/2}}{\sqrt2\pi}\int_{C'} du\ \Ga{\frac{1}{4}+\frac{u-\alpha}{2\pi i}} \Ga{\frac{1}{4}-\frac{u-\alpha}{2\pi i}} \ {:}{\cal Z}_-'(\alpha)F(u){:},\label{rel+} \end{eqnarray} and \begin{eqnarray} {\cal Z}_-'(z)&=&2e^{-2\gamma}\pi^{3/2}\int_{C'}du\ e^{\frac{u-z}{2}} \left[e^{\frac{\pi i}{4}} {\cal Z}'_+(\alpha)\widetilde E(u)+ e^{-\frac{\pi i}{4}}\widetilde E(u) {\cal Z}'_+(\alpha) \right]\nonumber} \def\gamma{\gamma\\ &=&\frac{e^{-3\gamma/2}}{\sqrt2\pi}\int_{C'} du\ \Ga{\frac{1}{4}+\frac{u-\alpha}{2\pi i}} \Ga{\frac{1}{4}-\frac{u-\alpha}{2\pi i}} \ {:}{\cal Z}_+'(\alpha)\widetilde E(u){:},\label{rel-} \end{eqnarray} where in both formulas the contour $C'$ goes from $-\infty$ to $\infty$ along the real axis such that \begin{equation}} \def\bn{\beq\label{contour'} {\rm Im}\,\,\alpha-\pi/2 < {\rm Im}\,\,u< {\rm Im}\,\,\alpha+\pi/2 \end{equation}} \def\ed{\eeq The proof of the fact that the relation \r{rel+} is equivalent to the relation \r{rel1} for $\varepsilon=-$ or vice versa \r{rel-} is equivalent to the relation \r{rel1} for $\varepsilon=+$ can be found in the Appendix \ref{app2}. The fact that the second set of the screening currents for the quantum Jost operators ${\cal Z}'_\pm(\alpha)$ coincided with scattering data operators ${\cal Z}_\pm(\theta)$ is specific to the FF point. But what is true in general is that we have the closed algebra of screening currents either for the pair $E(u)$, $F(u)$ or for the pair $\widetilde E(u)$, $\widetilde F(u)$. Both of these algebras can be obtained from the commutation relations \r{scr-com} using different Gauss decompositions of $L$-operators. \setcounter{equation}{0} \section{Algebra of screening currents} As we already said in the Introduction one of the goals of this paper is to explain the algebraic structures which allows to describe simultaneously two quantum group structures with different parameter of deformations developing ideas of the paper \cite{Lu1}. This algebra will be introduced and explained in this section. This is a non-abelian algebra of screening currents which can be defined using exact $S$-matrix of soliton-antisoliton scattering in SG model \cite{ZZ,KLP}. We define this algebra for the value of the of the renormalized coupling constant $1<\xi<\infty$ in so called breatherless regime. We will demonstrate that the representation theory of this algebra has smooth limit when $\xi\to1$, which corresponds to the FF point of SG model. Using the bosonization we will show that the intertwining operators of the level 1 highest weight modules for the screening current algebra coincide with the operators ${\cal Z}_\pm(\theta)$ and ${\cal Z}'_\pm(\theta)$ defined in the previous section from analysis of massive Dirac fermions in RRW. \subsection{$R$ and $S$ Matrices} Consider the following $\CR$-matrix. \begin{eqnarray} \CR^+(u,\xi)&=&\tau^+(u) \CR (u,\xi),\quad \CR(u,\xi)\ =\ r (u,\xi)\overline \CR (u,\xi)\ , \label{R-mat}\\ \overline \CR (u,\xi)&=& \left( \begin{array}{cccc} 1&0&0&0\\ 0&b (u,\xi)&c (u,\xi)&0\\ 0&\widetilde c (u,\xi)&b (u,\xi)&0\\ 0&0&0&1 \end{array} \right) \ , \nonumber} \def\gamma{\gamma\\ r (u,\xi)&=&{\Ga{\frac{1}{\xi}}\Ga{1+\frac{iu}{\pi\xi}}\over \Ga{\frac{1}{\xi}+\frac{iu}{\pi\xi}}} \prod_{p=1}^\infty {R _p(u,\eta) R _p(i\pi- u,\eta) \over R _p(0,\xi) R _p(i\pi,\xi)}\ , \label{inf-pr}\\ R _p(u,\eta)&=&{\Ga{\frac{2p}{\xi}+\frac{iu}{\pi\xi}} \Ga{1+\frac{2p}{\xi}+\frac{iu}{\pi\xi}}\over \Ga{\frac{2p+1}{\xi}+\frac{iu}{\pi\xi}} \Ga{1+\frac{2p-1}{\xi}+\frac{iu}{\pi\xi}}}\ ,\nonumber} \def\gamma{\gamma\\ b(u,\xi)&=& {{\rm sh}\,\frac{u}{\xi}\over{\rm sh}\,\frac{u-\pi i}{\xi}}\ ,\quad c(u,\xi)\ =\ - {{\rm sh}\,\frac{i\pi}{\xi} \over{\rm sh}\,\frac{u-\pi i}{\xi}}\ ,\quad \tau^+(u)=i{\rm cth}\sk{{u\over2}} . \nonumber} \def\gamma{\gamma \end{eqnarray} The scalar factor $r(u,\xi)$ has an integral representation \begin{equation}} \def\bn{\beq\label{fac-r} r(u,\xi)= \exp \left(2i\int_{0}^\infty{d\lambda\,\over\lambda} \frac{{\rm sh}\,\lambda/2}{{\rm sh}\,\lambda}\ \frac{{\rm sh}\,(\xi-1)\lambda/2}{{\rm sh}\,\xi\lambda/2} \ \sin\sk{\frac{\lambda u}{\pi}}\right),\quad -\pi<{\rm Im}\,\,u<\pi\ . \end{equation}} \def\ed{\eeq The $R$-matrix \r{R-mat} differs from the physical $S$-matrix which describe the soliton-antisoliton scattering by the transformation \begin{equation}} \def\bn{\beq\label{S-mat} S(\theta)=- r(\theta,\xi)\overline S(\theta),\quad \overline S(\theta)= (\sigma_z\ot1)\overline \CR(\theta,\xi)(1\otimes\sigma_z) \end{equation}} \def\ed{\eeq which change the sign in front of the elements $b(\theta,\xi)$. In the classical limit $\xi\to\infty$ such that $u/\xi$ is fixed the $R$-matrix $\CR(u,\xi)$ goes to identity, while the `physical' $S$-matrix goes to ${\rm diag}(-1,1,1,-1)$. On the other hand at the FF point $\xi\to1$ $S$-matrix becomes equal to $-1$ while $R$-matrix \r{R-mat} to ${\rm diag}(1,-1,-1,1)$. The matrix $\overline S(\theta)$ can be written using a multiplicative spectral parameter $z=e^{-\theta/\xi}$ and deformation parameter $q$ introduced by \r{s-mul} \begin{equation}} \def\bn{\beq\label{S-mul} \left(\begin{array}{cccc} 1&0&0&0\\ 0&\frac{z-z^{-1}}{zq-z^{-1}q^{-1}}&\frac{(q-q^{-1})}{zq-z^{-1}q^{-1}}&0\\ 0&\frac{(q-q^{-1})}{zq-z^{-1}q^{-1}}&\frac{z-z^{-1}}{zq-z^{-1}q^{-1}}&0\\ 0&0&0&1 \end{array}\right). \end{equation}} \def\ed{\eeq The physical $R$-matrix which describe the commutation relations of quantum monodromy matrices in SG model \r{YBt} and quantum Jost functions \r{PB-q} can be similarly written in terms of the matrix $\CR(u,\xi)$ \begin{equation}} \def\bn{\beq\label{RR-mat} R(\alpha)= r(\alpha,\xi+1)\overline R(\alpha),\quad \overline R(\alpha)= (\sigma_z\ot1)\overline \CR(-\alpha,-\xi-1)(1\otimes\sigma_z) \end{equation}} \def\ed{\eeq Note that matrices $\overline S(\theta)$ and $\overline R(\alpha)$ are related by the duality transformation \begin{equation}} \def\bn{\beq\label{dual-tr} \xi\to -\xi -1,\quad \theta\to-\alpha\ . \end{equation}} \def\ed{\eeq The fact is that the scalar factors of $S$ and $R$-matrices are related by the same transformation. To see this one should use simple identity $$ \frac{{\rm sh}\,\lambda/2}{{\rm sh}\,\lambda}\sk{ \frac{{\rm sh}\,(\xi+1)\lambda/2}{{\rm sh}\,\xi\lambda/2}+ \frac{{\rm sh}\,(\xi-1)\lambda/2}{{\rm sh}\,\xi\lambda/2}}=1 $$ to rewrite the scalar factor $-r(\theta,\xi)$ in the form \begin{equation}} \def\bn{\beq\label{fac-r-n} -r(u,\xi)= \exp \left(-2i\int_{0}^\infty{d\lambda\,\over\lambda} \frac{{\rm sh}\,\lambda/2}{{\rm sh}\,\lambda}\ \frac{{\rm sh}\,(\xi+1)\lambda/2}{{\rm sh}\,\xi\lambda/2} \ \sin\sk{\frac{\lambda u}{\pi}}\right). \end{equation}} \def\ed{\eeq Now one can see that functions $r(u,\xi+1)$ and $-r(u,\xi)$ transform to each other under \r{dual-tr}, although one should not think about this transformation literally. The point is that the quantization of the SG model is well defined for $0<\xi<\infty$, so in order to perform the dual transformation \r{dual-tr} we should first go to $+\infty$ and then come back to the negative axis from $-\infty$. During this path the properties of the model itself change drastically. \subsection{Algebra of screening currents} Set $1<\xi<\infty$. Let \begin{equation}} \def\bn{\beq L(u,\xi)=\left(\begin{array}{cc} L_{++}(u,\xi)&L_{+-}(u,\xi)\\ L_{-+}(u,\xi)&L_{--}(u,\xi) \end{array}\right) \label{L-op} \end{equation}} \def\ed{\eeq be a quantum $L$-operator whose matrix elements are treated as generating functions for the elements of the algebra given by the commutation relations: \begin{eqnarray} R^+(u_1-u_2,\xi+c)L_1(u_1,\xi)L_2(u_2,\xi)&=& L_2(u_2,\eta)L_1(u_1,\xi) R^+(u_1-u_2,\xi) \ , \label{RLL-univ}\\ \hbox{qdet} L(u)=L_{++}(u-i\tih)L_{--}(u)&-&L_{+-}(u-i\tih)L_{-+}(u)=1\ . \label{qdet=1} \end{eqnarray} Let \begin{equation}} \def\bn{\beq L(u) =\left(\begin{array}{cc} 1& f(u)\\0&1\end{array}\right) \left(\begin{array}{cc} k_1(u)&0\\ 0&k_2(u) \end{array}\right) \left(\begin{array}{cc} 1&0\\ e(u)&1\end{array}\right)\ , \label{GL-univ} \end{equation}} \def\ed{\eeq be the Gauss decomposition of the $L$-operator \r{L-op}. This Gauss decomposition corresponds to the algebra of screening currents $E(u)$ and $F(u)$ described in the previous section at the value $\xi=1$. To obtain the algebra related to the currents $\widetilde E(u)$ and $\widetilde F(u)$ we should start from another Gauss decomposition \begin{equation}} \def\bn{\beq \widetilde L(u) =\left(\begin{array}{cc} 1& 0\\\widetilde e(u)&1\end{array}\right) \left(\begin{array}{cc} \widetilde k_1(u)&0\\ 0&\widetilde k_2(u) \end{array}\right) \left(\begin{array}{cc} 1&\widetilde f(u)\\ 0&1\end{array}\right). \label{GL-univ-al} \end{equation}} \def\ed{\eeq The relation between the Gauss coordinates of both $L$-operators is complicated enough and can be described on the level of the bosonization of the $L$-operators by the relations similar to those described in the Appendix \ref{app1}. For the remainder of this paper we will work only with the operator \r{GL-univ}. One can deduce from (\ref{RLL-univ}), (\ref{qdet=1}) that $$ k_1(u)=(k_2(u+i\tih))^{-1}. $$ Let $ h(u)=k_1\left(u\right) k_2\left(u\right)^{-1}\ , \qquad h'(u)=k_2\left(u\right)^{-1} k_1\left(u\right)= \frac{\xi\sin\,(\pi/\xi)}{\xi'\sin\,(\pi/\xi')} h(u) \ , $ where by $\xi'$ we denote the combination $\xi+c$ of the parameter $\xi$ and the central element of the algebra ${\cal A}(\widehat{sl}_2)$. The Gauss coordinates $e(u)$, $f(u)$ and $h(u)$ of the $L$-operator \r{L-op} satisfy the following commutation relations ($u=u_1-u_2$): \begin{equation}} \def\bn{\beq e(u_1)f(u_2)-f(u_2)e(u_1) ={{\rm sh}\,(i\pi/\xi')\over{\rm sh}\, (u/\xi')}h'(u_1)- {{\rm sh}\,(i\pi/\xi)\over{\rm sh}\, (u/\xi)} h(u_2),\label{ef} \end{equation}} \def\ed{\eeq \begin{eqnarray} {\rm sh}\sk{\frac{u+i\tih}{\xi}}h(u_1)e(u_2)- {\rm sh}\sk{\frac{u-i\tih}{\xi}}e(u_2)h(u_1) &=&{\rm sh}\sk{\frac{i\pi}{\xi}} \{h(u_1),e(u_1)\},\label{he}\\ {\rm sh}\sk{\frac{u-i\tih}{\xi'}}h(u_1)f(u_2)- {\rm sh}\sk{\frac{u+i\tih}{\xi'}}f(u_2)h(u_1) &=&-{\rm sh}\sk{\frac{i\pi}{\xi'}} \{h(u_1),f(u_1)\},\label{hf}\\ {\rm sh}\sk{\frac{u+i\tih}{\xi}}e(u_1)e(u_2)- {\rm sh}\sk{\frac{u-i\tih}{\xi}}e(u_2)e(u_1) &=&{\rm sh}\sk{\frac{i\pi}{\xi}} \left(e(u_1)^2+e(u_2)^2\right),\label{ee-univ}\\ {\rm sh}\sk{\frac{u-i\tih}{\xi'}}f(u_1)f(u_2)- {\rm sh}\sk{\frac{u+i\tih}{\xi'}}f(u_2)f(u_1) &=&-{\rm sh}\sk{\frac{i\pi}{\xi'}} \left(f(u_1)^2+f(u_2)^2\right),\label{ff-univ} \end{eqnarray} \begin{equation}} \def\bn{\beq {{\rm sh}\sk{\frac{u-i\pi}{\xi'}} \over {\rm sh}\sk{\frac{u+i\pi}{\xi'}}} h(u_1)h(u_2) = h(u_2)h(u_1) {{\rm sh}\sk{\frac{u-i\pi}{\xi}} \over {\rm sh}\sk{\frac{u+i\pi}{\xi}}} \ .\label{hh} \end{equation}} \def\ed{\eeq In the next subsections we will describe the finite and infinite dimensional representations of the algbera ${\cal A}(\widehat{sl}_2)$. We will consider also the tensor products of the representations where the action of the algebra ${\cal A}(\widehat{sl}_2)$ is defined by the following comultiplication structure compatible with the commutation relations \r{RLL-univ}: \begin{eqnarray} \Delta\,c&=&c'+c''=c\otimes1+1\otimes c\ ,\nonumber} \def\gamma{\gamma\\ \Delta^{\rm op} L(u,\xi)&=&L(u- i\pi c^{(2)}/4,\xi+c^{(2)}) \ {\dot\otimes}\ L(u+ i\pi c^{(1)}/4,\xi) \label{comul-L-univ}\\ \Delta \sk{L(u,\xi)}^{-1}&=&\sk{L(u+i\pi c^{(2)}/4,\xi)} ^{-1} {\dot\otimes} \sk{L(u- i\pi c^{(1)}/4,\xi+c^{(1)})}^{-1},\nonumber} \def\gamma{\gamma \end{eqnarray} where the symbol $\dot\otimes$ signifies the matrix tensor product $$ \sk{A\dot\otimes B}_{ij}=\sum_k A_{ik}\otimes B_{kj}\ . $$ The comultiplications of the the Gauss coordinates of $L$-operators $e(u,\xi)$, $f(u,\xi)$ and $h(u,\xi)$ are \begin{eqnarray} \Delta e(u,\xi) &=& e(u+ i\pi c^{(2)}/4,\xi)\otimes 1+ \sum_{p=0} ^{\infty}(-1)^p \left(f(u+ i\pi c^{(2)}/4-i\tih,\xi)\right)^{p}\times\nonumber} \def\gamma{\gamma\\ &\times& h(u+ i\pi c^{(2)}/4,\xi)\otimes \left(e(u- i\pi c^{(1)}/4,\xi+c^{(1)}) \right)^{p+1}\ , \label{com-e-fu}\\ \Delta f(u,\xi)&=& 1\otimes f(u- i\pi c^{(1)}/4,\xi+c^{(1)}) + \sum_{p=0} ^{\infty} (-1)^p \left(f(u+ i\pi c^{(2)}/4,\xi)\right)^{p+1}\otimes\nonumber} \def\gamma{\gamma\\ &\otimes& \widetilde h(u- i\pi c^{(1)}/4,\xi+c^{(1)}) \left(e(u- i\pi c^{(1)}/4-i\tih,\xi+c^{(1)}) \right)^p\ , \label{com-f-fu}\\ \Delta h(u,\xi)&=&\sum_{p=0}^\infty (-1)^p {\sin\,(\pi (p+1)/\xi)\over \sin\,(\pi/\xi)} \left(f(u+ i\pi c^{(2)}/4-i\tih,\xi)\right)^{p} h(u+ i\pi c^{(2)}/4,\xi) \otimes\nonumber} \def\gamma{\gamma\\ &\otimes& h(u- i\pi c^{(1)}/4,\xi+c^{(1)}) \left(e(u- i\pi c^{(1)}/4-i\tih,\xi+c^{(1)}) \right)^p\ . \label{comul-h} \end{eqnarray} \subsection{Finite-dimensional representations and the intertwining operators}\label{3.3} Let $e$, $f$ and $h$ generators of the algebra $U_{i\pi/\xi}(sl_2)$ with the commutation relations: \begin{equation}} \def\bn{\beq\label{soot} [h,e]=2e,\qquad [h,f]=-2f,\qquad [e,f]= \frac{\sin\sk{\pi h/\xi}}{\sin\sk{\pi/\xi}}\ . \end{equation}} \def\ed{\eeq The following formulas describe the evaluation homomorphism of the algebra $\Ael$ at $c=0$ onto the algebra $U_{i\pi/\xi}(sl_2)$: \begin{eqnarray} {\cal E}v_z \sk{e(u)}&=& - {{\rm sh}\sk{i\pi/\xi} \over {\rm sh}\sk{\frac{u-z}{\xi}+\frac{i\pi(h-1)}{2\xi}} }\, e= - e\ {{\rm sh}\sk{i\pi\frac{\xi+1}{\xi}} \over {\rm sh}\sk{\frac{u-z}{\xi}+\frac{i\pi(h+1)}{2\xi}} }\ , \nonumber} \def\gamma{\gamma\\ {\cal E}v_z \sk{f(u)}&=&- {{\rm sh}\sk{i\pi/\xi} \over {\rm sh}\sk{\frac{u-z}{\xi}+\frac{i\pi(h+1)}{2\xi}} }\, f= -f\ {{\rm sh}\sk{i\pi\frac{\xi+1}{\xi}} \over {\rm sh}\sk{\frac{u-z}{\xi}+\frac{i\pi(h-1)}{2\xi}} }\ , \nonumber} \def\gamma{\gamma\\ {\cal E}v_z \sk{h(u)}&=& \cos\sk{\frac{i\pi h}{\xi}} - {\rm sh}\sk{\frac{i\pi}{\xi}}\left[ {\rm cth}\sk{\frac{u-z}{\xi}+\frac{i\pi(h-1)}{2\xi}}\ ef\right.\nonumber} \def\gamma{\gamma\\ &-&\left.{\rm cth}\sk{\frac{u-z}{\xi}+\frac{i\pi(h+1)}{2\xi}}\ fe \right].\label{eval-h-n} \end{eqnarray} Let $V_n$ be $(n+1)$-dimensional $U_{q}(sl_2)$-module with a basis $v_k$, $k=0,1,\ldots,n$ where the operators $h$, $e$ and $f$ act according to the rules \begin{equation}} \def\bn{\beq h\,v_k=(n-2k)\,v_k,\quad e\,v_k= \frac{\sin(\pi k \xi)}{\sin(\pi/\xi)}\,v_{k-1},\quad f\,v_k=\frac{\sin(\pi (n-k)/\xi)}{\sin(\pi/\xi)}\,v_{k+1}\ . \label{finite} \end{equation}} \def\ed{\eeq Combining these formulas with the evaluation homomorphism we can construct the level zero evaluation representations of the algebra ${\cal A}(\widehat{sl}_2)$ in the finite-dimensional space $V_n$. In particular, in what follows we need the evaluation representation of this algebra into two-dimensional space $V_1$. It is given by the formulas \begin{eqnarray} \pi_z\sk{e(u)}v_{+}&=&0,\quad \pi_z\sk{f(u)} v_{-}\, =\, 0\ , \label{eval-0}\\ \pi_z\sk{e(u)}v_{-}&=& - {{\rm sh}\sk{i \pi/\xi} \over {\rm sh}\sk{\frac{u-z}{\xi}} }\ v_{+}\ ,\quad \pi_z\sk{f(u)}v_{+}\ =\ - {{\rm sh}\sk{i \pi/\xi} \over {\rm sh}\sk{\frac{u-z}{\xi}} }\ v_{-}\ ,\label{eval-e-f1}\\ \pi_z\sk{h(u)}v_{\pm}&=&\cos\sk{\frac{\pi} {\xi}}\mp{\rm sh}\sk{\frac{i\pi}{\xi}} {\rm cth}\sk{\frac{u-z}{\xi}} v_{\pm} = {{\rm sh}\sk{\frac{u-z\mp i\pi}{\xi}} \over {\rm sh}\sk{\frac{u-z}{\xi}} } v_{\pm} \ .\label{eval-h1} \end{eqnarray} In these formulas we have identified $v_{+}=v_{0}$ and $v_{-}=v_{1}$. Using formulas \r{eval-0}--\r{eval-h1} we can define certain intertwining operators between level one highest weight modules over the algebra ${\cal A}(\widehat{sl}_2)$. It was shown in \cite{KLP} that the algebra ${\cal A}(\widehat{sl}_2)$ has the highest weight representations at the value of the central element $c=1$ which can be bosonized using one free continuous bosonic field. We denote this representation space by the symbol $\CH_R$ and will demonstrate in the next subsection that at the FF point it coincides with the bosonized version of the Hilbert space of the free massive Dirac field in the RRW $\CH_R^b$. In analogy with the group-theoretical description of the quantum integrable models on the infinite-dimensional lattice \cite{JM} we define four types of the {\it twisted} intertwining operators \begin{eqnarray} Z'(z)&:& \CH_R\to \CH_R\otimes V_{1}\ ,\quad Z^{\prime *}(z)\ :\ \CH_R\otimes V_{1} \to \CH_R\ ,\n \\ Z^{*}(z)&:&V_{1}\otimes \CH_R\to \CH_R\ ,\quad Z(z)\ :\ \CH_R\to V_{1}\otimes \CH_R\ . \label{inteq0 \end{eqnarray} The algebra ${\cal A}(\widehat{sl}_2)$ acts on the two-dimensional evaluation module $V_1$ by the formulas \r{eval-0}--\r{eval-h1}. We require that these operators commute with the action of the algebra ${\cal A}(\widehat{sl}_2)$ up to the the involution \r{L-oper-rel} \begin{eqnarray} Z(z) \iota(x) &=& \Delta(x) Z(z)\ ,\quad Z^{\prime *}(z)\Delta(x)\ =\ \iota(x)Z^{\prime *}(z)\ , \nonumber} \def\gamma{\gamma\\ Z^{*}(\theta)\Delta(x)&=& \iota(x)Z^{*}(z)\ ,\quad Z(z) \iota(x) \ =\ \Delta(x) Z(z)\ ,\quad \label{inteq} \end{eqnarray} where $x\in{\cal A}(\widehat{sl}_2)$. Due to the dimension of the module $V_1$ the intertwining operators have two components which are defined as follows: \begin{eqnarray} Z'(z) v &=& Z'_+(z)v\otimes v_+ + Z'_-(z)v\otimes v_-\ ,\quad Z^{\prime *}(z)(v\otimes v_\pm)\ =\ Z^{\prime *}_\pm(z)v \ , \nonumber} \def\gamma{\gamma\\ Z^{*}(z)(v_\pm\otimes v)&=& Z^{*}_\pm(z)v \ ,\quad Z(z) v \ =\ v_+\otimes Z_+(z)v + v_-\otimes Z_-(z)v\ , \nonumber} \def\gamma{\gamma \end{eqnarray} where $v\in\CH_R$. Using the coalgebraic structure of the algebra ${\cal A}(\widehat{sl}_2)$ we can rewrite the defining relations \r{inteq} for the components of the intertwining operators as commutativity with Gauss coordinates of $L$-operators. For $x=h(u)$ in \r{inteq} we have \begin{eqnarray} h(u)Z^*_-(\theta)h^{-1}(u)&=& \frac{{\rm sh}\sk{\frac{u-\theta+5i\pi/4}{\xi}}} {{\rm sh}\sk{\frac{u-\theta+ i\pi/4}{\xi}}}Z^*_-(\theta)\ ,\label{51a}\\ h(u)Z_+(\theta)h^{-1}(u)&=& \frac{{\rm sh}\sk{\frac{u-\theta+ i\pi/4}{\xi}}} {{\rm sh}\sk{\frac{u-\theta-3i\pi/4}{\xi}}}Z_+(\theta)\ ,\label{51b}\\ h(u)Z^{\prime*}_+(\alpha)h^{-1}(u)&=& \frac{{\rm sh}\sk{\frac{u-\theta-5i\pi/4}{\xi+1}}} {{\rm sh}\sk{\frac{u-\theta- i\pi/4}{\xi+1}}}Z^{\prime*}_+(\alpha)\ ,\label{51c}\\ h(u)Z'_-(\alpha)h^{-1}(u)&=& \frac{{\rm sh}\sk{\frac{u-\theta- i\pi/4}{\xi+1}}} {{\rm sh}\sk{\frac{u-\theta+3i\pi/4}{\xi+1}}}Z'_-(\alpha)\ .\label{51d} \end{eqnarray} For $e(u)$ we have \begin{eqnarray} {\rm sh}\,\sk{\frac{i\pi}{\xi}} Z^*_+(\theta)&=& {\rm sh}\,\sk{\frac{u-\theta+i\pi/4}{\xi}} e(u)Z_-^*(\theta)+\nonumber} \def\gamma{\gamma\\ &+& {\rm sh}\,\sk{\frac{u-\theta+5i\pi/4}{\xi}} Z_-^*(\theta)e(u)\ ,\label{52a}\\ {\rm sh}\,\sk{\frac{i\pi}{\xi}} Z_-(\theta)&=& {\rm sh}\,\sk{\frac{u-\theta-3i\pi/4}{\xi}} e(u)Z_+(\theta)+\nonumber} \def\gamma{\gamma\\ &+& {\rm sh}\,\sk{\frac{u-\theta+i\pi/4}{\xi}} Z_+(\theta)e(u)\ ,\label{52b}\\ \{e(u),Z_+^{\prime*}(\theta)\}&=&\{e(u),Z_-'(\theta)\}\ =\ 0\ ,\label{52c} \end{eqnarray} and finally for $f(u)$ \begin{eqnarray} {\rm sh}\,\sk{\frac{i\pi}{\xi+1}} Z^{\prime*}_-(\alpha)&=& {\rm sh}\,\sk{\frac{u-\alpha-i\pi/4}{\xi+1}} f(u)Z_+^{\prime*}(\alpha)+\nonumber} \def\gamma{\gamma\\ &+& {\rm sh}\,\sk{\frac{u-\alpha-5i\pi/4}{\xi+1}} Z_+^{\prime*}(\alpha)f(u)\ ,\label{53a}\\ {\rm sh}\,\sk{\frac{i\pi}{\xi+1}} Z'_+(\alpha)&=& {\rm sh}\,\sk{\frac{u-\alpha+3\pi/4}{\xi+1}} f(u)Z'_-(\alpha)+\nonumber} \def\gamma{\gamma\\ &+& {\rm sh}\,\sk{\frac{u-\alpha-i\pi/4}{\xi+1}} Z'_-(\alpha)f(u)\ ,\label{53b}\\ \{f(u),Z_-^*(\theta)\}&=&\{f(u),Z_+(\theta)\}\ =\ 0\ .\label{53c} \end{eqnarray} Note that in all commutation relations for the operators $Z_\pm(\theta)$ or $Z^*_\pm(\theta)$ appear only the trigonometric functions with the periods $2i\pi/\xi$ while in all those related to the operators $Z'_\pm(\alpha)$ or $Z^{\prime*}_\pm(\alpha)$ with the period $2i\pi/(\xi+1)$. This property is encoded into the comultiplication rules \r{com-e-fu}--\r{comul-h}. We did not write down all the relations following from \r{inteq} for the components of the intertwining operators but only independent ones. For example, the relation \r{51a} is obtained by applying \r{inteq} to the vector $v_-\otimes v\in V_1\otimes \CH_R$. If we apply it to the vector $v_+\otimes v$ we obtain the relation \begin{eqnarray} {\rm sh}\,\sk{\frac{2i\pi}{\xi}} Z^*_-(\theta)h(u)e(u-i\pi)&=& {\rm sh}\,\sk{\frac{u-\theta+i\pi/4}{\xi}} h(u)Z_+^*(\theta)-\nonumber} \def\gamma{\gamma\\ &-& {\rm sh}\,\sk{\frac{u-\theta-3i\pi/4}{\xi}} Z_+^*(\theta)h(u)\nonumber} \def\gamma{\gamma \end{eqnarray} which is a consequence of \r{52a} and \r{he}. Nevertheless, the defining relations \r{51a}--\r{53c} allow one to calculate some properties of the intertwining operators. For example, the commutation relations \begin{eqnarray} Z^*_{\nu_1}(\theta_1) Z^*_{\nu_2}(\theta_2) &=&\rho(\theta_1-\theta_2) \overline S_{\nu_1\nu_2}^{\nu'_1\nu'_2} (\theta_1-\theta_2,\xi) Z^*_{\nu'_2}(\theta_2) Z^*_{\nu'_1}(\theta_1)\ , \nonumber} \def\gamma{\gamma \\ Z'_{\varepsilon_2}(\alpha_2) Z'_{\varepsilon_1}(\alpha_1) &=&\rho'(\theta_1-\theta_2) \overline R_{\varepsilon_1\varepsilon_2}^{\varepsilon'_1\varepsilon'_2} (\alpha_1-\alpha_2,\xi) Z'_{\varepsilon'_1}(\alpha_1) Z'_{\varepsilon'_2}(\alpha_2)\ . \nonumber} \def\gamma{\gamma \end{eqnarray} can be proved using only \r{51a}-\r{53c} and supposing that the operators $Z^*_-(\theta_1)$, $Z^*_-(\theta_2)$ commute up to some scalar factor and analogously for the operators $Z'_-(\alpha_1)$, $Z'_-(\alpha_2)$. The defining relations \r{51a}--\r{53c} allow one to find the bosonization of the intertwining operators from the bosonization of the screening current algebra. Being specialized to the FF point $\xi=1$ the intertwining operators will coincide with the operators ${\cal Z}_\pm(\theta)$, and ${\cal Z}'_\pm(\alpha)$ constructed in the previous section modulo shifts in the spectral parameters. Also, all the scalar coefficients mentioned above can be fixed using these bosonizations. We will do this in the next subsection. \subsection{Bosonization of the screening operator algebra} The description of the infinite dimensional representations of the algebra ${\cal A}(\widehat{sl}_2)$ at non-zero value of the central element $C$ is divided into two steps. The first step is to rewrite the commutation relations \r{ef}--\r{hh} in terms of the total currents $E(u)$, $F(u)$ and $H(u)$: \begin{eqnarray} e\sk{u-\fract{ic\tih}{4}}+e\sk{u-i\pi\xi-\fract{ic\tih}{4}}&=& \xi\sin\,(\pi/\xi) E(u)\ ,\label{FDE}\\ f\sk{u+\fract{ic\tih}{4}}+f\sk{u-i\pi\xi'+\fract{ic\tih}{4}}&=& \xi'\sin\,(\pi/\xi') F(u)\ ,\label{FDF}\\ h(u)&=&2\pi\xi'\ \sin(\pi/\xi')\ H\sk{u+\frac{i\pi\xi}{2}+\frac{i\pi c}{4}} \ .\label{H-new} \end{eqnarray} We write the commutation relations for the total currents in the form adequate for the category of the highest weight representations: \begin{equation}} \def\bn{\beq\label{EF} [E(u),F(v)]= \left[ \delta\sk{u-v-\frac{i\pi c}{2}}H\sk{v+\frac{i\pi(\xi+c)}{2}} - \delta\sk{u-v+\frac{i\pi c}{2}}H\sk{v-\frac{i\pi(\xi+c)}{2}} \right], \end{equation}} \def\ed{\eeq \begin{eqnarray} { \Ga{\frac{1}{2}+\frac{1}{\xi}+\frac{i(u-v)}{\pi\xi}} \over \Ga{\frac{1}{2}-\frac{1}{\xi}+\frac{i(u-v)}{\pi\xi}} } H(u)E(v) &=& E(v)H(u) { \Ga{\frac{1}{2}+\frac{1}{\xi}-\frac{i(u-v)}{\pi\xi}} \over \Ga{\frac{1}{2}-\frac{1}{\xi}-\frac{i(u-v)}{\pi\xi}} }, \label{HE}\\ { \Ga{\frac{1}{2}-\frac{1}{\xi'}+\frac{i(u-v)}{\pi\xi'}} \over \Ga{\frac{1}{2}+\frac{1}{\xi'}+\frac{i(u-v)}{\pi\xi'}} } H(u)F(v) &=& F(v)H(u) { \Ga{\frac{1}{2}-\frac{1}{\xi'}-\frac{i(u-v)}{\pi\xi'}} \over \Ga{\frac{1}{2}+\frac{1}{\xi'}-\frac{i(u-v)}{\pi\xi'}} }, \label{HF}\\ { \Ga{1+\frac{1}{\xi}+\frac{i(u-v)}{\pi\xi}} \over \Ga{-\frac{1}{\xi}+\frac{i(u-v)}{\pi\xi}} } E(u)E(v) &=& -E(v)E(u) { \Ga{1+\frac{1}{\xi}-\frac{i(u-v)}{\pi\xi}} \over \Ga{-\frac{1}{\xi}-\frac{i(u-v)}{\pi\xi}} }, \label{EE}\\ { \Ga{1-\frac{1}{\xi'}+\frac{i(u-v)}{\pi\xi'}} \over \Ga{\frac{1}{\xi'}+\frac{i(u-v)}{\pi\xi'}} } F(u)F(v) &=& -F(v)F(u) { \Ga{1-\frac{1}{\xi'}-\frac{i(u-v)}{\pi\xi'}} \over \Ga{\frac{1}{\xi'}-\frac{i(u-v)}{\pi\xi'}} }, \label{FF} \end{eqnarray} \begin{eqnarray} &{ \Ga{1+\frac{1}{\xi}+\frac{i(u-v)}{\pi\xi}} \over \Ga{1-\frac{1}{\xi}+\frac{i(u-v)}{\pi\xi}} } { \Ga{1-\frac{1}{\xi'}+\frac{i(u-v)}{\pi\xi'}} \over \Ga{1+\frac{1}{\xi'}+\frac{i(u-v)}{\pi\xi'}} } H(u)H(v) =\nonumber} \def\gamma{\gamma\\ &\qquad=H(v)H(u) { \Ga{1+\frac{1}{\xi}-\frac{i(u-v)}{\pi\xi}} \over \Ga{1-\frac{1}{\xi}-\frac{i(u-v)}{\pi\xi}} } { \Ga{1-\frac{1}{\xi'}-\frac{i(u-v)}{\pi\xi'}} \over \Ga{1+\frac{1}{\xi'}-\frac{i(u-v)}{\pi\xi'}} }. \label{HH} \end{eqnarray} The commutation relations for the total currents \r{HE}--\r{HH} are written in the form of equalities of the meromorphic functions without poles and zeros \cite{DK,DKKP}. This means that the product of the currents has the structure of poles and zeros defined by the zeros and poles of the function which is in front of this product in the commutation relations \r{HE}--\r{HH}. For example, the product $E(u)E(v)$ has poles at the points $u=v-i\pi+i\pi\xi k$ and zeros at the points $u=v+i\pi+i\pi\xi (k+1)$, $k\geq0$. The Frenkel-Ding \cite{DF} formulas \r{FDE} and \r{FDF} for the total currents can be inverted solving the Riemann-Hilbert problem associated with the strips of the widths $\pi\xi$ and $\pi\xi'=\pi(\xi+c)$ for the currents $E(u)$ and $F(u)$ respectively \cite{KLP} \begin{eqnarray} e(u)&=&\sin\,\pi/\xi\ \int_{C} {dv\over2\pi i}\ { E(v)\over{\rm sh}\,\frac{u-v+ ic\tih/4}{\xi}}\ ,\label{2e}\\ f(u)&=&\sin\,\pi/\xi'\ \int_{C'} {dv\over2\pi i}\ { F(v)\over{\rm sh}\, \frac{u-v- ic\tih/4}{\xi'}}\ ,\label{2f} \end{eqnarray} where the contour $C'$ goes from $-\infty$ to $+\infty$, the points $u+ic\tih/4+ik\pi\xi'$ ($k\geq0$) are above the contour and the points $u-ic\tih/4-ik\pi\xi'$ ($k\geq0$) are below the contour. The contour $C$ also goes from $-\infty$ to $+\infty$ but the points $u-ic\tih/4+ik\pi\xi$ ($k\geq0$) are above the contour and the points $u+ic\tih/4-ik\pi\xi$ ($k\geq0$) are below the contour. The second step is the bosonization of the currents $E(u)$, $F(u)$ and $H(u)$. To describe the symmetries of the SG model we need the bosonization of the algebra ${\cal A}(\widehat{sl}_2)$ at the value of the central element $c=1$. To construct this bosonization we define the continuous Heisenberg operators $a_\lambda$ which satisfy the commutation relations \begin{equation}} \def\bn{\beq\label{gen-case} [a_\lambda,a_\mu]=\lambda\, \frac{{\rm sh}\,\frac{\pi\lambda}{2}}{{\rm sh}\,\pi\lambda}\, \frac{{\rm sh}\,\frac{\pi\lambda(\xi+1)}{2}}{{\rm sh}\,\frac{\pi\lambda\xi}{2}} \delta(\lambda+\mu)=c(\lambda)\delta(\lambda+\mu)\ . \end{equation}} \def\ed{\eeq The commutation relations of the currents $E(u)$, $F(u)$ and $H(u)$ are satisfied by the operators \begin{eqnarray} E(u)&=&\exp\sk{2i{{\cal Q}}+2\int_{-\infty}^\infty\hspace{-20.5pt}-\hspace{11.5pt}{d\lambda\over\lambda}\ e^{iu\lambda}{\rm ch}\sk{\pi\lambda/2}\ a_\lambda} \ ,\label{E-bos}\\ F(u)&=&\exp\sk{-\frac{2i\xi}{\xi+1}\ {{\cal Q}}-2 \int_{-\infty}^\infty\hspace{-20.5pt}-\hspace{11.5pt}{d\lambda\over\lambda}\ e^{iu\lambda} \frac{ {\rm ch}\sk{\pi\lambda/2} {\rm sh}\sk{\pi\lambda\xi/2}} {{\rm sh}\sk{\pi\lambda(\xi+1)/2}} \ a_\lambda} \ ,\label{F-bos}\\ H(u)&=&\exp\sk{\frac{2i}{\xi+1}\ {{\cal Q}}+ \int_{-\infty}^\infty\hspace{-20.5pt}-\hspace{11.5pt}{d\lambda\over\lambda}\ e^{iu\lambda} \frac{ {\rm sh}\sk{\pi\lambda}} {{\rm sh}\sk{\pi\lambda(\xi+1)/2}} \ a_\lambda} \ ,\label{H-bos} \end{eqnarray} To verify this statement we should use the normal ordering rule given by \r{normal} with the function $c(\lambda)$ specified in \r{gen-case} and formulas given in the Appendix \ref{app2}. Note that at the FF point the Heisenberg operators $a_\lambda$ become the same as \r{i15} of the nonlocal integrals of motion \r{ii6} so the bosonization of the current $E(u)$ coincides with the bosonization \r{E-scr}, the bosonization of the current $F(u)$ coincides with the bosonization of the scattering operator ${\cal Z}_-(u)$ and $H(u)$ with $\Lambda_+(u)$ (cf. \r{i17}). The commutation relations \r{HE}--\r{HH} become in this case $$ [H(u),E(v)]=[E(u),E(v)]=\{H(u),F(v)\}=\{F(u),F(v)\}=\{H(u),H(v)\}=0\ . $$ The commutation relation \r{EF} being multiplied by $e^{u-v}$ and integrated over the parameter $u$ becomes the relation \r{Z+Z-} which relates the components of the scattering data operators ${\cal Z}_\pm(\theta)$. The formulas \r{51a}--\r{53c} allow to bosonize the components of the intertwining operators. It is given by the following formulas \cite{Lu1,KLP} \begin{eqnarray} Z_+(\theta)&=& \exp\left(-i{{\cal Q}}-\int_{-\infty}^\infty\hspace{-20.5pt}-\hspace{11.5pt} {d\lambda\over\lambda}\ {\bf e}^{-i\lambda (\theta+\pi i/2)}\ a_\lambda\right) \ ,\label{Psi-}\\ Z_-(\theta) &=& \int_{C} {du\over2\pi}\ e^{(u-\theta)/\xi} \left[(q)^{1/2} E(u) Z_+(\theta)-(q)^{-1/2}Z_+(\theta)E(u)\right]\ ,\label{Psi+}\\ Z^*_\pm(\theta)&=&Z_{\mp}(\theta-i\tih)\ ,\label{Psidual}\\ Z'_-(\alpha)&=& \exp\sk{\frac{i\xi}{\xi+1}\ {{\cal Q}}+ \int_{-\infty}^\infty\hspace{-20.5pt}-\hspace{11.5pt}{d\lambda\over\lambda}\ e^{-i\lambda(\alpha-\pi i/2)} \frac{ {\rm sh}\sk{\pi\lambda\xi/2}} {{\rm sh}\sk{\pi\lambda(\xi+1)/2}} \ a_\lambda} ,\label{Phi-}\\ Z'_+(\alpha) &=& \int_{C'} {du\over2\pi}\ e^{(u-\alpha)/(\xi+1)} \left[(q')^{1/2} Z'_+(\alpha)F(u)-(q')^{-1/2}F(u)Z'_+(\alpha)\right]\ , \label{Phi+}\\ Z^{'*}_\pm(\alpha)&=&Z'_{\mp}(\alpha+i\tih)\ , \label{Phidual} \end{eqnarray} where $q$ and $q'$ are given by \r{s-mul} and \r{q'-def} respectively and the contour $C'$ goes from $-\infty$ to $+\infty$ along the real axis leaving the points $z+i\tih/2+ik\pi(\xi+1)$ ($k\geq0$) above the contour and the points $z-i\tih/2-ik\pi(\xi+1)$ ($k\geq0$) below the contour. The contour $C$ also goes from $-\infty$ to $+\infty$ but the points $z-i\tih/2+ik\pi\xi$ ($k\geq0$) are above the contour and the points $z+i\tih/2-ik\pi\xi$ ($k\geq0$) are below the contour. The formulas \r{Psi-}, \r{Psi+} and \r{Phi-}, \r{Phi+} demonstrate that at the FF point the operators $Z_\pm(\theta)$ coincide with the operators ${\cal Z}_\mp(\theta+\pi i/2)$ modulo certain normalization constants. The same is true for the relation between operators $Z'_\pm(\alpha)$ and ${\cal Z}'_\pm(\alpha-\pi i/2)$. The second remark concerns the form of the contour $C$ in the relation \r{Psi+}. The form of this contour is shown on the Fig.~2. \bigskip \bigskip \unitlength 1mm \linethickness{0.4pt} \begin{picture}(130.67,38.67) \bezier{148}(10.33,18.67)(42.67,18.67)(47.33,17.33) \bezier{40}(47.33,17.33)(52.33,15.67)(56.33,12.33) \bezier{44}(56.33,12.33)(59.67,9.33)(66.00,7.67) \bezier{68}(66.00,7.67)(77.00,5.33)(83.00,7.00) \bezier{84}(83.00,7.00)(92.67,12.33)(84.00,16.33) \bezier{92}(84.00,16.33)(73.67,20.33)(62.00,24.00) \bezier{80}(62.00,24.00)(53.33,28.67)(62.00,33.00) \bezier{92}(62.00,33.00)(73.67,37.33)(83.00,33.33) \bezier{48}(83.00,33.33)(91.00,30.00)(91.33,26.67) \bezier{32}(91.33,26.67)(93.00,22.33)(96.33,22.00) \bezier{76}(96.33,22.00)(105.67,21.00)(115.33,21.00) \put(130.33,21.00){\vector(1,0){0.2}} \put(115.33,21.00){\line(1,0){15.00}} \put(79.33,30.00){\makebox(0,0)[cc]{$\bullet$}} \put(79.33,11.33){\makebox(0,0)[cc]{$\circ$}} \put(75.50,13.00){\makebox(0,0)[rc]{$\theta_1-i\tih/2$}} \put(75.50,28.00){\makebox(0,0)[rc]{$\theta_1+i\tih/2$}} \put(130.67,25.67){\makebox(0,0)[cc]{$+\infty$}} \put(10.33,26.00){\makebox(0,0)[cc]{$-\infty$}} \put(10.00,9.33){\makebox(0,0)[cc]{Fig.~2.}} \put(79.33,38.67){\makebox(0,0)[cc]{$\circ$}} \put(75.50,38.67){\makebox(0,0)[rc]{$\theta_1-i\pi/2+i\pi\xi$}} \put(79.33,0.67){\makebox(0,0)[cc]{$\bullet$}} \put(75.50,0.67){\makebox(0,0)[rc]{$\theta_1+i\pi/2-i\pi\xi$}} \put(85.33,30.00){\makebox(0,0)[cc]{$\bullet$}} \put(92.33,30.00){\makebox(0,0)[lc]{$\theta_2+i\pi/2$}} \put(85.33,0.67){\makebox(0,0)[cc]{$\bullet$}} \put(91.33,0.67){\makebox(0,0)[lc]{$\theta_2+i\pi/2-i\pi\xi$}} \put(85.33,-2.67){\makebox(0,0)[rc]{$\vdots$}} \put(79.33,-2.67){\makebox(0,0)[rc]{$\vdots$}} \put(79.33,43.67){\makebox(0,0)[rc]{$\vdots$}} \end{picture} \bigskip \bigskip \noindent We can see from this picture that in the limit to the FF point $(\xi\to1)$ there is a double pinching of the integral which leads to the relation \r{Z+Z-} where the integrals are calculated as the residues in the points $u=\theta\pm\pi i/2$. Moreover, this figure demonstrates that the product $Z_+(\theta_1)Z_-(\theta_2)$ has the pole when $\theta_1\to\theta_2+\pi i$ because of the pinching of the contour in the integral representation of this quantity. The origin of this pole due to pinching does not yield the restriction on the domain of the definition of this product and $Z_+(\theta_1)Z_-(\theta_2)$ is an meromorphic function of the variable $\theta_1-\theta_2$ in the domain ${\rm Im}\,(\theta_1-\theta_2)>-\pi i/2$ with a simple pole in the point $\theta_1\to\theta_2+\pi i$. Using standard techniques \cite{Lu1,KLP} we can find the properties of the intertwining operators: \begin{eqnarray} Z_{\nu_1}(\theta_1) Z_{\nu_2}(\theta_2) &=&S_{\nu_1\nu_2}^{\nu'_1\nu'_2} (\theta_1-\theta_2,\xi) Z_{\nu'_2}(\theta_2) Z_{\nu'_1}(\theta_1)\ , \label{ZFII} \\ Z'_{\varepsilon_2}(\alpha_2) Z'_{\varepsilon_1}(\alpha_1) &=& R_{\varepsilon_1\varepsilon_2}^{\varepsilon'_1\varepsilon'_2} (\alpha_1-\alpha_2,\xi) Z'_{\varepsilon'_1}(\alpha_1) Z'_{\varepsilon'_2}(\alpha_2)\ , \label{ZFI} \\ Z_{\nu}(\theta) Z'_{\varepsilon}(\alpha) &=&\nu\varepsilon\,{\rm tg}\left( {i(\theta-\alpha)\over2} -{\pi\over4} \right) Z'_{\varepsilon}(\alpha) Z_{\nu}(\theta)\ , \label{ZFI-II}\\ \label{complet} \sum_{\varepsilon=\pm} Z^{\prime*}_{\varepsilon}(\alpha)Z'_{\varepsilon}(\alpha) &=&g'(\xi)\ {\rm id}\ ,\\ \label{orthI} Z'_{\varepsilon_1}(\alpha)Z^{\prime*}_{\varepsilon_2}(\alpha) &=&g'(\xi)\delta_{\varepsilon_1\varepsilon_2}\ {\rm id}\ ,\\ \label{orthII} Z_{\varepsilon_1}(\theta_1)Z^{*}_{\varepsilon_2}(\theta_2) &=&{g(\xi)\delta_{\varepsilon_1\varepsilon_2}\, {\rm id} \over \theta_1-\theta_2} + o(z_1-z_2)\ , \end{eqnarray} where the $S$ and $R$-matrices is given by \r{S-mat} and \r{RR-mat} respectively and the normalization constants $g(\xi)$, $g'(\xi)$ can be expressed through double $\Gamma$-functions using the formulas given in the Appendix \ref{app2} \cite{KLP}. \setcounter{equation}{0} \section{Angular quantization} Before starting this section we would like to fix the terminology and explain what we mean by the angular quantization in the context of integrable quantum field theory. By this term we mean the possibility to represent the states and operators in the total Hilbert space of the model associated with total space-time as some operators acting in the Hilbert space associated with RRW. So, considering the free fermion in RRW in the second section we did not really consider the angular quantization but did only some preliminary work. The angular quantization of SG model will be considered in this section. But first we would like to recall the angular quantization in lattice integrable models inspired by Baxter's corner transfer matrix method. \subsection{Angular quantization on the lattice} In a series of papers, see e.g. \cite{XXZ,JM} the precise mathematical description of anti-ferroelectric XXZ model in thermodynamic limit was developed in terms of representation theory of quantum affine Lie algebra $U_q(\widehat{sl}_2)$ with the real deformation parameter $-1<q<0$. This description, based on Baxter's corner transfer matrix method, looks as follows. The total Hilbert space of the theory is identified with the space of endomorphisms ${\rm End}(\Lambda_0\oplus\Lambda_1)$ of direct sum of the level one irreducible $U_q(\widehat{sl}_2)$ modules with (complex linear) scalar product given by the natural prescription \begin{equation}} \def\bn{\beq\label{Ksc} (A,B)={\rm Tr}\,_{\Lambda_0\oplus\Lambda_1}AB\ . \end{equation}} \def\ed{\eeq Two components of degenerated vacuum are identified, up to the constant, with $(-q)^{D^{(1)}}$, where $D^{(1)}$ is principal gradation operator for quantum affine algebra, multiplied by the projection to $\Lambda_i$. The representation theory of $U_q(\widehat{sl}_2)$ provides two types of operators \begin{equation}} \def\bn{\beq\label{vertex} \Phi(\zeta): \Lambda_i\to \Lambda_{1-i}\otimes V_\zeta,\quad \Psi^*(\zeta): V_\zeta\otimes\Lambda_i\to \Lambda_{1-i},\quad i=0,1 \end{equation}} \def\ed{\eeq which commute with the action of $U_q(\widehat{sl}_2)$. Here $V_\zeta$ is a two-dimensional representation of $U_q(\widehat{sl}_2)$ with basis $v_\pm$ evaluated at the point $\zeta^2$. The transfer matrix $T(\zeta)$ of the theory acts on the state $A\in{\rm End}(\Lambda_0+\Lambda_1)$ as \begin{equation}} \def\bn{\beq\label{trans} T(\zeta)\cdot A=\sum_{\varepsilon=\pm}\Phi_\varepsilon(\zeta)\ A\ \Phi_{-\varepsilon}(\zeta) \end{equation}} \def\ed{\eeq and the eigenvectors of the transfer matrix are described in terms of the second type intertwining operators: \begin{equation}} \def\bn{\beq\label{st-lat} |\xi_n,\ldots,\xi_1\rangle_{\varepsilon_n,\ldots,\varepsilon_1;(i)}= c(n)\Psi^*_{\varepsilon_n}(\xi_n)\ldots \Psi^*_{\varepsilon_1}(\xi_1) (-q)^{D^{(i)}}\ . \end{equation}} \def\ed{\eeq The local spin operators $\sigma_n^\pm$ acting on the $n$th cite of the lattice can be described in terms of operators $\Phi(\zeta)$. Due to the definition of scalar product it gives the expressions of the correlation functions of finite products of operators $\sigma_n^\pm$ and of the form-factors of a local operator in terms of traces of products operators $\Phi(\zeta)$ and $\Psi^*(\xi)$ in the Fock space $\Lambda_0\oplus\Lambda_1$ \cite{JM}. Moreover, the adjoint (in a sense of Hopf algebra) action of $U_q(\widehat{sl}_2)$ equips the space of states with a structure of level 0 $U_q(\widehat{sl}_2)$-module, such that $n$-particle states form $n$-fold tensor products of the two-dimensional representations of $U_q(\widehat{sl}_2)$. \subsection{Angular quantization in the 2d field theory} A counterpart of the CTM ideology in the integrable models of the 2d quantum field theory in the infinite volume looks as follows \cite{BrL}. Let $\CH_R$ be a Hilbert space of canonical quantization of a theory in the RRW, where boost operator $K=-i\partial_\alpha$ is considered as Hamiltonian. Here $\partial_\alpha$ is the differentiation with respect to the angular `time' or, what is the same, with respect of the spectral parameters (see \r{stpr}). The total Hilbert space $\CH$ of the model is supposed to be a properly defined subspace of ${\rm End}\,\CH_R$ with the scalar product $(A,B)={\rm Tr}\,_{\CH_R}A\cdot B$. The vacuum state in $\CH$ is identified with the operator $e^{\pi K}$ in $\CH_R$ and the definition of the transfer matrix refers to certain quantum version of Jost functions \cite{Lu1}, $Z'_\pm(\alpha)$ (here $\alpha$ is the spectral parameter) whose precise construction on the quantum level is not known. On the classical level these objects in SG theory were introduced by S.Lukyanov in \cite{Lu2} using zero curvature representation of SG equation in RRW: $[\partial_r-A_r, \partial_\alpha-A_\alpha]=0$. The asymptotic states $|\theta_1,\ldots,\theta_n\rangle_{\varepsilon_1,\ldots,\varepsilon_n}$ are presented by the products of the operators \begin{equation}} \def\bn{\beq\label{states} |\theta_1,\ldots,\theta_n\rangle_{\varepsilon_1,\ldots,\varepsilon_n}= Z^*_{\varepsilon_1}(\theta_1)\ldots Z^*_{\varepsilon_n}(\theta_n)e^{\pi K}, \end{equation}} \def\ed{\eeq analytically continued to the real line, where $Z^*_{\pm}(\theta)$ are certain operators acting in the RRW Hilbert space $\CH_R$. They can be represented by the bosonized expressions \r{Psi-}-\r{Psidual}. The conjugated states are given by the product of the operators $Z_\pm(\theta)=Z^*_\mp(\theta+\pi i)$: \begin{equation}} \def\bn{\beq\label{states-co} _{\varepsilon_1,\ldots,\varepsilon_n}\langle\theta_1,\ldots,\theta_n|=e^{\pi K} Z_{\varepsilon_1}(\theta_1)\ldots Z_{\varepsilon_n}(\theta_n). \end{equation}} \def\ed{\eeq Analogously to the lattice case, one can assume that any local operator in the theory can be presented in this language in terms of left and right multiplications of certain combinations of the operators $Z'_{\varepsilon}(\alpha)$ and thus form-factor of operator $O$ can be given by some trace formula \begin{equation}} \def\bn{\beq\label{tracecon} _{\rm ph}\langle\mbox{vac}| O |\theta_1,\ldots,\theta_n\rangle_{\varepsilon_1,\ldots,\varepsilon_n}= {\rm Tr}\,_{\CH_R}\sk{e^{2\pi K}\widetilde O Z^*_{\varepsilon_1}(\theta_1)\ldots Z^*_{\varepsilon_n}(\theta_n) }, \end{equation}} \def\ed{\eeq where $\widetilde O$ is some operator acting in $\CH_R$ and corresponding to the original operator $O$. The problem to find an expression for the operator $\widetilde O$ in terms of the quantum Jost operators $Z'_\pm(\alpha)$ is a complicated problem and has no general solution for the arbitrary operator $O$ although for some simple operators it can be solved by comparing the form factors obtained in the framework of the bootstrap program with those obtained by means of the formula \r{tracecon} (see \cite{Lu1,NPT} for the simplest examples in case of the $SU(2)$-invariant Thirring model). We understand the trace in \r{tracecon} as properly regularized to produce the known form factor formulas in SG theory given in \cite{Sbook} (see the paper \cite{Ni} for the alternative formulation of a continuum analogue of the Baxter corner matrix method). The possibility to present the matrix element $_{\rm ph}\langle\mbox{vac}| O |\theta_1,\ldots,\theta_n\rangle_{\varepsilon_1,\ldots,\varepsilon_n}$ as a trace \r{tracecon}, the relation \r{Psidual} and the fact that the operators $Z_\pm(\theta)$ commute with the operators $\widetilde O$ up to numbers related to the locality index \cite{Lu1} allows to demonstrate easily the crossing symmetry of these matrix elements. We have \begin{eqnarray}\label{crossing} &_{\varepsilon'_1,\ldots,\varepsilon'_{n'}}\langle\theta'_1,\ldots,\theta'_{n'}| O |\theta_1,\ldots,\theta_n\rangle_{\varepsilon_1,\ldots,\varepsilon_n}=\nonumber} \def\gamma{\gamma\\ &\qquad= {\rm Tr}\,_{\CH_R}\sk{e^{2\pi K} Z_{\varepsilon'_1}(\theta'_1)\ldots Z_{\varepsilon'_{n'}}(\theta'_{n'}) \widetilde O Z^*_{\varepsilon_1}(\theta_1)\ldots Z^*_{\varepsilon_n}(\theta_n) }=\nonumber} \def\gamma{\gamma\\ &\qquad=\ \ _{\rm ph}\langle\mbox{vac}| O |\theta_1,\ldots,\theta_{n}, \theta'_1-i\pi,\ldots,\theta'_{n'}-i\pi\rangle _{\varepsilon_1,\ldots,\varepsilon_{n},-\varepsilon'_1,\ldots,-\varepsilon'_{n'}}\ . \end{eqnarray} Using the trace formulas we can also verify the completeness of the space of states \r{states} and \r{states-co} with respect to the scalar product given by the trace over RRW Hilbert space $\CH_R$. First of all we observe that the matrix element \r{tracecon} of the unity operator vanishes identically because after substitution of the integral representations of the operators $Z_+^*(\theta)$ \r{Psi+} in \r{tracecon} we obtain the integral with the integrand being the total difference which leads to the vanishing of the integral \cite{NPT}. On the other hand the pairing of the states \r{states} and \r{states-co} does not vanish identically but is proportional to some combinations of the $\delta$-functions. In particular, the simplest pairing of the one-particle states is equal to $$ _{\varepsilon'}\langle\theta'|\theta\rangle_{\varepsilon}=\delta_{\varepsilon\ep'}\delta(\theta-\theta'). $$ The delta-function in this formula appears because the trace ${\rm Tr}\,_{\CH_R}\sk{e^{\sigma K} Z_{\varepsilon'}(\theta')Z^*_{\varepsilon}(\theta) }$ has two simple poles in the points $\theta'=\theta$ due to \r{orthII} and in the point $\theta=\theta'+\sigma + 2\pi i$ due to the trace properties. When the parameter $\sigma$ tends to the value $-2\pi i$ these two poles form the $\delta$-function (see \cite{JM} for the detailed description of this mechanism in a case of lattice integrable models). We would like to note here that the same mechanism is responsible for the fact that form factors of the local operators satisfy the annihilation axiom \cite{Sbook}. For XXZ model this fact was established in \cite{P}. These are the general features of the angular quantization approach in the 2d integrable field theory. In order for the angular quantization approach be the self-consistent, in particular the traces \r{tracecon} satisfy all the axioms in the form-factor approach \cite{Sbook}, the operators $Z^*_\pm(\theta)$ and $Z'_\pm(\alpha)$ should satisfy the properties \r{ZFII}--\r{orthII}. Since the representation theory of the algebra ${\cal A}(\widehat{sl}_2)$ contains the operators which satisfy such properties we claim that this algebra is the dynamical symmetry algebra of the SG model in the sense claimed in \cite{XXZ} for the quantum XXZ model. Using the properties of the operators $Z'_\pm(\alpha)$ we can find representations of the commutation relations \r{YBt} for the quantum monodromy matrices and interpret its trace as the generating function of the local integrals of motion through the asymptotical expansion. This will be done in the next subsection. Moreover, we can define appropriate adjoint action of the algebra ${\cal A}(\widehat{sl}_2)$ onto the Hilbert space of the SG model $\CH$ which describes the known symmetries of this space of states related to the quantum affine algebra $U_q(\widehat{sl}_2)$ \cite{BL} and interpret these symmetries as level zero action of the algebra ${\cal A}(\widehat{sl}_2)$ in the Hilbert space of states. In the last subsection we will demonstrate that these symmetries being specialized to the FF point become the symmetries governed by the classical affine algebra at level zero and associated with the strip \cite{KLPST}. \subsection{Properties of the monodromy matrix in SG model} In this and next subsections we will understand by the operators $Z^*_\pm(\theta)$, $Z_\pm(\theta)$, $Z'_\pm(\alpha)$ and $Z^{'*}_\pm(\alpha)$ the intertwining operators of the algebra ${\cal A}(\widehat{sl}_2)$ which satisfy the properties \r{ZFII}--\r{orthII}. A monodromy matrix of the model acting on any state $X_k\in \CH_k$ of the total Hilbert space $\CH= \CH_0\oplus\CH_1$, where $k=0,1$ corresponds to the subspaces of $\CH$ of the even and odd number of particles respectively, is defined as follows \begin{equation}} \def\bn{\beq\label{monod1} {\cal T}_{\varepsilon\ep'}(\alpha)\cdot X_k= \sk{g'(\xi)}^{-1}\ \varepsilon^k Z'_\varepsilon(\alpha)\cdot X_k\cdot Z'_{-\varepsilon'}(\alpha),\quad k=0,1\ . \end{equation}} \def\ed{\eeq The commutation relations \r{ZFI} allow to find the commutation relation for this matrices: \begin{equation}} \def\bn{\beq\label{com-mon} R(\alpha_1-\alpha_2,\xi){\cal T}_1(\alpha_1){\cal T}_2(\alpha_2)= {\cal T}_2(\alpha_2){\cal T}_1(\alpha_1) R(\alpha_1-\alpha_2,\xi)\ \end{equation}} \def\ed{\eeq which coincides with \r{YBt}. The trace of the monodromy matrix or the transfer matrix $T(\alpha)$ is \begin{equation}} \def\bn{\beq\label{tr-monod} T(\alpha)\cdot X_k=\sk{g'(\xi)}^{-1}\ \sum_{\varepsilon=\pm}\varepsilon^k Z'_\varepsilon(\alpha)\cdot X_k\cdot Z'_{-\varepsilon}(\alpha)\ . \end{equation}} \def\ed{\eeq The inverse transfer matrix is given in terms of the operators $Z^{\prime*}_\pm(\alpha)$: \begin{equation}} \def\bn{\beq\label{tr-monod-in} T^{-1}(\alpha)\cdot X_k=\sk{g'(\xi)}^{-1}\ \sum_{\varepsilon=\pm}\varepsilon^k Z^{\prime*}_\varepsilon(\alpha) \cdot X_k\cdot Z^{\prime*}_{-\varepsilon}(\alpha)\ . \end{equation}} \def\ed{\eeq The fact that operators \r{tr-monod} and \r{tr-monod-in} are inverse to each other is a direct consequence of the properties \r{complet} and \r{orthI}. The same properties allow to prove that the physical vacuum vector $|\mbox{vac}\rangle_{\rm ph}\in\CH$ is stable under the action of the operators $T(\alpha)$ and $T^{-1}(\alpha)$: \begin{equation}} \def\bn{\beq\label{vac-inv} T(\alpha)|\mbox{vac}\rangle_{\rm ph}=\sk{g'(\xi)}^{-1}\ \sum_{\varepsilon=\pm}Z'_\varepsilon(\alpha)\cdot e^{\pi K}\cdot Z'_{-\varepsilon}(\alpha)\rav{Phidual} e^{\pi K}\sum_{\varepsilon=\pm}Z^{\prime*}_{-\varepsilon}(\alpha) Z'_{-\varepsilon}(\alpha) \rav{complet}|\mbox{vac}\rangle_{\rm ph}\ . \end{equation}} \def\ed{\eeq Here and below we will often use the formulas \begin{equation}} \def\bn{\beq\label{boost-act} Z^*_\pm(\theta) e^{\pi K}= e^{\pi K}Z_\mp(\theta),\quad Z'_\pm(\alpha) e^{\pi K}= e^{\pi K}Z^{\prime*}_\mp(\theta) \end{equation}} \def\ed{\eeq which are consequences of the definition of the boost operator and \r{Psidual}, \r{Phidual}. The commutation relations \r{com-mon} imply the commutativity \begin{equation}} \def\bn{\beq\label{tr-mo-co} [T(\alpha_1),T(\alpha_2)]=0 \end{equation}} \def\ed{\eeq which signifies that the operator $T(\alpha)$ can be considered as the generating function of the local integrals of motion. Using the property \r{ZFI-II} we can calculate the action of the generating function $T(\alpha)$ onto $n$-particle state $|\theta_1,\ldots,\theta_n\rangle_{\varepsilon_1,\ldots,\varepsilon_n}$: \begin{equation}} \def\bn{\beq\label{tr-act} T(\alpha)\cdot |\theta_1,\ldots,\theta_n\rangle_{\varepsilon_1,\ldots,\varepsilon_n} =\prod_{j=1}^n \varepsilon_j\ {\rm ctg}\sk{\frac{\pi}{4}+\frac{\alpha-\theta_j}{2i}} |\theta_1,\ldots,\theta_n\rangle_{\varepsilon_1,\ldots,\varepsilon_n}\ . \end{equation}} \def\ed{\eeq Using this equality we can see that the quantity \begin{equation}} \def\bn{\beq\label{gen-int} I(\alpha)=\sum_s I_se^{s\alpha}= \frac{1}{2i}\ T^{-1}(\alpha)\frac{\partial T(\alpha)}{\partial\alpha}= \frac{1}{2i}\ \frac{\partial\ln T(\alpha)}{\partial\alpha}, \end{equation}} \def\ed{\eeq has an eigenvalue on the states $|\theta_N,\ldots,\theta_1\rangle_{\varepsilon_N,\ldots,\varepsilon_1}$ $$ I(\alpha)|\theta_N,\ldots,\theta_1\rangle_{\varepsilon_N,\ldots,\varepsilon_1} =\sum_{j=1}^N\frac{2}{{\rm ch}(\alpha-\theta_j)} |\theta_N,\ldots,\theta_1\rangle_{\varepsilon_N,\ldots,\varepsilon_1} $$ and is a generating function of the local integrals of motion \r{int-n} $I_n$ and $\overline I_n$ for odd indeces $n$: \begin{equation}} \def\bn{\beq\label{int-act} I(\alpha) |\theta_N,\ldots,\theta_1\rangle_{\varepsilon_N,\ldots,\varepsilon_1}= \left\{ \begin{array}{c} \sum_{s\geq0}(-1)^s e^{-(2s+1)\alpha}I_{2s+1} |\theta_N,\ldots,\theta_1\rangle_{\varepsilon_N,\ldots,\varepsilon_1},\quad \alpha\to+\infty \\ \sum_{s\geq0}(-1)^s e^{(2s+1)\alpha}\overline I_{2s+1} |\theta_N,\ldots,\theta_1\rangle_{\varepsilon_N,\ldots,\varepsilon_1},\quad \alpha\to-\infty \end{array} \right. \end{equation}} \def\ed{\eeq where \begin{eqnarray} I_{2s+1} |\theta_N,\ldots,\theta_1\rangle_{\varepsilon_N,\ldots,\varepsilon_1}&=& \sum_{j=1}^N e^{(2s+1)\theta_j} |\theta_N,\ldots,\theta_1\rangle_{\varepsilon_N,\ldots,\varepsilon_1},\nonumber} \def\gamma{\gamma\\ \overline I_{2s+1} |\theta_N,\ldots,\theta_1\rangle_{\varepsilon_N,\ldots,\varepsilon_1}&=& \sum_{j=1}^N e^{-(2s+1)\theta_j} |\theta_N,\ldots,\theta_1\rangle_{\varepsilon_N,\ldots,\varepsilon_1},\quad s\geq0\ .\label{int-val} \end{eqnarray} It follows from \r{vac-inv} that \begin{equation}} \def\bn{\beq\label{vacinv} I_{2s+1}|\mbox{vac}\rangle_{\rm ph}=\overline I_{2s+1}|\mbox{vac}\rangle_{\rm ph}=0\ . \end{equation}} \def\ed{\eeq It is clear that the form factors of the quantum integrals $I_{2s+1}$ and $\overline I_{2s+1}$ vanish identically, but using these quantities we can partially solve the problem of the reconstructing the map $O\to\widetilde O$ of the local operators into the operators acting in the Hilbert space of the angular quantization. Suppose we know this identification for some particular local operator $\CO\to \widetilde\CO$. Then we can immediately find this identification for arbitrary descendant of the operator $\CO$ with respect to all integrals of motion $I_{2s+1}$ and $\overline I_{2s+1}$: $\CO(\alpha)=[\CO,I(\alpha)]$ \cite{Lu1}. The answer is \begin{equation}} \def\bn{\beq\label{ans11} \CO(\alpha)\mapsto \widetilde\CO(\alpha)=\widetilde\CO\, \widetilde I(\alpha)- \widetilde I(\alpha+2\pi i)\,\widetilde \CO\ , \end{equation}} \def\ed{\eeq where \begin{equation}} \def\bn{\beq\label{til-int} \widetilde I(\alpha)=\frac{1}{2ig'(\xi)} \sum_{\varepsilon=\pm}Z'_\varepsilon(\alpha+i\pi)\partial_\alpha Z'_{-\varepsilon}(\alpha)= \sum_{s>0}\widetilde I_{\pm(2s+1)}e^{\mp(2s+1)\alpha}\quad \hbox{when}\quad \alpha\to\pm\infty\ . \end{equation}} \def\ed{\eeq The prove of this statement is based on the cyclic property of the trace \r{tracecon} and looks as follows: \begin{eqnarray} &_{\rm ph}\langle\mbox{vac}| \CO(\alpha) |\theta_1,\ldots,\theta_n\rangle_{\varepsilon_1,\ldots,\varepsilon_n} =\ \ _{\rm ph}\langle\mbox{vac}| [\CO, I(\alpha)] |\theta_1,\ldots,\theta_n\rangle_{\varepsilon_1,\ldots,\varepsilon_n} \nonumber} \def\gamma{\gamma\\ &\rav{vacinv} _{\rm ph}\langle\mbox{vac}| \CO I(\alpha) |\theta_1,\ldots,\theta_n\rangle_{\varepsilon_1,\ldots,\varepsilon_n} = {\rm Tr}\,_{\CH_R}\sk{e^{2\pi K}\widetilde\CO\widetilde I(\alpha) Z^*_{\varepsilon_1}(\theta_1)\ldots Z^*_{\varepsilon_n}(\theta_n)}-\nonumber} \def\gamma{\gamma\\ &-\frac{1}{2ig'(\xi)^2} \sum_{\varepsilon,\mu=\pm}(\varepsilon\mu)^{n} {\rm Tr}\,_{\CH_R}\sk{e^{\pi K}\widetilde \CO Z^{\prime*}_\varepsilon(\alpha)Z'_\mu(\alpha) Z^*_{\varepsilon_1}(\theta_1)\ldots Z^*_{\varepsilon_n}(\theta_n) e^{\pi K} Z'_{-\mu}(\alpha)\partial_\alpha Z^{\prime*}_{-\varepsilon}(\alpha)}.\n \end{eqnarray} The last line in the previous calculation can be transformed as follows: \begin{eqnarray} &\sum_{\varepsilon,\mu=\pm}(\varepsilon\mu)^{n} {\rm Tr}\,_{\CH_R}\sk{e^{\pi K}\widetilde \CO Z^{\prime*}_\varepsilon(\alpha)Z'_\mu(\alpha) Z^*_{\varepsilon_1}(\theta_1)\ldots Z^*_{\varepsilon_n}(\theta_n) e^{\pi K} Z'_{-\mu}(\alpha)\partial_\alpha Z^{\prime*}_{-\varepsilon}(\alpha)}=\nonumber} \def\gamma{\gamma\\ &\quad\rav{ZFI-II} \sum_{\varepsilon,\mu=\pm} {\rm Tr}\,_{\CH_R}\sk{e^{\pi K} \widetilde \CO Z^*_{\varepsilon_1}(\theta_1)\ldots Z^*_{\varepsilon_n}(\theta_n) Z^{\prime*}_\varepsilon(\alpha) e^{\pi K} Z^{\prime*}_{-\mu}(\alpha) Z'_{-\mu}(\alpha)\partial_\alpha Z^{\prime*}_{-\varepsilon}(\alpha)}=\nonumber} \def\gamma{\gamma\\ &\quad\rav{complet} g'(\xi)\sum_{\varepsilon=\pm} {\rm Tr}\,_{\CH_R}\sk{ Z^{\prime*}_\varepsilon(\alpha) e^{\pi K} \partial_\alpha Z^{\prime*}_{-\varepsilon}(\alpha) e^{\pi K} \widetilde \CO Z^*_{\varepsilon_1}(\theta_1)\ldots Z^*_{\varepsilon_n}(\theta_n) }=\nonumber} \def\gamma{\gamma\\ &\quad\rav{Phidual} g'(\xi){\rm Tr}\,_{\CH_R}\sk{ e^{2\pi K} \widetilde I(\alpha+2\pi i) \widetilde \CO Z^*_{\varepsilon_1}(\theta_1)\ldots Z^*_{\varepsilon_n}(\theta_n) } \end{eqnarray} so \r{ans11} is proved. The generating function \r{til-int} has also another meaning. Namely, it was proved in \cite{NPT} that after substitution to the trace \r{tracecon} the coefficients $\widetilde I_{1}$ and $\widetilde I_{-1}$ of the asymptotical expansion \r{til-int} one obtains the known form factors \cite{Sbook} of the stress energy tensor in $SU(2)$-invariant Thirring model (this model can be obtained from SG model in the limit $\xi\to+\infty$). This allows to conjecture that for the finite $\xi$ the corresponding coefficients of the quantity $\widetilde I(\alpha)$ will also generate the form factors of stress-energy tensor in the SG model. \subsection{Symmetries of the model} In this subsection we will prove the following three statements. ($\iota$) The adjoint action of the algebra ${\cal A}(\widehat{sl}_2)$ \r{adj} on the total Hilbert space is given by the level zero action of this algebra \begin{equation}} \def\bn{\beq\label{repres1} \overline \CR(u_1-u_2,\xi){\rm Ad}_{L_1(u_1;\xi)}{\rm Ad}_{L_2(u_2;\xi)}= {\rm Ad}_{L_2(u_2;\xi)}{\rm Ad}_{L_1(u_1;\xi)} \overline \CR(u_1-u_2,\xi)\ . \end{equation}} \def\ed{\eeq ($\iota\iota$) The subspace of the $n$-particle states carries the finite-dimensional representation of the algebra ${\cal A}(\widehat{sl}_2)$ given by the the formulas \begin{equation}} \def\bn{\beq\label{homoc} \sk{{\rm id}\otimes\iota\otimes{\rm id}\otimes\iota\otimes\ldots\iota^{n-1}}\Delta^{(n-1)}(x) |\theta_1,\ldots, \theta_1\rangle_{\varepsilon_1,\ldots,\varepsilon_n},\quad x=e(u),\ f(u),\ h(u)\ , \end{equation}} \def\ed{\eeq where $\Delta^{(n)}(x)$ is $n$th power of the comultiplication maps \r{com-e-fu}--\r{comul-h} defined inductively $$\Delta^{(1)}\equiv\Delta,\quad \Delta^{(n)}(x)=\sk{\Delta\otimes{\rm id}}\Delta^{(n-1)}\ , $$ where the action of the Gauss coordinates $e(u)$, $f(u)$ and $h(u)$ on the one-particle states is defined by the formulas \r{e-res1} and \r{e-res2}. ($\iota\iota\iota$) The commutation relations of the algebra ${\cal A}(\widehat{sl}_2)$ in the form \r{ef}--\r{hh} allow to define certain asymptotical operators ${\bf e}_i$, ${\bf f}_i$ and ${\bf h}$ such that their commutation and comultiplication relations correspond to those of the Chevalley generators of the quantum affine algebra $U_q(\widehat{sl}_2)$ at level zero with the parameter of deformation $q=\exp\sk{\pi i\frac{\xi+1}{\xi}}$. The first statement is a simple consequence of the commutation relation \r{scr-com} and the fact of commutativity $[\CR(u,\xi),\sigma_z\otimes \sigma_z]=0$. The second statement is a direct consequence of the defining relations \r{51a}, \r{52a} and \r{53c} for the operators $Z^*_\pm(\theta)$. We start from one particle states $|\theta+\pi i/2\rangle_\pm$ and prove that they realize the spin $1/2$ representation \r{eval-0}--\r{eval-h1}. {}From the definition of the adjoint action \r{adj} we have \begin{eqnarray} {\rm Ad}_{k_1(u)^{-1}}\cdot Z^*_\pm(\widetilde\theta)&=& k_1(\widetilde u)^{-1}Z^*_\pm(\widetilde\theta)k_1(\widetilde u)+ k_1(\widetilde u)^{-1}\{Z^*_\pm(\widetilde\theta),f(\widetilde u)\}k_2(\widetilde u) e(\widetilde u)\label{ad11}\\ {\rm Ad}_{k_2(u)^{-1}}\cdot Z^*_\pm(\widetilde\theta)&+& {\rm Ad}_{e(u)k_1(u)^{-1}f(u)}\cdot Z^*_\pm(\widetilde\theta)=\nonumber} \def\gamma{\gamma\\ &=&k_2(\widetilde u)^{-1}Z^*_\pm(\widetilde\theta)k_2(\widetilde u)+ e(\widetilde u)k_1(\widetilde u)^{-1}\{Z^*_\pm(\widetilde\theta),f(\widetilde u)\} k_2(\widetilde u)\label{ad22}\\ -{\rm Ad}_{k_1(u)^{-1}f(u)}\cdot Z^*_\pm(\widetilde\theta)&=& k_1(\widetilde u)^{-1}\{Z^*_\pm(\widetilde\theta),f(\widetilde u)\} k_2(\widetilde u)\label{ad12}\\ -{\rm Ad}_{e(u)k_1(u)^{-1}}\cdot Z^*_\pm(\widetilde\theta)&=& e(\widetilde u)k_1(\widetilde u)^{-1}Z^*_\pm(\widetilde\theta)k_1(\widetilde u)+ k_2(\widetilde u)^{-1}Z^*_\pm(\widetilde\theta)k_2(\widetilde u)e(\widetilde u)+\nonumber} \def\gamma{\gamma\\ &+&e(\widetilde u) k_1(\widetilde u)^{-1}\{Z^*_\pm(\widetilde\theta),f(\widetilde u)\} k_2(\widetilde u)e(\widetilde u)\label{ad21} \end{eqnarray} where we denote $\widetilde u=u+\pi i/4$ and $\widetilde \theta=\theta+\pi i/2$. The calculation of the adjoint action of the Gauss coordinates of $L$-operator onto the state $|\widetilde \theta\rangle_-$ is an easy part. Indeed, using formulas \r{53c} we observe first that the anticommutator $\{Z^*_-(\widetilde\theta),f(\widetilde u)\}$ vanishes in \r{ad11}--\r{ad22} and using then \r{51a}, \r{52a} we obtain \begin{equation}} \def\bn{\beq\label{e-res1} {\rm Ad}_{f(u)}|\widetilde \theta\rangle_-=0, \quad {\rm Ad}_{e(u)}|\widetilde \theta\rangle_-=-\frac{{\rm sh}\,i\pi/\xi} {{\rm sh}\, \sk{ \frac{u-\theta}{\xi} } } |\widetilde \theta\rangle_+,\quad {\rm Ad}_{h(u)}|\widetilde \theta\rangle_-=\frac{{\rm sh}\,\sk{\frac{u-\theta+i\pi}{\xi} }} {{\rm sh}\,\sk{\frac{u-\theta}{\xi}}}|\widetilde \theta\rangle_- \end{equation}} \def\ed{\eeq which obviously coincide with the analogous formulas from \r{eval-0}--\r{eval-h1}. Let us demonstrate how the second formula in \r{e-res1} is obtained. Combining \r{ad11} and \r{ad21} and taking into account \r{52c} we obtain \begin{eqnarray} {\rm Ad}_{e(u)}\cdot Z_-^*(\widetilde\theta)&=& -e(\widetilde u) Z_-^*(\widetilde\theta) - h(\widetilde u)Z_-^*(\widetilde\theta) h(\widetilde u)^{-1}e(\widetilde u)=\nonumber} \def\gamma{\gamma\\ &\rav{51a}&-e(\widetilde u) Z_-^*(\widetilde\theta) - {{\rm sh}\frac{u-\theta+i\pi}{\xi}\over{\rm sh}\frac{u-\theta}{\xi}}Z_-^*(\widetilde\theta) e(\widetilde u)=\nonumber} \def\gamma{\gamma\\ &\rav{52a}& - {{\rm sh}\frac{i\pi}{\xi}\over{\rm sh}\frac{u-\theta}{\xi}}Z_+^*(\widetilde\theta)\ .\nonumber} \def\gamma{\gamma \end{eqnarray} The calculation of the adjoint action of the Gauss coordinates onto the state $|\widetilde \theta\rangle_+$ is more complicated but straightforward. The main trick is to use the formula \r{52a} to replace the operator $Z^*_+(\widetilde\theta)$ by the combination of the products $e(v)Z^*_-(\widetilde\theta)$ and $Z^*_-(\widetilde\theta)e(v)$. Using then the commutation relations of the algebra ${\cal A}(\widehat{sl}_2)$ in terms of the Gauss coordinates \r{ef}--\r{hh} we will find that the dependence on the spectral parameter $v$ is canceled out and we obtain \begin{equation}} \def\bn{\beq\label{e-res2} {\rm Ad}_{e(u)}|\widetilde \theta\rangle_+=0, \quad {\rm Ad}_{f(u)}|\widetilde \theta\rangle_+=-\frac{{\rm sh}\,i\pi/\xi} {{\rm sh}\,\sk{\frac{u-\theta}{\xi}}}|\widetilde \theta\rangle_-,\quad {\rm Ad}_{h(u)}|\widetilde \theta\rangle_+=\frac{{\rm sh}\,\sk{\frac{u-\theta-i\pi}{\xi}}} {{\rm sh}\,\sk{\frac{u-\theta}{\xi}}}|\widetilde \theta\rangle_+ \end{equation}} \def\ed{\eeq which coincide with the rest of the formulas \r{eval-0}--\r{eval-h1}. To find the action of the Gauss coordinates $e(u)$, $f(u)$ and $h(u)$ on the $n$-particle states we use the same formulas \r{ad11}--\r{ad21} with $Z^*_\pm(\widetilde\theta)$ replaced by the $n$-fold product of these operators. For example, the adjoint action of the Gauss coordinate $e(u)$ on the two-particle state is given by the formula \begin{eqnarray} {\rm Ad}_{e(u)}\cdot|\theta_1,\theta_2\rangle_{-,-}&=& -{{\rm sh}\sk{\frac{i\pi}{\xi}}\over{\rm sh}\sk{\frac{u-\theta_1}{\xi}}} |\theta_1,\theta_2\rangle_{+,-} + {{\rm sh}\sk{\frac{u-\theta_1+i\pi}{\xi}}\over{\rm sh}\sk{\frac{u-\theta_1}{\xi}}} {{\rm sh}\sk{\frac{i\pi}{\xi}}\over{\rm sh}\sk{\frac{u-\theta_2}{\xi}}} |\theta_1,\theta_2\rangle_{-,+}=\nonumber} \def\gamma{\gamma\\ &=&\widehat\Delta(e(u))\ |\theta_1,\theta_2\rangle_{-,-} \ , \label{examp1} \end{eqnarray} where we denoted by $\widehat\Delta=\sk{{\rm id}\otimes\iota}\Delta$ the composition of the comultiplication of the algebra ${\cal A}(\widehat{sl}_2)$ \r{com-e-fu}--\r{comul-h} and the involution \r{L-oper-rel}. Repeating these arguments inductively we prove the formula \r{homoc} where the action of the Gauss coordinates $e(u)$, $f(u)$ and $h(u)$ on the one-particle states are given by the formulas \r{e-res1} and \r{e-res2}. The commutation relations of the algebra ${\cal A}(\widehat{sl}_2)$ \r{ef}--\r{hh} at the zero central element demonstrate that the Gauss coordinates of $L$-operators have following asymptotics when ${\rm Re}\,\,u\to\pm\infty$: \begin{equation}} \def\bn{\beq e(u)\stackreb{\sim}{} \exp\sk{-\frac{|u|}{\xi}},\quad f(u)\stackreb{\sim}{} \exp\sk{-\frac{|u|}{\xi}},\quad h(u)\stackreb{\sim}{} h(\pm\infty)\equiv h_\pm\ . \label{asympHH} \end{equation}} \def\ed{\eeq It follows from \r{he} and \r{hf} that Cartan asymptotical generators $h_\pm$ have the following commutation relations with Gauss coordinates $e(u)$ and $f(u)$: \begin{equation}} \def\bn{\beq h_\pm e(u) h_\pm^{-1}=\exp\sk{\pm\frac{2\pi i}{\xi}}e(u),\quad h_\pm f(u) h_\pm^{-1}=\exp\sk{\mp\frac{2\pi i}{\xi}}f(u) \ . \label{q-comm}\end{equation}} \def\ed{\eeq The comultiplication rule \r{comul-h} yields that the asympotical Cartan elements are primitive and group-like: $\Delta h_\pm=h_\pm\otimes h_\pm$. The commutation relations \r{q-comm} yields that the product $h_+ h_-$ is central and also group-like primitive. Due to this we can put this central element to be equal to one so the asymptotical Cartan operators are inverse to each other: $h_+=h_-^{-1}$. Let us define the logarithmic Cartan operator ${\bf h}$ as follows: \begin{equation}} \def\bn{\beq\label{as-Car} h_\pm=\exp\sk{\pm i\pi \frac{\xi+1}{\xi}\ {\bf h}} \end{equation}} \def\ed{\eeq where the operator ${\bf h}$ has standard commutation relation with the Gauss coordinates $[{\bf h},e(u)]=2e(u)$ and $[{\bf h},f(u)]=-2f(u)$. Define also the asymptotical operators \begin{eqnarray} {\bf e}_\pm&=&\frac{1}{2}\ {\rm sh}\sk{i\pi\frac{\xi+1}{\xi}}^{-1} \lim{{\rm Re}\,\,u\to\pm\infty}e^{\pm u/\xi}e(u)\ ,\nonumber} \def\gamma{\gamma\\ {\bf f}_\pm&=&\frac{1}{2}\ {\rm sh}\sk{i\pi\frac{\xi+1}{\xi}}^{-1} \lim{{\rm Re}\,\,u\to\pm\infty}e^{\pm u/\xi}f(u)\ . \label{asy-gen} \end{eqnarray} {}From the commutation relations \r{ef}--\r{hf} we can obtain the commutation relations of these operators: \begin{eqnarray} [{\bf h},{\bf e}_\pm]&=&\pm2{\bf e}_\pm,\qquad [{\bf h},{\bf f}_\pm]=\mp2{\bf f}_\pm,\qquad [{\bf e}_\pm,{\bf f}_\mp]=\pm \frac{\sin\sk{\pi {\bf h}(\xi+1)/\xi}}{\sin\sk{\pi(\xi+1)/\xi}}= \pm\frac{q^{\bf h}-q^{-{\bf h}}}{q-q^{-1}}\ ,\nonumber} \def\gamma{\gamma\\ q{\bf e}_+{\bf e}_-&=&{q^{-1}}{\bf e}_-{\bf e}_+,\quad q^{-1}{\bf f}_+{\bf f}_-\ =\ {q}{\bf f}_-{\bf f}_+\ ,\nonumber} \def\gamma{\gamma\\ q^{\mp3}{\bf e}^3_\pm{\bf f}_\pm&-&\frac{q^3-q^{-3}}{q-q^{-1}} \sk{q^{\mp1}{\bf e}^2_\pm{\bf f}_\pm{\bf e}_\pm -q^{\pm1}{\bf e}_\pm{\bf f}_\pm{\bf e}^2_\pm}-q^{\pm3}{\bf f}_\pm{\bf e}^3_\pm=0\ ,\nonumber} \def\gamma{\gamma\\ q^{\pm3}{\bf f}^3_\pm{\bf e}_\pm&-&\frac{q^3-q^{-3}}{q-q^{-1}} \sk{q^{\pm1}{\bf f}^2_\pm{\bf e}_\pm{\bf f}_\pm -q^{\mp1}{\bf f}_\pm{\bf e}_\pm{\bf f}^2_\pm}-q^{\mp3}{\bf e}_\pm{\bf f}^3_\pm=0\ , \label{soot1} \end{eqnarray} where $q=\exp\sk{i\pi\frac{\xi+1}{\xi}}$. These commutation relations allows to identify the asymptotical operators with the Chevalley generators of the affine quantum algebra $U_q(\widehat{sl}_2)$ at level zero. Using formulas \r{e-res1}, \r{e-res2} and the rule of the Gauss coordinates actions onto multi-particle states \r{homoc} we can obtain the action of the asymptotical operators ${\bf e}_\pm$, ${\bf f}_\pm$ and ${\bf h}$ onto multiparticle states and prove that it is given by the comultiplication of the algebra $U_q(\widehat{sl}_2)$. To do this we first slightly modify the action of these generators following \cite{RS} when they act on the one-particle states $|\theta\rangle_\pm$: \begin{equation}\label{as-ge-mo} {\bf e}_\pm\mapsto \exp\sk{\mp\frac{\theta}{\xi}} {\bf e}_\pm,\quad {\bf f}_\pm\mapsto \exp\sk{\pm\frac{\theta}{\xi}} {\bf f}_\pm,\quad {\bf h} \mapsto {\bf h}. \end{equation} By the straightforward calculation using the definition of the adjoint action on the multiple-particle states \r{adj} and the formulas \r{51a}--\r{53c} we obtain that this action can be formulated through the comultiplication \begin{equation}\label{co-as} \Delta_0 {\bf e}_\pm= {\bf e}_\pm\otimes 1+q^{\mp{\bf h}}\otimes {\bf e}_\pm,\quad \Delta_0 {\bf f}_\pm= 1\otimes {\bf f}_\pm + {\bf f}_\pm\otimes q^{\pm{\bf h}},\quad \Delta_0 {\bf h}={\bf h}\otimes 1+1\otimes {\bf h}\ , \end{equation} which can be formally obtained from the comultiplication formulas for the Gauss coordinates \r{com-e-fu}--\r{comul-h} using \r{asympHH}. The action of the asymptotical operators on the one particle states are defined as follows \begin{equation}} \def\bn{\beq\label{act-asy} {\bf e}_\pm|\theta\rangle_+={\bf f}_\pm|\theta\rangle_-=0,\quad {\bf e}_\pm|\theta\rangle_-=|\theta\rangle_+,\quad {\bf f}_\pm|\theta\rangle_+=|\theta\rangle_-,\quad {\bf h}|\theta\rangle_\pm=\pm|\theta\rangle_\pm\ . \end{equation}} \def\ed{\eeq For example, let us demonstrate the origin of this comultiplication on the two-particle state. From \r{examp1} we have \begin{eqnarray} {\rm Ad}_{{\bf e}_+}\cdot|\theta_1,\theta_2\rangle_{-,-}\!\!\!\!&=&\!\!\!\!\!\!\!\! \lim{{\rm Re}\, u\to+\infty}\sk{{\exp\sk{\frac{u-\theta_1}{\xi}}\over 2{\rm sh}\sk{\frac{u-\theta_1}{\xi}}}|\theta_1,\theta_2\rangle_{+,-} - {{\rm sh}\sk{\frac{u-\theta_1+i\pi}{\xi}}\over{\rm sh}\sk{\frac{u-\theta_1}{\xi}}} {\exp\sk{\frac{u-\theta_2}{\xi}}\over2{\rm sh}\sk{\frac{u-\theta_2}{\xi}}} |\theta_1,\theta_2\rangle_{-,+}}=\nonumber} \def\gamma{\gamma\\ &=&|\theta_1,\theta_2\rangle_{+,-}-\exp\sk{\frac{i\pi}{\xi}} |\theta_1,\theta_2\rangle_{-,+}=\nonumber} \def\gamma{\gamma\\ &\rav{act-asy}&\sk{{\bf e}_+\ot1+q^{-{\bf h}}\otimes{\bf e}_+}|\theta_1,\theta_2\rangle_{-,-} =\Delta_0({\bf e}_+)\ |\theta_1,\theta_2\rangle_{-,-}\ .\nonumber} \def\gamma{\gamma \end{eqnarray} We would like to note here that the set of the asymptotical generators ${\bf e}_+$, ${\bf f}_-$ and ${\bf h}$ or ${\bf e}_-$, ${\bf f}_+$ and ${\bf h}$ cannot be identified with the set $e$, $f$ and $h$ used in the construction of the evaluation homomorphism from the algebra ${\cal A}(\widehat{sl}_2)$ onto $U_{i\pi/\xi}(sl_2)$, because the first ones are the subalgebras while the second one is factor subalgebra. In particular, the action \r{act-asy} cannot be obtained from the adjoint action onto one-particle states \r{e-res1} and \r{e-res2}. The consideration presented above prove that the adjoint action of the finite-dimensional subalgebra of ${\cal A}(\widehat{sl}_2)$ onto the total Hilbert space of the SG model describe the symmetries of this space investigated in \cite{RS,AL,BL}. \subsection{Symmetries of the model at the FF point} Now we would like to demonstrate how the quantum symmetries of the Hilbert space $\CH$ of the SG model become the classical ones (i.e. correspond to undeformed current algebra) at the FF point. It is clear that the finite-dimensional representations of the algebra ${\cal A}(\widehat{sl}_2)$ at the value $\xi=1$ degenerate. Moreover, the operator $h_+$ becomes the central element of the algebra (cf. \r{q-comm}) and takes the value $(-1)^{k+1}$, $k=0,1$ on the subspace $\CH_k$ of the even and odd number of particles of the total Hilbert space $\CH$. In order to obtain the nontrivial action of the algebra ${\cal A}(\widehat{sl}_2)$ at the FF point on the Hilbert space of states we introduce the rescaled operators \begin{equation}} \def\bn{\beq\label{res-op} \widehat e(u)=\left.-\frac{e(u)}{{\rm sh}(i\pi/\xi)}\right|_{\xi=1},\quad \widehat f(u)=\left.-h_+\frac{f(u)}{{\rm sh}(i\pi/\xi)}\right|_{\xi=1},\quad \widehat h(u)=\left.-\frac{h(u)h_+^{-1}-1}{{\rm sh}(i\pi/\xi)}\right|_{\xi=1}. \end{equation}} \def\ed{\eeq The nontrivial commutation relations of the algebra ${\cal A}(\widehat{sl}_2)$ reads as follows: \begin{eqnarray} {[}\widehat h(u), \widehat e(v){]}&=& 2{\rm cth}\,(u-v) \widehat e(v)- 2{\widehat e(u)\over {\rm sh}\,(u-v)} \ ,\label{the}\\ {[}\widehat h(u), \widehat f(v){]}&=& -2{\rm cth}\,(u-v) \widehat f(v)+ 2{\widehat f(u)\over {\rm sh}\,(u-v)} \ ,\label{thf}\\ {[}\widehat e(u), \widehat f(v){]}&=& {\widehat h(u)-\widehat h(v) \over {\rm sh}\,(u-v) }\ .\label{thh} \end{eqnarray} The algebra \r{the}--\r{thh} coincides with the classical current algebra $\widehat{sl}_2$ on the line \cite{KLPST}. Formulas \r{e-res1} and \r{e-res2} of the adjoint action of the operators \r{res-op} becomes \begin{eqnarray} {\rm Ad}_{\widehat f(u)}|\widetilde \theta\rangle_-&=& {\rm Ad}_{\widehat e(u)}|\widetilde \theta\rangle_+=0, \quad {\rm Ad}_{\widehat h(u)}|\widetilde \theta\rangle_\pm=\pm{\rm cth}(u-\theta) \theta\rangle_\pm\ ,\nonumber} \def\gamma{\gamma\\ {\rm Ad}_{\widehat e(u)}|\widetilde \theta\rangle_-&=&\frac{1} {{\rm sh}\, \sk{u-\theta} } |\widetilde \theta\rangle_+,\quad {\rm Ad}_{\widehat f(u)}|\widetilde \theta\rangle_+=-\frac{1} {{\rm sh}\,\sk{u-\theta}}|\widetilde \theta\rangle_-\ ,\label{e-res0} \end{eqnarray} and on the multi-particle states are \begin{equation}} \def\bn{\beq \widehat\Delta(x)=x\ot1+1\otimes x,\quad x=\widehat e(u),\ \widehat f(u),\ \widehat h(u)\ , \label{com-clas} \end{equation}} \def\ed{\eeq where in order to obtain \r{e-res0} and \r{com-clas} we used the fact that operator $h_+$ equal to $-1$ on the one-particle state. The phenomena that quantum symmetries of the Hilbert space of state for the SG model becomes the classical ones at the FF point is a consequence of the fact that $S$-matrix in this limit yields the classical $r$-matrix \r{r-clas}: \begin{equation}} \def\bn{\beq\label{r-clas2} r(u)=\lim{\xi\to1}\frac{S(u;\xi)+1}{\pi i(1-\xi)}\ . \end{equation}} \def\ed{\eeq This phenomena was observed in reflectionless SG theory \cite{LN} and was used to investigate the space of the local operators in SG model at FF point \cite{Le}. \setcounter{equation}{0} \section{Discussion} In this paper we further developed the method of angular quantization for the Sine-Gordon model. Technically the application of this method splits into two parts. First, one should explicitly describe canonical quantization $\CH_R$ of the model in right Rindler wedge, where the boost plays the role of hamiltonian. Then the space of states and local operators of the theory on the line are described in terms of certain operators acting in $\CH_R$. We studied the SG theory at the free fermion point where the canonical quantization in RRW can be done explicitly. We investigated the integrals of motion and found that the usual local integrals of motion diverge. This forced us to consider nonlocal integrals of motion which are a certain analytical continuation (in the space of eigenvalues of Lorentz boost) of the usual charges and the only possibility to close them into a quadratic current algebra is to use charges with different monodromy properties. They form the specialization of the scaling elliptic algebra ${\cal A}(\widehat{sl}_2)$ \cite{KLP} into free fermion point. The bosonization \cite{Lu1} naturally appears in terms of scattering data. This indicates that angular quantization of SG model can be done in terms of the representation theory of the algebra ${\cal A}(\widehat{sl}_2)$. Starting from level one representations of this algebra in the bosonic Fock space we managed to construct the space of asymptotical states of SG model and some local operators acting into this space of states, in particular, the transfer matrix and the commuting set of the integrals of motion, and demonstrate the mechanism of trace calculations of the form factors of local operators. This approach is an extension of the ideas presented in \cite{JM} for XXZ model. The algebra ${\cal A}(\widehat{sl}_2)$ is not a Hopf algebra, but we were able to define the adjoint action of this algebra on the space of states, such that $n$-particle states with given rapidities form $n$-fold tensor product of two-dimensional representations of the algebra ${\cal A}(\widehat{sl}_2)$. Contrary to the integrable models on the lattice local integrals and local operators of the SG theory appear as coefficients of the asymptotical expansions of certain currents which are constructed explicitly. In particular, the asymptotical expansion of the level zero adjoint action of the algebra ${\cal A}(\widehat{sl}_2)$ on the space of states produce the action of Chevelley generators of quantum affine algebra, which was known before. At the free fermion point we get in this way the action of the classical affine algebra $\widehat{sl}_2$ which was constructed in the framework of the radial quantization in the paper \cite{Le}. Nevertheless, the understanding of the angular quantization method of SG model for generic value of the renormalized coupling constant $\xi$ is far from being complete. In particular, there is no rigorous construction of the quantum analogs of the Jost functions introduced in \cite{Lu1,Lu2} without referring to bosonization. SG model admits also natural analog of `new level zero action' (see \cite{JMKKP} and references therein) which is given in terms of $L$-operators as follows \begin{equation}} \def\bn{\beq\label{adj-d} {\rm Ad}'_{L(u)}\cdot X= L(u)\,X\,L(u)^{-1}\ \end{equation}} \def\ed{\eeq and depends on the dual deformation parameter $q'=\exp\sk{i\pi\frac{\xi}{\xi+1}}$. It will be interesting to extend the results on the spinon bases in conformal field theories investigated in \cite{Spinon} to the massive integrable models. As we mentioned already that the algebra ${\cal A}(\widehat{sl}_2)$ is quasi-Hopf algebra, but belonging to a family of dynamical elliptic algebra. The definition of adjoint action, used in this paper, did not refer to the axiomatics of this family. It would be interesting to fill this gap. Finally, it would also be interesting to further explore the role of the duality transformation \r{dual-tr}. In terms of the SG coupling $\beta$, this is an electric/magnetic duality $\beta \to 2/\beta$ familiar from the conformal field theory of a compactified free boson. As described in the paper this duality relates the $q'$-deformation parameter of the algebra of the monodromy matrix with the $q$-deformation parameter of the physical S-matrix. Though we did not present this here, one can define a dual monodromy matrix by the formal replacement $\beta \to 2/\beta$ in the usual monondromy matrix, and show that formally this dual monodromy matrix commutes with the original monodromy matrix. This would imply that the dual monodromy matrix generates additional integrals of motion, presumably related to the quantum affine symmetry described in \cite{BL}. \section{Acknowledgment} This investigation was basically done during the visits of two of the authors (S.Kh. and S.P.) to Cornell University organized in the framework of the CRDF Award No. RM2-150. We would like to acknowledge hospitality at the Newmann Laboratory of Cornell University and to thank S.~Lukyanov and A.~Zamolodchikov for the hospitality and useful discussions during S.Kh. and S.P. visit to Rutgers University in November 1998. We also would like to acknowledge the discussions with our colleagues S.~Kharchev, D.~Lebedev and V.~Shadura. The research described in this publication was made possible in part by grants RFBR-97-01-01041 (S.~Pakuliak), INTAS 93-10183 (S.~Khoroshkin), INTAS--OPEN--97--1312, by Award No. RM2-150 of the U.S. Civilian Research \& Development Foundation (CRDF) for the Independent States of the Former Soviet Union, and by the grants for promotion of French-Russian scientific cooperation: the CNRS grant PICS No. 608 and the RFBR grant No. 98-01-2033 (S.~Khoroshkin and S.~Pakuliak). \app{The algebra of bosons $a_\lambda$ and $\widetilde a_\lambda$} \label{app1} The normal ordering with respect to the vacuum vectors \r{Fock} and \r{lFock} \begin{equation}} \def\bn{\beq \label{j7} b_\pm(\nu)b_\mp(\nu')={:}b_\pm(\nu)b_\mp(\nu'){:} + \langle b_\pm(\nu)b_\mp(\nu')\rangle,\qquad \langle b_\pm(\nu)b_\mp(\nu')\rangle = \delta(\nu+\nu')\Theta(-\nu'), \end{equation}} \def\ed{\eeq where $\Theta(\nu)$ is a `continuous' step function \begin{equation}} \def\bn{\beq\label{step} \Theta(\nu)=\left\{ \begin{array}{c} 1,\quad\hbox{for}\quad \nu>0\\ 1/2,\quad\hbox{for}\quad \nu=0\\ 0,\quad\hbox{for}\quad \nu<0 \end{array}\right.,\qquad \Theta(\nu)+\Theta(-\nu)\equiv1 \end{equation}} \def\ed{\eeq allows to observe that the commutation relations between operators $a_\lambda$ and $\widetilde a_\lambda$ is not closed in a sense that the commutator $[a_\lambda,\widetilde a_\mu]$ cannot be presented as a linear combinations of the same operators with $\Bbb{C}$-number coefficients and $\Bbb{C}$-valued functions. The idea is to consider this commutator as a new bosonic operator and try to close the extended by this operator algebra. Fortunately this extended algebra is closed. To describe its commutation relations we introduce new operators: \begin{equation}} \def\bn{\beq\label{j14a} t_{\lambda,\mu}= \int_{-\infty}^\infty d\nu\ \frac{\Gamma\sk{\frac{1}{2}-i(\lambda+\nu)}} {\Gamma\sk{\frac{1}{2}+i(\mu-\nu)}}\ {\rm th}\,\pi\nu\ {:}b_-(\mu-\nu)b_+(\lambda+\nu){:}\ , \end{equation}} \def\ed{\eeq \begin{equation}} \def\bn{\beq\label{j14} \widetilde t_{\lambda,\mu}= \int_{-\infty}^\infty d\nu\ \frac{\Gamma\sk{\frac{1}{2}-i(\mu-\nu)}} {\Gamma\sk{\frac{1}{2}+i(\lambda+\nu)}}\ {\rm th}\,\pi\nu\ {:}b_-(\mu-\nu)b_+(\lambda+\nu){:}\ . \end{equation}} \def\ed{\eeq The bosonic operators $a_\lambda$ and $\widetilde a_\lambda$ are related to the new operators $t_{\lambda,\mu}$ and $\widetilde t_{\lambda,\mu}$ by the linear transformation: \begin{eqnarray} \widetilde a_\lambda&=&{\rm cth}\,\pi\lambda\ \widetilde t_{\lambda,0} - \frac{1}{{\rm sh}\,\pi\lambda}\ t_{0,\lambda},\nonumber} \def\gamma{\gamma\\ a_\lambda&=& \frac{1}{{\rm sh}\,\pi\lambda}\ \widetilde t_{\lambda,0} -{\rm cth}\,\pi\lambda\ t _{0,\lambda}. \label{j17} \end{eqnarray} Using simple trigonometric algebra we conclude that the set of the operators $t_{\lambda,\mu}$ and $\widetilde t_{\lambda,\mu}$ is not independent. For example, the following relation is valid: \begin{equation}} \def\bn{\beq\label{j19} {\rm sh}\,\pi\lambda\ (t_{\lambda,\mu}- t_{\lambda+\mu,0} = {\rm sh}\,\pi\mu\ (\widetilde t_{0,\lambda+\mu}-\widetilde t_{\lambda,\mu}) \end{equation}} \def\ed{\eeq therefore we can conclude that complete algebra of the bosonic operators reads as follows: \begin{eqnarray} [a_\lambda,a_\mu{]}&=&\lambda\delta(\lambda+\mu)\label{j21}\\ {[}a_\lambda,t_{\mu,\rho}{]}&=& t_{\mu,\rho+\lambda}-t_{\mu+\lambda,\rho}+\delta(\lambda+\mu+\rho) \int_{0}^\lambda d\nu\ {\rm th}\,\pi(\nu+\rho)\label{j15}\\ {[}t_{\lambda,\mu},t_{\lambda',\mu'}{]} &=& {\rm cth}\,\pi(\lambda+\mu')(t_{\lambda+\lambda'+\mu',\mu}-t_{\lambda',\lambda+\mu'+\mu})+\nonumber} \def\gamma{\gamma\\ &+& {\rm cth}\,\pi(\lambda'+\mu)(t_{\lambda,\lambda'+\mu'+\mu}-t_{\lambda'+\lambda+\mu,\mu'})+ \nonumber} \def\gamma{\gamma\\ &+&\delta(\lambda+\lambda'+\mu+\mu')\int_0^{\lambda+\mu} d\nu \ {\rm th}\,\pi(\nu-\lambda)\ {\rm th}\,\pi(\nu+\mu'). \label{j16} \end{eqnarray} This algebra can be understood as an algebraic realization of the complicated integral transform which relate the operators $\widetilde a_\lambda$ and $a_\lambda$. Indeed, using these commutation relations we can verify that the combination \begin{equation}} \def\bn{\beq\label{j20} \widetilde a_\lambda = {\rm ch}\,\pi\lambda\ a_\lambda - {\rm sh}\,\pi\lambda\ t_{0,\lambda} \end{equation}} \def\ed{\eeq also has the commutation relations of the Heisenberg algebra as $a_\lambda$ do (cf. \r{j21}). \app{Quantum Jost functions at the FF point} \label{app2} We will prove the equivalence of \r{rel+} to \r{rel1}. The second case can be treated analogously. First, we write down explicitly all the normal ordering rules which follows from \r{normal} \begin{eqnarray} Z'_-(z)F(u)&=&\frac{e^{\gamma/2}}{(2\pi)^{3/2}} { \Ga{\frac{1}{4}+\frac{u-z}{2\pi i}} \over \Ga{\frac{3}{4}+\frac{u-z}{2\pi i}} }\ {:}Z'_-(z)F(u){:}\nonumber} \def\gamma{\gamma\\ F(u)Z'_-(z)&=&\frac{e^{\gamma/2}}{(2\pi)^{3/2}} { \Ga{\frac{1}{4}-\frac{u-z}{2\pi i}} \over \Ga{\frac{3}{4}-\frac{u-z}{2\pi i}} }\ {:}Z'_-(z)F(u){:}\nonumber} \def\gamma{\gamma\\ F(u)F(v)&=& -\frac{e^{2\gamma}}{2\pi i}(u-v) {:}F(u)F(v){:}\nonumber} \def\gamma{\gamma\\ Z'_+\sk{\alpha_1-\frac{i\pi}{2}} Z'_+\sk{\alpha_2+\frac{i\pi}{2}}&=&g(\alpha_1-\alpha_2){:}Z'_+\sk{\alpha_1-\frac{i\pi}{2}} Z'_+\sk{\alpha_2+\frac{i\pi}{2}}{:}\nonumber} \def\gamma{\gamma \end{eqnarray} where in the last formula the function $g(\alpha)$ is given in terms of double $\Gamma$-functions $$ g(\alpha)=\frac{e^{5\gamma/4}}{2\pi} \frac{\Gamma_2^2(3\pi+i\alpha)}{\Gamma_2(2\pi+i\alpha)\Gamma_2(4\pi+i\alpha)}. $$ For the double and usual $\Gamma$-functions we use the integral representations \cite{Bar,JM1} \begin{eqnarray} \int_{\widetilde C}{d\lambda\,\ln(-\lambda)\over2\pi i\lambda}\ {e^{-x\lambda}\over1-e^{-\lambda/\eta}}&=&\ln\Gamma(\eta x)+ \left(\eta x-\fract{1}{2}\right)(\gamma-\ln\eta)-\fract{1}{2}\ln2\pi\ , \n \\ \int_{\widetilde C}{d\lambda\,\ln(-\lambda)\over2\pi i\lambda}\ {e^{-x\lambda}\over(1-e^{-\lambda\omega_1}) (1-e^{-\lambda\omega_2})}&=&\ln\Gamma_2(x\mid\omega_1,\omega_2)-{\gamma\over2} B_{2,2}(x\mid\omega_1;\omega_2)\ , \n \end{eqnarray} where $B_{2,2}(x\mid\omega_1;\omega_2)$ is the double Bernulli polynomial of the second order $$ B_{2,2}(x\mid\omega_1;\omega_2) ={1\over\omega_1\omega_2} \left[x^2-x(\omega_1+\omega_2)+{\omega^2_1+3\omega_1\omega_2+\omega^2_2 \over6}\right]. $$ The constant $g$ in the relation \r{rel1} is the value of the function $g(\alpha)$ at the point $\alpha=0$. We have \begin{eqnarray} &g^{-1} Z'_+\sk{\alpha-\frac{i\pi}{2}} Z'_+\sk{\alpha+\frac{i\pi}{2}} =\nonumber} \def\gamma{\gamma\\ &=\frac{i}{16\pi^6}\int_{C_1}\int_{C_2} du_1du_2\ {:} Z'_-\sk{\alpha-\frac{i\pi}{2}} Z'_-\sk{\alpha+\frac{i\pi}{2}} F(u_1)F(u_2){:}\times\nonumber} \def\gamma{\gamma\\ &\quad\times \Ga{\frac{1}{2}+\frac{u_1-\alpha}{2\pi i}} \Ga{-\frac{u_1-\alpha}{2\pi i}} \Ga{\frac{1}{2}-\frac{u_2-\alpha}{2\pi i}} \Ga{\frac{u_2-\alpha}{2\pi i}} \times\nonumber} \def\gamma{\gamma\\ &\quad \times (u_1-u_2)\ { \Ga{\frac{1}{2}+\frac{u_2-\alpha}{2\pi i}} \Ga{\frac{1}{2}-\frac{u_1-\alpha}{2\pi i}} \over \Ga{1+\frac{u_2-\alpha}{2\pi i}} \Ga{1-\frac{u_1-\alpha}{2\pi i}}}.\label{r4} \end{eqnarray} The contours $C_1$ and $C_2$ in \r{r4} go from $-\infty$ to $+\infty$ and \begin{equation}} \def\bn{\beq\label{co1} {\rm Im}\,\,\alpha-\pi < {\rm Im}\,\,u_1< {\rm Im}\,\,\alpha ,\quad {\rm Im}\,\,\alpha < {\rm Im}\,\,u_2< {\rm Im}\,\,\alpha+\pi \end{equation}} \def\ed{\eeq Using the elementary properties of the $\Gamma$-functions we can rewrite the integrand in \r{r4} in the form \begin{equation}} \def\bn{\beq\frac{i}{8\pi^2} \int_{C_1}\int_{C_2} du_1du_2 \left[{1\over u_2-\alpha} -{1\over u_1-\alpha}\right] {{:} Z'_-\sk{\alpha-\frac{i\pi}{2}} Z'_-\sk{\alpha+\frac{i\pi}{2}} F(u_1)F(u_2){:}\over {\rm ch}\sk{\frac{u_1-z}{2}} {\rm ch}\sk{\frac{u_2-z}{2}} }\label{integra}\end{equation}} \def\ed{\eeq where contours $C_1$ and $C_2$ are specified in \r{co1}. Using the fact that integrand in \r{integra} is antisymmetric function with respect to variables $u_1$ and $u_2$ we conclude: \begin{equation}} \def\bn{\beq\label{final} g^{-1} Z'_+\sk{\alpha-\frac{i\pi}{2}} Z'_+\sk{\alpha+\frac{i\pi}{2}}=\frac{1}{4\pi}\int_{-\infty}^\infty du \frac{F(u)} {{\rm ch}\sk{\frac{u-\alpha}{2}}}=\Lambda_-(\alpha) \end{equation}} \def\ed{\eeq since the current $F(u)$ coincide with the scattering data operator ${\cal Z}_-(u)$.
\section{Overview} The Einstein equations for the structure of spacetime were first published in 1916 when Einstein introduced his famous general theory of relativity. This theory of gravity has remained essentially unchanged since its discovery, and it provides the underpinnings of modern theories of astrophysics and cosmology. The theory is essential in describing phenomena such as black holes, compact objects, supernovae, and the formation of structure in the Universe. Unfortunately, the equations are a set of highly complex, coupled, nonlinear partial differential equations involving 10 functions of 4 independent variables. They are among the most complicated equations in mathematical physics. For this reason, in spite of more than 80 years of intense study, the solution space to the full set of equations is essentially unknown. Most of what we know about this fundamental theory of physics has been gleaned from linearized solutions, highly idealized solutions possessing a high degree of symmetry (e.g., static, or spherically or axially symmetric), or from perturbations of these solutions. Over the last 30 years a growing research area, called Numerical Relativity, has developed, where computers are employed to construct numerical solutions to these equations. Although much has been learned through this approach, progress has been slow due to the complexity of the equations and inadequate computational power. For example, an important astrophysical application is the 3D spiraling coalescence of two black holes (BH) or neutron stars (NS), which will generate strong sources of gravitational waves. As has been emphasized by Flanagan and Hughes, one of the best candidates for early detection by the laser interferometer network is increasingly considered to be BH mergers\cite{Flanagan97b,Flanagan97a}. The imminent arrival of data from the long awaited gravitational wave interferometers (see, e.g., Ref.~\cite{Flanagan97b} and references therein) has provided a sense of urgency in understanding these strong sources of gravitational waves. Such understanding can be obtained only through large scale computer simulations using the full machinery of numerical relativity. Furthermore, the gravitational wave signals are likely to be so weak by the time they reach the detectors that reliable detection may be difficult without prior knowledge of the merger waveform. These signals can be properly interpreted, or perhaps even detected, only with a detailed comparison between the observational data and a set of theoretically determined ``waveform templates''. In most cases, these waveform templates needed for gravitational wave data analysis have to be generated by large scale computer simulations, adding to the urgency of developing numerical relativity. However, a realistic 3D simulation based on the full Einstein equations is a highly non-trivial task---based on axisymmetric black hole calculations performed during late 1980's and algorithms available at the time---one can estimate the time required for a reasonably accurate 3D simulation of, say, the coalescence of a compact object binary, to be at least 100,000 Cray Y-MP hours! But there is good reason for optimism that such problems can be solved within the next decade. Scalable parallel computers, and efficient algorithms that exploit them, are quickly revolutionizing computational science, and numerical relativity is a great beneficiary of these developments. Over the last years the community has developed 3D codes designed to solve the complete set of Einstein equations that run very efficiently on large scale parallel computers. We will describe below one such code, called ``Cactus'', that has achieved 142 GFlops on a 1024 node Cray T3E-1200, which is more than 2000 times faster than 2D codes of a few years ago running on a Cray Y-MP (which also had only about 0.5\% the memory capacity of the large T3E). Such machines are expected to scale up rapidly as faster processors are connected together in even higher numbers, achieving Teraflop performance on real applications in a few years. Numerical relativity requires not only large computers and efficient codes, but also a wide variety of numerical algorithms for evolving and analyzing the solution. Because of this richness and complexity of the equations, and the interesting applications to problems such as black holes and neutron stars, natural collaborations have developed between applied mathematicians, physicists, astrophysicists, and computational scientists in the development of a single code to attack these problems. There are various large scale collaborative effort in recent years in this direction, including the NSF Black Hole Grand Challenge Project (recently concluded), the NASA Neutron Star Grand Challenge Project and the NCSA/Potsdam/Wash U numerical relativity collaboration. We introduce in this paper a code called ``Cactus'', which is developed by the NCSA/Potsdam/Wash U collaboration, and is employed in the NASA Neutron Star Grand Challenge Project. We will describe some of the algorithms and capabilities of this code in this paper. In the next sections we will first give a brief description of the numerical formulation of the theory of general relativity, and discuss particular difficulties associated with numerical relativity. The discussion will necessarily be brief. Examples are mostly drawn from work carried out by our NCSA/Potsdam/Wash U numerical relativity collaboration. We also provide URL addresses for web pages containing graphics and movies of some of our results. To conclude this brief introduction, a statement of where we stand in terms of simulating general relativistic compact objects is in order. The NSF black hole grand challenge project and related work achieved long term stable evolution of single black hole spacetimes under certain conditions~\cite{Daues96a,Cook97a,Gomez98a}, but there is still a long way to go before the spiraling coalescence can be computed. The presently on-going NASA neutron star grand challenge project recently succeeded in evolving grazing collision of two neutron stars using the full Einstein-relativistic hydrodynamic system of equations, with a simple equation of state. While the inspiral coalescences of two neutron stars is {\it not} a stated goal of the NASA project, we expect to be able to carry out preliminary studies of the inspiral coalescences in the next few years. The final goal of a full solution of the problem including radiation transport and magneto-hydrodynamics for comparison between numerical simulations and observations in gravitational wave astronomy (waveform templates) and high energy astronomy ($gamma$ ray burst) will take many more years, hopefully building on the effort described in this paper. The Nakamura group also reports preliminary success in evolving several orbits with a fully relativistic GR-hydro code~\cite{Nakamura99a}. \section{Einstein Equations for Relativity} The generality and complexity of the Einstein equations make them an excellent and fertile testing ground for a variety of broadly significant computing issues. They form a system of dozens of coupled, nonlinear equations, with thousands of terms, of mixed hyperbolic-elliptic type, and even undefined types, depending on coordinate conditions. This rich and general structure of the equations implies that the techniques developed to solve our problems will be immediately applicable to a large family of diverse scientific applications. The system of equations breaks up naturally into a set of constraint equations, which are elliptic in nature, evolution equations, which are ``hyperbolic'' in nature (more on this below), and gauge equations, which can be chosen arbitrarily (often leading to more elliptic equations). The evolution equations guarantee (mathematically) that the elliptic constraints are satisfied at all times provided the initial data satisfied them. This implies that the initial data are not freely specifiable. Moreover, although the constraints are satisfied mathematically during evolution, it will not be so numerically. These problems are each discussed in turn below. First, however, we point out that a much simpler theory, familiar to many, has all of these same features. Maxwell's equations describing electromagnetic radiation have: (a) elliptic constraint equations, demanding that in vacuum the divergence of the electric and magnetic fields vanish at all times; (b) evolution equations, determining the time development of these fields, given suitable initial data that satisfies the elliptic constrain equations; and (c) gauge conditions that can be applied freely to certain variables in the theory, such as some components of the vector potential. Some choices of vector potential lead to hyperbolic evolution equations for the system, and some do not. We will find all of these features present in the much more complicated Einstein equations, so it is useful to keep Maxwell's equations in mind when reading the next sections. In the standard 3+1 ADM approach to general relativity,\cite{York79}, the basic building block of the theory---the spacetime metric---is written in the form \begin{equation} d s^2 = -(\alpha^{2} -\beta ^{a}\beta _{a}) d t^2 + 2 \beta _{a} d x^{a} d t +\gamma_{ab} d x^{a} d x^{b}\; , \end{equation} using geometrized units such that the gravitational constant $G$ and the speed of light $c$ are both equal to unity. Throughout this paper, we use Latin indices to label spatial coordinates, running from 1 to 3. The ten functions $(\alpha, \beta ^{a}, \gamma_{ab})$ are functions of the spatial coordinates $x^{a}$ and time $t$. Indices are raised and lowered by the ``spatial 3--metric'' $\gamma_{ab}$. Notice that the geometry on a 3D spacelike hypersurface of constant time (i.e., $dt=0$) is determined by $\gamma_{ab}$. As we will see below, the Einstein equations control the evolution in time of this 3D geometry described by $\gamma_{ab}$, given appropriate initial conditions. The lapse function $\alpha$ and the shift vector $\beta^{a}$ determine how the slices are threaded by the spatial coordinates. Together, $\alpha$ and $\beta^a$ represent the coordinate degrees of freedom inherent in the covariant formulation of Einstein's equations, and can therefore be chosen, in some sense, ``freely'', as discussed below. This formulation of the equations assumes that one begins with an everywhere spacelike slice of spacetime, that should be evolved forward in time. Due to limited space, we will not discuss promising alternate treatments, based on either characteristic, or null foliations of spacetime\cite{Bishop98a}, or on asymptotically null slices of spacetime\cite{Friedrich81a,Friedrich81b,Friedrich96,Huebner96}. \subsection{Constraint Equations} \label{constraints} The constraints can be considered as the relativistic generalization of the Poisson equation of Newtonian gravity, but instead of a single linear elliptic equation there are now four, coupled, highly nonlinear elliptic equations, known as the hamiltonian and momentum constraints. Under certain conditions, the equations decouple and can be solved independently and more easily, and this is how they have been usually treated. Recently, techniques have been developed that allow one to solve the constraints in a more general setting, without making restrictive assumptions that lead to decoupling\cite{Bernstein97x,Laguna97x,Thornburg98,Miller98x}. In such a system the four constraint equations are solved simultaneously. This may prove useful in generating new classes of initial data. However, at present there is no satisfactory algorithm for controlling the physics content of the data generated. The major remaining work in this direction is to develop a scheme that is capable of constructing the initial data that describe a {\it given} physical system. That is, although we have schemes available to solve many variations on the initial value problem, it is difficult to specify in advance, for example, what are the precise spins and momenta of two black holes in orbit, or even if the hole {\em are} in orbit. This can generally only be determined after the equations have been solved and analyzed. The elliptic operators for these equations are usually symmetric, but they are otherwise the most general type, with all first and mixed second derivative terms present. The boundary conditions, which can break the symmetry, are usually linear conditions that involve derivatives of the fields being solved. In any case, once the initial value equations have been solved, initial data for the evolution problem result. We illustrate the central idea of constructing initial data with vacuum spacetimes for simplicity. The application of the algorithm presented here to a general spacetime with matter source is currently routine in numerical relativity. The full 4D Einstein equations can be decomposed into six evolution equations and four constraint equations. The constraints\index{constraints!in relativity} may be subdivided, in turn, into one Hamiltonian (or energy) constraint equation, \begin{equation} \label{Hamiltonian constraint} R+({\mathrm tr}\, K)^{2}-K^{ab}K_{ab}=0\; , \end{equation} and three momentum constraint equations (or one vector equation), \begin{equation} \label{momentum constraint} D_{b}(K^{ab}-\gamma^{ab}{\mathrm tr}\, K)=0\; . \end{equation} In these equations $K_{ab}$ is the extrinsic curvature of the slice, related to the time derivative of $\gamma_{ab}$ by \begin{equation} K_{ab} = -\frac{1}{2 \alpha} (\partial_t \gamma_{ab} - D_a \beta_b -D_b \beta_a)\; . \label{kdef} \end{equation} Here we have introduced the 3D spatial covariant derivative operator $D_a$ associated with the 3--metric $\gamma_{ab}$ (i.e. $D_a \gamma_{bc}=0$), and the 3D scalar curvature $R$ computed from $\gamma_{ab}$. These four constraint equations can be used to determine initial data for $\gamma_{ab}$ and $K_{ab}$, which are to be evolved with the evolution equations to be discussed below. These equations (2,3) are referred to as constraints because, as in the case of electrodynamics, they contain no time derivatives of the fundamental fields $\gamma_{ab}$ and $K_{ab}$, and hence do not propagate the solution in time. Next, we will sketch the standard method for obtaining a solution to these constraint equations. We follow York and coworkers (e.g., \cite{York89}) by writing the 3--metric and extrinsic curvature in ``conformal form'', and also make use of the simplifying assumption $\mathrm{tr}K = 0$ which causes the Hamiltonian and momentum constraints to completely decouple (note that actually the equations decoupled with $\mathrm{tr}K = {\rm const.}$ but we will discuss only the simplest case here). We write \begin{equation} \label{conformal form} \gamma_{ab}=\Psi^{4}\hat{\gamma}_{ab}, K_{ab} = \Psi^{-2}\hat{K}_{ab}\; , \end{equation} where $\hat{\gamma}_{ab}$ and the transverse tracefree part of $\hat{K}_{ab}$ is regarded as given, i.e., chosen to represent the physical system that we want to study. Under the conformal transformation, with $\mathrm{tr}K = 0$ we find that the momentum constraint becomes \begin{equation} \hat{D}_b \hat{K}^{ab} = 0\; , \label{eq10} \end{equation} where $\hat{D}_a$ is the 3D covariant derivative associated with $\hat{\gamma}_{ab}$ (i.e., $\hat{D}_a \hat{\gamma}_{ab}=0$). In vacuum, black hole spacetimes $\hat{K}_{ab}$ can often be solved analytically. For more details on how to solve the momentum constraints in complicated situations, please see \cite{York79,Cook93,Nakamura89}. The remaining unknown function $\Psi$, must satisfy the Hamiltonian constraint. The conformal transformation of the scalar curvature is \begin{equation} R=\Psi^{-4}\hat{R}-8\Psi^{-5}\hat{\Delta}\Psi\; , \end{equation} where $\hat{\Delta} = \hat{\gamma}^{ab} \hat{D}_a \hat{D}_b$ and $\hat{R}$ is the scalar curvature of the known metric $\hat{\gamma}_{ab}$. Plugging this back in to the Hamiltonian constraint and dividing through by $-8 \Psi^{-5}$, we obtain \begin{equation} \hat\Delta\Psi-\frac{1}{8}\Psi \hat{R}+\frac{1}{8} \Psi^{-7} \left( \hat{K}_{ab} \hat{K}^{ab}\right) = 0\; , \label{psieq} \end{equation} an elliptic equation for the conformal factor $\Psi$. To summarize, one first specifies $\hat{\gamma}_{ab}$ and the transverse tracefree part of $\hat{K}_{ab}$ ``at will'', choosing them to be something ``closest'' to the spacetime one wants to study. Then one solves (\ref{eq10}) for the conformal extrinsic curvature $\hat{K}_{ab}$. Finally, (\ref{psieq}) is solved for the conformal factor $\Psi$, so the full solution $\gamma_{ab}$ and $K_{ab}$ can be reconstructed. In this process the elliptic equations are solved by standard techniques, e.g., the conjugate gradient~\cite{Ashby90} or multigrid methods~\cite{Cook89}. In situations where there is a black hole singularity, there could be added complications in solving the elliptic equations, and special treatments would have to be introduced, e.g., the ``puncture'' treatment of \cite{Brandt97b}, or employing an ``isometry'' operation to provide boundary conditions on black hole throats, ensuring identical spatial geometries inside and outside the throat (see, e.g.,~\cite{Cook91,Cook93}, or~\cite{Seidel96a} for more details). While this is a well established process for generating an initial data set for numerical study, there is a fundamental difficulty in using this approach to generate initial data corresponding to a physical system one wants to evolve, e.g., a coalescing binary system. It is not clear how to choose the ``closest'' $\hat{\gamma}_{ab}$, and the corresponding free components in $\hat{K}_{ab}$, so that the resulting $\gamma_{ab}$ and $K_{ab}$ represents the inspiraling system at its late stage of inspiral. This late stage is the so-called ``intermediate challenge problem'' of binary black holes~\cite{Brady98a}, an area of much current interest. \subsection{Evolution Equations} \label{evolution} \subsubsection{The standard evolution system} With the initial data $\gamma_{ab}$ and $K_{ab}$ specified, we now consider their evolution in time. There are six evolution equations for the 3--metric $\gamma_{ab}$ that are second order in time, resulting from projections of the full 4D Einstein equations onto the 3D spacelike slice \cite{York79}. These are most often written as a first-order-in-time system of twelve evolution equations, usually referred to as the ``ADM'' evolution system~\cite{Arnowitt62,York79}: \begin{eqnarray} \partial_{t}\gamma_{ab} & = & -2\alpha K_{ab}+ D_{a}\beta_{b}+D_{b}\beta_{a}\; \label{metric evolution}\\ \partial_{t}K_{ab} & = & -D_{a}D_{b}\alpha+\alpha \left[ R_{ab}+\mbox({tr}K)K_{ab}-2K_{ac}K^{c}{}_{b} \right] \nonumber\\ & & +\beta^{c}D_{c}K_{ab}+ K_{ac}D_{b}\beta^{c}+K_{cb}D_{a}\beta^{c}\; .\label{excurv evolution} \end{eqnarray} Here $R_{ab}$ is the Ricci tensor of the 3D spacelike slice labeled by a constant value of $t$. Note that these are quantities defined only on a $t=const$ hypersurface, and require only the 3--metric $\gamma_{ab}$ in their construction. Do not confuse them with the conventional 4D objects! The complete set of Einstein equations are contained in constraint equations (\ref{Hamiltonian constraint}), (\ref{momentum constraint}) and the evolution equations (\ref{excurv evolution}), (\ref{metric evolution}). Note that (\ref{metric evolution}) is simply the definition of the extrinsic curvature $K_{ab}$ (\ref{kdef}). These equations are analogous to the evolution equations for the electric and magnetic fields of electrodynamics. Given the ``lapse'' $\alpha$ and ``shift'' $\beta^a$, discussed below, they allow one to advance the system forward in time. \subsubsection{Hyperbolic evolution systems} The evolution equations (\ref{excurv evolution}), (\ref{metric evolution}) have been presented in the ``standard ADM form'', which has served numerical relativity well over the last few decades. However, the equations are enormously complicated; the complication is hidden in the definition of the curvature tensor $R_{ab}$ and the covariant differentiation operator $D_a$. In particular, although they describe physical information propagating with a finite speed, the system does not form a hyperbolic system, and is not necessarily the best for numerical evolution. Other fields of physics, in particular hydrodynamics, have developed very mature numerical methods that are specially designed to treat the well studied flux conservative, hyperbolic system of balance laws having the form \begin{equation} \partial_t {\bf u} + \partial_k F^k_-{\bf u} = S_-{\bf u} \label{3Dbalance} \end{equation} where the vector ${\bf u}$ displays the set of variables and both ``fluxes'' $F^k$ and ``sources'' $S$ are vector valued functions. In hydrodynamic systems, it often turns out that the characteristic matrix $\partial F / \partial {\bf u} $ projected into any spacelike direction can often be diagonalized, so that fields with definite propagation speeds can be identified (the eigenvectors and the eigenvalues of the projected characteristic matrix). One important point is that in (\ref{3Dbalance}) all spatial derivatives are contained in the flux terms, with the source terms in the equations containing no derivatives of the eigenfields. All of these features can be exploited in numerical finite difference schemes that treat each term in an appropriate way to preserve important physical characteristics of the solution. Amazingly, the complete set of Einstein equations can also be put in this ``simple'' form (the source terms still contain thousands of terms however). Building on earlier work by Choquet-Bruhat and Ruggeri\cite{Choquet83}, Bona and Mass\'o began to study this problem in the late 1980's, and by 1992 they had developed a hyperbolic system for the Einstein equations with a certain specific gauge choice\cite{Bona92} (see below). Here by hyperbolic, we mean simply that the projected characteristic matrix has a complete set of eigenfields with real eigenvalues. This work was generalized recently to apply to a large family of gauge choices\cite{Bona94b,Bona97a}. The Bona-Mass\'o system of equations is available in the 3D ``Cactus'' code~\cite{Bona98b,Alcubierre98a}, as is the standard ADM system, where both are tested and compared on a number of spacetimes. The Bona-Mass\'o system is now one among many hyperbolic systems, as other independent hyperbolic formulations of Einstein's equations were developed\cite{Fritelli94,Choquet95,Abrahams95a,Fritelli95,MVP95,Abrahams97b} at about the same time as Ref.~\cite{Bona94a}. Among these other formulations only the one originally devised in Ref.~\cite{Abrahams95a} has been applied to spacetimes containing black holes\cite{Scheel97}, although still only in the spherically symmetry 1D case (a 3D version is under development\cite{CookScheelPrivateComm}.) Hence, of the many hyperbolic variants, only the Bona-Mass\'o family and the formulations of York and co-workers have been tested in any detail in 3D numerical codes. Notably among the differences in the formulations, the Bona-Mass\'o and Fritelli families contain terms equivalent to second time derivatives of the three metric $\gamma_{ab}$, while many other formulations go to a higher time derivative to achieve hyperbolicity. Another comment worth making is that for harmonic slicing, both the Bona-Mass\'o and York families have characteristic speeds of either zero, or light speed. For maximal slicing, they both reduce to a coupled elliptic-hyperbolic system. The Bona-Mass\'o system (at least) also allows for an additional family of explicit algebraic slicings, with the lapse proportional to an explicit function of the determinant of the three--metric, and in those cases one can also identify gauge speeds which can be different from light speed (harmonic slicing is one example of this family where the gauge speed corresponds to light speed). Some of these slicings, such as ``1+log''~\cite{Anninos94c}, have been found to be very useful in 3D numerical evolutions. This information about the speed of gauge and physical propagation can be very helpful in understanding the system, and can also be useful in developing numerical methods. Only extensive numerical studies will tell if the various hyperbolic formulations live up to their promise. Reula has recently reviewed, from the mathematical point of view, most of the recent hyperbolic formulations of the Einstein equations\cite{Reula98a} (This article, in the online journal ``Living Reviews in Relativity'', will be periodically updated). It is important to realize that the mathematical relativity field has been interested in hyperbolic formulations of the Einstein equations for many years and some systems that could have been suitable for numerical relativity were already published in the 1980's\cite{Choquet83,Friedrich85}. However, these developments were generally not recognized by the numerical relativity community until recently. \subsubsection{Numerical techniques for the evolution equations} Most of what has been attempted in numerical relativity evolution schemes is built on explicit finite difference schemes. Implicit and iterative evolution schemes have been occasionally attempted, but the extra cost associated has made them less popular. We now describe the basic approach that has been tried for both the standard ADM formulation and more recent hyperbolic formulations of the equations. \paragraph{ADM evolutions} The ADM system of evolution equations is often solved using some variation of the leapfrog method, similar to that described in have been used successfully. The most extensively tested is the ``staggered leapfrog'', detailed in axisymmetric cases in Ref.~\cite{Bernstein93b} and in 3D in Ref.~\cite{Anninos94c}, but other successful versions include full leapfrog implementations used in 3D by \cite{Bruegmann96} and \cite{Bona98b}. For the ADM system, the basic strategy is to use centered spatial differences everywhere, march forward according to some explicit time scheme, and hope for the best! Generally, this technique has worked surprisingly well until large gradients are encountered, at which time the methods often break down. The problem is that the equations in this ADM form are difficult to analyze, and hence ad hoc numerical schemes are often tried without detailed knowledge of how to treat specific terms in the equations, or how to treat instabilities when they arise. A recent development is that of the ``deloused'' leapfrog, which amounts to filtering the solution\cite{New98}. Also recently, the iterative Crank-Nicholson scheme has been found effective in suppressing some instabilities that occur~\cite{Huq98x}. \paragraph{Hyperbolic evolutions} The hyperbolic formulations are on a much firmer footing numerically than the ADM formulation, as the equations are in a much simpler form that has been studied for many years in computational fluid dynamics. However, the application of such methods to relativity is quite new, and hence the experience with such methods in this community is relatively limited. Furthermore, the treatment of the highly nonlinear source terms that arise in relativity is very much unexplored, and the source terms in Einstein's equations are much more complicated than those in hydrodynamics. A standard technique for equations having flux conservative form is to split Eq.~(\ref{3Dbalance}) into two separate processes. The transport part is given by the flux terms \begin{equation} \partial_t {\bf u} + \partial_k F^k_-{\bf u} = 0 \;\; . \label{3Dtransport} \end{equation} The source contribution is given by the following system of {\em ordinary} differential equations \begin{equation} \partial_t {\bf u} = S_-{\bf u} \;\; . \label{3Dsources} \end{equation} Numerically, this splitting is performed by a combination of both flux and source operators. Denoting by $E(\Delta t)$ the numerical evolution operator for system (\ref{3Dbalance}) in a single timestep, we implement the following combination sequence of subevolution steps: \begin{equation} E(\Delta t) = S(\Delta t/2)\;T(\Delta t)\;S(\Delta t/2) \label{splitting} \end{equation} where $T$, $S$ are the numerical evolution operators for systems (\ref{3Dtransport}) and (\ref{3Dsources}), respectively. This is known as ``Strang splitting''~\cite{Press86}. As long as both operators $T$ and $S$ are second order accurate in $\Delta t$, the overall step of operator $E$ is also second order accurate in time. This choice of splitting allows easy implementation of different numerical treatments of the principal part of the system without having to worry about the sources of the equations. Additionally, there are numerous computational advantages to this technique, as discussed in \cite{Clune98a}. The sources can be updated using a variety of ODE integrators, and in ``Cactus'' the usual technique involves second order predictor-corrector methods. Higher order methods for source integration can be easily implemented, but this will not improve the overall order of accuracy. However, in special cases where the evolution is largely source driven\cite{Masso92}, it may be important to use higher order source operators, and this method allows such generalizations. The details can be found in Ref.~\cite{Bona98b}. The implementation of numerical methods for the flux operator is much more involved, and there are many possibilities, ranging from standard choices to advanced shock capturing methods\cite{Leveque92,Bona96a,Bona97a}. Among standard methods, the MacCormack method, which has proven to be very robust in the computational fluid dynamics field (see, e.g., Ref.~\cite{Yee88} and references therein), and a directionally split Lax-Wendroff method have been implemented and tested extensively in ``Cactus''. These schemes are fully second order in space and time. Shock capturing methods have been shown to work extremely well in 1D problems in numerical relativity~\cite{Bona94b,Bona96a}, but their application in 3D is an active research area full of promise, but as yet, unfulfilled. The details of these methods, as they are applied to the Bona-Mass\'o formulation of the equations, can be found in Refs.~\cite{Bona96a,Bona98b}. \subsubsection{The Role of Constraints} If the constraints are satisfied on the initial hypersurface, the evolution equations then guarantee that they remain satisfied on all subsequent hypersurfaces. Thus once the initial value problem has been solved, one may advance the solution forward in time by using only the evolution equations. This is the same situation encountered in electrodynamics as discussed above. However, in a numerical solution, the constraints\index{constraints!in relativity} will be violated at some level due to numerical error. They hence provide useful indicators for the accuracy of the numerical spacetimes generated. Traditional alternatives to this approach, which is often referred to as ``free evolution'', involve solving some or all of the constraint equations on each slice for certain metric and extrinsic curvature components, and then simply monitoring the ``left over'' evolution equations. This issue is discussed further by Choptuik in \cite{Choptuik91}, and in detail for the Schwarzschild spacetime in \cite{Bernstein89}. New approaches to this problem of constraint vs. evolution equations are currently being pursued by Lee~\cite{Lee93,Lee94a}, among others. This approach is to advance the system forward using the evolution equations, and then adjust the variables slightly so that the constraints are satisfied (to some tolerance), i.e., the solution is projected onto the constraint surface. Because there are many variables that go into the constraints, there is not a unique way to decide which ones to adjust and by how much. But one can compute the ``minimum'' perturbation to the system, which corresponds to projecting to the {\em closest} point on the constraint surface. Other approaches, similar in spirit to each other, have been suggested by Detweiler~\cite{Detweiler87} and Brodbeck et al~\cite{Brodbeck98}. The Detweiler approach restricts the numerical evolution to the constraint surface by adding terms to the evolution equations (\ref{metric evolution}), (\ref{excurv evolution}) terms which are proportional to the constraints. Numerical tests of the scheme using gravitational wave spacetimes have recently been carried out, showing promising results~\cite{Lai98}. \subsubsection{Gauge Conditions} \label{gauge} Kinematic conditions for the lapse function $\alpha$ and shift vector $\beta^i$ have to be specified for the evolution equations (\ref{metric evolution}), and (\ref{excurv evolution}). With $\gamma_{ab}$ and $K_{ab}$ satisfying the constraint on the initial slice, the lapse and shift can be chosen {\em arbitrarily} on the initial slice and thereafter. These are referred to as gauge choices\index{gauge conditions!in relativity}, analogous to the choice of the gauge function $\Lambda$ in electrodynamics. Einstein did not specify these quantities; they are up to the numerical relativist to choose at will. \paragraph{Lapse.} The choice of lapse corresponds to how one chooses 3D spacelike hypersurfaces in the 4D spacetime. The ``lapse'' of {\em proper} time along the normal vector of one slice to the next is given by $\alpha d t$, where $d t$ is the {\em coordinate} time interval between slices. As $\alpha(x,y,z)$ can be chosen at will on a given slice, some regions of spacetime can be made to evolve farther into the future than others. There are many motivations for particular choices of lapse. A primary concern is to ensure that it leads to a stable long term evolution. It is easy to see that a naive choice of the lapse, e.g., $\alpha =1$, the so-called geodesic slicing, suffers from a strong tendency to produce coordinate singularities~\cite{Smarr78a,Smarr78b}. A related concern is that one would like to cover the region of interest in an evolution, say, where gravitational waves generated by some process could be detected, while avoiding troublesome regions, say, inside black holes where singularities lurk (the so-called ``singularity avoiding'' time slicings). Another important motivation is that some choices of $\alpha$ allow one to write the evolution equations in forms that are especially suited to numerical evolution. Finally, computational considerations also play important role in choice of the lapse; one prefers a condition for $\alpha$ that does not involve great computational expense, while also providing smooth, stable evolution. Some ``traditional'' choices of the lapse used in the numerical construction of spacetimes are~\cite{Piran83}: (1.) Lagrangian slicing, in which the coordinates are following the flow of the matter in the simulation. This choice simplifies the matter evolution equations, but it is not alway applicable, e.g., in a vacuum spacetime or when the fluid flow pattern becomes complicated. (2.) Maximal slicing,~\cite{Smarr78a,Smarr78b} in which the trace of the extrinsic curvature is required to be zero always, i.e, $K(t=0) = 0 = \partial_t K$. The evolution equations of the extrinsic curvature then lead to an elliptic equation for the lapse \begin{equation} D^a D_a \alpha - \alpha (R + K^2) =0 . \label{maximal} \end{equation} The maximal slicing has the nice property of causing the lapse to ``collapse'' to a small value at regions of strong gravity, hence avoiding the region that a curvature singularity is forming. It is one of the so-called ``singularity avoiding slicing conditions''. Maximal slicing is easily the most studied slicing condition in numerical relativity. (3.) Constant mean curvature, where we let $K = constant$ different from zero, a choice often used in constructing cosmological solutions. (4.) Algebraic slicing, where the lapse is given as an algebraic function of the determinant of the three metric. Algebraic slicing can also be singularity avoiding~\cite{Bona88}. As there is no need to solve an elliptic equation as in the case of maximal slicing, algebraic slicing is computationally efficient. Some algebraic slicings (e.g., the harmonic slicing in which $\alpha$ is set proportional to the square root of the determinant of the 3-metric $g_{ab}$) also make the mathematical structure of the evolution equations simpler. However, the local nature of the choice of the lapse could lead to noise in the lapse~\cite{Anninos94c} and the formation of ``shock'' like features in numerical evolutions~\cite{Alcubierre97a,Alcubierre97b}. The former problem can be dealt with by turning the algebraic slicing equation to an evolution equation with a diffusion term~\cite{Anninos94c}, but the latter problem does not seem to have a simple solution. In addition to these most widely used ``traditional'' choices of the lapse, there are also some newly developed slicing conditions whose use in numerical relativity though promising remain to be largely unexplored~\cite{Tobias96} : (5.) K-driver. This is a generalization of the maximal slicing in which the extrinsic curvature, instead of being set to zero, is required to satisfy the condition \begin{equation} \partial_t K = - c K , \label{kdrive} \end{equation} where c is some positive constant. This was first brought up by Eppley,~\cite{Eppley79} and recently investigated in of the extrinsic curvature, when numerical inaccuracy causes it to drift away from zero, is ``driven'' back to zero exponentially. When combined with the evolution equations, (\ref{kdrive}) again leads to an elliptic equation for the lapse. This choice of the lapse is shown~\cite{Balakrishna96a} to lead to a much more stable numerical evolution in cases where one wants to avoid large values of the extrinsic curvature. The optimal choice of the constant $c$ as well as a number of variations on this ``driver'' scheme are presently being studied. (6.) $\gamma$- driver. This is another use of the ``driver'' idea. In this case, the time rate of change of the determinant of the three metric $det(g_{ab})$ is driven to zero ~\cite{Balakrishna96a}. In the absent of a shift vector or if the shift has zero divergence, this reduces to the K-driver. This choice of the lapse, which has the unique property of being able to respond to the choice of the shift, demands extensive investigations and evaluations. \paragraph{Shift.} The shift vector describes the ``shifting'' of the coordinates from the normal vector as one moves from one slice to the next. If the shift vanishes, the coordinate point $(x,y,z)$ will move normal to a given 3D time slice to the next slice in the future. Please refer to York~\cite{York79} or Cook~\cite{Cook90}, for details and diagrams. The choice of shift is perhaps less well developed than the choice of lapse in numerical relativity, and many choices need to be explored, particularly in 3D. The main purpose of the shift is to ensure that the coordinate description of the spacetime remains well behaved throughout the evolution. With an inappropriate or poorly chosen shift, coordinate lines may move toward each other, or become very stretched or sheared, leading to pathological behavior of the metric functions that may be difficult to handle numerically. It may even cause the code to crash, if for example, two coordinate lines ``touch'' each other creating a ``coordinate singularity'' (i.e., the metric becomes singular as the distance $ds$ between two coordinate lines goes to zero). Two important considerations for appropriate shift conditions are the ability to prevent large shearing or drifting of coordinates during an evolution, and the ability to control the coordinate location of a physical object, e.g., the horizon of a black hole. These considerations are discussed below. The development of appropriate shift conditions for full 3D evolution, for systems without symmetries, is an important research area that needs much attention. Geometrical shift conditions that can be formulated without reference to specific coordinate systems or symmetries seem to be desirable. The basic idea is to develop a condition that minimizes the stretching, shearing, and drifting of coordinates in a general way. A few examples have been devised which partially meet these goals, such as ``minimal distortion'', ``minimal strain'', and variations \cite{York79}, but much more investigations are needed. New gauge conditions, based on these earlier proposals, have recently been proposed but not yet tested in numerical simulations~\cite{Brady98a}. It is important to emphasize that the lapse and shift {\em only} change the way in which the slices are chosen through a spacetime and where coordinates are laid down on every slice, and do {\em not}, in principle, affect any physical results whatsoever. They {\em will} affect the value of the metric quantities, but not the physics derived from them. In this respect the freedom of choice in the lapse and shift is analogous to the freedom of gauge in electromagnetic systems. On the other hand, it is also important to emphasize that proper choices of lapse and shift are crucial for the numerical construction of a spacetime in the Einstein theory of general relativity, in particular in a general 3D setting. In a general 3D simulation without symmetry assumption, there is no preferred choice of the form of the metric (e.g., a diagonal 3-metric, or $g_{\theta \theta} = r ^ 2 $ as in spherical symmetry), hence forcing us to deal with the gauge degree of freedom in relativity in full. This, when coupled with the inevitable lower resolution in 3D simulations, often leads to development of coordinate singularities, when evolved without a sophisticated choice of lapse and shift. Indeed the success of the ``driver'' idea suggested~\cite{Balakrishna96a} that in order to obtain a stable evolution over a long time scale, it is important to ensure that the coordinate conditions used are not only suitable for the geometry of the spacetime being evolved, but also that {\it the conditions themselves are stable}. That is, when the condition is perturbed, e.g., by numerical inaccuracy, there is no long term secular drifting. We regard the construction of an algorithm for choosing a suitable lapse and shift for a general 3D numerical simulation to be one of the most important issues facing numerical relativity at present. \subsection{General Relativistic Hydrodynamics} In order to make numerical relativity a tool for computational general relativistic astrophysics, it is important to combine numerical relativity with traditional tools of computational astrophysics, and in particular relativistic hydrodynamics. While a large amount of 3D studies in numerical relativity have been devoted to the {\em vacuum} Einstein equations, the spacetime dynamics with a non-vanishing source term remains a large uncharted territory. As astrophysics of compact objects that needs general relativity for its understanding is attracting increasing attention, general relativistic hydrodynamics will become an increasingly important subject as astrophysicists begin to study more relativistic systems, as relativists become more involved in studies of astrophysical sources. This promises to be one of the most exciting and important areas of research in relativistic astrophysics in the coming years. Previously, most work in relativistic hydrodynamics has been done on fixed metric backgrounds. In this approximation the fluid is allowed to move in a relativistic manner in strong gravitational fields, say around a black hole, but its effect on the spacetime is not considered. Over the last years very sophisticated methods for general relativistic hydrodynamics have been developed by the Valencia group led by Jos\'e M. Ib\'a\~{n}ez~\cite{Marti91,Font94,Banyuls97,Donat98}. These methods are based on a hyperbolic formulation of the hydrodynamic equations, and are shown to be superior to traditional artificial viscosity methods for highly relativistic flows and strong shocks. However, just fixed background approximation is inadequate in describing a large class of problems which are of most interest to gravitational wave astronomy, namely those with substantial matter motion generating gravitational radiation, like the coalescences of neutron star binaries. As will be discussed in more detail below, we are constructing a multi-purpose 3D code for the NASA Neutron Star Grand Challenge Project~\cite{Nasa98} that contains the full Einstein equations coupled to general relativistic hydrodynamics. The spacetime part of the code is based on the ``Cactus'' code; the hydrodynamic part consists of both an artificial viscosity module,~\cite{Wang97} and a module (MAHC HYPERBOLIC\_HYDRO) based on modern shock capturing schemes~\cite{Font98b}. The ``MAHC'' general relativistic hydro code at present contains three hydro evolution methods~\cite{Font98b}: a flux split method, Roe's approximate Riemann solver~\cite{Roe81} and Marquina's approximate Riemann solver~\cite{Donat96,Donat98}. All three are based on finite-difference schemes employing approximate Riemann solvers to account explicitly for the characteristic information of the equations. These schemes are particularly suitable for astrophysics simulations that involve matter in (ultra)relativistic speeds and strong shock waves. In the flux split method, the flux is decomposed into the part contributing to the eigenfields with positive eigenvalues (fields moving to the right) and the part with negative eigenvalues (fields moving to the left). These fluxes are then discretized with one sided derivatives (which side depends on the sign of the eigenvalue). The flux split method presupposes that the equation of state of the fluid has the form $P = P(\rho,\epsilon) = \rho f(\epsilon)$, which includes, e.g., the adiabatic equation of state. The second scheme, Roe's approximate Riemann solver~\cite{Roe81} is by now a ``traditional'' method for the integration of non-linear hyperbolic systems of conservation laws.~\cite{Font94,Eulderink94,Banyuls97} This method makes no assumption on the equation of state, and, is more flexible than the flux split methods. The third method, the Marquina's Method, is a promising new scheme.\cite{Donat96} It is based on a flux formula which is an extension of Shu and Osher's entropy-satisfying numerical flux~\cite{Shu89} to systems of hyperbolic conservation laws. In this scheme there are no artificial intermediate states constructed at each cell interface. This implies that there are no Riemann solutions involved (either exact or approximate); moreover, the scheme has been proved to alleviate several numerical pathologies associated to the introduction of an {\it averaged} state (as Roe's method does) in the local diagonalization procedure (see~\cite{Donat96,Donat98}). For a detailed comparison of the three schemes and their coupling to dynamical evolution of spacetimes, see ~\cite{Font98b}. The availability of the hyperbolic hydro treatment and its coupling to the spacetime evolution code is particularly noteworthy. With the development of a hyperbolic formulation of the Einstein equations described above, the {\em entire} system can be treated as a single system of hyperbolic equations, rather than artificially separating the spacetime part from the fluid part. Such a unified treatment based on the ``MAHC'' module is presently under construction by our NCSA/Potsdam/WashU collaboration. \subsection{Boundary Conditions} Appropriate conditions for the outer boundary have yet to be derived for 3D numerical relativity. In 1D and 2D relativity codes, the outer boundary is generally placed far enough away that the spacetime is nearly flat there, and static or flat (i.e., copying data from the next-to-last zone to the outer edge) boundary conditions can usually be specified for the evolved functions. However, due to the constraints placed on us by limited computer memory, this is not currently possible in 3D. Adaptive mesh refinement will be of great use in this regard, but will not substitute for proper physical treatment. Most results to date have been computed with the evolved functions kept static at the outer boundary, even if the boundaries are too close for comfort in 3D! There are several other approaches under development that promise to improve this situation greatly that we will not have room to explore in detail here, but should be mentioned. Generally, one has in mind using Cauchy evolution in the strong field, interior region where, say, black holes are colliding. This outer part of this region will be matched to some exterior treatment designed to handle what is primarily expected to be outgoing radiation. Two major approaches have been developed by the NSF Black Hole Grand Challenge Alliance, a large US collaboration working to solve the black hole coalescence problem, and other groups. First, by using perturbation theory, it is possible to identify quantities in the numerically evolved metric functions that obey the Regge-Wheeler and Zerilli wave equations that describe gravitational waves propagating on a black hole background. These can be used to provide boundary conditions on the metric and extrinsic curvature functions in an actual evolution, as described in a recent paper~\cite{Abrahams97a}. This is an excellent step forward in outer boundary treatments that should work to minimize reflections of the outgoing wave signals from the outer boundary. In tests with weak waves, a full 3D Cauchy evolution code has been successfully matched to the perturbative treatment at the boundary, permitting waves to escape from the interior region with very little reflection. Alternatively, ``Cauchy-Characteristic matching'' attempts to match spacelike slices in the Cauchy region to null slices at some finite radius, and the null slices can be carried out to null infinity. 3D characteristic evolution codes have progressed dramatically in recent years, and although the full 3D matching remains to be completed, tests of the scheme in specialized settings show promise\cite{Bishop98a}. One can also use the hyperbolic formulations of the Einstein equations to find eigenfields, for which outgoing conditions can in principle be applied\cite{Bona94b} in 1D. In 3D this technique is still under development, but it shows promise for future work. Finally, there is another hyperbolic approach which uses conformal rescaling to move the boundary to infinity~\cite{Friedrich81a,Friedrich81b,Friedrich96,Huebner96}. These methods have different strengths and weaknesses, but all promise to improve boundary treatments significantly, helping to enable longer evolutions than are presently possible. \subsection{Special Difficulties with Black Holes} The techniques described so far are generic in their application in numerical relativity. But in this section we describe a few problems that are characteristic of black holes, and special algorithms under development to handle them. Black hole spacetimes all have in common one problem: singularities lurk within them, which must be handled numerically. Developing suitable techniques for doing so is one of the major research priorities of the community at present. If one attempts to evolve directly into the singularity, infinite curvature will be encountered, causing any numerical code to break down. Traditionally, the singularity region is avoided by the use of ``singularity avoiding'' time slices, that wrap up around the singularity. Consider the evolution shown in Fig.~\ref{singularity}. A star is collapsing, a singularity is forming, and time slices are shown which avoid the interior while still covering a large fraction of the spacetime where waves will be seen by a distant observer. However, these slicing conditions by themselves do not solve the problem; they merely serve to delay the onset of instabilities. As shown in Fig.~\ref{singularity}, in the vicinity of the singularity these slicings inevitably contain a region of abrupt change near the horizon, and a region in which the constant time slices dip back deep into the past in some sense. This behavior typically manifests itself in the form of sharply peaked profiles in the spatial metric functions~\cite{Smarr78b}, ``grid stretching''~\cite{Shapiro86} or large coordinate shift~\cite{Bernstein89} on the BH throat, etc. Numerical simulations will eventually fail due to these pathological properties of the slicing. \begin{figure} \label{singularity} \hspace{1.5cm}\epsfysize=8cm \epsfbox{wrap.ps} \caption{A spacetime diagram showing the formation of a BH, and time slices traditionally used to foliate the spacetime in traditional numerical relativity with singularity avoiding time slices. As the evolution proceeds, pathologically warped hypersurfaces develop, leading to unresolvable gradients that cause numerical codes to crash.} \end{figure} \subsubsection{Apparent Horizon Boundary Conditions (AHBC)} The cosmic censorship conjecture suggests that in physical situations, singularities are hidden inside BH horizons. Because the region of spacetime inside the horizon is causally disconnected from the region of interest outside the horizon, one is tempted numerically to cut away the interior region containing the singularity, and evolve only the singularity-free region outside, as originally suggested by Unruh\cite{Unruh84}. This has the consequence that there will be a region inside the horizon that simply has no numerical data. To an outside observer no information will be lost since the regions cut away are unobservable. Because the time slices will not need such sharp bends to the past, this procedure will drastically reduce the dynamic range, making it easier to maintain accuracy and stability. Since the singularity is removed from the numerical spacetime, there is in principle no physical reason why BH codes cannot be made to run indefinitely without crashing. We spoke innocently about the BH horizon, but did not distinguish between the {\em apparent} and {\em event} horizon. These are very different concepts! While the event horizon, which is roughly a null surface that never reaches infinity and never hits the singularity, may hide singularities from the outside world in many situations, there is no guarantee that the apparent horizon, which is the (outermost) surface that has instantaneously zero expansion everywhere, even exists on a given slice! (By ``zero expansion'' we mean that the surface area of outgoing bundles of photons normal to the surface is constant. Hence, the surface is ``trapped.'') Methods for finding event horizons in numerical spacetimes are now known, and will be discussed below. But event horizons can only be found after examining the {\em history} of an evolution that has been already been carried out to sufficiently late times\cite{Anninos94f,Libson94a}. Hence they are useless in providing boundaries as one integrates {\em forward} in time. On the other hand the apparent horizon, if it exists, can be found on any given slice by searching for closed 2--surfaces with zero expansion. Although one should worry that in a generic BH collision, one may evolve into situations where no apparent horizon actually exists, let us cross that bridge if we come to it. Methods for finding apparent horizons will also be discussed below, but for now we assume that such a method exists. Given these considerations, there are two basic ideas behind the implementation of the apparent horizon boundary condition (AHBC), also known as black hole excision: {\em (a)} It is important to use a finite differencing scheme which respects the causal structure of the spacetime. Since the horizon is a one-way membrane, quantities on the horizon can be affected only by quantities outside but not inside the horizon: all quantities on the horizon can in principle be updated solely in terms of known quantities residing on or outside the horizon. There are various technical details and variations on this idea, which is called ``Causal Differencing''\cite{Seidel92a} or ``Causal Reconnection''\cite{Alcubierre94a}, but here we focus primarily on the basic ideas and results obtained to date. {\em (b)} A shift is used to control the motion of the horizon, and the behavior of the grid points outside the BH, as they tend to fall into the horizon if uncontrolled. An additional advantage to using causal differencing is that it allows one to follow the information flow to create grid points with proper data on them, as needed inside the horizon, even if they did not exist previously. (Remember above that we have cut away a region inside the horizon, so in fact we have no data there.) One example is to let a BH move across the computational grid. If a BH is moving physically, it may also be desirable to have it move through coordinate space. Otherwise, all physical movement will be represented by the ``motion'' of the grid points. For a single BH moving in a straight line, this may be possible (though complicated), but for spiraling coalescence this will lead to hopelessly contorted grids. The immediate consequence of this is that as a BH moves across the grid, regions in the wake of the hole, now in its exterior, must have previously been inside it where no data exist! But with AHBC and causal differencing this need not be a problem. Does the AHBC idea work? Preliminary indications are very promising. In spherical symmetry (1D), numerous studies show that one can locate horizons, cut away the interior, and evolve for essentially unlimited times ($t \propto 10^{3-4}M$, where $M$ is the black hole mass). The growth of metric functions can be completely controlled, errors are reduced to a very low level, and the results can be obtained with a large variety of shift and slicing conditions, and with matter falling in the BH to allow for true dynamics even in spherical symmetry\cite{Seidel92a,Anninos94e,Scheel94,Marsa96}. In 3D, the basic ideas are similar but the implementation is much more difficult. The first successful test of these ideas to a Schwarzschild BH in 3D used horizon excision and a shift provided from similar simulations carried out with a 1D code\cite{Anninos94c}. The errors were found to be greatly reduced when compared even to the 1D evolution with singularity avoiding slicings. Another 3D implementation of the basic technique was provided by Br{\"u}gmann~\cite{Bruegmann96}. This was a proof of principle, but more general treatments are following. Daues extended this work to a full range of shift conditions~\cite{Daues96a}, including the full 3D minimal distortion shift~\cite{York79}. He also applied these techniques to dynamic BH's, and collapse of a star to form a BH, at which point the horizon is detected, the region interior to the horizon excised, and the evolution continued with AHBC. The focus of this work has been on developing general gauge conditions for single BH's without movement through a grid. Under these conditions, BH's have been accurately evolved well beyond $t=100M$. The NSF Black Hole Grand Challenge Alliance had focussed on development of 3D AHBC techniques for moving Schwarzschild BH's\cite{Cook97a}. In this work, analytic gauge conditions are provided, which are chosen to make the evolution static, although the numerical evolution is allowed to proceed freely. This moving hole is the first successful 3D test of populating grid points with data as they emerge in the BH wake. These new results are significant achievements, and show that the basic techniques outlined above are not only sound, but are also practically realizable in a 3D numerical code. However, there is still a significant amount of work to be done! The techniques have yet to be applied carefully to distorted BH's, with tests of the waveforms emitted. They have not been applied to rotating BH's of any kind; they have not been applied to colliding BH's with horizon topology change, and moving black holes have yet to be evolved in AHBC with a nonanalytic gauge choice. There are still clearly many steps to be taken before the techniques will be successfully applied to the general BH merger problem. \section{Tools for Analyzing the Numerical Spacetimes} We now turn to the description of several important tools that have been developed to analyze the results of a numerical evolution, carried out by some numerical evolution scheme. The evolution will generally provide metric functions on a grid, but as described above these functions are highly dependent on both the coordinate system and gauge in which the system is evolved. Determining {\em physical} information, such as whether a black hole exists in the data, or what gravitational waveforms have been emitted, are the subjects of this section. \subsection{Horizon Finders} As described above, black holes are defined by the existence of an event horizon (EH), the surface of no return from which nothing, not even light, can escape. The event horizon is the boundary that separates those null geodesics that reach infinity from those that do not. The global character of such a definition implies that the position of an EH can only be found if the whole history of the spacetime is known. For numerical simulations of black hole spacetimes in particular, this implies that in order to locate an EH one needs to evolve sufficiently far into the future, up to a time where the spacetime has basically settled down to a stationary solution. Recently, methods have been developed to locate and analyze black hole horizons in numerically generated spacetimes, with a number of interesting results obtained~\cite{Anninos94f,Libson94a,Hughes94a,Matzner95a,Masso95a,Shapiro95a}. In contrast, an apparent horizon (AH) is defined locally in time as the outer-most marginally trapped surface~\cite{Hawking73a}, i.e. a surface on which the expansion of out-going null geodesics is everywhere zero. An AH can therefore be defined on a given spatial hypersurface. A well known result~\cite{Hawking73a} guarantees that if an AH is found, an EH must exist somewhere outside of it and hence a black hole has formed. \subsection{Locating the apparent horizons} The expansion $\Theta$ of a congruence of null rays moving in the outward normal direction to a closed surface can be shown to be~\cite{York89} \begin{equation} \Theta = \nabla_i s^i + K_{ij} s^i s^j - {\rm tr} K , \label{eqn:expansion} \end{equation} where $\nabla_i$ is the covariant derivative associated with the 3-metric $\gamma_{ij}$, $s^{i}$ is the normal vector to the surface, $K_{ij}$ is the extrinsic curvature of the time slice, and ${\rm tr} K$ is its trace. An AH is then the outermost surface such that \begin{equation} \Theta = 0. \label{eqn:horizon1} \end{equation} This equation is not affected by the presence of matter, since it is purely geometric in nature. We can use the same horizon finders without modification for vacuum as well as non-vacuum spacetimes. The key is to find a closed surface with normal vector $s^{i}$ satisfying this equation. \subsubsection{Minimization Algorithms} As apparent horizons are defined by the vanishing of the expansion on a surface, a fairly obvious algorithm to find such a surface involves minimizing a suitable norm of the expansion below some tolerance while adjusting test surfaces. Minimization algorithms for finding apparent horizons were among the first methods developed~\cite{Brill63,Eppley77}. More recently, a 3D minimization algorithm was developed and implemented by the Potsdam/NCSA/WashU group, applied to a variety of black hole initial data and 3D numerically evolved black hole spacetimes~\cite{Libson94b,Libson95a,Libson93a,Camarda97a,Camarda97c}. Essentially the same algorithm was also implemented independently by Baumgarte {\em et.al.}~\cite{Baumgarte96}. The basic idea behind a minimization algorithm is to assume the surface can be represented by a function $F(x^i)=0$, expand it the in terms of some set of basis functions, and then minimize the integral of the square of the expansion $\Theta^2$ over the surface. For example, one can parameterize a surface as \begin{equation} F(r,\theta,\phi) = r - h(\theta,\phi) . \label{eqn:F} \end{equation} The surface under consideration will be taken to correspond to the zero level of $F$. The function $h(\theta,\phi)$ is then expanded in terms of spherical harmonics: \begin{equation} h(\theta,\phi) = \sum_{l=0}^{l_{\rm max}} \sum_{m=-l}^{l} a_{lm} Y_{lm}(\theta,\phi) . \end{equation} Similar techniques were developed by~\cite{Nakamura84}. At an AH the expansion integral over the surface should vanish, and we will have a global minimum. Of course, since numerically we will never find a surface for which the integral vanishes exactly, one must set a given tolerance level below which a horizon is assumed to have been found. Minimization algorithms for finding AH's have a few drawbacks: First, the algorithm can easily settle down on a local minimum for which the expansion is not zero, so a good initial guess is often required. Moreover, when more than one marginally trapped surface is present, as is often the case, it is very difficult to predict which of these surfaces will be found by the algorithm: The algorithm can often settle on an inner horizon instead of the true AH. Again, a good initial guess can help point the finder towards the correct horizon. Finally, minimization algorithms tend to be very slow when compared with `flow' algorithms of the type described in the next section. Typically, if $N$ is the total number of terms in the spectral decomposition, a minimization algorithm requires of the order of a few times $N^2$ evaluations of the surface integrals (where in our experience `a few' can sometimes be as high as 10). This algorithm has been implemented in the ``Cactus'' code for 3D numerical relativity~\cite{Bona98b}. For more details of the application of this algorithm, see Refs.~\cite{Libson94b,Libson95a,Baumgarte96,Libson93a}. \subsubsection{3D fast flow algorithm} A second method that has been implemented in the ``Cactus'' code is the ``fast flow'' method proposed by Gundlach~\cite{Gundlach97a}. Starting from an initial guess for the $a_{lm}$, it approaches the apparent horizon through the iteration \begin{equation} \label{SpectralAHF} a_{lm}^{(n+1)} = a_{lm}^{(n)} -{A\over 1 + B l(l+1)} \left(\rho \Theta \right)_{lm}^{(n)}. \end{equation} where $(n)$ labels the iteration step, $\rho$ is some positive definite function (``a weight''), and $(\rho \Theta)_{lm}$ are the harmonic components of the function $(\rho \Theta)$. Various choices for the weight $\rho$ and the coefficients $A$ and $B$ parameterize a family of such methods. The fast flow algorithm in Cactus uses \begin{equation} \label{rho} \rho = 2 \; r^2 |\nabla F| \left[ \left(g^{ij}-s^i s^j \right) \left( \bar g_{ij}-\nabla_i r \nabla_j r \right) \right]^{-1} , \end{equation} where $\bar g_{ij}$ is the flat background metric associated with the coordinates $(r,\theta,\phi)$, and \begin{equation} \label{alphabeta} A = {\alpha\over l_{\rm{max}}(l_{\rm{max}}+1)} + \beta, \quad B = {\beta\over \alpha}. \end{equation} with $\alpha=c$ and $\beta=c/2$. Here $c$ is a variable step size, with a typical value of $c\sim 1$. $l_{\rm{max}}$ is the maximum value of $l$ one chooses to use in describing the surface. The iteration procedure is a finite difference approximation to a parabolic flow, and the adaptive step size is chosen to keep the finite difference approximation roughly close to the flow limit to prevent overshooting of the true apparent horizon. The adaptive step size is determined by a standard method used in ODE integrators: we take one full step and two half steps and compare the resulting $a_{lm}$. If the two results differ too much one from another, the step size is reduced. Other methods for finding apparent horizons, based directly on computing the jacobian of the finite differenced horizon equation, have been developed\cite{Thornburg95,Huq95} and successfully used in 3D. For details, please see these references. \subsection{Locating the event horizons} The AH is defined locally in time and hence is much easier to locate than the event horizon (EH) in numerical relativity. The EH is a global object in time; it is traced out by the path of outgoing light rays that {\em never} propagate to future null infinity, and {\em never} hit the singularity. (It is the boundary of the causal past of future null infinity $\dot{\cal{J}}^{{-}}(\cal{I}^{+})$.) In principle one needs to know the entire time evolution of a spacetime in order to know the precise location of the EH. However, in spite of the global properties of the EH, hope is not lost for finding it very accurately, even in a numerical simulation of finite duration. Here we discuss a method to find the EH, given a numerically constructed black hole spacetime that eventually settles down to an approximately stationary state at late times. In principle, as the EH is a null surface, it can be found by tracing the path of null rays through time. Outward going light rays emitted just outside the EH will diverge away from it, escaping to infinity, and those emitted just inside the EH will fall away from it, towards the singularity. In a numerical integration it is difficult to follow accurately the evolution of a horizon generator forward in time, as small numerical errors cause the ray to drift and diverge rapidly from the true EH. It is a physically unstable process. But we can actually use this property to our advantage by considering the time-reversed problem. In a global sense in time, any outward going photon that begins near the EH will be {\em attracted} to the horizon if integrated {\em backward} in time~\cite{Anninos94f,Libson94b}. In integrating backwards in time, it turns out that it suffices to start the photons within a fairly broad region where the EH is expected to reside. Such a horizon-containing region, as we call it, is often easy to determine after the spacetime has settled down to approximate stationarity. The crucial point is that when integrated backward in time along null geodesics, this horizon-containing region shrinks rapidly in ``thickness'', leading to a very accurate determination of the location of the EH at earlier times. Note that it is the earlier time when the black hole is under highly dynamical evolution that we are really interested in. Although one can integrate individual null geodesics backward in time, we find that there are various advantages to integrate the entire null surface backward in time. A null surface, if defined by $f(t,x^i)=0$ satisfies the condition \begin{equation} g^{\mu\nu} \partial_{\mu} f \partial_{\nu} f = 0\; . \label{nullsurface} \end{equation} Hence the evolution of the surface can be obtained by a simple integration, \begin{equation} \partial_t f = \frac{ - g^{ti} \partial_i f + \sqrt{(g^{ti}\partial_i f)^2 - g^{tt} g^{ij} \partial_i f \partial_j f} }{g^{tt}} \; . \label{evolveh} \end{equation} The inner and outer boundary of the horizon containing region when integrated backward in time, will rapidly converge to practically a single surface to within the resolution of the numerically constructed spacetime, i.e., a small fraction of a grid point. An accurate location of the event horizon is hence obtained. We henceforth shall represent the horizon surface as the function $f _ H (t,x^i)$. Aside from the simplicity of this method, there are a number of technical advantages as discussed in~\cite{Anninos94f}. One particularly noteworthy point is that this method is capable of giving the caustic structure of the event horizon if there is any; for details see~\cite{Anninos94f}. The function $f _ H (t,x^i)$ provides the complete coordinate location of the EH through the spacetime (or a very good approximation of it, as shown in~\cite{Libson94a}). This function by itself directly gives us the topology and location of the EH. When combined with the induced metric function on the surface, which is recorded throughout the evolution, it gives the intrinsic geometry of the EH. When further combined with the spacetime metric, all properties of the EH including its embedding can be obtained\index{horizons!event}. Moreover, as the normal of $f _ H (t,x^i) =0 $ gives the null generators of the horizon, it is an easy further step to determine the null generators, and hence the complete dynamics of the horizon in this formulation. As described in a series of papers, the event horizon, once found with such a method, can be analyzed to provide important information about the dynamics of black holes in a numerically generated spacetime~\cite{Anninos94f,Libson94a,Hughes94a,Matzner95a,Masso95a,Shapiro95a}. \subsection{Wave Extraction} The gravitational radiation emitted is one of the most important quantities of interest in many astrophysical processes. The radiation is generated in regions of strong and dynamic gravitational fields, and then propagated to regions far away where it will someday be detected. We take the approach of computing the generation and evolution of the fields in a fully nonlinear way, while analyzing the radiation with a perturbation formulation in the regions where it can be so treated. The theory of black hole perturbations is well developed. One identifies certain perturbed metric quantities that evolve according to wave equations on the black hole background. These perturbed metric functions are also dependent of the gauge in which they are computed. We use a {\em gauge-invariant} prescription for isolating wave modes on black hole background, developed first by Moncrief~\cite{Moncrief74}. The basic idea is that although the perturbed metric functions transform under coordinate transformations (gauge transformations), one can identify certain linear combinations of these functions that are invariant to first order of the perturbation. These gauge-invariant functions are the quantities that carry true physics, which does not depend on coordinate systems. They obey the wave equations describing waves propagating on the fixed blackhole background. There are two independent wave modes, even- and odd-parity, corresponding to the two degrees of freedom, or polarization modes, of the waves. A {\em waveform extraction} procedure has been developed that allows one to process the metric and to identify the wave modes. The gravitational wave function (often called the ``Zerilli function'' for even-parity or the ``Regge-Wheeler function'' for odd-parity) can be computed by writing the metric as the sum of a background black hole part and a perturbation: \begin{equation} g_{\alpha\beta}=\stackrel{o}{g}_{\alpha\beta}+h_{\alpha \beta}(Y_{\ell,m}), \end{equation} where the perturbation $h_{\alpha\beta}$ is expanded in spherical harmonics and their tensor generalizations and the background part $\stackrel{o}{g}_{\alpha\beta}$ is spherically symmetric. To compute the elements of $h_{\alpha\beta}$ in a numerical simulation, one integrates the numerically evolved metric components $g_{\alpha\beta}$ against appropriate spherical harmonics over a coordinate 2--sphere surrounding the black hole, making use of the orthogonality properties of the tensor harmonics. This process is performed for each $\ell , m$ mode for which waveforms are desired. The resulting functions $h_{\alpha \beta}(Y_{\ell,m})$ can then be combined in a gauge-invariant way, following the prescription given by Moncrief\cite{Moncrief74}. For each $\ell , m$ mode, this gauge invariant gravitational waveform can be extracted when the wave passes through ``detectors'' at some fixed radius in the computational grid. This procedure has been described in detail in~\cite{Brandt94a,Brandt94b,Brandt94c}, and more generally in Refs.~\cite{Allen97a,Allen98a,Camarda97c}. It works amazingly well, allowing extraction of waves that carry very small energies (of order $10^{-7}M$ or less, with $M$ being the mass of the source) away from the source. The procedure should apply to any isolated source of waves, such as colliding black holes, neutron stars, etc. If the sources are rotating, this procedure should be generalized to use the Teukolsky formalism describing perturbations about a Kerr black hole, but this has not yet been done. Instead, the spherical perturbation theory (with a few minor modifications) has been applied to distorted rotating black holes with satisfactory results~\cite{Brandt94a,Brandt94b,Brandt94c}. \section{Computational Science, Numerical Relativity, and the ``Cactus'' Code} \subsection{The Computational Challenges of Numerical Relativity} Before we describe our computational methods in the following subsections, we summarize the computational challenges of numerical relativity discussed above. It is in response to these challenges that we have devised the computational methods. \noindent $\bullet$ Computational challenges due to the complexity of the physics involved: The Einstein equations are probably the most complex partial differential equations in all of physics, forming a system of dozens of coupled, nonlinear equations, with thousands of terms, of mixed hyperbolic, elliptic, and even undefined types in a general coordinate system. The evolution has elliptic constraints that should be satisfied at all times. In simulations without symmetry, as would be the case for realistic astrophysical processes, codes can involve hundreds of 3D arrays, and ten of thousands of operations per grid point per update. Moreover, for simulations of astrophysical processes, we will ultimately need to integrate numerical relativity with traditional tools of computational astrophysics, including hydrodynamics, nuclear astrophysics, radiation transport and magneto-hydrodynamics, which govern the evolution of the source terms (i.e., the right hand side) of the Einstein equations. This complexity requires us to push the frontiers of massively parallel computation. \noindent $\bullet$ Challenge in Collaborative Technology: The integration of numerical relativity into computational astrophysics is a multi-disciplinary development, partly due to the complexity of the Einstein equations, and partly due to the physical systems of interest. Solving the Einstein equations on massively parallel computers involves gravitational physics, computational science, numerical algorithm and applied mathematics. Furthermore, for the numerical simulations of realistic astrophysical systems, many physics disciplines, including relativity, astrophysics, nuclear physics, and hydrodynamics are involved. It is therefore essential to have the numerical code software engineered to allow co-development by different research groups {\it and} groups with different expertise. \noindent $\bullet$ The numerical construction of a spacetime itself presents unique challenges: According to the singularity theorems of general relativity, regions of strong gravity often generate spacetime singularities. Due to the need to avoid spacetime singularities~\cite{Seidel96a,Seidel96b}, and to obtain long term stability in the numerical simulations, sophisticated control of the coordinate system is needed for the construction of a numerical spacetime. This dynamic interplay between the spacetime being constructed and the computational coordinate choice itself (``gauge choice'') is a unique feature of general relativity that makes the numerical simulations much more demanding. Besides extra computational power, advanced visualization tools, preferably real time interactive ``window into the oven'' visualization, are particularly useful in the numerical construction. \noindent $\bullet$ The multi-scale problem: Astrophysics of strongly gravitating systems inherently involves many length and time scales. The microphysics of the shortest scale (the nuclear force), controls macroscopic dynamics on the stellar scale, such as the formation and collapse of neutron stars (NSs). On the other hand, the stellar scale is at least 10 times {\it less} than the wavelength of the gravitational waves emitted, and many orders of magnitude less than the astronomical scales of their accretion disk and jets; these larger scales provide the directly observed signals. Numerical studies of these systems, aiming at direct comparison with observations, fundamentally require the capability of handling a wide range of dynamical time and length scales. All of these issues lead to important research questions in computational science. Here we give an overview of some of our effort in these directions, focusing on performance and coding issues on parallel machines, and on the development of a community code that incorporates all the mathematical and computational techniques described above (and many more), in a collaborative infrastructure for numerical relativity. \subsection{Code Generation and Data Parallel Fortran} When expanded out in a particular coordinate system the evolution equations for the full Einstein equations in the 3$+$1 formulation have many thousands of terms. These are usually derived and coded in Fortran with a symbolic manipulator package such as Mathematica or Macsyma. However, these packages often generate Fortran expressions that are unsuitable for most compilers, even on traditional supercomputers. We often exceed internal compiler limits on length of expression, number of continuation lines, number of arguments to a subroutine, number of nested parentheses, and so forth. So our code generation techniques need to be carefully massaged before an efficient, working code is generated. The evolution equations are generally written out using explicit finite difference schemes, which are very popular for hyperbolic systems of equations. These equations are good candidates for the ``SIMD'' style approach to programming parallel machines. (SIMD stands for ``Single Instruction Multiple Data'', which means an operation like ``add arrays A and B together'' can be carried out completely in parallel, with the same instruction (add) on multiple data elements in memory. This is also a so-called ``data parallel'' operation, since the same operation is applied simultaneously to all data elements of arrays A and B in parallel, and no communications are required between processors.) Until recently, in our research group 3D codes have been generally written in this style using data parallel Fortran90 and CMFortran style languages. With this approach, communications between processors, required for example when computing derivatives (which require knowledge of neighboring data points in memory), are handled by the compilers without need for the user to do anything. We have used the C-preprocessor to incorporate a few different code blocks so that we can maintain a single source file for several machines. (For an excellent review of many modern approaches to parallel computing, including further information on many of the concepts and acronyms common in computational science, see, e.g.,~\cite{Foster95a}, available both in print and on-line). Using this global approach we previously developed a single code, called H3expresso, that achieved over 15 Gflops on the 512 node CM-5 and about 8 Gflops on the 16 processor Cray C-90. This code was one of the fastest applications on either machine~\cite{Hillis93}. We performed a detailed comparative study of this code on many architectures, including the C-90, Convex SPP-1200, T3D, CM-5, SGI Power Challenge, and SP-2, achieving excellent scaling all machines. These results are possible because of the very high computation/communication ratio inherent in the Einstein equations. The hyperbolic equations contain thousands of terms to be evaluated, while the only communications required are in computing finite differences for numerical derivatives. \subsection{Data Parallel Fortran Evolves to MPI} However, this data parallel approach is not the best one to follow with more modern microprocessor based scalable supercomputers, such as the SGI/Cray Origin 2000 and Cray T3D and T3E, due largely to the use of caches that boost performance of a single node. It is worth commenting on how we have adapted the H3expresso code to a ``message passing'' language like MPI, with single processor optimizations, which then led to to the development of the new Cactus code described below. (MPI stands for ``Message Passing Interface'', a standard communications library now available on most parallel computers, that allows the user to explicitly control the communication of data between processors when required~\cite{Foster95a}. This can be more efficient than allowing the compiler to handle this automatically.) Due to the data-parallel nature of the code, many of the temporaries evolved in solving the hyperbolic equations (\ref{3Dbalance}), notably the sources and the fluxes, are created as 3D arrays. This allows fairly easy parallelization of the code with MPI. Since the only finite differencing in the code is on the fluxes, they are the only variables which need communication, and thus we can easily do an MPI-based communication with these variables during our update loop. Unfortunately, one of the major problems of the data parallel programming model is that it requires the creation of large numbers of 3D temporary arrays to store source or flux terms. On a system like the CM-5, this technique was crucial in obtaining performance; the arrays were distributed and were stored on the vector units, so the system could operate on them quickly and communicate them transparently. However, with single statements for entire arrays with large degrees of complexity, each assignment requires a sweep through the complete memory space. Cache locality is impossible, and the code performs very poorly in an out-of-cache regime. Hence, to achieve high single processor performance on more modern microprocessor based architectures special attention has to be paid to rewriting expressions to maximize the use of the processor cache. Using the experience gained from exploring issues of parallelism and single processor performance with the H3expresso code, we have developed from the scratch a new 3D Einstein equation code, the ``Cactus'' code, which integrates important design decisions for modern HPC (standard acronym for ``High Performance Computing'') architectures from the outset: \begin{itemize} \item The numerical kernals for the Einstein equations, needed by all users, are highly optimized for modern microprocessor based architecture. \item Other routines (e.g., waveform extraction) are written by the community of users in either C or Fortran. \item It obtains parallelism through MPI, not through compiler technologies. In the following, we describe some details of single processor performance, parallelism and the code's collaborative infrastructure. \end{itemize} \subsubsection{Parallelism using MPI} In ``Cactus'', we achieve parallelism using ghost-zone domain decomposition. That is, we decompose a global domain over our processors, and place an overlap region on each processor. Then each single processor is responsible for updating the interior of their local region, and a MPI communication synchronizes the boundary zones after communications. We have also optimized for the particular structure of our equations. The straightforward way to set up our communication patterns would be an algorithm like \begin{verbatim} for (it = 0 to maxit) { update interior for a time step dt update ghost zones for all grid functions } \end{verbatim} However, many of our variables have no flux. Since we generally use a strang split in the hyperbolic evolution, which allows us to update our source and flux evolution separately, the integration of these flux-free variables is a completely spatially de-coupled point local ODE. These variables need no communication whatsoever, in absence of flux. Thus, we can re-write the above loop as \begin{verbatim} for (it = 0 to maxit) { evolve sources for dt/2 evolve fluxes for dt update ghost zones for all grid functions which have a flux evolve sources for dt/2 } \end{verbatim} In practice, this algorithm allows us to reduce our communication cost, resulting in excellent scaling, in addition to excellent single processor performance. \begin{figure} \label{fig:nt_scal} \hspace{1.5cm}\epsfysize=8cm \epsfbox{nt.eps} \caption{We show scaling of the Cactus code on two very different architectures: an SGI/Cray Origin 2000 DSM architecture with 128 processors, and a cluster of 300Mhz Compaq workstations running Windows NT. The data is obtained with all "thorns" that are needed to evolve a gravitational wave packet using the vacuum Einstein equation.} \end{figure} \begin{figure} \label{fig:cactus_scal} \hspace{1.5cm}\epsfysize=8cm \epsfbox{t3e.eps} \caption{We show the scaling of the Cactus code on the Cray T3E-600, giving the total Mflops/sec as a function of the number of processors. Results are shown for single precision calculations, and include calculations performed on ghost zones. The grid size per processor is kept constant as the number of processors in increased (so the total problem size scales with the size the machine). The performance data is obtained for an evolution using the Einstein equations with the hydrodynamic source terms, and the relativistic hydrodynamics equations in high resolution shock capturing treatment as described in the paper. The inclusion of the hydrodynamics does not change the performance in any substantial manner. } \end{figure} These techniques have enabled ``Cactus'' to be a highly portable and efficient code for numerical relativity and astrophysics. It has been tested and performed very well on three very different parallel architectures: the SGI Origin 2000 system, the SGI/Cray T3E system, and a cluster of 128 NT workstations, developed at NCSA, running Pentium II processors. In Fig.~\ref{fig:cactus_scal} we show scaling results achieved on a Cray T3E-600, and in Fig.~\ref{fig:nt_scal} we show the scaling results achieved on both the Origin 2000 and the NT cluster. Presently, the SGI Origin, with 250MHz R10000 processors, has more than three times the per processor performance of the 300MHz Pentium II in the NT cluster, but the trend is very encouraging. These results indicate that codes like this can be run efficiently in parallel on a wide variety of machines. Finally, we have recently tested the code on a 1024 node T3E-1200 (provided for the NASA Neutron Star Grand Challenge Project~\cite{Nasa98} for performance tests), achieving 142GFlops and linear scaling up to 1024 nodes. The version of the code tested is the so-called GR3D code developed for the NASA Grand Challenge Project. GR3D is a version of ``Cactus'' code for spacetime evolution coupled to a Riemann solver based relativistic hydrodynamic code (MAHC HYPERBOLIC\_HYDRO) that has recently been released (available at http://wugrav.wustl.edu/Codes/GR3D/). Performance information for the ``Cactus'' code will also be kept up to date at http://cactus.aei-potsdam.mpg.de. In the following we give a summary of the performance test. The test was carried out with the released version without special tuning for this 1024 node machine. We note that the full set of the Einstein equations with the perfect fluid source, as solved in this code, involved a large number of 3D arrays. The huge number of grid points used (644 x 644 x 1284, or 500 x 500 x 996 respectively for 32 and 64 bits) is made possible by reduced memory footprint of the code. \noindent {\em \underline{Machine Configuration}}: 1024 node T3E-1200 with 512MB per node \noindent {\em \underline{Date tested}}: May 10, 1998 \begin{verbatim} 32 bit 64 bit -------------------------------------------------------------------- Grid Size per Processor 84x84x84 66x66x66 Processor topology 8 x8 x16 8 x8 x16 Total Grid Size 644 x 644 x 1284 500 x 500 x 996 Single Proc MFlop/sec 144.35 118.33 Aggregate GFlop/sec 142.2 115.8 Scaling efficiency 96.2\% 95.6\% -------------------------------------------------------------------- \end{verbatim} Previously, the largest production simulations in 3D numerical relativity have been limited to about $300^{3}$ or less, and applied to distorted 3D black hole systems~\cite{Camarda97b,Camarda97c,Allen97a,Allen98a}. When such large machines as the test T3E described above are made available for routine production simulations, we expect to further improve the results of the such black hole simulations, and perform more general 3D calculations involving distorted rotating black holes, coalescences of neutron stars, gravitational waves, as well as other interesting problems in general relativistic astrophysics. \subsection{Collaborative Infrastructure} While ``Cactus'' is our third generation 3D numerical relativity code, it is our first generation of code in which we paid special attention to the difficult software engineering problem of collaborative code development, maintenance and management. Our earlier generations of 3D numerical relativity codes (the so-called ``G'' and ``H'' codes, described in~\cite{Anninos94c,Anninos94d}) involving about a dozen researchers in the Potsdam/NCSA/Wash U collaboration, have made us keenly aware of the issues and difficulties involved in distributed collaborative code development. For the ``Cactus'' code, a collaborative infrastructure has been essential. As of this writing, several dozen collaborators at 7 institutions are using the code for various research projects, and we aim at further making it a truly community code for the investigation of general relativistic astrophysics. ``Cactus'' is hence designed to minimize barriers to the community development and use of the code, including the complexity associated with both the code itself and the networked supercomputer environments in which simulations and data analyses are performed. This complexity is particularly noticeable in large multidisciplinary simulations such as ours, because of the range of disciplines that must contribute to code development (relativity, hydrodynamics, astrophysics, numerics, and computer science) and because of the geographical distribution of the people and computer resources involved in simulation and data analysis. The collaborative technologies that we are developing within Cactus include: \noindent $\bullet$ {\em A modular code structure and associated code development tools}. Cactus defines coding rules that allow one, with only a working knowledge of Fortran or C, to write new code modules that are easily plugged in as ``thorns'' to the main Cactus code (the ``flesh''). The ``flesh'' contains basic computational infrastructure needed to develop and connect parallel modules into a working code. All told, the ``flesh'' is thousands of lines of highly optimized C and Fortran, not counting some utility libraries, makefile, and perl scripts. It allows one to plug in not only different physics modules, such as the basic Einstein solver, other formulations of the equations, analysis routines, etc., but also different parallel domain decomposition modules, I/O modules, utilities, elliptic solvers, and so forth. A thorn may be any code that the user wants, in order to provide different initial data, different matter fields, different gauge conditions, visualization modules, etc. Thorns need not have anything to do with relativity: the flesh could be used as basic infrastructure for any set of PDE's, from Newtonian hydrodynamics equations to Yang Mills equations, that are coded as thorns. The user inserts the hook to their thorn into the flesh code in a way that the thorn will not be compiled unless it is designated to be active. We have developed a makefile and perl-based thorn management system that, through the use of preprocessor macros that are appropriately expanded to the arguments of the flesh and additional arguments defined by each thorn, is able to create a code which can configure itself to have a variety of different modules. This ensures that {\em only} those variables needed for a particular simulation are actually used, and that no conflicts can be created in subroutine argument calling lists. \noindent $\bullet$ {\em A consistency test suite library.} An increased number of thorns makes the code more attractive to its community but also increases the risk of incompatibilities. Hence, we provide technology that allows each developer to create a test/validation suite for their own thorn. These tests are run prior to any check in of code to the repository, ensuring that new code reproduces results consistent with previous one, and hence cannot compromise the work of other developers relying on a given thorn. So, how does a user use the code? A detailed user guide will be available with the code when it is released during 1999 (see http://cactus.aei-potsdam.mpg.de), but in a nutshell, one specifies which physics modules, and which computational/parallelism modules, are desired in a configuration file, and makes the code on the desired architecture, which can be any one of a number of machines from SGI/Cray Origin or T3E, Dec Alpha, Linux workstations or clusters, NT clusters, and others. The make system automatically detects the architecture and configures the code appropriately. Control of run parameters is then provided through an input file. Additional modules can be selected from a large community-developed library, or new modules may be written and used in conjunction with the pre-developed modules. Our experiences with Cactus up to now suggest that these techniques are effective. It allows a code of many tens of thousands of lines, but with a compact flesh that is possible to maintain despite the large number of people contributing to it. The common code base has enhanced the collaborative process, having important and beneficial effects on the flow of ideas between remote groups. This flexible, open code architecture allows, for example, a relativity expert to contribute to the code without knowing the details of, say, the computational layers (e.g., message passing or AMR libraries) or other components (e.g., hydrodynamics). This is an area that we plan to invest more effort into in the coming few years. \subsection{Adaptive Mesh Refinement} 3D simulations of Einstein's equations are very demanding computationally. In this section we outline the computational needs, and techniques designed to reduce them. We will need to resolve waves with wavelengths of order $5M$ or less, where $M$ is the mass of the BH or the neutron star. Although for Schwarzschild black holes, the fundamental $\ell=2$ quasinormal mode wavelength is $16.8M$, higher modes, such as $\ell=4$ and above, have wavelengths of $8M$ and below. The BH itself has a radius of $2M$. More important, for very rapidly rotating Kerr BH's, which are expected to be formed in realistic astrophysical BH coalescence, the modes are shifted down to significantly shorter wavelengths\cite{Flanagan97a,Flanagan97b}. As we need at least 20 grid zones to resolve a single wavelength, we can conservatively estimate a required grid resolution of about $\Delta x = \Delta y = \Delta z \approx 0.2M$. For simulations of time scales of order $t \propto 10^{2}-10^{3}M$, which will be required to follow coalescence, the outer boundary will probably be placed at a distance of roughly $R \propto 100M$ from the coalescence, requiring a Cartesian simulation domain of about $200M$ across. This leads to about $10^{3}$ grid zones in each dimension, or about $10^{9}$ grid zones in total. As 3D codes to solve the full Einstein equations have typically 100 variables to be stored at each location, and simulations are performed in double precision arithmetic, this leads to a memory requirement of order 1000 Gbytes! (In fairness to some groups that use spectral methods instead of finite differences (e.g., the Meudon group), we should point out highly accurate 3D simulations can now be achieved on problems that are well suited to such techniques, using much less memory!~\cite{Bonazzola98a}). The largest supercomputers available to scientific research communities today have only about $\frac{1}{20}$ of this capacity, and machines with such capacity will not be routinely available for some years. Furthermore, if one needs to double the resolution in each direction for a more refined simulation, the memory requirements increase by an order of magnitude. Although such estimates will vary, depending on the ultimate effectiveness of inner or outer boundary treatments, gauge conditions, etc., they indicate that barring some unforeseen simplification, some form of adaptive mesh refinement (AMR) that places resolution only where it is required is not only desirable, but essential. The basic idea of AMR is to use some set of criteria to evaluate the quality of the solution on the present time step. If there are regions that require more resolution, then data are interpolated onto a finer grid in those regions; if less resolution is required, grid points are destroyed. Then the evolution proceeds to the next time step on this hierarchy of grids, where the process is repeated. These rough ideas have been refined and applied in many applications now in computational science. There are several efforts ongoing in AMR for relativity. Choptuik was the early pioneer in this area, developing a 1D AMR system to handle the resolution requirements needed to follow scalar field collapse to a BH\cite{Choptuik89}. As an initially regular distribution of scalar field collapses, it will require more and more resolution as its density builds up. The grid density required to resolve the initial distribution may not even see the final BH. Further, as pulses of radiation propagate back out from the origin, they, too may have to be resolved in regions where there was previously a coarse grid. Choptuik's AMR system, built on early work of Berger and Oliger\cite{Berger84}, was able to track dynamically features that develop, enabling him to discover and accurately measure BH critical phenomena that have now become so widely studied\cite{Choptuik93}. Based on this success and others, and on the general considerations discussed above, full 3D AMR systems are under development to handle the much greater needs of solving the full set of 3D Einstein equations. A large collaboration, begun by the NSF Black Hole Grand Challenge Alliance, has been developing a system for distributing computing on large parallel machines, called Distributed Adapted Grid Hierarchies, or DAGH. DAGH was developed to provide MPI-based parallelism for the kinds of computations needed for many PDE solvers, and it also provides a framework for parallel AMR. It is one of the major computational science accomplishments to come out of the Alliance. Developed by Manish Parashar and Jim Browne, in collaboration with many subgroups within and without the Alliance, it is now being applied to many problems in science and engineering. One can find information about DAGH online at http://www.cs.utexas.edu/users/dagh/. At least two other 3D software environments for AMR have been developed for relativity: one is called HLL, or Hierarchical Linked Lists, developed by Lee Wild and Bernard Schutz\cite{Wild98a}; another, called BAM, was the first AMR application in 3D relativity developed by Br{\"u}gmann~\cite{Bruegmann96}. The HLL system has recently been applied to the test problem of the Zerilli equation (discussed above) describing perturbations of black holes\cite{Papadopoulos98a}. This nearly 30 year old linear equation is still providing a powerful model for studying BH collisions, and it is also being used as a model problem for 3D AMR. In this work, the 1D Zerilli equation is recast as a 3D equation in cartesian coordinates, and evolved within the AMR system provided by HLL. Even though the 3D Zerilli equation is a single linear equation, it is quite demanding in terms of resolution requirements, and without AMR it is extremely difficult to resolve both the initial pulse of radiation, the blue shifting of waves as they approach the horizon, and the scattering of radiation, including the normal modes, far from the hole. \section{Summary} In this article we have attempted to review the essential mathematical and computational elements needed for a full scale numerical relativity code that can treat a variety of problems in relativistic astrophysics and gravitation. Various formulations of the Einstein equations for evolving spacelike time slices, techniques for providing initial data, the basic ideas of gauge conditions, several important analysis tools for discovering the physics contained in a simulation, and numerical algorithms for each of these items have been reviewed. Unfortunately, we have only been able to cover the basics of such a program, and in addition many promising alternative approaches have necessarily been left out. As one can see, the solution to a single problem in numerical relativity requires a huge range of computational and mathematical techniques. It is truly a large scale effort, involving experts in computer and computational science, mathematical relativity, astrophysics, and so on. For these reasons, aided by collaborations such as the NSF Black Hole Grand Challenge Alliance and the NCSA/Potsdam/WashU collaboration, there has been a great focusing of effort over the last years. A natural byproduct of this focusing has been the development of codes that are used and extended by large groups. A code must have a large arsenal of modules at its disposal: different initial data sets, gauge conditions, horizon finders, slicing conditions, waveform extraction, elliptic equation solvers, AMR systems, boundary modules, different evolution modules, etc. Furthermore, these codes must run efficiently on the most advanced supercomputers available. Clearly, the development of such a sophisticated code is beyond any single person or group. In fact, it is beyond the capability of a single community! Different research communities, from computer science, physics, and astrophysics, must work together to develop such a code. As an example of such a project, we described briefly the ``Cactus'' code, developed by a large international collaboration\cite{Bona98b}. This code is an outgrowth of the last 5 years of 3D numerical relativity development primarily at NCSA/Potsdam/WashU, and builds heavily on the experience gained in developing the so-called ``G'' and ``H'' codes~\cite{Anninos94c,Anninos94d,Bona98b}. Presently, it is being developed collaboratively by these groups in collaboration with groups at Palma, Valencia, Physical Research Laboratory in India, and computational science groups at U. of Illinois, and Argonne National Lab. The code has a very modular structure, allowing different physics, analysis, and computational science modules to be plugged in. In fact, versions of essentially all the modules listed above are already developed for the code. For example, several formulations of Einstein's equations, including the ADM formalism and the Bona-Mass\'o hyperbolic formulation, can be chosen as input parameters, as can different gauge conditions, horizon finders, hydrodynamics evolvers, etc. It is being tested on BH spacetimes, such as those described above, as well as on pure wave spacetimes, self-gravitating scalar fields and hydrodynamics. It has also been designed to connect to DAGH ultimately for parallel AMR. This code was also designed as a community code. After first developing and testing it within our rather large community of collaborators, it will be made available with full documentation via a public ftp server maintained at AEI and a mirroring site at WashU. By having an entire research community using and contributing to such a code, we hope to accelerate the maturation of numerical relativity. Information about the code is available online, and can be accessed at http://cactus.aei-potsdam.mpg.de. {\bf Acknowledgments} {It is a pleasure to acknowledge many friends and colleagues who have contributed to the work described in this article, some of which was derived from papers we have written together. The Cactus code was originally started by Joan Mass\'o and Paul Walker at AEI, and the MAHC general relativistic hydrodynamic module was started by Mark Miller at Wash U; these were then opened up to development by our entire research groups at Potsdam, Washington University, and NCSA (the NCSA/Potsdam/Wash U collaboration), and elsewhere, notably the University of the Balearic Islands in Spain, the University of Valencia, Argonne National Lab, and Physical Research Lab (PRL) in India. Without the contributions from people at all these institutions, the work described here would not have been possible. Thanks to Tom Goodale and Ed Evans for carefully reading the manuscript and suggesting improvements, to Tom Clune for performance tuning and scaling results for the T3E, and to staff at NCSA for helping study the performance of the code on the Origin 2000 and NT workstations. This work has been supported by AEI, NCSA, NSF grant No. PHY-96-00507, NASA HPCC/ESS Grand Challenge Applications Grant No. NCCS5-153, NSF MRAC Allocation Grant No.~MCA93S025, and Hong Kong RGC Grant CUHK 4189/97P.}
\section*{References}
\section{Introduction} \label{introsect} The appearance of macroscopic organized states, or coherent structures, in the midst of turbulent small-scale fluctuations is a common feature of many fluid and plasma systems \cite{Hasegawa}. Perhaps the most familiar example is the formation of large-scale vortex structures in a turbulent, large Reynolds' number, two-dimensional fluid \cite{McWilliams,SegreKida}. A similar phenomenon occurs in slightly dissipative magnetofluids in two and three dimensions, where coherent structures emerge in the form of magnetic islands with flow \cite{BW,KMT}. In the present work, we shall be concerned with another nonlinear partial differential equation whose solutions in the long-time limit tend to form large-scale coherent structures while simultaneously exhibiting intricate fluctuations on very fine spatial scales. Namely, we shall consider the class of one-dimensional wave systems governed by a nonlinear Schr\"odinger equation of the form \begin{equation} \label{nls1} i \psi_t + \psi_{xx} + f(|\psi|^2)\psi = 0\,, \end{equation} for various nonlinearities $f$ chosen so that the dynamics is nonintegrable and free of wave collapse. Numerical simulations of the NLS equation (\ref{nls1}) in a bounded spatial interval with either periodic or Dirichlet boundary conditions demonstrate that, after a sufficiently long time, the field $\psi$ evolves into a state consisting of a time-periodic coherent structure on the large spatial scales coupled with turbulent fluctuations, or radiation, on the small scales \cite{DZPSY,ZPSY,JJ}. For a focusing nonlinearity $f$ ($f(a) > 0, f'(a) > 0$ for $a > 0$) the coherent structure takes the form of a spatially localized, solitary wave. This organization into a soliton-like state is observed from generic initial conditions for the focusing NLS equation \cite{JJ}. At intermediate times a typical solution consists of a collection of the soliton-like structures, which as time evolves undergo a succession of collisions, or interactions. In these collisions the larger of the solitons increases in amplitude and smaller waves are radiated. This interaction of the solitons continues until eventually a single soliton of large amplitude survives in a sea of small-scale turbulent waves. (Even though (\ref{nls1}) does not possess solitons in the strict sense unless $f(|\psi|^2) = |\psi|^2$, it does have spatially localized solitary wave solutions in the generic focusing case. As a matter of convenience, and in keeping with the terminology in \cite{DZPSY,ZPSY}, we therefore simply refer to the coherent states that emerge in the simulations in this case as solitons). Figure (1) below illustrates this behavior for the particular saturated focusing nonlinearity $f(|\psi|^2) = |\psi|^2/(1 + |\psi|^2 ) $. For a defocusing $f$, coherent structures also emerge and persist, but they are less conspicuous since they are not spatially localized. The reader is referred to \cite{JJ} for a detailed description of numerical investigations of the long-time behavior of the system (\ref{nls1}) for various nonlinearities $f$. Theoretical arguments and numerical experiments suggest that these coherent structures for the NLS dynamics (\ref{nls1}) correspond to ground states of the system. That is, a coherent structure assumes the form $\psi(x,t) = \phi(x) \exp(-i\lambda t)$, where $\phi(x)$ is a solution of the ground state equation \begin{equation} \label{gse} \phi_{xx} + f(|\phi|^2)\phi + \lambda \phi = 0\,, \end{equation} In turn, these ground states $\phi$ are the minimizers of the Hamiltonian \begin{equation} \label{ham1} H(\psi)= \frac{1}{2} \int_{\Omega} |\psi_x|^2\,dx - \frac{1}{2} \int_{\Omega}F(|\psi|^2)\,dx \, \end{equation} subject to the constraint that the particle number \begin{equation} \label{part1} N(\psi) = \frac{1}{2} \int_{\Omega} |\psi|^2\,dx\ \end{equation} be equal to given initial value $N^0$. We shall refer to the first and second terms in (\ref{ham1}) as the {\em kinetic energy} and the {\em potential energy}, respectively. The potential $F$ is defined by $F(a)= \int_0^a f(a') da'$. The parameter $\lambda$ arising in (\ref{gse}) is the Lagrange multiplier in the variational principle: minimize $H(\phi)$ subject to $N(\phi) = N^0$. For focusing nonlinearities $f$, the ground states $\phi$ are spatially localized. By contrast, for periodic boundary conditions and for defocusing nonlinearities $f$, the ground states are spatially uniform; this can be proved by a straightforward application of Jenson's inequality when the potential $F$ is strictly concave. Nevertheless, independent of the properties of the ground states, the coherent structures are expected to be constrained minimizers of the Hamiltonian given the particle number, whenever these minimizers exist. Yankov and collaborators \cite{KY,DZPSY,ZPSY} have suggested that the tendency of the solution of the NLS system in the focusing case to approach a coherent soliton state coupled with small-scale radiation can be understood within a statistical mechanics framework. They argue that the solutions to the ground state equation that realize the minimum of the Hamiltonian subject to the constraint on the particle number (\ref{gse}) are ``statistical attractors'' to which the solutions of the NLS equation (\ref{nls1}) tend to relax. In this argument, the process which increases the amplitude of the solitons as the number of solitons decreases is thought to be ``thermodynamically advantageous'', in the sense that it increases the ``entropy'' of the system. This entropy seems to be directly related to the amount of kinetic energy contained in the radiation, or the small-scale fluctuations of the field $\psi$, so that the transference of kinetic energy to the fluctuations is accompanied by an increase in the entropy. Similar ideas have been expounded by Pomeau \cite{Pomeau} who has argued, based on weak turbulence theory, that the defocusing cubic NLS equation ($f(|\psi|^2) = - |\psi|^2$) in a bounded two--dimensional spatial domain should exhibit the tendency to approach a long-time state consisting of solution of the ground state equation (\ref{gse}) plus small-scale radiation. Our primary purpose in the present paper is to construct a statistical equilibrium model of the coherent structures and the turbulent fluctuations inherent in the long--time behavior of system (\ref{nls1}). With this model we seek to translate the various intuitive ideas about the balance between order and disorder in the NLS dynamics into explicit, verifiable calculations. In particular, we provide strong support to the notion set forth in \cite{KY,DZPSY,ZPSY,Pomeau} that the ground states, which minimize the energy for a given particle number, are statistically preferred states. In addition, we furnish a definite meaning to the concept that random fluctuations absorb the remainder of the energy. We derive our equilibrium statistical model from a mean--field approximation of the conserved Hamiltonian $H$ and particle number $N$ for a finite--dimensional truncation of the NLS equation (\ref{nls1}). This approximation relies on the fact that the fluctuations of $\psi$ about its mean, the coherent structure, become asymptotically small in the continuum limit, provided that the mean energy and mean particle number remain finite as the number of modes in the spectral truncations goes to infinity. In constructing the mean--field theory, therefore, we choose that probability measure on the finite--dimensional phase space which maximizes entropy subject to approximated constraints on the mean energy and particle number. Specifically, we approximate the means of the particle number and the potential energy term in the Hamiltonian in terms of the mean state $\langle \psi \rangle$. This approach gives rise to a Gaussian model, which is accordingly easy to analyze. Thus, we find that to any initial value $N^0$ of the particle number there corresponds a coherent structure, which is precisely the mean state, or most probable state, for the statistical model. Moreover, we show that this mean state is indeed a solution to the ground state equation that minimizes the Hamiltonian over all states with the same particle number $N^0$. Furthermore, we clarify the role of fine-scale fluctuations by demonstrating that the gradient field $\psi_x$ has finite variance at every point. Thus, we see that the difference between the initial energy $H^0$ and the energy of the mean state, $H(\langle \psi \rangle$), resides entirely in the kinetic energy of infinitesimally fine-scale fluctuations. This result, which makes precise the ideas of Yankov {\em et al.} \cite{DZPSY,ZPSY} and Pomeau \cite{Pomeau}, explains how it is possible for an ergodic solution to the Hamiltonian system (\ref{nls1}) to approach a solution to the ground state equation (\ref{gse}) without violating the conservation of energy or particle number. In the present paper the principal justification for our mean--field approximation is the evidence of extensive numerical simulations, which show that the fluctuations in $\psi$ go to zero over longer and longer time intervals \cite{DZPSY,ZPSY,JJ}. Nevertheless, we also include an {\em a posteriori} justification of this approximation by proving that, in the continuum limit, the probability measures defining the mean--field theory with approximately conserved quantities $H$ and $N$ concentrate about the phase space manifold on which $H=H^0$ and $N=N^0$. In other words, we establish a form of the equivalence of ensembles, showing that our mean-field ensemble becomes equivalent in a certain sense to the microcanonical ensemble, which is a measure concentrated on the manifold $H=H^0, N=N^0$. Besides connecting our mean-field approach to the standard principles of equilibrium statistical mechanics \cite{Balescu}, this result shows that the mean--field theory is asymptotically exact in the continuum limit. This property of the mean--field ensembles is a common feature of turbulent continuum systems in which the mean values of conserved quantities are fixed while the number of degrees of freedom goes to infinity \cite{MWC,JT}. These topics are taken up after a quick review of NLS theory and previous work on invariant measures for NLS. \section{The NLS Equation and its Conserved Quantities} \label{NLSsect} The NLS equation (\ref{nls1}) describes the slowly-varying envelope of a wave train in a dispersive conservative system and so arises in many branches of physics. It models, for example, gravity waves on deep water \cite{AS}, Langmuir waves in plasmas \cite{Pesceli}, pulse propagation along optical fibers \cite{HK}, and self-induced motion of vortex filaments \cite{Hasimoto,Majda}. In particular, the cubic NLS equation, which corresponds to the nonlinearity $f(|\psi|^2) = \pm |\psi|^2,$ has garnered much attention in the mathematics and physics literature. On the whole real line, or on a periodic interval, the cubic NLS equation is completely integrable via the inverse scattering transform \cite{LM,ZS}. The NLS equation with any other nonlinearity or boundary conditions, however, is not known to be integrable. In the present article we are interested in the NLS equation (\ref{nls1}) as a generic model of nonlinear wave turbulence. For this reason, we wish to consider nonlinearities $f$ and boundary conditions such that the dynamics is well-posed for all time and yet is nonintegrable. For instance, a natural class of prototypes of this kind consists of those NLS equations with bounded nonlinearities $f$ on a bounded spatial interval with Dirichlet boundary conditions. Bounded nonlinearities arising in physical applications are often called saturated nonlinearities, of which $f(|\psi|^2) = |\psi|^2/(1 + |\psi|^2)$ and $f(|\psi|^2) = 1-\exp(-|\psi|^2)$ are examples. These amount to corrections to the focusing cubic nonlinearity for large wave amplitudes, and they have been proposed as a models of nonlinear self-focusing of laser beams \cite{AB}, propagation of nonlinear optical waves through dielectric and metallic layered structures \cite{MNF}, and self-focusing of cylindrical light beams in plasmas due to the ponderomotive force \cite{max}. Such nonlinearities certainly satisfy the requirements set forth by Zhidkov to guarantee well--posedness of the NLS equation (\ref{nls1}) in the space $L^2$ for Dirichlet boundary conditions \cite{Zhidkov}. In light of these properties, we shall adopt the NLS equation with homogeneous Dirichlet boundary conditions on a bounded interval $\Omega = [0, L]$ and with a (focusing or defocusing) bounded nonlinearity $f$ as the generic system throughout our main development. Precisely, we impose on $f$ is the following {\em boundedness condition} \begin{equation} \label{boundcond} \sup_{a \in R} \left ( |f(a)| + |(1 + a) f'(a)| \right ) \, < \, \infty \, . \end{equation} In Section 9 we indicate how our main results can be extended to a natural class of unbounded nonlinearities. Also, when it is convenient for comparison with some numerical simulations, we can instead use periodic boundary conditions. The necessary modifications, which are straightforward, are left to the reader. The NLS equation (\ref{nls1}) may be cast in the Hamiltonian form \begin{displaymath} \frac{i}{2}\psi_t =\frac{\delta H}{\delta\psi^*} \,, \end{displaymath} where $\psi^*$ is the complex conjugate of the field $\psi$ and $H$ is the Hamiltonian defined by (\ref{ham1}). For our class of generic NLS systems, the total energy $H$ and the particle number $N$, or $L^2$ norm squared, are the only known dynamical invariants for (\ref{nls1}). These invariants play the leading role in the statistical equilibrium theory and the values $H^{0}$ and $N^{0}$, derived from a given initial state, say, alone determine the statistical ensemble. For later use, we write the NLS equation (\ref{nls1}) as the following coupled system of partial differential equations for the real and imaginary components $u(x,t)$ and $v(x,t)$ of the complex field $\psi(x,t)$: \begin{equation} \label{nls2} u_t + v_{xx} + f(u^2 + v^2)v = 0\,,\;\;\; v_t - u_{xx} - f(u^2+v^2)u=0\,. \end{equation} Similarly, the Hamiltonian can be written in terms of the fields $u$ and $v$ as \begin{equation} \label{ham2} H(u,v)= K(u,v) \, + \, \Theta (u,v) \end{equation} with kinetic and potential energy terms \[ K(u,v) = \frac{1}{2} \int_{\Omega} ( u_x^2 + v_x^2)\, dx\,, \;\;\;\;\;\;\;\; \Theta(u,v) =-\frac{1}{2} \int_{\Omega} F(u^2+v^2) \, dx\,, \] and the particle number is given by \begin{equation} \label{part2} N(u,v)= \frac{1}{2} \int_{\Omega} (u^2+v^2)\, dx\,. \end{equation} In the same way, the ground state equation (\ref{gse}) for $\phi = u + i v$ can be expressed as the following coupled system of differential equations: \begin{equation} \label{gse2} u_{xx} + f(u^2 + v^2)u + \lambda u = 0\,,\;\;\;\; v_{xx} + f(u^2 + v^2)v + \lambda v = 0\,. \end{equation} \section{Invariant Measures for NLS} \label{invsect} Several authors have constructed invariant measures for the NLS equation (\ref{nls1}). Zhidkov \cite{Zhidkov} considered nonlinearities $f$ satisfying the boundedness condition (\ref{boundcond}) and showed that the Gibbs measure \begin{equation} \label{gibbs} P_{\beta}(dudv) = Z^{-1}\exp(-\beta H(u,v)) \prod_{x \in \Omega} du(x) dv(x)\,, \end{equation} for fixed $\beta > 0$ is normalizable and has a rigorous interpretation as an invariant measure for the NLS system on the phase space $L^2$. Similar results have been obtained by Bidegary \cite{Bid}. These results hold for both periodic and Dirichlet boundary conditions. However, $P_{\beta}$ as constructed by Zhidkov fails to account for the invariance of the particle number (\ref{part1}) under the NLS dynamics. Earlier, Lebowitz {\em et al.} \cite{LRS} considered the problem of constructing invariant measures for NLS on the periodic interval with the focusing power law nonlinearities $f(|\psi|^2) = |\psi|^s$. For such nonlinearities the Hamiltonian $H$ is not bounded below, and therefore the measure $P_{\beta} $ defined by (\ref{gibbs}) is not normalizable. As an alternative, Lebowitz {\em et al.} conditioned the Gibbs measure on the particle number and studied the measures \begin{equation} \label{mgibbs} P_{\beta,N^0}(dudv) = Z^{-1} \exp(-\beta H(u,v)) I(N(u,v) \le N^0) \prod_{x \in \Omega} du(x) dv(x)\,, \end{equation} where $I(A)$ is the characteristic function of the set $A$. They showed that $P_{\beta,N^0}$ is normalizable for $1\le s < 4$ for any $N^0$, and for $s=4$ if $N^0$ is sufficiently small. For $s>4$, though, $P_{\beta,N^0}$ fails to be normalizable. The normalizability of the conditioned Gibbs ensemble (\ref{mgibbs}) is closely related to the well-posedness of the NLS equation as an initial value problem \cite{LRS}. Indeed, for the focusing power law nonlinearities it can be shown that the NLS equation (\ref{nls1}) with periodic or Dirichlet boundary conditions is well-posed, or free of wave collapse, in the Sobolev space $H^1(\Omega)$ as long as $s < 4$, while blow-up can occur from smooth initial conditions when $s \ge 4$ \cite{LRS,Bourgain,strauss}. Bourgain \cite{Bourgain} proved that under the conditions on $s$ and $N^0$ considered by Lebowitz {\em et al.} the measures $P_{\beta,N^0}$ are rigorously invariant under the NLS dynamics and he thereby obtained a new global existence theorem for the initial value problem with these nonlinearities. Analogous results have also been established by McKean \cite{McKean}. A common feature of all these invariant measures for the NLS equation is that for any fixed $\beta > 0$ the kinetic energy $(1/2) \int (u_x^2 + v_x^2)\,dx$ is infinite with probability $1$, and hence the Hamiltonian is also infinite with probability 1 \cite{LRS}. This difficulty occurs whenever classical Gibbs statistical mechanics is applied to a system with infinitely many degrees of freedom and is related to the well-known Jeans ultraviolet catastrophe \cite{Balescu}. Consequently, these ensembles are appropriate when the typical state of the continuum system is such that each of its modes contains a finite energy. They are not appropriate, on the other hand, to situations in which a typical state of the system is realized by an ergodic evolution from an ensemble of initial conditions having finite energy $H^0$ and particle number $N^0$, both of which partition amongst the many modes of the system. Since the purpose of the current paper is to construct ensembles which provide meaningful predictions about the long--time behavior of solutions of the NLS equation (\ref{nls1}), we are interested in statistical equilibrium ensembles which, in the continuum limit, have finite mean energy and mean particle number. In view of this goal, we scale the inverse temperature parameter $\beta$ in the Gibbs measure with the number of degrees of freedom $n$ and thereby maintain the mean energy at a finite value as $n \rightarrow \infty$. The continuum limit we obtain under this scaling is degenerate in certain ways, but this is hardly surprising given that the known invariant measures for NLS are supported on fields with infinite kinetic energy. On the other hand, the analysis of the scaled continuum limit is greatly simplified by the fact that the random fluctuations of $(u,v)$ at each point tend to zero. This fortuitous property implies that a natural mean--field approximation is asymptotically exact. Moreover, the mean-field ensemble concentrates on the microcanonical ensemble for the invariants $H$ and $N$ in the scaled continuum limit. For this reason, we can avoid the canonical ensemble constructed from $H$ and $N$, which fails to be normalizable for focusing nonlinearities because the linear combination $H - \lambda N$ always has a direction in which it goes to $-\infty$, as we shall demonstrate in Section 5. \vspace{1.5ex} \section{Discretization of NLS} \label{discretesect} We now introduce a finite--dimensional approximation of the NLS equation (\ref{nls1}) with homogeneous Dirichlet boundary conditions on the interval $\Omega = [0,L]$. For this purpose we shall use a standard spectral truncation, even though many other approximation schemes could be invoked. Let $e_k(x)= \sqrt{2/L} \sin (\sqrt{\lambda_k} x)$ and $\lambda_k= ( k \pi /L)^2, k=1,2,\ldots$, be the eigenfunctions and eigenvalues of the operator $- \frac{d^2}{dx^2}$ on $\Omega$ with homogeneous Dirichlet boundary conditions. The eigenfunctions $e_k$ form an orthonormal basis for the real space $L^2 (\Omega) $. For given $n$, let $X_n = \mbox{span } \{e_1, \ldots, e_n \} \subset L^2$, define $\un$ and $v^{(n)}$ in $X_{n}$ by \begin{equation} \label{unvn} u^{(n)}(x,t) = \sum_{k=1}^n u_k(t) e_k(x) \,,\;\; v^{(n)}(x,t) = \sum_{k=1}^n v_k(t) e_k(x) \,, \end{equation} and set $\psi^{(n)}(x,t) = u^{(n)}(x,t) + i v^{(n)}(x,t)$. We consider the following evolution equation for $\psi^{(n)}$: \begin{equation} \label{spectrunc1} i \psi^{(n)}_t + \psi^{(n)}_{xx} + P^{(n)} \left ( f(|\psi^{(n)}|^2)\psi^{(n)} \right ) = 0\,, \end{equation} where $P^{(n)}$ denotes the orthogonal projection from $L^{2}(\Omega)$ onto $X_n$. Equation (\ref{spectrunc1}) is clearly a spectral truncation of the NLS equation (\ref{nls1}). It is equivalent to the following coupled system of ordinary differential equations for the Fourier coefficients $u_k$ and $v_k, k= 1,\ldots,n$: \begin{equation}\label{spectrunc2} \frac{du_k}{dt} - \lambda_k v_k + \int_{\Omega} f( (u^{(n)})^2 + (v^{(n)})^2) v^{(n)} e_{k}\,dx = 0\,, \end{equation} \begin{equation}\label{spectrunc3} \frac{dv_k}{dt} + \lambda_k u_k - \int_{\Omega} f( (u^{(n)})^2 + (v^{(n)})^2) u^{(n)} e_{k}\,dx = 0, \end{equation} where we have used the fact that $(e_k)_{xx}= - \lambda_k e_k$. Whether viewed as the evolution of $(\un,v^{(n)})$ on $X_{n}\times X_{n}$ or in terms of the Fourier coefficients $(u_{k},v_{k})$ on $R^{2n}$, it can be shown \cite{Zhidkov,Bid} that this finite--dimensional dynamical system has Hamiltonian structure, with Hamiltonian $H_n=K_n+\Theta_n$, where \begin{equation}\label{kn} K_n(\un,v^{(n)}) = \frac{1}{2} \int_{\Omega} [(u^{(n)}_{x})^{2} + (v^{(n)}_{x})^{2}] \,dx = \frac{1}{2} \sum_{k=1}^n \lambda_k (u_k^2 + v_k^2)\,, \end{equation} is the kinetic energy, and \begin{equation} \label{tn} \Theta_n(\un,v^{(n)}) = -\frac{1}{2}\int_{\Omega} F((u^{(n)})^2 + (v^{(n)})^2)\, dx\,, \end{equation} is the potential energy. The functionals $H_n, K_n$ and $\Theta_n$ are just the restrictions to $X_n \times X_n$ of the functionals $H, K$ and $\Theta$, which are defined by (\ref{ham2}). The Hamiltonian $H_n$ is, of course, an invariant of the dynamics. The particle number \begin{equation}\label{dispn} N_n(\un,v^{(n)}) = \frac{1}{2} \int_{\Omega} [(u^{(n)})^{2} + (v^{(n)})^{2}] \,dx = \frac{1}{2} \sum_{k=1}^n (u_k^2 + v_k^2) \end{equation} is also conserved by this dynamics, as may be verified by direct calculation. Also, $N_n$ is the restriction of the functional $N$, which is defined by (\ref{part2}), to the space $X_n \times X_n$. The Hamiltonian system (\ref{spectrunc2})-(\ref{spectrunc3}) as a dynamical system on $R^{2n}$ satisfies the Liouville property \cite{Bid}. In other words, the Lebesgue measure $\prod_{k=1}^n du_k dv_k$ on $R^{2n}$ is invariant under the phase flow for this dynamics. \section{Mean-Field Ensembles for NLS} \label{idealsect} We now proceed to construct a statistical model that describes the long--time behavior of the truncated NLS system (\ref{spectrunc2})-(\ref{spectrunc3}). According to the fundamental principles of equilibrium statistical mechanics, the Liouville property and the ergodicity of the dynamics provide the usual starting point for such a description, and the microcanonical ensemble is the appropriate statistical distribution (that is, probability measure on phase space) with which to calculate averages for an isolated Hamiltonian system \cite{Balescu,PB}. In the specific system under consideration, the natural canonical random variables are $u_{k}$ and $v_{k}$, which comprise the phase space $R^{2n}$. The random fields $\un(x) = \sum u_{k} e_{k}(x)$, $v^{(n)}(x) = \sum v_{k} e_{k}(x)$, and $\psi^{(n)}(x)=\un(x)+i v^{(n)}(x)$, as well as the functions $H_{n}=H(\un,v^{(n)})$ and $N_{n}=N(\un,v^{(n)})$, are determined by these canonical variables. The microcanonical ensemble for the truncated NLS system therefore takes the form \begin{equation} \label{microcan} P_{H^0,N^0}(dudv) \;=\; W^{-1} \delta (H_n-H^0) \, \delta (N_n-N^0) \, \prod_{k=1}^n du_k dv_k\, , \end{equation} where $W=W(H^0,N^0)$ is the structure function or normalizing factor. Under the ergodic hypothesis, expectations with respect to this ensemble equal long-time averages over the spectrally-truncated dynamics from initial conditions with prescribed, finite values $H^0$ and $N^0$ of the invariants $H_n$ and $N_n$. While the microcanonical ensemble is a well-defined invariant measure for the spectrally-truncated dynamics, it is a cumbersome to analyze, and the calculation of ensemble averages for finite $n$ is not feasible. In statistical mechanics the usual procedure to overcome this difficulty is to introduce another invariant measure, the canonical ensemble, and prove that in the limit as $n \rightarrow \infty$, it becomes equivalent to the microcanonical ensemble \cite{Balescu,PB}. For the NLS system the canonical Gibbs ensemble is \begin{equation} \label{can} P_{\beta,\lambda} (dudv) \;=\; Z^{-1} \exp(-\beta [H_{n} - \lambda N_{n}]) \, \prod_{k=1}^n du_k dv_k, \end{equation} where $\beta$ and $\lambda$ are chosen such that the averages of $H_n$ and $N_n$ with respect to $P_{\beta,\lambda}$ equal $H^0$ and $N^0$, respectively, and $Z = Z(\beta,\lambda)$ is the partition function or normalizing factor. Unfortunately, for general focusing nonlinearities $f$, even those satisfying the boundedness condition (\ref{boundcond}), this canonical measure is not normalizable. To see this degeneracy, let us consider a nonlinearity $f$ that is both positive and strictly increasing and an associated ground state $\phi$ that is a nontrivial solution to (\ref{gse}). Then, for any positive scale factor $\sigma$, a straightforward calculation yields the identity \begin{eqnarray*} (H - \lambda N) ( \sigma \phi) &=& \frac{\sigma^2}{2} \int_{\Omega}[ |\phi_x|^2 \, - \, \lambda | \phi|^2 ] \, dx \,-\, \frac{1}{2} \int_{\Omega} F( \sigma^2 |\phi|^2) \, dx \\ &=& - \frac{1}{2} \int_{\Omega} dx \, \int_0^{ \sigma^2 |\phi|^2} [ f(a) - f(|\phi|^2) ] \, da \; . \end{eqnarray*} It follows from this expression that, as $\sigma$ increases to infinity, the value of $H - \lambda N$ at $\sigma \phi$ tends to $- \infty$. The same reasoning applies to the discretized system. Consequently, the function $H_n - \lambda N_n$ in the Gibbs weight has a direction in which is goes to $- \infty$, resulting in the divergence of the partition function $Z$ in (\ref{can}). In essence, this divergence reflects the fact that such a ground state $\phi$ is a critical point, but not a minimizer, of $H - \lambda N$. Locally, the first variation $\delta H - \lambda \delta N$ vanishes at $\phi$, but the second variation $\delta^2 H - \lambda \delta^2 N$ is positive only on those variations $\delta \phi$ for which $\delta N = 0$. When $\delta \phi = (1 - \sigma) \phi$ for $\sigma$ near $1$, and thus $\delta N \ne 0$, the corresponding second variation of $H - \lambda N$ is negative. The above calculation shows that this local behavior extends to a global degeneracy for a wide class of nonlinearities. As Lebowitz {\it et al.} discuss in \cite{LRS}, this defect of the canonical ensemble is even more severe when the nonlinearity $f$ is unbounded. These considerations compel us to find another way to approximate the microcanonical ensemble and thus to derive a tractable statistical equilibrium model. First, we note that since the microcanonical measure (\ref{microcan}) depends on $\un$ and $v^{(n)}$ through $H_{n}$ and $N_{n}$ only, it is invariant with respect to multiplicative constants of the form $e^{i \theta}$. In view of this phase invariance, we can factor the random field $\psi^{(n)}$ in the form $e^{i \theta} \phi^{(n)}$, where $\theta$ is uniformly distributed on $[0,2\pi]$ and is independent of $\phi^{(n)}$. In what follows, we shall write $\phi^{(n)}=\un+iv^{(n)}$ with the phase normalization \[ \mbox{ arg } \, \int_{\Omega} \phi^{(n)}(x) \, dx \;=\;0\,, \] and consider the microcanonical distribution conditioned by this constraint. It is this conditional distribution that we seek to approximate with a mean-field approximation. In constructing the approximation, we can therefore describe distributions for $\phi^{(n)}$ that need not possess invariance under a change of phase, with the understanding that we can recover the random field $\psi^{(n)}$ by setting $\psi^{(n)}=e^{i \theta} \phi^{(n)}$. The key to the mean-field construction is supplied by direct numerical simulations of the evolving microstates $\psi$ governed by (\ref{nls1}) \cite{ZPSY,JJ}. Since the particle number and the Hamiltonian are well conserved in these simulations, we may think of these numerical experiments as realizations of the microcanonical ensemble. The simulations clearly demonstrate that the amplitude of fluctuations of the random field $\phi$ (the phase normalized field associated with $\psi$) at each point $x \in \Omega$ become small in the long-time limit, and that they contribute little to the conserved $L^2$ norm of $\phi$ \cite{ZPSY,JJ} (also, see Figure 1 below). Moreover, this effect becomes more apparent as the spatial resolution of the numerical simulations is improved \cite{JJ}. In particular, these fluctuations become negligible compared to the magnitude of the coherent structure that emerges in the mean field $\phi$, which in the focusing cases takes the form of a single, localized soliton. On the basis of these properties of the simulated microstates, we shall adopt the hypothesis that, in the limit of infinite resolution, the fluctuations of $\phi$ are infinitesimal. Precisely, we shall make the following {\em vanishing of fluctuations hypothesis}: \vspace*{.2in} \noindent With respect to the phase-invariance conditioned microcanonical ensemble on $2n$-dimensional phase-space, there holds \begin{equation} \label{hyp1} \lim_{n \rightarrow \infty} \int_{\Omega} [\mbox{Var}( \un(x)) + \mbox{Var}(v^{(n)}(x))] \, dx \;=\; \lim_{n \rightarrow \infty} \sum_{k=1}^n [ \mbox{Var}(u_{k}) + \mbox{Var}(v_{k})] \;=\, 0 \, , \end{equation} where Var denotes the variance of the indicated random variable with respect to that ensemble. The first equality follows from the identity $ \int_{\Omega} [ \mbox{Var}(\un(x)) + \mbox{Var}(v^{(n)}(x)) ] \, dx = \sum_{k=1}^n [ \mbox{Var}(u_k) + \mbox{Var}(v_k) ]$. \vspace{1.5ex} Our strategy is to use the vanishing of fluctuations hypothesis to build a mean-field ensemble that is equivalent in an appropriate sense to the microcanonical ensemble (\ref{microcan}) in the continuum limit as $n \rightarrow \infty$, but is easy to analyze for finite $n$. Rather than attempt to verify this hypothesis {\it a priori}, we shall show {\it a posteriori} that the resulting mean-field ensemble concentrates on the microcanonical manifold at fixed $H^0$ and $N^0$. The analysis of the fundamental microcanonical ensemble itself will be taken up in a subsequent publication. We note however that the numerical simulations of the NLS system which support the hypothesis (\ref{hyp1}) also strongly indicate that the fluctuations of the gradients $u^{(n)}_x, v^{(n)}_x$ are not negligible in the limit of infinite resolution. Figure 2 illustrates the typical long-time behavior of $|\psi_x|^2$ for the particular nonlinearity $f(|\psi|^2)= |\psi|^2/(1 + |\psi|^2)$, in which finite fluctuations in the gradient $\psi_x$ persist. This property of the microcanonical ensemble is compatible with the vanishing of fluctuations hypothesis, as can be seen from the straightforward identity \begin{equation} \label{gradfluct} \int_{\Omega} [\mbox{Var}(u^{(n)}_{x}(x)) + \mbox{Var}(v^{(n)}_{x}(x))] dx = \sum_{k=1}^n \lambda_{k} [ \mbox{Var}(u_{k}) + \mbox{Var}(v_{k})]\,. \end{equation} In fact, as we shall see, this quantity remains finite in the continuum limit according to our statistical theory, and it represents the portion of kinetic energy absorbed by infinitesimally fine-scale fluctuations. We now proceed to construct a mean-field ensemble for each $n$ based on approximations derived from the hypothesis (\ref{hyp1}). This ensemble is determined by a probability density $\rho^{(n)}$ on $R^{2n}$ with respect to Lebesgue measure $\prod_{k=1}^n du_k dv_k$. (Unlike the microcanonical ensembles, the mean-field ensembles are taken to be absolutely continuous with respect to Lebesgue measure for each $n$.) In order to define that $\rho^{(n)}$ which governs the mean-field theory, we appeal to standard statistical-mechanical and information-theoretic principles \cite{Jaynes,Balescu,PB} and choose $\rho^{(n)}$ so that it maximizes the Gibbs--Boltzmann entropy functional \begin{equation} \label{entropy} S(\rho) = -\int_{R^{2n}} \rho(u_1,\ldots, u_n, v_1, \ldots, v_n) \log \rho(u_1,\ldots, u_n, v_1, \ldots, v_n)\prod_{k=1}^n du_k dv_k\,, \end{equation} subject to some appropriate constraints. The entropy $S$ has the well-known interpretation as a measure of the number of {\em microstates} $(u, v)$ corresponding to the {\em macrostate} $\rho$ \cite{Balescu}. The form of the entropy as a functional of $\rho$ is dictated by the Liouville property of the dynamics (\ref{spectrunc2})--(\ref{spectrunc3}), which requires that the entropy be relative to the invariant measure $\prod_{k=1}^n du_k dv_k$. Alternatively, from the information theoretic point of view, maximizing the entropy $S$ amounts to finding the least biased distribution on $2n$-dimensional phase space compatible with the constraints and the uniform prior distribution $\prod_{k=1}^n du_k dv_k$ \cite{Jaynes}. The constraints imposed in the Maximum Entropy Principle (MEP) are the crucial ingredient in the determination of the ensemble $\rho^{(n)}$. As is well-known, the microcanonical ensemble is produced when they are taken to be the exact constraints $H_n=H^0$ and $N_n=N^0$; the canonical ensemble, which however is ill-defined in the focusing case, corresponds to average constraints $\langle H_n \rangle = H^0$ and $\langle N_n \rangle = N^0$. Throughout our discussion of (MEP) the angle brackets $\langle \cdot \rangle$ denote expectation with respect to the admissible density $\rho$ in (MEP). For our mean-field theory we seek constraints that are intermediate between those giving the microcanonical and the canonical ensembles. To this end we invoke the vanishing of fluctuations hypothesis (\ref{hyp1}) and impose the following constraints: \begin{equation}\label{mfc1} \frac{1}{2} \sum_{k=1}^n (\langle u_k \rangle^2 + \langle v_k \rangle^2) = N^0\,. \end{equation} \begin{equation}\label{mfc2} \frac{1}{2} \sum_{k=1}^n \lambda_k (\langle u_k^2 \rangle + \langle v_k^2 \rangle) -\frac{1}{2} \int_{\Omega} F(\langle u^{(n)} \rangle^2 + \langle v^{(n)} \rangle^2)\,dx = H^0 , \end{equation} which we shall refer to as the {\em mean--field constraints} for obvious reasons. We note that these constraints involve only the first and second moments of the fields $\un$ and $v^{(n)}$ with respect to $\rho$. We can motivate this choice of constraints in (MEP) by approximating the values of the conserved quantities $H_n$ and $N_n$ in terms of means with respect to the microcanonical ensemble. An immediate implication of (\ref{hyp1}) applied to the definition of $N_n$ given in (\ref{dispn}) is that \begin{eqnarray}\label{Napprox} N^0 &=& \frac{1}{2} \int_{\Omega} (\langle u^{(n)}(x) \rangle^{2} + \langle v^{(n)}(x) \rangle^{2}) \,dx \;+\; \frac{1}{2} \int_{\Omega} [\mbox{Var}(\un(x)) + \mbox{Var}(v^{(n)}(x))] \, dx \nonumber \\ &=& N_n(\langle u^{(n)} \rangle, \langle v^{(n)} \rangle) \;+\; o(1) \end{eqnarray} as $n \rightarrow \infty$. When we drop the error term in this approximation at finite $n$ and we replace expectation with respect to the microcanonical distribution by expectation with respect to an admissible density $\rho$, we obtain the mean-field constraint (\ref{mfc1}). Similarly, we can obtain the mean-field constraint (\ref{mfc2}) from the vanishing of fluctuations hypothesis (\ref{hyp1}) by analyzing $ H_n$ with respect to the microcanonical distribution for large $n$. The kinetic energy (\ref{kn}) in $ H^0 = \langle H_n \rangle$ is retained exactly as $\langle K_n \rangle$. On the other hand, the potential energy (\ref{tn}) is approximated by expanding $F$ about the mean $(\langle u^{(n)} \rangle, \langle v^{(n)} \rangle)$, which yields \begin{eqnarray} \label{tnexp} \Theta_n(\un,v^{(n)}) & = & -\frac{1}{2} \int_{\Omega} F(\langle u^{(n)} \rangle^2 + \langle v^{(n)} \rangle^2)\,dx \nonumber \\ &-& \int_{\Omega} f(\langle u^{(n)} \rangle^2 + \langle v^{(n)} \rangle^2)\left ( \langle u^{(n)} \rangle (u^{(n)} - \langle u^{(n)} \rangle ) + \langle v^{(n)} \rangle (v^{(n)} - \langle v^{(n)} \rangle ) \right )\, dx \nonumber \\ & - & \frac{1}{4} \int_{\Omega} \left ( \begin{array}{c} (u^{(n)} - \langle u^{(n)} \rangle) \\ (v^{(n)} - \langle v^{(n)} \rangle) \end{array} \right )^{T} J(\tilde{u}^{(n)}, \tilde{v}^{(n)}) \left ( \begin{array}{c} (u^{(n)} - \langle u^{(n)} \rangle) \\ (v^{(n)} - \langle v^{(n)} \rangle) \end{array} \right )\,dx\,, \end{eqnarray} where \begin{equation} \label{matrix} J(u,v) = \left ( \begin{array}{cc} 2f(u^2 + v^2) + 4u^2f'(u^2 + v^2) & 4uv f'(u^2 + v^2)\\ 4uv f'(u^2 + v^2) & 2f(u^2 + v^2) + 4v^2 f'(u^2 + v^2) \end{array} \right ) \end{equation} is the matrix of second partial derivatives of $F(u^2 + v^2)$ with respect to $u$ and $v$ and $(\tilde{u}^{(n)}, \tilde{v}^{(n)})$ lies between $(u^{(n)}, v^{(n)})$ and $(\langle u^{(n)} \rangle, \langle v^{(n)} \rangle)$. Because we are considering nonlinearities $f$ satisfying (\ref{boundcond}) each element of the matrix $J(\tilde{u}^{(n)}, \tilde{v}^{(n)})$ is bounded over the domain $\Omega$ independently of $n$. Therefore, taking the expectation on both sides of equation (\ref{tnexp}) and noting that the second term on the right hand side of this equation has zero mean, we obtain \begin{displaymath} \langle \Theta_n \rangle = -\frac{1}{2} \int_{\Omega} F(\langle u^{(n)} \rangle^2 + \langle v^{(n)} \rangle^2)\,dx + R_n\,, \end{displaymath} where the remainder $R_n$ satisfies \begin{eqnarray*} |R_n| & \le & C \left \langle \int_{\Omega} [(u^{(n)} - \langle u^{(n)} \rangle)^2 + (v^{(n)} -\langle v^{(n)} \rangle)^2 ]\,dx \right \rangle \\ & = & C \int_{\Omega} [\mbox{Var}(\un(x)) + \mbox{Var}(v^{(n)}(x))] \, dx \end{eqnarray*} for some constant $C$ independent of $n$. The vanishing of fluctuations hypothesis (\ref{hyp1}) thus implies that $R_n \rightarrow 0$ as $n \rightarrow \infty$, and hence that \begin{equation} \label{Happrox} H^0 \;=\; \frac{1}{2} \sum_{k=1}^n \lambda_k (\langle u_k^2 \rangle + \langle v_k^2 \rangle) -\frac{1}{2} \int_{\Omega} F(\langle u^{(n)} \rangle^2 + \langle v^{(n)} \rangle^2)\,dx \;+\; o(1). \end{equation} The mean-field constraint (\ref{mfc2}) results from neglecting the error term in this expression for finite $n$ and replacing microcanonical expectations by expectations with respect to $\rho$. In summary, the mean-field theory is defined by a probability density $\rho^{(n)}$ on $R^{2n}$ that maximizes the entropy $S$ given in (\ref{entropy}) subject to the constraints (\ref{mfc1}) and (\ref{mfc2}). These mean-field ensembles, which solve the governing (MEP), are analyzed in detail in the subsequent sections. Here we merely note that they enjoy some properties not shared by either the microcanonical or the canonical ensembles. First, unlike the microcanonical ensemble, the mean-field ensembles are analytically tractable because the densities $\rho^{(n)}$ are Gaussian. This desirable property is a simple consequence of the fact that the constraints (\ref{mfc1}) and (\ref{mfc2}) involve only the first and second moments of $\rho^{(n)}$. Second, in contrast to the canonical ensemble, the mean-field ensemble exists in both the focusing and defocusing cases. This crucial property depends on the fact that fluctuations in $N_n$ are suppressed in the mean-field ensemble. However, because of the approximations made in deriving the constraints (\ref{mfc1}) and (\ref{mfc2}), the mean-field ensemble $\rho^{(n)}$ is not an invariant measure for the truncated NLS dynamics (\ref{spectrunc2})-(\ref{spectrunc3}) at finite $n$, even with the random phase $e^{i \theta}$ included. Nevertheless, it becomes consistent with the fundamental microcanonical ensemble in the limit as $n \rightarrow \infty$, in the sense that $ \langle N_n \rangle \rightarrow N^0$, $\langle H_n \rangle \rightarrow H^0$, and the variances of $N_n$ and $H_n$ converge to 0 . These properties, which are proved in Section 8, imply that the ensembles $\rho^{(n)}$ concentrate about the phase space manifold $H_n=H^0, N_n=N^0$, and thereby justify the approximations made in deriving the mean-field theory. \section{Equilibrium States for the Mean-Field Theory} \label{approxsect} We now proceed to calculate and analyze the solutions $\rho^{(n)}$ to the maximum entropy principle (MEP). First, we calculate the density $\rho^{(n)}$ which maximizes entropy subject to the mean-field constraints (\ref{mfc1})-(\ref{mfc2}) by the method of Lagrange multipliers. If we denote the left-hand sides of (\ref{mfc1}) and (\ref{mfc2}) by $\tilde{N}_{n}$ and $\tilde{H}_{n}$, then we have \begin{equation}\label{lmr} \delta S = \mu^{(n)} \delta \tilde{N}_n + \beta^{(n)} \delta \tilde{H}_n\,, \end{equation} where $\delta$ denotes variation with respect to the density variable $\rho$ and $\mu^{(n)}$ and $\beta^{(n)}$ are the Lagrange multipliers to enforce the constraints (\ref{mfc1}) and (\ref{mfc2}) on $\tilde{N_n}$ and $\tilde{H_n}$. The variation of $S$ is \begin{equation}\label{delS} \delta S = - \int_{R^{2n}} \log (\rho) \delta \rho \prod_{k=1}^n du_k dv_k \,, \end{equation} for variations $\delta \rho$ satisfying $\int_{R^{2n}} \delta \rho \prod_{k=1}^n du_k dv_k =0$. Similarly, the variations of $\tilde{N}_n$ and $\tilde{H}_n$ are \begin{equation} \label{delN} \delta \tilde{N}_n = \sum_{k=1}^n \int_{R^{2n}} ( \langle u_k \rangle u_k + \langle v_k \rangle v_k)\, \delta \rho \prod_{k=1}^n du_k dv_k\,, \end{equation} and \begin{eqnarray} \label{delH} \lefteqn{ \delta \tilde{H}_n = \sum_{k=1}^n \frac{1}{2}\int_{R^{2n}} \lambda_k ( u_k^2 + v_k^2)\, \delta \rho \prod_{k=1}^n du_k dv_k } \nonumber \\ &-& \int_{R^{2n}} \left ( \int_{\Omega} f(\langle u^{(n)} \rangle^2 + \langle v^{(n)} \rangle^2) \left [\langle u^{(n)} \rangle \un + \langle v^{(n)} \rangle v^{(n)} \right ]\, dx \right) \delta \rho \prod_{k=1}^n du_k dv_k\, \end{eqnarray} using the relation $F' = f$. Equating the coefficients of $\delta \rho$ in (\ref{lmr}) and recalling that $\un = \sum_{k=1}^{n} u_{k} e_{k}$ and $v^{(n)} = \sum_{k=1}^{n} v_{k} e_{k}$, we discover after some algebraic manipulations that the entropy--maximizing density $\rho^{(n)}$ has the form \begin{equation}\label{rho1} \rho^{(n)}(u_1, \ldots, u_n, v_1, \ldots, v_n) = \prod_{k=1}^n \rho_k(u_k, v_k)\,, \end{equation} where, for $k=1,\ldots, n$, \begin{equation}\label{rho2} \rho_k(u_k,v_k) \,=\, \frac{\beta^{(n)} \lambda_k}{2 \pi} \exp \left \{ -\frac{\beta^{(n)} \lambda_k}{2} \left [ (u_{k} - \langle u_k \rangle )^2 + (v_{k} - \langle v_k \rangle )^2 \right ] \right\}\,, \end{equation} with \begin{equation}\label{rho3} \langle u_k \rangle = \frac{1}{\lambda_k} \int_{\Omega} f(\langle u^{(n)}(x) \rangle^2 + \langle v^{(n)}(x) \rangle^2) \langle u^{(n)}(x) \rangle e_k(x) \, dx -\frac{\mu^{(n)}}{\beta^{(n)} \lambda_k}\langle u_k \rangle\,, \end{equation} \begin{equation}\label{rho4} \langle v_k \rangle = \frac{1}{\lambda_k} \int_{\Omega} f(\langle u^{(n)}(x) \rangle^2 + \langle v^{(n)}(x) \rangle^2) \langle v^{(n)}(x) \rangle e_k(x) \, dx -\frac{\mu^{(n)}}{\beta^{(n)} \lambda_k}\langle v_k \rangle\,. \end{equation} We see immediately that $u_{1}, \ldots, u_{n}, v_{1}, \ldots, v_{n}$ are mutually independent Gaussian random variables with means satisfying (\ref{rho3})--(\ref{rho4}), and variances \begin{equation}\label{vark} \mbox{Var}(u_k) = \mbox{Var}(v_k) = \frac{1}{\beta^{(n)} \lambda_k}\,. \end{equation} The multiplier $\beta^{(n)}$ is therefore necessarily positive. The equations (\ref{rho3})--(\ref{rho4}) for the means $\langle u_k \rangle$ and $\langle v_k \rangle$ can be written as an equivalent complex equation for the mean field $\langle \phi^{(n)} \rangle = \langle u^{(n)}\rangle + i \langle v^{(n)} \rangle$: \begin{equation} \label{mfeqn} \langle \phi^{(n)} \rangle_{xx} \,+\, P^{(n)} \left (f(|\langle \phi^{(n)} \rangle |^2) \langle \phi^{(n)} \rangle \right ) \,+\, \lambda^{(n)} \langle \phi^{(n)}\rangle \, = \, 0 \, , \end{equation} where we introduce the real parameter $ \lambda^{(n)} = -\mu^{(n)}/\beta^{(n)}$. Recalling that $P^{(n)}$ is the projection onto the span $X_{n}$ of the first $n$ eigenfunctions $e_1, \ldots, e_n$, we recognize this equation as the spectral truncation of the ground state equation (\ref{gse}) for the continuous NLS system (\ref{nls1}). In other words, the mean field $\langle \phi^{(n)} \rangle$ at each finite $n$ is a critical point of the functional $ H_n - \lambda^{(n)} N_n = 0$, by virtue of (\ref{mfeqn}). Thus, we draw the important conclusion that the mean field predicted by the statistical equilibrium theory is a solution of the ground state equation for the NLS system. Since the maximum entropy distribution $\rho^{(n)}$ satisfies the mean--field Hamiltonian constraint (\ref{mfc2}), a direct calculation reveals that \begin{eqnarray} \label{hameqn1} H^0 & = & \frac{1}{2}\sum_{k=1}^n \lambda_k ( \mbox{Var }u_k + \mbox{Var }v_k ) + \frac{1}{2}\sum_{k=1}^n \lambda_k ( \langle u_k \rangle^2 + \langle v_k \rangle^2 ) - \frac{1}{2} \int_{\Omega} F( \langle u^{(n)}\rangle^2 + \langle v^{(n)}\rangle^2)\, dx \nonumber \\ & = & \frac{n}{\beta^{(n)}} + H_n ( \langle u^{(n)} \rangle, \langle v^{(n)} \rangle)\,, \end{eqnarray} where (\ref{vark}) is invoked to obtain the second equality. In this expression the term $n/\beta^{(n)}$ in represents the contribution of fluctuations to the energy, while $ H_n ( \langle u^{(n)} \rangle,\langle v^{(n)} \rangle)$ is the energy of the mean state, or coherent structure. We can rearrange (\ref{hameqn1}) to get the following expression for $\beta^{(n)}$ in terms of the number of modes $n$ and the energy of the mean state: \begin{equation} \label{betan} \beta^{(n)} = \frac{n}{H^0 - H_n ( \langle u^{(n)} \rangle,\langle v^{(n)} \rangle)}\,. \end{equation} Now we come to the following central result, which enables us to characterize completely the mean--field statistical ensembles $\rho^{(n)}$. \vspace{1.5ex} \noindent {\bf Theorem 1.} The mean field pair $(\langle u^{(n)} \rangle, \langle v^{(n)}\rangle)$ corresponding to any solution $\rho^{(n)}$ of the constrained variational principle (MEP) is an absolute minimizer of the Hamiltonian $H_n$ over all $(u^{(n)},v^{(n)}) \in X_n \times X_n$ which satisfy the particle number constraint $N_n (u^{(n)},v^{(n)}) = N^0$. \vspace{1.5ex} \noindent {\bf Proof:} To prove this assertion, we calculate the entropy of a solution $\rho^{(n)}$ of (MEP). Referring to equations (\ref{rho1})--(\ref{vark}), which define any solution of (MEP) we immediately obtain \begin{eqnarray*} S(\rho^{(n)}) & = & -\int_{R^{2n}} \rho^{(n)} \log \rho^{(n)}\prod_{k=1}^n du_k dv_k \\ & = & -\sum_{k=1}^n \int_{R^2} \rho_k(u_k, v_k) \log \rho_k(u_k,v_k)\, du_k dv_k \\ & = & -\sum_{k=1}^n \left ( \log \frac{\beta^{(n)} \lambda_k}{2 \pi} -1 \right ) \\ & = & C(n) + n \log\left( \frac{L^2}{\beta^{(n)}} \right)\,, \end{eqnarray*} where $C(n) = n - \sum_{k=1}^n \log ( k^2 \pi/2)$ depends only on the number of Fourier modes $n$. Here, we have used the identity $\lambda_k = (k \pi/L)^2$. From this calculation and equation (\ref{betan}), it follows that \begin{equation} \label{entmax} S(\rho^{(n)}) = C(n) + n \log \left ( \frac{ L^2\, [ H^0 - H_n ( \langle u^{(n)} \rangle,\langle v^{(n)} \rangle )] }{n} \right ) \,. \end{equation} Evidently, the entropy $ S(\rho^{(n)})$ is maximized if and only if the mean field pair $(\langle u^{(n)} \rangle, \langle v^{(n)} \rangle)$ corresponding to $\rho^{(n)}$ minimizes the Hamiltonian $H_n$ over all fields $(u^{(n)},v^{(n)}) \in X_n \times X_n$ that satisfy the constraint $N_n (u^{(n)} , v^{(n)}) = N^0$. That such minimizers exist is proved in the Appendix. \vspace{1.5ex} The equation (\ref{entmax}) has the interesting interpretation that, up to additive and multiplicative constants, the entropy of the mean-field equilibrium $\rho^{(n)}$ is the logarithm of the kinetic energy contained in the fluctuations about the mean state, which is the coherent structure. Because the potential energy is determined entirely by the mean, the difference between the prescribed total energy $H^0$ and the energy of the mean must be accounted for entirely by the contribution of the fluctuations to the kinetic energy. The same conclusion also follows from equation (\ref{hameqn1}). This result provides a new and definite interpretation to the notion set forth by Yankov {\em et al.} \cite{KY,DZPSY,ZPSY} and Pomeau \cite{Pomeau} that the ``entropy'' of the NLS system is directly related to the amount of kinetic energy contained in the small--scale fluctuations, or radiation, and that the dynamical tendency of solutions to approach a ground state that minimizes energy at a conserved particle number is ``thermodynamically advantageous.'' The parameter $\lambda^{(n)}$ is determined as a Lagrange multiplier for the mean-field variational problem: minimize $H_n$ subject to the constraint that $N_n=N^0$. The multiplier associated with any minimizer $ \langle \phi^{(n)} \rangle = \langle u^{(n)} \rangle + i \langle v^{(n)} \rangle$ is unique, even though the minimizer itself may be nonunique. Indeed, the family of solutions $e^{i \theta} \langle \phi^{(n)} \rangle$ for arbitrary (constant) phase shifts $\theta$ exists for every such minimizer. Moreover, the ground state equation is a nonlinear eigenvalue equation, whose solutions can bifurcate. These bifurcations play the role of phase transitions in the mean-field statistical theory. Some analysis of the solutions of the mean-field equation is provided in the Appendix. Once a solution pair $( \langle \phi^{(n)} \rangle ,\lambda^{(n)} )$ is found, with $H_n ( \langle \phi^{(n)} \rangle ) = H_n^*$ being the minimum value of $H_n$ allowed by the particle number constraint $N_n = N^0$, the ``inverse temperature'' $\beta^{(n)}$ is uniquely determined by \begin{equation} \label{betan2} \beta^{(n)} = \frac{n}{H^0 - H_n^*}\,. \end{equation} Indeed, this is a restatement of (\ref{betan}). Since $ H_n^*$ approaches a finite limit as $n \rightarrow \infty$, we find that $\beta^{(n)}$ scales linearly with the number of modes $n$. This asymptotic scaling of inverse temperature, $\beta^{(n)} \sim n \beta^*$, where $\beta^*$ is finite in the continuum limit, is what distinguishes our mean-field theory from the statistical equilibrium theories discussed in Section 3. In those theories the inverse temperature is fixed and the mean energy is allowed to go to infinity in the continuum limit. Our rescaling of inverse temperature with $n$ is necessary in order that the expectation of Hamiltonian with respect to $\rho^{(n)}$ remain finite in the limit as $n \rightarrow \infty$. The implications of this scaling of $\beta^{(n)}$ are most evident in the the particle number and kinetic energy spectral densities. Substituting the expression (\ref{betan2}) for $\beta^{(n)}$ into (\ref{vark}), we obtain the following formulas for the variances of the Fourier components $u_k$ and $v_k$: \begin{equation} \label{vark2} \mbox{Var }(u_k) = \mbox{Var }(v_k) = \frac{H^0 - H_n^*}{n \lambda_k}\,. \end{equation} From these equilibrium expressions we obtain a sharp form of the vanishing of fluctuations hypothesis (\ref{hyp1}), which we adopted to derive the mean-field ensembles; namely, \begin{equation} \label{partfluct} \frac{1}{2} \sum_{k=1}^n [ \mbox{Var}(u_k) + \mbox{Var}(v_k)] \;=\; \frac{H^0 - H_n^*}{n} \sum_{k=1}^n \frac{1}{\lambda_k} \;=\; O\left(\frac{1}{n}\right)\,\;\; \mbox{as } n \rightarrow \infty\, , \end{equation} using the fact that $\lambda_k^{-1} = O (k^{-2})$ in the estimate. Thus, we have the following prediction for the particle number spectral density \begin{equation} \label{partspectra} \frac{1}{2}\langle u_k^2 + v_k^2 \rangle = \frac{1}{2} ( \langle u_k \rangle ^2 + \langle v_k \rangle^2 ) + \frac{H^0 - H_n^*}{n \lambda_k}\,. \end{equation} Since the mean-field is a smooth solution of the ground state equation, its spectrum decays rapidly in $k$, and we have the approximation $\frac{1}{2}\langle u_k^2 + v_k^2 \rangle \approx (H^0 - H_n^*)/(n \lambda_k)$ for $k >>1/L$. In the same way, the kinetic energy spectral density is \begin{equation} \label{kinspectra} \frac{1}{2} \lambda_k(\langle u_k^2 + v_k^2 \rangle) = \frac{1}{2}\lambda_k ( \langle u_k \rangle ^2 + \langle v_k \rangle^2 ) + \frac{H^0 - H_n^*}{n}\,, \end{equation} This expression shows that, as may be anticipated from a statistical equilibrium ensemble, the contribution to the kinetic energy from the fluctuations is equipartitioned among the $n$ spectral modes. These predictions are tested against the results of direct numerical simulations in \cite{JJ}, where a good agreement with the mean-field theory is documented. \section{Continuum Limit of Random Fields} \label{continuumsect} In this section we investigate the limits as $n \rightarrow \infty$ of the random fields $u^{(n)}$, $v^{(n)}$, and their gradients. The random fields $u^{(n)}$ and $v^{(n)}$ will be seen to converge in a uniform way to deterministic limits $u^{*}$ and $v^{*}$, while the gradients $u^{*}_{x}$ and $v^{*}_{x}$ converge in a weaker sense to random fields with finite variance but no spatial coherence. A simple form of the convergence of $u^{(n)}$ can be seen from the following easy estimate: \begin{eqnarray} \label{varn} \mbox{ Var}(\un(x)) &=& \sum_{k=1}^n \mbox{Var}(u_k) \, (e_k(x))^2 \nonumber \\ &=& \frac{1}{\beta^{(n)}} \sum_{k=1}^n \frac{1}{\lambda_k} (e_k(x))^2 \nonumber \\ &=& O(n^{-1}), \end{eqnarray} and similarly for $v^{(n)}(x)$. Thus, the variance at each point goes to zero as $n \rightarrow \infty$. A more detailed analysis is complicated by the fact that the mean fields $(\langle u^{(n)} \rangle, \langle v^{(n)} \rangle )$ are known to converge only on appropriately selected subsequences. This result is the content of Theorem A1 proved in the Appendix. It states that such subsequences of mean fields converge in the $C^1$-norm to ground states $(u^*, v^*)$ that solve the continuum equation (\ref{gse}); these ground states minimize the Hamiltonian $H$ over all $(u,v) \in H^1_0(\Omega)$ that satisfy the particle number constraint $N(u,v) = N^0$. The possible nonuniqueness of solutions to the ground state equation necessitates the introduction of subsequences, since in general there may exist a {\em set} of limits $(u^*, v^*)$ for a specified constraint value $N^0$. Accordingly, we shall assume throughout this section that $n$ goes to infinity along a subsequence for which the limiting ground state $(u^*, v^*)$ exists, without relabeling $n$. The detailed properties of the continuum limit are expressed in the following theorem. \vspace{1.5ex} \noindent {\bf Theorem 2.} The (sub)sequence of random fields $(u^{(n)}, v^{(n)})$ whose distribution is the mean-field ensemble with density $\rho^{(n)}$ converges in distribution to a non--random field $(u^*, v^*)$ which minimizes the Hamiltonian $H$ given the particle number constraint $N=N^0$. In addition, the gradients $(u^{(n)}_x, v^{(n)}_x)$ converge in the sense of finite-dimensional distributions to a Gaussian random field $(U'(x), V'(x))$ on $\Omega=[0,L]$ with the following properties: $U'$ and $V'$ are statistically independent, the mean of $(U',V')$ is $(u^*_x, v^*_x)$, and $U'$ and $V'$ each have the covariance function \begin{equation} \label{Gcovar} \Gamma^*(x,y) = \frac{H^{0}-H^{*}}{L} \left\{ \begin{array}{cl} 0, & \mbox{if $x \ne y$} \\ 1, & \mbox{if $x = y$ and $0 < x < L$} \\ 2, & \mbox{if $x = y$ and $x= 0, L$} \end{array} \right. \end{equation} where $H^*$ is the minimum value of $H$ given $N=N^0$. \vspace{1.5ex} {\bf Proof.} Recall that for each $n$, the mean $(\langle u^{(n)} \rangle, \langle v^{(n)} \rangle)$ under $\rho^{(n)}$ is a solution to (\ref{mfeqn}) which minimizes $H_n$ over $X_n \times X_n$ subject to the constraint $N_n = N^0$. By Theorem A1, the (sub)sequence $(\langle u^{(n)} \rangle, \langle v^{(n)} \rangle)$ converges in $C^1(\Omega)$ to $(u^{*}, v^{*})$, where $(u^{*}, v^{*})$ is a solution of (\ref{gse2}) that minimizes $H$ subject to the constraint $N(u,v) = N^0$. Recalling (\ref{vark}) and the independence of the Fourier coefficients $u_{k}$, we may use the following representation of $u^{(n)}$: \begin{eqnarray} \label{gg} u^{(n)} = \sum_{k=1}^{n} u_{k} e_{k} & = & \sum_{k=1}^{n} \left [ \langle u_{k} \rangle + \frac{1}{\sqrt{\lambda_{k} \beta^{(n)}}} Z_{k} \right ] e_{k} \nonumber \\ & = & \langle u^{(n)} \rangle + \frac{1}{\sqrt{\beta^{(n)}}} \sum_{k=1}^{n} \frac{L}{\pi k} Z_{k} e_{k}, \end{eqnarray} where $Z_{1}, Z_{2}, \ldots$ are independent Gaussian random variables with mean 0 and variance 1. The sum in the last line is the $n^{\mbox{th}}$ partial sum of the Karhunen-Lo\`{e}ve expansion of the Brownian bridge on $\Omega$, which is known to converge to its limit $B(x), x \in \Omega$, uniformly with probability one as $n \rightarrow \infty$ (\cite{Adler}, Theorem 3.3.3). That is, for each function $h: C(\Omega) \rightarrow R$ which is continuous in the supremum norm, we have, with probability one, \begin{equation} h\left(\sum_{k=1}^{n} \frac{L}{\pi k} Z_{k} e_{k}\right) \rightarrow h(B) \;\; \mbox{as} \;\; n \rightarrow \infty. \end{equation} It is clear from this demonstration that for such $h$ we have \begin{equation} h(u^{(n)}) \rightarrow h(u^{*}) \;\; \mbox{as} \;\; n \rightarrow \infty \end{equation} with probability one, since along this (sub)sequence the means converge uniformly and $\beta^{(n)}$ goes to infinity. Restricting to bounded $h$ and taking averages, we have \begin{equation} \langle h(u^{(n)}) \rangle \rightarrow h(u^{*}) \;\; \mbox{as} \;\; n \rightarrow \infty, \end{equation} which is the definition of convergence in distribution of the random fields $u^{(n)}$ to $u^{*}$. The argument that $v^{(n)}$ converges to $v^{*}$ is identical. The analysis of the gradient fields $u^{(n)}_{x}$ and $v^{(n)}_{x}$ is simpler. First note that they are independent and differ only in their means, and consequently it is sufficient to consider $u^{(n)}_{x}$, say. From the expression \begin{equation} u^{(n)}_{x}(x) = \sum_{k=1}^{n} u_{k} (e_{k})_{x}(x) \, , \end{equation} it is evident that $u^{(n)}_{x}$ is a continuous Gaussian random field whose mean $\langle u^{(n)}_{x} \rangle$ converges uniformly to $u^{*}_{x}$ as $n \rightarrow \infty$. Its covariance is then calculated to be \begin{eqnarray} \label{covargradn1} \Gamma_n(x,y) & = & \frac{1}{\beta^{(n)}} \sum_{k=1}^n \frac{1}{\lambda_k} (e_k)_{x}(x)(e_k)_{x}(y)\nonumber \\ & = & \frac{1}{2 L \beta^{(n)} } \left [ D_n ( \pi (x+y)/L) + D_n ( \pi(x-y)/L ) -2 \right ]\,, \end{eqnarray} where \begin{displaymath} D_n(\theta) = \left\{ \begin{array}{cl} \frac{\sin (n + \frac{1}{2})\theta}{\sin\frac{1}{2}\theta}, & \mbox{for } \theta \neq 0, \pm 2\pi, \pm 4\pi, \ldots \\ 2n +1, & \mbox{for } \theta = 0, \pm 2\pi, \pm 4\pi, \ldots \end{array} \right. \end{displaymath} is the Dirichlet kernel \cite{CHQZ}. The asymptotic behavior $\beta^{(n)} \sim n/(H^0 - H^*)$ implied by (\ref{betan2}) combined with the properties of $D_n(\theta)$ then produce the desired result that, as $n \rightarrow \infty$, the covariance function $\Gamma_n(x,y)$ converges pointwise to the function $\Gamma^*(x,y)$ given in the statement of the theorem. Now let $U'$ be the Gaussian random field with mean $u^{*}_{x}$ and covariance $\Gamma^{*}$. For convergence of finite-dimensional distributions, it remains to show that for each $m$ and any $x_{1}, \ldots, x_{m}$ in $\Omega$, the joint distribution on $R^{m}$ of the vector $(u^{(n)}_{x}(x_{1}), \ldots, u^{(n)}_{x}(x_{m}))$ converges weakly to the joint distribution of $(U'(x_{1}), \ldots, U'(x_{m}))$ as $n \rightarrow \infty$. But this follows by the L\`{e}vy Continuity Theorem \cite{Billingsley}, since the characteristic function of the former vector converges to the characteristic function of the latter vector, simply because the means and covariances converge. The argument that $v^{(n)}_{x}$ converges to $V'$ is similar, as is the extension to the convergence of the pair $(u^{(n)}_{x}, v^{(n)}_{x})$. \vspace{1.5ex} The results of Theorem 2 may be paraphrased as follows: The mean-field distributions $\rho^{(n)}$ converge weakly as measures on $C(\Omega)$ (along a subsequence) to a Dirac mass concentrated at a minimizer of the Hamiltonian subject to the particle number constraint. In this limit, the variance of the field $\phi^{(n)}$ vanishes while the variance of its gradient $\phi^{(n)}_x$ is uniform over the interval $\Omega$, apart from the endpoints. This behavior is in good qualitative agreement with the long--time behavior observed in numerical simulations \cite{JJ}. Nonetheless, the continuum limit is degenerate in two ways. First, the limiting field $\phi^*$ is deterministic. Second, the limiting gradient field, whose mean is $\phi^*_x$ and covariance function is $\Gamma^*$, is so rough that it does not even take values in the space of measurable functions (that is, it does not have a measurable version) \cite{RevuzYor}. These unusual properties of the limiting distribution are a direct consequence of the scaling of the inverse temperature $\beta^{(n)}$ with $n$, which is required to maintain finite mean energy in the continuum limit. \section{The Concentration Property} \label{concpropsect} We shall establish in this section the important result that the mean-field ensembles $\rho^{(n)}$, which solve the maximum entropy principle (MEP), become equivalent in a certain sense to the microcanonical ensemble (\ref{microcan}) in the continuum limit $n \rightarrow \infty$. Specifically, we shall prove the following {\em concentration property}: \vspace{1.5ex} \noindent {\bf Theorem 3.} Let $\rho^{(n)}, n=1,2,\cdots$, be a sequence of solutions of the constrained variational principle (MEP). Then \begin{equation}\label{th3.1} \lim_{n \rightarrow \infty} \langle N_n \rangle = N^0\,,\;\;\; \lim_{n \rightarrow \infty} \mbox{Var}(N_n) = 0\,, \end{equation} and \begin{equation}\label{th3.2} \lim_{n \rightarrow \infty} \langle H_n \rangle = H^0\,,\;\;\; \lim_{n \rightarrow \infty} \mbox{Var}(H_n) = 0\,. \end{equation} \vspace{1.5ex} This theorem has the interpretation that the mean-field ensembles $\rho^{(n)}$ concentrate on the microcanonical constraint manifold $N_n= N^0, H_n=H^0$ in the continuum limit. As we have emphasized above, it is a well-accepted axiom of statistical mechanics that the microcanonical ensemble constitutes the appropriate statistical equilibrium description for an isolated ergodic system \cite{Balescu,PB}. Thus, this concentration property provides an important theoretical justification for our mean-field theory. In the proof of Theorem 3, we will make use of the following elementary facts concerning Gaussian random variables: If $W$ is a Gaussian random variable with mean $\mu$ and variance $\sigma^2$, then \begin{equation} \label{gaussfact} \langle ( W - \mu )^4 \rangle = 3 \sigma^4\,,\;\;\; \langle W^4 \rangle - \langle W^2 \rangle^2 = 2 \sigma^4 + 4 \sigma^2 \mu^2\,. \end{equation} We will also repeatedly make use of the result that $1/\beta^{(n)} = O(n^{-1})$, which follows from equation (\ref{betan2}). \vspace{1.5ex} \noindent {\bf Proof of Theorem 3: } The first conclusion in (\ref{th3.1}) is easy to establish. Indeed, using (\ref{mfc1}) and the calculation (\ref{partfluct}), we have \begin{eqnarray*} \langle N_n \rangle & = &\frac{1}{2} \sum_{k=1}^n (\langle u_k \rangle^2 + \langle v_k \rangle^2 ) + \sum_{k=1}^n [ \mbox{Var}(u_k) + \mbox{Var}(v_k) ] \\ & = & N^0 + O(n^{-1})\,,\; \mbox{as } n \rightarrow \infty\,. \end{eqnarray*} To verify the second assertion in (\ref{th3.1}), we calculate using (\ref{dispn}) \begin{eqnarray} \label{th3.3} \mbox{Var}(N_n) & = & \frac{1}{4} \left \langle \left ( \sum_{k=1}^n \left [ ( u_k^2 - \langle u_k^2 \rangle ) + ( v_k^2 - \langle v_k^2 \rangle ) \right ] \right )^2 \right \rangle \nonumber \\ & = & \frac{1}{4} \sum_{k=1}^n \left [ (\langle u_k^4 \rangle - \langle u_k^2 \rangle^2 ) + (\langle v_k^4 \rangle - \langle v_k^2 \rangle^2) \right ]\,, \end{eqnarray} where, to obtain the second equality in this calculation, we have made use of the mutual statistical independence of the random variables $u_1,\ldots,u_n, v_1, \ldots, v_n$. But, as $u_k$ and $v_k$ are Gaussian with means $\langle u_k \rangle$ and $\langle v_k \rangle$ and variances given by (\ref{vark2}), we have from (\ref{gaussfact}) that \begin{equation} \label{th3.4} \langle u_k^4 \rangle - \langle u_k^2 \rangle^2 = \frac{2}{(\beta^{(n)} \lambda_k)^2} + \frac{4 \langle u_k \rangle^2}{\beta^{(n)} \lambda_k}\,,\;\; \langle v_k^4 \rangle - \langle v_k^2 \rangle^2 = \frac{2}{(\beta^{(n)} \lambda_k)^2} + \frac{4 \langle v_k \rangle^2}{\beta^{(n)} \lambda_k}\,. \end{equation} Upon substituting (\ref{th3.4}) into (\ref{th3.3}), we obtain \begin{eqnarray*} \mbox{Var}(N_n) & =& \frac{1}{(\beta^{(n)})^2} \sum_{k=1}^n \frac{1}{\lambda_k^2} + \frac{1}{\beta^{(n)}}\sum_{k=1}^n \left [ \frac {\langle u_k \rangle^2 +\langle v_k \rangle^2} {\lambda_k} \right ] \\ & \rightarrow & 0\;\;\; \mbox{as }\; n \rightarrow \infty\,, \end{eqnarray*} since $\sum_{k=1}^n \lambda_k^{-2}$ converges as $n \rightarrow \infty$, and $\sum_{k=1}^n \lambda_k^{-1} ( \langle u_k \rangle^2 +\langle v_k \rangle^2 ) \le 2\lambda_1^{-1} N^0$ for all $n$. To establish the first part of (\ref{th3.2}), observe that from (\ref{mfc2}), there holds \begin{equation}\label{th3.4.5} \left | \langle H_n \rangle - H^0 \right | = \left | \langle \Theta_n(\un, v^{(n)}) \rangle - \Theta_n (\langle \un \rangle, \langle v^{(n)} \rangle) \right |\,, \end{equation} where the potential energy $\Theta_n$ is defined by (\ref{tn}). Mimicking the calculations following equation (\ref{Napprox}) and preceding equation (\ref{Happrox}), we find that \begin{equation}\label{th3.5} \left | \langle \Theta_n(\un, v^{(n)}) \rangle - \Theta_n (\langle \un \rangle, \langle v^{(n)} \rangle) \right | \le C \sum_{k=1}^n \left [ \mbox{Var}(u_k) + \mbox{Var}(v_k) \right ]\,, \end{equation} where $C$ is a constant independent of $n$. Thus, according to (\ref{th3.4.5}), (\ref{th3.5}) and (\ref{partfluct}), there holds \begin{displaymath} \left | \langle H_n \rangle - H^0 \right | = O(n^{-1})\,,\;\; \mbox{as }\, n \rightarrow \infty\,. \end{displaymath} Next, we have that \begin{equation} \label{th3.7} \mbox{Var}(H_n) = \mbox{Var}( K_n + \Theta_n) \le 2\; \mbox{Var}(K_n) + 2\; \mbox{Var}(\Theta_n)\,, \end{equation} where $K_n$ is the kinetic energy defined as in (\ref{kn}). Using the statistical independence properties noted above, we calculate \begin{eqnarray*} \mbox{Var}(K_n) & = & \frac{1}{4} \left \langle \left ( \sum_{k=1}^n \lambda_k \left [ (u_k^2 - \langle u_k^2 \rangle) + (v_k^2 - \langle v_k^2 \rangle) \right ] \right )^2 \right \rangle \\ & = & \frac{1}{4} \sum_{k=1}^n \lambda_k^2 \left [ (\langle u_k^4 \rangle - \langle u_k^2 \rangle^2) + (\langle v_k^4 \rangle - \langle v_k^2 \rangle^2) \right ]\,, \end{eqnarray*} and using (\ref{th3.4}), we arrive at \begin{displaymath} \mbox{Var}(K_n) =\frac{n}{(\beta^{(n)})^2} + \frac{1}{\beta^{(n)}} \sum_{k=1}^n \lambda_k (\langle u_k \rangle^2 + \langle v_k \rangle^2)\,. \end{displaymath} This last expression vanishes as $n \rightarrow \infty$, because the sum $\sum_{k=1}^n \lambda_k (\langle u_k \rangle^2 + \langle v_k \rangle^2)$, which represents twice the kinetic energy of the mean, is bounded independently of $n$. Finally, we show that the variance of the potential energy $\Theta_n$ tends to 0 in the continuum limit. Using the definition of $\Theta_n$ and the Cauchy--Schwarz inequality, we have \begin{eqnarray} \label{th3.8} \mbox{Var}(\Theta_n) & = & \frac{1}{4}\left \langle \left ( \int_{\Omega} \left [ F((\un)^2 + (v^{(n)})^2) - \left \langle F((\un)^2 + (v^{(n)})^2) \right \rangle \right ]\, dx \right )^2 \right \rangle \nonumber \\ & \le & \frac{L}{4} \int_{\Omega} \mbox{Var}\left ( F((\un)^2 + (v^{(n)})^2) \right ) \, dx\,. \end{eqnarray} Now the potential $F$ may be expanded about the mean $(\langle \un \rangle, \langle v^{(n)} \rangle)$ as \begin{eqnarray*} F((\un)^2 + (v^{(n)})^2)& = & F(\langle \un \rangle^2 + \langle v^{(n)}\rangle^2) \\ & + & 2 f(\langle u^{(n)} \rangle^2 + \langle v^{(n)} \rangle^2)\left [ \langle u^{(n)} \rangle (u^{(n)} - \langle u^{(n)} \rangle ) + \langle v^{(n)} \rangle (v^{(n)} - \langle v^{(n)} \rangle ) \right ] \\ &+& \frac{1}{2} \left ( \begin{array}{c} (u^{(n)} - \langle u^{(n)} \rangle) \\ (v^{(n)} - \langle v^{(n)} \rangle) \end{array} \right )^{T} J(\tilde{u}^{(n)}, \tilde{v}^{(n)}) \left ( \begin{array}{c} (u^{(n)} - \langle u^{(n)} \rangle) \\ (v^{(n)} - \langle v^{(n)} \rangle) \end{array} \right )\,, \end{eqnarray*} where the matrix $J$ is defined by (\ref{matrix}) and $(\tilde{u}^{(n)}, \tilde{v}^{(n)})$ lies between $(\un, v^{(n)})$ and $(\langle \un \rangle, \langle v^{(n)} \rangle)$. Using the boundedness condition (\ref{boundcond}) on $f$ and the statistical independence of $\un$ and $v^{(n)}$, we obtain after a straightforward analysis, that pointwise over $\Omega$ \begin{eqnarray} \label{th3.9} \mbox{Var}\left ( F((\un)^2 + (v^{(n)})^2) \right ) &\le & C \left [ \langle u^{(n)} \rangle^2\; \mbox{Var}(\un) + \langle v^{(n)} \rangle^2\; \mbox{Var}(v^{(n)}) \right ] \nonumber \\ & + & C \left [\langle ( u^{(n)} - \langle u^{(n)} \rangle )^4 \rangle + \langle ( v^{(n)} - \langle v^{(n)} \rangle )^4 \rangle \right ]\,. \end{eqnarray} Here, $C$ is a constant that is independent of both $n$ and $x \in \Omega$. As we have demonstrated in (\ref{varn}), \begin{equation} \label{th3.10} \mbox{Var}(\un (x)) = \mbox{Var}(v^{(n)}(x)) = \frac{1}{\beta^{(n)}} \sum_{k=1}^n \frac{e_k^2(x)}{\lambda_k}\,, \end{equation} and since $\un(x)$ and $v^{(n)}(x)$ are Gaussian variables, it follows from (\ref{th3.10}) and (\ref{gaussfact}) that \begin{equation}\label{th3.11} \left \langle ( u^{(n)}(x) - \langle u^{(n)}(x) \rangle )^4 \right \rangle = \left \langle ( v^{(n)}(x) - \langle v^{(n)}(x) \rangle )^4 \right \rangle = \frac{3}{(\beta^{(n)})^2} \left ( \sum_{k=1}^n \frac{e_k^2(x)}{\lambda_k} \right )^2\,. \end{equation} Thus, from (\ref{th3.8})--(\ref{th3.11}), the fact that the eigenfunctions $e_k$ are uniformly bounded over $\Omega$ independently of $k$, and the fact that $\int_{\Omega} ( \langle u^{(n)} \rangle^2 + \langle v^{(n)} \rangle^2 )\, dx = 2N^0$ for all $n$, we have \begin{displaymath} \mbox{Var}(\Theta_n) = O(n^{-1})\,,\;\; \mbox{as }\; n \rightarrow \infty. \end{displaymath} This completes the proof of Theorem 3. \section{Extension to Unbounded Nonlinearities} \label{unboundedsect} \vspace{1.5ex} In this section, we briefly indicate how our mean-field statistical theory, can be extended to a class of unbounded nonlinearities, which includes the focusing power law nonlinearities $f(|\psi|^2) = |\psi|^s, 0 < s < 4$. As illustrated in \cite{JJ,DZPSY,ZPSY}, numerical simulations for such power law nonlinearities exhibit the same phenomena as seen for bounded nonlinearities such as $f(|\psi|^2) = |\psi|^2/(1+|\psi|^2)$, for which the results of numerical simulations are displayed in Figures (1)-(2). That is, the field $\psi$ approaches a long-time state consisting of a coherent soliton structure coupled with radiation or fluctuations of very small amplitude. As in the case of the bounded nonlinearities, the gradient $\psi_x$ exhibits fluctuations of nonnegligible amplitude for all time. In fact, we expect, that this general behavior occurs as long as the nonlinearity $f$ is such that the NLS equation (\ref{nls1}) is nonintegrable and free of collapse (i.e., such that finite time singularity does not occur). Hence, we would like to apply our statistical theory to NLS with such nonlinearities, as well. Let us recall that to motivate the mean-field constraints (\ref{mfc1}) and (\ref{mfc2}) for nonlinearities $f$ satisfying the boundedness condition ({\ref{boundcond}), we invoked the vanishing of fluctuations hypothesis (\ref{hyp1}). However, for nonlinearities such that the potential $F(|\psi|^2)$ grows more rapidly than $C |\psi|^2$ as $|\psi| \rightarrow \infty$ (as it does for the power law nonlinearities), the hypothesis (\ref{hyp1}) is not sufficient to guarantee {\em a priori} that $\langle \Theta_n (\un,v^{(n)}) \rangle$ converges to $\Theta_n (\langle \un \rangle, \langle v^{(n)} \rangle)$ as $n \rightarrow \infty$. Thus the mean-field Hamiltonian constraint (\ref{mfc2}) can not be derived from the vanishing of fluctuations hypothesis (\ref{hyp1}) alone for such $f$. Note that by making a stronger vanishing of fluctuations hypothesis, we could have weakened the assumptions on $f$ and arrived at the same mean-field constraints (\ref{mfc1})-(\ref{mfc2}). In any case, we could simply impose these mean-field constraints and investigate the resulting maximum entropy ensembles $\rho^{(n)}$. If it can be shown that these ensembles exist and satisfy the concentration property expressed in Theorem 3, then we will consider this approach to be justified {\em a posteriori}. It is clear from Theorem 1 and Theorem 2 that, in order for our statistical theory to be well-defined, it is necessary that there exist minimizers (in $H^1_0(\Omega)$, say) of the Hamiltonian $H$ given the particle number constraint $N=N^0$. This places restrictions on the class of nonlinearities that we can consider. However for nonlinearities $f$ such that the potential $F$ satisfies \begin{equation} \label{potcond} \mbox{There exists } C > 0\;\; \mbox{such that } F(|\psi|^2) \le C (|\psi|^2 + |\psi|^q )\,,\; \mbox{for some }\; 2 \le q < 6\,, \end{equation} it may be shown that such minimizers exist for any $N^0$ \cite{CL}. The condition (\ref{potcond}) is also crucial in establishing the well-posedness of the NLS equation (\ref{nls1}) as an initial value problem in the Sobolev space $H^1_0(\Omega)$ \cite{strauss}. Clearly, the focusing power law nonlinearities $f(|\psi|^2) = |\psi|^s$ satisfy this condition when $0 < s < 4$. Thus, assuming that $F$ satisfies the growth condition (\ref{potcond}), we may investigate the ensembles $\rho^{(n)}$ which maximize entropy subject to the mean-field constraints (\ref{mfc1})-(\ref{mfc2}). It is easy to see that the analysis of Section 6 goes through without change. Thus, under the the maximum entropy ensemble $\rho^{(n)}$, the Fourier coefficients $u_1, \cdots\, u_n, v_1, \cdots\, v_n$ are mutually independent Gaussian variables, the mean-field minimizes the Hamiltonian given the particle number constraint $N_n = N^0$, and the variances of the $u_k$ and $v_k$ are given by (\ref{vark2}). It may be shown that the continuum limit of Section 7 and the important concentration property of Section 8 also hold, if we impose, in addition to (\ref{potcond}), a lower bound on the potential $F$ of the form \begin{equation} \label{potcond2} \mbox{There exists } c \ge 0\;\; \mbox{such that } -c |\psi|^r \le F(|\psi|^2)\,,\; \mbox{for some } r \ge 0\,. \end{equation} That is, Theorems 2 and 3 still hold for $F$ satisfying (\ref{potcond}) and (\ref{potcond2}), but the proofs are complicated by the unboundedness of $f$. Since our goal in this paper has been to emphasize conceptual issues rather than technical details, we shall not present the proofs here. However, we refer the interested reader to \cite{jtz2}, where some of the analysis has been carried out for nonlinearities satisfying (\ref{potcond}) and (\ref{potcond2}). \section*{Acknowledgments} The research of R. Jordan has been supported in part by an NSF Postdoctoral Research Fellowship and also by DOE through grants to the Center for Nonlinear Studies at Los Alamos National Laboratory. The research of B. Turkington is partially supported by the NSF through the grant NSF-9600060. R. Jordan thanks Christophe Josserand and Pieter Swart for valuable discussions and suggestions. \section*{Appendix} In this Appendix, we provide an analysis of the variational principle that defines the mean field corresponding to the maximum entropy ensemble $\rho^{(n)}$. In particular, we prove that for (smooth) nonlinearities $f$ satisfying the condition (\ref{boundcond}), solutions of this variational principle exist, and we investigate the convergence properties of these mean fields as $n \rightarrow \infty$. First, we consider the variational principle \begin{equation} \label{VP} H(\psi) \rightarrow \mbox{min}\,,\;\; \mbox{subject to } \psi \in A^0\,, \end{equation} where the admissible set $A^0$ is defined as \begin{equation} \label{A0} A^0 = \left \{ \psi :\Omega \rightarrow {\bf C}\; |\; \psi \in H^1_0(\Omega)\,,\; N(\psi) = N^0 \right \}\,. \end{equation} The Hamiltonian $H$ and the particle number $N$ are defined by equations (\ref{ham1}) and (\ref{part1}), respectively, and, as in the main text, $\Omega=[0,L]$. Recall also that the potential $F$ is defined as $F(a) = \int_0^a f(a')\, da'$. \vspace{1.5ex} \noindent {\bf Lemma A1.} There exists a solution $\psi \in H^1_0(\Omega)$ of the variational problem (\ref{VP}). \vspace{1.5ex} \noindent {\bf Proof:} As $f$ satisfies the boundedness condition (\ref{boundcond}), there exists a constant $K > 0$ such that $|F(|\psi|^2)| \le K |\psi|^2$. Thus, for all $\psi \in A^0$, there holds \begin{eqnarray} \label{a1} H(\psi) & = & \frac{1}{2} \int |\psi_x|^2\, dx - \frac{1}{2}\int F(|\psi|^2)\, dx \nonumber \\ & \ge & \frac{1}{2} \int |\psi_x|^2\, dx - \frac{K}{2}N^0\,. \end{eqnarray} Here, and in the remainder of this Appendix, all integrals are over the domain $\Omega$. It follows from (\ref{a1}) that $H$ is bounded below on $A^0$, so that we may choose a minimizing sequence $ (\psi_n)_{n \in {\bf N}} \subset A^0$ for $H$. The inequality (\ref{a1}) also implies that the sequence $\psi_n$ is uniformly bounded in $H^1_0(\Omega)$. Hence, by the Sobolev Imbedding Theorem \cite{Adams}, there exists a subsequence, still denoted by $\psi_n$, and a function $\psi$ such that $\psi_n$ converges weakly to $\psi$ in $H^1_0(\Omega)$ and $\psi_n$ converges in $C(\Omega)$ to $\psi$. Because of the convergence in $C(\Omega)$, we have that $\int F(|\psi|^2)\,dx = \lim_{n \rightarrow \infty} \int F(|\psi_n|^2)\,dx$. Also, since $\psi_n$ converges weakly to $\psi$ in $H^1_0(\Omega)$, there holds $\int |\psi_x|^2\, dx \le \liminf_{n \rightarrow \infty} \int |(\psi_n)_x|^2\,dx$. Putting these two results together, we obtain that \begin{equation} \label{a2} H(\psi) \le \liminf_{n \rightarrow \infty} H(\psi^{(n)}) = \inf_{\phi \in A^0} H(\phi)\,. \end{equation} But, as $\psi_n$ converges to $\psi$ in $C(\Omega)$, it follows that $N(\psi)=N^0$, so that $\psi \in A^0$. We conclude from (\ref{a2}), therefore, that $\psi$ is a solution of the variational problem (\ref{VP}). \vspace{1.5ex} \noindent {\bf Lemma A2.} Any solution $\psi$ of the variational problem (\ref{VP}) is a solution of the ground state equation (\ref{gse}) for some real $\lambda$. \vspace{1.5ex} \noindent {\bf Proof: } The conclusion of this lemma follows from an application of the Lagrange multiplier rule for constrained optimization. The parameter $\lambda$ in the ground state equation (\ref{gse}) is the Lagrange multiplier which enforces the constraint $N(\psi)=N^0$. \vspace{1.5ex} We have demonstrated in Section (\ref{approxsect}) that the mean fields $\langle \phi^{(n)} \rangle = \langle u^{(n)} \rangle + i \langle v^{(n)} \rangle $ corresponding to the maximum entropy ensembles $\rho^{(n)}$ are solutions of the following variational problem \begin{equation} \label{VPn} H(\psi) \rightarrow \mbox{min}\,,\;\; \mbox{subject to } \psi \in A_n^0\,, \end{equation} where the admissible set $A_n^0$ is defined as \begin{equation} \label{An0} A_n^0 = \left \{ \psi \in \Xi_n\; |\; N(\psi) = N^0 \right \}\,. \end{equation} Here, $\Xi_n$ is the set of all complex fields $\psi$ on $\Omega$ having the form $\psi = u + iv$, where the real fields $u$ and $v$ are elements of the $n$-dimensional space $X_n$ spanned by the first $n$ eigenfunctions $e_1, \cdots, e_n$ of the operator $-d^2/dx^2$ on $\Omega$ with homogeneous Dirichlet boundary conditions. $\Xi_n$ may be thought of as a closed subspace of the Sobolev space $H^1_0(\Omega)$. The sequence of variational problems (\ref{VPn}) corresponds to a Ritz--Galerkin scheme for approximating solutions of the variational principle (\ref{VP}). The proof of the next lemma follows from arguments analogous to those that were used to prove Lemmas A1 and A2. The conclusion about the smoothness of solutions of (\ref{VPn}) is obvious, because any such solution is necessarily a finite linear combination of the basis functions $e_k(x)=\sqrt{2/L} \sin(k \pi x/L)$. \vspace{1.5ex} \noindent {\bf Lemma A3.} For each $n$, there exist solutions of the variational problem (\ref{VPn}). Any such solution $\psi^{(n)}$ is a solution of the differential equation \begin{equation} \label{gsen} \psi^{(n)}_{xx} + P^{(n)} \left (f(|\psi^{(n)}|^2) \psi^{(n)} \right ) + \lambda^{(n)} \psi^{(n)} = 0\,, \end{equation} where $P^{(n)}$ is the projection from $H^1_0(\Omega)$ onto $\Xi_n$, and $\lambda^{(n)}$ is a Lagrange multiplier which enforces the constraint $N(\psi^{(n)})=N^0$. In addition, any solution $\psi^{(n)}$ of (\ref{VPn}) is in $C^{\infty}(\Omega)$. \vspace{1.5ex} We wish to investigate the convergence of solutions of (\ref{VPn}) to those of (\ref{VP}). It is not our objective to obtain the strongest possible convergence results. For our purposes, it is sufficient to prove that any sequence of solutions of (\ref{VPn}) has a subsequence which converges in $C^1(\Omega)$ to a solution of (\ref{VP}). Such a result is all that is needed for the analysis of the continuum limit of the mean field ensembles $\rho^{(n)}$ in Section (\ref{continuumsect}). In general, we do not know that solutions of the variational principle are unique, so that subsequences can not be avoided. We shall now prove the following theorem. \vspace{1.5ex} \noindent {\bf Theorem A1.} If $\psi^{(n)}, n=1,2,\cdots$ is a sequence of solutions of the variational problem (\ref{VPn}), then there exists a subsequence $\psi^{(n')}$ of $\psi^{(n)}$ which converges in $C^1({\Omega})$ as $n' \rightarrow \infty$ to a solution of the variational problem (\ref{VP}). \vspace{1.5ex} \noindent {\bf Proof:} Let $\psi^{(n)}$ be a sequence of solutions of (\ref{VPn}), and let $H_n^* = H(\psi^{(n)})$. Because $A_n^0 \subset A_{n+1}^0, n=1,2,\cdots$, and $A^0 = {\overline{\bigcup_{n=1}^\infty A_n^0}}$, we have that $H_n^* \ge H_{n+1}^*$, and $H_n^* \searrow H^*$ as $n \rightarrow \infty$, where $H^*$ is the minimum value of $H$ over $A^0$. Thus, $\psi^{(n)}$ is a minimizing sequence for the variational problem (\ref{VP}), and so following the proof of Lemma A1, there exists a subsequence $\psi^{(n')}$ and a solution $\psi$ of the variational problem (\ref{VP}) such that $\psi^{(n')}$ converges weakly in $H^1_0(\Omega)$ to $\psi$ and $\psi^{(n')}$ converges in $C(\Omega)$ to $\psi$. Now, as $\psi^{(n')}$ is a solution of the variational problem (\ref{VPn}), it satisfies equation (\ref{gsen}) for some Lagrange multiplier $\lambda^{(n')}$. We shall now show that the sequence $\lambda^{(n')}, n' \rightarrow \infty$ is bounded independently of $n'$. For this purpose, we multiply equation (\ref{gsen}) for $\psi^{(n')}$ by the complex conjugate ${\overline{\psi^{(n')}}}$ and integrate over $\Omega$ to obtain \begin{equation} \label{thA11} -\int |(\psi^{(n')})_x|^2\, dx + \int P^{(n')} ( f(|\psi^{(n')}|^2) \psi^{(n')}) {\overline{\psi^{(n')}}}\, dx + \lambda^{(n')} \int |\psi^{(n')}|^2\, dx = 0\,. \end{equation} It follows immediately from (\ref{thA11}) and the fact that $\int |\psi^{(n')}|^2\, dx = 2N^0$ that \begin{equation} \label{thA12} |\lambda^{(n')}| \le \frac{1}{2N^0} \left [ \int |(\psi^{(n')})_x|^2\, dx + \left |\int P^{(n')} ( f(|\psi^{(n')}|^2) \psi^{(n')}) {\overline{\psi^{(n')}}}\, dx \right | \right ]\,. \end{equation} Because $\psi^{(n')}$ converges weakly to $\psi$ in $H^1_0(\Omega)$, the first integral on the right hand side of (\ref{thA12}) is bounded independently of $n'$. The second integral on the right hand side of (\ref{thA12}) may be estimated as follows: \begin{eqnarray*} \left |\int P^{(n')} ( f(|\psi^{(n')}|^2) \psi^{(n')}) {\overline{\psi^{(n')}}}\, dx \right | &\le & \left ( \int |\psi^{(n')}|^2\, dx \right )^{\frac{1}{2}} \left ( \int |P^{(n')} ( f(|\psi^{(n')}|^2) \psi^{(n')})|^2\, dx \right)^{\frac{1}{2}} \\ & \le & \left ( \int |\psi^{(n')}|^2\, dx \right )^{\frac{1}{2}} \left ( \int (f(|\psi^{(n')}|^2))^2 |\psi^{(n')})|^2\, dx \right)^{\frac{1}{2}} \\ &\le & 2N^0 \sup_{x \in \Omega} |f(|\psi^{(n')}|^2)|\,. \end{eqnarray*} To obtain the first line of this display, we have used the Cauchy--Schwarz inequality. The second line follows from the fact that $\int |P^{(n)} \phi|^2\, dx \le \int |\phi|^2\, dx$ for all $\phi \in H^1_0(\Omega), n \in {\bf N}$, and to obtain the third line, we have once again used the identity $\int |\psi^{(n')}|^2\, dx =2N^0$. Now, owing to (\ref{boundcond}), $ \sup_{x \in \Omega} |f(|\psi^{(n')}|^2)|$ is bounded independently of $n'$, so the preceding calculation demonstrates that the second integral on the right hand side of (\ref{thA12}) can also be bounded independently of $n'$. This implies that $\lambda^{(n')}$ is uniformly bounded in $n'$. The strategy now is to use the uniform boundedness of the eigenvalues $\lambda^{(n')}$ to establish that the subsequence $\psi^{(n')}$ is uniformly bounded in the Sobolev space $H^2_0(\Omega)$. From this result and the Sobolev Imbedding Theorem \cite{Adams}, we may conclude that the $\psi^{(n')}$ converges in $C^1(\Omega)$ to the function $\psi$ as above, which is a solution of the variational problem (\ref{VP}). This is the desired conclusion. To prove that $\psi^{(n')}$ is, in fact, uniformly bounded in $H^2_0(\Omega)$, we multiply the equation (\ref{gsen}) for $\psi^{(n')}$ by the complex conjugate of $(\psi^{(n')})_{xx}$ and integrate over $\Omega$. This yields, after integrating by parts and using the homogeneous Dirichlet boundary conditions, \begin{equation} \label{thA13} \int |(\psi^{(n')})_{xx}|^2\, dx = \lambda^{(n')} \int |(\psi^{(n')})_x|^2\, dx + \int {\overline{\psi^{(n')}}}_x (P^{(n')} ( f(|\psi^{(n')}|^2) \psi^{(n')}))_x\, dx\,. \end{equation} Using the Cauchy--Schwarz inequality, and the inequality $\int |(P^{n}(\phi))_x|^2\, dx \le \int |\phi_x|^2\, dx$, which is valid for all $ n\in {\bf N}$ and all $\phi \in H^1_0(\Omega)$, we find that \begin{equation} \label{thA14} \left |\int {\overline{\psi^{(n')}}}_x (P^{(n')} ( f(|\psi^{(n')}|^2) \psi^{(n')}))_x\, dx \right | \le \left (\int |(\psi^{(n')})_x|^2\, dx \right )^{\frac{1}{2}} \left ( \int |(f(|\psi^{(n')}|^2) \psi^{(n')})_x|^2\, dx \right )^{\frac{1}{2}}\,. \end{equation} But, a straightforward calculation yields that, pointwise on $\Omega$, \begin{eqnarray} \label{thA15} |(f(|\psi^{(n')}|^2) \psi^{(n')})_x|^2 &\le& 2 (f(|\psi^{(n')}|^2))^2 |(\psi^{(n')})_x|^2 +8 (f'(|\psi^{(n')}|^2))^2|\psi^{(n')}|^4 |(\psi^{(n')})_x|^2 \nonumber \\ &\le& C |(\psi^{(n')})_x|^2\,, \end{eqnarray} for a constant $C$ independent of $n'$. The second line of this calculation follows from the boundedness condition (\ref{boundcond}) on $f$. Taken together, (\ref{thA13})-(\ref{thA15}) imply that \begin{equation}\label{thA16} \int |(\psi^{(n')})_{xx}|^2\, dx \le (C + |\lambda^{(n')}|) \int |(\psi^{(n')})_x|^2\, dx\,. \end{equation} As we have established above, $\lambda^{(n')}$ and $\int |(\psi^{(n')})_x|^2\, dx$ are bounded independently of $n'$, so that from (\ref{thA16}), we conclude that the subsequence $\psi^{(n')}$ is uniformly bounded in $H^2_0(\Omega)$. This concludes the proof of Theorem A1.
\section{Introduction} The discovery of a heavy top quark, with a mass of $m_t = 175 \pm 6$ GeV \cite{CDF}, which is far larger than that of all other quarks, opens up the possibility that the top quark may have properties very different from those of the other quarks. Observation of these properties might even signal new physics beyond the standard model. Several efforts in the past few years have gone into the investigation of the potential of different experiments to study possible new interactions of the top quark. In particular, possible anomalous couplings of the top quark to electroweak gauge bosons\footnote{References to the voluminous literature on this subject can be found, for example, in \cite{ttbargamma}.} and to gluons \cite{ttbarg, rizzog} have also been discussed. Top polarization is especially useful in such studies \cite{ps,hp,gh,yuan}, because with a mass around 175 GeV, the top quark decays before it can hadronize \cite{bigi}, and all spin information is preserved in the decay distributions. In this paper, we investigate the potential of $e^+e^-$ experiments at a future linear collider with centre-of-mass (c.m.) energies of 500~GeV or higher, to study anomalous chromomagnetic and chromoelectric dipole couplings of the top quark to gluons. So far, a considerable amount of earlier work on the topic of anomalous gluon couplings has concentrated on hadron colliders. But also high energy $e^+e^-$ experiments with sufficiently high luminosities would provide a relatively clean environment to probe the standard model for anomalous gluon couplings. While earlier efforts in the context of $e^+e^-$ colliders are mainly based on an analysis of the gluon distribution \cite{rizzog} in $e^+e^-\to t\overline{t} g$, we look at the possible information that could be obtained from studying the total cross section, and the polarization of $t$ and $\overline{t}$ separately. This has the advantage over $t$ and $\overline{t}$ spin correlations that, because the polarization of only one of $t$ and $\overline{t}$ is analyzed by means of a definite decay channel, the other is free to decay into any channel. This leads to much better statistics compared to the case when $t-\overline{t}$ spin correlations are considered, where definite $t$ and $\overline{t}$ channels have to be used as analyzers. We find that there are three independent quantities, viz., the cross section, and the CP-even and CP-odd linear combinations of the $t$ and $\overline{t}$ polarizations, which can be used to probe separately the CP-even chromomagnetic dipole moment coupling and the CP-odd chromoelectric dipole coupling. Of these, the CP-odd combination provides the most sensitive probe of the imaginary parts of the anomalous couplings. We obtain here the 90\% confidence level (C.L.) limits that would be possible at a typical $e^+e^-$ linear collider, with integrated luminosity 50 fb$^{-1}$. We have considered two possible centre-of-mass (c.m.) energies, viz., 500~GeV and 1000~GeV. We have also considered the effect of beam polarization on the sensitivity of the measurements. Our results on the constraints on the CP-violating chromoelectric dipole moments were presented in Ref.~\cite{RT} . We include in this paper also the limits that would be obtainable on CP-conserving chromomagnetic dipole moments, and combinations of couplings. The paper is organized as follows: In Section 2, we introduce the effective action for the general massive anomalous $q\bar{q}g$ vertex. The subsequent discussion presents the full framework necessary for obtaining the analytical extensions that modify the standard QCD one-loop predictions for quark-antiquark production. We then give explicit combinations of the total polarized cross sections which are sensitive to the chromomagnetic and chromoelectric dipole moments, respectively. Next, Section 3 focuses on the numerical estimates of these observables for various kinematical regions readily accessible at a future $e^+ e^-$ collider. The final Section 4 is the Conclusion. \section{Calculation of cross section and top polarization with anomalous couplings} An effective $t\overline{t} g$ vertex can be written in the form \begin{equation} \Gamma^a_{t\overline{t} g} = - g_s T^a \left[ \gamma^{\mu}\epsilon_{\mu} + \frac{i\mu} {m_t} \sigma^{\mu\nu}q_{\mu}\epsilon_{\nu} - \frac{d}{m_t} \sigma^{\mu\nu}\gamma_5 q_{\mu}\epsilon_{\nu} \right], \label{effL} \end{equation} where \begin{equation} T^a = \Frac{1}{2} \lambda^a; \end{equation} $\lambda^a$ being the $SU(3)$ Gell-Mann matrices, and $q_{\mu}$ and $\epsilon_{\mu}$ are respectively the momentum and polarization four-vectors of the gluon. This is the most general Lorentz- and colour-invariant trilinear coupling (additional quadrilinear terms are needed for local colour invariance, but we do not need them here). The $\mu$ and $d$ terms are the chromomagnetic and chromoelectric dipole terms, respectively. These dipole couplings $\mu$ and $d$ are in fact momentum-dependent form factors, and complex in general. They parameterize the effects arising at the loop level, which in principle could arise from both, interactions within the Standard Model (SM) and possible new interactions. If the measured values of the corresponding form factors deviate from the theoretical predictions of the conventional SM, this would indicate the presence of additional, new physics. Note that a possible experimental determination of non-vanishing $\mathop{\mbox{\rm Im}}\nolimits(\mu)$ [$\mathop{\mbox{\rm Im}}\nolimits(d\,)$] does not necessarily imply the existence of CP-conserving [CP-violating] new physics. We use Eq.~({\ref{effL}) to calculate the $t\overline{t}$ total cross section and the $t$ and $\overline{t}$ polarizations to order $\alpha_s\equiv g_s^2/(4\pi)$. The extra diagrams contributing to this order are shown in Fig.~1, where the large dots represent anomalous couplings. The anomalous couplings enter in the amplitudes for soft and collinear gluon emissions. The infrared divergences in the amplitudes for soft gluon emission and in the virtual gluon corrections cancel as in standard QCD, since the anomalous terms vanish in the infrared limit. Moreover, we do not include anomalous couplings in the loop diagrams. The latter are included merely to regulate the infrared divergences. We thus study how anomalous couplings at tree-level would modify standard QCD predictions at order $\alpha_s$. For heavy-quark production, the standard QCD one-loop corrections to the total cross section and the longitudinal spin polarization~\cite{kpt} were calculated in closed analytic form before. Those results have been extended here to the case when anomalous couplings are present. The total unpolarized production rate is given only in terms of the $VV$ and $AA$ parity-parity combinations for the Born contributions: \begin{equation} \sigma_{Born}\left(e^+ e^-\to\gamma,Z\to q\bar{q}\right) = \Frac{1}{2}v(3-v^2)\sigma^{VV}+v^3\sigma^{AA}, \end{equation} where the mass parameters are $v=\sqrt{1-\xi\,}$ and $\xi=4m_q^2/s$. The $O(\alpha_s)$ unpolarized case has the cross section \begin{equation} \sigma\left(e^+ e^-\to\gamma,Z\to q\bar{q}\right) = \Frac{1}{2}v(3-v^2)\sigma^{VV}c^{VV}+v^3\sigma^{AA}c^{AA}, \end{equation} where the $VV$ and $AA$ factors that multiply with the appropriate Born terms are given below. \begin{equation} c^{VV} = 1+{\alpha_s\over2\pi}C_{\scriptscriptstyle F}\left[\,\tilde{\Gamma}-v{\xi\over2+\xi}\ln\left( {1+v\over1-v}\right)-{4\over v}\I{2}-{\xi\over v}\II{3}+ {4\over v(2+\xi)}+{2-\xi\over v}\II{5}+\Delta_0^{VV}\,\right]. \label{cvv} \end{equation} In this equation, the contribution from the virtual gluon loop is denoted as \begin{eqnarray} \tilde{\Gamma} &=& \left[\,2-{1+v^2\over v}\ln\left({1+v\over1-v}\right)\right]\, \ln\left(\Frac{1}{4}\xi\right)+{1+v^2\over v}\left[\, \mbox{$\mbox{\rm Li}_2$}\left(-{2v\over1-v}\right)-\mbox{$\mbox{\rm Li}_2$}\left({2v\over1+v}\right)+ \pi^2\,\right] \nonumber\\ &&+3v\ln\left({1+v\over1-v}\right)-4. \end{eqnarray} Here, the $q\bar{q}g$ phase-space integrals are abbreviated by $\I{i}$, and $\II{i}$ specify the results after the (soft) IR divergences have canceled. The explicit analytical expressions for these phase-space integrals may be found in \cite{tbp}. The additional component stemming from the anomalous gluon bremsstrahlung is given by \begin{equation} \Delta_0^{VV} = {8\over(2+\xi)v}\left[\,\mathop{\mbox{\rm Re}}\nolimits(\mu)(\I{1}+\I{4})+{2\over\xi} \left(|\mu|^2+|d|^2\right)(\I{1}-2\I{8})\,\right]. \end{equation} For the $AA$ contribution we find: \begin{equation} c^{AA} = 1+{\alpha_s\over2\pi}C_{\scriptscriptstyle F}\left[\,\tilde{\Gamma}+2{\xi\over v}\ln\left( {1+v\over1-v}\right)+{\xi\over v^3}\I{1}-{4\over v}\I{2}- {\xi\over v}\II{3}+{2+\xi\over v^3}\I{4}-{2-\xi\over v}\II{5} +\Delta_0^{AA}\,\right], \label{caa} \end{equation} with the following anomalous part \begin{eqnarray} \Delta_0^{AA} &=& {2\over v^3}\left[\,\mathop{\mbox{\rm Re}}\nolimits(\mu)\Big\{-(4-\xi)\I{1}+(2+\xi)\I{4}\Big\} +\left(|\mu|^2+|d|^2\right)\left\{\left({4\over\xi}+\xi-6\right)\I{1}+ \xi\I{4} \right.\right.\nonumber\\&&\left.\left. -{4\over\xi}(2-\xi)\I{8}+{4\over\xi}\I{9}\right\}\,\right]. \end{eqnarray} The remaining $V\!A$ and $AV$ parts are identical and only contribute to the spin-dependent cross section. In the absence of anomalous couplings $(\mu=d=0)$, the cross section for longitudinally polarized quarks of helicity $\pm \Frac{1}{2}$ is given by \begin{equation} \sigma\left(e^+ e^-\to\gamma,Z\to q(\lambda_\pm)\,\bar{q}\right) = \Frac{1}{2}v(3-v^2)\sigma^{VV}c^{VV}+v^3\sigma^{AA}c^{AA} \pm v^2\sigma^{V\!A}_S c_\pm^{VA}. \end{equation} The multiplication factors $c^{ij}$ are expressed in terms of phase-space integrals of type $\S{i}$: \begin{equation} c^{V\!A,AV}_\pm = 1+{\alpha_s\over2\pi}C_{\scriptscriptstyle F}\left[\,\tilde{\Gamma}+{\xi\over v}\ln\left({1+v\over1-v} \right)+\Delta^{V\!A,AV}_{\mu=d=0}\,\right], \label{cva} \end{equation} where \begin{eqnarray} \Delta^{V\!A,AV}_{\mu=d=0} &=& \Frac{1}{2}\Big[\,(4-\xi)\S{1}-(4-5\xi)\S{2}-2(4-3\xi)\S{4}- \xi(1-\xi)(\SS{3}+\SS{5})+\xi(\S{6}-\S{7}) \nonumber\\ && -2\S{8}+(2-\xi)\S{9}+(6-\xi)\S{10}-2\S{11}+ 2(1-\xi)(2-\xi)\S{12}\,\Big]. \end{eqnarray} The full analytic forms of all relevant $S$ integrals are too lengthy to be exhibited here. Most of them are compiled in Ref.~\cite{tbp}, except for the four additional integrals $S_{14}$, $S_{15}$, $S_{16}$, and $S_{17}$, which are discussed in more detail in the appendix. Including spins for the quark or the antiquark introduces additional spin-flip terms in the $O(\alpha_s)$ $c$ factors given in Eqs.~(\ref{cvv}), (\ref{caa}) and (\ref{cva}). For longitudinal quark polarization we find \begin{equation} c^{ij}_\pm = \Frac{1}{2}\left[\,c^{ij}\pm{\alpha_s\over2\pi}C_{\scriptscriptstyle F}\Delta^{ij}_S\,\right]. \end{equation} The individual parity-parity combinations are \begin{eqnarray} \label{sf1} (2+\xi)v\,\Delta^{VV}_S &=& 8\mathop{\mbox{\rm Im}}\nolimits\left(\mu^*d\right)\Bigg[\, -2\left(1-{2\over\xi}\right)\S{1}+\left(1-{4\over\xi}\right)\S{8}-\S{9}- \S{10}+\S{11} \nonumber\\ &&\hskip2.4cm+2\left(1-{4\over\xi}\right)\S{13}+{4\over\xi}(\S{15}+\S{17})\, \Bigg] \nonumber\\ &&+\mathop{\mbox{\rm Im}}\nolimits(d)\Bigg[\,8(1-\xi)\S{1}-\Big\{8-3\xi(2-\xi)\Big\}\S{2}+ \Big\{8+\xi(2-3\xi)\Big\}\S{4} \nonumber\\ &&\hskip1.75cm-\xi(2+\xi)\S{6}-4\S{8}+4(1-\xi)\S{9}-4(3+\xi)\S{10}+4\S{11} \nonumber\\ &&\hskip1.75cm+\xi(2-\xi)\S{14} \,\Bigg], \\[.5cm] \label{sf2} v^3 \Delta^{AA}_S &=& 4\mathop{\mbox{\rm Im}}\nolimits\left(\mu^*d\right)\Bigg[\, {2\over\xi}(1-\xi)(2-\xi)\S{1}+\left(5-{4\over\xi}\right)\S{8}- (1-\xi)(\S{9}+\S{10})+\S{11} \nonumber\\ &&+2\left(3-{4\over\xi}\right)\S{13}- 2\left(1-{2\over\xi}\right)\S{15}-{4\over\xi}\S{16}- 2\left(1-{4\over\xi}\right)\S{17}\,\Bigg] \nonumber\\ &&+\mathop{\mbox{\rm Im}}\nolimits(d)\Bigg[\, 2\xi\S{1}-(1-\xi)(4-\xi)(\S{2}-\S{4})-\xi(1-\xi)\S{6}-(2-\xi)\S{8} \nonumber\\ &&+(2-3\xi)\S{9}-(6-5\xi)\S{10}+(2+\xi)(\S{11}+2\S{13})+\xi(1-\xi)\S{14} \Bigg],\\[.5cm] v^2 \Delta^{V\!A,AV}_S &=& \mathop{\mbox{\rm Re}}\nolimits(\mu)\Bigg[\, -\xi(\S{1}+\S{7})-2\S{8}+(2-\xi)(\S{9}+\S{10})-2\S{11}\,\Bigg] \nonumber\\ &&\pm2\,i\mathop{\mbox{\rm Im}}\nolimits(\mu)\Bigg[\, (2-\xi)\S{1}-\xi\S{10}-2\S{13}\,\Bigg] \\ &&+\left(|\mu|^2+|d|^2\right)\Bigg[\,-(4-\xi)\S{1}-\xi\S{7}+4\S{8}+\xi\S{9}+ (4-\xi)\S{10}-4\S{11}\,\Bigg]. \nonumber \end{eqnarray} Using charge conjugation in the final state, one can readily obtain the corresponding expressions for (longitudinal) antiquark polarization. In the following, we denote the antiquark results by an additional bar, {\it i.e.\/} $\Deltab{ij}_S$: \begin{eqnarray} \Deltab{VV}_S &=& \Delta^{VV}_S, \\ \Deltab{AA}_S &=& \Delta^{AA}_S, \end{eqnarray} where the following identities hold \begin{eqnarray} \Deltab{V\!A} &=& -\Delta^{V\!A}_S\ =\ \left(\Deltab{AV}_S\right)^*, \\ \hbox{with}\qquad \Deltab{AV}_S &=& -\Delta^{AV}_S. \end{eqnarray} Considering the above expressions, we can construct the following combinations of polarization asymmetries of $t$ and $\overline{t}$, \begin{equation} \Delta\sigma^{(+)} = \Frac{1}{2} \Big[\, \sigma(\uparrow ) - \sigma(\downarrow) - \overline{\sigma}(\uparrow) + \overline{\sigma}(\downarrow) \,\Big], \end{equation} \begin{equation} \Delta\sigma^{(-)} = \Frac{1}{2} \Big[\, \sigma(\uparrow) - \sigma(\downarrow) + \overline{\sigma}(\uparrow) - \overline{\sigma}(\downarrow) \,\Big], \end{equation} where $\sigma (\uparrow)$, $\overline{\sigma} (\uparrow)$ refer respectively to the cross sections for top and antitop with positive helicity, and $\sigma (\downarrow)$, $\overline{\sigma} (\downarrow)$ are the same quantities with negative helicity. Of these, $\Delta\sigma^{(+)}$ is CP even and $\Delta\sigma^{(-)}$ is CP odd. This is obvious from the fact that under C, $\sigma$ and $\overline{\sigma}$ get interchanged, while under P, the helicities of both $t$ and $\overline{t}$ get flipped. Consequently, $\sigma$ and $\Delta\sigma^{(+)}$, both nonzero in standard QCD, receive contributions from combinations of anomalous couplings which are CP even, viz., Im$(\mu)^2 + \vert d\vert ^2$ and Re$(\mu)$. On the other hand, $\Delta\sigma^{(-)}$ vanishes in standard QCD, and in the presence of anomalous couplings it depends only on the CP-odd variables Im$(\mu^*d)$ and Im$(d)$. That $\Delta\sigma^{(-)}$ depends on the imaginary parts rather than the real parts of a combination of couplings follows from the fact that it is even under naive time reversal T$_{\rm N}$, i.e., reversal of all momenta, without change in helicities, and without interchange of initial and final states (as would have been required by genuine time reversal). As a consequence, it is odd under CPT$_{\rm N}$, and imaginary parts of couplings have to appear in order to avoid conflict with the CPT theorem. \section{Numerical results} Figs.~2a and 2b show the dependence of $\Delta\sigma^{(+)}$ on $\sqrt{{\rm Im}(\mu)^2+\vert d\vert^2}$ and Re$(\mu)$ at $\sqrt{s}=500$ GeV and 1000 GeV, respectively. Figs.~3a and 3b depict $\Delta\sigma^{(-)}$ plotted against Im$(\mu^*d)$ and Im$(d)$. We would now like to see how sensitive experiments at a future linear collider would be to the anomalous quantities $\mu$ and $d$. Since these are determined through $\Delta\sigma^{(+)}$ and $\Delta\sigma^{(-)}$, we should know how well these latter quantities can be measured experimentally. The quantity $\sigma(\uparrow)- \sigma(\downarrow)$ is simply the polarization $P_t$ of the top quark in the production process, and it can be determined by looking at the angular distribution of the decay products of the top. For example, in the decay mode $t\rightarrow W^+b$, the angular distribution of the $b$ quark in the $t$ quark rest frame is \begin{equation} \frac{1}{\Gamma}\frac{d\Gamma}{d\Omega_{b}} = \frac{1}{4\pi} \Big(1+P_t\,\alpha\,\hat{p}_b\cdot\vec{s}\,\Big), \label{tpol} \end{equation} where $\vec{p}_b$ is the $b$-quark momentum and $\vec{s}$ is the top spin. This angular distribution can be used to determine $P_t$. In Eq.~(\ref{tpol}), the parameter $\alpha$ ($\vert\alpha\vert\leq 1$) is a constant known as the {\it analyzing power} for the decay channel. In this particular case, $\alpha$ is given by \begin{equation} \alpha=\frac{m_t^2-2m_W^2}{m_t^2+2m_W^2}\ . \end{equation} There is an analogous expression for the decay $\overline{t}\rightarrow W^-\overline{b}$. If the angular distribution of a lepton or a jet arising from $W$ decay is used to determine $P_t$, the corresponding analyzing power would be different. The efficiency with which the top or antitop polarization can be measured will depend not only on the analyzing power of the channel, but also on the detection efficiency of the observed particles, like the $b$ quark. While the tagging efficiency of the $b$ would be much better at $e^+e^-$ colliders than at hadron colliders, because of the lower hadronic activity, it would nevertheless depend on the degree to which backgrounds are understood and eliminated. A more detailed analysis in the context of specific experimental conditions would be needed to obtain the overall top polarization detection efficiency, which we parameterize as $\epsilon$ in what follows. We use our expressions to obtain simultaneous 90\% confidence level (CL) limits that could be obtained at a future linear collider with an integrated luminosity of 50 fb$^{-1}$. We do this by equating the magnitude of the difference between the values for a quantity with and without anomalous couplings to 2.15 times the statistical error expected. Thus, the limiting values of Im$(\mu)^2 + \vert d\vert ^2$ and Re$(\mu)$ for an integrated luminosity $L$ and a top detection efficiency of $\epsilon$ are obtained from \begin{equation} \epsilon\, L\,\vert\,\sigma(\mu,d)-\sigma_{\scriptscriptstyle SM}\,\vert = 2.15\;\sqrt{L\,\sigma_{\scriptscriptstyle SM}\,} \label{lims} \end{equation} and \begin{equation} \epsilon\, L\,\left\vert\,\Delta\sigma^{(+)}(\mu,d)-\Delta\sigma^{(+)}_{\scriptscriptstyle SM} \,\right\vert = 2.15 \,\sqrt{L\,\left\vert\,\sigma_{\scriptscriptstyle SM} (\uparrow)-\overline{\sigma}_{\scriptscriptstyle SM}(\uparrow)\,\right\vert\:}, \label{lim+} \end{equation} whereas the limiting values of Im$(\mu^* d)$ and Im$(d)$ are obtained from \begin{equation} \epsilon\, L\,\left\vert\,\Delta\sigma^{(-)}(\mu,d)-\Delta\sigma^{(-)}_{\scriptscriptstyle SM} \,\right\vert = 2.15 \,\sqrt{L\,\left\vert\,\sigma_{\scriptscriptstyle SM} (\uparrow)+\overline{\sigma}_{\scriptscriptstyle SM}(\uparrow)\,\right\vert\:}. \label{lim-} \end{equation} In the above expressions, the subscript ``SM'' denotes the value expected in the standard model, with $\mu=d=0$. We use $L=50$ fb$^{-1}$ and $\epsilon = 0.1$ in our numerical estimates. The value of $\epsilon$ depends on the details of the detector, as well as the kinematical cuts employed. The value we use is only representative, and it would be easy to obtain limits for any other value of $\epsilon$ by appropriate scaling, because of the simple dependence on $\epsilon$ in Eqs.~(\ref{lims}), (\ref{lim-}), and (\ref{lim+}). For the running of the strong coupling, we choose $\alpha_s^{(5)}(M_{\scriptscriptstyle Z})=0.118$ (with $M_{\scriptscriptstyle Z}=91.178$~GeV) in the modified minimal subtraction scheme and use the appropriate conditions to match for six active flavours\footnote{It is common to indicate the number of active flavours as superscript of $\alpha_s$. For practical purposes one usually selects bottom production as reference. Here, our choice for $\alpha_s^{(5)}(M_{\scriptscriptstyle Z})$ translates to $\alpha_s^{(6)}(M_t=172.1{\rm GeV})=0.10811$}. Eqs.~(\ref{lims}) and (\ref{lim+}) are used to obtain contours in the plane of $\sqrt{{\rm Im}(\mu)^2+\vert d\vert^2}$ and Re$(\mu)$. The contours from (\ref{lims}) are shown in Figs.~4a and 4b for $\sqrt{s}=500$~GeV and 1000~GeV, respectively. The contours obtained from (\ref{lim+}) are shown in Figs.~5a and 5b. Eq.~(\ref{lim-}) gives contours in the plane of Im$(\mu^*d)$ and Im$(d)$, which are displayed in Figs.~6a and 6b. For different $e^-$ longitudinal beam polarizations $P_-$, the corresponding contours are presented in Figs.~4--6. In Figs.~4 and 5, the allowed regions are the ones shown below the respective contours. In Figs.~6a,b, the allowed regions are bands lying between the upper and lower straight lines. In principle, the measurement of two independent quantities $\sigma$ and $\Delta\sigma^{(+)}$ could have given independent limits on Im$(\mu)^2+\vert d\vert^2$ and Re$(\mu)$. However, a comparison of Figs.~4 and 5 shows that this possibility is not realized. A superposition of Figs.~4a,b and 5a,b is shown in Fig.~7a. It can be seen that the allowed region coming from $\sigma$ lies entirely within the allowed region from $\Delta\sigma^{(+)}$, with no intersection between the two. Thus $\sigma$ is more sensitive to the anomalous CP-even couplings as compared to $\Delta\sigma^{(+)}$. Moreover, each contour for $\sqrt{s}=1000$~GeV lies within the corresponding contour for $\sqrt{s}=500$~GeV, showing a uniform increase in sensitivity with c.m.\ energy. The other conclusion that can be drawn from Figs.~4 and 5 is that a large left-handed polarization leads to increase in sensitivity. The best limits obtainable from $\sigma$ at $\sqrt{s}=500$ GeV are $-3\stackrel{<}{~} {\rm Re}(\mu) \stackrel{<}{~} 2$ for Im$(\mu)=0=d$, and $\sqrt{{\rm Im}(\mu)^2 + \vert d \vert ^2} \stackrel{<}{~} 2.25$ for Re$(\mu)= 0 $. These limits are improved by about a factor 2 in going to $\sqrt{s}=1000$~GeV for the same integrated luminosity. In the case of CP-odd combinations of the couplings Im$(\mu^*d)$ and Im$(d)$, there is only one measurable CP-odd quantity $\Delta\sigma^{(-)}$ at each c.m.\ energy, and therefore independent limits on the two combinations are not possible. Fig.~6a shows that for $\sqrt{s}=500$ GeV, the best limits which can be obtained are for $P=-1$, viz., $-3.6 < {\rm Im} (\mu^* d) < 3.6$ for ${\rm Im} (d) = 0$ and $-10 < {\rm Im} (d) < 10$ for ${\rm Im} (\mu^* d) = 0$. The corresponding limits for $\sqrt{s}=1000$ GeV, as seen from Fig.~6b, are $-0.4 < {\rm Im} (\mu^* d) < 0.4$ for ${\rm Im} (d) = 0$ and $-20 < {\rm Im} (d) < 20$ for ${\rm Im} (\mu^* d) = 0$. However, if a measurement of $\Delta\sigma^{(-)}$ is made at two c.m.\ energies, a relatively narrow allowed range can be obtained, allowing {\em independent} limits to be placed on both Im$(\mu^*d)$ and Im$(d)$. This is demonstrated in Fig.~7b. In fact, the improvement in the limit on Im$(\mu^*d)$ in going from $\sqrt{s}=500$~GeV to $\sqrt{s}=1000$~GeV is considerable. The possible limits are \begin{equation} -0.8 < {\rm Im} (\mu^* d) < 0.8, -11 < {\rm Im} (d) < 11. \end{equation} These limits may be compared with the limits obtainable from gluon jet energy distribution in $e^+e^- \to t\overline{t} g$ \cite{rizzog}. While our proposal for the CP-even case seems to fare worse, for the CP-odd case, our proposal can be competitive. It should however be emphasized that in the case of the CP-odd couplings, we are proposing the measurement of a genuinely CP-violating quantity, whereas the analysis in \cite{rizzog} is merely based on the energy spectrum resulting from both CP-odd and CP-even couplings. In case of $\Delta\sigma^{(-)}$, the dependence on $e^-$ beam polarization is rather mild. \section{Conclusions} It is worthwhile noting that we have used a rather conservative value of $\epsilon = 0.1$ for top detection and polarization analysis. A better efficiency would lead to an improvement in the limits, as would a higher luminosity. We have not considered the effect of initial-state radiation in this work. We have also ignored possible effects of collinear gluon emission from one of the decay products of $t$ or $\overline{t}$. A complete analysis should indeed incorporate these effects, as well as a study of $t$ and $\overline{t}$ decay distributions which can be used to measure the polarizations. However, we do not expect our conclusions to change drastically when these effects are taken into account. In summary, we have examined the capability of total cross section and single quark polarization in $e^+e^- \to t\overline{t}$ to measure or put limits on anomalous chromomagnetic and chromoelectric dipole couplings. While the total cross section measurements can give, for the luminosities assumed, limits of order 1 on the CP-even couplings, only the CP-violating combination of top and antitop polarizations is sensitive to anomalous couplings, and can yield a limit of the order of 1 on a CP-odd combination of anomalous couplings. \vskip.75cm {\bf Acknowledgments.} We are both grateful for the kind hospitality and stimulating scientific atmosphere we enjoyed at the Departament de F\'{\i} sica Te\`orica during the principal stages of this work. We thank Bar-Shalom Shaouly for pointing out the update of Figures~6. This work has been supported by the DGICYT under Grants Ns.\ PB95-1077 and SAB95-0175, as well as by the TMR network ERBFMRXCT960090 of the European Union (S.D.R.). M.M.T.\ acknowledges support by CICYT Grant AEN-96/1718 and the Max-Kade Foundation, New York, NY. \newpage \setcounter{equation}{0} \renewcommand\theequation{A.\arabic{equation}} \section*{Appendix: Additional Phase-Space Integrals} The complicated three-body phase-space for the process $e^+ e^-\to q(\uparrow)\bar{q}g$ is best solved analytically by using the kinematic variables $y=1-p_1\cdot q/q^2$ and $z=1-p_2\cdot q/q^2$, which are natural dimensionless parameters referring to the quark and antiquark energies (including the radiated gluon) in the centre-of-mass system. Then, the relevant phase-space integrands all result in simple rational functions containing polynomials in $y$ and $z$, and the corresponding integration boundary is described by the symmetric solution of a $(y,z)$-biquadratic form~\cite{kpt,tbp}. Apart from the detailed integral list of Ref.~\cite{tbp}, in this particular calculation four new integrals emerged. For completeness, we give here the following full analytical results: \begin{eqnarray} S_{14} &=& \int{dy\,dz\over\sqrt{\vphantom{\big|}(1-y)^2-\xi\,}\:}\kern.1cm {z\over y^2} \nonumber \\[.25cm] &=& {2\over\xi}-{2+\mbox{$\sqrt\xi$}\over2(2-\mbox{$\sqrt\xi$})}-\ln(2-\mbox{$\sqrt\xi$})+\Frac{1}{2} \ln\xi-\Frac{1}{2}, \\ \nonumber\\ S_{15} &=& \int{dy\,dz\over\sqrt{\vphantom{\big|}(1-y)^2-\xi\,}\:}\kern.1cm\;y^2 \nonumber \\[.25cm] &=& \Frac{1}{32}\xi^3\Big[\Frac{1}{2}\ln\xi-\ln(2-\mbox{$\sqrt\xi$})\Big]- \mbox{$\sqrt\xi$}\left(4+\Frac{1}{3}\xi+\Frac{1}{4}\xi^2\right) \nonumber\\ & & +\Frac{1}{8}\left(7-\Frac{1}{2}\xi\right)\xi+\Frac{1}{3}, \\ \nonumber\\ S_{16} &=& \int{dy\,dz\over\sqrt{\vphantom{\big|}(1-y)^2-\xi\,}\:}\kern.1cm\;y^2z \nonumber \\[.25cm] &=& -{\xi^4(4-\xi)\over512(2-\mbox{$\sqrt\xi$})^2}+ \Frac{1}{16}\left(\Frac{3}{16}\xi-1\right)\ln(2-\mbox{$\sqrt\xi$}) -\Frac{3}{8}\xi\left(1+\Frac{3}{8}\xi\right) \nonumber\\ & & +\Frac{1}{512}\xi^3\Big[4-\xi+(16-3\xi)\ln\xi\Big] +\Frac{1}{4}\left(\Frac{7}{3}- \Frac{1}{2}\xi+\Frac{1}{16}\xi^2\right)+\Frac{1}{24}, \\ \nonumber\\ S_{17} &=& \int{dy\,dz\over\sqrt{\vphantom{\big|}(1-y)^2-\xi\,}\:}\kern.1cm\;y\,z \nonumber \\[.25cm] &=& \Frac{1}{32}\xi^2(6-\xi)\Big[-\Frac{1}{2}\ln\xi+\ln(2-\mbox{$\sqrt\xi$})\Big] +\Frac{1}{128}(4-\xi){\xi^3\over(2-\mbox{$\sqrt\xi$})^2} \nonumber\\ & & +\Frac{1}{4}\xi^{3\over2}\left(\Frac{5}{3}-\Frac{1}{8}\xi\right) -\Frac{1}{2}\xi\left(1-\Frac{1}{64}\xi^2\right)+\Frac{1}{12}, \end{eqnarray} where the usual mass parameters are $v=\sqrt{1-\xi}$ and $\xi=4m_q^2/s$. Note that there are no soft divergences in these particular integrals. The collinear divergences contained in these integrals arise from the massless character of the fermion field, and are easily identified by observing the limit $\xi\to 0$. In our full analytical expressions for the polarized cross sections with $VV$ and $AA$ parity-parity combinations, integral $S_{14}$ is multiplied by the mass factor $\xi$, viz. Eqs.~(\ref{sf1}) and (\ref{sf2}). This produces additional finite contributions, which originate from a collinear helicity-flip with anomalous chromoelectric couplings, and are absent in a naive massless model. \newpage
\section{introduction} The majority of nuclear phenomena have been successfully described in relativistic mean-field theory using only hadronic degrees of freedom\cite{Walecka75,Serot86,Serot92}. However, due to the observations which revealed the medium modification of the internal structure of the baryons\cite{EMC}, it has become essential to explicitly incorporate the quark-gluon degrees of freedom while respecting the established model based on the hadronic degrees of freedom in nuclei. One of the first models put forward along these lines is the quark-meson coupling (QMC) model, proposed by Guichon\cite{Guichon}, which describes nuclear matter as a collection of non-overlapping MIT bags interacting through the self-consistent exchange of scalar $\sigma$ and vector $\omega$ mesons in the mean field approximation with the meson fields directly coupled to the quarks\cite{Blunden,Saito94}. The scalar $\sigma $ meson is supposed to simulate the exchange of correlated pairs of pions and may represent a very broad resonance observed in $\pi\pi$ scattering, while the vector $\omega$ meson is identified with the actual meson having a mass of about 780 MeV. In the chiral models, the scalar $\sigma$ and vetcor $\omega$ mean fields represent the $u,d$ quark condensates\cite{Papa98,Papa99}. The QMC model thus incorporates explicitly the quark degrees of freedom and this has nontrivial consequences\cite{Blunden,Saito94}. It also have been extended to study superheavy finite nuclei\cite{Tsushima,Saito97}. In the so-called modified quark meson coupling model (MQMC)\cite{Jin,Jin2}, it has been further suggested that including a medium-dependent bag parameter may be essential for the success of relativistic nuclear phenomenology. It was found that when the bag parameter is significantly reduced in the nuclear medium with respect to its free-space value, large cancelling isoscalar Lorentz scalar and vector potentials for the nucleon in nuclear matter emerge naturally. Such potentials are comparable to those suggested by relativistic nuclear phenomenology and finite density QCD sum rules\cite{Cohen}. The density-dependence of the bag parameter is introduced by coupling it to the scalar meson field\cite{Jin}. This coupling is motivated by invoking the nontopological soliton model for the nucleon\cite{soliton}. In this model a scalar soliton field provides the confinement of the quarks. This effect of the soliton field is, roughly speaking, mimicked by the introduction of the bag parameter in the Bag Model. When a nucleon soliton is placed in a nuclear environment, the scalar soliton field interacts with the scalar mean field. It is thus reasonable to couple the bag parameter to the scalar mean fields\cite{Jin,Jin2}. The QMC model was extended to finite temperatures to investigate the liquid-gas phase transition in nuclear matter\cite{Song}. Recently, the MQMC model has been also extended to finite temperature\cite{Panda,Jaqaman,Jaqamanb} and has been applied to the study of the properties of nuclear matter where it was found that the bag parameter decreases appreciably above a critical temperature $T_c\approx 200$ MeV indicating the onset of quark deconfinement\cite{Jaqaman}. The effect of glueball exchange as well as a realization of the broken scale invariance of quantum chromodynamics has also been investigated in the MQMC model through the introduction of a dilaton field\cite{Jaqamanb}. It was found that the introduction of the dilaton potential improves the shape of the saturation curve at T=0 and affects hot nuclear matter significantly\cite{Jaqamanb}. In the present work, we extend the MQMC model to hot hypernuclear matter by introducing the scalar $\zeta$ and vector $\phi$ mean fields that are coupled to the $s$-quark in addition to the $\sigma$ and $\omega$ fields which are coupled to the $u,d$-quarks. The $\zeta$ and $\phi$ fields are identified with the real mesons having masses $m_{\zeta}=975$ and $m_{\phi}=1020$ MeV respectively \cite{Papa98,Papa99,Schaffner}. Hypernuclear matter is considered to contain the octet $p,n,\Lambda,\Sigma^+,\Sigma^0,\Sigma^-,\Xi^0$ and $\Xi^-$ baryons\cite{Mueller98}. We introduce an ideal gas of kaons to keep zero net strangeness density $\rho_S=0$. We simplify the calculations by considering symmetric hypernuclear matter whereby the octet baryons reduce to 4 species: $2N, \Lambda, 3\Sigma, 2\Xi$. The outline of the paper is as follows. In section II, we present the MQMC model for hypernuclear matter at finite temperature, together with the details of the self-consistency conditions for the vector and scalar mean fields. In section III, we discuss our results and present our conclusions. \section{The Quark Meson Coupling Model for Hypernuclear Matter} The quark field $\psi_{q}(\vec{r},t)$ inside a bag of radius $R_i$ representing a baryon of species $i$ satisfies the Dirac equation \begin{eqnarray} \left[ i\gamma^{\mu}\partial_{\mu}- m_{q}^{0} +(g_{\sigma}^{q}\sigma -g_{\omega}^{q}\omega_{\mu}\gamma^{\mu})\delta_{qr} +(g_{\zeta}^{q}\zeta-g_{\phi}^{q}\phi_{\mu}\gamma^{\mu})\delta_{qs} \right]\psi_{q}(\vec{r},t)=0. \label{Dirac} \end{eqnarray} where $m_{q}^{0}$ is the current mass of the quarks of flavor $q$ and where $r$ refers to the up or down quarks and $s$ refers to the strange quark. The Kronecker deltas insure that the u, d quarks are coupled only to the $\sigma$ and $\omega$ fields while the s-quark is coupled only to the $\zeta$ and $\phi$ fields. In the mean field approximation all the meson fields are treated classically and, for nuclear matter in this approximation, these fields are translationally invariant. Moreover, because of rotational invariance, the space-like components of the vector fields vanish so that we have $\omega_{\mu}\gamma^{\mu}=<\omega_0>\gamma^{0}=\omega\gamma^{0}$ and $\phi_{\mu}\gamma^{\mu}=<\phi_0>\gamma^{0}=\phi\gamma^{0}$. The single-particle quark and antiquark energies in units of $R_i^{-1}$ for quark flavor $q$ are given by \begin{eqnarray} {\epsilon_q}^{n\kappa}_\pm=\Omega_q^{n\kappa} \pm \left( g_{\omega}^{q}\omega R_i\delta_{qr} + g_{\phi}^{q}\phi R_i\delta_{qs} \right) \label{EPN} \end{eqnarray} where \begin{eqnarray} \Omega_q^{n\kappa}=\sqrt{ {x^q_{n\kappa}}^{2} + R_i^{2}{m^{*}}^{2}_{q} } \label{Omegnk} \end{eqnarray} and \begin{eqnarray} m^{*}_{q}=m^{0}_{q}-g^{q}_{\sigma}\sigma\delta_{qr} -g^{q}_{\zeta}\zeta\delta_{qs}, \label{effmass} \end{eqnarray} are the effective quark kinetic energy and effective quark mass, respectively. The boundary condition for each quark of flavor $q$ at the bag surface is given by \begin{eqnarray} i\gamma\cdot \hat{n} \psi_{q}^{n\kappa}({x^q_{n\kappa}}) =\psi_{q}^{n\kappa}({x^q_{n\kappa}}), \label{roots} \end{eqnarray} which determines the quark momentum $x^q_{n\kappa}$ in the state characterized by specific values of $n$ and $\kappa$. For a given value of the bag radius $R_i$ for baryon species $i$ and the scalar fields $\sigma$ and $\zeta$, the quark momentum $x^q_{n\kappa}$ is determined by the boundary condition Eq. (\ref{roots}) which, for quarks of flavor $q$ in a spherical bag, reduces to $j_0({x^{q }_{n\kappa}})=\beta_{q} j_1({x^{q}_{n\kappa}})$, where \begin{eqnarray} \beta_{q}= \sqrt{ \frac{ {\Omega^{*}}_{q}^{n\kappa}(\sigma,\zeta) -R_i m^{*}_{q}(\sigma,\zeta) } {{\Omega^{*}}_{q}^{n\kappa}(\sigma,\zeta) +R_i m^{*}_{q}(\sigma,\zeta) }}. \label{beta} \end{eqnarray} The quark chemical potential $\mu_{q}$, assuming that there are three quarks in the baryon bag, is determined from $3 =\sum_{q}n_q$ where $n_q$ is the number of quarks of flavor q and is determined by\cite{Panda,Jaqaman,Jaqamanb} \begin{eqnarray} n_q ={\sum_{n\kappa}}\left[ \frac{1}{e^{({\epsilon_q}^{n\kappa}_{+}/R-\mu_{q})/T}+1} -\frac{1}{e^{({\epsilon_q}^{n\kappa}_{-}/R+\mu_{q})/T}+1} \right]. \label{nq} \end{eqnarray} The total energy from the quarks and antiquarks of each baryon of species $i$ is \begin{eqnarray} E^i_{tot}=\sum_{n\kappa q} n_{q}\frac{\Omega_{q}^{n\kappa}}{R}\left[ \frac{1}{e^{({\epsilon_{q}}^{n\kappa}_{+}/R_i-\mu_{q})/T}+1} +\frac{1}{e^{({\epsilon_{q}}^{n\kappa}_{-}/R_i+\mu_{q})/T}+1} \right]. \label{Etot} \end{eqnarray} The bag energy for baryon species $i$ is given by \begin{eqnarray} E^i_{bag}=E^i_{tot} - \frac{Z_i}{R_i} +\frac{4\pi}{3}R_i^{3} B_i(\sigma,\zeta). \label{Ebag} \end{eqnarray} where $B_i=B_i(\sigma,\zeta)$ is the bag parameter. In the simple QMC model, the bag parameter $B$ is taken as $B_{0}$ corresponding to its value for a free baryon. The medium effects are taken into account in the MQMC model \cite{Jin,Jin2} by coupling the bag parameter to the scalar meson fields. In the present work we generalize the coupling suggested in the latter references to the case of two scalar meson fields by using the following ansatz for the bag parameter \begin{eqnarray} B_i=B_{0}\exp\left[-\frac{4}{3} \frac{((n_{u}+n_{d})g^{B}_{\sigma}\sigma +n_{s}g^{B}_{\zeta}\zeta)}{ M_{i}}\right] \label{bagCon} \end{eqnarray} with $g^{B}_{\sigma}$ and $g^{B}_{\zeta}$ as additional parameters. The spurious center-of-mass momentum in the bag is subtracted to obtain the effective baryon mass\cite{Fleck} \begin{eqnarray} M^{*}_{i}=\sqrt{{E^i_{bag}}^2 - <{p^{2}_{cm}}>^i}, \label{MNSTAR} \end{eqnarray} where \begin{eqnarray} <{p^{2}_{cm}}>^i=\frac{<x^{2}>^{i}}{R_i^{2}} \label{PCM} \end{eqnarray} and \begin{eqnarray} <x^{2}>^{i}= \sum_{n\kappa q} n_{q}{x^{q}_{n\kappa}}^{2} \left[ \frac{1}{e^{({\epsilon_{q}}^{n\kappa}_{+}/R_i-\mu_{q})/T}+1} +\frac{1}{e^{({\epsilon_{q}}^{n\kappa}_{-}/R_i+\mu_{q})/T}+1} \right]. \label{xseq} \end{eqnarray} The bag radius $R_{i}$ for baryon species $i$ is obtained through the minimization of the baryon mass with respect to the bag radius\cite{Saito94,Jin,Panda,Jaqaman,Jaqamanb} \begin{eqnarray} \frac{\partial M^{*}_{i}}{\partial R_{i}}=0. \label{MNR} \end{eqnarray} The total energy density at finite temperature $T$ and at finite baryon density $\rho_{B}$ reads \begin{eqnarray} \varepsilon&=& \sum^{Baryons}_i \frac{\gamma_i}{(2\pi)^{3}} \int d^{3}k\sqrt{k^{2}+{M_{i}^{*}}^{2}}(f_{i}+\overline{f}_{i}) +\frac{1}{2}m^{2}_{\omega} \omega^2 +\frac{1}{2}m^{2}_{\phi} \phi^2 +\frac{1}{2}m^{2}_{\sigma}\sigma^{2} +\frac{1}{2}m^{2}_{\zeta}\zeta^{2} \nonumber \\ &+& \varepsilon^{id}_{K}, \label{Edensity} \end{eqnarray} where $\gamma_i$ is the spin-isospin degeneracy factor of baryon species $i$ and where the last term corresponds to the energy density of the $K$-mesons treated here as an ideal gas. In Eq. (\ref{Edensity}) $f_{i}$ and $\overline{f}_{i}$ are the Fermi-Dirac distribution functions for the baryons and antibaryons of species $i$, \begin{eqnarray} f_{i}=\frac{1}{e^{(\epsilon_i^{*}-\mu^{*}_{i})/T}+1}, \label{fB} \end{eqnarray} and \begin{eqnarray} \overline{f}_{i}=\frac{1}{e^{(\epsilon_i^{*}+\mu^{*}_{i})/T}+1}, \label{fBAR} \end{eqnarray} where $\epsilon_i^{*}$ and $\mu^{*}_{i}$ are, respectively, the effective energy and effective chemical potential of baryon species $i$. These are given by $\epsilon_i^{*}=\sqrt{ k^{2}+{M^{*}_{i}}^{2} }$ and $\mu^{*}_{i}= B_i\mu_B+S_i\mu_S-\left(g_{\omega i}\omega+g_{\phi i}\phi \right)$ where $B_i$ and $S_i$ are the baryon and strangeness quantum numbers and where $g_{\omega i}=(n_{u}+n_{d})g^{q}_{\omega i}$ and $g_{\phi i}=n_{s}g^{q}_{\phi i}$ . The chemical potentials $\mu_B, \mu_S$ are determined by the self-consistency equations for the total baryonic density \begin{eqnarray} \rho_{B}=\frac{1}{(2\pi)^{3}} \sum_i B_i \gamma_i \int d^{3}k(f_{i}-\overline{f}_{i}), \label{rhoB1} \end{eqnarray} and the total strangeness density \begin{eqnarray} \rho_{S}=\frac{1}{(2\pi)^{3}} \sum^{Baryons}_i S_i \gamma_i \int d^{3}k(f_{i}-\overline{f}_{i}) +\sum^{Kaons}_i S_i {\rho^{id}_{Ki}} = 0 \label{rhoS1} \end{eqnarray} where in the last equation we have introduced the contribution of an ideal gas of $K$ and $K^{*}$ mesons to make the total strangeness density vanish identically. The vector mean fields are determined by \begin{eqnarray} \omega=\sum_i \frac{g_{\omega i}}{m^{2}_{\omega}}B_i\rho_{i}, \label{vecomeg} \end{eqnarray} and \begin{eqnarray} \phi=\sum_i \frac{g_{\phi i}}{ m^{2}_{\phi}}B_i\rho_{i}. \label{vecphi} \end{eqnarray} The pressure is the negative of the grand thermodynamic potential density and is given by \begin{eqnarray} P&=& \frac{1}{3}\sum_i\frac{\gamma_i}{(2\pi)^{3}}\int d^{3} k \frac{k^{2}}{\epsilon_i^{*}}(f_{i}+\overline{f}_{i}) +\frac{1}{2}m^{2}_{\omega}\omega^{2} +\frac{1}{2}m^{2}_{\phi}\phi^{2} -\frac{1}{2}m^{2}_{\sigma}\sigma^{2} -\frac{1}{2}m^{2}_{\zeta}\zeta^{2} +P^{id}_{K}, \label{pressurD} \end{eqnarray} where the summation $i$ runs over the 8 species of the baryon octet which reduces for symmetric hypernuclear matter to 4 species with $\gamma_i=4,2,6$ and $4$ for $N$, $\Lambda$, $\Sigma$ and $\Xi$, respectively. In Eq. (\ref{pressurD}) $P^{id}_{K}$ is the pressure of the ideal gas of $K$ mesons. The density of the $K$-mesons of species $i$ is given by \begin{eqnarray} {\rho^{id}_{K}}_i=\frac{\gamma^K_i}{(2\pi)^{3}} \int d^{3}k (b_i - \overline{b}_i) \label{Mesdens} \end{eqnarray} where the spin-isospin degeneracy $\gamma^K_i=4,6$ for $K, K^{*}$, respectively. The Bose-Einstein distribution functions for the $K$ mesons are given by \begin{eqnarray} b_i=\frac{1}{e^{[\sqrt{k^2+M^2_i}-\mu_i]/T}-1} \label{boson} \end{eqnarray} and \begin{eqnarray} \overline{b}_i=\frac{1}{e^{[\sqrt{k^2+M^2_i}+\mu_i]/T}-1} \label{bosonbar} \end{eqnarray} where $\mu_i=S_i\mu_S$ is the chemical potential of $K$-meson of species $i$. The total energy density and total pressure of the $K$-meson ideal gas are given by \begin{eqnarray} \varepsilon^{id}_{K}= \sum_i \frac{\gamma^K_i}{(2\pi)^{3}} \int d^{3}k \sqrt{k^{2}+M^{2}_i} (b_i-\overline{b}_i), \label{EdensK} \end{eqnarray} and \begin{eqnarray} P^{id}_{K}=\frac{1}{3} \sum_i \frac{\gamma^K_i}{(2\pi)^{3}} \int d^{3}k \frac{k^2}{\sqrt{k^{2}+M^{2}_i}} (b_i-\overline{b}_i), \label{PdensK} \end{eqnarray} respectively. The scalar mean fields $\sigma$ and $\zeta$ are determined through the minimization of the thermodynamic potential or the maximizing of the pressure with respect to these fields. The pressure depends explicitly on the scalar mean fields through the $\frac{1}{2}m_{\sigma}^2\sigma^2$ and $\frac{1}{2}m_{\zeta}^2\zeta^2$ terms in Eq. (\ref{pressurD}). It also depends on the baryon effective masses $M^{*}_{i}$ which in turn also depend on $\sigma$ and $\zeta$. If we write the pressure as a function of $M_{i}^{*}$, $\sigma$ and $\zeta$\cite{Jaqaman,Jaqamanb}, the extremization of $P(M^{*}_{i},\sigma, \zeta)$ with respect to the scalar mean field $\sigma$ can be written as \begin{eqnarray} \left( \frac{\partial P}{\partial \sigma} \right)_{\zeta}= \sum_i \left( \frac{\partial P}{\partial M^{*}_{i}} \right)_{\mu_{B},T} \left( \frac{\partial M^{*}_{i}}{\partial \sigma} \right)_{\zeta} +\left(\frac{\partial P}{\partial \sigma}\right)_{\{M^{*}_{i}\}}=0, \label{dP1} \end{eqnarray} where \begin{eqnarray} \left(\frac{\partial P}{\partial \sigma}\right)_{\{M^{*}_{i}\}}= - m^{2}_{\sigma}\sigma, \label{dPdseg} \end{eqnarray} with a similar experssion for the extremization of $P(M^{*}_{i},\sigma, \zeta)$ with respect to the scalar mean field $\zeta$. The derivative of the pressure with respect to effective mass $M^{*}_{i}$ reads \begin{eqnarray} \left( \frac{\partial P}{\partial M^{*}_{i}} \right)_{\mu_{i},T}= &-&\frac{1}{3}\frac{\gamma_i}{(2\pi)^{3}} \int d^{3} k \frac{k^{2}}{{\epsilon^{*}}^{2}}\frac{M^{*}_{i}}{\epsilon_i^{*}} \left[f_{i}+\overline{f}_{i}\right] \nonumber \\ &-&\frac{1}{3}\frac{\gamma_i}{(2\pi)^{3}} \frac{1}{T}\int d^{3} k \frac{k^{2}}{\epsilon_i^{*}}\frac{M^{*}_{i}}{\epsilon_i^{*}} \left[f_{i}(1-f_{i})+\overline{f}_{i}(1-\overline{f}_{i})\right] \nonumber \\ &-&\frac{1}{3}\frac{\gamma_i}{(2\pi)^{3}} \frac{1}{T}g_{\omega i} \left(\frac{\partial \omega}{\partial M^{*}_{i}}\right)_{\mu_{i},T} \int d^{3} k \frac{k^{2}}{\epsilon_i^{*}} \left[f_{i}(1-f_{i})-\overline{f}_{i}(1-\overline{f}_{i})\right] \nonumber \\ &-&\frac{1}{3}\frac{\gamma_i}{(2\pi)^{3}} \frac{1}{T}g_{\phi i} \left(\frac{\partial \phi}{\partial M^{*}_{i}}\right)_{\mu_{i},T} \int d^{3} k \frac{k^{2}}{\epsilon_i^{*}} \left[f_{i}(1-f_{i})-\overline{f}_{i}(1-\overline{f}_{i})\right] \nonumber \\ &+& m^{2}_{\omega} \omega \left(\frac{\partial \omega}{\partial M^{*}_{i}}\right)_{\mu_{i},T} \nonumber \\ &+&m^{2}_{\phi} \phi \left(\frac{\partial \phi}{\partial M^{*}_{i}}\right)_{\mu_{i},T}. \label{dPdMef} \end{eqnarray} Since the baryon chemical potential $\mu_{B}$ and temperature are treated as input parameters, the variation of the vector mean field $\omega$ with respect to the effective baryon mass $M^{*}_{i}$ at a given value of the baryon density $\rho_{B}$ reads \begin{eqnarray} \left(\frac{\partial \omega}{\partial M_{i}^{*}}\right)_{\mu_{i},T}= -\frac{ \left[g_{\omega i} / m^{2}_{\omega}\right] \left[\gamma_i/(2\pi)^{3}\right]\left[1/T\right]\int d^{3}k \frac{M_{i}^{*}}{\epsilon^{*}_i} \left[f_{i}(1-f_{i})-\overline{f}_{i}(1-\overline{f}_{i})\right]} {1+\sum_j \left[ g^{2}_{\omega j}/ m^{2}_{\omega} \right] \left[\gamma_j/ (2\pi)^{3}\right] \left[1 /T \right] \int d^{3}k \left[f_{j}(1-f_{j})+\overline{f}_{j}(1-\overline{f}_{j})\right]}. \label{OME} \end{eqnarray} with a similar expression for $\left({\partial \phi}/{\partial M_{i}^{*}}\right)_{\mu_{i},T}$. The coupling of the scalar mean fields $\sigma$ and $\zeta$ with the quarks in the non-overlapping MIT bags through the solution of the point like Dirac equation should satisfy the self-consistency condition. These constraints are essential to obtain the correct solution of the scalar mean fields $\sigma$ and $\zeta$. \section{Results and Discussions} We have studied hypernuclear matter at finite temperature using the modified quark meson coupling model which takes the medium-dependence of the bag into account. We choose a direct coupling of the bag parameter to the scalar mean fields $\sigma$ and $\zeta$ in the form given in Eq. (\ref{bagCon}). The bag parameter is taken as that adopted by Jin and Jennings\cite{Jin} $B^{1/4}_{0}=188.1$ MeV and the free nucleon bag radius $R_{0}=0.60$ fm. We have taken the current quark masses to be $m_{u}=m_{d}=0$ and $m_{s}= 150$ MeV. For $g_{\sigma}^{q}=1$, the values of the vector meson coupling constant and the parameter $g_{\sigma}^{B}$, as fitted from the saturation properties of nuclear matter, are given as $g^{2}_{\omega}/4\pi=(3g^q_{\omega})^2/4\pi =5.24$ and ${g^{B}_{\sigma}}^{2}/4\pi=({3g^B_\sigma}^q)^{2}/4\pi$=3.69. The $Z_i=2.03, 1.814, 1.629$ and $1.505$ are chosen to reproduce the baryon masses at their experimental values $M_{N,\Lambda,\Sigma,\Xi}=$939, 1157, 1193, and 1313 MeV respectively. Normal nuclear matter saturation density is taken as ${\rho_B}_0=0.17$ fm$^{-3}$. The extra coupling constants needed to couple the scalar and vector mean fields $\zeta$ and $\phi$ to the $s$-quark are chosen to satisfy $SU(6)$ symmetry where $|g_\zeta^q|=\sqrt{2} |g_\sigma^q|$ and $|g_\phi^q|=\sqrt{2} |g_\omega^q|$ and $|{g_\zeta^B}^q|=\sqrt{2} |{g_\sigma^B}^q|$. If it is assumed that the mean fields $\zeta$ and $\phi$ are positive definite, then all the coupling constants are positive and the absolute value signs become redundant. The $\sigma$ mean field is supposed to simulate the exchange of correlated pairs of pions and may represent a very broad resonances observed in $\pi \pi$ scattering and fixed at $m_\sigma=550$ MeV, while the vector $\omega$ meson is identified with the actual meson whose mass is $m_\omega=783$ MeV. Since the mean fields, $\sigma$ and $\omega$, are considered as $<u\overline{d}>$ condensates, they interact only with $u,d$-quark in the baryons. On the other hand, the scalar and vector mean fields $\zeta, \phi$ are considered as actual mesons with $m_\zeta=975$ MeV and $m_\phi=1020$ MeV, respectively. They are considered as $<s\overline{s}>$ condensates and interact only with the $s$-quarks in the baryons. This picture is consistent with the chiral models\cite{Papa98,Papa99}. We fixed the total strangeness density $\rho_S$ to $0$ by introducing an ideal gas of $K$ and $ K^*$ mesons where $m_K=$ 495 MeV and $m_{K^*}=$ 892 MeV. The contribution of other $K$ mesons was found to be negligible. It is supposed in the ideal gas limit that the Kaons do not interact with the mean fields $\zeta$ and $\phi$. The extension to the case that the Kaons interact with the $\zeta$ and $\phi$ fields will be considered in a future work. We first solve Eqs. (\ref{rhoB1}), (\ref{rhoS1}), (\ref{vecomeg}) and (\ref{vecphi}) self consistently for given values of temperature $T$ and densities $\rho_B$ and $\rho_S$ to determine the baryonic and strangeness chemical potentials $\mu_B$ and $\mu_S$, respectively. These constraints are given in terms of the effective baryon masses $M^*_i$ which depend on the bag radii $R_i$, the quark chemical potentials $\mu^i_q$ and the mean fields. For given values of the scalar fields $\sigma, \zeta$ and vector fields $\omega, \phi$, the quark chemical potential and bag radius of species $i$ are obtained using the self consistency conditions Eqs. (\ref{nq}) and (\ref{MNR}), respectively. The pressure is evaluated for specific values of temperature $T$ and chemical potentials $(\mu_B, \mu_S)$ which now become input parameters. We then determine the values of $\sigma$ and $\zeta$ by using the extermization conditions as given in Eq. (\ref{dP1}). These constraints take into account the coupling of the quark with the scalar mean fields in the frame of the point like Dirac equation exactly\cite{Saito94,Jin}. The dependence of the baryon effective masses $M_{N,\Lambda,\Sigma,\Xi}^{*}$ on the total baryonic density $\rho_B$ and temperature is shown in Fig. 1 where it is seen that the baryon masses decrease with baryonic density except at the highest temperatures where the effective masses become almost density-independent. Moreover, for a given baryonic density $\rho_{B}$ it is seen that as the temperature is increased the mass $M_{i}^{*}$ of species $i$ first increases slightly up to about $T=150$ MeV and then decreases rather rapidly for $T>150$ MeV. These results are displayed in a different manner in Fig. 2 which plots the baryon masses $M^{*}_i$ as a function of $T$ for $\rho_B=0$ fm $^{-3}$ and ${\rho_B}_0=0.17$ fm $^{-3}$. It is seen that the effective baryon masses $M_{N,\Lambda,\Sigma,\Xi}^{*}$ with ${\rho_B}_0=0.17$ fm $^{-3}$ are less than those with zero baryonic density. Moreover, the effective baryonic masses $M_{N,\Lambda,\Sigma,\Xi}^{*}$ increase only slightly, if at all, with temperature up to about $T= 150$ MeV beyond which they decrease rapidly. This behaviour is qualitatively similar to our earlier results for normal nuclear matter\cite{Furnstahl,Jaqaman,Jaqamanb} where the rapid decrease in the nucleon's effective mass was, however, found to start at rather higher temperatures $T>200$ MeV. This rapid decrease of $M_{i}^{*}$ with increasing temperature resembles a phase transition at high temperatures and low density, when the system becomes a dilute gas of baryons in a sea of baryon-antibaryon pairs\cite{Furnstahl}. In Fig. 3, we display the scalar mean fields $\sigma$ and $\zeta$ as functions of the total baryonic density $\rho_B$ for various temperatures. It is seen that the value of $\sigma$ initially decreases with increasing temperature for temperatures less than 150 MeV. The scalar mean field $\zeta$ is almost negligible for such low temperatures. However, as the temperature reaches 150 MeV there are indications of an increase in $\sigma$ at low baryon densities where it attains a nonzero value at $\rho_{B}=0$. For still higher temperatures, the situation is more dramatic with the value of $\sigma$ increasing with temperature for all values of $\rho_{B}$. This is also qualitatively similar to our earlier results for $SU(2)$ nuclear matter\cite{Jaqaman,Jaqamanb}. Furthermore, the scalar mean field $\zeta$ becomes important for $T>150$ MeV and also increases with temperature for all values of $\rho_{B}$. An interesting new feature here is that the scalar mean fields $\sigma$ and $\zeta$ tend to take almost constant values irrespective of density at high temperatures which was not seen in our earlier calculations for $SU(2)$ nuclear matter even at temperatures as high $240$ MeV . Fig. 4 displays $\sigma$ and $\zeta$ versus $T$ for $\rho_B=0$ and ${\rho_B}_0=0.17$ fm $^{-3}$. It is seen that the $\sigma$ field has a nonzero (and almost constant) value only at the higher density until the temperature reaches about $T=150$ MeV when a rapid increase sets in at both densities. The increase at zero density is actually more dramatic and the $\sigma$ field rapidly attains values equal to those occuring at ${\rho_B}_0=0.17$ fm $^{-3}$. This is another indication of a phase transition to a system of baryon-antibaryon pairs. The behaviour of the $\zeta$ field is qualitatively similar except that its value is negligible at both densities for temperatures less than about $T=150$ MeV. In Fig. 5, we display the baryonic density dependence of the bag parameters for $N,\Lambda,\Sigma,\Xi$ for different values of the temperature. For each baryon, the bag parameter increases with temperature for temperatures less than 150 MeV. However, the situation is completely reversed after the phase transition takes place. For temperatures $T>150$ MeV the bag parameters display a dramatic decrease with temperature for all densities. This can be seen more clearly in Fig. 6 which displays $B_i$ vs $T$ for $\rho_B=0$ fm $^{-3}$ and ${\rho_B}_0=0.17$ fm $^{-3}$. The bag parameters are almost constant until the temperature exceeds $T=150$ MeV when they start to decrease rapidly. This indicates the onset of quark deconfinement above the critical temperature: at high enough temperature and/or baryon density there is a phase transition from the baryon-meson phase to the quark-gluon phase. This behaviour is also in qualitative agreement with our earlier results for $SU(2)$ nuclear matter\cite{Jaqaman} except that the decrease here is more dramatic indicating that the phase transition is much stronger than in ordinary nuclear matter. Our results are also comparable to those obtained from lattice QCD calculations which have so far only explored the zero baryon density axis of the phase diagram in a meaningful way. The lattice results with 2 light quark flavors, indicate that the transition from hadronic matter to a quark-gluon plasma occurs at a temperature $T\approx 140$ MeV and that for low densities it may not be a phase transition at all but what is called a rapid crossover\cite{PRLcross}. Fig. 7 displays the relative abundance $\rho_i/\rho_B$ for each baryon species. At low temperatures the nucleons $N$ are almost the only constituents. However, the contribution of the hyperons starts to be noticeable when the temperature reaches $100$ MeV and becomes more important when the temperature is increased to $T>150$ MeV. At temperatures $T>200$ MeV the contribution of the nucleons falls down to about half the total baryonic density. This can also be seen in Fig. 8 where we display the ratio of the baryonic strangeness density to the total baryonic density. As mentioned earlier, we keep the net strangeness of the system fixed to zero by introducing Kaons so that $\rho_S=\rho^{Baryons}_S+\rho^{Mesons}_S=0$. It is seen that the net strangeness of the baryons $\rho^{Baryons}_S$ is small at low temperatures $T<150$ MeV (note the logarithmic scale). However, as the temperature increases the net strangeness of the baryon octet becomes significant. For $T>150$ MeV, the ratio $\rho^{Baryons}_S/\rho_B$ becomes of order one. Therefore, the $K$-meson plays a significant role at high temperatures and probably also in the phase transition. In Fig. 9, we display the relative abundances of the various baryon species $\rho_i/\rho_B$ versus temperature at normal nuclear matter density ${\rho_B}_0=0.17$ fm $^{-3}$. It is seen that the the nucleons are dominant at low temperatures and their relative abundance $\rho_N/\rho_B$ decreases very slowly at first and deviates very little from $1$ until the temperature reaches about 100 MeV when $\rho_N/\rho_B$ starts to decrease rapidly and becomes negligibly small for temperatures larger than about 280 MeV. On the other hand, the abundance of the hyperons is negligible at low temperatures and increases significantly as the temperature is increased beyond $T=100$ MeV. They become more abundant than the nucleons for temperatures larger than 200 MeV. The relative abundance of hyperons in high energy heavy ion collisions can therefore be used as a simple thermometer to measure the temperature of the hot nuclear matter produced in the reaction and probably to study the phase transition to the quark gluon plasma. In conclusion, we have generalized the MQMC model to the case of hot hypernuclear matter by introducing two new meson fields that couple to the strange quarks and using $SU(6)$ symmetry to relate the coupling constants. The results are qualitatively similar to those obtained for $SU(2)$ nuclear matter including the onset of quark deconfinement. It is observed, however, that in this model quark deconfinement in $SU(3)$ (or hypernuclear) matter is much stronger than in $SU(2)$ (or normal) nuclear matter and that hyperons become more abundant than nucleons at high temperatures $T>200$ MeV. An ideal gas of Kaons was introduced to keep zero net strangeness density. The interaction of the Kaons with the meson fields will be considered in a future work. \acknowledgments Financial support by the Deutsche Forschungsgemeinschaft through the grant GR 243/51-1 is gratefully acknowledged.
\section{Discussions} \input{bib-sks} \end{document} \section{Introduction} \label{sec-intro} In this paper we investigate the existence and uniqueness of the solution to the Kuramoto-Sivashinsky (K--S) equation subject to a random forcing term. Specifically, the solution of \begin{equation} du + (u_{xxxx} +u_{xx} + u u_x) dt - dw =0, \label{stkeq} \end{equation} where $w$ is a Q--Wiener process in a probability space $(\Omega, {\cal F}, \mbox{${\rm I\!P}$})$. The Wiener process $w$ takes value in a Hilbert space to be specified later. The distributional derivative of $w(t)$ represents an external random force. The usual K--S equation ((\ref{stkeq}) without the $dw$ term) has been studied as a prototypical example for an infinite dimensional dynamical system. It possesses a finite dimensional maximal attractor (\cite{nic891,Collet,Ilynko,Goodman}) and inertial manifold (\cite{Foias,Jolly,Temam_Wang,Robinson}). Equation (\ref{stkeq}) arises in the modeling of surface erosion via ion sputtering in amorphous materials \cite{cue951}. The random forcing term in the model accounts for the fluctuations in the flux of the bombarding particles. Herein we confine our attention to the case of $u$ restricted to the interval $I := (-l , l)$, subject to the given initial condition $u_{0}$, and homogeneous Dirichlet boundary conditions, i.e. \begin{equation} u(0 , x) \ = \ u_{0}( x ) \, , -l < x < l \, , \mbox{ and } u(t , -l) \ = \ u(t, l) \ = 0 \, \mbox{ for } t > 0 . \label{bc1} \end{equation} We show that for any $T > 0$ there exists a unique solution to (\ref{stkeq}),(\ref{bc1}) for $0< t < T$, and establish a priori estimates for the solution. The approach we follow is similar to that for establishing existence and uniqueness for parabolic differential equations. Firstly we establish local existence (with respect to time) and then show that the solution remains bounded for any $T > 0$. For (\ref{stkeq}) these steps are preceeded by the introduction of a change of variable which enables us to consider, instead of the stochastic differential equation, a related deterministic equation. Local existence and uniqueness is then established via an application of a fixed point argument over a suitably defined space. The application of the fixed point theorem necessitaties expressing the nonlinear solution operator of the derived deterministic equation as a mapping from a space $\mbox{${\mathcal E}$ }$ into itself. To achieve this we must show that the extensions of two operators, which arise in the analysis, are well defined. This effort is the major part of section \ref{sec-lclex}. In section \ref{sec-glbex} we show that the local solution remains bounded for any $T > 0$ which implies global existence of the solution. We begin our discussion by presenting in the next section several definitions and basic results which we use later in our analysis. \section{Preliminaries} \label{sec-prel} As usual, we denote by $L^{p}(I)$, $p = 1, 2, \ldots $ the closure of $C^{\infty}(I)$ (the space of infinitely differentiable functions on $I$) with respect to the $L^{p}(I)$ norm: \[ \| f \|_{L^{p}(I)} \ = \ \left( \int_{I} \ | f |^{p} \ dx \right)^{1/p} \ . \] Also, $H_{0}^{k}(I)$, $k = 1, 2, \ldots $, denotes the closure of $C_{0}^{\infty}(I)$ (the space of infinitely differentiable functions on $I$ which vanish at the the endponts) with respect to the $H_{0}^{k}(I)$ norm: \[ \| f \|_{H^{k}_0(I)} \ = \ \left( \int_{I} \ | f |^{2} \ dx \ + \ \int_{I} \ | f' |^{2} \ dx \ + \ \ldots \ + \ \int_{I} \ | f^{(k)} |^{2} \ dx \ \right)^{1/2} \ . \] For convenience we use \[ H \ := \ L^{2}(I) \ \mbox{ and } \ V \ := \ H_{0}^{1}(I) \, . \] To account for the temperal dependence we use the Banach spaces $L^{p}( 0 , T ; L^{q}( I ) )$, with the associated norm: \[ \| f \|_{L^{p}( 0 , T ; L^{q}( I ) )} \ := \ \left( \int_{0}^{T} \, \left( \int_{I} \, | f |^{q} \, dx \right)^{p/q} dt \right)^{1/p} \ . \] \textbf{Note}: The spaces $L^{p}( 0 , T ; H_{0}^{k}( I ) )$ are defined analogously. A central role in the analysis below is played by the space $E$ defined by \[ E \ := \ L^{4}( 0 , T ; L^{4}( I ) ) \ . \] We remark that this choice for $E$ arises from the proof of lemma \ref{lmaGx} and is dictated by the nonlinear term $u u_{x}$. We begin by establishing the following embedding result which we combine with lemma \ref{lmaregy} to establish the setting for the application of the Banach contraction mapping theorem (lemma \ref{lmaBan}). \begin{lemma} \label{lmainc} For any $T > 0$ we have \begin{equation} L^{\infty}(0 , T; H) \cap L^{2}(0 , T; V) \subset E \ , \label{eqinc} \end{equation} and there exists a constant $K$, independent of $T > 0$, such that \begin{equation} \| u \|_{E} \le K \left( \| u \|_{L^{\infty}(0 , T; H)} \ + \ \| u \|_{L^{2}(0 , T; V)} \right) \ , \ u \in E \ . \label{eqinc2} \end{equation} \end{lemma} \textbf{Proof}: We have by the Sobolev embedding theorem, (see \cite{ada751}, pg. 217), that \[ H^{1/2}(D) \subset H^{1/4}(D) \subset L^{4}(D) \] and there exists a constant $C_{1} > 0$ such that \begin{equation} \| v \|_{L^{4}(D)} \ \le \ C_{1} \| v \|_{H^{1/2}(D)} \ . \label{eqbd1} \end{equation} Using the interpolation inequality for ${H^{1/2}(D)}$ in terms of $L^{2}(D)$ and ${H^{1}(D)}$ we have for some constant $C_{2} > 0$ and all $t \in [0 , T]$ \begin{equation} \| u( t ) \|_{H^{1/2}(D)} \le C_{2} \, \| u( t ) \|_{L^{2}(D)}^{1/2} \ \| u( t ) \|_{H^{1}(D)}^{1/2} \ . \label{eqbd2} \end{equation} Raising both sides of (\ref{eqbd2}) to the fourth power and integrating (\ref{eqbd2}) over the interval $[0 , T]$, equation (\ref{eqinc2}) follows using standard inequalitites and the definitions of the norms, with $K \ = \ C_{1} \, C_{2} / 2$ . \\ $\mbox{ }$ \hfill \hbox{\vrule width 6pt height 6pt depth 0pt} Essential to establishing the local existence is the following contraction mapping theorem. \begin{lemma} (\cite{dap961}, Pg. 290) \label{lmaBan} Let $\mbox{${\mathcal F}$ }$ denote a transformation from a Banach space $\mbox{${\mathcal E}$ }$ into $\mbox{${\mathcal E}$ }$, $\tilde{a}$ an element of $\mbox{${\mathcal E}$ }$ and $\alpha > 0$ a positive number. If $\mbox{${\mathcal F}$ }(0) = 0$, $\| \tilde{a} \| \le \frac{1}{2} \alpha$ and \begin{equation} \| \mbox{${\mathcal F}$ }(z_{1}) - \mbox{${\mathcal F}$ }(z_{2}) \|_{\mathcal{E}} \le \frac{1}{2} \| z_{1} - z_{2} \|_{\mathcal{E}} \mbox{ for } \| z_{1} \|_{\mathcal{E}} \le \alpha , \ \| z_{2} \|_{\mathcal{E}} \le \alpha , \label{eqaF} \end{equation} then the equation \begin{equation} z \ = \ \tilde{a} \ + \ \mbox{${\mathcal F}$ }(z) , \ \ z \in \mbox{${\mathcal E}$ } , \label{eqfp} \end{equation} has a unique solution $z \in \mbox{${\mathcal E}$ }$ satisfying $\| z \|_{\mathcal{E}} \le \alpha $. \end{lemma} $\mbox{ }$ \hfill \hbox{\vrule width 6pt height 6pt depth 0pt} Below we use the following lemma and corollary, which describe the regularity of the solution to a negative self--adjoint operator. \begin{lemma} (\cite{zei901} pg. 424) \label{lmaregy} Assume that $A$ is a negative self--adjoint operator on $H$ and \[ V \ = \ D( ( -A )^{1/4} ) \subset H \subset V' \ . \] Then $A$ and $S(t) \ = \ e^{t A}$ has a continuous extension from $V$ to $V'$. If \[ y(t) \ = \ y(t ; g) \ = \ S( t ) y_{0} \ + \ \int_{0}^{t} \, e^{(t - s) A} \, g(s) \, ds \ , t \in [0 , T] , \] for $y_{0} \in H$, and $ g \in L^{2}(0 , T; V') $, then \[ y \in L^{\infty}(0 , T; H) \cap L^{2}(0 , T; V) , \] and for some constant $L > 0$, independent of $T > 0$, \begin{equation} \| y \|_{L^{\infty}(0 , T; H)} \ + \ \| y \|_{L^{2}(0 , T; V)} \ \le \ L \left( \| y_{0} \|_{H} \ + \ \| g \|_{L^{2}(0 , T; V')} \right) \ . \label{ctyS} \end{equation} \end{lemma} $\mbox{ }$ \hfill \hbox{\vrule width 6pt height 6pt depth 0pt} \begin{corollary} \label{correg} For $A$, $S(t)$, and $y_{0}$ as described in lemma \ref{lmaregy}, we have that \begin{equation} \| S(t) y_{0} \|_{E} \le 8 K T^{1/4} \left( \| S(t) y_{0} \|_{L^{\infty}(0, T; H)}^{4} \ + \ \| S(t) y_{0} \|_{L^{\infty}(0, T; V)}^{4} \right)^{1/4} \ . \label{estSy} \end{equation} \end{corollary} \textbf{Proof}: Using (\ref{eqbd1}), (\ref{eqbd2}), (\ref{ctyS}) and, for notation convenience, $u \ := \ S(t) y_{0}$, we have that \begin{eqnarray*} \int_{0}^{T} \ \| u \|_{L^4}^{4} \ dt &\le& ( C_{1} \, C_{2} )^{4} \ \int_{0}^{T} \ \| u \|_{H}^{2} \ \| u \|_{V}^{2} \ dt \\ &\le& ( C_{1} \, C_{2} )^{4} / 2 \left( \int_{0}^{T} \ \| u \|_{H}^{4} \ dt \ + \ \int_{0}^{T} \ \| u \|_{V}^{4} \ dt \right) \\ &\le& ( C_{1} \, C_{2} )^{4} T / 2 \left( \ \| u \|_{L^{\infty}(0, T; H)}^{4} \ + \ \| u \|_{L^{\infty}(0, T; V)}^{4} \right) \ . \end{eqnarray*} Taking the fourth root of both sides yields (\ref{estSy}) for $K \ = \ C_{1} \, C_{2} / 2 $. \\ $\mbox{ }$ \hfill \hbox{\vrule width 6pt height 6pt depth 0pt} \textbf{Note}: From lemma \ref{lmaregy} we have that $S(t) \, y_{0} \in L^{\infty}(0, T; H) \cap L^{2}(0, T; V)$ which guarantees that the right hand side of (\ref{estSy}) is finite. \section{Local Existence and Uniqueness} \label{sec-lclex} Our first step in establishing local existence and uniqueness of the stochastic differential equation is to introduce a change of variable to reduce (\ref{stkeq}) to a deterministic equation. Denote by $A$ the self-adjoint operator \begin{equation} A u \ := \ - u_{xxxx} \ - u_{xx} \ - c \, u \ . \label{defA} \end{equation} We assume that $c$ is chosen sufficiently large such that A is a strictly negative operator on the space $H^{4}_{0}( I )$. Observe that as A is a strictly negative, self--adjoint, operator we can define $(-A)^{\alpha}$ via Fourier analysis, with domain $D( (-A)^{\alpha} ) \ = \ H^{4 \alpha}_{0}( I )$. (See \cite{tem881} pg.55 for details.) In view of (\ref{defA}) note that (\ref{stkeq}) can be rewritten in the form \begin{equation} du \ = \ ( A u \ - u u_{x} \ + c u ) \ dt \ + \ dw \ , \label{stkmeq} \end{equation} where the Wiener process $w$ takes value in the separable Hilbert space $H = L^2(I)$ and it has the covariance operator $Q$. With $S( t ) \ := \ e^{t A}$, $t \ge 0$, we define $w_{A}( t )$ via the stochastic integral \begin{equation} w_{A}( t ) \ := \ \int_{0}^{t} S( t - s ) \, dw(s) \ . \label{defwa} \end{equation} Using the substitution \begin{equation} y(t , x) \ := \ u(t , x) \ - \ w_{A}(t , x) \ , \ t \in [0 , T] \ \mbox{${\rm I\!P}$} \mbox{--} a.s. \ , \label{defy} \end{equation} (\ref{stkeq}) reduces to the deterministic problem \begin{equation} y_{t} \ = \ A y - (y \ + w_{A}) (y \ + w_{A})_{x} \ + c (y \ + w_{A}) \ , \label{deteq} \end{equation} subject to \begin{equation} y(0 , x) \ = \ u_{0}(x) \ \mbox{ and } y(t , -l) \ = \ y(t , l) \ = \ 0 \ . \label{bcy} \end{equation} \textbf{Note}: The assumption that $dw$ in (\ref{stkeq}) denotes a Q--Wiener process, together with the fact that $A$ is a strictly negative self--adjoint operator, ensures that $w_{A}(t)$ given by (\ref{defwa}) has a version which is H\"{o}lder continuous with values in $D( (-A)^{\alpha} )$ for $0 \le \alpha < 1/4$, with H\"{o}lder exponent less than $(1/4 - \alpha)$, (see \cite{dap961}, pg. 60). Thus, with $\alpha \ = \ 1 / 8$, in view of (\ref{eqbd1}), we conclude that $w_{A}(t)$ has a continuous version in $L^{4}(I)$. Below we take $w_{A}(t)$ to denote this continuous version. The solution $y$ satisfying (\ref{deteq}) may be expressed in integral form as \begin{eqnarray} y(t) & = & S( t ) u_{0} \ + \ \nonumber \\ & & \hspace{-0.3in} \int_{0}^{t} S( t - s ) \left[ - ( y( s ) \ + \ w_{A}( s ) ) ( y( s ) \ + \ w_{A}( s ) )_{x} \ + \ c ( y( s ) \ + \ w_{A}( s ) ) \right] \ ds \label{eqint1} \\ & = & S( t ) u_{0} \ + \ F( y \ + \ w_{A} )( t ) \ , \ \ t \in [0 , T] \ . \label{eqint2} \end{eqnarray} In the following we show the existence and uniqueness of the solution $y$ to this integral equation (\ref{eqint2}). This gives a so-called (mild) solution $u$ for the stochastic Kuramoto-Sivashinsky equation (\ref{stkeq}). This is the definition of `solution' used in this paper. In (\ref{eqint2}) $F : E \rightarrow E $ is a continuous extension of the operator \[ F_{0} : C^{1}([0 , T] ; V) \rightarrow E \] defined by \begin{equation} ( F_{0} u )( t ) \ = \ \int_{0}^{t} S(t - s) ( G_{0} u )( s ) \ ds \ , t \in [0 , T] , \label{defF0} \end{equation} where \[ G_{0} : C^{1}([0 , T] ; V) \rightarrow E \] is given by \begin{equation} ( G_{0} u )( t ) \ = \ - u u_{x}(t) \ + \ c u(t) \ , t \in [0 , T] . \label{defG0} \end{equation} In view of (\ref{eqint2}), and assuming that $F$ is well defined --- a non--trivial point whose discussion occupies the later part of this section --- , on applying lemma \ref{lmaBan} we have local existence of the solution to (\ref{stkeq}),(\ref{bc1}). \textbf{Note}: The value of $\tau$ for which we establish local existence and uniqueness of the solution on the interval $ [0 , \tau] $, depends upon the particular realization. \begin{theorem} \label{locex} For $u_{0}$ in $H$ there exists a random variable $\tau$ taking values $\mbox{${\rm I\!P}$}$--$a.s.$ in $(0 , T]$ such that equations (\ref{stkeq})(\ref{bc1}) have a unique solution $u$ on the interval $[0 , \tau]$. \end{theorem} Note that by a general result in \cite{dap921}, page 72, the solution $u$ has a measurable modification. In the following the solution $u$ refers to this measurable version. \textbf{Proof}: Observe that with $z(t) \ = \ y(t) \ + \ w_{A}(t) \ - \ S(t) \, u_{0} $, equation (\ref{eqint2}) may be rewritten as \begin{equation} z \ = \ \tilde{a} \ + \ \mbox{${\mathcal F}$ }(z) \label{defzeq} \end{equation} for $ \tilde{a} \ = \ w_{A}(t)$, and $\mbox{${\mathcal F}$ }(z) \ = \ F(z \ + \ S(t) \, u_{0})$. Thus the existence and uniqueness of the solution to (\ref{eqint2}) is equivalent to that for (\ref{defzeq}). Let $\alpha = 1/6 M$, and $\tau_{1}$ be given by \begin{equation} \tau_{1} \ = \ \left(6 M \, [ c \, (2 l)^{1/4} \ + \ 16 K \, ( \| S(t) u_{0} \|_{L^{\infty}(0, T; H)}^{4} \ + \ \| S(t) u_{0} \|_{L^{\infty}(0, T; V)}^{4} )^{1/4} \, ] \right)^{-4} \ , \label{dt1} \end{equation} for $K$ defined in lemma \ref{lmainc}, and $M$ defined in lemma \ref{lmextF}. The $\tau_{1}$ is well-defined and it is so chosen that it will guarantee that $\cal F$ is a contraction mapping (see below). As $w_{A}(t)$ is continuous with $w_{A}(0) \ = \ 0$, there exists $\tau_{2}$ such that \[ \int_{0}^{t} \, \| w_{A}(s) \|_{L^{4}(I)}^{4} \, ds \ \le \ \alpha / 2 \, , \ \mbox{ for } 0 \le t \le \tau_{2} \ . \] Let $\tau := \min \{\tau_{1} \, , \, \tau_{2} \}$ and analogous to the definition for $E$ introduce $\mbox{${\mathcal E}$ }$ as \[ \mbox{${\mathcal E}$ } \ := \ L^{4}(0 , \tau ; L^{4}(I)) \ . \] With $z_{1}$ and $z_{2}$ satisfying $\| z_{i} \|_{\mbox{${\mathcal E}$ }} \le \alpha $ ( = 1 / 6 M ) for $i = 1 , 2$, we have using lemma \ref{lmextF}, (\ref{estSy}), and the definition of $\tau$ \begin{eqnarray*} \| \mbox{${\mathcal F}$ }(z_{1}) - \mbox{${\mathcal F}$ }(z_{2}) \|_{\mbox{${\mathcal E}$ }} &\le& M \left( \| z_{1} \ + \ S(t) u_{0} \|_{\mbox{${\mathcal E}$ }} \ + \ \| z_{2} \ + \ S(t) u_{0} \|_{\mbox{${\mathcal E}$ }} \ + \ c ( 2 l \tau )^{1/4} \right) \ \| z_{1} \ - \ z_{2} \|_{\mbox{${\mathcal E}$ }} \\ &\le& M \left( \| z_{1} \|_{\mbox{${\mathcal E}$ }} \ + \ \| z_{2} \|_{\mbox{${\mathcal E}$ }} \ + \ 2 \| S(t) u_{0} \|_{\mbox{${\mathcal E}$ }} \ + \ c ( 2 l \tau )^{1/4} \right) \ \| z_{1} \ - \ z_{2} \|_{\mbox{${\mathcal E}$ }} \\ &\le& \frac{1}{2} \| z_{1} \ - \ z_{2} \|_{\mbox{${\mathcal E}$ }} \ . \end{eqnarray*} Finally, applying lemma \ref{lmaBan} we establish the existence and uniqueness of $z(t)$, and consequently $y(t)$, on the interval $[0 , \tau]$. \\ $\mbox{ }$ \hfill $\hbox{\vrule width 6pt height 6pt depth 0pt}$ What remains is to establish the regularity result used for $F$ in the proof of theorem \ref{locex}. However we must first show that $F$ is well defined by showing $G_{0}$ and $F_{0}$ defined by (\ref{defG0}) and (\ref{defF0}) have appropriate extensions. \begin{lemma} \label{lmaGx} The operator $G_{0}$ defined by (\ref{defG0}) can be continuously extended to \[ G : \ E \ \rightarrow \ L^{2}(0 , T ; V') \ , \] satisfying \[ \| G(u) \ - \ G(v) \|_{L^{2}(0 , T ; V')} \le 27^{1/4} ( \| u \|_{E} \ + \| v \|_{E} \ + c (2 l T)^{1/4} ) \| u - v \|_{E} \ \ u , v \in E \ . \] \end{lemma} \textbf{Proof}: Let $u, \ v, \ \psi \in L^{2}(0 , T; V)$. Denoting the duality mapping between $L^{2}(0 , T; V)$ and $L^{2}(0 , T; V')$ by $\langle \cdot , \cdot \rangle$, we have \begin{eqnarray*} \langle G_{0}( u ) - G_{0}( v ) , \psi \rangle & = & \int_{0}^{T} \int_{I} \left( - u u_{x} \ + v v_{x} \right) \psi \ + \ c (u - v) \psi \ dx \ dt \\ & = & \int_{0}^{T} \int_{I} \frac{1}{2} (u + v)(u - v) \psi_{x} \ + \ c (u - v) \psi \ dx \ dt \ , \end{eqnarray*} i.e. \begin{eqnarray*} | \langle G_{0}( u ) - G_{0}( v ) , \psi \rangle | & \le & \left( \int_{0}^{T} \int_{I} (|u| \ + \ |v| + \ c)^{2}(u \ - \ v)^{2} \ dx \ dt \right)^{1/2} \ \cdot \\ & & \hspace{1.0in} \left( \int_{0}^{T} \int_{I} (\frac{1}{2} |\psi_{x}| \ + \ |\psi|)^{2} \ dx \ dt \right)^{1/2} \\ & \le & \left( \int_{0}^{T} \int_{I} (|u| \ + \ |v| + \ |c|)^{4} \ dx \ dt \right)^{1/4} \ \cdot \\ & & \hspace{0.5in} \left( \int_{0}^{T} \int_{I} (u \ - \ v )^{4} \ dx \ dt \right)^{1/4} \ \| \psi \|_{L^{2}(0 , T ; V)} \\ & \le & \left( 27 ( \| u \|_{E}^{4} \ + \ \| v \|_{E}^{4} \ + \ \| c \|_{E}^{4} ) \right)^{1/4} \cdot \\ & & \hspace{1.0in} \|(u \ - \ v )\|_{E} \ \| \psi \|_{L^{2}(0 , T ; V)} \\ & \le & 27^{1/4} ( \| u \|_{E} \ + \ \| v \|_{E} \ + \ c (2 l T)^{1/4} ) \cdot \\ & & \hspace{1.0in} \ \|(u \ - \ v )\|_{E} \ \| \psi \|_{L^{2}(0 , T ; V)} \ , \end{eqnarray*} from which the result follows. $\mbox{ }$ \hfill $\hbox{\vrule width 6pt height 6pt depth 0pt}$ For the extension of $F_{0}$ and the regularity of $F$ we have: \begin{lemma} \label{lmextF} The transformation $F_{0}$ described by (\ref{defF0}) can be continuously extended to $F : \ E \rightarrow E$. Moreover there exists a constant $M > 0$, independent of $T > 0$, such that \begin{equation} \| F(u) \ - \ F(v) \|_{E} \le M ( \| u \|_{E} \ + \ \| v \|_{E} \ + \ c (2 l T)^{1/4} ) \|(u \ - \ v )\|_{E} \ , \ u, \, v \in E \ . \label{regF} \end{equation} \end{lemma} \textbf{Proof}: In view of the definition of $F_{0}$ in (\ref{defF0}) and lemma \ref{lmaregy} we have that \[ F( u ) (t) \ = \ y(t ; G(u)) \in L^{\infty}(0 , T; H) \cap L^{2}(0 , T; V) \ . \] Moreover from lemma \ref{lmainc} we have \begin{eqnarray*} \| F(u) \ - \ F(v) \|_{E} & \le & K \left( \| y(\cdot ; G(u)) \ - \ y(\cdot ; G(v)) \|_{L^{\infty}(0 , T; H)} \ + \right. \\ & & \hspace*{1.5in} \left. \| y(\cdot ; G(u)) \ - \ y(\cdot ; G(v)) \|_{L^{2}(0 , T; V)} \right) \\ & \le & K L \| G(u) - G(v) \|_{L^{2}(0 , T; V')} \ , \mbox{ using lemma \ref{lmaregy} } , \\ & \le & M ( \| u \|_{E} \ + \ \| v \|_{E} \ + \ c (2 l T)^{1/4} ) \|(u \ - \ v )\|_{E} \ , \ \ \mbox{ using lemma \ref{lmaGx} } , \end{eqnarray*} for $M = 27^{1/4} K L $ . \\ $ \mbox{ } $ \hfill $\hbox{\vrule width 6pt height 6pt depth 0pt}$ \section{Global Existence} \label{sec-glbex} We now extend the local existence of theorem \ref{locex} established in the previous section to global existence. Local existence establishes that the solution, $u$, lies in the solution space, $E$, for some initial time period. We establish global existence by showing that for any time $T$ the $E$--norm of $u$ is finite and hence $u$ still lies in the space $E$. To do this we first establish that the solutions are continuous with respect to the initial data. This enables us to restrict our attention to showing that \underline{strong solutions} remain bounded in $E$. Following directly the proof of lemma \ref{lmaGx} with the inner product (only) taken over the spatial domain, $I$, we have: \begin{corollary} \label{corGnmVp} The operator $G$ defined in lemma \ref{lmaGx} satisfies for $\ u , \ v \in E$ \begin{equation} \label{GnmVp} \| G(u) \ - \ G(v) \|_{V'} \le 27^{1/4} ( \| u \|_{L^{4}(I)} \ + \| v \|_{L^{4}(I)} \ + c (2 l)^{1/4} ) \ \| u - v \|_{_{L^{4}(I)}} \ . \end{equation} \end{corollary} \mbox{ } \hfill \hbox{\vrule width 6pt height 6pt depth 0pt} \begin{lemma} \label{ctdep} The solution, $y(t)$, of (\ref{eqint2}) depends continuous on the initial data $u_{0} \in H$, and the random forcing term $w_{A}(t) \in E$. \end{lemma} Note that the continuous dependence on $w_{A}(t)$ is needed in the proof of the next lemma where we approximate $w_{A}(t)$ by regular processes. \textbf{Proof}: Let $y_{0}$, and $y_{1}$ denote solutions of (\ref{eqint2}) generated by $u_{0}$, $w_{A}^{0}(t)$, and $u_{1}$, $w_{A}^{1}(t)$, respectively. Then, on the common existence interval of $y_{0}$ and $y_{1}$, \begin{eqnarray*} y_{0} \ - \ y_{1} & = & S( t ) ( u_{0} \ - \ u_{1} ) \ + \ \\ & & \int_{0}^{t} \left[ S( t - s ) G(y_{0} + w_{A}^{0})(s) \ - \ S( t - s ) G(y_{1} + w_{A}^{1})(s) \right] \ ds \ . \end{eqnarray*} From lemma \ref{lmaregy} we have $y_{0}$, and $y_{1}$ $\in L^{2}(0,T;V)$ thus $(y_{0} \ - \ y_{1} )(t) \in V , \mu \ a.e.$, (i.e. for almost all $t \in (0 , T)$). Using the continuity of $S(t)$ and lemma \ref{corGnmVp} we have that there exits constants $L_{1}$ and $C_{1}$ such that \begin{eqnarray*} \| y_{0} \ - \ y_{1} \|_{V} & \le & L_{1} \| u_{0} \ - \ u_{1} \|_{H} \ + \ \int_{0}^{t} C_{1} \| G(y_{0} + w_{A}^{0})(s) \ - \ G(y_{1} + w_{A}^{1}) \|_{V'} \ ds \\ & \le & L_{1} \| u_{0} \ - \ u_{1} \|_{H} \ + \ \\ & & \hspace*{-1.0in} C_{1} \int_{0}^{t} 27^{1/4} ( \| y_{0} + w_{A}^{0} \|_{L^{4}(I)} \ + \| y_{1} + w_{A}^{1} \|_{L^{4}(I)} \ + c (2 l)^{1/4} ) \| ( y_{0} + w_{A}^{0} ) \ - \ ( y_{1} + w_{A}^{1} ) \|_{_{L^{4}(I)}} \ ds \ , \\ & & \mu \ a.e. \end{eqnarray*} Using the Sobolev embedding theorem, the existence of $y_{0}$, and $y_{1}$ $\in E$, implies there exists constants $C_{2}$ and $C_{3}$ such that \[ \| y_{0} \ - \ y_{1} \|_{L^{4}(I)} \le C_{2} \left( \| u_{0} \ - \ u_{1} \|_{H} \ + \ \| w_{A}^{0} \ - \ w_{A}^{1} \|_{E} \right) \ + \ C_{3} \int_{0}^{t} \| y_{0} \ - \ y_{1} \|_{L^{4}(I)} \ ds \ \ \mu \ a.e. \] Applying Gronwall's inequality then yields \[ \| y_{0} \ - \ y_{1} \|_{L^{4}(I)} \le C_{2} \left( \| u_{0} \ - \ u_{1} \|_{H} \ + \ \| w_{A}^{0} \ - \ w_{A}^{1} \|_{E} \right) \ e^{C_{3} t} \ \ \ \mu \ a.e. \] from which the stated conclusion follows. \mbox{ } \hfill \hbox{\vrule width 6pt height 6pt depth 0pt} Next we establish appropriate norm estimates for the solution. \begin{lemma} \label{glnmest} Let $u_{0} \in H$, $w_{A}$ be given by (\ref{defwa}), and $y$ denote the solution of \begin{equation} y(t) \ = \ S( t ) u_{0} \ + \ F( y \ + \ w_{A} )( t ) \ , \ \ t \in [0 , T] \ . \label{eqgex1} \end{equation} Then, $y$ satisfies \begin{equation} \sup_{t \in [0 , T]} \, \| y( t ) \|_{H}^{2} \le \| u_{0} \|^2_{H} \, e^{\int_{0}^{T} \, f( s ) \, d s} \ + \ \int_{0}^{T} \, e^{\int_{0}^{T} \, f( s ) \, d s} \ g( \tau ) \, d \tau \ , \label{Hnmest} \end{equation} and \begin{equation} \int_{0}^{T} \, \| y ( t ) \|_{V}^{2} \, dt \le \| u_{0} \|^2_{H} \ + \ \sup_{s \in [0 , T]} \, \| y( s ) \|_{H}^{2} \, \int_{0}^{T} \, f( \tau ) \, d \tau \ + \ \int_{0}^{T} \, g( \tau ) \, d \tau \ , \label{Vnmest} \end{equation} where \begin{eqnarray*} f( t ) & = & \frac{1}{2} \left( 2 \, c \ + \ (2 \, + \, \frac{3}{4} C_{1} C_{2})^{2} \ + \ C_{1} C_{2} \| w_{A} \|_{L^{4}}^{4} \right) \ , \\ \mbox{ and } & & \hfill \\ g( t ) & = & c \, \| w_{A} \|_{H}^{2} \ + \ \frac{1}{4} \| w_{A} \|_{L^{4}}^{4} \ . \end{eqnarray*} \end{lemma} \textbf{Proof}: As $D( A )$ and $C(0, T; H_{0}^{2}(I))$ are dense in $H$ and $E$, respectively, and from lemma \ref{ctdep} we have established continuous dependence of the solution, it suffices to establish (\ref{Hnmest}),(\ref{Vnmest}) for the (strong) solution of the differential equation \begin{eqnarray} \frac{d y(t)}{d t} & = & A y(t) \ - \ (y(t) \ + \ w_{A}(t)) (y(t) \ + \ w_{A}(t))_{x} \ + \ c (y(t) \ + \ w_{A}(t)) \ , \label{eqde1} \\ y_{0} & = & u_{0} \ . \nonumber \end{eqnarray} We first show that \begin{equation} \frac{d}{d t} \| y(t) \|_{H}^{2} \ + \ \| y(t) \|_{V}^{2} \le f(t) \, \| y \|_{H}^{2} \ + \ c \, \| w_{A} \|_{H}^{2} \ + \ \frac{1}{4} \| w_{A} \|_{L^{4}}^{4} \ . \label{nmieq1} \end{equation} Multiplying (\ref{eqde1}) by $y(t)$ and integrating over $I$ we obtain \begin{equation} \frac{1}{2} \frac{d}{d t} \| y \|_{H}^{2} \ + \ \| y_{xx} \|_{H}^{2} \ - \ \| y_{x} \|_{H}^{2} \ = \ - \int_{-l}^{l} \, y (y \ + \ w_{A}) (y \ + \ w_{A})_{x} \, dx \ + \ c \int_{-l}^{l} \, y w_{A} dx \ . \label{nmieq2} \end{equation} We obtain a lower bound for $\| y_{xx} \|_{H}$ via: \begin{eqnarray*} \| y_{x} \|_{H}^{2} & = & \int_{-l}^{l} y_{x} \, y_{x} \, dx \ = \ -\int_{-l}^{l} \, y_{x x} \, y \, dx \\ & \le & \left( \int_{-l}^{l} \, y_{x x}^{2} \, dx \right)^{1/2} \, \left( \int_{-l}^{l} \, y^{2} \, dx \right)^{1/2} \\ & \le & \frac{1}{(2 \, + \, \frac{3}{4} C_{1} C_{2})} \, \| y_{xx} \|_{H}^{2} \ + \ \frac{(2 \, + \, \frac{3}{4} C_{1} C_{2})}{4} \, \| y \|_{H}^{2} \ . \end{eqnarray*} Rearranging yields \begin{equation} \| y_{xx} \|_{H}^{2} \ \ge \ (2 \, + \, \frac{3}{4} C_{1} C_{2}) \, \| y_{x} \|_{H}^{2} \ - \ \frac{(2 \, + \, \frac{3}{4} C_{1} C_{2})^{2}}{4} \, \| y \|_{H}^{2} \ . \label{nmieq3} \end{equation} Next consider the first integral on the r.h.s. of (\ref{nmieq2}) in three pieces: \begin{eqnarray} -\int_{-l}^{l} \, y \, y \, y_{x} \, dx & = & -\int_{-l}^{l} \, \frac{1}{3} \frac{d}{d x}( y^{3} ) \, dx \ = \ 0 \ , \label{nmieq4} \\ | -\int_{-l}^{l} \, y \, w_{A} \, {w_{A}}_{x} \, dx \, | & = & | \frac{1}{2} \int_{-l}^{l} \, y_{x} \, w_{A}^{2} \, dx \, | \ \le \frac{1}{2} \| y \|_{V}^{2} \ + \ \frac{1}{8} \| w_{A} \|_{L^{4}}^{4} \ , \label{nmieq5} \\ \mbox{ and } \hspace*{2in} & & \nonumber \\ -\int_{-l}^{l} \, \left(y \, y_{x} \, w_{A} \ + \ y^{2} \, {w_{A}}_{x} \right) \, dx \, & = & \int_{-l}^{l} \, y \, y_{x} \, w_{A} \, dx \, . \nonumber \end{eqnarray} This last term is estimated as follows. \begin{eqnarray} | \int_{-l}^{l} \, y \, y_{x} \, w_{A} \, dx | & \le & \left( \int_{-l}^{l} \, y_{x}^{2} \, dx \right)^{1/2} \, \left( \int_{-l}^{l} \, y^{2} \, w_{A}^{2} \, dx \right)^{1/2} \, \nonumber \\ & \le & \left( \int_{-l}^{l} \, y_{x}^{2} \, dx \right)^{1/2} \, \left( \int_{-l}^{l} \, y^{4} \, dx \right)^{1/4} \, \left( \int_{-l}^{l} \, w_{A}^{4} \, dx \right)^{1/4} \, \nonumber \\ & \le & \| y \|_{V} \, \| y \|_{L^{4}} \, \| w_{A} \|_{L^{4}} \nonumber \\ & \le & C_{1} \, C_{2} \| y \|_{V}^{3/2} \, \| y \|_{H}^{1/2} \,\| w_{A} \|_{L^{4}} \mbox{ ( using (\ref{eqbd1}),(\ref{eqbd2}) )} \nonumber \\ & \le & \frac{3 \, C_{1} \, C_{2}}{4} \| y \|_{V}^{2} \ + \ \frac{C_{1} \, C_{2}}{4}\| y \|_{H}^{2} \,\| w_{A} \|_{L^{4}}^{4} \label{nmieq7} \\ & & \hspace*{2in} \mbox{ ( using Young's Inequality )} . \nonumber \end{eqnarray} For the remaining term on the r.h.s. of (\ref{nmieq2}) we have \begin{eqnarray} | c \int_{-l}^{l} \, y \, w_{A} \, dx | & \le & c \left( \int_{-l}^{l} \, y^{2} \, dx \right)^{1/2} \, \left( \int_{-l}^{l} \, w_{A}^{2} \, dx \right)^{1/2} \nonumber \\ & \le & \frac{c}{2} \| y \|_{H}^{2} \ + \ \, \frac{c}{2} \| w_{A} \|_{H}^{2} \ . \label{nmieq8} \end{eqnarray} Combining (\ref{nmieq2})---(\ref{nmieq8}) yields (\ref{nmieq1}). The estimate for $\| y \|_{H}$ in (\ref{Hnmest}) now follows from the observation that (\ref{nmieq1}) is a first order differential inequality for $\| y \|_{H}$. The bound involving $\| y \|_{V}$ in (\ref{Vnmest}) is established by integrating (\ref{nmieq1}) from $0$ to $T$. \\ \mbox{ } \hfill \hbox{\vrule width 6pt height 6pt depth 0pt} We are now in a position to establish the global existence of the solution. \begin{theorem} \label{glbex} For $u_{0} \in H=L^2(I)$, there exists $\mbox{${\rm I\!P}$}$ a.s. a unique solution $u(\cdot , x) \in E$ of (\ref{stkeq}),(\ref{bc1}). \end{theorem} \textbf{Proof}: From theorem \ref{locex} we have existence of the solution $u(\cdot , x) \in E$, $\mbox{${\rm I\!P}$}$ a.s., for the interval $[0 , \tau]$. In view of (\ref{defy}) and lemma \ref{glnmest} we conclude that $u(t , x)$ remains bounded in $E$, $\mbox{${\rm I\!P}$}$ a.s. for all $t \ge 0$, which implies global existence of the solution to (\ref{stkeq}),(\ref{bc1}). \mbox{ } \hfill \hbox{\vrule width 6pt height 6pt depth 0pt} \bigskip Finally we remark that by following the same argument as in Brannan et al. \cite{Duan-SQG}, we can show that the solution is actually H\"{o}lder continuous in space with exponent less than $\frac18$. It is also possible to consider multiplicative noise in equation (1.1). The approach in this paper should also apply to other similar parabolic type stochastic partial differential equations.
\section{The paper has been withdrawed.} \end{document}
\section{Systems of conservation laws. Equations for the conserved densities} We consider systems of conservation laws \begin{equation} u^i_t=f^i(u)_x=v^i_j(u) u^j_x, ~~~~~ v^i_j=\frac{\partial f^i}{\partial u^j}. \label{cons} \end{equation} Eigenvalues $\lambda^i$ of the matrix $v^i_j$ are called the characteristic velocities of system (\ref{cons}). In what follows system (\ref{cons}) is assumed to be strictly hyperbolic, so that $\lambda^i$ are real and pairwise distinct. Let $\stackrel{\rightarrow}{\xi_i}=(\xi_i^1(u), ..., \xi_i^n(u))^t$ be the corresponding right eigenvectors: $$ v\ \stackrel{\rightarrow}{\xi_i}=\lambda^i \stackrel{\rightarrow}{\xi_i}, ~~ {\rm or, ~ in ~ the ~ components,} ~~ v^s_k\ \xi^k_i=\lambda^i\ \xi^s_i. $$ We denote by $L_i=\xi^k_i\frac{\partial}{\partial u^k}$ the Lie derivative along the vector field $\stackrel{\rightarrow}{\xi_i}$ and introduce commutator expansions $$ [L_i, L_j]=L_iL_j-L_jL_i=c^k_{ij}\ L_k, $$ where $c^k_{ij}$ are certain functions of $u$. Let $$ h(u)_t=g(u)_x $$ be any conservation law of system (\ref{cons}). Its density $h$ and flux $g$ satisfy the equations $$ \frac{\partial g}{\partial u^k}=\frac{\partial h}{\partial u^s}\ v^s_k. $$ Contraction with $\stackrel{\rightarrow}{\xi_i}=(\xi_i^1, ..., \xi_i^n)^t$ results in $$ \frac{\partial g}{\partial u^k}\ \xi^k_i=\frac{\partial h}{\partial u^s}\ v^s_k \ \xi^k_i, $$ or \begin{equation} L_i g= \lambda^i L_i h, ~~~ i=1,..., n. \label{int} \end{equation} Equations (\ref{int}) are the defining equations for the conserved densities $h$ and the corresponding fluxes $g$. The compatibility conditions of (\ref{int}) are of the form $$ L_i(L_jg)-L_j(L_ig)=c^k_{ij}L_kg, $$ or, taking into account (\ref{int}), $$ L_i(\lambda^jL_jh)-L_j(\lambda^iL_ih)=c^k_{ij}\lambda^kL_kh. $$ This results in the following linear second-order system for the conserved densities $h$: \begin{equation} L_iL_jh=\frac{L_j\lambda^i}{\lambda^j-\lambda^i}\ L_ih+ \frac{L_i\lambda^j}{\lambda^i-\lambda^j}\ L_jh+ c^k_{ij} \ \frac{\lambda^i-\lambda^k}{\lambda^i-\lambda^j}\ L_kh, ~~~ i\ne j. \label{h} \end{equation} In particular, $h=u^1, ..., u^n$ satisfy (\ref{h}). It should be pointed out that in the generic situation (to be more precise, in the case $c^k_{ij}\ne 0$ for any $i\ne j\ne k$) the overdetermined system (\ref{h}) possesses at most finite-dimensional linear space of solutions. In what follows we will make use of the equations satisfied by the ratio $\varphi =\frac{g}{h}$, which can be obtained by rewriting (\ref{int}) in the form $$ L_i(\varphi \ h)=\lambda^i\ L_ih, $$ or, equivalently, \begin{equation} L_i \ln h=\frac{L_i\varphi}{\lambda^i-\varphi}. \label{hphi} \end{equation} The compatibility conditions of (\ref{hphi}) imply the following nonlinear second-order system for $\varphi$: \begin{equation} \begin{array}{c} L_iL_j\varphi= \left(\frac{1}{\varphi-\lambda^i}+\frac{1}{\varphi-\lambda^j}\right) L_i\varphi \ L_j\varphi+ \frac{L_j\lambda^i}{\lambda^j-\lambda^i} \frac{\varphi-\lambda^j}{\varphi-\lambda^i}L_i\varphi+ \frac{L_i\lambda^j}{\lambda^i-\lambda^j} \frac{\varphi-\lambda^i}{\varphi-\lambda^j}L_j\varphi+ \\ \ \\ c^k_{ij} \ \frac{\lambda^i-\lambda^k}{\lambda^i-\lambda^j} \frac{\varphi-\lambda^j}{\varphi-\lambda^k} L_k\varphi. \end{array} \label{phi} \end{equation} Formula (\ref{hphi}) establishes an equivalence between the linear system (\ref{h}) and the nonlinear system (\ref{phi}). The ratio $\varphi=\frac{g}{h}$ naturally arises in projective differential geometry (describing surfaces conjugate to a congruence -- see sect.2), and in the Lie sphere geometry (parametrizing Ribaucour congruences of spheres -- see the Appendix). \section{Commuting flows} System of conservation laws \begin{equation} u^i_{\tau}=q^i(u)_x=w^i_j(u) u^j_x, ~~~~~ w^i_j=\frac{\partial q^i}{\partial u^j}, \label{cons1} \end{equation} is called the commuting flow of system (\ref{cons}), if $u_{t\tau}^i= u^i_{\tau t}$, or, equivalently, $$ \left( \frac{\partial f^i}{\partial u^j} \frac{\partial q^j}{\partial u^k} u^k_x \right)_x = \left( \frac{\partial q^i}{\partial u^j} \frac{\partial f^j}{\partial u^k} u^k_x \right)_x. $$ Equating the coefficients at $u^k_{xx}$, we arrive at the commutativity of matrices $v=v^i_j$ and $w=w^i_j$. Thus, they possess coinciding eigenvectors $\stackrel{\rightarrow}{\xi_i}$. Let $\mu^i$ be the characteristic velocities of system (\ref{cons1}): $$ w\ \stackrel{\rightarrow}{\xi_i}=\mu^i \stackrel{\rightarrow}{\xi_i}. $$ According to sect.1, conserved densities $h$ of system (\ref{cons1}) satisfy the equations \begin{equation} L_iL_jh=\frac{L_j\mu^i}{\mu^j-\mu^i}\ L_ih+ \frac{L_i\mu^j}{\mu^i-\mu^j}\ L_jh+ c^k_{ij} \ \frac{\mu^i-\mu^k}{\mu^i-\mu^j}\ L_kh. \label{h1} \end{equation} Since both systems (\ref{h}) and (\ref{h1}) share a common set of $n$ functionally independent solutions $h=u^1,..., u^n$, their coefficients must coincide identically (if this were not the case, there would be a first-order relation between $L_ih$, contradicting the functional independence of $u^1,..., u^n$). Thus, \begin{equation} \frac{L_j\mu^i}{\mu^j-\mu^i}=\frac{L_j\lambda^i}{\lambda^j-\lambda^i} ~~~~ {\rm for ~ any} ~~ i\ne j, \label{comm1} \end{equation} and \begin{equation} c^k_{ij}\left(\frac{\mu^i-\mu^k}{\mu^i-\mu^j}- \frac{\lambda^i-\lambda^k}{\lambda^i-\lambda^j}\right)=0 ~~~~ {\rm for ~ any} ~~ i\ne j\ne k. \label{comm2} \end{equation} In this form equations governing commuting flows of system (\ref{cons}) have been set down in \cite{Sevennec}. If $n=2$, equations (\ref{comm2}) are absent. Let us consider the case $n=3$ and assume that at least one of the coefficients $c^k_{ij}$ (with three distinct indices $i, j, k$) is nonzero. Then equations (\ref{comm2}) imply $$ \mu^i=\lambda^i b-a $$ for appropriate $b$ and $a$. Substitution of this representation in (\ref{comm1}) implies, however, that $a$ and $b$ must be constants, so that the commuting flow is trivial. Hence, for $n=3$, only systems with zero $c^k_{ij}$ (for distinct $i, j, k$) may possess nontrivial commuting flows. Similarly, in the case $n\geq 3$, the presence of "sufficiently many" nonzero coefficients $c^k_{ij}$ prevents the existence of nontrivial commuting flows. \section{Diagonalizable systems of conservation laws} Let us assume that all coefficients $c^k_{ij}$ (with distinct $i, j, k$) are zero. In this case one can normalise eigenvectors $\stackrel{\rightarrow}{\xi_i}$ in such a way that the Lie derivatives $L_i$ will pairwise commute: $[L_i, L_j]=0$, so that the remaining coefficients $c^j_{ij}$ will also be zero. The commutativity of $L_i$ implies the existense of the coordinates $R^1(u), ..., R^n(u)$, such that $L_i$ become partial derivatives: $$ L_i=\partial_i = \partial / \partial R^i. $$ In the coordinates $R^i$ equations (\ref{cons}) assume diagonal form \begin{equation} R_t^i=\lambda^i(R)\ R^i_x, ~~~~ i=1,...,n. \label{riemann} \end{equation} Variables $R^i$ are called the Riemann invariants of system (\ref{cons}). Systems (\ref{cons}), possessing Riemann invariants, are called diagonalizable. Let $$ u_t=f_x $$ be a conservation law of system (\ref{riemann}). In the diagonalizable case equations (\ref{int}) assume the form $$ \partial_i f=\lambda^i\partial_i u, ~~~~ i=1,..., n, $$ while system (\ref{h}) for the conserved densities $u$ simplifies to \begin{equation} \partial_i\partial_j u=a_{ij}\ \partial_i u+a_{ji}\ \partial_j u, ~~~ i\ne j, \label{u} \end{equation} where $a_{ij}=\frac{\partial_j\lambda^i}{\lambda^j-\lambda^i}$. The compatibility conditions of system (\ref{u}) are of the form \begin{equation} \partial_ka_{ij}=a_{ik}\ a_{kj}+a_{ij}\ a_{jk}-a_{ij}\ a_{ik}, ~~~~ i\ne j\ne k; \label{semiham} \end{equation} they must be identically satisfied if we want system (\ref{u}) to possess $n$ functionally independent solutions $u=u^1,..., u^n$. In fact, conditions (\ref{semiham}) imply the existence of infinitely many conservation laws parametrized by $n$ arbitrary functions of one variable. Systems (\ref{riemann}) satisfying (\ref{semiham}) are called semihamiltonian. We refer to \cite{Tsarev}, \cite{DN}, \cite{Sevennec} for further information concerning integrability, differential geometry and applications of semihamiltonian systems of conservation laws. Semihamiltonian systems possess infinitely many commuting flows $$ R^i_{\tau}=\mu^i \ R^i_x $$ with the characteristic velocities $\mu^i$ governed by the equations $$ \frac{\partial_j\mu^i}{\mu^j-\mu^i}= \frac{\partial_j\lambda^i}{\lambda^j-\lambda^i}=a_{ij}, ~~~~ i\ne j. $$ We point out that any semihamiltonian system possesses infinitely many different conservative representations. \section{Systems of conservation laws and line congruences. Hypersurfaces conjugate to a congruence} With any system of conservation laws $$ u^i_t=f^i(u)_x $$ we associate an $n$-parameter family of lines \begin{equation} \begin{array}{c} y^1=u^1\ y^0-f^1(u), \\ ................... \ \\ y^n=u^n\ y^0-f^n(u), \end{array} \label{cong} \end{equation} in the $(n+1)$-dimensional space $A^{n+1}$ with the coordinates $y^0, y^1, ..., y^n$. We refer to \cite{Fer1}, \cite{Fer2} for motivation and the most important properties of this correspondence. In the case $n=2$ we obtain two-parameter family of lines, or a congruence of lines in $A^3$. From the beginning of the 19th century theory of congruences was one of the most popular chapters of classical differential geometry -- see e.g. \cite{Finikov}. We keep the name "congruence" for $n$-parameter family of lines (\ref{cong}). Any congruence possesses $n$ focal hypersurfaces $\stackrel{\rightarrow}{{\bf r}_i}, ~~ i=1, ..., n$, ~ with the parametric equations $$ \stackrel{\rightarrow}{{\bf r}_i}=(y^0, y^1, ..., y^n)= \left(\lambda^i, \ u^1 \lambda^i-f^1, ..., u^n \lambda^i-f^n\right); $$ here $\lambda^1, ..., \lambda^n$ are the characteristic velocities of system (\ref{cons}) -- see \cite{Fer1}, \cite{Fer2}. A line (\ref{cong}) is tangent to $\stackrel{\rightarrow}{{\bf r}_i}$ in the point with $y^0=\lambda^i$. Let us consider a hypersurface $M^n$ with the radius-vector $\stackrel{\rightarrow}{{\bf r}}$ parametrized as follows: \begin{equation} \stackrel{\rightarrow}{{\bf r}}=(y^0, y^1, ..., y^n)= \left(\varphi, \ u^1 \varphi-f^1, ..., u^n \varphi-f^n\right); \label{conjugate} \end{equation} here $\varphi (u)$ is an arbitrary function which is assumed to be different from $\lambda^i$ so that $M^n$ is not focal. A line (\ref{cong}) meets $M^n$ in the point with $y^0=\varphi$. Obviously, any hypersurface $M^n\in A^{n+1}$ can be parametrized in the form (\ref{conjugate}) for an appropriate function $\varphi$. We say that hypersurface $M^n$ is conjugate to congruence (\ref{cong}) if and only if $$ L_iL_j\stackrel{\rightarrow}{{\bf r}} \in TM^n ~~~ {\rm for \ any} ~ i \ne j. $$ Geometrically, this means that the developable surfaces of congruence (\ref{cong}) meet $M^n$ in the curves of a conjugate net. In 3-space the notion of conjugacy between a surface and a congruence was introduced by Guichard (see \cite{Eisenhart}, chapter 1; \cite{Finikov}). \begin{theorem} Hypersurface (\ref{conjugate}) is conjugate to a congruence if and only if $\varphi$ is representable in the form $\varphi=\frac{g}{h}$, where $h_t=g_x$ is a conservation law of system (\ref{cons}). \end{theorem} \centerline{Proof:} The tangent space of $M^n$ is spanned by the vectors \begin{equation} L_j\stackrel{\rightarrow}{\bf r}=(L_j\varphi)\stackrel{\rightarrow}{\bf U}+(\varphi-\lambda^j)L_j\stackrel{\rightarrow}{\bf U}, \label{tangent} \end{equation} where $\stackrel{\rightarrow}{\bf U}$ denotes the $(n+1)$-vector $(1, u^1, ..., u^n)$. Hence, \begin{equation} L_j\stackrel{\rightarrow}{\bf U}=\frac{L_j\varphi}{\lambda^j-\varphi} \stackrel{\rightarrow}{\bf U} ~~ {\rm mod} ~ TM^n. \label{tm} \end{equation} Let us compute $L_iL_j\stackrel{\rightarrow}{\bf r}$: $$ L_iL_j\stackrel{\rightarrow}{\bf r}=(L_iL_j\varphi) \stackrel{\rightarrow}{\bf U} +(L_j\varphi)L_i\stackrel{\rightarrow}{\bf U}+L_i(\varphi-\lambda^j)L_j\stackrel{\rightarrow}{\bf U}+(\varphi-\lambda^j)L_iL_j\stackrel{\rightarrow}{\bf U}. $$ Inserting here $L_iL_j\stackrel{\rightarrow}{\bf U}$ from (\ref{h}) and keeping in mind (\ref{tm}), we arrive at $$ \begin{array}{c} L_iL_j\stackrel{\rightarrow}{\bf r}= (L_iL_j\varphi+L_j\varphi \frac{L_i\varphi}{\lambda^i-\varphi}+L_i(\varphi-\lambda^j) \frac{L_j\varphi}{\lambda^j-\varphi}+ \\ \ \\ (\varphi-\lambda^j)( \frac{L_j\lambda^i}{\lambda^j-\lambda^i}\frac{L_i\varphi}{\lambda^i-\varphi} +\frac{L_i\lambda^j}{\lambda^i-\lambda^j}\frac{L_j\varphi}{\lambda^j-\varphi} +c^k_{ij} \frac{\lambda^i-\lambda^k}{\lambda^i-\lambda^j} \frac{L_k\varphi}{\lambda^k-\varphi} ) ) \stackrel{\rightarrow}{\bf U} ~~{\rm mod} ~ TM^n. \end{array} $$ Hence, $L_iL_j\stackrel{\rightarrow}{\bf r}\in TM^n$ if and only if the coefficient at $\stackrel{\rightarrow}{\bf U}$ vanishes. The resulting system for $\varphi$ identically coincides with (\ref{phi}). Thus, hypersurfaces conjugate to a congruence (\ref{cong}) are parametrized by conservation laws of system (\ref{cons}). According to \cite{Eisenhart}, two hypersurfaces conjugate to one and the same congruence are said to be in relation F (or related by a Fundamental transformation). {\bf Remark 1}. The case $\varphi=\lambda^i$ requires a special treatment. In this case $M^n$ coincides with the i-th focal hypersurface of a congruence. A direct computation shows that the i-th focal hypersurface is conjugate to a congruence if and only if $c^i_{jk}=0$ for any $j, k\ne i$ (i is fixed!). This is equivalent to the existence of a function $R^i(u)$ (called the i-th Riemann invariant) such that $$ R^i_t=\lambda^i\ R^i_x; $$ in particular, all focal hypersurfaces are conjugate to a congruence if and only if system (\ref{cons}) possesses $n$ Riemann invariants. The proof and some further details can be found in \cite{Fer2}, see also \cite{Akivis}. {\bf Remark 2}. If conservation law $h_t=g_x$ is a linear combination of conservation laws (\ref{cons}), hypersurface $M^n$ degenerates into a hyperplane (which is automatically conjugate to any congruence). Thus, only "additional" conservation laws give rise to nontrivial conjugate hypersurfaces. {\bf Remark 3}. Conjugate hypersurfaces always appear in 1-parameter families since, for a fixed density $h$, one can add a constant $c$ to the flux $g$. The corresponding family of conjugate hypersurfaces $\stackrel{\rightarrow}{\bf r}_c$ determined by $\varphi _c =\frac{h}{g+c}$ forms a parallel family, that is, the directions $L_i \stackrel{\rightarrow}{\bf r}_c$ are independent of $c$. This immediately follows from (\ref{tangent}), since the ratio $\frac{L_j \varphi}{\varphi - \lambda^j}=-\frac{L_jh}{h}$ does not depend on $c$. \section{Surfaces harmonic to a congruence} In this section we consider 2-component systems of conservation laws \begin{equation} \begin{array}{c} u_t^1=f^1_x, \\ u_t^2=f^2_x, \end{array} \label{cons2} \end{equation} and the associated congruences of lines in $A^3$: \begin{equation} \begin{array}{c} y^1=u^1\ y^0-f^1, \\ y^2=u^2\ y^0-f^2. \end{array} \label{cong2} \end{equation} Let \begin{equation} \begin{array}{c} u_{\tau}^1=q^1_x, \\ u_{\tau}^2=q^2_x, \end{array} \label{comm3} \end{equation} be a commuting flow of system (\ref{cons2}). In Riemann invariants $R^1, R^2$ (we point out that any two-component system is diagonalizable) equations (\ref{cons2}) and (\ref{comm3}) assume the forms $$ \begin{array}{c} R_{t}^1=\lambda^1\ R^1_x, \\ R_{t}^2=\lambda^2\ R^2_x, \end{array} $$ and $$ \begin{array}{c} R_{\tau}^1=\mu^1\ R^1_x, \\ R_{\tau}^2=\mu^2\ R^2_x, \end{array} $$ respectively. The densities $u=(u^1, u^2)$ and the fluxes $f=(f^1, f^2)$, $q=(q^1, q^2)$ satisfy the equations $$ \partial_i f=\lambda^i\ \partial_i u, ~~~~ \partial_i q=\mu^i\ \partial_i u, ~~~~ i=1, 2. $$ With the commuting flow (\ref{comm3}) we associate $2$-parameter family of planes in $A^3$ defined by the equations \begin{equation} \frac{y^1-u^1\ y^0+f^1}{q^1} = \frac{y^2-u^2\ y^0+f^2}{q^2}. \label{plane2} \end{equation} The family of planes (\ref{plane2}) has the following remarkable properties: {\it 1. Each plane $\pi$ from family (\ref{plane2}) contains a line $l$ of congruence (\ref{cong2}). 2. The congruence of lines $l_1=\pi \cap \partial_1 \pi $ is conjugate to the focal surface $\stackrel{\rightarrow}{{\bf r}_1}$ of congruence (\ref{cong2}). Similarly, the congruence of lines $l_2=\pi \cap \partial_2 \pi $ is conjugate to $\stackrel{\rightarrow}{{\bf r}_2}$. Lines $l_1$ and $l_2$ are called the characteristics of plane $\pi$. Characteristic $l_1$ (resp, $l_2$), meets the line $l$ in the point of tangency of $l$ with the focal surface $\stackrel{\rightarrow}{{\bf r}_1}$ (resp, $\stackrel{\rightarrow}{{\bf r}_2}$). } The proof follows from the explicit parametrization of congruences $l_1$, $l_2$: \centerline{Congruence $l_1$} $$ \begin{array}{c} y^1=\left(u^1-\frac{q^1}{\mu^1}\right)y^0- \left(f^1-\frac{\lambda^1q^1}{\mu^1}\right), \\ \ \\ y^2=\left(u^2-\frac{q^2}{\mu^1}\right)y^0- \left(f^2-\frac{\lambda^1q^2}{\mu^1}\right). \end{array} $$ \centerline{Congruence $l_2$} $$ \begin{array}{c} y^1=\left(u^1-\frac{q^1}{\mu^2}\right)y^0- \left(f^1-\frac{\lambda^2q^1}{\mu^2}\right), \\ \ \\ y^2=\left(u^2-\frac{q^2}{\mu^2}\right)y^0- \left(f^2-\frac{\lambda^2q^2}{\mu^2}\right). \end{array} $$ Obviously, the line $l_1$ passes through the point $$ (y^0, y^1, y^2)= \left(\lambda^1, \ u^1 \lambda^1-f^1, \ u^2 \lambda^1-f^2\right) $$ of the focal surface $\stackrel{\rightarrow}{{\bf r}_1}$. Similarly, the line $l_2$ passes through the point $$ (y^0, y^1, y^2)= \left(\lambda^2, \ u^1 \lambda^2-f^1, \ u^2 \lambda^2-f^2\right) $$ of the focal surface $\stackrel{\rightarrow}{{\bf r}_2}$. The point of intersection $l_1\cap l_2\in \pi$ has the coordinates \begin{equation} \begin{array}{c} y^0=\frac{\lambda^2 \mu^1-\lambda^1 \mu^2}{\mu^1-\mu^2}, \\ \ \\ y^1=\frac{\lambda^2 \mu^1-\lambda^1 \mu^2}{\mu^1-\mu^2}\ u^1+ \frac{\lambda^1-\lambda^2 }{\mu^1-\mu^2}\ q^1-f^1, \\ \ \\ y^2=\frac{\lambda^2 \mu^1-\lambda^1 \mu^2}{\mu^1-\mu^2}\ u^2+ \frac{\lambda^1-\lambda^2 }{\mu^1-\mu^2}\ q^2-f^2, \end{array} \label{harm} \end{equation} and sweeps a surface in $A^3$. By a construction, surface (\ref{harm}) is the envelope of the family of planes (\ref{plane2}). It has the following geometric properties: {\it 1. Each tangent plane $\pi$ of surface (\ref{harm}) contains a line $l$ of congruence (\ref{cong2}). By a construction, $\pi$ and $l\in \pi$ correspond to the same values of parameters $R^1, R^2$. Thus, one can speak of the correspondence between lines (\ref{cong2}) and points of surface (\ref{harm}). 2. The net $R^1, R^2$ on surface (\ref{harm}) is conjugate. In other words, developable surfaces of congruence (\ref{cong2}) correspond to a conjugate net on surface (\ref{harm}). } Surfaces, satisfying the properties 1, 2, are called harmonic to congruence (\ref{cong2}) -- see \cite{Finikov}, p.251. Formulae (\ref{harm}) provide an explicit parametrization of surfaces, harmonic to congruence (\ref{cong2}), by commuting flows of system (\ref{cons2}). Conversely, any surface harmonic to congruence (\ref{cong2}) is representable in the form (\ref{harm}). \section{L\'evy transformations of semihamiltonian systems} Let us consider semihamiltonian system (\ref{riemann}) in Riemann invariants: $$ R_t^i=\lambda^i(R)\ R^i_x, ~~~~ i=1,...,n. $$ Conservation laws $$ u_t=f_x $$ of system (\ref{riemann}) satisfy the equations $$ \begin{array}{c} \partial_i f=\lambda^i\ \partial_i u, ~~~~~ i=1,...,n, \\ \ \\ \partial_i\partial_j u=a_{ij}\ \partial_i u+a_{ji}\ \partial_j u, ~~~ i\ne j. \end{array} $$ Let us choose particular conservation law $$ h_t=g_x $$ of system (\ref{riemann}) and introduce new variable $U$ by the formula \begin{equation} U=u-\frac{h}{\partial_{\alpha} h}\ \partial_{\alpha} u, \label{levi} \end{equation} where ${\alpha}$ is fixed. Transformations of this type originate from projective differential geometry of conjugate nets and are known as the transformations of L\'evy \cite{Levy}, \cite{Forsyth}, p.94, \cite{Eisenhart}, chapter 1. In paper \cite{Doliva} transformations of L\'evy have been identified with the vertex operators of the multicomponent KP hierarchy. Their geometric interpretation will be clarified in the second half of this section. We will refer to (\ref{levi}) as to the transformation of L\'evy ${\cal L}_{\alpha}$. A direct calculation shows that $U={\cal L}_{\alpha}(u)$ satisfies the equations of the same form as $u$: \begin{equation} \partial_i\partial_j U=A_{ij}\partial_i U+A_{ji}\partial_j U \label{U} \end{equation} where the new coefficients $A={\cal L}_{\alpha}(a)$ are given by the formulae $$ \begin{array}{c} A_{{\alpha}i}=\left(1-\frac{a_{i{\alpha}}h}{\partial_{\alpha}h}\right) \frac{\partial_ih}{h}, ~~~ i\ne {\alpha}, \\ \ \\ A_{ij}=a_{ij}+\partial_j \ln \left(1-\frac{a_{i{\alpha}}h}{\partial_{\alpha}h}\right), ~~~ i\ne {\alpha}, ~~ j ~ {\rm is \ arbitrary}.\\ \end{array} $$ Transformations ${\cal L}_{\alpha}$ can be pulled back to the transformations of the corresponding hydrodynamic type systems: let us introduce the system \begin{equation} R_T^i=\Lambda^i(R)\ R^i_X, ~~~~ i=1,...,n \label{newriemann} \end{equation} with the characteristic velocities \begin{equation} \begin{array}{c} \Lambda^{\alpha}=\frac{g}{h}, \\ \ \\ \Lambda^i=\frac{\displaystyle \lambda^i\partial_{\alpha}h-a_{i{\alpha}}g} {\displaystyle \partial_{\alpha}h-a_{i{\alpha}}h}, ~~~ i\ne {\alpha}. \end{array} \label{newlambda} \end{equation} \begin{theorem} Conservation laws $$ U_T=F_X $$ of system (\ref{newriemann}), (\ref{newlambda}) are the ${\cal L}_{\alpha}$-transforms of conservation laws $$ u_t=f_x $$ of system (\ref{riemann}): $$ \begin{array}{c} U={\cal L}_{\alpha}(u)=u-\frac{h}{\partial_{\alpha}h}\ \partial_{\alpha} u, \\ \ \\ F={\cal L}_{\alpha}(f)=f-\frac{g}{\partial_{\alpha}g}\ \partial_{\alpha} f. \end{array} $$ \end{theorem} Formally, the proof of this theorem follows from the identities $$ \partial_iF=\Lambda^i\ \partial_iU, ~~~~ A_{ij}=\frac{\partial_j \Lambda^i}{\Lambda^j-\Lambda^i}, $$ which can be verified by a direct calculation. Geometric constructions underlying these formulae will be discussed below. System (\ref{newriemann}), (\ref{newlambda}) will be called the ${\cal L}_{\alpha}$-transform of system (\ref{riemann}). Obviously, transformations ${\cal L}_{\alpha}$ preserve the semihamiltonian property. We also include L\'evy transformations of the Lame coefficients $h_i$ defined by the formulae $$ \partial_j \ln h_i= a_{ij}, ~~~ j\ne i. $$ The ${\cal L}_{\alpha}$-transformed Lame coefficients are given by $$ \begin{array}{c} H_{\alpha}=h_{\alpha} \ \frac{h}{\partial_{\alpha}h}, \\ \ \\ H_i=h_i\ \left(1-\frac{a_{i{\alpha}}h}{\partial_{\alpha}h} \right), ~~~ i\ne {\alpha}. \end{array} $$ One can check directly that $$ \partial_j \ln H_i= A_{ij}, ~~~ j\ne i. $$ L\'evy transformations of hydrodynamic type systems in Riemann invariants are closely related to the transformations of Laplace discussed recently in \cite{Fer3}, \cite{Kamran}. We recall that Laplace transformation $S_{{\alpha}{\beta}}$ of system (\ref{u}) is defined by the formula $$ U=S_{{\alpha}{\beta}}(u)=u-\frac{\partial_{\alpha}u}{a_{{\beta}{\alpha}}}, $$ where both indices ${\alpha}\ne {\beta}$ are fixed. Laplace transformations also induce transformations of the characteristic velocities $\lambda^i$, the explicit form of which has been set down in \cite{Fer3}. One can check directly that the L\'evy transformation ${\cal L}_{\alpha}$ of system (\ref{riemann}) is related to its L\'evy transformation ${\cal L}_{\beta}$ via the Laplace transformation $S_{{\alpha}{\beta}}$: $$ {\cal L}_{\alpha}=S_{{\alpha}{\beta}}\circ {\cal L}_{\beta}. $$ To clarify geometric picture underlying transformations ${\cal L}_{\alpha}$ we choose an arbitrary conservative representation $$ u_t^i=f^i_x $$ of system (\ref{riemann}) and introduce the associated congruence $$ \begin{array}{c} y^1=u^1y^0-f^1, \\ ............ \\ y^n=u^ny^0-f^n. \end{array} $$ Let $M^n$ be hypersurface conjugate to this congruence. Following sect.3, we represent the radius-vector $\stackrel{\rightarrow }{\bf r}$ of $M^n$ in the form $$ \stackrel{\rightarrow}{\bf r}=\left(\varphi, ~ u^1\varphi - f^1, ..., \ u^n\varphi - f^n \right), ~~~~ \varphi = \frac{g}{h}, $$ where $h_t=g_x$ is a conservation law of system (\ref{riemann}). Coordinate system $R^1, ..., R^n$ on $M^n$ is conjugate, so that $$ \partial_i\partial_j \stackrel{\rightarrow}{\bf r}\in TM^n ~~ {\rm for \ any} ~~ i\ne j. $$ Let us introduce a new congruence, formed by the tangents to the $R^{\alpha}$-curves on hypersurface $M^n$. Parametrically, its lines can be represented in the form $$ \stackrel{\rightarrow}{\bf r}+t\ \partial_{\alpha}\stackrel{\rightarrow}{\bf r}, $$ or, in the components, $$ \begin{array}{c} y^0=\varphi+t\ \partial_{\alpha}\varphi, \\ \ \\ y^1=u^1\varphi-f^1+t\ (u^1\partial_{\alpha}\varphi+(\varphi-\lambda^{\alpha}) \partial_{\alpha}u^1), \\ ............................................... \\ y^n=u^n\varphi-f^n+t\ (u^n\partial_{\alpha}\varphi+(\varphi-\lambda^{\alpha}) \partial_{\alpha}u^n). \end{array} $$ Inserting $t=\frac{y^0-\varphi}{\partial_{\alpha} \varphi}$ in the last n equations, we arrive at the new congruence \begin{equation} \begin{array}{c} y^1=U^1y^0-F^1, \\ ................ \\ y^n=U^ny^0-F^n, \end{array} \label{newcongr} \end{equation} where $$ \begin{array}{c} U^1=u^1+\frac{\varphi -\lambda^{\alpha}}{\partial_{\alpha}\varphi} \partial _{\alpha}u^1, ~~~ F^1=f^1+\varphi \frac{\varphi -\lambda^{\alpha}}{\partial_{\alpha}\varphi} \partial_{\alpha}u^1, \\ ...................................... \\ U^n=u^n+\frac{\varphi -\lambda^{\alpha}}{\partial_{\alpha}\varphi} \partial _{\alpha}u^n, ~~~ F^n=f^n+\varphi \frac{\varphi -\lambda^{\alpha}}{\partial_{\alpha}\varphi} \partial_{\alpha}u^n. \end{array} $$ Since $\frac{\varphi-\lambda^{\alpha}}{\partial_{\alpha}\varphi} =-\frac{h}{\partial_{\alpha}h}$, these formulae can be rewritten in the form $$ U=u-\frac{h}{\partial_{\alpha}h}\ \partial_{\alpha}u, ~~~ F=f-\frac{g}{\partial_{\alpha}g}\ \partial_{\alpha}f. $$ Congruence (\ref{newcongr}) will be called the ${\cal L}_{\alpha}$-transform of the initial congruence. The corresponding system of conservation laws $$ U^i_T=F^i_X $$ has the same Riemann invariants $R^1, ..., R^n$: $$ R^i_T=\Lambda^i\ R^i_X, $$ where $\Lambda^i$ can be computed as follows: $\Lambda^i=\partial_iF/\partial_iU$. A direct calculation results in formulae (\ref{newlambda}). Note that the final expressions for $\Lambda^i$ do not depend on the particular conservative representation $u^i_t=f^i_x$ of system (\ref{riemann}). If, for $M^n$, we choose any of the focal hypersurfaces of the congruence (which are all conjugate to a congruence if the system possesses Riemann invariants), the above construction gives transformations of Laplace. Formula (\ref{U}) shows that the density $u=h$ belongs to the kernel of the L\'evy transformation ${\cal L}_{\alpha}$. Nevertheless, transformations ${\cal L}_{\alpha}$ can be explicitely inverted, as we will demonstrate in the next section. Let us conclude with the formula for the composition of the L\'evy transformations $$ {\cal L}={\cal L}_n\circ ... \circ {\cal L}_2\circ {\cal L}_1 $$ corresponding to $n$ particular linearly independent conservation laws $h^i_t=g^i_x, ~~ i=1,...,n$ of system (\ref{riemann}). The composition is understood as follows. Let $u_t=f_x$ be an arbitrary conservation law of system (\ref{riemann}). First of all, we apply to $u_t=f_x$ transformation ${\cal L}_1$, corresponding to the first conservation law $h^1_t=g^1_x$. Secondly, we apply to the result of the first step transformation ${\cal L}_2$, corresponding to the ${\cal L}_1$-transform of conservation law $h^2_t=g^2_x$. Proceeding in this way, we obtain the ${\cal L}$-transformed density $U={\cal L}(u)$ and the flux $F={\cal L}(f)$ in the following compact form: \begin{equation} U=\frac{det \left( \begin{array}{cccc} u & \partial_1u & ... & \partial_nu \\ h^1 & \partial_1 h^1 & ... & \partial_n h^1 \\ \ \\ h^n & \partial_1 h^n & ... & \partial_n h^n \end{array} \right)} {det \left( \begin{array}{ccc} \partial_1 h^1 & ... & \partial_n h^1 \\ \ \\ \partial_1 h^n & ... & \partial_n h^n \end{array} \right)}, ~~~~~ F=\frac{det \left( \begin{array}{cccc} f & \partial_1f & ... & \partial_nf \\ g^1 & \partial_1 g^1 & ... & \partial_n g^1 \\ \ \\ g^n & \partial_1 g^n & ... & \partial_n g^n \end{array} \right)} {det \left( \begin{array}{ccc} \partial_1 g^1 & ... & \partial_n g^1 \\ \ \\ \partial_1 g^n & ... & \partial_n g^n \end{array} \right)}. \label{UF} \end{equation} Geometrically, the composition ${\cal L}_n\circ ... \circ {\cal L}_2\circ {\cal L}_1$ corresponds to the following construction (compare with \cite{Finikov}, p.255-256): choose an arbitrary conservative representation $$ u^i_t=f^i_x $$ of system (\ref{riemann}) and introduce the corresponding congruence (\ref{cong}): $$ y^i=u^iy^0-f^i. $$ Let $M_i, ~i=1, ...,n$, be $n$ hypersurfaces conjugate to congruence (\ref{cong}). According to sect.2, they are parametrized by $n$ particular conservation laws $h^i_t=g^i_x$ of system (\ref{riemann}). Let $TM_i$ be the tangent hyperplanes of hypersurfaces $M_i$ in the points of intersection with line (\ref{cong}). The intersection $$ TM_1\cap ... \cap TM_n $$ defines a new line $$ y^i=U^iy^0-F^i; $$ one can check directly, that the formulae for $U=U^i$ and $F=F^i$ coincide with (\ref{UF}). \section{The adjoint transformations of L\'evy} We again consider semihamiltonian systems (\ref{riemann}) $$ R_t^i=\lambda^i(R)\ R^i_x $$ with conservation laws $$ u_t=f_x $$ satisfying the equations $$ \begin{array}{c} \partial_i f=\lambda^i\ \partial_i u, \\ \ \\ \partial_i\partial_j u=a_{ij}\ \partial_i u+a_{ji}\ \partial_j u, \end{array} $$ where $a_{ij}=\frac{\partial_j\lambda^i}{\lambda^j-\lambda^i}$. Let \begin{equation} R^i_{\tau}=\mu ^i(R)\ R^i_x \label{mu} \end{equation} be a commuting flow of system (\ref{riemann}): $$ \frac{\partial_j \mu^i}{\mu^j-\mu^i}=a_{ij}. $$ Let $q$ be the flux of density $u$, corresponding to this commuting flow: $$ u_{\tau}=q_x. $$ The flux $q$ and the density $u$ satisfy the equations $$ \partial_i q=\mu^i\ \partial_i u. $$ Let us introduce new variable $U$ by the formula \begin{equation} U=u-\frac{q}{\mu^{\alpha}}, \label{U1} \end{equation} where ${\alpha}$ is fixed. We will refer to (\ref{U1}) as to the adjoint transformation of L\'evy ${\cal L}^{*}_{\alpha}$. A direct calculation shows that $U={\cal L}^{*}_{\alpha}(u)$ satisfies the equations of the same form as $u$: $$ \partial_i\partial_j U=A_{ij}\partial_i U+A_{ji}\partial_j U $$ where the new coefficients $A={\cal L}^{*}_{\alpha}(a)$ are given by the formulae $$ \begin{array}{c} A_{{\alpha}i}=a_{{\alpha}i}+ \partial_i\ln\frac{\partial_{\alpha}\mu^{\alpha}}{\mu^{\alpha}}, ~~~ i\ne {\alpha}, \\ \ \\ A_{ij}=a_{ij}+\partial_j \ln \left(1-\frac{\mu^i}{\mu^{\alpha}}\right), ~~~ i\ne {\alpha}, ~~ j ~ {\rm is \ arbitrary}.\\ \end{array} $$ Transformations ${\cal L}^{*}_{\alpha}$ can be pulled back to the transformations of the corresponding hydrodynamic type systems: let us introduce the system \begin{equation} R_T^i=\Lambda^i(R)\ R^i_X, ~~~~ i=1,...,n \label{newriemann1} \end{equation} with the characteristic velocities \begin{equation} \begin{array}{c} \Lambda^{\alpha}= \frac{\displaystyle \lambda^{\alpha}\partial_{\alpha}\mu^{\alpha}- \mu^{\alpha}\partial_{\alpha}\lambda^{\alpha}} {\displaystyle \partial_{\alpha}\mu^{\alpha}}, \\ \ \\ \Lambda^i=\frac{\displaystyle \lambda^i\mu^{\alpha}-\lambda^{\alpha}\mu^i} {\displaystyle \mu^{\alpha}-\mu^i}, ~~~ i\ne {\alpha}. \end{array} \label{newlambda1} \end{equation} \begin{theorem} Conservation laws $$ U_T=F_X $$ of system (\ref{newriemann1}), (\ref{newlambda1}) are the ${\cal L}^{*}_{\alpha}$-transforms of conservation laws $$ u_t=f_x $$ of system (\ref{riemann}): $$ \begin{array}{c} U={\cal L}^{*}_{\alpha}(u)=u-\frac{q}{\mu^{\alpha}}, \\ \ \\ F={\cal L}^{*}_{\alpha}(f)=f-\frac{\lambda^{\alpha}q}{\mu^{\alpha}}. \end{array} $$ \end{theorem} Formally, the proof of this theorem follows from the identities $$ \partial_iF=\Lambda^i\ \partial_iU, ~~~~ A_{ij}=\frac{\partial_j \Lambda^i}{\Lambda^j-\Lambda^i}, $$ which can be verified by a direct calculation. Geometric constructions underlying these formulae will be discussed below. System (\ref{newriemann1}), (\ref{newlambda1}) will be called the ${\cal L}^{*}_{\alpha}$-transform of system (\ref{riemann}). Obviously, transformations ${\cal L}^{*}_{\alpha}$ preserve the semihamiltonian property. We also include ${\cal L}^{*}_{\alpha}$-transforms of the Lame coefficients $h_i$ defined by the formulae $$ \partial_j \ln h_i= a_{ij}, ~~~ j\ne i. $$ The ${\cal L}^{*}_{\alpha}$-transformed Lame coefficients are given by $$ \begin{array}{c} H_{\alpha}=h_{\alpha} \ \frac{\partial_{\alpha}\mu^{\alpha}}{\mu^{\alpha}}, \\ \ \\ H_i=h_i\ \left(1-\frac{\mu^i}{\mu^{\alpha}} \right), ~~~ i\ne {\alpha}. \end{array} $$ One can check directly that $$ \partial_j \ln H_i= A_{ij}, ~~~ j\ne i. $$ Transformations ${\cal L}^{*}_{\alpha}$ and the Laplace transformations $S_{{\alpha}{\beta}}$ satisfy the identities $$ {\cal L}^{*}_{\alpha}={\cal L}^{*}_{\beta}\circ S_{{\beta}{\alpha}}. $$ To clarify geometric picture underlying transformations ${\cal L}^{*}_{\alpha}$ we choose an arbitrary conservative representation $$ u_t^i=f^i_x $$ of system (\ref{riemann}) and introduce the associated congruence $$ \begin{array}{c} y^1=u^1y^0-f^1, \\ ............ \\ y^n=u^ny^0-f^n. \end{array} $$ Let $$ u^i_{\tau}=q^i_x $$ be a commuting flow of system (\ref{riemann}) with the characteristic velocities $\mu$, so that $$ \partial_iq=\mu^i\ \partial_iu, $$ (the last identity holding for any $q=q^k, ~ u=u^k$). Let us introduce $n$-parameter family of 2-planes in $A^{n+1}$ defined by the equations \begin{equation} \frac{y^1-u^1y^0+f^1}{q^1} = ... = \frac{y^n-u^ny^0+f^n}{q^n}. \label{plane} \end{equation} The family of planes (\ref{plane}) possesses the following three important properties: {\it 1. Each plane $\pi$ from family (\ref{plane}) contains a line $l$ of the initial congruence. 2. Each plane $\pi$ intersects the plane $\partial_i\pi$ along a line $l_i$: $$ l_i=\pi \cap \partial_i\pi, $$ (we point out that two planes in $A^{n+1}$ do not necessarily intersect along a line unless $n=2$). Geometrically, this property implies that each 1-parameter subfamily of (\ref{plane}), specified by fixing the values of $R^k, k\ne i$, envelopes a developable surface in $A^{n+1}$. Lines $l_i, ~ i=1,...,n$, are called the characteristics of plane $\pi$. 3. Congruence $l_i$ is conjugate to the i-th focal hypersurface $$ \stackrel{\rightarrow}{{\bf r}_i}= \left(\lambda^i, \ u^1 \lambda^i-f^1, ..., u^n \lambda^i-f^n\right) $$ of the initial congruence $l$.} Conversely, one can show that any $n$-parameter family of 2-planes satisfying the properties 1 -- 3 is necessarily of the form (\ref{plane}) for an appropriate commuting flow $u^i_{\tau}=q^i_x$. Congruence $l_{\alpha}$ will be called the ${\cal L}^{*}_{\alpha}$-transform of the initial congruence $l$. A direct calculation shows that $l_{\alpha}$ is representable in the form $$ \begin{array}{c} y^1=U^1y^0-F^1, \\ ................ \\ y^n=U^ny^0-F^n, \end{array} $$ where $$ \begin{array}{c} U^1=u^1-\frac{q^1}{\mu^{\alpha}}, ~~~ F^1=f^1-\frac{\lambda^{\alpha}q^1}{\mu^{\alpha}}, \\ ...................................... \\ U^n=u^n-\frac{q^n}{\mu^{\alpha}}, ~~~ F^n=f^n-\frac{\lambda^{\alpha}q^n}{\mu^{\alpha}}, \end{array} $$ (compare with Theorem 4). Line $l_{\alpha}$ meets the focal hypersurface $\stackrel{\rightarrow}{{\bf r}_{\alpha}}$ in the point $$ \left(\lambda^{\alpha}, \ u^1 \lambda^{\alpha}-f^1, ..., u^n \lambda^{\alpha}-f^n\right). $$ The corresponding system of conservation laws $$ U^i_T=F^i_X $$ has the same Riemann invariants $R^1, ..., R^n$: $$ R^i_T=\Lambda^i\ R^i_X, $$ (in fact, this is the analytic manifestation of the above property 3), where the transformed characteristic velocities $\Lambda^i=\partial_iF/\partial_iU$ coincide with (\ref{newlambda1}). Note that the final expressions for $\Lambda^i$ do not depend on the particular conservative representation $u^i_t=f^i_x$ of system (\ref{riemann}). Obviously, the inverse transformation $l_{\alpha} \to l$ is the transformation ${\cal L}_{\alpha}$ of L\'evy. Indeed, $l_{\alpha}$ is conjugate to hypersurface $\stackrel{\rightarrow}{{\bf r}_{\alpha}}$, while the initial congruence $l$ consists of the $R^{\alpha}$-tangents to hypersurface $\stackrel{\rightarrow}{{\bf r}_{\alpha}}$. Thus, transformations of L\'evy ${\cal L}_{\alpha}$ are the inverses of ${\cal L}^{*}_{\alpha}$. This can be demonstrated analytically as well: Let us consider a system $$ R^i_t=\lambda^i\ R^i_x $$ along with its L\'evy transform ${\cal L}_{\alpha}$ defined by formulae (\ref{newriemann}), (\ref{newlambda}). The transformed system (\ref{newriemann}), (\ref{newlambda}) possesses commuting flow $$ \begin{array}{c} \mu^{\alpha}=\frac{1}{h}, \\ \ \\ \mu^i=\frac{\displaystyle a_{i{\alpha}}} {\displaystyle a_{i{\alpha}}h-\partial_{\alpha}h}, ~~~ i\ne {\alpha}, \end{array} $$ (which can be obtained by a shift $g\to g+1$ in formulae (\ref{newlambda})). Applying to the transformed system (\ref{newriemann}), (\ref{newlambda}) transformation ${\cal L}^{*}_{\alpha}$ (generated by the above commuting flow), we return to the initial system $$ R^i_t=\lambda^i\ R^i_x. $$ Conversely, let us consider transformation ${\cal L}^{*}_{\alpha}$. The transformed system (\ref{newriemann1}), (\ref{newlambda1}) possesses conservation law $$ h_T=g_X, ~~~~ h=\frac{1}{\mu^{\alpha}}, ~~ g=\frac{\lambda^{\alpha}}{\mu^{\alpha}} $$ (which can be obtained by a shift $q\to q-1$ in formula (\ref{U1})). Applying to (\ref{newriemann1}), (\ref{newlambda1}) transformation ${\cal L}_{\alpha}$ (generated by this particular $h$), we also return back to the initial system. \section{Appendix: Ribaucour congruences of spheres} Let $M^n$ be a hypersurface in the Euclidean space $E^{n+1}$ parametrized by coordinates $u^1, ..., u^n$. Let $\stackrel{\rightarrow}{\bf r}$ and $\stackrel{\rightarrow}{\bf n}$ be the radius-vector and the unit normal of $M^n$, respectively. The Weingarten formulae $$ \frac{\partial \stackrel{\rightarrow}{\bf n}}{\partial u^j}= w^i_j(u) \ \frac{\partial \stackrel{\rightarrow}{\bf r}}{\partial u^i} $$ define the so-called Weingarten (shape) operator of hypersurface $M^n$. Its eigenvalues and eigenvectors are called the principal curvatures and the principal directions of $M^n$, respectively. Let us consider a hypersphere $S$ of radius $R$ and the centre $\stackrel{\rightarrow}{\bf r}-R\stackrel{\rightarrow}{\bf n}$, which is tangent to $M^n$ at the point $\stackrel{\rightarrow}{\bf r}$. Specifying $R$ as a function of $u$, we obtain $n$-parameter family of hyperspheres (or a congruence of hyperspheres) enveloped by hypersurface $M^n$. Let $\tilde M^n$ be the second sheet of the envelope. Clearly, there exists a point correspondence between both sheets $M^n$ and $\tilde M^n$: a point $p\in M^n$ corresponds to $\tilde p \in \tilde M^n$ if $p$ and $\tilde p$ are the two points of tangency of one and the same hypersphere from the family $S(u)$. {\bf Definition}. Family of hyperspheres $S(u)$ is called the family of Ribaucour if the principal distributions of $M^n$ correspond to the principal distributions of $\tilde M^n$. Let us introduce the system of hydrodynamic type \begin{equation} u^i_t=w^i_j(u) \ u^j_x, \label{w} \end{equation} where $w^i_j$ is the Weingarten operator of $M^n$. We refer to \cite{Fer4} for the general discussion of the correspondence between hypersurfaces and systems of hydrodynamic type. Let $$ h(u)_t=g(u)_x $$ be a conservation law of system (\ref{w}). \begin{theorem} Congruence $S(u)$ is the congruence of Ribaucour if and only if $R(u)$ is representable in the form $$ R(u)=\frac{h(u)}{g(u)} $$ for some conservation law of system (\ref{w}). \end{theorem} In the case $n=2$ this result (stated in a somewhat different form) can be found in \cite{Eisenhart}. It should be emphasized that this theorem equally applies to hypersurfaces which do not possess a curvatute-line parametrization (for $n=2$ such parametrization is always possible). We hope to present the details elsewhere. \section{Acknowledgements} I would like to thank M.V.~Pavlov for usefull discussions. This research was partially supported by INTAS 96-0770 and the RFFI grant 99-01-00010.
\section{Introduction} Consider the stationary Schr\"{o}dinger operator \begin{equation} H_0=-\Delta +q(x) \label{H0} \end{equation} in $L^2(\@subst@obsolete{amsfonts}\bf\mathbb{R}^n)$ ($n=2$ or $3$), where the real potential $q(x)\in L^\infty (\@subst@obsolete{amsfonts}\bf\mathbb{R}^n)$ is periodic with respect to the integer lattice $\@subst@obsolete{amsfonts}\bf\mathbb{Z% }^n$: $q(x+l)=q(x)$ for all $l\in \@subst@obsolete{amsfonts}\bf\mathbb{Z}^n$, $x\in \@subst@obsolete{amsfonts}\bf\mathbb{R}^n$. The spectrum of this operator is absolutely continuous and has the well known band-gap structure (see \cite{E}, \cite{Gelf}, \cite{K}, \cite{OK}, \cite{RS}, \cite {T}, \cite{W}): \[ \sigma (H_0)=\stackunder{i\geq 1}{\cup }[a_i,b_i], \] where $a_i<b_i$, and $\lim\nolimits_{i\rightarrow \infty }a_i=\infty $. Let us introduce also a perturbed operator with an impurity potential $v(x)$: \begin{equation} H=H_0+v(x)=-\Delta +q(x)+v(x), \label{H} \end{equation} where $v(x)$ is compactly supported or sufficiently fast decaying at infinity (the exact conditions will be introduced later). It is known (see \cite{B2}, Section 18 in \cite{G}, and references therein) that the continuous spectrum of $H$ coincides with the spectrum of $H_0$, and only some additional ``impurity'' point spectrum $\{\lambda _j\}$ can arise. A general understanding is that eigenvalues $\lambda _j$ should normally arise only in the gaps of the continuous spectrum (i.e. in the gaps of $\sigma (H_0)$). In other words, the case of embedded eigenvalues when one of the eigenvalues $\lambda _j$ belongs to one of the open segments $(a_i,b_i)$ is prohibited (at least for a sufficiently fast decaying impurity potential $% v(x)$). Physically speaking, the existence of an embedded eigenvalue means a strange situation when an electron is confined, in spite of having enough energy for propagation. There are known examples of embedded eigenvalues, first of which was suggested by J. von Neumann and E. Wigner (see \cite{EK} and section XIII.13 in \cite{RS} for discussion of this topic and related references). The problem of the absence of embedded eigenvalues has been intensively studied. There are many known results on non-existence of embedded eigenvalues for the case of zero underlying potential $q(x)$ (see, for instance, books \cite{EK}, \cite{H} Section 14.7, and \cite{RS} Section XIII.13). The case of a periodic potential $q(x)$ is much less studied. In \cite{EK} a one-dimensional example is provided where a not very fast decaying perturbation of a periodic potential creates embedded eigenvalues. There are many papers devoted to studying the behavior of the point spectrum in the gaps of the continuous spectrum (see, for instance, \cite{AADH} - \cite{B3}, \cite{D}, \cite{GS}, \cite{He}, \cite{L}-\cite{L3}, \cite{Ra}, and \cite{RB} - \cite{S}). However, apparently the only known result on the absence of embedded eigenvalues relates to the one-dimensional case of the Hill's operator \[ H_0=-\frac{d^2}{dx^2}+q(x)+\nu (x). \] (see \cite{RB}, \cite{RB2}). The purpose of this paper is to address the multi-dimensional case. Section 2 contains some necessary notions, auxiliary information, and our main condition on the periodic potential. It is conjectured that the condition is always satisfied. According to \cite{BKT}, this condition is satisfied at least when the potential is separable in 2D or of the form $q_1(x_1)+q_2(x_2,x_3)$ in 3D. In Section 3 we prove a conditional statement (in dimension less than four) on the absence of embedded eigenvalues for a periodic potential satisfying this condition. Finally, the Section 4 contains the main unconditional result. In our considerations we follow an approach that was suggested long ago by the second author for the case of zero underlying potential $q(x)$ (see \cite{V}% ). Some recent developments made it possible to adjust this method to the periodic case. In particular, results on allowed rate of decay of solutions obtained in \cite{FH} and \cite{M} play a crucial role. A more limited version of the main result was announced by the authors without proof in \cite{KV}. \section{\label{Bloch}Bloch and Fermi varieties} Let us introduce some notions and notations. We assume that $q(x)\in L^\infty (\@subst@obsolete{amsfonts}\bf\mathbb{R}^n)$ is a periodic potential. \begin{definition} The (complex) \textbf{Bloch variety} $B(q)$ of the potential $q$ consists of all pairs $(k,\lambda )\in \@subst@obsolete{amsfonts}\bf\mathbb{C}^{n+1}$ ( where $k\in \@subst@obsolete{amsfonts}\bf\mathbb{C}^n$ is a \textbf{quasimomentum} and $\lambda \in \@subst@obsolete{amsfonts}\bf\mathbb{C}$ is an eigenvalue) for which there exists a non-zero solution of the equation \begin{equation} H_0u=\lambda u \label{spect_0} \end{equation} satisfying the so called Floquet-Bloch condition: \begin{equation} u(x+l)=e^{ik\cdot l}u(x),\;l\in \@subst@obsolete{amsfonts}\bf\mathbb{Z}^n,\;x\in \@subst@obsolete{amsfonts}\bf\mathbb{R}^n. \label{Fl} \end{equation} Here $k\cdot l=\sum k_il_i$ is the standard dot-product. \end{definition} \begin{definition} The projection $F_c(q)$ onto $\@subst@obsolete{amsfonts}\bf\mathbb{C}^n$ of the intersection of the Bloch variety $B(q)$ with the subspace $\lambda =c$ is called the (complex) \textbf{Fermi variety}\textit{\ of} the potential $q$ at the level of energy $c$: \[ F_c(q)=\left\{ k\in \@subst@obsolete{amsfonts}\bf\mathbb{C}^n|\;(k,c)\in B(q)\right\} . \] In other words, $F_c(q)$ consists of all quasimomenta $k\in \@subst@obsolete{amsfonts}\bf\mathbb{C}^n$ for which there exists a non-zero solution of the equation \[ H_0u=cu \] satisfying (\ref{Fl}). \end{definition} When $c=0$ we will use the notation $F(q)$ for $F_0(q)$. Let us notice the following obvious relation between the Bloch and Fermi varieties: \begin{equation} \left\{ (k,c)\in B(q)\right\} \Longleftrightarrow \left\{ k\in F_c(q)\right\} \Longleftrightarrow \left\{ k\in F(q-c)\right\} . \label{equiv} \end{equation} We will use the following notations for the real parts of the above varieties: \[ B_{\@subst@obsolete{amsfonts}\bf\mathbb{R}}(q)=B(q)\cap \@subst@obsolete{amsfonts}\bf\mathbb{R}^{n+1},F_{\@subst@obsolete{amsfonts}\bf\mathbb{R},\@subst@obsolete{amsfonts}\bf\mathbb{\lambda }% }(q)=F_{\,\lambda }(q)\cap \@subst@obsolete{amsfonts}\bf\mathbb{R}^n. \] The varieties $B_{\@subst@obsolete{amsfonts}\bf\mathbb{R}}(q)$ and $F_{\@subst@obsolete{amsfonts}\bf\mathbb{R},\lambda }(q)$ are called the \textbf{real Bloch variety} and the \textbf{real Fermi variety} respectively. The following statement is contained in Theorems 3.17 and 4.4.2 of \cite{K} (with remarks of Section 3.4.D in \cite{K} about conditions on potentials taken into account). \begin{lemma} \label{analyt}The Bloch variety $B(q)\subset \@subst@obsolete{amsfonts}\bf\mathbb{C}^{n+1}$ is the set of all zeros of a non-zero entire function $f(k,\lambda )$ on $\@subst@obsolete{amsfonts}\bf\mathbb{C}^{n+1}$ of order $n$ (see \cite{Lel}), i.e. \[ \left| f(k,\lambda )\right| \leq C_{\,\epsilon }\exp \left( (\left| k\right| +\left| \lambda \right| )^{n+\epsilon }\right) ,\;\forall \;\epsilon >0 . \] The Fermi variety $F_{\,\lambda }(q)$ is the set of all zeros of a non-zero entire function of order $n$ on $\@subst@obsolete{amsfonts}\bf\mathbb{C}^n$. \end{lemma} Lemma \ref{analyt} implies in particular that both Bloch and Fermi varieties are examples of what is called in complex analysis \textbf{analytic sets} (see for instance \cite{Ch}, \cite{GR}, and \cite{N}). Moreover, these are \textbf{principal} analytic sets in the sense that they are sets of all zeros of single analytic functions, while general analytic sets might require several analytic equations for their (local) description. Analyticity of these varieties (without estimates on the grows of the defining function) was obtained in \cite{W}. \begin{definition} An analytic set $A\subset \@subst@obsolete{amsfonts}\bf\mathbb{C}^m$ is said to be \textbf{irreducible}, if it cannot be represented as the union of two proper analytic subsets. \end{definition} Irreducibility of the zero set of an analytic function can be understood as absence of non-trivial factorizations of this function (i.e., of a factorization into analytic factors that have smaller zero sets). \begin{definition} A point of an analytic set $A\subset \@subst@obsolete{amsfonts}\bf\mathbb{C}^m$ is said to be \textbf{regular% }, if in a neighborhood of this point the set $A$ can be represented as an analytic submanifold of $\@subst@obsolete{amsfonts}\bf\mathbb{C}^m$. The set of all regular points of $A$ is denoted by $regA$. \end{definition} We collect in the following lemma several basic facts about analytic sets that we will need later. The reader can find them in many books on several complex variables. In particular, all these statements are proven in sections 2.3, 5.3, 5.4, and 5.5 of Chapter 1 of \cite{Ch}. \begin{lemma} \label{property}Let $A$ be an analytic set. a) The set $regA$ is dense in $A$. Its complement in $A$ is closed and nowhere dense in $A$. b) The set $A$ can be represented as a (maybe infinite) locally finite union of irreducible subsets \[ A=\stackunder{i}{\cup }A_i \] called its \textbf{irreducible components}. c) Irreducible components are closures of connected components of $regA$. In particular, set $A$ is irreducible if and only if $regA$ is connected. d) Let $A$ be irreducible and $A_1$ be another analytic set such that $A\cap A_1$ contains a non-empty open portion of $A$. Then $A\subset A_1$. In particular, if $f$ is an analytic function that vanishes on an open portion of $A$, then $f$ vanishes on $A$. e) Any analytic set $A$ has a stratification $A=\cup A_j$ into disjoint complex analytic manifolds (strata) $A_j$ such that the union $\cup A_j$ is locally finite, the closure $\overline{A_j}$ of each $A_j$ and its boundary $% \overline{A_j}\backslash A_j$ are analytic subsets, and such that if the intersection $A_j\cap \overline{A_k}$ of two different strata is not empty, then $A_j\subset \overline{A_k}$ and $dimA_j<dimA_k$. \end{lemma} We will need the following simple corollary from this lemma. \begin{corollary} \label{cor}Let $A$ be an irreducible proper analytic subset of $\@subst@obsolete{amsfonts}\bf\mathbb{C}^n$ such that the intersection $A\cap \@subst@obsolete{amsfonts}\bf\mathbb{R}^n$ contains an open part of a smooth $(n-1)$-dimensional submanifold $M\subset \@subst@obsolete{amsfonts}\bf\mathbb{R}^n$. If $f$ is an analytic function in a complex neighborhood of $M$ such that $f=0$ on $M$, then $f=0$ on an open subset of $A$. In particular, if $B$ is another analytic subset of $\@subst@obsolete{amsfonts}\bf\mathbb{C}^n$ such that $M\subset B$, then $A\subset B$. \end{corollary} \textbf{Proof}. Considering stratification $A=\cup A_j$ (see the last statement in the preceding lemma), one can conclude that only strata of dimension $n-1$ can contain open pieces of $\dot{M}$. In fact, if there is an $A_j$ of complex dimension at most $n-2$ containing an open piece of $M$, we get the contradiction as follows. Consider a point of $M$ and the tangent space to $A_j$ at this point. This is a complex linear subspace of complex dimension at most $n-2$, which contains a real subspace (tangent space to $M$% ) of real dimension $n-1$. This is obviously impossible. So, now we can assume that instead of $A$ we are dealing with one of its strata of dimension $n-1$, i.e. we may assume that $A$ is smooth. In appropriate analytic coordinates in a complex neighborhood of a point of $M$ one can represent $A$ locally as a complex hyperplane with the real part $M$. Then standard uniqueness theorem for analytic functions implies that any function $f$ analytic on $A$ and vanishing on $M$ is identically equal to zero on $A$% . If now $B$ is an analytic subset of $\@subst@obsolete{amsfonts}\bf\mathbb{C}^n$ such that $M\subset B$, then each of the analytic functions locally defining $B$ has the properties of the function $f$ above. Hence, we conclude that $B$ contains an open part of $A$. Then statement d) of the lemma implies that $A\subset B$. This finishes the proof of the corollary. After this brief excursion into complex analysis we return now to Bloch and Fermi varieties. The next statement follows by inspection of (\ref{Fl}): \begin{lemma} The sets $B(q)$ and $F_{\,\lambda }(q)$ are periodic with respect to the quasimomentum $k$ with the lattice of periods $2\pi \@subst@obsolete{amsfonts}\bf\mathbb{Z}^n\subset \@subst@obsolete{amsfonts}\bf\mathbb{C}% ^n$ (i.e., the dual lattice to $\@subst@obsolete{amsfonts}\bf\mathbb{Z}^n$). \end{lemma} We choose as a fundamental domain of the group $2\pi \@subst@obsolete{amsfonts}\bf\mathbb{Z}^n$ acting on $% \@subst@obsolete{amsfonts}\bf\mathbb{R}^n$ the following set called the (first) \textbf{Brillouin zone}: \[ B=\left\{ k=(k_1,...,k_n)\in \@subst@obsolete{amsfonts}\bf\mathbb{R}^n|\;0\leq k_j\leq 2\pi ,\;j=1,...,n\right\} . \] It has been known for a long time (though not always formulated in these terms) that one can tell where the spectrum $\sigma (H_0)$ lies by looking at the Fermi variety. We will now briefly remind the reader this classical result and, at the same time, the definition of spectral bands. Consider the problem (\ref{spect_0}) - (\ref{Fl}). It has a discrete real spectrum $\{\lambda _j(k)\}$, where $\lambda _j\rightarrow \infty $ as $% j\rightarrow \infty $. We number the eigenvalues in the increasing order. This way we obtain a sequence of continuous functions $\lambda _j(k)$ on the Brillouin zone $B$. They are usually called \textbf{band functions}, or branches of the \textbf{dispersion relation}. The values of the function $% \lambda _j(k)$ for a fixed $j$ span the $j$th band $[a_j,b_j]$ of the spectrum $\sigma (H_0)$. A reformulation of this statement is the following theorem, which can be found in equivalent forms in \cite{E} (Section 6.6), \cite{Gelf}, \cite{K} (Theorem 4.1.1), \cite{OK}, and \cite{RS} (Theorem XII.98). \begin{theorem} A point $\lambda $ belongs to the spectrum $\sigma (H_0)$ of the operator $% H_0=-\Delta +q(x)$ if and only if the real Fermi variety $F_{\@subst@obsolete{amsfonts}\bf\mathbb{R},\lambda }(q)$ is non-empty. \end{theorem} In other words, by changing $\lambda $ one observes the Fermi variety $% F_{\,\lambda }(q)$ and notices the moments when it touches the real subspace. This set of values of $\lambda $ is the spectrum. Now the question arises how can one distinguish the interiors of the spectral bands. The natural idea is that when $\lambda $ is in the interior of a spectral band, then the real Fermi variety will be massive. ``Massive'' means here ``of dimension $n-1$'', i.e. of maximal possible dimension for a proper analytic subset in $\@subst@obsolete{amsfonts}\bf\mathbb{R}^n$. This is confirmed by the following statement. \begin{lemma} \label{big}If $\lambda $ belongs to the interior of a spectral band, then the real Fermi variety $F_{\@subst@obsolete{amsfonts}\bf\mathbb{R},\lambda }(q)$ contains an open piece of a submanifold $M$ of dimension $n-1$ in $\@subst@obsolete{amsfonts}\bf\mathbb{R}^n$. \end{lemma} Proof of the lemma follows from the stratification of the real Fermi variety $F_{\@subst@obsolete{amsfonts}\bf\mathbb{R},\lambda }(q)$ into smooth manifolds (see for instance propositions 17 and 18 of the Chapter V in \cite{N}) and from the obvious remark that when $\lambda $ belongs to the interior of a spectral band, then $F_{\@subst@obsolete{amsfonts}\bf\mathbb{R},\lambda }(q)$ must separate $\@subst@obsolete{amsfonts}\bf\mathbb{R}^n$. These two observations imply existence of a smooth piece in $F_{\@subst@obsolete{amsfonts}\bf\mathbb{R},\lambda }(q)$ of dimension at least $(n-1)$. Now we introduce our basic condition: \begin{condition} \label{condition}Assume that for any $\lambda $ that belongs to the interior of a spectral band of the Schr\"{o}dinger operator \[ H_0=-\Delta +q(x) \] any irreducible component of the Fermi variety $F_{\,\lambda }(q)$ intersects the real space $\@subst@obsolete{amsfonts}\bf\mathbb{R}^n$ by a subset of dimension $n-1$ (i.e. by a subset that contains a piece of a smooth hypersurface). \end{condition} In fact, the Lemma \ref{big} says that for $\lambda $ in the interior of a spectral band the Fermi variety $F_{\,\lambda }(q)$ does intersect the real space $\@subst@obsolete{amsfonts}\bf\mathbb{R}^n$ over a subset of dimension $n-1$. We, however, need more, that every irreducible component of $F_{\,\lambda }(q)$ does the same. In other words, there are no ``hidden'' components that do not show up on the real subspace in any significant way. Thus, Lemma \ref{big} and irreducibility of $F_{\,\lambda }(q)/2\pi \@subst@obsolete{amsfonts}\bf\mathbb{Z}^n$ would imply validity of the Condition \ref{condition}. In fact, we believe that (modulo the action of the dual lattice) the Fermi surface is irreducible. The following conjecture formulated in \cite{BKT} is probably correct: \begin{conjecture} \label{conj1}For any periodic potential (from an appropriate functional class, for instance continuous, or locally square integrable) the surface $% F_{\,\lambda }(q)/2\pi \@subst@obsolete{amsfonts}\bf\mathbb{Z}^n$ is irreducible. \end{conjecture} It looks like this conjecture is very hard to prove (see related discussion in \cite{BKT} and \cite{KT}). As the rest of the paper shows, proving it would lead to a result on the absence of embedded eigenvalues. The following weaker conjecture must be easier to prove: \begin{conjecture} \label{conj2}A generic periodic potential (from an appropriate functional class) satisfies the Condition \ref{condition}. \end{conjecture} The word ``generic'' could mean ``from a residual set'', or something of this sort. One can easily prove using separation of variables and simple facts about the Hill's equation irreducibility of $F_{\,\lambda }(q)/2\pi \@subst@obsolete{amsfonts}\bf\mathbb{Z}^n$ for separable periodic potentials $q(x)=\sum_iq_i(x_i)$ in any dimension \cite {BKT}. The paper \cite{BKT}, however, also contains a stronger non-trivial result: \begin{lemma} \label{irred} (\cite{BKT}) If $d=3$ and $q(x)=q_1(x_1)+q_2(x_2,x_3)$, then $% F_{\,\lambda }(q)/2\pi \@subst@obsolete{amsfonts}\bf\mathbb{Z}^n$ is irreducible. \end{lemma} The proof of this lemma relies on the deep study of the Bloch variety done in \cite{KT}. \section{The conditional result} We are ready to prove our main conditional statement. \begin{theorem} \label{cond}If a real periodic potential $q(x)\in L^\infty (\@subst@obsolete{amsfonts}\bf\mathbb{R}^n)$ ($% n\leq $ $3$) satisfies the Condition \ref{condition} and an impurity potential $v(x)$ is measurable and satisfies the estimate \begin{equation} \left| \nu (x)\right| \leq Ce^{-\left| x\right| ^r},r>4/3 \label{estimate} \end{equation} almost everywhere in $\@subst@obsolete{amsfonts}\bf\mathbb{R}^n$, then the spectrum of $H$ contains no embedded eigenvalues. In other words, \[ \left\{ \lambda _j\right\} \cap \stackunder{i\geq 1}{\cup }% (a_i,b_i)=\emptyset , \] where $\left\{ \lambda _j\right\} $ is the impurity point spectrum of $H$, and \[ \stackunder{i\geq 1}{\cup }[a_i,b_i]=\sigma (H_0) \] is the band structure of the essential spectrum of $H$. \end{theorem} \textbf{Proof}. Let us assume that there exists a $\lambda $ that belongs to some $(a_i,b_i)$ and to the point spectrum of $H$ simultaneously. Then there exists a non-zero function $u(x)\in L_2(\@subst@obsolete{amsfonts}\bf\mathbb{R}^n)$ (an eigenfunction) such that \[ -\Delta u+qu+vu=\lambda u, \] or \[ (H_0-\lambda )u=-vu. \] Let us denote the function in the right hand side by $\psi (x):$% \[ \psi (x)=-v(x)u(x). \] Consider the fundamental domain $\mathcal{K}$ of the group $\@subst@obsolete{amsfonts}\bf\mathbb{Z}^n$ of periods: \[ \mathcal{K}=\{(x_1,...,x_n)\in R^n|\;0\leq x_i\leq 1,\;i=1,...,n\}. \] Then (\ref{estimate}) implies that the function $\psi (x)$ satisfies the estimate \begin{equation} \left| \left| \psi \right| \right| _{L^2(\mathcal{K}+l)}\leq Ce^{-\left| l\right| ^r} \label{decay} \end{equation} for all $l\in \@subst@obsolete{amsfonts}\bf\mathbb{Z}^n$, where $r>4/3$. Consider the following Floquet transform of functions defined on $\@subst@obsolete{amsfonts}\bf\mathbb{R}^n$ (see section 2.2 in \cite{K} about properties of this transform): \[ \mathcal{F}:\;f(x)\longrightarrow \widehat{f}(k,x)=\sum\limits_{l\in \@subst@obsolete{amsfonts}\bf\mathbb{Z}% ^n}f(x-l)e^{-ik\cdot (x-l)}. \] The result is a function, which is $\@subst@obsolete{amsfonts}\bf\mathbb{Z}^n$-periodic with respect to $x$. We can consider the resulting function as a function of $k$ with values in a space of functions on the $n$-dimensional torus $\@subst@obsolete{amsfonts}\bf\mathbb{T}^n=\@subst@obsolete{amsfonts}\bf\mathbb{R}^n/\@subst@obsolete{amsfonts}\bf\mathbb{Z}% ^n$. If we apply this transform to the function $\psi (x)$, then we get a function of $k\in \@subst@obsolete{amsfonts}\bf\mathbb{C}^n$ \begin{equation} \phi (k)=\widehat{\psi }(k,x) \label{funct} \end{equation} with values in the space $L^2(\@subst@obsolete{amsfonts}\bf\mathbb{T}^n)$. \begin{lemma} \label{order}Function $\phi (k)$ is an entire function on $\@subst@obsolete{amsfonts}\bf\mathbb{C}^n$ of the order $s=r/(r-1)<4$. \end{lemma} \textbf{Proof of the lemma}. From (\ref{decay}) we get \[ \left| \left| \psi (x-l)e^{-ik\cdot (x-l)}\right| \right| _{L^2(\mathcal{K}% )}\leq Ce^{C\left| k\right| +Im(k\cdot l)-\left| l\right| ^r}. \] Then \[ \left| \left| \phi (k)\right| \right| _{L^2(\mathcal{K})}\leq Ce^{C\left| k\right| }\stackunder{l\in \@subst@obsolete{amsfonts}\bf\mathbb{Z}^n}{\sum }e^{-0.5\left| l\right| ^r}e^{% Im(k\cdot l)-0.5\left| l\right| ^r}. \] A simple Legendre transform type estimate (finding extremal values of the exponent) shows that \[ e^{Im(k\cdot l)-0.5\left| l\right| ^r}\leq Ce^{C\left| k\right| ^{r/(r-1)},} \] which proves that the function $\phi (k)$ is an entire function of the order $s=r/(r-1)$ in $\@subst@obsolete{amsfonts}\bf\mathbb{C}^n$. The lemma is proven. As it is well known (see e.g. Chapter 2 of \cite{K} or XIII in \cite{RS}), the transform $\mathcal{F}$ leads to the following operator equation with a parameter $k$: \begin{equation} \left( H_0(k)-\lambda \right) \widehat{u}(k)=\phi (k), \label{operator} \end{equation} where \[ \widehat{u}(k)=\left( \mathcal{F}u\right) (k,x) \] is defined for $k\in \@subst@obsolete{amsfonts}\bf\mathbb{R}^n$ and \[ H_0(k)=(i\nabla -k)^2+q(x) \] is considered as an operator on the torus $\@subst@obsolete{amsfonts}\bf\mathbb{T}^n$. According to Theorem 2.2.5 in \cite{K}, \[ \widehat{u}(k)\in L_{loc}^2(\@subst@obsolete{amsfonts}\bf\mathbb{R}^n,L_2(\@subst@obsolete{amsfonts}\bf\mathbb{T}^n)). \] In fact, standard interior elliptic estimates show that \[ \widehat{u}(k)\in L_{loc}^2(\@subst@obsolete{amsfonts}\bf\mathbb{R}^n,H^2(\@subst@obsolete{amsfonts}\bf\mathbb{T}^n)), \] where $H^2(\@subst@obsolete{amsfonts}\bf\mathbb{T}^n)$ is the Sobolev space of order two on $\@subst@obsolete{amsfonts}\bf\mathbb{T}^n$. Theorem 3.1.5 of \cite{K} implies that the operator \[ \left( H_0(k)-\lambda \right) :H^2(\@subst@obsolete{amsfonts}\bf\mathbb{T}^n)\rightarrow L_2(\@subst@obsolete{amsfonts}\bf\mathbb{T}^n) \] is invertible if and only if $k\notin F_{\@subst@obsolete{amsfonts}\bf\mathbb{R},\lambda }(q)$. As it is shown in the proof of Theorem 3.3.1 and in Lemma 1.2.21 of \cite{K} (see also comments in Section 3.4.D on reducing requirements on the potential, in particular Theorem 3.4.2), the inverse operator can be represented as follows: \begin{equation} \left( H_0(k)-\lambda \right) ^{-1}=B(k)/\zeta (k), \label{represent} \end{equation} where $B(k)$ is a bounded operator from $L_2(\@subst@obsolete{amsfonts}\bf\mathbb{T}^n)$ into $H^2(\@subst@obsolete{amsfonts}\bf\mathbb{T}% ^n) $ and $B(k)$ and $\zeta (k)$ are correspondingly an operator and a scalar entire functions of order $n$ in $\@subst@obsolete{amsfonts}\bf\mathbb{C}^n$. Besides, the zeros of $% \zeta (k)$ constitute exactly the Fermi variety $F_{\,\lambda }(q)$. We conclude now that the following representation holds on the set $\@subst@obsolete{amsfonts}\bf\mathbb{R}% ^n\backslash F_{\@subst@obsolete{amsfonts}\bf\mathbb{R},\lambda }(q)$: \[ \widehat{u}(k)=\frac{B(k)\phi (k)}{\zeta (k)}=\frac{g(k)}{\zeta (k)}, \] where $g(k)$ is a $H^2(\@subst@obsolete{amsfonts}\bf\mathbb{T}^n)$-valued entire function of order $n$. Now we need the following auxiliary result: \begin{lemma} \label{divis}Let $Z$ be the set of all zeros of an entire function $\zeta (k) $ in $\@subst@obsolete{amsfonts}\bf\mathbb{C}^n$ and $Z_j$ be its irreducible components. Assume that the real part $Z_{j,\@subst@obsolete{amsfonts}\bf\mathbb{R}}=Z_j\cap \@subst@obsolete{amsfonts}\bf\mathbb{R}^n$ of each $Z_j$ contains a submanifold of real dimension $n-1$. Let also $g(k)$ be an entire function in $\@subst@obsolete{amsfonts}\bf\mathbb{C}^n$ with values in a Hilbert space $H$ such that on the real subspace $\@subst@obsolete{amsfonts}\bf\mathbb{R}^n$ the ratio \[ \widehat{u}(k)=\frac{g(k)}{\zeta (k)} \] belongs to $L_{2,loc}(\@subst@obsolete{amsfonts}\bf\mathbb{R}^n,H)$. Then $\widehat{u}(k)$ extends to an entire function with values in $H$. \end{lemma} \textbf{Proof of the lemma}. Applying linear functionals, one can reduce the problem to the case of scalar functions $g$, so we will assume that $g(k)\in \@subst@obsolete{amsfonts}\bf\mathbb{C}$. According to Lemma \ref{property}, the sets of regular points of components $% Z_j$ are disjoint. Hence, the traces of these sets on $\@subst@obsolete{amsfonts}\bf\mathbb{R}^n$ are also disjoint. Consider one component $Z_j$. The intersection of $regZ_j$ with the real space $\@subst@obsolete{amsfonts}\bf\mathbb{R}^n$ contains a smooth manifold of dimension $(n-1)$. Namely, we know that $Z_{j,\@subst@obsolete{amsfonts}\bf\mathbb{R}}$ contains such a manifold, which we will denote $M_j$. The only alternative to our conclusion would be that the whole $M_j$ sits inside the singular set of $Z_j$. The proof of Corollary \ref{cor} shows that this is impossible, since lower dimensional strata cannot contain any open pieces of $M_j$. Let us denote by $m_j$ the minimal order of zero of function $\zeta (k)$ on $% Z_j$. (We remind the reader that the order of zero of an analytic function at a point is determined by the order of the first non-zero term of the function's expansion at this point into homogeneous polynomials.) Since the condition that an analytic function has a zero of order higher than a given number can be written down as a finite number of analytic equations, one can conclude that the order of zero of $\zeta (k)$ equals $m_j$ on a dense open subset of $Z_j$, whose complement is an analytic subset of lower dimension. As it was explained in the proof of Corollary \ref{cor}, lower dimensional strata cannot contain $(n-1)$-dimensional submanifolds of $\@subst@obsolete{amsfonts}\bf\mathbb{R}^n$. This means that one can find a point $k^j\in Z_{j,\@subst@obsolete{amsfonts}\bf\mathbb{R}}$ such that $Z_{j,\@subst@obsolete{amsfonts}\bf\mathbb{R% }}$ is a smooth hypersurface in a neighborhood $U$ of $k^j$ in $\@subst@obsolete{amsfonts}\bf\mathbb{R}^n$ and such that the order of zero of $\zeta (k)$ on $Z_{j,\@subst@obsolete{amsfonts}\bf\mathbb{R}}$ equals $% m_j $ in this neighborhood. Let us now prove that $g(k)$ has zeros on $Z_{j,% \@subst@obsolete{amsfonts}\bf\mathbb{R}}\cap U$ of at least the same order as $\zeta (k)$. If this were not so, then in appropriate local coordinates $k=(k_1,...,k_n)$ the ratio $% g/\zeta $ would have a singularity of at least the order $k_1^{-1}$, which implies local square non-integrability of the function $\widehat{u}(k)$. This contradicts our assumption, so we conclude that $g(k)$ has zeros on $% Z_{j,\@subst@obsolete{amsfonts}\bf\mathbb{R}}\cap U$ of at least the same order as $\zeta (k)$. Consider the (analytic) set $A$ of all points in $\@subst@obsolete{amsfonts}\bf\mathbb{C}^n$ where $g(k)$ has zeros of order at least $m_j$. We have just proven that the intersection of $A$ with $% Z_j$ contains an $(n-1)$-dimensional smooth submanifold of $R^n$. Then Corollary \ref{cor} implies that $Z_j\subset A$, or $g(k)$ has zeros on $Z_j$ of at least the order $m_j$. Hence, according to Proposition 3 in section 1.5 of Chapter 1 in \cite{Ch} the ratio $g/\zeta $ is analytic everywhere in $\@subst@obsolete{amsfonts}\bf\mathbb{C}^n$ except maybe at the union of subsets of $Z_j$, where the function $\zeta $ has zeros of order higher than $m_j$. This subset, however, is of dimension not higher than $n-2$. Now a standard analytic continuation theorem (see for instance Proposition 3 in Section 1.3 of the Appendix in \cite{Ch}) guarantees that $g/\zeta $ is an entire function. This concludes the proof of the lemma. Returning to the proof of the theorem and using the result of the Lemma \ref {big} and the assumption that the Condition \ref{condition} is satisfied for the potential $q(x)$, we conclude that $\widehat{u}(k)$ is an entire function and that it is the ratio of two entire functions of order at most $% w=\max (n,s)<4$, where $s$ is defined in Lemma \ref{order}. Using (in the radial directions) the estimate of entire functions from below contained in section 8 of Chapter 1 in \cite{Le} (see also Theorem 1.5.6 and Corollary 1.5.7 in \cite{K} or similar results in \cite{Bo}), we conclude that $% \widehat{u}(k)$ is itself an entire function of order $w$ with values in $% H^2(\@subst@obsolete{amsfonts}\bf\mathbb{T}^n)$. Now Theorem 2.2.2 of \cite{K} claims that the solution $u(x) $ satisfies the decay estimate: \[ \left| \left| u\right| \right| _{H^2(K+a)}\leq C_p\exp (-c_p\left| a\right| ^p) \] for any $p<w/(w-1)$, where $K$ is an arbitrary compact in $\@subst@obsolete{amsfonts}\bf\mathbb{R}^n$, and $% a\in \@subst@obsolete{amsfonts}\bf\mathbb{R}^n$. Using standard embedding theorems, we conclude that \[ \left| u(x)\right| \leq C_p\exp (-c_p\left| x\right| ^p),\;\forall \;p<w/(w-1). \] Since $w<4$, one can choose a value $p$ such that \[ 4/3<p<w/(w-1). \] However, Remark 2.6 in \cite{FH} and Theorem 1 in \cite{M} state that there is no non-trivial solution of the equation \[ -\Delta u+qu+vu-\lambda u=0 \] with the rate of decay \[ \left| u(x)\right| \leq C\exp (-c\left| x\right| ^p),\;p>4/3. \] This contradiction concludes the proof of the theorem.% \begin{remark} The condition $p>4/3$ (and hence the dimension restriction $n<4$) is essential for the validity of the result of \cite{FH} and \cite{M}, so this is the place where our argument breaks down for dimensions four and higher even if we require compactness of support of the perturbation potential $v(x) $. The rest of the arguments stay intact (the embedding theorem argument, which also depends on dimension, is not really necessary). \end{remark} \section{Separable potentials} The result of the previous section leads to the problem of finding classes of periodic potentials that satisfy the Condition \ref{condition}. As we stated in conjectures \ref{conj1} and \ref{conj2}, we believe that all (or almost all) of periodic potentials satisfy this condition. Although we were not able to prove these conjectures, as an immediate corollary of the Lemma \ref{irred} and of the Theorems \ref{cond} and \ref{sep} we get the following result: \begin{theorem} \label{combin}If for $n<4$ the background periodic potential $q(x)\in L^\infty (\@subst@obsolete{amsfonts}\bf\mathbb{R}^n)$ is separable for $n=2$ or $q(x)=q_1(x_1)+q_2(x_2,x_3)$ for $n=3$, and the perturbation potential $v(x)$ satisfies the estimate \ref {estimate}, then there are no eigenvalues of the operator $H$ in the interior of the bands of the continuous spectrum. \end{theorem} \section{Comments} 1. We have only proven the absence of eigenvalues embedded into the interior of a spectral band. It is likely that eigenvalues cannot occur at the ends of the bands either (maybe except the bottom of the spectrum), if the perturbation potential decays fast enough. This was shown in the one-dimensional case in \cite{RB} under the condition \[ \int (1+\left| x\right| )\left| v(x)\right| dx<\infty \] on the perturbation potential. On the other hand, if $v(x)$ only belongs to $% L^1$, then the eigenvalues at the endpoints of spectral bands can occur \cite {RB2}. 2. Most of the proof of the conditional Theorem \ref{cond} does not require the unperturbed operator to be a Schr\"{o}dinger operator. One can treat general selfadjoint periodic elliptic operators as well. The only obstacle occurs at the last step, when one needs to conclude the absence of fast decaying solutions to the equation. Here we applied the results of \cite{FH} and \cite{M}, which are applicable only to the operators of the Schr\"{o}dinger type. Carrying over these results to more general operators would automatically generalize Theorem \ref{cond}. The restriction that the dimension $n$ is less than four also comes from the allowed rate of decay stated in the result of \cite{FH} and \cite{M}. 3. It might seem that the irreducibility condition \ref{condition} arises only due to the way the proof is done. We believe that this is not true, and that the validity of the condition is essentially equivalent to the absence of embedded eigenvalues. To be more precise, we conjecture that existence for some $\lambda $ in the interior of a spectral band of an irreducible component $A$ of the Fermi surface such that $A\cap \@subst@obsolete{amsfonts}\bf\mathbb{R}^n=\emptyset $ implies existence of a localized perturbation of the operator that creates an eigenvalue at $\lambda $. As a supporting evidence of this one can consider fourth order periodic differential operators, where the Fermi surface contains four points. In this case one can have $\lambda $ in the continuous spectrum, while some points of the Fermi surface being complex. Then one can use these components of the Fermi surface ``hidden'' in the complex domain to cook up a localized perturbation that does create an eigenvalue at $\lambda $ \cite{P}. \begin{center} ACKNOWLEDGMENTS \end{center} The authors express their gratitude to Professors A. Figotin, T. Hoffmann-Ostenhof, H. Kn\"{o}rrer, S. Molchanov, and V. Papanicolaou for helpful discussions and information. The work of P. Kuchment was partly supported by the NSF Grant DMS 9610444 and by a DEPSCoR Grant. P. Kuchment expresses his gratitude to NSF, ARO, and to the State of Kansas for this support. The work of B. Vainberg was partly supported by the NSF Grant DMS-9623727. B. Vainberg expresses his gratitude to NSF for this support. The content of this paper does not necessarily reflect the position or the policy of the federal government, and no official endorsement should be inferred.
\section{Introduction and main results} \subsection{Universal ${\Bbb R}$-trees as functional spaces} The study of the ${\Bbb R}$-trees and their applications to different areas of geometry and topology has been very intensive in the last years (see a recent review [Bes] and references within). In particular, as it was pointed out by Gromov, the ${\Bbb R}$-trees appear naturally in the asymptotic geometry of hyperbolic metric spaces ([Gr1],[Gr2]). The aim of this paper is to present some explicit constructions related to this observation. Recall that a metric space $T$ is {\it geodesic} (see [GH]) if for every two points $t_1,t_2 \in T$ with the distance $a=d_T(t_1,t_2)$ between them, there exists an isometric inclusion $F:[0,a]\to T$, such that $F(0)=t_1$, $F(a)=t_2$. The image $F([0,a])=[t_1,t_2]$ is called a {\it geodesic segment} joining the points $t_1$ and $t_2$. A metric space $T$ is a {\it real tree} (or ${\Bbb R}$-{\it tree} for short) if any pair of its points can be connected with a unique geodesic segment, and if the union of any two geodesic segments $[x,y], [y,z]\subset T$ having the only endpoint $y$ in common is the geodesic segment $[x,z]\in T$. The {\it valency} of a point $t\in T$ is the cardinal number of the set of connected components of $T\setminus t$. An ${\Bbb R}$-tree is {\it $\mu$-universal} (see [MNO]) if any ${\Bbb R}$-tree with the valency not greater than $\mu$ at every its point can be isometrically embedded into it. We introduce an explicit construction of $\mu$-universal ${\Bbb R}$-trees as spaces of functions. For every cardinal $\mu \ge 2$ let $C_\mu$ be a set such that $\mbox{Card}(C_\mu)=\mu$ if $C_\mu$ is an infinite set and $\mbox{Card}(C_\mu)=\mu-1$ otherwise. Let us assume that each set $C_\mu$ contains $0$ as one of it elements. \begin{definition} \label{def1} Let $A_\mu$ be a set of functions $f:(-\infty,\rho_f) \to C_\mu$ with the following properties: 1)~for every function $f$ there exists $\tau_f\le \rho_f$ such that $f(t)\equiv~0 \, \, \forall t<\tau_f$; 2)~all functions $f$ are ``piecewise--constant'' from the right, i.e. for each $t\in (-\infty,\rho_f)$ there exists $\varepsilon > 0$ such that $f|_{[t,t+\varepsilon]}\equiv \mbox{const.}$ Denote the distance between two functions in $A_\mu$ by \begin{equation} \label{dA} d_{A_\mu}(f_1,f_2)=(\rho_{f_1}-s)+(\rho_{f_2}-s), \end{equation} where $s$ is the {\it separation moment} between the functions $f_1$,$f_2$ (see [PSh]): $$ s=\sup \{ t|f_1(t')=f_2(t') \quad \forall t'<t \}. $$ \end{definition} A relation similar to (\ref{dA}) in [MNO] is called a "railroad-track" equation. For every $\mu$ the space $A_\mu$ is a complete real tree such that every its point has valency $\mu$ (see section 2.1). In particular, if $\mu=2$, i.e. if the set $C_\mu=\{0\}$, the space $A_\mu$ is isometric to the real line. Functional spaces similar to the space $A_\mu$ were also considered in [PSh], [Ber], [Sh]. \begin{theorem} $\operatorname{(cf. [N], [MNO]).}$ \label{tree} Consider any cardinal $\mu \ge 2$. (i) (universality) The metric space $A_\mu$ is a $\mu$-universal ${\Bbb R}$-tree. (ii) (uniqueness) Any complete ${\Bbb R}$-tree with valency $\mu$ at every its point is isometric \hskip0.7truecm to the space $A_\mu$. (iii) (homogeneity) The metric space $A_\mu$ is homogeneous. \end{theorem} This theorem is proved in section 2.2 using simpler arguments than the analogous statements in ([N], [MNO]). Universal ${\Bbb R}$-trees can be also constructed in a slightly different way. \begin{definition} \label{def2} Let $\mu$ be an infinite cardinal and let $\tilde A_\mu$ be a set of functions $f:[0,\rho_f) \to C_\mu$, $\rho_f\ge 0$ with the metric (\ref{dA}) which are piecewise-constant from the right. In particular, if $\rho_f=0$ we consider the function $f=f_{\emptyset}\in \tilde A_\mu$. For $\mu=2^{\aleph_0}$ we set $C_\mu=[0,1)$ and denote $\tilde A_\mu$ simply by $A$ (note that the definition of the space $A$ given in [DP] contains an inaccuracy). \end{definition} Such a representation of universal ${\Bbb R}$-trees is more convenient for applications to asymptotic geometry of hyperbolic spaces. Let us mention that for any infinite cardinal $\mu$ the spaces $A_\mu$ and $\tilde A_\mu$ are isometric due to part (ii) of Theorem \ref{tree} being complete ${\Bbb R}$-trees with the valency $\mu$ at every point. Note that this would fail for $\mu$ finite: in this case the valency of the point $f_{\emptyset} \in \tilde A_\mu$ would be equal to $\mu-1$. \subsection{Isometric embedding at infinity} Let $X$ be a metric space with a distance function $d_X$. Intuitively, structure at infinity of the space $X$ is what is seen if one looks at the space $X$ from an infinitely far point (see [Gr2]). One of the possible ways to treat this notion rigorously is as follows. \begin{definition} \label{embed} A metric space $(T, d_T)$ admits an {\it isometric embedding at infinity} into the space $(X, d_X)$ if there exists a sequence of positive scaling factors $\lambda_i \to \infty$ such that for every point $t \in T$ one may find an infinite sequence $\{x_t^i\}$, $i=1,2,..$ of points in $X$ satisfying the relation \begin{equation} \lim_{i\to\infty}\frac{d_X(x_{t_1}^i, x_{t_2}^i)}{\lambda_i}= d_T(t_1, t_2) \end{equation} for every $t_1,t_2 \in T$. \end{definition} In other words, every point of the space $T$ corresponds to a sequence of points in $X$ tending to infinity, such that the ``normalized'' pairwise distances between the sequences in $X$ tend to the distances between the corresponding points in $T$. Isometric embedding at infinity is a strengthening of the notion of an {\it asymptotic subcone} of a metric space (see [Gr1], [GhH]). \begin{theorem} \label{main} A $2^{\aleph_0}$-universal ${\Bbb R}$ tree can be isometrically embedded at infinity into: (a) a complete simply connected manifold of negative curvature; (b) a non-abelian free group. \end{theorem} As usual, it is assumed in (a) that the curvature of a complete simply connected manifold is separated from zero. In the sequel we refer to such manifolds just as to {\it manifolds of negative curvature}. We prove part (a) in section 3.2. Part (b) of this theorem is proved in section 3.1. We believe that its statement should remain true for any non-elementary hyperbolic group. Note that eventhough there exists a bilipschitz embedding of a non-abelian free group into a non-elementary hyperbolic group (see [Kap]), this is not enough to generalize (b) since isometric embeddings at infinity are not (at least, automatically) preserved by bilipschitz mappings. Let us mention that Theorem \ref{main} gives a complete description of metric spaces admitting an isometric embedding at infinity into negatively curved manifolds or non-abelian free groups: any such (geodesic) space is a subtree of a $2^{\aleph_0}$-universal ${\Bbb R}$-tree (cf. [Gr1], [GhH]). \subsection{Asymptotic cones} Let us recall the notion of an asymptotic cone of a metric space. A {\it non-principal ultrafilter $\omega$ } is a finitely additive measure on subsets of ${\Bbb N}$ such that each subset has measure either $0$ or $1$ and all finite subsets have measure $0$. For any bounded function $h : {\Bbb N}\to {\Bbb R}$ its limit $h(\omega)$ with respect to a non-principal ultrafilter $\omega$ is uniquely defined by the following condition: for every $\varepsilon>0$ $$\omega(\{i\in I| \, |h(i)-h(\omega)\|<\varepsilon\})=1.$$ \begin{definition} (See [Gr2], [DrW], [Dru]). Fix a sequence of basepoints $O_i\in X$ and a sequence of scaling factors $\lambda_i \to \infty$, $\lambda_i \in {\Bbb R}$. Consider a set of sequences $g:{\Bbb N}\to X$ such that $d_X(O_i, g(i))\le \mbox{const}_g \cdot \lambda_i$. To any pair of such sequences $g_1,g_2$ one may correspond a function $$h_{g_1,g_2}(i)=\frac{d_X(g_1(i),g_2(i))}{\lambda_i}.$$ We say that the sequences $g_1$, $g_2$ are equivalent if the limit $h_{g_1,g_2}(\omega)=0$. The set $T$ of all equivalence classes with the distance $d_T(g_1,g_2)=h_{g_1,g_2}(\omega)$ is an {\it asymptotic cone} of $X$ with respect to the non-principal ultrafilter $\omega$, sequence of basepoints $O_i$ and scaling factors $\lambda_i$. We denote it by $T=\mbox{Con}_{\omega}(X,O_i,\lambda_i)$. \end{definition} It was proved by M.~Gromov that any asymptotic cone of a hyperbolic metric space is an ${\Bbb R}$-tree (see [Gr1],[Gr2]) \begin{theorem} \label{con} Any asymptotic cone of a manifold of negative curvature is isometric to the $2^{\aleph_0}$-universal ${\Bbb R}$-tree. \end{theorem} We prove this theorem in section 2.3. Theorem \ref{con} together with Theorem \ref{main} imply that any asymptotic cone of a manifold of negative curvature can be isometrically embedded at infinity into such manifold. It follows from Theorem \ref{con} that asymptotic cones of manifolds of negative curvature depend neither on a manifold nor an ultrafilter or the sequence of scaling factors. The statement of Theorem \ref{con} is folklore for any non-elementary hyperbolic group (cf. [KapL], [Dru]). Let us note, however, that in general asymptotic cones do depend on the choice of ultrafilters: in particular, there exists a finitely generated group with non-homeomorphic asymptotic cones corresponding to different ultrafilters (see [ThV]). Main results of the present paper were announced in [DP]. \section{Universal ${\Bbb R}$-trees and asymptotic cones} \subsection{Properties of the space $A_\mu$} \begin{lemma} \label{v} The space $A_\mu$ is a real tree such that the valency of every its point is~$\mu$. \end{lemma} \noindent {\bf Proof.} It is easy to see that the space $A_\mu$ with the metric (\ref{dA}) is a real tree. Let us prove that the valency of each point is $\mu$. Consider a function $f:(-\infty,\rho_f) \to C_\mu$ in $A_\mu$. For any $c \in C_\mu$ take a set of functions $f_\delta:(-\infty,\rho_f+\delta) \to C_\mu$ such that $f_\delta(t)= f(t)$ if $\quad -\infty < t < \rho_f$ and $f_\delta(t)\equiv c$ if $\quad \rho_f \le t < \rho_f+\delta.$ For different constants $c$ we get non--intersecting rays in $A_\mu$ starting from the point $f$ (note that if $\mu$ is finite the total number of rays emanating from $f$ is exactly $\operatorname{Card}(C_\mu)+1=\mu$). Since $A_\mu$ is a real tree, these rays lie in different connected components of $A_\mu\setminus f$. Therefore, $A_\mu$ has valency $\mu$ at every its point. \qed \begin{lemma} \label{c} The metric space $A_\mu$ is complete. \end{lemma} \noindent {\bf Proof.} Consider a fundamental sequence $f_i:(-\infty,\rho_{f_i})\to C_\mu$ in $A_\mu$. Then $\rho_i=\rho_{f_i}$ is also fundamental, since $d(f_i,f_j) \ge |\rho_i - \rho_j|$. Denote $\rho=\lim_{i\to \infty} \rho_i$. Then $\lim_{i,j \to \infty}s_{ij}=\rho$, where $s_{ij}$ is a separation moment of functions $f_i,f_j$. Indeed, $d_{A_\mu}(f_i, f_j)=(\rho_i - s_{ij})+(\rho_j - s_{ij})$ and hence $$ s_{ij}=\frac{\rho_i + \rho_j - d_{A_\mu}(f_i,f_j)}{2} \to \rho. $$ Therefore for each $\rho' < \rho$ the functions $f_i$ coincide for sufficiently large $i$ on the interval $(-\infty,\rho')$. Define $f$ by $f(x)=\lim_{i\to \infty} f_i(x)$ for any $x \in (-\infty,\rho)$. Note that $f \in A_\mu$, since for any $x \in (-\infty,\rho)$ there exist $I\in {\Bbb N}$ and $\varepsilon >0$, such that $f|_{(-\infty,x+\varepsilon]}\equiv f_I|_{(-\infty,x+\varepsilon]}$ and hence $f$ is ``piecewise--constant'' from the right. \qed \subsection{Proof of Theorem \ref{tree}} (i) Let $\Psi$ be a tree of cardinality $card(\Psi)\le \mu$. We say that a subtree $T\subset \Psi$ is {\it good} if it satisfies the following two properties: a) $T$ is nonempty and connected; b)$\forall t \in T$ either $T$ has a nonempty intersection with all connected components of the set $\Psi\setminus t$ (in this case we say that $t$ is {\it full-valented}), or $T$ has a non--empty intersection with at most $2$ connected components of $\Psi\setminus t$ (then we call $t$ {\it few--valented}). Note that union of the nested good subtrees is also a good subtree. Consider pairs $P=(T,F)$ such that $T$ is a good subtree and $F:T \to A_\mu$ is an isometric inclusion. Let us introduce a partial order on such pairs: $(T_1,F_1) \le (T_2,F_2)$ if $T_1 \subset T_2$ and $F_2|_{T_1}=F_1$. The set of such pairs is nonempty since it contains all inclusions of a one-point set. Any linearly ordered set of pairs $P_\alpha=(T_\alpha,F_\alpha)$ has a maximal element $(T',F')$, where $T'=\cup T_\alpha$ and $F'|_{T_{\alpha}}=F_{\alpha}$. Then by Zorn's lemma there exists a maximal element in the whole set of pairs: $P_{max}=(T_{max}, F_{max})$. Let us show that $T_{max}=\Psi$. If not, there exists $\gamma \in \Psi\setminus T_{max}$. Take an arbitrary $t \in T_{max}$ and consider the geodesic segment $[t,\gamma]$. The intersection $[t,\gamma]\cap T_{max}$ is either a closed segment $[t,t']$ or a semi-interval $[t,t')$. Consider each of these cases separately. {\it Case 1.} $T_{max}\cap [t,\gamma]=[t,t').$ Take $T_{max}'=T_{max}\cup t'$. It is a good subtree since $T_{max}'$ intersects just with a single connected component of $\Psi\setminus t'$ (since $T_{max}$ is connected), and for the points from $T_{max}$ the condition b) remains true. Due to completeness of $A_\mu$, the isometric inclusion of $T_{max}$ can be continued to the isometric inclusion of $T_{max}'$. This contradicts with the maximality of $P_{max}$. {\it Case 2.} $T_{max} \cap [t,\gamma]=[t,t']$. Then $T_{max}$ does not intersect with the connected component of $\Psi\setminus t'$ containing $\gamma$, therefore $t'$ is few--valented. In each component of $\Psi\setminus t'$ non-intersecting with $T_{max}$ we fix a point $t_\nu$ and consider geodesic segments $[t', t_\nu]$. Take $T_{max}'=\cup_{\nu} [t',t_\nu]\cup T_{max}$. This is a good subtree. Indeed, $t'$ is full-valented. For the points $T_{max}\setminus t'$ the condition b) remains true, and the points of $T_{max}'\setminus T_{max}=\cup_\nu (t',t_\nu]$ are few--valented. We build the continuation of $F$ to $T'_{max}$. Consider $F(t') \in A_\mu$. The point $F(t')$ is an endpoint for $\mu$ rays lying in different connected components of $A_\mu\setminus F(t')$. Points of $F(T_{max})$ may lie in at most two of them. To every segment $[t',t_\nu]$ we isometrically correspond a segment $[F(t'), F(t_\nu)]$ on one of the ``free'' rays (i.e. having no points of $T_{max}$) of $A_\mu\setminus F(t')$. This is possible since cardinality of the set of the ``free'' rays is always greater or equal than of the set of points $t_\nu$. Once again, we get a contradiction with the maximality of $P_{max}$. Therefore, $T_{max}=\Psi$ and $F_{max}$ is its isometric emebedding into the space $A_\mu$. This completes the proof of (i). \noindent (ii) Let $\Psi$ be a complete real tree such that the valency of every its point has cardinality $\mu$. We call the pair $(T,F)$ {\it good}, if $T$ is a good subtree and $F$ preserves full-valentness. The same as in (i), the set of good pairs is non-empty and one may introduce partial order on it. It may be shown analogously that every linear ordered set of good pairs has a maximal element and hence by Zorn's lemma there exists a maximal pair $(T_{max}, F_{max})$, where $T_{max}=\Psi$. Let us prove that $F_{max}$ is a surjection. If not, there exists $a \in A_\mu$ not lying in $F_{max}(T_{max})$. Take an arbitrary point $p \in F_{max}(T_{max})$ and consider $[a,p] \cap F_{max}(T_{max})$. Since $F_{max}(T_{max})$ is complete (being an isometric image of a complete metric space $T_{max}=\Psi$), it is a closed segment $[a,p']$. But in this case $F_{max}(T_{max})$ intersects not with all connected components of $A_\mu\setminus p'$ while $F^{-1}_{max}(p') \in \Psi$ is clearly full--valented --- we got a contradiction with the fact that $(T_{max}, F_{max})$ is a good pair. Therefore, $F_{max}$ is an isometry between $A_\mu$ and~$\Psi$. \medskip \noindent (iii) Let us note that in the proof of (ii) we could consider only such good pairs $(T,F)$ that satisfy $(T_0, F_0)\le (T,F)$, where $T_0={\gamma_0}$, $F_0(\gamma_0)=a_0$ for some arbitrary fixed $\gamma_0 \in \Psi$ and $a_0 \in A_\mu$. Then finally we would obtain an isometry between $\Psi$ and $A_\mu$ which moves $\gamma_0$ into $a_0$. Setting $\Psi=A_\mu$ implies that $A_\mu$ is homogeneous. \qed \subsection {Asymptotic cones of manifolds of negative curvature} In this section we prove Theorem \ref{con} stating that any asymptotic cone of a manifold of negative curvature is isometric to the space $A$ (see Definition \ref{def2}). \noindent {\bf Proof of Theorem \ref{con}.} Let $O_i\in X$ be some sequence of basepoints and $\lambda_i\to \infty$ be some sequence of scaling factors. An asymptotic cone $\operatorname{Con}_\omega (X,O_i,\lambda_i)$ of a manifold of negative curvature $X$ is a complete geodesic metric space. Let us show that the valency at every its point is $2^{\aleph_0}$. Consider a point $\xi \in \operatorname{Con}_\omega (X,O_i,\lambda_i)$ and let $\{x_i\}$ be one of the sequences in $X$ corresponding to this point. For every $\alpha \in [0,2\pi)$ and for every $\rho>0$ consider a sequence $\{x_i^{\alpha,\rho}\}$ constructed according to the following rule. Consider the geodesic segment $[O_i,x_i]$. The segment $[x_i, x_i^{\alpha,\rho}]$ is of length $\lambda_i\rho$ and the angle between it and $[O_i,x_i]$ is equal to $\alpha$. It clearly follows from the triangle inequality that the distances between $O_i$ and $x_i^{\alpha,\rho}$ grow not faster than linearly. Let us show that the distance between two points of the asymptotic cone $\xi^{\alpha_j,\rho_j}$ corresponding to the sequences $\{x_i^{\alpha_j,\rho_j}\}$, $j=1,2$, is equal to $\rho_1+\rho_2$ if $\alpha_1 \ne \alpha_2$ and $|\rho_1-\rho_2|$ if $\alpha_1=\alpha_2$ The latter is due to the fact that $\alpha_1=\alpha_2$ implies $d_X(x_i^{\alpha_1,\rho_1},x_i^{\alpha_2,\rho_2})=\lambda_i|\rho_1-\rho_2|$. If $\alpha_1 \ne \alpha_2$ we need the following lemmas. \begin{lemma} \label{26} Let $ABC$ and $A'B'C'$ be geodesic triangles in $\Bbb H$ and in $X$ respectively, such that $|AB|=|A'B'|=u$, $|AC|=|A'C'|=v$, $|BC|=w$, $|B'C'|=w'$, $\angle A = \angle A'=\varphi$, where $|\cdot|$ denotes the length of a segment in the corresponding metric. Then $w <w'$. \end{lemma} \noindent {\bf Proof.} Consider a comparison triangle $PQR \in {\Bbb H}$ for $ABC$: $|PQ|=u$, $|PR|=v$, $|QR|=w$ and let $\angle P=\psi$. By Alexandrov comparison theorem (see [GhH) $\psi>\varphi$. On the other hand by hyperbolic cosine rule (see [Bea]) $$\cosh w =\cosh u\cosh v-\cos \varphi \sinh u \sinh v <$$ $$<\cosh u\cosh v-\cos \psi \sinh u \sinh v = \cosh w',$$ and hence $w<w'$. \qed \begin{lemma} \label{nep} Consider a point $O\in X$ and two geodesic rays $l_1$ and $l_2$ starting from $O$ with a non-zero angle between them. Let $A_i\in l_1$ and $B_i\in l_2$ be two sequences of points such that $d_X(O,A_i)=\lambda_i\rho_1$ and $d_X(O,B_i)=\lambda_i\rho_2$, where $\lambda_i \to \infty$. Then \begin{equation} \label{pum} \lim_{i\to \infty}\frac{d_X(A_i,B_i)}{\lambda_i}=\rho_1+\rho_2. \end{equation} \end{lemma} \noindent {\bf Proof.} Let $O'A_i'B_i'$ be the triangle in ${\Bbb H}$ corresponding to $OA_iB_i$ as in lemma \ref{26}. Then due to triangle inequality and lemma \ref{26} we have: $$ \rho_1+\rho_2 \ge \frac{d_X(A_i,B_i)}{\lambda_i} \ge \frac{d_{\Bbb H}(A_i',B_i')}{\lambda_i} \to \rho_1+\rho_2, $$ where the limit at the right follows from a simple asymptotic analysis of distance formulas on the hyperbolic plane. This proves the relation (\ref{pum}). \qed \smallskip Lemma \ref{nep} implies that \begin{equation} \frac{d_X(x_i^{\alpha_1,\rho_1},x_i^{\alpha_2,\rho_2})}{\lambda_i} \to \rho_1+\rho_2 \end{equation} Therefore $\{x_i^{\alpha_1,\rho_1}\}=\{x_i^{\alpha_2,\rho_2}\}$ iff $\alpha_1=\alpha_2$ and $\rho_1=\rho_2$. For every $\alpha \in [0,2\pi)$ we get an infinite ray starting from the point $\xi$ and for different $\alpha$ these rays have the only point $\xi$ in common and hence lie in different connected components of $\operatorname{Con}_{\omega}(X,O_i,\lambda_i)\setminus \xi$, i.e. the valency of the point $\xi$ is $2^{\aleph_0}$. Since $\xi$ is an arbitrary point in the asymptotic cone $\operatorname{Con}_{\omega}(X,O_i,\lambda_i)$ we have proved that it is isometric to the metric space~$A$.~\qed \section{Proof of Theorem \ref{main}} \subsection{Part(b)} We start with the easier part of Theorem \ref{main}, namely, we prove that the space $A$ can be isometrically embedded at infinity into a non-abelian free group. Let us fix some sequence of scaling factors $\lambda_i\to \infty$ (for simplicity we set $\lambda_i=i$). Consider a Cayley graph corresponding to a subgroup generated by two arbitrary generators $\gamma_1,\gamma_2$ of the group $\Gamma$, and let $G$ be its part corresponding to non-negative powers of $\gamma_1,\gamma_2$. Consider the standard coding of vertices of $G$ with binary sequences of $\gamma_1$-s and $\gamma_2$-s taking the unity as the initial vertex. Setting $\gamma_1=0$ and $\gamma_2=1$ we correspond to every vertex of $G$ a binary number in a unique way. For any point $t\in [0,\rho_f)$ let $[t_l, t_r)\ni t$ be the maximal interval where the function $f\in A$ is constant. For every $\varepsilon > 0$ consider the following ``cut-off'' function: $f^\varepsilon(t)=f(t)$ if $t_r-t_l\ge \varepsilon$ and $f^\varepsilon(t)=0$ otherwise. Fix some bijection $F$ from $[0,1)$ to the set of all binary sequences. Our aim is to build a sequence of vertices $\{ x_i \} \in G$ (which are actually elements of the group $\Gamma$) corresponding to an arbitrary function $f\in A$. Consider a function $f^{1/\sqrt{i}}$, and let $l^i_1,...,l^i_{k_i}$ be the lengths of the segments on which it is constant, and $0=a^i_1,...,a^i_{k_i}$ be the values of $f^{1/\sqrt{i}}$ on them. Consider the binary sequence $F(a_1^i)$ and take its first $[l_1\cdot i]$ elements, where $[\cdot]$ denotes the integer part. We get a binary number which corresponds to a unique point $x_1^i \in G$. From the point $x_1^i$ we go along the sequence $F(a_2^i)$ for the time $[l_2\cdot i]$ and so on until $x^i_{k_i}=x_i$. Let us check that $$\frac{d_{G}(x_i,e)}{i} \to \rho_f.$$ Note that $k_i \le 2\sqrt{i}\rho_f+1$ since the number of intervals on which the function $f^{1/\sqrt{i}}$ does not vanish is not greater than $\sqrt{i}\cdot \rho_f$. Therefore $$d_{G}(x_i, e)=\sum_{r=1}^{k_i} [l^i_r\cdot i]\ge \sum_{r=1}^{k_i}(l^i_r \cdot i -1) = \rho_f\cdot i - k_i.$$ Thus, $$\rho_f =\sum_{r=1}^{k_i} l^i_r\ge \frac{d_{G}(x_i,e)}{i}\ge \rho_f-\frac{k_i}{i},$$ and since $k_i/i\to 0$ we get $d_{G}(x_i,e)/i \to \rho_f$. Consider now two arbitrary functions $f,g \in A$. Let $s$ be their separation moment, and let $s_i$ be moments of separation of the segments $[e,x^f_i]$ and $[e,x^g_i]$. Note that for sufficiently small $\varepsilon$ the moment of separation between $f^\varepsilon$ and $g^\varepsilon$ coincides with the separation moment of $f$ and $g$, and $f^\varepsilon|_{[0,s]}=(f|_{[0,s]})^\varepsilon.$ Applying previous arguments to the function $f|_{[0,s]}$ we get that $d_{G}(s_i,e)/i \to s$, since for the sufficiently large $i$ we have $s_i=x_i^{f|_{[0,s]}}.$ Hence $$\frac{d_{G}(x_i^f,x_i^g)}{i}= \frac{d_{G}(x_i^f,e)}{i}+ \frac{d_{G}(x_i^g,e)}{i}- \frac{d_{G}(s_i,e)}{i} \to \rho_f+\rho_g-2s=d_A(f,g).$$ This completes the proof of part (b) of the theorem.\qed \subsection{Part(a)} In this section we prove that the metric space $A$ can be isometrically embedded at infinity into any manifold of negative curvature. First let us mention that in case of the hyperbolic plane ${\Bbb H}$ (or, more general, in case of a complete simply connected manifold of {\it constant} negative curvature) one may isometrically embed the space $A$ into it in a very explicit and simple way. Consider the Poincare disc model of ${\Bbb H}$ and let us introduce the hyperbolic polar coordinates $(r, \varphi)$ centered at some $O\in {\Bbb H}$. To every function $f(t)\ne f_\emptyset \in A$, $f:[0,\rho) \to [0,1)$ we correspond a sequence $x_n=(\rho_i, \varphi_i)\in {\Bbb H}$, where $\rho_i=\rho i$ and $$\varphi_i=2\pi f(0)+\int_0^{\rho}e^{-ti}f(t)dt.$$ To the ``empty function'' $f_\emptyset$ we correspond a constant sequence $x_i=O$. We also set $\lambda_i=i$. Using some asymptotic analysis of distance formulas on the hyperbolic plane one can prove that this is indeed an isometric embedding at infinity. \smallskip Let us prove the theorem in the general case. We divide the proof into several steps. \subsection*{Step 1.} Let us fix a sequence of scaling factors (as before, let $\lambda_i=i$). Let $O \in X$ be a basepoint and $\l_0 \in X$ some fixed geodesic ray emanating from $O$. For every $0 \le \varphi < 2\pi$ let us fix a ray $l_\varphi \subset X$ starting from the point $O$ such that the angle between $\angle(l_0,l_\varphi)=\varphi$. As in the case of the hyperbolic plane, let us correspond to $f_\emptyset$ the constant sequence $x_i=O$. We prove that for any $f:[0,\rho)\to [0,1)$ and for any $i>0$ there exists a unique naturally--parametrized path $\gamma_i: [0,\rho\cdot i] \to X$ with the following properties: \noindent (i)if $a,b$ are such that $f|_{[a,b]}=\operatorname{const.}$ then $\gamma_i|_{[a\cdot i,b\cdot i]}$ is a geodesic segment; \noindent (ii) if $a$ is a point of discontinuity of the function $f$ then the angle between $[O,\gamma_i(a\cdot i)]$ and $[\gamma_i(a\cdot i),\gamma_i((a+\varepsilon)i)]$ is equal to $1/100 + f(a)$, where $\varepsilon$ is such that $f|_{[a,a+\varepsilon]}=\operatorname{const};$ \noindent (iii) if $f|_{[0,a]} \equiv f(0)$ then $\gamma(a)\in l_{2\pi f(0)}$. \subsection*{Step 2.} There is always a unique way to start constructing $\gamma$ due to the property (iii), since for every function there exists a non--zero interval $[0,\varepsilon]$ where it is constant and is equal to $f(0)$. If $\gamma$ is constructed on $[0,r\cdot i)$ it may be continued in a unique way to the point $r$ by setting $\gamma_i(ri)=\lim_{z\to ri-0}\gamma_i(z)$. If $\gamma_i$ is constructed on $[0,r\cdot i]$ for some $r<\rho$ we may take such $\varepsilon>0$ that $f|_{[r,r+\varepsilon]}=\operatorname{const.}$ There are two possible cases: if $r$ is a point of discontinuity of $f$ then we apply (ii) in order to continue $\gamma_i$ up to the point $(r+\varepsilon)i$, and if $f$ is continuous at $r$ we may apply (i). Therefore it is possible to construct a unique path $\gamma_i$ on the whole $[0,\rho i]$. This path is a broken geodesic line with possibly infinite number of links. We set $\{x_i^f\}=\gamma_i(\rho \cdot i)$. Let us prove that the correspondence $f\in A \to \{x_i^f\}\subset X$ satisfies Definition \ref{embed}. \subsection*{Step 3.} Our next aim is to show that for any $f,g\in A$ we have $$\frac{d_X(x_i^f,x_i^g)}{i}\to d_A(f,g).$$ Let us prove first that \begin{equation} \label{sim} \frac{d_X(x_i^f,O)}{i} \to \rho. \end{equation} Denote a function $f_r=f|_{[0,r]}$. Consider the set $$R=\{r\in [0,\rho)|\, \forall r'\le r \quad \frac{d_X(x_i^{f_r},O)}{i} \to r'\}.$$ We want to verify that $R=[0,\rho]$. We do it in the same way as we have shown the existence of the path $\gamma_i$ in the previous Step. Since for some $\varepsilon>0$ $f|_{[0,\varepsilon]}=\operatorname{const.}$ therefore $\varepsilon \in R$. If every $r_1 <r$ belongs to $R$ then $r\in R$. This follows from the triangle inequality: $$ r \ge \frac{d_X(x_i^{f_r}, O)}{i} \ge \frac{d_X(x_i^{f_{r_1}}, O)}{i}- (r-r_1), $$ and since we may choose $r_1$ arbitrary close to $r$ we get $$\frac{d_X(x_i^{f_r}, O)}{i}\to r.$$ Finally, if $r \in R$ for some $r<\rho$ then $r+\varepsilon \in R$ for some $\varepsilon>0$. Indeed, let $r'$ be the left end of the interval containing $r$ on which $f$ is constant. Due to (ii) the angle $\angle ([O, \gamma_i(r')], [\gamma_i(r'), \gamma_i(r+\varepsilon)]) = 1/100 + f(r') > 0$ and hence using Lemma \ref{nep} we obtain that $d_X(x_i^{f_{r+\varepsilon}}, O)/i \to r+\varepsilon$. This completes the proof of the fact that $R=[0,\rho]$. \subsection*{Step 4.} In order to proceed we need the following lemma. \begin{lemma} \label{29} Let $A_iB_iC_i$ be a triangle in $X$ such that $|A_iB_i|=u\cdot i$, $|A_iC_i|=v \cdot i$, $u,v >0$. Consider a path $\zeta_i$ of length $|\zeta_i|$ connecting $C_i$ and $B_i$ such that \begin{equation} \label{cond} \lim_{i\to \infty}\frac{u\cdot i}{|\zeta_i|+v\cdot i}=1. \end{equation} Then $\lim_{i\to\infty}\angle A_i = 0$. \end{lemma} \noindent {\bf Proof.} Suppose that the angle $\angle A_i$ does not tend to zero. Then there exist a subsequence $i_k\to \infty$ such that $\angle A_{i_k}>\alpha>0$. Then due to Lemma \ref{nep} $d_X(B_{i_k},C_{i_k})/i_k \to u+v$. On the other hand, $|\zeta_{i_k}|\ge d_X(B_{i_k},C_{i_k})$ and hence $$\lim_{i_k\to \infty}\frac{|\zeta_{i_k}|+v\cdot i_k}{u\cdot i_k} \ge \frac{u+2v}{u} >1, $$ which contradicts with (\ref{cond}). This completes the proof of Lemma \ref{29}. \qed \subsection*{Step 5.} Consider a function $f\in A$ which is discontinious at the point $s$. We say that it has a {\it discontinuity of the first type} at $s$ if there exists $\delta>0$ such that $f_{[s-\delta,s]} \equiv \operatorname{const.}$ Otherwise we say that $f$ has a {\it discontinuity of the second type} at $s$. Actually, discontinuities of the second type are accumulation points of discontinuities of the first type. Now, let $f,g \in A$ be arbitrary functions and $s$ be their separation moment. At least one of the functions should be discontinuous at $s$, let it be the function $f$. Note that if it has a discontinuity of the first type at $s$, then $g$ is either continuous, or has a discontinuity of the first type, and if $f$ has a discontinuity of the second type at $s$, the function $g$ has also a discontinuity of the second type at this point. One may check in each case that due to the property (ii) and Lemma \ref{29} the paths corresponding to functions $f$ and $g$ for large $i$ separate exactly at the point $x_i^s=x_i^{f_s}=x_i^{g_s}$, and that the angle $\angle([x_i^s,x_i^f],[x_i^s,x_i^g)$ is separated from zero. Indeed, if $g$ is continuous at $s$ then $f$ has a discontinuity of the first type (see Figure 1, left); the angles marked with black are close to zero by Lemma \ref{29}, the angle $\alpha$ by property (ii) is neither $0$ nor $\pi$ and does not depend on $i$. \begin{figure} \noindent\centering\hfill\epsffile{fig1.1}\hfill\epsffile{fig2.1}\hfill \caption{} \end{figure} If $g$ is discontinuous at $s$ (see Figure 1, right) then similarly the angles marked with black are close to zero by Lemma \ref{29} and the angle $\beta$ by property (ii) is neither $0$ nor $\pi$ and does not depend on $i$. Thus it can be shown analogously to (\ref{sim}) that $d_X(O,x_i^s)/i \to~s$, $d_X(x_i^s,x_i^f)/i \to \rho_f-s$, $d_X(x_i^s,x_i^g)/i \to \rho_g-s$. Therefore, applying Lemma \ref{nep} we finally obtain: $$ \lim_{i\to \infty}\frac{d_X(x_i^f,x_i^g)}{i}= \lim_{i\to \infty}\frac{d_X(x_i^f,x_i^s)+d_X(x_i^s,x_i^g)}{i}= (\rho_f-s)+(\rho_g-s)=d_A(f,g), $$ which completes the proof of the theorem. \qed \subsection* {Acknowledgments.} We would like to thank Professor A.~Shnirelman for fruitful discussions which stimulated us to write this paper. We are also grateful to Professors B.~Bowditch, F.~Paulin, L.~Polterovich and O.~Schramm for helpful remarks, and to V.~Kondrat'ev for TeX-ing Figure 1. \section*{References} \noindent [Bea] A. Beardon, The geometry of discrete groups, Springer--Verlag, 1983 \noindent [Ber] V.N. Berestovskii, Quasicones at infinity of Lobavhevski spaces, Announcement on the Maltsev International Algebraic Conference, Novosibirsk, 1989. \noindent [Bes] M. Bestvina, ${\Bbb R}$-trees in topology, geometry, and group theory, Preprint, 1999. \noindent [DrW] L. van den Dries, A.J. Wilkie, On Gromov's theorem concerning groups of polynomial growth and elementary logic, J. of Algebra 89, 1984, 349-374. \noindent [Dru] C. Drutu, R\'eseaux des groupes semisimple et invariants de quasi-isometrie, Ph.D. Thesis, Universit\'e d'Orsay, 1996. \noindent [DP] A.G. Dyubina, I.V. Polterovich, Structures at infinity of hyperbolic spaces, Russian Mathematical Surveys, vol. 53, No. 5, 1998, 239-240. \noindent [GhH] E. Ghys, P. de la Harpe, Sur le groupes hyperboliques apres Mikhael Gromov, Birkh\"auser, 1990. \noindent [Gr1] M. Gromov, Hyperbolic Groups, in: Essays in group theory, ed. S.M.Gersten, M.S.R.I. Publ. 8 , Springer--Verlag, 1987, 75-263. \noindent [Gr2] M. Gromov, Asymptotic invariants of infinite groups, Geometric group theory. Vol. 2 (Sussex, 1991), London Math. Soc. Lecture Note Ser. 182, Cambridge Univ. Press, 1993, 1-295. \noindent [Kap] I. Kapovich, A non-quasiconvexity embedding theorem for hyperbolic groups, Math. Proc. Camb. Phil. Soc. 127 (1999), no. 3, 461-486. \noindent [KapL] M. Kapovich, B. Leeb, On asymptotic cones and quasi-isometry classes of fundamental groups of 3-manifolds, GAFA, vol. 3, No. 5, 1995, 583-603. \noindent [MNO] J. Mayer, J. Nikiel, L. Oversteegen, Universal spaces for ${\Bbb R}$-trees, Trans. AMS, vol. 334, No. 1, 1992, 411-432. \noindent [N] J. Nikiel, Topologies on pseudo-trees and applications, Memoirs AMS, No. 416, 1989. \noindent [PSh] I. Polterovich, A. Shnirelman, An asymptotic subcone of the Lobachevskii plane as a space of functions, Russian Math. Surveys, vol. 52 No. 4, 1997, 842-843. \noindent [Sh] A. Shnirelman, On the structure of asymptotic space of the Lobachevsky plane, Preprint, 1997, 1-23. \noindent [ThV] S. Thomas, B. Velickovic, Asymptotic cones of finitely generated groups, Bull. London. Math. Soc., vol. 32, no. 2, 2000, 203-208. \end{document}
\section{Minimal error probability for the parity bit.} This appendix is about the difficulty of finding the parity bit in quantum cryptography. It completes some work did by Bennett, Mor and Smolin in 1996 \cite{bms96}. The setting is the following. Consider two states $\psi(0)$ and $\psi(1)$ at an angle $\Omega$. We have $m$ bits $a_1 \ldots a_m$ encoded by Alice into the product state $\psi(a_1) \ldots \psi(a_m)$ which is sent to Bob. Bob wants to obtain a guess $A^*$ on $A = \oplus_j a_j$ so that $PE(m) = \Pr(A^* \neq A)$ is minimal. We evaluate exactly the minimum of $PE(m)$ over all possible POVM. Next, we use it to easily bound the probability of obtaining a conclusive outcome $PC(m)$ using $PC(m) \leq 1 - 2 PE(m)$, which comes from the fact that if you obtain a conclusive outcome with probability $PC(m)$ you will guess the parity bit with probability of error $(1 - PC(m))/2$, which by definition must be larger than $PE(m)$. It is easy to determine the POVM that minimises the probability and to actually compute exactly $PE(m) = \Pr(A^* \neq A)$ for that POVM. We only do the case where $m$ is odd. The case where $m$ is even is similar. The states $\psi(0)$ and $\psi(1)$ are conveniently written $\psi(0) = c|0\rangle + s|1\rangle$ and $\psi(1) = c|0\rangle - s |1\rangle$, where $c = \cos(\Omega/2)$ and $s = \sin(\Omega/2)$. The two density matrices for the parity $A = 0$ and $A=1$ respectively are block diagonal (every block is a 2x2 matrix). If we reorganise the order of rows and collums properly, for $k = 0,...., (m-1)/2$, there are ${m \choose k}$ blocks like this one: \[ \left( \begin{array}{cc} c^{2(m-k)} s^{2k} & \pm c^m s^m \\ \pm c^m s^m & c^{2k} s^{2(m-k)} \end{array} \right) \] The plus sign is for $A = 0$ and the negative sign for $A = 1$. In particular, one can easily check that \[ \sum_{k=0}^{(m-1)/2} {m \choose k} [c^{2(m-k)}s^{2k} + c^{2k}s^{2(m-k)}] = 1. \] Therefore, in both cases, for the best guess or to maximise the probability of a conclusive outcome, the best strategy is to first find out in which block we have obtained, and then try to find out the parity given that we have obtained that block. Given that we have one of these blocks the task is easy because every block has the shape \[ \left( \begin{array}{cc} a^2 & \pm ab \\ \pm ab & b^2 \end{array} \right) \] which corresponds to the pure states $\phi_\pm = a|0\rangle \pm b|1\rangle$, where $a = \cos(\Theta_k/2)$ and $b = \sin(\Theta_k/2)$. The angle $\Theta_k$ is the angle between the two states $\phi_+$ and $\phi_-$. Given that we have one of these blocks, the best guess has probability of error $PE = \sin^2(\pi/4 - \Theta_k/2)$. One obtains $PE = (1/2)[b - a]^2 = (1/2)[ 1 - 2ab] = 1/2 - ab$. Let $p_k$ be the trace of the block of type $k$, which corresponds to the probability of each block of this type. We have that $ab = (c^m s^m)/p_k$. Using this simple observation, one obtains that the average probability of error over the blocks is \begin{eqnarray*} PE(m)& = & \sum_{k=0}^{(m-1)/2} {m \choose k} p_k [1/2 - (c^m s^m)/p_k]\\ & = & (1/2)[1 - (2cs)^m]. \end{eqnarray*} Therefore, we have obtaines $PE(m) = (1/2)[1 - \sin^m(\Omega)]$. Finally, we obtain $PC(m) \leq \sin^m(\Omega)$. Note that $PC(1) = 1 - \cos(\Omega)$, and we have $1 - \cos(\Omega) \leq \sin(\Omega)$ by the triangle inequality. }
\section{Introduction} \subsection{The principles involved. One of the outstanding problems of particle physics is whether the Higgs' field exists; these hypothetical fields are conjectured to make non-Abelian gauge fields $A^\alpha_i$, describe massive particles as needed by the standard particle physics models. There are \cite{bi:mdr89} alternatives to Higgs' scalar fields which use fluids instead. Apart from it being a good idea to have alternative models, the reason for producing these particular fluid alternatives can be addressed after outlining the principles below: \bd \item[The principle that: ''All fundamental fields are gauge fields''.] The success of non-Abelian gauge vector fields in describing fundamental interactions suggests the principle that ``{\sc All fundamental field are gauge fields.}'' Taken at face value this implies that the Higgs fields used in symmetry breaking should also be a gauge field: as it is a scalar field at first sight this seems impossible. A way around this is to have a collection of scalar fields with the properties that: {\it firstly} they encompass the standard Higgs mechanism, {\it secondly} they form a gauge system. This can be realized by using the scalar field description of a perfect fluid, and then ``charging'' it via $\partial_a\rightarrow D_a=\partial_a+ieA_a$, as described below. \item[The: "Extreme principle of equivalence."] The extreme principle of equivalence is that "{\sc There is only one concept of mass in physics.}" \cite{bi:mdr89}, and this extended principle requires the equivalence of these masses to all other masses, including those generated by symmetry breaking mechanisms. \item[The: "Redundancy of {\it ad hoc} procedures."] This principle is that "{\sc Any principled explanation (perhaps not only scientific) of anything is preferable to an {\it ad hoc} calculational procedure which produces an equivalent or inferior result.}" It leaves open what to choose if the {\it ad hoc} calculational procedure produces a superior result, or how to address the quality of the result. In the present case Higgs' scalar fields are {\it ad hoc} whereas fluids might arise from the statistical properties of the non-Abelian gauge fields. \end{description} \subsection{Application of these principle to symmetry breaking. \paragraph{Re-writing procedure. There are well known techniques by which stresses involving scalar fields can be rewritten as fluids, so that it is possible to re-write the standard model with fluids instead of Higgs scalars. My {\em first} attempt at using fluids for symmetry breaking \cite{bi:mdr89} was essentially to deploy these rewriting procedures to convert the scalar fields to fluids; the drawback of this approach is the fluids that result are somewhat unphysical, however an advantage is that symmetry breaking occurs with a change of state of the fluid. \paragraph{Velocity potential method. My {\em second} attempt \cite{bi:mdr97} used the decomposition of a perfect fluid vector into several scalar parts called vector potentials and identifying one of these with the radial Higgs scalar; thus the theory is a generalization of the standard theory, so that any experimental verification of the standard theory does not negate it. The gauge fields are then introduced by using the usual covariant substitutions for the partial derivatives of the scalars, for example $\phi_a=\partial_a\phi\rightarrow\nabla_a\phi=\partial_a\phi+ie\phi A_a$, and calling the resulting fluid the {\sc covariantly interacting fluid}. This results in an elegant extension of standard Higgs symmetry breaking, but with some additional parameters present. Previous work on the vector potentials shows that some of these have a thermodynamic interpretation, it is hoped that this is inherited in the fluid symmetry breaking models. The additional parameters can be partially studied with the help of an explicit Lagrangian, see below. Both of my approaches have been so far restricted to abelian gauge fields. \subsection{The lagrangian formulation of fluids. A perfect fluid has a Lagrangian formulation in which the Lagrangian is the pressure $p$. Variation is achieved by using the first law of thermodynamics \begin{equation} dp=n~dh-nT~ds, \label{eq:1.1} \end{equation} where $n$ is the particle number, $T$ is the temperature, $s$ is the entropy, and $h$ the enthalpy. In four dimensions a vector can be decomposed into four scalars, however the five scalar decomposition \begin{equation} hV_\mu=W_\mu=\sigma_\mu+\sum_{i}\theta^{(i)}s_{(i)\mu},~~~~~~~~V_\rho V^\rho=-1, \label{eq:1.2} \end{equation} $(i)=1,2$ is often used, because for $i=1$, $s$ and $\theta=\int T~d\tau$ have interpretation, Roberts (1997) \cite{bi:mdr97}, as the entropy and the thermasy respectively. From now on the index $(i)$ is suppressed as it is straightforward to reinstate. Replacing the first law with $dp=-V_\rho dW^\rho-nT~ds$, variation gives the familiar stress and scalar evolution equations. Previously, Roberts (1997) \cite{bi:mdr97}, all partial derivatives in the above have been replaced with vector covariant derivatives \begin{equation} D_\mu=\partial_\mu+ie~A_\mu, \label{eq:1.3} \end{equation} to obtain a generalization of scalar electrodynamics called fluid electrodynamics. For many interacting fluids the interaction terms can be disregarded, Anile (1990) \cite{bi:anile}, however here all interaction terms are kept. Quantization of these models is not looked at here although fluids can be quantized, Roberts (1999) \cite{bi:mdr99}. Here instead of using the first law for variation a specific explicit Lagrangian is assumed. This fixes the equation of state. Since previous work three things are approached differently. The {\it first} is the best conventions for the scalar decomposition are \ref{eq:1.2}. This is because the spacetime index is put on $s$ as this will allow easier generalization to second order non-equilibrium thermodynamics, Israel and Stewart (1979) \cite{bi:IS}. The $+$ convention is used for $\theta$ and $s$ in \ref{eq:1.2} (rather than $-\theta_\mu s$). The {\it second} is that the fluid remains isentropic after charging. The {\it third} is that two vector normalization conditions are required after charging. Equivalences between fluids and scalar fields were first studies by Tabensky and Taub (1973) \cite{bi:TT}. Thermodynamical quantities can be introduced into the standard Higg's model via the partition function Kapusta (1989) \cite{bi:kapusta}. Clearly the entropy, temperature and so on cannot occur from both the vector \ref{eq:1.2} and the partition function, perhaps a suitable partition function for the present model might allow both to be identified. An approach which dispenses with Higgs fields is that of Nicholson and Kennedy (2000) \cite{NK}, another approach which involves gravity is that of Kakushadze and Langfelder (2000) \cite{KL}. \section{Explicit Lagrangians} Consider the Lagrangian \begin{equation} L(q,q_\rho)=\beta(-W_\rho W^\rho)^{r}-Q(q), \label{eq:2.1} \end{equation} where $r$ is called the equation of state parameter (see equation \ref{eq:2.6}), $\beta$ is called the sign parameter and $Q$ is the potential. Varying with respect to the metric and then assuming the stress is of the form of a perfect fluid implies \begin{equation} -2(-1)^{r}\beta rW_\rho^{2r-2}W_\mu W_\nu=(p+\mu)V_\mu V_\nu. \label{eq:2.2} \end{equation} Normalization $V_\rho^{2}=-1$ and $hV_\mu=W_\mu$ implies \begin{equation} W_\rho^{2r}=(-1)^{r}h^{2r}. \label{eq:2.3} \end{equation} Recall that $L=p$, together with the above three equations this gives \begin{equation} p=\beta h-Q,~~~ \mu=(2r-1)\beta h^{2r}+Q,~~~ n=2\beta rh^{2r-1}. \label{eq:2.4} \end{equation} Varying $L$ with respect to $\sigma$, $\theta$ and $s$ respectively \begin{equation} (nV^\rho)_\rho-Q_{\sigma}= 0,~~~ -n\dot{s}-Q_{\theta}=0,~~~ (n\theta V^\rho)_\rho-Q_{\sigma}=0, \label{eq:2.5} \end{equation} This is a perfect fluid; when $Q=0$ it has the $\gamma$-equation of state \begin{equation} 1<\gamma=\frac{2r}{2r-1}<2,~~~ \frac{\gamma}{2\gamma-2}=r<\frac{1}{2}. \label{eq:2.6} \end{equation} Particular cases are $r=1$ which implies $\gamma=2$ which is the equation of state for coherent radiation; and $r=2$ which implies $\gamma=\frac{4}{3}$ which is the equation of state for incoherent radiation. The canonical momenta are \begin{equation} \Pi^{\sigma}=-n,~~~ \Pi^{\theta}=0,~~~ \Pi^{s}=n\theta. \label{eq:2.7} \end{equation} The constrained Hamiltonian is \begin{equation} H_{\lambda}=\Pi^{\sigma}(\dot{\sigma}+\theta\dot{s}) +\lambda^{1}(\Pi^{s}-\theta\Pi^{\sigma}) +\lambda^{2}\Pi^{\theta}-L, \label{eq:2.8} \end{equation} where $\lambda^{1}$ and $\lambda^{2}$ are the Lagrange multipliers. The momenta and Hamiltonian are the same as in the general case. In the general case the Euler equations and the canonical stress vanish identically; however for explicit Lagrangians the Euler equations are the same as \ref{eq:2.5} and the canonical stress is the same as the stress. \section{Charged Explicit Lagrangian.} All partial derivatives are replaced by vector covariant derivative \ref{eq:1.1}. Capital letters are used for the new quantities and small letters for the uncharged quantities. The Lagrangian becomes \begin{equation} L\rightarrow L=\beta(-W_\rho W^{*\rho})^{r}-Q(q)-\frac{1}{4}F^{2}, \label{eq:3.1} \end{equation} where "*" denotes complex conjugate. The vector field becomes \begin{equation} W_\mu=\sigma_\mu+\theta s_\mu+i(\sigma+\theta s)eA_\mu=w_\mu+i\bar{e} A_\mu, \label{eq:3.2} \end{equation} where $\bar{e}\equiv(\sigma+\theta s)e$. Under the global transformations \begin{equation} \sigma\rightarrow exp(-ie\lambda)\sigma,~~~ s\rightarrow exp(-ie\lambda)s,~~~ \theta\rightarrow\theta,~~~ A_\mu\rightarrow A_\mu+\lambda_\mu, \label{eq:3.3} \end{equation} the vector field changes to $W_\mu\rightarrow exp(-ie\lambda)W_\mu$ so that $W_\rho W^{*\rho}$ is invariant, implying that the Lagrangian is invariant. One can choose that $A_\mu=0$ and then these transformations are local. The N\"other current is \begin{equation} J^\mu=\beta r(-W_\rho W^{*\rho})^{r-1}\left\{ -i(\bar{e}^*w^\mu-\bar{e}w^{*\mu})+2\bar{e}\bar{e}^*A^\mu\right\}. \label{3.3b} \end{equation} One can introduce two normalization conditions \begin{equation} V_\mu V^{*\mu}=-1,~~~ v_\mu v^\mu=-1, \label{eq:3.5} \end{equation} which imply \begin{equation} H^{2}=h^{2}-e^{2}A_\rho^{2}, \label{eq:3.6} \end{equation} Proceeding as before \begin{equation} P=\beta H^{2r}-Q-\frac{1}{4}F^{2},~~~ \check{\mu}=\beta(2r-1)H^{2r}+Q+\frac{1}{4}F^{2},~~~ N=2\beta rH^{2r-1}, \label{eq:3.7} \end{equation} and the canonical momenta are \begin{equation} \Pi^{\sigma}=-N,~~~ \Pi^{\theta}=0,~~~ \Pi^{s}=-N\theta,~~~ \Pi^{A_\mu}=0, \label{eq:3.8} \end{equation} the last of which is surprising; because not all the components of $\Pi^{A_\mu}$ vanish in usual gauged scalar electrodynamics. The stress is \begin{eqnarray} T_{\mu\nu}&=&NH(V_\mu V^*_\nu+V^*_\mu V_\nu)+Pg_{\mu\nu},\nonumber\\ &=& \frac{N}{H}(w_\mu w_\nu+e^2A_\mu A_\nu)+Pg_{\mu\nu},\nonumber\\ T&=&3P-\check{\mu},\nonumber\\ v^\alpha v^\beta T_{\alpha\beta}&=&\mu+e\frac{N}{H}\left(A_\alpha^{2}+(v^\alpha A_\alpha)^2\right). \label{eq:3.9} \end{eqnarray} The conservation law is \begin{equation} T_{.\mu.;\beta}^\beta=w_\mu Q_{\sigma} +N\frac{h}{H}\dot{w}_\mu +e^{2}(\frac{N}{H}A_\mu A^\beta)_\beta +P_\mu. \label{eq:3.10} \end{equation} The constrained Hamiltonian is \begin{equation} H_{\lambda}=\Pi^{\sigma}(\dot{\sigma}+\theta\dot{s})+\lambda^{1}(\Pi^{s}-\theta\Pi^{\sigma}) +\lambda^{2}\Pi^{\theta} +\lambda^{3}\Pi^{A_{a}}-L, \label{eq:3.11} \end{equation} Variation with respect to the scalar and vector fields gives \begin{eqnarray} \delta\sigma&:&Q_{\sigma}=(N\frac{h}{H}v^\alpha)_\alpha,\nonumber\\ \delta\theta&:&-Q_{\theta}=N\frac{h}{H}\dot{s},\nonumber\\ \delta s&:&Q_{s}=(N\frac{h}{H}v^\alpha)_\alpha,\nonumber\\ \delta A_\mu&:&F_{\mu;\beta}^\beta=e^2\frac{N}{H}A_\mu. \label{eq:3.12} \end{eqnarray} Note that if one starts with the electrodynamical Lagrangian ${\cal{L}}=-F^2/4$ and tries to implement a substitution scheme from this via $A_\mu\rightarrow A_\mu+kW_\mu$ one cannot recover the above type of charged fluids because $F^2\rightarrow F^2+$ second derivatives in the scalar fields, which do not occur in the above. Thus there is no mirror mechanism. \section{Comparison with the Higgs' Model} To recover scalar electrodynamics first note that \begin{equation} W_\alpha W^{*\alpha}=w_\alpha^2+\bar{e}^2A_\alpha^2. \label{eq:4.1} \end{equation} Now set the state parameter $r=1$, the sign parameter $\beta=-1$, and the potential $Q(q)=V(\rho^{2})$ so that \begin{equation} L=w_\alpha^2+\bar{e}^2A_\alpha^2-V(\rho^2)-\frac{1}{4}F^{2}, \label{eq:4.2} \end{equation} which has no cross term $w_\alpha A^\alpha$. Now set $\sigma=\rho$, $\theta=s= 0$. Changing the gauge $A'_\mu=\nu_\mu+A_\mu$ and dropping the prime gives {\it scalar electrodynamics} in radial form. This has lagrangian, stress and equations of motion \begin{eqnarray} L&=&\rho_\alpha^2+(\hat{\nabla}_\alpha\nu)^{2}-V(\rho^{2})-\frac{1}{4}F^{2},\nonumber\\ \hat{\nabla}_\mu\nu&=&\rho(\nu_\mu+eA_\mu),\nonumber\\ T_{\mu\nu}&=&2\phi_\mu\phi_\nu+2\hat{\nabla}_\mu\nu\hat{\nabla}_\nu\nu +F_{\mu\gamma}F_{\nu.}^{.\gamma}+g_{\mu\nu}L,\nonumber\\ F^{\mu\beta}_{..\beta}&+&2e\rho\hat{\nabla}^\mu\nu=0,\nonumber\\ 2[\stackrel{2}{\nabla}&+&(\nu_\beta+eA_\beta)^{2}+V']\rho=0,\nonumber\\ 2[\stackrel{2}{\nabla}\nu&+&eA^\beta_\beta]=0, \label{eq:4.3} \end{eqnarray} respectively. Giving the Higgs field a non-zero expectation value $<0|\rho|0>= a$. $\rho$ transfers to \begin{equation} \rho \rightarrow \rho ' = \rho +a, \label{eq:4.4} \end{equation} and the lagrangian becomes \begin{equation} L=\rho_\beta^{2}+(\rho+a)^2e^2A_\beta^2-V\left((\rho+a)^{2}\right) -\frac{1}{4}F^{2}, \label{eq:4.5} \end{equation} the term $a^2e^2A_\beta^2$ is a constant mass term. For a fluid one can perform a 'scalar fluid rotation' \begin{equation} \sigma \rightarrow \sigma '=\sigma+\theta s. \label{eq:4.6} \end{equation} Starting with scalar electrodynamics and identifying $\rho$ with $\frac{\sigma}{e}$ does not give one of the above fluids as gradients in $\theta$ appear. Alternatively there is the 'vector fluid rotation' \begin{equation} \sigma_\mu=\rho_\mu\rightarrow w_\mu. \label{eq:4.7} \end{equation} Starting with scalar electrodynamics and performing this rotation and replacing $\rho$ with $\frac{\sigma}{e}$ gives a fluid of the above type. One can then assume a VEV $<0|\frac{\sigma}{e}|0>$ to generate a constant mass term as above, the difference being that some account of thermodynamics has been achieved. \section{Acknowledgements} I would like to thank Prof.T.W.B.Kibble for his interest in this work.
\section{Introduction} Progress in understanding nonleptonic weak decays of bottom baryons has been very slow both theoretically and experimentally. While some new data of charmed baryon nonleptonic weak decays have become available, the experimental situation for hadronic weak decays of bottom baryons is meagre. The decay mode $\Lambda_b \to \Lambda J/\psi $ is the first successful measurement \cite{ref00} of the exclusive hadronic decay rate of bottom baryons. In the near future, one can expect new data on exclusive bottom baryon decays calling for a comprehensive theoretical analysis of these decay modes. These decay processes can provide useful information for QCD effects in weak decays and indirect CP asymmetry which involve the CKM-Wolfenstein parameters $\rho $ and $\eta $. However, a rigorous and reliable approach suitable for analyzing these decays does not exist at this time, and one must rely on some approximation to deal with them. The well-known factorization approach,\cite{ref1} which has been applied successfully to nonleptonic $B$ meson decays \cite{ref2} can also be applicable to bottom baryon decays. In this approximation the effects of final state interactions (FSI) are neglected. These interactions are thought to be less important in bottom baryon decay, since the decay particles in the two-body final state are energetic and moving fast, allowing little time for significant final state interactions. In previous papers, we made analyses of exclusive semi-leptonic\cite{ref6p} and non-leptonic\cite{Bmeson} decays of $B$ mesons by using the covariant oscillator quark model(COQM),\cite{ref3} leading to satisfactory results. In this paper we discuss the nonleptonic decays $\Lambda_b \to \Lambda_c P(V) $ and $\Lambda_b \to \Lambda J/\psi $ using this model. One of the most important motives for COQM is to covariantly describe the centre-of-mass motion of hadrons, retaining the considerable successes of the non-relativistic quark model on the static properties of hadrons. As a result, in COQM we can treat both static and non-static problems simultaneously. A common systematic treatment of the systems with any quark flavor configuration is another feature of COQM. A key point in COQM for doing this is to treat directly the squared masses of hadrons in contrast to the mass itself, as is done in conventional approaches. This makes the covariant treatment simple. The COQM has been applied to various problems \cite{ref4} with satisfactory results. Here we would like to note that, when applied to heavy-to-heavy baryon transitions, our model predictions satisfy the constraints imposed by the heavy quark symmetry,\cite{ref6} suggesting that COQM is reliable. Moreover, our model is also applicable for heavy-to-light transitions, which are beyond the scope of heavy quark effective theory (HQET). We organise the paper as follows. Assuming factorization, in \S 2 we present the expressions for the decay rates and the up-down asymmetry parameter $\alpha $. We use the covariant oscillator quark model to evaluate the baryonic form factors. Results and discussion are presented in \S 3. \section{Methodology} \subsection{Effective weak Hamiltonian and factorized amplitude} Neglecting the penguin contribution, the effective Hamiltonian describing the decays under consideration is given by \begin{eqnarray} &&{\cal H}_{\rm eff} = \frac{G_F}{\sqrt{2}}\; V_{cb}\;V_{q_i q_j}\; [a_1~ O_1 + a_2~ O_2 ] \label{eq:1} \end{eqnarray} with \begin{eqnarray} && O_1=(\bar q_i q_j)^{\mu}\; (\bar c b)_{\mu}~~~~~~~~~~ \mbox{and}~~~~~~~~~~~ O_2= (\bar q_i b)^{\mu} (\bar c q_j)_{\mu}\;, \end{eqnarray} where $a_1$ and $a_2$ correspond to the external and internal $W$ emission amplitudes. The quark current $(\bar q_i q_j )_{\mu}=\bar q_i \gamma_\mu(1+\gamma_5) q_j$ denotes the usual $(V-A)$ current. $q_i$ and $q_j $ are two types of quark flavors that are hadronized to $P$ or $V$ mesons. Here Eq.~(\ref{eq:1}) is to be understood as an effective Hamiltonian after considering the Fierz rearrangement. In the factorization approximation, the baryon decay amplitude is factorized into a product of two matrix elements of the form $\langle B^\prime|J_\mu | B \rangle $, which involves the form factors, and $\langle P(V) |J^\mu |0 \rangle $, which is expressed by the meson decay constant. The factorized amplitude for the decay processes $\Lambda_b \to \Lambda_c P(V) $, which proceed via the external $W$ emission is given as \begin{equation} {\cal M}(\Lambda_b \to \Lambda_c P(V) )= \frac{G_F}{\sqrt 2} V_{cb} V_{q_i q_j}~a_1~\langle P(V)|(\bar q_i q_j)^{\mu}| 0 \rangle \langle \Lambda_c | (\bar c b)_\mu |\Lambda_b \rangle \;.\label{eq:eqn8} \end{equation} Similarly, the amplitude for the decay mode $\Lambda_b \to \Lambda J/\psi $, which is described by the internal $W$ emission, is given by Eq. (\ref{eq:eqn8}) in which $a_1$ is replaced by $a_2 $. The current matrix elements between a pseudoscalar/vector/ axialvector meson\footnote{ We also treat the decay $\Lambda_b\to\Lambda_c+a_1$, $a_1$ being the axial-vector meson. } ($P/V/a_1 $) and the vacuum are related to the corresponding decay constants as \begin{eqnarray} &&\langle P(p) |(\bar q_i q_j)^{\mu} |0 \rangle = -if_P~ p^\mu ,\nonumber\\ &&\langle V(p,\epsilon) |(\bar q_i q_j)^{\mu} |0 \rangle = M_V~f_V~ \epsilon^\mu ,\nonumber\\ &&\langle a_1(p,\epsilon) |(\bar q_i q_j)^{\mu} |0 \rangle = M_{a_1}~f_{a_1}~ \epsilon^\mu\;, \label{eq4} \end{eqnarray} where $f_P$, $f_V$ and $f_{a_1}$ are the respective decay constants. To evaluate the baryon form factors, we use the covariant oscillator quark model, which is explicitly presented in the next section. \subsection{ Model framework and the Hadronic form factors} The general treatment of COQM may be called the ``boosted $LS$-coupling scheme,'' and the wavefunctions, being tensors in $\tilde U(4) \times O(3,1) $-space, reduce to those in the $SU(2)_{\rm spin} \times O(3)_{\rm orbit} $-space in the nonrelativistic quark model in the hadron rest frame. The spinor and space-time portion of the wave functions separately satisfy the respective covariant equations, the Bargmann-Wigner (BW) equation for the former and the covariant oscillator equation for the latter. The model parameters and form of the wave function are determined completely through the analysis of mass spectra.\cite{ref3,ref4} In COQM all the non-exotic $ qqq $ baryons are described\cite{ref6pp} by tri-local fields $ \Phi_{A_1 A_2 A_3} (x_1,x_2,x_3) $, where the $x_{i}$ are Lorentz four vectors representing the space time coordinates of constituent quarks, $A_1=(a,\alpha)\;A_2=(b,\beta)~A_3=(c,\gamma)$, describing the flavor and covariant spinor of the quarks. The baryon fields are assumed to satisfy wave equations of the Klein-Gordon type and expanded in terms of Fierz components (that is, the eigenfunctions of ${\cal M}^2$), and are written as \begin{equation} \Phi(X,r \cdots ) =\sum_{P,n}\left (e^{i P_n X}\Psi_n^{(+)} (r \cdots,P_n)+e^{-iP_n X}\Psi_n^{(-)}(r \cdots ,P_n)\right )\;, \label{eq:eqn1} \end{equation} \begin{equation} {\cal M}^2(r_{\mu} \cdots , \partial/\partial r_{\mu} \cdots ) \Psi_n^{(\pm)}(r \cdots , P_n)=M_n^2 \Psi_n^{(\pm)} (r \cdots ,P_n )\;, \end{equation} where $X_{\mu}(r_\mu \cdots )$ is the centre-of-mass (relative) coordinate, and ${\cal M}^2$ is the squared mass operator depending on relative coordinate variables. The first (second) term of Eq. (\ref{eq:eqn1}) corresponds to the positive (negative) frequency part of the centre-of-mass plane-wave motion with definite total four momentum $P_n$ and mass $M_n=\sqrt{-P_n^2} $. The spin portions $U_n$ $(\bar U_n)$ of positive (negative) frequency internal wave functions $\Psi_n^{(+)}(r \cdots ,P_n) \equiv U_n(P) f_n(r \cdots, P_n ) $ $(\Psi_n^{(-)} \equiv \bar U_n f_n )$, satisfy the respective Bargmann-Wigner (BW) equations. The BW spinor functions are reducible and decompose into the irreducible components as \begin{eqnarray} U_{n,ABC}&=&\frac{1}{2 \sqrt{2}}[-\gamma_5(1+i v \cdot \gamma)C^{-1}]_{\alpha \beta}~u_{n,~abc;~\gamma}^{(A)} +\frac{1}{2 \sqrt{2}}[i\gamma_\mu (1+i v \cdot \gamma)C^{-1}]_{\alpha \beta} \nonumber\\ & & \times\left [u_{n,~abc;~\gamma,~\mu} ^{(S^*)}+\frac{-1}{\sqrt 3}\left \{(i \gamma_{\mu} +v_{\mu}) \gamma_5~u_{n,~abc}^{(S)}\right \}_\gamma \right ]\;,\label{eq:eqn2} \end{eqnarray} where $v_{\mu}$ is the four velocity of the hadron ($v_{\mu} =P_{\mu}/M $) and $C$ is the antiparticle conjugate matrix. The quantities $u_n^{(A)} $, $u_{n, \mu}^{(S^*)} $ and $u_n^{(S)}$ denote the spin-1/2 `Dirac spinor', spin-3/2 `Rarita-Schwinger vector-spinor', and the spin-1/2 `Dirac spinor' Fierz components, respectively. It may be noted that in the $\Lambda_Q$-type baryons, the two light quarks are in the flavor antisymmetric and spin-0 state. Thus the BW spinor function for $\Lambda_Q$-type baryons is given by the first term of Eq. (\ref{eq:eqn2}). The oscillator space-time wave function for the ground state baryons is given\footnote{ In this paper we apply the pure-confining approximation, neglecting the effect of the one-gluon-exchange potential $U_{\rm OGE}$. This approximation is expected, aside from the spin-dependent structure, to be good for the light-light $q\bar q$ and heavy-light $Q\bar q$ meson systems, where the reduced mass of the system is small comparatively and the effect of the central potential out of $U_{\rm OGE}$ is expected to be not so large. A similar situation is also expected to be valid for the relevant $Qqq$ (or $qqq$) baryon systems. The well-known phenomenological fact of a linearly-rising Regge trajectory for the $q\bar q$ and $qqq$ systems seems to support this conjecture. (See the papers referred to in Ref. \citen{ref4}, published in 1993 and 1994.) } by \begin{equation} f(P,\rho, \lambda)=f_{\rho}(P;\rho)~f_{\lambda}(P;\lambda), \end{equation} where \begin{equation} f_{\rho}(P;\rho)=\frac{\beta_{\rho}}{\pi} \exp\left ( -\frac{\beta_\rho }{2} \left (\rho^2+2\frac{(P\cdot \rho)^2}{M^2}\right ) \right )\;, \end{equation} and \begin{equation} f_{\lambda}(P;\lambda)=\frac{\beta_{\lambda}}{\pi} \exp\left ( -\frac{\beta_\lambda }{2} \left (\lambda^2+2\frac{(P\cdot \lambda)^2} {M^2}\right ) \right )\;, \end{equation} with \begin{equation} \beta_\rho = \frac{\sqrt{3 m K}}{4}~~~~~~~~~\mbox{and} ~~~~~~~~~~\beta_{\lambda} =\sqrt{\frac{mM K}{2m+M}}\;.\label{eq:eqn3} \end{equation} In the above equation, $m$ and $M$ denote the mass of the light (heavy) quark, and $K$ is the universal spring constant for all hadronic systems with the value\footnote{ In this paper we apply the values of $\beta$ corresponding to the case (A) in Ref. \citen{ref6p}. The values of the branching ratios with $\beta$ corresponding to the case (B) generally become slightly smaller (by $\stackrel{<}{\sim}20\%$). } $K$ =0.106 $\mbox{GeV}^3 $.\cite{ref8} The internal relative coordinates are defined as \begin{equation} \rho_{\mu}=x_{2 \mu}-x_{1 \mu}~~~~~~~\mbox{and}\;\;\;\;\;\; \lambda_{\mu}=x_{3 \mu} - \frac{x_{1\mu}+x_{2 \mu}}{2}\;. \end{equation} The effective action for weak interactions of baryons with $W$-bosons is given by \begin{eqnarray} S_W &=& \int d^4 x_1 d^4 x_2 d^4 x_3~ \langle \bar \Phi_{F, P^\prime} (x_1,x_2,x_3)^{C^\prime B A}~ i\gamma_{\mu} (1+\gamma_5{)_{\gamma^\prime}}^{\gamma}\nonumber\\ & & \times \Phi_{I, P}(x_1,x_2,x_3)_{ABC} \rangle W_{\mu,q}(x_3)\;, \end{eqnarray} where we have omitted the CKM matrix elements and the coupling constants. This is obtained from the consideration of covariance, assuming a quark current with the standard $V-A $ form. Here $\Phi_{I,P}~(\bar \Phi_{F,P^\prime}) $ denotes the initial (final) baryons with definite four momentum $P_\mu~(P_{\mu}^{\prime}) $, and $q_\mu $ is the momentum of the $W$ boson. The notation $\langle~~\rangle $ represents the trace of Dirac spinor indices. Concentrating only on $\Lambda_Q (Qud) \to \Lambda_{Q^\prime} (Q^\prime ud) $, the relevant effective currents $J_{\mu}(X)_ {P^\prime,P} $ are obtained by identifying the above action with \begin{equation} S_W =\int d^4 X J_{\mu}(X)_{P^\prime,P}\;W_{\mu}(X)_q\;. \end{equation} Then $J_{\mu}(X=0)_{P^\prime,P} \equiv J_\mu $ is explicitly given as,\cite{ref6pp} \begin{equation} J_{\mu}^{Q Q^\prime}=I_{ud}^{Q^\prime Q}(w)~ \bar u_A^{Q^\prime q_2 q_1}(v^\prime)~\frac{w+1}{2}~ i \gamma_{\mu}(1+\gamma_5)~u_{A,Qq_1 q_2 }(v)\;. \label{eq15} \end{equation} The quantity $I_{ud}^{Q^\prime Q}(w) (w\equiv -v\cdot v')$, the overlapping of the initial and final wave functions, represents the universal form factor. It describes the confined effects of quarks and is given by \begin{equation} I_{ud}^{Q^\prime Q}(w)=\frac{1}{w}~\frac{4 \beta_{\lambda} \beta_{\lambda}^\prime} {\beta_{\lambda}+\beta_{\lambda}^\prime } \frac{1}{\sqrt{C(w)}} \exp(-G(w))\;, \label{eqnI} \end{equation} where \begin{equation} C(w)=(\beta_\lambda-\beta_\lambda^{\prime})^2+ 4\beta_\lambda \beta_\lambda^{\prime}~w^2\;, \end{equation} and \begin{equation} G(w)= \frac{4 m_q^2 (\beta_\lambda+\beta_\lambda^\prime)~w (w-1)} {(\beta_\lambda - \beta_\lambda^\prime )^2 + 4 \beta_\lambda \beta_\lambda^\prime~w^2}\;. \end{equation} The expression for $\beta_\lambda $ is given by Eq. (\ref{eq:eqn3}) The form factor function $I_{ud}^{cb}(w)$ for $ \Lambda_b \to \Lambda_c $ decays corresponds to the baryonic Isgur-Wise function $\eta(w) $ in HQET.\cite{ref6} At the zero recoil point $w=1$, the value of $I_{ud}^{cb}(w)$ is given by \begin{equation} I_{ud}^{cb}(w=1)=\frac{1}{w}\frac{4 \beta_{\lambda} \beta_{\lambda}^\prime} {(\beta_{\lambda}+\beta_{\lambda}^\prime)^2}\;.\label{eq:eqn7} \end{equation} In the heavy quark symmetry limit we have $\beta_\lambda= \beta_{\lambda}^\prime $, so that Eq. (\ref{eq:eqn7}) correctly reproduces\cite{ref6p} the normalization condition of HQET, i.e., $\eta(w=1)=1 $. However, HQET as it is predicts nothing about the Isgur-Wise function except for the zero recoil point, while in COQM the form factor functions can be derived at any kinematical point of interest. In addition, the functional form Eq.~(\ref{eqnI}) of the COQM form factor $I_{ud}^{Q^\prime Q}(w)$ also applies to the heavy-to-light transition processes, whereas HQET does not provide anything for this sector. \subsection{Decay rates and aymmetry parameters} After obtaining the effective current in the COQM with Eqs.~(\ref{eq:eqn8}) and (\ref{eq15}), one can write the transition amplitude for the decay mode $\Lambda_b \to \Lambda_c P $ as \begin{eqnarray} {\cal M}(\Lambda_b(v) \to \Lambda_c(v^\prime) P(p))&=& \frac{G_F}{2\sqrt 2} V_{cb} V_{q_i q_j}~ a_1~f_P~p^{\mu}~I_{ud}^{cb} (w)~(w+1)\nonumber\\ & & \times \bar u_A^{cud}(v^\prime)~ \gamma_{\mu}(1+\gamma_5)~u_{A,bud}(v)\;,\label{eq:eqn4} \end{eqnarray} and the corresponding decay width as \begin{eqnarray} \Gamma (\Lambda_b \to \Lambda_c P )&=& \frac{G_F^2}{32 \pi M_{\Lambda_b}^2} |V_{cb} V_{q_i q_j}|^2~a_1^2~ f_P^2~ (I_{ud}^{cb}(w))^2 (w+1)^2~|{\bf p}|\nonumber\\ & & \times \left [(M_{\Lambda_b}^2 -M_{\Lambda_c}^2)^2 -M_P^2(M_{\Lambda_b}^2+M_{\Lambda_c}^2)\right ]\;, \end{eqnarray} where ${\bf p}$ is the c.m.~momentum of the emitted particles in the rest frame of the parent $\Lambda_b $ baryon. To obtain the asymmetry parameter $\alpha $, we write Eq. (\ref{eq:eqn4}), substituting $p^\mu=M_{\Lambda_b} v^{\mu} - M_{\Lambda_c} v^{\prime \mu} $ as \begin{equation} {\cal M}(\Lambda_b \to \Lambda_c P)=f_P~ \bar u_{A}^{cud} ~(G_1+\gamma_5G_2)~u_{A,bud}(v)\;, \end{equation} where \begin{equation} G_1=\lambda(M_{\Lambda_b}-M_{\Lambda_c})\;,~~~~~~~~~~~ G_2=-\lambda(M_{\Lambda_b}+M_{\Lambda_c})\;, \end{equation} with \begin{equation} \lambda=\frac{G_F}{2 \sqrt 2} V_{cb} V_{q_i q_j} ~a_1 ~I_{ud}^{cb}(w)~ (1+w)\;. \end{equation} In terms of the new form factors $G_1$ and $G_2$, the asymmetry parameter $\alpha$ is given by \cite{ref91} \begin{equation} \alpha=\frac{2 G_1 G_2 |{\bf p}|}{(E_{\Lambda_c}+M_{\Lambda_c}) G_1^2+(E_{\Lambda_c}-M_{\Lambda_c})G_2^2}\;. \end{equation} To obtain the decay rate for the $\Lambda_b \to \Lambda_c V $ transition, we write the transition amplitude for the process as \begin{eqnarray} {\cal M}(\Lambda_b(v) \to \Lambda_c(v^\prime) V(p,\epsilon))&=& i\frac{G_F}{2\sqrt 2} V_{cb} V_{q_i q_j}~ a_1~f_V~M_V~ \epsilon^{\mu}~I_{ud}^{cb} (w)~(w+1)\nonumber\\ & & \times \bar u_A^{cud}(v^\prime)~ \gamma_{\mu}(1+\gamma_5)~u_{A,bud}(v)\;.\label{eq:eqn5} \end{eqnarray} The corresponding decay width is given by \begin{eqnarray} \Gamma (\Lambda_b \to \Lambda_c V )&=& \frac{G_F^2}{32 \pi M_{\Lambda_b}^2} |V_{cb} V_{q_i q_j}|^2~a_1^2~ f_V^2~ (I_{ud}^{cb}(w))^2 (w+1)^2~|{\bf p}|\nonumber\\ & & \times \left [(M_{\Lambda_b}^2 -M_{\Lambda_c}^2)^2 +M_V^2(M_{\Lambda_b}^2+M_{\Lambda_c}^2-2M_V^2)\right ]\;. \label{eqn27} \end{eqnarray} To obtain the asymmetry parameter, we compare Eq. (\ref{eq:eqn5}) to the general form \cite{ref09} in the rest frame of $\Lambda_b $, \begin{equation} \chi_{\Lambda_c}^{\dagger}[ S {\bf \sigma} +P_1 {\hat{\bf p}} +i P_2 ({\hat{\bf p}} \times {\bf \sigma})+D({\bf \sigma}\cdot {\hat{\bf p}}) {\hat{\bf p}}] \cdot {\bf \epsilon}~ \chi_{\Lambda_b}\;, \end{equation} with ${\hat{\bf p}}$ now a unit vector in the direction of the $\Lambda_c $ baryon. The values for the four amplitudes $ S,~ P_1,~P_2$ and $D$ are given as \begin{equation} S=i \frac{G_F}{2 \sqrt 2}~ V_{cb} V_{q_i q_j}~ a_1~ f_V~M_V~(I_{ud}^{cb}(w))(1+w)\;, \end{equation} \begin{equation} P_1/S= - \left ( \frac{M_{\Lambda_b}+M_{\Lambda_c}} {E_{\Lambda_c}+M_{\Lambda_c}} \right ) \left ( \frac{|{\bf p}|}{E_V} \right ) , \end{equation} \begin{equation} P_2/S= \left ( \frac{|{\bf p}| } {E_{\Lambda_c}+M_{\Lambda_c}} \right ) , \end{equation} \begin{equation} D/S= \left ( \frac{|{\bf p}|^2 } {E_V(E_{\Lambda_c}+M_{\Lambda_c})} \right ) . \end{equation} From these expressions, the up-down asymmetry $\alpha $ of the final $\Lambda_c $ with respect to $\Lambda_b $ polarization is given by \begin{equation} \alpha=2\mbox{Re}~\frac{[(1+D/S)^*P_1/S+2(P_2^*/S) M_V^2/E_V^2]} {K}\;, \end{equation} where \begin{equation} K=[|1+D/S|^2+|P_1/S|^2+2(1+|P_2/S|^2)M_V^2/E_V^2]\;. \end{equation} The decay rate for $\Lambda_b \to \Lambda J/\psi $ is given in analogy to Eq. (\ref{eqn27}) with $a_1$ replaced by $a_2$ and $f_V $ replaced by $f_{J/\psi}$, with the relevant CKM matrix elements as $|V_{cb} V_{cs}|^2$. The COQM is applicable not only for heavy-to-heavy transitions but also for heavy-to-light transitions, where the final baryon moves extremely relativistically, as was mentioned in \S 1. Here we consider the case of the decay mode $\Lambda_b \to \Lambda \bar D^0 $ to obtain the CKM-Wolfenstein parameter $(\rho^2+\eta^2)$ from the ratio of the the decay widths $\Gamma (\Lambda_b \to \Lambda \bar D^0) / \Gamma (\Lambda_b \to \Lambda J/\psi)$. The decay width for the process $\Lambda_b \to \Lambda \bar D^0 $ is given as \begin{eqnarray} \Gamma (\Lambda_b \to \Lambda \bar D^0 )&=& \frac{G_F^2}{32 \pi M_{\Lambda_b}^2} |V_{ub} V_{cs}|^2~a_2^2~ f_D^2~ (I_{ud}^{sb}(w))^2 (w+1)^2~|{\bf p}|\nonumber\\ & & \times \left [(M_{\Lambda_b}^2 -M_{\Lambda}^2)^2 -M_D^2(M_{\Lambda_b}^2+M_{\Lambda}^2)\right ]\;. \end{eqnarray} \section{Results and conclusion} \begin{table} \caption{Branching ratios (in percent) for different nonleptonic $\Lambda_b$ decay processes and their asymmetry parameters $\alpha$.} \vspace {0.2 true in} \begin{center} \begin{tabular}{lllll} \hline \multicolumn{1}{c}{Decay processes}& \multicolumn{1}{c}{$\alpha $} & \multicolumn{1}{c}{Br. ratio in $\%$}\\ \hline $\Lambda_b \rightarrow \Lambda_c^+ \pi^-$ & $-0.999$ & 0.175\\ $\Lambda_b \rightarrow \Lambda_c^+ K^-$ & $-1.000$ & 0.013\\ $\Lambda_b \rightarrow \Lambda_c^+ D^-$ & $-0.987$ & 0.030 \\ $\Lambda_b \rightarrow \Lambda_c^+ D_s^-$ & $-0.984$& 0.77 \\ $\Lambda_b \rightarrow \Lambda_c^+ \rho^-$ & $-0.898 $ & 0.491 \\ $\Lambda_b \rightarrow \Lambda_c^+ K^{*-}$ & $-0.865$ & 0.027 \\ $\Lambda_b \rightarrow \Lambda_c^+ D^{*-}$ & $-0.459$ & 0.049\\ $\Lambda_b \rightarrow \Lambda_c^+ D_s^{*-} $ & $-0.419$ & 1.414 \\ $\Lambda_b \to \Lambda_c^+ a_1 $& $-0.758 $& 0.532 \\ $\Lambda_b \rightarrow \Lambda J/\psi$& $-0.208$ & $2.55 \times 10^{-2}$\\ \hline \end{tabular} \end{center} \end{table} \begin{table} \caption{Branching ratios (in percent) for different $\Lambda_b$ processes and comparison with other calculations.} \vspace {0.2 true in} \begin{center} \begin{tabular}{llllll} \hline \multicolumn{1}{c}{Decay processes}& \multicolumn{1}{c}{This work}& \multicolumn{1}{c}{Ref. \citen{ref05} }& \multicolumn{1}{c}{Ref. \citen{ref06}} & \multicolumn{1}{c}{Ref. \citen{ref07} } & \multicolumn{1}{c}{Ref. \citen{ref08} }\\ \multicolumn{1}{c}{}& \multicolumn{1}{c}{}& \multicolumn{1}{c}{}& \multicolumn{1}{c}{Large $N_c$} & \multicolumn{1}{c}{Large $N_c$} & \multicolumn{1}{c}{Large $N_c$}\\ \hline $\Lambda_b \rightarrow \Lambda_c^+ \pi^-$ & 0.175 &0.38 & - &0.391 & 0.503\\ $\Lambda_b \rightarrow \Lambda_c^+ K^-$ & 0.013 & - &- &- & 0.037 \\ $\Lambda_b \rightarrow \Lambda_c^+ D^-$ & 0.030 &-&- &- &- \\ $\Lambda_b \rightarrow \Lambda_c^+ D_s^-$ & 0.77&1.1 &2.23& 1.291 &- \\ $\Lambda_b \rightarrow \Lambda_c^+ \rho^-$ & 0.491&0.54 & - &1.082 &0.723 \\ $\Lambda_b \rightarrow \Lambda_c^+ K^{*-}$ & 0.027 & - &- &- & 0.037 \\ $\Lambda_b \rightarrow \Lambda_c^+ D^{*-}$ & 0.049&- &-&- &- \\ $\Lambda_b \rightarrow \Lambda_c^+ D_s^{*-} $ & 1.414&0.91&3.26 & 1.983 & - \\ $\Lambda_b \to \Lambda_c a_1 $ & 0.532 &-&-&-&-\\ $\Lambda_b \rightarrow \Lambda J/\psi$& $2.49 \times 10^{-2}$ &$1.6 \times 10^{-2} $ & $6.037 \times 10^{-2}$ &-&-\\ \hline \end{tabular} \end{center} \end{table} \begin{table} \caption{Asymmetry parameter $\alpha$ for different $\Lambda_b$ processes and comparison with other calculations.} \vspace {0.2 true in} \begin{center} \begin{tabular}{lllll} \hline \multicolumn{1}{c}{Decay processes}& \multicolumn{1}{c}{This work}& \multicolumn{1}{c}{Ref. \citen{ref05} }& \multicolumn{1}{c}{Ref. \citen{ref06}} & \multicolumn{1}{c}{Ref. \citen{ref08}}\\ \multicolumn{1}{c}{}& \multicolumn{1}{c}{}& \multicolumn{1}{c}{}& \multicolumn{1}{c}{Large $N_c$}& \multicolumn{1}{c}{Large $N_c$} \\ \hline $\Lambda_b \rightarrow \Lambda_c^+ \pi^-$ & $-0.999$ &$-0.99$ & - &$-1.00$\\ $\Lambda_b \rightarrow \Lambda_c^+ K^-$ & $-1.000$ & - & -&$-1.00$ \\ $\Lambda_b \rightarrow \Lambda_c^+ D^-$ & $-0.987$&-&- &- \\ $\Lambda_b \rightarrow \Lambda_c^+ D_s^-$ & $-0.984$& $-0.99 $&$-0.98$&- \\ $\Lambda_b \rightarrow \Lambda_c^+ \rho^-$ & $-0.898$ &$-0.88$ & - &$- 0.885$ \\ $\Lambda_b \rightarrow \Lambda_c^+ K^{*-}$ & $-0.865$ & - & - &$-0.885$ \\ $\Lambda_b \rightarrow \Lambda_c^+ D^{*-}$ & $-0.459$&- &- &- \\ $\Lambda_b \rightarrow \Lambda_c^+ D_s^{*-} $ & $-0.419$ &$-0.36 $&$-0.40$ & - \\ $\Lambda_b \to \Lambda_c a_1 $ & $-0.758 $& -& -& -\\ $\Lambda_b \rightarrow \Lambda J/\psi$& $-0.208$ &$-0.1$& $-0.18$ &-\\ \hline \end{tabular} \end{center} \end{table} In order to make a numerical estimate, we use the following values of various quantities. The quark masses (in GeV) are taken as $m_u=m_d=0.4$, $m_s=0.51$, $m_c=1.7$ and $m_b=5$, which are determined from the analysis\cite{refmR} of meson mass spectra. The particle masses and lifetimes are taken from Ref. \citen{ref10}. The relevant CKM parameters\cite{ref10} used are $V_{cb}=0.0395$, $V_{cs}=1.04 $, $V_{cd}=0.224$, $V_{ud}=0.974$ and $V_{us} =0.2196$. The decay constants are taken as $f_{\pi}=130.7$, $f_K=159.8$; $f_{K^*}=214$;\cite{refa} $f_{\rho}=210$;\cite{refb} $f_D=220$, $f_{D^*}=230$, $f_{D_s}=240$, $f_{D_s^*}=260$\cite{refc} and $f_{a_1}=205$\cite{refd} (in MeV). The decay constant $f_{J/\psi}$ is determined from the value of $\Gamma(J/\psi \to e^+ e^- )$:\cite{ref10} \begin{equation} f_{J/\psi}=\sqrt{\frac{9}{4} \left (\frac{3}{4\pi \alpha^2} \right ) \Gamma(J/\psi \to e^+ e^-)M_{J/\psi}}=404.5 ~\mbox{MeV}\;. \end{equation} The parameters $a_1$ and $a_2$ appearing in these decays have recently been extracted from the CLEO data and turn out to be $a_1=1.05$ and $a_2=0.25$.\cite{ref02} Using these values we have obtained the values of the branching ratios and the asymmetry parameter $\alpha$ for several nonleptonic $\Lambda_b $ decays. These values are tabulated in Table I. The asymmetry parameter $\alpha $ in all these decay modes is found to be negative, which indicates the $V-A$ nature of the weak current. Our predicted branching ratio for the decay process $\Lambda_b \to \Lambda J/\psi $ ($2.55 \times 10^{-4})$ is consistent with the recent experimental data \cite{ref10} ($4.7 \pm 2.7 \times 10^{-4}$): Here it may be worthwhile to note that the final $\Lambda$ in this process experiences relativistic motion with velocity $v_\Lambda =0.84c$, and the form factor function plays a significant role taking a value $I(w)=0.11$. From the ratio of the decay widths $\Lambda_b \to \Lambda \bar D^0 $ and $\Lambda_b \to \Lambda J/\psi $, we obtain \begin{equation} \frac{\Gamma(\Lambda_b \to \Lambda \bar D^0)}{\Gamma (\Lambda_b \to \Lambda J/\psi)}=10.235 \times 10^{-2}~ |V_{ub}/V_{cb}|^2=4.936 \times 10^{-3}~ (\rho^2 +\eta^2 )\;.\label{eq:eqn9} \end{equation} Substituting the value $|V_{ub}/V_{cb} |=0.08\pm 0.02$, which is measured in charmless $b$ decays,\cite{ref03} into this relation, we obtain from Eq. (\ref{eq:eqn9}) \begin{equation} (\rho^2+\eta^2)^{1/2}=0.364\pm 0.091, \end{equation} which is in excellent agreement with the recent prediction \cite{ref04} $(\rho^2+\eta^2)^{1/2} =0.36 \pm 0.09 $. In this paper we have calculated the branching ratios of the exclusive nonleptonic decays of $\Lambda_b$ baryons using the covariant oscillator quark model, on the basis of the factorization approximation. The manifestly covariant weak currents are given by the overlapping integrals between the initial and final hadron wave functions. These currents are represented by a common form factor function, and for heavy $\to $ heavy baryon transitions this form factor function corresponds to the Isgur-Wise function of HQET and has similar properties at the zero recoil point. Using these currents we have derived the decay rates and up-down asymmetry parameter $\alpha $ for various nonleptonic weak decays of $\Lambda_b $ baryons. Our predicted result for the branching ratio $Br (\Lambda_b \to \Lambda J/\psi )$ is in agreement with the currently available experimental data. Recently, these decay processes have been studied using the quark model \cite{ref05,ref06} and HQET\cite{ref07,ref08} in the large $N_c $ limit. However, our predicted results for the branching ratios are smaller than the previous values, as can be seen from Table II. Future experimental data from the colliders are expected to verify and distinguish the various results. However, the values of the asymmetry parameter in all these calculations are nearly the same. Furthermore, the CKM-Wolfenstein parameter $(\rho^2+\eta^2)^{1/2}$ obtained in the framework of our model agrees very well with the recent prediction. \acknowledgements R. M. would like to thank CSIR, Govt. of India, for a fellowship. A. K. G. and M. P. K. acknowledge financial support from DST, Govt. of India.
\section{Introduction} The field of quantum computation studies the power of computers based on quantum mechanical principles. So far, most quantum algorithms---and {\em all} physically implemented ones---have operated in the so-called {\em black-box} setting. In the black-box model, the input of the function $f$ that we want to compute can only be accessed by means of queries to a ``black-box''. This returns the $i$th bit of the input when queried on $i$. The complexity of computing $f$ is measured by the required number of queries. In this setting we want quantum algorithms that use significantly fewer queries than the best classical algorithms. Examples of quantum black-box algorithms that are provably better than any classical algorithm can be found in \cite{deutsch&jozsa,simon:power,grover:search,bht:collision,bhmt:countingj,betal:distinctness}. Even Shor's quantum algorithm for period-finding, which is the core of his efficient factoring algorithm~\cite{shor:factoring}, can be viewed as a black-box algorithm~\cite{cleve:orderfinding}. We restrict our attention to computing total Boolean functions $f$ on $N$ variables. The query complexity of $f$ depends on the kind of errors one allows. For example, we can distinguish between exact computation, zero-error computation (a.k.a.~Las Vegas), and bounded-error computation (Monte Carlo). In each of these models, {\em worst-case} complexity is usually considered: the complexity is the number of queries required for the ``hardest'' input. Let $D(f)$, $R(f)$ and $Q(f)$ denote the worst-case query complexity of computing $f$ for classical deterministic algorithms, classical randomized bounded-error algorithms, and quantum bounded-error algorithms, respectively. More precise definitions will be given in the next section. Since quantum bounded-error algorithms are at least as powerful as classical bounded-error algorithms, and classical bounded-error algorithms are at least as powerful as deterministic algorithms, we have $Q(f)\leq R(f)\leq D(f)$. The main quantum success here is Grover's algorithm~\cite{grover:search}. It can compute the \mbox{\rm OR}-function with bounded-error using $\Theta(\sqrt{N})$ queries (which is optimal~\cite{bbbv:str&weak,bbht:bounds,zalka:grover}). Thus $Q(\mbox{\rm OR})\in\Theta(\sqrt{N})$, whereas $D(\mbox{\rm OR})=N$ and $R(\mbox{\rm OR})\in\Theta(N)$. This is the biggest gap known between quantum and classical worst-case complexities for total functions. (In contrast, for {\em partial} Boolean functions the gap can be much bigger~\cite{deutsch&jozsa,simon:power,cleve:orderfinding}.) In fact, it is known that the gap between $D(f)$ and $Q(f)$ is at most polynomial for {\em every} total $f$: $D(f)\in O(Q(f)^6)$~\cite{bbcmw:polynomials}. This is similar to the best known relation between classical deterministic and randomized algorithms: $D(f)\in O(R(f)^3)$~\cite{nisan:pram&dt}. Given some probability distribution $\mu$ on the set of inputs $\{0,1\}^N$ one may also consider {\em average-case} complexity instead of worst-case complexity. Average-case complexity concerns the {\em expected} number of queries needed when the input is distributed according to $\mu$. If the hard inputs receive little $\mu$-probability, then average-case complexity can be significantly smaller than worst-case complexity. Let $D^\mu(f)$, $R^\mu(f)$, and $Q^\mu(f)$ denote the average-case analogues of $D(f)$, $R(f)$, and $Q(f)$, respectively, to be defined more precisely in the next section. Again $Q^\mu(f)\leq R^\mu(f)\leq D^\mu(f)$. The objective of this paper is to compare these measures and to investigate the possible gaps between them. Our main results are: \begin{itemize} \item Under uniform $\mu$, $Q^\mu(f)$ and $R^\mu(f)$ can be super-exponentially smaller than $D^\mu(f)$. \item Under uniform $\mu$, $Q^\mu(f)$ can be exponentially smaller than $R^\mu(f)$. Thus the polynomial relation that holds between quantum and classical query complexities in the case of worst-case complexity~\cite{bbcmw:polynomials} does not carry over to the average-case setting. \item Under non-uniform $\mu$ the gap can be even larger: we give distributions $\mu$ where $Q^\mu(\mbox{\rm OR})$ is constant, whereas $R^\mu(\mbox{\rm OR})$ is almost $\sqrt{N}$. \item For every $f$ and $\mu$, $R^\mu(f)$ is lower bounded by the expected {\em block sensitivity} $E_\mu[bs(f)]$ and $Q^\mu(f)$ is lower bounded by $E_\mu[\sqrt{bs(f)}]$. \item For the MAJORITY-function under uniform $\mu$, we have that $Q^\mu(f)\in O(\sqrt{N}(\log N)^2)$ and $Q^\mu(f)\in\Omega(\sqrt{N})$. In contrast, $R^\mu(f)\in\Omega(N)$. \item For the PARITY-function, the gap between $Q^\mu$ and $R^\mu$ can be quadratic, but not more. Under uniform $\mu$, PARITY has $Q^\mu(f)\in\Omega(N)$. \end{itemize} \section{Definitions}\label{secdefs} Let $f:\{0,1\}^N\rightarrow\{0,1\}$ be a Boolean function. This function is {\em symmetric} if $f(X)$ only depends on $|X|$, the Hamming weight (the number of 1s) of $X$. We will in particular consider the following symmetric functions: $\mbox{\rm OR}(X)=1$ iff $|X|\geq 1$; $\mbox{\rm MAJ}(X)=1$ iff $|X|>N/2$; $\mbox{\rm PARITY}(X)=1$ iff $|X|$ is odd. If $X\in\{0,1\}^N$ is an input and $S$ a set of (indices of) variables, we use $X^S$ to denote the input obtained by flipping the values of the $S$-variables in $X$. The {\em block sensitivity} $bs_X(f)$ of $f$ on an input $X$ is the maximal number $b$ for which there are $b$ disjoint sets of variables $S_1,\ldots,S_b$ such that $f(X)\neq f(X^{S_i})$ for all $1\leq i\leq b$. The block sensitivity $bs(f)$ of $f$ is $\max_X bs_X(f)$. We are interested in the question how many bits of the input have to be queried in order to compute $f$, either for the worst-case or average-case input. We assume familiarity with classical computation and briefly sketch the definition of quantum query algorithms. For a general introduction to quantum computing, see the book of Nielsen and Chuang~\cite{nielsen&chuang:qc}. For more details about (quantum) query complexity we refer to~\cite{buhrman&wolf:dectreesurvey}. An $m$-qubit state is a $2^m$-dimensional unit vector of complex numbers, written $\sum_{x\in\{0,1\}^m}\alpha_x\ket{x}$. The complex number $\alpha_x$ is called the {\em amplitude} of the basis state $\ket{x}$. A $T$-query quantum algorithm corresponds to a unitary transformation $$ A=U_TOU_{T-1}O\ldots U_1OU_0. $$ Here the $U_j$ are unitary transformations on $m$ qubits. These $U_j$ are independent of the input. Each $O$ corresponds to a query to the input $X\in\{0,1\}^N$, formalized as the unitary transformation $$ \ket{i,b,z}\rightarrow\ket{i,b\oplus x_i,z}. $$ Here $i\in\{1,\ldots,N\}$, $b\in\{0,1\}$, $\oplus$ is addition modulo 2, and $z\in\{0,1\}^{m-\log N-1}$ is the workspace, which remains unaffected by the query. Intuitively, $O$ just gives us the bit $x_i$ when queried on $i$. We will sometimes use the word ``oracle'' to refer to $X$ as well as to the corresponding $O$. The initial state of the algorithm is the all-zero state $\ket{0^m}$. The final state is $A\ket{0^m}$, which depends on the input $X$ via the $T$ queries that are made. A measurement of a dedicated {\em output bit} of the final state will yield the output. It can be shown that this linear-algebraic quantum model is at least as strong as classical randomized computation: any classical $T$-query randomized algorithm can be simulated by a $T$-query quantum algorithm having the same error probabilities. As described above, the quantum algorithm will make exactly $T$ queries on {\em every} input $X$. Since we are interested in average-case number of queries and the required number of queries will depend on the input $X$, we need to allow the algorithm to give an output after fewer than $T$ queries. We will do that by measuring, after each $U_j$, a dedicated {\em flag-qubit} of the intermediate state at that point (this measurement may alter the state). This bit indicates whether the algorithm is already prepared to stop and output a value. If this bit is 1, then we measure the output bit, output its value $A(X)\in\{0,1\}$ and stop; if the flag-bit is 0 we let the algorithm continue with the next query $O$ and $U_{j+1}$. Note that the number of queries that the algorithm makes on input $X$ is now a random variable, since it depends on the probabilistic outcome of measuring the flag-qubit after each step. We use $T_A(X)$ to denote the {\em expected} number of queries that $A$ makes on input $X$. The Boolean output $A(X)$ of the algorithm is a random variable as well. We mainly focus on three kinds of algorithms for computing $f$: classical {\em deterministic}, classical {\em randomized} bounded-error, and {\em quantum} bounded-error algorithms. Let ${\cal D}(f)$ denote the set of classical {\em deterministic} algorithms that compute $f$. Let ${\cal R}(f)=\{\mbox{classical }A\mid\forall X\in\{0,1\}^N:\mbox{\rm Pr}[A(X)=f(X)]\geq 2/3\}$ be the set of classical {\em randomized} algorithms that compute $f$ with bounded error probability. The error probability $1/3$ is not essential; it can be reduced to any small $\varepsilon$ by running the algorithm $O(\log(1/\varepsilon))$ times and outputting the majority answer of those runs. Similarly we let ${\cal Q}(f)=\{\mbox{quantum }A\mid\forall X\in\{0,1\}^N:\mbox{\rm Pr}[A(X)=f(X)]\geq 2/3\}$ be the set of bounded-error {\em quantum} algorithms for $f$. We define the following worst-case complexities: \begin{eqnarray*} D(f) & = & \min_{A\in{\cal D}(f)}\max_{X\in\{0,1\}^N}T_A(X)\\ R(f) & = & \min_{A\in{\cal R}(f)}\max_{X\in\{0,1\}^N}T_A(X)\\ Q(f) & = & \min_{A\in{\cal Q}(f)}\max_{X\in\{0,1\}^N}T_A(X) \end{eqnarray*} $D(f)$ is also known as the {\em decision tree complexity} of $f$ and $R(f)$ as the {\em bounded-error} decision tree complexity of $f$. Since quantum computation generalizes randomized computation and randomized computation generalizes deterministic computation, we have $Q(f)\leq R(f)\leq D(f)\leq N$ for all $f$. The three worst-case complexities are polynomially related: $D(f)\in O(R(f)^3)$~\cite{nisan:pram&dt} and $D(f)\in O(Q(f)^6)$~\cite{bbcmw:polynomials} for all total $f$. Let $\mu:\{0,1\}^N\rightarrow[0,1]$ be a probability distribution. We define the {\em average-case complexity} of an algorithm $A$ with respect to a distribution $\mu$ as: $$ T_A^\mu=\sum_{X\in\{0,1\}^N}\mu(X)T_A(X). $$ The average-case deterministic, randomized, and quantum complexities of $f$ with respect to $\mu$ are \begin{eqnarray*} D^\mu(f) & = & \min_{A\in{\cal D}(f)}T^\mu_A\\ R^\mu(f) & = & \min_{A\in{\cal R}(f)}T^\mu_A\\ Q^\mu(f) & = & \min_{A\in{\cal Q}(f)}T^\mu_A \end{eqnarray*} Note that the algorithms still have to satisfy the appropriate output requirements (such as outputting $f(X)$ with probability $\geq 2/3$ in case of $R^\mu$ or $Q^\mu$) on {\em all} inputs $X$, even on $X$ that have $\mu(X)=0$. Clearly $Q^\mu(f)\leq R^\mu(f)\leq D^\mu(f)\leq N$ for all $\mu$ and $f$. Our goal is to examine how large the gaps between these measures can be, in particular for the uniform distribution $\mbox{\it unif\/}(X)=2^{-N}$. The above treatment of average-case complexity is the standard one used in average-case analysis of algorithms~\cite{vitter&flajolet:av}. One counter-intuitive consequence of these definitions, however, is that the average-case performance of polynomially related algorithms can be superpolynomially apart (we will see this happen in Section~\ref{secnonunifgap}). This seemingly paradoxical effect makes these definitions unsuitable for dealing with polynomial-time reducibilities and average-case complexity classes, which is what led Levin to his alternative definition of ``polynomial time on average''~\cite{levin:av}.% \footnote{We thank Umesh Vazirani for drawing our attention to this.} Nevertheless, we feel our definitions are the appropriate ones for our query complexity setting: they {\em are} just the average numbers of queries that one needs when the input is drawn according to distribution $\mu$. \section{Super-Exponential Gap between $D^{\mbox{\scriptsize\it unif}}(f)$ and $Q^{\mbox{\scriptsize\it unif}}(f)$} Before comparing the power of classical and quantum computing, we first compare the power of {\em deterministic} and {\em bounded-error} algorithms. It is not hard to show that $D^{\mbox{\scriptsize\it unif}}(f)$ can be much larger then $R^{\mbox{\scriptsize\it unif}}(f)$ and $Q^{\mbox{\scriptsize\it unif}}(f)$: \begin{theorem} Define $f$ on $N$ variables such that $f(X)=1$ iff $|X|\geq N/10$. Then $Q^{\mbox{\scriptsize\it unif}}(f)$ and $R^{\mbox{\scriptsize\it unif}}(f)$ are $O(1)$ and $D^{\mbox{\scriptsize\it unif}}(f)\in\Omega(N)$. \end{theorem} \begin{proof} Suppose we randomly sample $k$ bits of the input. Let $a=|X|/N$ denote the fraction of 1s in the input and $\tilde{a}$ the fraction of 1s in the sample. The Chernoff bound (see e.g.~\cite{alon&spencer:probmethod}) implies that there is a constant $c>0$ such that $$ \mbox{\rm Pr}[\tilde{a}<2/10\mid a\geq 3/10]\leq 2^{-ck}. $$ Now consider the following randomized algorithm for $f$: \begin{enumerate} \item Let $i=100$. \item Sample $k_i=i/c$ bits. If the fraction $\tilde a_i$ of 1s is $\geq 2/10$, then output 1 and stop. \item If $i<\log N$, then increase $i$ by 1 and repeat step 2. \item If $i\geq\log N$, then count $|X|$ exactly using $N$ queries and output the correct answer. \end{enumerate} It is easy to see that this is a bounded-error algorithm for $f$. Let us bound its average-case complexity under the uniform distribution. If $a\geq 3/10$, the expected number of queries for step 2 is $$ \sum_{i=100}^{\log N} \mbox{\rm Pr}[\tilde a_1\leq 2/10, \ldots, \tilde a_{i-1}\leq 2/10 \mid a\geq 3/10]\cdot\frac{i}{c} \leq $$ $$ \sum_{i=100}^{\log N} \mbox{\rm Pr}[\tilde a_{i-1}\leq 2/10 \mid a \geq3/10]\cdot \frac{i}{c} \leq \sum_{i=100}^{\log N} 2^{-(i-1)} \cdot\frac{i}{c} \in O(1). $$ The probability that step 4 is needed (given $a\geq 3/10$) is at most $2^{-c \log N/c}=1/N$. This adds $\frac{1}{N} N=1$ to the expected number of queries. Under the uniform distribution, the probability of the event $a<3/10$ is at most $2^{-c' N}$ for some constant $c'$. This case contributes at most $2^{-c' N} (N+(\log N)^2) \in o(1)$ to the expected number of queries. Thus in total the algorithm uses $O(1)$ queries on average, hence $R^{\mbox{\scriptsize\it unif}}(f)\in O(1)$. Since $Q^{\mbox{\scriptsize\it unif}}(f)\leq R^{\mbox{\scriptsize\it unif}}(f)$, we also have $Q^{\mbox{\scriptsize\it unif}}(f)\in O(1)$. Since a deterministic classical algorithm for $f$ must be correct on every input $X$, it is easy to see that it must make at least $N/10$ queries on every input, hence $D^{\mbox{\scriptsize\it unif}}(f)\geq N/10$. \end{proof} Accordingly, we can have huge gaps between $D^{\mbox{\scriptsize\it unif}}(f)$ and $Q^{\mbox{\scriptsize\it unif}}(f)$. However, this example tells us nothing about the gaps between quantum and classical bounded-error algorithms. In the next section we exhibit an $f$ where $Q^{\mbox{\scriptsize\it unif}}(f)$ is exponentially smaller than the classical bounded-error complexity $R^{\mbox{\scriptsize\it unif}}(f)$. \section{Exponential Gap between $R^{unif}(f)$ and $Q^{unif}(f)$}\label{secgapunif} \subsection{The Function} We use the following modification of Simon's problem~\cite{simon:power}:% \footnote{The preprint~\cite{hhz:aesuperiority} independently proves a related but incomparable result about another Simon-modification.} \medskip \noindent {\bf Input:} $X=(x_1,\ldots,x_{2^n})$, where each $x_i\in\{0,1\}^n$. \noindent {\bf Output:} $f(X)=1$ iff there is a non-zero $k\in\{0,1\}^n$ such that for all $i\in\{0,1\}^n$ we have $x_{i\oplus k}=x_i$. \medskip Here we treat $i\in\{0,1\}^n$ both as an $n$-bit string and as a number between $1$ and $2^n$, and $\oplus$ denotes bitwise XOR. Note that this function is total (unlike Simon's). Formally, $f$ is not a Boolean function because the variables are $\{0,1\}^n$-valued. However, we can replace every variable $x_i$ by $n$ Boolean variables and then $f$ becomes a Boolean function of $N=n 2^n$ variables. The number of queries needed to compute the Boolean function is at least the number of queries needed to compute the function with $\{0,1\}^n$-valued variables (because we can simulate a query to the Boolean oracle by means of a query to the $\{0,1\}^n$-valued input-variables, just ignoring the $n-1$ bits that we are not interested in) and at most $n$ times the number of queries to the $\{0,1\}^n$-valued oracle (because one $\{0,1\}^n$-valued query can be simulated using $n$ Boolean queries). As the numbers of queries are so closely related, it does not make a big difference whether we use the $\{0,1\}^n$-valued oracle or the Boolean oracle. For simplicity we count queries to the $\{0,1\}^n$-valued oracle. We are interested in the average-case complexity of this function. The main result is the following exponential gap, to be proven in the next sections: \begin{theorem}\label{thuniformgap} For $f$ as above, $Q^{unif}(f)\leq 22n+1$ and $R^{unif}(f)\in\Omega(2^{n/2})$. \end{theorem} \subsection{Quantum Upper Bound} The quantum algorithm is similar to Simon's. Start with the 2-register superposition $\sum_{i\in\{0,1\}^n}\ket{i}\ket{0}$ (for convenience we ignore normalizing factors). Apply the oracle once to obtain $$ \sum_{i\in\{0,1\}^n}\ket{i}\ket{x_i}. $$ Measuring the second register gives some $j$ and collapses the first register to $$ \sum_{i:x_i=j}\ket{i}. $$ A {\em Hadamard transform} $H$ maps bits $\ket{b}\rightarrow\frac{1}{\sqrt{2}}(\ket{0}+(-1)^b\ket{1})$. Applying this to each qubit of the first register gives \begin{equation}\label{eqnsupsimon} \sum_{i:x_i=j}\sum_{i'\in\{0,1\}^n}(-1)^{(i,i')}\ket{i'}. \end{equation} Here $(a,b)$ denotes inner product mod 2; if $(a,b)=0$ we say $a$ and $b$ are orthogonal. If $f(X)=1$, then there is a non-zero $k$ such that $x_i=x_{i\oplus k}$ for all $i$. In particular, $x_i=j$ iff $x_{i\oplus k}=j$. Then the final state~(\ref{eqnsupsimon}) can be rewritten as \begin{eqnarray*} \sum_{i'\in\{0,1\}^n}\sum_{i:x_i=j}(-1)^{(i,i')}\ket{i'} & = & \sum_{i'\in\{0,1\}^n}\left(\sum_{i:x_i=j}\frac{1}{2}((-1)^{(i,i')}+(-1)^{(i\oplus k,i')})\right)\ket{i'}\\ & = & \sum_{i'\in\{0,1\}^n}\left(\sum_{i:x_i=j}\frac{(-1)^{(i,i')}}{2}(1+(-1)^{(k,i')})\right)\ket{i'}. \end{eqnarray*} Notice that $\ket{i'}$ has non-zero amplitude only if $(k,i')=0$. Hence if $f(X)=1$, then measuring the final state gives some $i'$ orthogonal to the unknown $k$. To decide if $f(X)=1$, we repeat the above process $m=22 n$ times. Let $i_1,\ldots,i_m\in\{0,1\}^n$ be the results of the $m$ measurements. If $f(X)=1$, there must be a non-zero $k$ that is orthogonal to all $i_r$. Compute the subspace $S\subseteq\{0,1\}^n$ that is generated by $i_1,\ldots,i_m$ (i.e.~$S$ is the set of binary vectors obtained by taking linear combinations of $i_1,\ldots,i_m$ over $GF(2)$). If $S=\{0,1\}^n$, then the only $k$ that is orthogonal to all $i_r$ is $k=0^n$, so then we know that $f(X)=0$. If $S\neq\{0,1\}^n$, we just query all $2^n$ values $x_{0\ldots 0},\ldots,x_{1\ldots 1}$ and then compute $f(X)$. Of course, this latter step is very expensive, but it is needed only rarely: \begin{lemma} Assume that $X=(x_{0\ldots 0},\ldots,x_{1\ldots 1})$ is chosen uniformly at random from $\{0,1\}^N$. Then, with probability at least $1-2^{-n}$, $f(X)=0$ and the measured $i_1, \ldots, i_m$ generate $\{0,1\}^n$. \end{lemma} \begin{proof} It can be shown by a small modification of~\cite[Theorem~5.1, p.91]{alon&spencer:probmethod} that with probability at least $1-2^{-c2^{n}}$ ($c>0$), there are at least $2^n/8$ values $j$ such that $x_i=j$ for exactly one $i\in\{0,1\}^n$ (and hence $f(X)=0$). We assume that this is the case in the following. If $i_1,\ldots,i_m$ generate a proper subspace of $\{0,1\}^n$, then there is a non-zero $k\in\{0,1\}^n$ that is orthogonal to this subspace. We estimate the probability that this happens. Consider some fixed non-zero vector $k\in\{0,1\}^n$. The probability that $i_1$ and $k$ are orthogonal is at most $\frac{15}{16}$, as follows. With probability at least 1/8, the measurement of the second register gives $j$ such that $f(i)=j$ for a unique $i$. In this case, the measurement of the final superposition~(\ref{eqnsupsimon}) gives a uniformly random $i'$. The probability that a uniformly random $i'$ has $(k,i')\neq 0$ is 1/2. Therefore, the probability that $(k,i_1)=0$ is at most $1-\frac{1}{8}\cdot\frac{1}{2}=\frac{15}{16}$. The vectors $i_1,\ldots,i_m$ are chosen independently. Therefore, the probability that $k$ is orthogonal to each of them is at most $(\frac{15}{16})^m=(\frac{15}{16})^{22 n}<2^{-2n}$. There are $2^n-1$ possible non-zero $k$, so the probability that there is a $k$ which is orthogonal to each of $i_1,\ldots,i_m$, is $\leq(2^n-1)2^{-2n}<2^{-n}$. \end{proof} Note that this algorithm is actually a {\em zero-error} algorithm: it always outputs the correct answer. Its expected number of queries on a uniformly random input is at most $m=22n$ for generating $i_1, \ldots, i_m$ and at most $\frac{1}{2^n} 2^n=1$ for querying all the $x_i$ if the first step does not give $i_1,\ldots,i_m$ that generate $\{0,1\}^n$. This completes the proof of the first part of Theorem~\ref{thuniformgap}. In contrast, in the appendix we show that the {\em worst-case} zero-error quantum complexity of $f$ is $\Omega(N)$, which is near-maximal. \subsection{Classical Lower Bound} Let $D_1$ be the uniform distribution over all inputs $X\in\{0,1\}^N$ and $D_2$ be the uniform distribution over all $X$ for which there is a unique $k\neq 0$ such that $x_i=x_{i\oplus k}$ (and hence $f(X)=1$). We say an algorithm $A$ {\em distinguishes} between $D_1$ and $D_2$ if the average probability that $A$ outputs 0 is $\geq 2/3$ under $D_1$ and the average probability that $A$ outputs 1 is $\geq 2/3$ under $D_2$. \begin{lemma}\label{clb1} If there is a bounded-error algorithm $A$ that computes $f$ with $m=T_A^{unif}$ queries on average, then there is an algorithm that distinguishes between $D_1$ and $D_2$ and uses $O(m)$ queries on all inputs. \end{lemma} \begin{proof} Without loss of generality we assume $A$ has error probability $\leq 1/10$. To distinguish $D_1$ and $D_2$, we run $A$ until it stops or makes $10m$ queries. If it stops, we output the result of $A$. If it makes $10m$ queries and has not stopped yet, we output 1. Under $D_1$, the probability that $A$ outputs 1 is at most $1/10+o(1)$ ($1/10$ is the maximum probability of error on an input with $f(X)=0$ and $o(1)$ is the probability of getting an input with $f(X)=1$), so the probability that $A$ outputs 0 is at least $9/10-o(1)$. The average probability (under $D_1$) that $A$ does not stop before $10m$ queries is at most $1/10$, for otherwise the average number of queries would be more than $\frac{1}{10}(10m)=m$. Therefore the probability under $D_1$ that $A$ outputs 0 after at most $10m$ queries, is at least $(9/10-o(1))-1/10=4/5-o(1)$. In contrast, the $D_2$-probability that $A$ outputs 0 is $\leq 1/10$ because $f(X)=1$ for any input $X$ from $D_2$. This shows that we can distinguish $D_1$ from $D_2$. \end{proof} \begin{lemma}\label{clb2} A classical randomized algorithm $A$ that makes $m\in o(2^{n/2})$ queries cannot distinguish between $D_1$ and $D_2$. \end{lemma} \begin{proof} For a random input from $D_1$, the probability that all answers to $m$ queries are different is $$ 1\cdot \left(1-\frac{1}{2^n}\right) \cdots \left(1-\frac{(m-1)}{2^n}\right)\geq 1-\sum_{i=1}^{m-1}\frac{i}{2^n}=1-\frac{m(m-1)}{2^{n+1}} = 1-o(1). $$ For a random input from $D_2$, the probability that there is an $i$ such that $A$ queries both $x_i$ and $x_{i\oplus k}$ ($k$ is the hidden vector) is $\leq {m\choose 2}/(2^n-1)\in o(1)$, since: \begin{enumerate} \item for every pair of distinct $i,j$, the probability that $i=j\oplus k$ is $1/(2^n-1)$ \item since $A$ queries only $m$ of the $x_i$, it queries only ${m\choose 2}$ distinct pairs $i,j$ \end{enumerate} If no pair $x_i$, $x_{i\oplus k}$ is queried, the probability that all answers are different is $$ 1\cdot \left(1-\frac{1}{2^{n-1}}\right) \cdots \left(1-\frac{(m-1)}{2^{n-1}}\right) = 1-o(1). $$ It is easy to see that all sequences of $m$ different answers are equally likely. Therefore, for both distributions $D_1$ and $D_2$, we get a uniformly random sequence of $m$ different values with probability $1-o(1)$ and something else with probability $o(1)$. Thus $A$ cannot ``see'' the difference between $D_1$ and $D_2$ with sufficient probability to distinguish between them. \end{proof} The second part of Theorem~\ref{thuniformgap} now follows: a classical algorithm that computes $f$ with an average number of $m$ queries can be used to distinguish between $D_1$ and $D_2$ with $O(m)$ queries (Lemma~\ref{clb1}), but then $O(m)\in\Omega(2^{n/2})$ (Lemma~\ref{clb2}). \section{Super-Exponential Gap for Non-Uniform $\mu$}\label{secnonunifgap} The last section gave an exponential gap between $Q^\mu$ and $R^\mu$ under uniform $\mu$. Here we show that the gap can be even larger for non-uniform $\mu$. Consider the average-case complexity of the \mbox{\rm OR}-function. It is easy to see that $D^{unif}(\mbox{\rm OR})$, $R^{unif}(\mbox{\rm OR})$, and $Q^{unif}(\mbox{\rm OR})$ are all $O(1)$, since the average input will have many 1s under the uniform distribution. Now we give some examples of non-uniform distributions $\mu$ where $Q^\mu(\mbox{\rm OR})$ is super-exponentially smaller than $R^\mu(\mbox{\rm OR})$: \begin{theorem}\label{thoravgap} If $\alpha\in(0,1/2)$ and $\mu(X)=c/{N\choose |X|}(|X|+1)^\alpha(N+1)^{1-\alpha}$ ($c\approx 1-\alpha$ is a normalizing constant), then $R^\mu(\mbox{\rm OR})\in\Theta(N^\alpha)$ and $Q^\mu(\mbox{\rm OR})\in\Theta(1)$. \end{theorem} \begin{proof} Any classical algorithm for \mbox{\rm OR}\ requires $\Theta(N/(|X|+1))$ queries on an input $X$. The upper bound follows from random sampling, the lower bound from a block-sensitivity argument~\cite{nisan:pram&dt}. Hence (omitting the intermediate $\Theta$s): $$ R^\mu(\mbox{\rm OR})=\sum_X\mu(X)\frac{N}{|X|+1}= \sum_{t=0}^N\frac{cN^\alpha}{(t+1)^{\alpha+1}} \in\Theta(N^\alpha), $$ where the last step can be shown by approximating the sum over $t$ with an integral. Similarly, for a quantum algorithm $\Theta(\sqrt{N/(|X|+1))}$ queries are necessary and sufficient on an input $X$~\cite{grover:search,bbht:bounds}, so $$ Q^\mu(\mbox{\rm OR})=\sum_X\mu(X)\sqrt{\frac{N}{|X|+1}}= \sum_{t=0}^N\frac{cN^{\alpha-1/2}}{(t+1)^{\alpha+1/2}} \in\Theta(1). $$ \end{proof} In particular, for $\alpha=1/2-\varepsilon$ we have the very large gap of $O(1)$ quantum versus $\Omega(N^{1/2-\varepsilon})$ classical. Note that we obtain this super-exponential gap by weighing the complexity of two algorithms (classical and quantum \mbox{\rm OR}-algorithms) which are only quadratically apart on each input $X$. This is the phenomenon we referred to at the end of Section~\ref{secdefs}. \section{General Bounds for Average-Case Complexity} In this section we prove some general bounds. First we make precise the intuitively obvious fact that if an algorithm $A$ is faster on every input than another algorithm $B$, then it is also faster on average under any distribution: \begin{theorem} If $\phi:\mbox{\bf R}\rightarrow\mbox{\bf R}$ is a concave function and $T_A(X)\leq \phi(T_B(X))$ for all $X$, then $\displaystyle T^\mu_A\leq \phi\left(T^\mu_B\right)$ for every $\mu$. \end{theorem} \begin{proof} By Jensen's inequality, if $\phi$ is concave then $E_\mu[\phi(T)]\leq\phi(E_\mu[T])$, hence $$ T^\mu_A =\sum_{X\in\{0,1\}^N}\mu(X)T_A(X) \leq\sum_{X\in\{0,1\}^N}\mu(X)\phi(T_B(X)) \leq\phi\left(\sum_{X\in\{0,1\}^N}\mu(X)T_B(X)\right)=\phi\left(T^\mu_B\right). $$ \end{proof} In words: taking the average cannot make the complexity-gap between two algorithms smaller. For instance, if $T_A(X)\leq\sqrt{T_B(X)}$ (say, $A$ is Grover's algorithm and $B$ is a classical algorithm for \mbox{\rm OR}), then $T^\mu_A\leq\sqrt{T^\mu_B}$. On the other hand, taking the average {\em can} make the gap much larger, as we saw in Theorem~\ref{thoravgap}: the quantum algorithm for \mbox{\rm OR}\ runs only quadratically faster than any classical algorithm on each input, but the {\em average-case} gap between quantum and classical can be much bigger than quadratic. We now prove a general lower bound on $R^\mu$ and $Q^\mu$. The classical case of the following lemma was shown in~\cite{nisan:pram&dt}, the quantum case in~\cite{bbcmw:polynomials}: \begin{lemma} Let $A$ be a bounded-error algorithm for some function $f$. If $A$ is classical then $T_A(X)\in\Omega(bs_X(f))$, and if $A$ is quantum then $T_A(X)\in\Omega(\sqrt{bs_X(f)})$. \end{lemma} A lower bound in terms of the $\mu$-expected block sensitivity follows: \begin{theorem}\label{thavgbsbound} For all $f$, $\mu$: $R^\mu(f)\in\Omega(E_\mu[bs_X(f)])$ and $Q^\mu(f)\in\Omega(E_\mu[\sqrt{bs_X(f)}])$. \end{theorem} \section{Average-Case Complexity of MAJORITY} Here we examine the average-case complexity of the MAJORITY-function. The hard inputs for majority occur when $t=|X|\approx N/2$. Any quantum algorithm needs $\Omega(N)$ queries for such inputs~\cite{bbcmw:polynomials}. Since the uniform distribution puts most probability on the set of $X$ with $|X|$ close to $N/2$, we might expect an $\Omega(N)$ average-case complexity as well. However, we will prove that the complexity is nearly $\sqrt{N}$. For this we need the following result about approximate quantum counting, which is Theorem~13 of~\cite{bhmt:countingj} (this is the upcoming journal version of~\cite{bht:counting} and~\cite{mosca:eigen}; see also~\cite[Theorem~1.10]{nayak&wu:median}): \begin{theorem}[Brassard, H{\o}yer, Mosca, Tapp]\label{thqcounting} There exists a quantum algorithm {\bf QCount} with the following property. For every $N$-bit input $X$ (with $t=|X|$) and number of queries $T$, and any integer $k\geq 1$, {\bf QCount} uses $T$ queries and outputs a number $\tilde{t}$ such that $$ |t-\tilde{t}|\leq 2\pi k\frac{\sqrt{t(N-t)}}{T}+\pi^2k^2\frac{N}{T^2} $$ with probability at least $8/\pi^2$ if $k=1$ and probability $\geq 1- 1/2(k-1)$ if $k>1$. \end{theorem} Using repeated applications of this quantum counting routine we can obtain a quantum algorithm for majority that is fast on average: \begin{theorem}\label{thavqmaj} $Q^{\mbox{\scriptsize\it unif}}(\mbox{\rm MAJ})\in O(\sqrt{N}(\log N)^2)$. \end{theorem} \begin{proof} For all $i\in\{1,\ldots,\log N\}$, define $A_i=\{X\mid N/2^{i+1}< \left| |X|-N/2 \right|\leq N/2^i\}$. The probability under the uniform distribution of getting an input $X\in A_i$ is $\mu(A_i)\in O(\sqrt{N}/2^i)$, since the number of inputs $X$ with $k$ 1s is ${N\choose k}\in O(2^N/\sqrt{N})$ for all $k$. The idea of our algorithm is to have $\log N$ runs of the quantum counting algorithm, with increasing numbers of queries, such that the majority value of inputs from $A_i$ is probably detected around the $i$th counting stage. We will use $T_i=100\cdot 2^i\log N$ queries in the $i$th counting stage. Our MAJORITY-algorithm is the following: \begin{quote} For $i=1$ to $\log N$ do:\\[1mm] \hspace*{7mm}quantum count $|X|$ using $T_i$ queries (call the estimate $\tilde{t}_i$)\\ \hspace*{7mm}if $|\widetilde{t}_i-N/2|>N/2^i$, then output whether $\widetilde{t}_i>N/2$ and stop.\\[1mm] Classically count $|X|$ using $N$ queries and output its majority. \end{quote} Let us analyze the behavior of the algorithm on an input $X\in A_i$. For $t=|X|$, we have $|t-N/2|\in(N/2^{i+1},N/2^i]$. By Theorem~\ref{thqcounting}, with probability $>1-1/10\log N$ we have $\left|\widetilde{t}_i-t\right|\leq N/2^i$, so with probability $(1-1/10\log N)^{\log N}\approx e^{-1/10}>0.9$ we have $\left|\widetilde{t}_i-t\right|\leq N/2^i$ for all $1\leq i\leq N$. This ensures that the algorithm outputs the correct value with high probability. We now bound the expected number of queries the algorithm needs on input $X$. Consider the $(i+2)$nd counting stage. With probability $1-1/10\log N$ we will have $|\tilde{t}_{i+2}-t|\leq N/2^{i+2}$. In this case the algorithm will terminate, because $$ |\tilde{t}_{i+2}-N/2|\geq |t-N/2|-|\tilde{t}_{i+2}-t|>N/2^{i+1}-N/2^{i+2}=N/2^{i+2}. $$ Thus with high probability the algorithm needs no more than $i+2$ counting stages on input $X$. Later counting stages take exponentially more queries ($T_{i+2+j}=2^jT_{i+2}$), but are needed only with exponentially decreasing probability $O(1/2^j\log N)$: the probability that $|\tilde{t}_{i+2+j}-t|>N/2^{i+2}$ goes down exponentially with $j$ precisely because the number of queries goes up exponentially. Similarly, the last step of the algorithm (classical counting) is needed only with negligible probability. Now the expected number of queries on input $X$ can be upper bounded by $$ \sum_{j=1}^{i+2}T_i + \sum_{k=i+3}^{\log N}T_k\cdot O\left(\frac{1}{2^{k-i-3}\log N}\right) < 100\cdot 2^{i+3}\log N + \sum_{k=i+3}^{\log N}100\cdot 2^{i+3} \in O(2^i\log N). $$ Therefore under the uniform distribution the average expected number of queries can be upper bounded by $\sum_{i=1}^{\log N} \mu(A_i)O(2^i\log N)\in O(\sqrt{N}(\log N)^2).$ \end{proof} The nearly matching lower bound is: \begin{theorem}\label{thavqmajlower} $Q^{\mbox{\scriptsize\it unif}}(\mbox{\rm MAJ})\in\Omega(\sqrt{N})$. \end{theorem} \begin{proof} Let $A$ be a bounded-error quantum algorithm for MAJORITY. It follows from the worst-case results of~\cite{bbcmw:polynomials} that $A$ uses $\Omega(N)$ queries on the hardest inputs, which are the $X$ with $|X|=N/2\pm 1$. Since the uniform distribution puts $\Omega(1/\sqrt{N})$ probability on the set of such $X$, the average-case complexity of $A$ is at least $\Omega(1/\sqrt{N})\Omega(N)=\Omega(\sqrt{N})$. \end{proof} What about the {\em classical} average-case complexity of MAJORITY? Alonso, Reingold, and Schott~\cite{ars:avmaj} prove the bound $D^{\mbox{\scriptsize\it unif}}(\mbox{\rm MAJ})=2N/3-\sqrt{8N/9\pi}+O(\log N)$ for deterministic classical computers. We can also prove a linear lower bound for the {\em bounded-error} classical complexity, using the following lemma: \begin{lemma} Let $\Delta\in\{1,\ldots,\sqrt{N}\}$. Any classical bounded-error algorithm that computes MAJORITY on inputs $X$ with $|X|\in\{N/2,N/2+\Delta\}$ must make $\Omega(N)$ queries on all such inputs. \end{lemma} \begin{proof} We will prove the lemma for $\Delta=\sqrt{N}$, which is the hardest case. We assume without loss of generality that the algorithm queries its input $X$ at $T(X)$ random positions, and outputs 1 if the fraction of 1s in its sample is at least $(N/2+\Delta)/N=1/2+1/\sqrt{N}$. We do not care what the algorithm outputs otherwise. Consider an input $X$ with $|X|=N/2$. The algorithm uses $T=T(X)$ queries and should output 0 with probability at least $2/3$. Thus the probability of output 1 on $X$ must be at most $1/3$, in particular $$ \mbox{\rm Pr}[\mbox{ at least $T(1/2+1/\sqrt{N})$ 1s in sample of size $T$}]\leq 1/3. $$ Since the $T$ queries of the algorithm can be viewed as sampling without replacement from a set containing $N/2$ 1s and $N/2$ 0s, this error probability is given by the hypergeometric distribution $$ \mbox{\rm Pr}[\mbox{ at least $T(1/2+1/\sqrt{N})$ 1s in sample of size $T$}]= \frac{\displaystyle \sum_{i=T(1/2+1/\sqrt{N})}^{T}{{N/2}\choose i}\cdot {{N/2}\choose{T-i}}} {\displaystyle {N\choose T}}. $$ We can approximate the hypergeometric distribution using the normal distribution, see e.g.~\cite{nicholson:hyper}. Let $z_k=(2k-T)/\sqrt{T}$ and $\Phi(z)=\int_{-\infty}^z\frac{1}{\sqrt{2\pi}}e^{-t^2/2}dt$, then the above probability approaches $$ \Phi(z_T)-\Phi(z_{T(1/2+1/\sqrt{N})}). $$ Note that $\Phi(z_T)=\Phi(\sqrt{T})\rightarrow 1$ and that $\Phi(z_{T(1/2+1/\sqrt{N})})=\Phi(2\sqrt{T/N})\rightarrow 1/2$ if $T\in o(N)$. Thus we can only avoid having an error probability close to 1/2 by using $T\in\Omega(N)$ queries on $X$ with $|X|=N/2$. A similar argument shows that we must also use $\Omega(N)$ queries if $|X|=N/2+\Delta$. \end{proof} It now follows that: \begin{theorem} $R^{\mbox{\scriptsize\it unif}}(\mbox{\rm MAJ})\in\Omega(N)$. \end{theorem} \begin{proof} The previous lemma shows that any algorithm for MAJORITY needs $\Omega(N)$ queries on inputs $X$ with $|X|\in[N/2,N/2+\sqrt{N}]$. Since the uniform distribution puts $\Omega(1)$ probability on the set of such $X$, the theorem follows. \end{proof} Accordingly, {\em on average} a quantum computer can compute MAJORITY almost quadratically faster than a classical computer, whereas for the {\em worst-case} input quantum and classical computers are about equally fast (or slow). \section{Average-Case Complexity of PARITY} Finally we prove some results for the average-case complexity of PARITY. This is in many ways the hardest Boolean function. Firstly, $bs_X(f)=N$ for all $X$, hence by Theorem~\ref{thavgbsbound}: \begin{corollary} For every $\mu$, $R^\mu(\mbox{\rm PARITY})\in\Omega(N)$ and $Q^\mu(\mbox{\rm PARITY})\in\Omega(\sqrt{N})$. \end{corollary} With high probability we can obtain an exact count of $|X|$, using $O(\sqrt{(|X|+1)N})$ quantum queries~\cite{bhmt:countingj}. Combining this with a $\mu$ that puts $O(1/\sqrt{N})$ probability on the set of all $X$ with $|X|>1$ and distributes the remaining probability arbitrarily over the $X$ with $|X|\leq 1$, we obtain a distribution $\mu$ such that $Q^\mu(\mbox{\rm PARITY})\in O(\sqrt{N})$. We can prove $Q^\mu(\mbox{\rm PARITY})\leq N/6$ for any $\mu$ by the following algorithm: with probability $1/3$ output 1, with probability $1/3$ output 0, and with probability $1/3$ run the exact quantum algorithm for PARITY, which has worst-case complexity $N/2$~\cite{bbcmw:polynomials,fggs:parity}. This algorithm has success probability $2/3$ on every input and has expected number of queries equal to $N/6$. More than a linear speed-up on average is not possible if $\mu$ is uniform: \begin{theorem} $Q^{\mbox{\scriptsize\it unif}}(\mbox{\rm PARITY})\in\Omega(N)$. \end{theorem} \begin{proof} Let $A$ be a bounded-error quantum algorithm for PARITY. Let $B$ be an algorithm that flips each bit of its input $X$ with probability $1/2$, records the number $b$ of actual bitflips, runs $A$ on the changed input $Y$, and outputs $A(Y)+ b\mbox{ mod }2$. It is easy to see that $B$ is a bounded-error algorithm for PARITY and that it uses an {\em expected} number of $T_A^\mu$ queries on {\em every} input. Using standard techniques, we can turn this into an algorithm for PARITY with {\em worst-case} $O(T_A^\mu)$ queries. Since the worst-case lower bound for PARITY is $N/2$~\cite{bbcmw:polynomials,fggs:parity}, the theorem follows. \end{proof} \subsection*{Acknowledgments} We thank Harry Buhrman for suggesting this topic, and him, Lance Fortnow, Lane Hemaspaandra, Hein R\"ohrig, Alain Tapp, and Umesh Vazirani for helpful discussions. Also thanks to Alain for sending a draft of~\cite{bhmt:countingj}.
\section{Introduction} The AdS/CFT correspondence\cite{Mald} relates string theories and M-theory solutions with an $AdS$ factor to the infrared conformal field theory living on the brane system. This was further elaborated upon in Refs. \cite{Wi,GKP} where it was explained how one calculates the spectrum and correlation functions of the conformal field theory within a supergravity description. Tests of the conjecture involve comparing the supergravity correlation functions and dimensions of operators with the ones associated to the low energy dynamics of the brane. In order to construct tests that probe beyond BPS states and the large $g^2N$ limit, it is necessary to understand the full spectrum of excitations in these backgrounds. The first example\footnote{For earlier work, see Refs. \cite{morerefs}.} was discussed in \cite{GKS} where a weak coupling regime exists at finite $N$ (see also \cite{BORT,newKS}). Further examples have appeared in Refs. \cite{KLL,EFGT}. In general, in order to probe physics where supersymmetry is less restrictive but still a powerful tool, we want to develop a large collection of spacetimes that are good string backgrounds and that can be understood in a simple fashion. It is the purpose of this paper to give a large class of backgrounds for which this program is possible in principle. Our construction gives rise to spaces with $N=2$ spacetime supersymmetry, and they are derived from a worldsheet $N=2$ superconformal field theory. After we completed this paper, we learned of the results of Ref. \cite{GR} which overlaps some of our results. \section{$AdS_3$ backgrounds} We wish to discuss general $NSNS$ backgrounds of Type II superstrings which lead to spacetime supersymmetry on $AdS_3$. We will require that a large radius limit of the geometry exists, such that a semiclassical analysis can be made valid; this is done in order to be able to compare with low energy supergravity. We will restrict ourselves to coset conformal field theories, for which the semiclassical limit corresponds to taking all levels of the corresponding algebras to infinity simultaneously. Moreover, we will be interested in theories with $N=2$ worldsheet superconformal invariance, which will leads naturally to spacetime supersymmetry. These two requirements lead to the following constraints. The spacetime is of the form $AdS_3 \times X$, with $X$ a seven-dimensional manifold/orbifold. $\hbox{N=2}$ worldsheet supersymmetry is obtained if $X$ is chosen to be a circle bundle over a complex manifold, which itself is a product of hermitian symmetric spaces constructed as in Kazama-Suzuki\cite{KZ}. (A possible generalization would be to take more general complex manifolds $X/U(1)$ with constant positive curvature, but this is less well understood.) This bundle structure allows for a decomposition of the energy momentum tensor in terms of a $U(1)$ CFT and a hermitian symmetric space coset. This $U(1)$ pairs with the Cartan of the $SL(2)$ current algebra into an $N=2$ free system, and the remaining $SL(2,\BR)/U(1)$ is also a hermitian symmetric non-compact coset. This splitting of the $U(1)$ piece of $X$ is required if we want to bosonize the fermions in the $SL(2,\BR)$ current algebra. In total, we have a system with $N=2$ worldsheet supersymmetry, \begin{equation} \left( SL(2,\hbox{\xiiss R\kern-.45emI}}\newcommand{\BR}{\RR)\slash U(1)\right)\times \left( X\slash U(1)\right)\times U(1)^2 \end{equation} This construction possesses a ``natural'' GSO projection (preserving the $N=2$ structure) which leads to spacetime supersymmetry. Of course, there are other possible GSO projections, such as that used in the construction of the NS vacua of the $N=4$ geometries considered in \cite{GKS,KLL,EFGT}. Each of these GSO projections will be considered below. The collection of models constructed in this fashion is a finite family of spaces, and they should be interpreted as conifolds where a set of D1-D5 branes is localized. A table with a list of compact cosets of dimension $d\leq 7$ is provided here, where we include the central charge of the model in terms of the levels of the different groups that form the coset. Other spaces like $SU(3)/U(1)^2$ can be constructed as $SU(3)/(SU(2)\times U(1))\times SU(2)/U(1)$, so we don't list them. \renewcommand{\arraystretch}{1.5} \begin{table}[h] \begin{tabular}{c|c|c|c} Coset & Central charge & Group& Dimension\\ \hline $S^1$& $\frac 32$ & U(1) & 1\\ $\hbox{\xiiss C\kern-.4emIP}^1$ & $3-\frac 6 k$ & $SU(2)_k\slash U(1)$ & 2\\ $S^3$&$\frac 92 -\frac 6k$ & $SU(2)_k$&3\\ $\hbox{\xiiss C\kern-.4emIP}^2$ & $6-\frac {18} k$ & $SU(3)_k/(SU(2)_k\times U(1))$ & 4\\ $S^5$ & $\frac{15}2 - \frac {18} k$ & $SU(3)_k/SU(2)_k$ & 5 \\ $T^{pq}$ & $\frac {15}2 - \frac 6k -\frac 6{k'}$ & $(SU(2)_k\times SU(2)_{k'})/U(1)$&5\\ $\hbox{\xiiss C\kern-.4emIP}^3$ & $9 -\frac {36}k$ & $SU(4)_k/(SU(3)_k\times U(1))$ & 6\\ $S^7$ & $\frac{21}2 -\frac {36}k$ & $SU(4)_k/SU(3)_k$&7\\ $T^{pqr}$ & $\frac {21}2 - \frac 6 k -\frac 6{k'}-\frac 6{k''}$& $(SU(2)_k\times SU(2)_{k'}\times SU(2)_{k''})/U(1)^2$&7 \end{tabular} \end{table} Our objective in the next section will be to give a detailed construction of spacetime supersymmetry in these backgrounds, and to show why the standard choice of using the $N=2$ $U(1)_R$ worldsheet current is not appropriate. We begin by defining our conventions. In general, we will have an $SL(2,\BR)_k$ superconformal current algebra $(J^a,\psi^a)$, realized by free fields. There is a free $U(1)$ system, with currents $(K,\chi)$. Finally, there is a coset $X/U(1)$ with an $N=2$ $U(1)_R$ current which we will denote by $J_c$. We require the total central charge $c_T=15$ such that it may be coupled to the standard superconformal ghost system to define a consistent string theory. The total $U(1)_R$ current may then be written as \newcommand{J_R}{J_R} \begin{equation} J_R = (\psi^0 \chi)+\left( :\psi^+\psi^-: +\frac{2}{k} J^0 + J_c \right)= i\partial H_1 +i\partial H_2 \end{equation} We have bosonized the currents in parenthesis in favor of two bosons\footnote{Note the normalization $H_2(z)H_2(0)=-4\ln z$.} $H_1,H_2$. This is the canonical choice for generic $X$, where we require the bosons to be integral.\footnote{In special cases, $X$ will be such that $J_R$ splits integrally in terms of more bosons. In these cases, spacetime supersymmetry will be enhanced.} It is straightforward to verify that the following OPEs are regular \begin{eqnarray} J_R(z) J^0(z') \sim 0\\ J_R(z) K(z') \sim 0 \end{eqnarray} and moreover \begin{eqnarray} J_R(z) J_R(z')&\sim& \frac {5}{(z-z')^2}\label{eq:RRope}\\ J^0(z) J^0(z') &\sim & -{k/2\over (z-z')^2}\\ K(z) K(z') &\sim & \frac {k'}{(z-z')^2} \end{eqnarray} where $k$ is the level of the $SL(2)$ current algebra and $k'$ is related to the radius of compactification of the free boson leading to the $U(1)$ supercurrent. In deriving \eq{RRope}, we have used $c=15$ (and thus $c(X/U(1))=9-6/k$). The canonical choice for spacetime supercharges is \beql{standQ} Q^{\pm\pm} = \oint dz\ e^{\pm iH_1/2\pm iH_2/2} e^{-\phi/2} \end{equation} where $\phi$ is the superconformal ghost. These supercharges have the property that they are BRST invariant, while mutual locality requires that we keep only $Q^{+\pm}$. Modulo picture changing\cite{FMS}, we find the commutator \begin{equation} \{Q^{++}, Q^{+-}\} = J^0 -K \end{equation} Thus, \eq{standQ} leads naturally to spacetime supersymmetry, but unfortunately, the expected spacetime bosonic symmetries (such as $SL(2)$) are not recovered. (problems of this nature were discussed in \cite{GKS}; there the problems were much worse, as the algebra was doubled, etc.) We'll consider this in more detail below, and give the correct construction. \section{Target space consistency} If $X/U(1)$ is a direct product $\sim \oplus_i {\cal A}_i$, where each factor has central charge $c_i=b_i-a_i/k_i$, then the condition $c=15$ forces the constraint \begin{equation}\label{eq:sevencon} \frac 6 k = \sum_i \frac{a_i}{k_i} \end{equation} which relates the level of the $SL(2)$ algebra to the level of the other cosets in the construction, and it is easy to see that taking $k_i\to \infty$ for all $i$ forces $k\to\infty$. This means that we have a good geometric picture of the coset as a target space, since in the large $k_i$ limit the cosets behave semiclassically. In writing \eq{sevencon}, we have assumed that $X$ has classical dimension seven. Changing this would modify \eq{sevencon} by terms of order one, so the AdS spacetime would never become semiclassical. \newcommand{{\cal J}}{{\cal J}} Now let us analyse the spacetime isometries. We would like to see that the isometries have appropriate commutators with the spacetime supercharges. In particular, given the worldsheet $SL(2)$ currents, we can construct charges that should generate spacetime symmetries; consider the spacetime $SL(2)$ raising operator ${\cal J}^+ = \oint J^+ dz$. This operator has the following OPE with $J_R$, \begin{equation} {\cal J}^+ J_R(z) \sim -\psi^+ \chi +2\psi_0\psi^+ - \frac {2}{k} J^+ \end{equation} and the last term implies that the OPE ${\cal J}^+ Q^{++}$ is non-local.\footnote{To see this, it is useful to realize $Q\sim e^{\int J_R}$.} Also, the $Q$'s do not carry $U(1)_R$ charges, which is in conflict with the spacetime $N=2$ algebra if we want to identify $\oint K$ as the $U(1)_R$ current. That is, the existence of the operator $Q^{++}$ leads to the wrong target space symmetries. Instead of doing this, we will modify the supersymmetry generator in order that the spacetime Virasoro algebra is recovered. It is simple to see that the problem comes from the factor of $J^0/{2k}$ in the $U(1)_R$ current of $SL(2,\BR)/U(1)$. To solve the problem then, it is sufficient to redefine the worldsheet $U(1)_R$ current \begin{equation} \tildeJ_R = J_R +\frac 1{2k}J^0 = i\partial H_1+i\partial \tilde H_2 \end{equation} However, the OPE $\tildeJ_R(z)\tildeJ_R(z') \sim (5 - \frac 1{4k})/(z-z')^2$ implies that the spin operator \begin{equation} \Sigma^{++} = \exp(+\frac {iH_1}{2}+\frac{i\tilde H_2}{2}) \end{equation} has the wrong ($\neq 5/8$) conformal dimension. This means we need to correct again $\tildeJ_R$. That is, we define \begin{equation} J_R' = \tildeJ_R + \frac F{2k} \end{equation} with $F$ a current such that $F(z)F(z') = - J^0(z) J^0(z')$ and that is orthogonal to $J_0$, so that the bosonization of $J'$ leads to the correct conformal weight for the spin operator. We have a canonical choice for $F$ in our models, as we have the extra free $U(1)$, $K$. Hence, if $K$ has the same level as $SL(2)$ , then \begin{equation} J_R' = J_R + \frac{ J^0+K}{2k} \end{equation} will lead to the correct conformal weight for the spin operator. Notice that the current $J^0+K$ is null, and orthogonal to $J_R$, so that the natural spin operators constructed from $J_R'$ have automatically the correct conformal dimension. This also gives a charge to the spin operators under the spacetime $U(1)_R$ generated by $\oint K$. The full current is a sum of three integral pieces now \begin{equation} J_R' = (\psi_0 \chi) + (\psi^+\psi^-) + (\frac {K}{2k} + J_c) = i\partial(H'_1+H_2'+H'_3) \end{equation} Let us check that we get the $N=2$ superconformal algebra in spacetime. The BRST invariant spin operators will be given by \begin{equation} S = \exp(\epsilon_i\frac{iH'_i}2)\exp(-\phi/2) \end{equation} with $\epsilon_i= \pm 1$. Mutual locality and BRS invariance lead to the constraint $\prod \epsilon_i = 1$, and the spacetime supersymmetry operators are given by \begin{eqnarray} Q^{+\pm} &=& \oint \exp\left({i\frac {H'_1}{2} \pm i\frac {H'_2}{2}\pm i\frac {H'_3}{2}}\right)\\ Q^{-\pm} &=& \oint \exp\left({-i\frac {H'_1}{2} \pm i\frac {H'_2}{2}-\pm i\frac {H'_3}{2}}\right) \end{eqnarray} The spacetime supersymmetry algebra (modulo picture changing) is now seen to be given by \begin{eqnarray} \{Q^{++},Q^{-+}\} &=& J^+\\ \{Q^{+-},Q^{--}\} &=& J^-\\ \{Q^{-+},Q^{--}\} &=& J^0+K\\ \{Q^{+-},Q^{++}\} &=& J^0 - K \end{eqnarray} which is exactly the $NS$ sector of two dimensional $N=2$ supersymmetry. That is, this provides a general construction of $N=2$ spacetime supersymmetry whenever we have $N=2$ worldsheet supersymmetry. Motions in the $U(1)$ fiber are the generators of the $R$ symmetry spacetime current, and the modifications we made give charge to the supersymmetry generators, which is a requirement of the algebra. We want to complete the construction with the rest of the Virasoro algebra in spacetime. That is, we can write the rest of the modes of the spacetime supercurrent by applying the other Virasoro generators to the spin fields \cite{GKS}. It is a straightforward exercise to show that the algebra closes with a central charge given by $c_{st} = k p$, where $p$ is the winding number of the strings located at infinity. It is also clear that the fact that the level of the $U(1)$ is chosen to be equal to that of the $SL(2)$ current implies a quantization condition on the radius of the $U(1)$ bundle to be of the same order of magnitude as the radius of the AdS spacetime. \section{Comments on $AdS_3\times S_3\times T^4$ and $AdS_3\times S^3\times S^3\times U(1)$} Now that we have a general recipe, we can go back and analyze previous results \cite{GKS, EFGT} in a new light. We confine ourselves to a few comments. In particular, the current $K$ is chosen in a very special way. For example, in $AdS_3\times S^3\times T^4$, we could have chosen either the Cartan of $SU(2)$, or one of the free $U(1)$'s. Since we have only analyzed the effects of the spacetime raising operator $J^+$, we can always guarantee that we get the spacetime Virasoro algebra with either choice. On the other hand, if we want to keep the full $SU(2)$ isometries of the target space, we actually have to also eliminate a similar problem with the locality of the $SU(2)$ raising operator, and the modification in the currents is such that both effects get cancelled simultaneously if we choose to pair the Cartan element of $SU(2)$ with the one of $SL(2,\BR)$, so the choice of null field solves both problems at the same time. For $AdS_3\times S^3\times S^3\times U(1)$, there is one linear combination which we want to add to the Cartan of $SL(2,\BR)$ such that all of the problems with the isometries go away simultaneously. This choice is the diagonal $U(1)$ which cancels the terms from both $AdS_3$ terms, and a posteriori ends up having the correct quantization condition. This chooses a complex structure on the four free $U(1)$ directions. Finally, let us remark on $AdS_3\times S^3\times \hbox{\xiiss C\kern-.4emIP}^2$, where $\hbox{\xiiss C\kern-.4emIP}^2$ is understood as the coset $SU(3)/(SU(2)\times U(1))$. In this case, it can be seen that the Cartan of the $SU(2)$ coming from $S^3$ can not cancel the non-locality of the raising operator of $SU(2)$, so that we do not preserve the isometries, and the correct quantization condition squashes the sphere, so it leaves us with $N=2$ spacetime supersymmetry as opposed to $N=4$ which one may have naively guessed. \section{Vertex Operators and free fields.} We have shown that it is straightforward to compute the spacetime NS sector of supersymmetry. Naturally, as we have a field of conformal dimension zero, represented by $\gamma$, we can multiply $Q$ by any fractional powers of $\gamma$. BRST invariance will choose a linear combination of $Q\gamma^\alpha$ with the same spacetime quantum numbers. Closure of the fermionic algebra on the bosonic generators provides that $\alpha\in \BZ$ or $\alpha\in \BZ+1/2$, but not both simultaneously, as then the bosonic generators will not be single valued when $\gamma\to e^{2\pi i}\gamma$. That is, we have a choice of two super selection sectors for the supersymmetry algebra. One is to be taken as the Neveu-Schwarz sector as we already saw, and the other is the Ramond sector of the spacetime supersymmetry. This is a feature of the free field realization for the $SL(2)$ current algebra, and it actually lets us write vertex operators for the $SL(2)$ current with free fields. The same can be done with the coset constructions, as they admit free field realizations. Now, we want to analyze the chiral GSO projection. This needs to be done for each case, as we are not using the $N=2$ structure of the worldsheet CFT directly, but a modified version of it. Let us fix some notation. As is well known, coset conformal field theories admit free field realizations \cite{KOS,KOS2}. For each of the raising operators of the algebra we use a $B,C$ system, with OPE given by \begin{equation} B(z) C(z') = \frac 1{z-z'-\theta\theta'} = \frac{1}{Z-Z'} \end{equation} these can be super-bosonized into a set of two null (lightcone) scalars $T,U$ with OPE \begin{equation} T(z) U(z') = \log(Z-Z') \end{equation} by taking $B = a DU \exp (a^{-1}T)$, $C = \exp (-a^{-1}T)$. To the Cartan elements we associate free supersymmetric bosons $H$. The total collection of free bosons will be labeled by $\phi$, and the fermions by $\psi$. For the bosons we want to take operators (in the left moving sector lets say) which have an odd number of free fermion insertions in the $[-1]$ picture. \begin{equation} {\cal O}(\psi^\alpha, \partial\phi, \partial)e^{ip \cdot \phi} \end{equation} In this basis the embedding of the group theory (allowed lattice of values for $p$) is not manifest. This requirement corresponds to the chiral GSO projection. The mass shell condition \begin{equation} p\cdot (p-\lambda) +N = \frac 12 \end{equation} may be equated to $-\frac {j(j+1)}{2k} + m^2 +1/2$, where $j$ is the spacetime $SL(2)$ quantum number and $m^2$ is to be understood as the $AdS_3$ mass of the state. $\lambda$ is the curvature coupling of the free scalars. The GSO projection implies that we get a positive value for $j$; that is, the theory does not have tachyons. The number of free superfields is ten, and the total field with a worldsheet curvature coupling of the $AdS$ and $X$ theories combined together is a null field. In this sense, all theories have the same underlying structure. It is the choice of lattice (modular invariant) which makes them different from one another. The ten dimensional massless vertex operators are the most interesting as they predict the supergravity spectrum of the compactification. In this case the polynomial $S$ reduces to one free fermion operator. Amongst these, we will find all the chiral operators of the spacetime conformal field theory which can be described by string vertex operators. The fermion vertex operators are constructed by acting with the spacetime supersymmetry generators on the spacetime boson vertex operators that we described. As we have a choice of super selection sector, they will look different in each of the cases. It is also clear that this difference is in the powers of $\gamma$, and therefore the OPE of two fermion vertex operators closes on the ones of bosonic type. The advantage of having a free field realization is that we have a choice of ten free fermions $\psi^i$, which can be bosonized into five scalars, and writing spin vertex operators is straightforward. One has to remember that in order to get the right charges for the supersymmetries, they will be multiplied by powers of $e^{ip\cdot\phi}$ with $p^2 = 0$. From the bosonization, we find that each $B,C$ system contributes two lightcone scalars, i.e., their signature is $(1,1)$. Hence the lattice that we obtain has a signature of $(n,m)$ with $n+m=10$, and $n,m\geq 2$. The difficulty in writing the partition function lies in finding the constraints on the lattice and the screening operators of the system, so that at the end we can recover unitarity and modular invariance. \section{Discussion} We have given a complete construction of $N=2$ spacetime supersymmetry on $AdS_3$ spacetime for type II NS NS backgrounds, which are constrained to admit a version of $N=2$ supersymmetry on the worldsheet. It is clear that as we have exact conformal field theories on the worldsheet, these are solutions to all orders in $\alpha'$ of the string equations of motion. Moreover, usually when we have enough spacetime supersymmetry, this is what we need to guarantee that there are no non-perturbative corrections to the target space. In principle, if we ask questions that can be answered perturbatively, this approach should give a complete set of calculational tools. On the other hand, further work is required. In particular, the coset model predictions certainly go beyond supergravity and compute the full spectrum of the one-particle states of the string theory propagating on these spacetimes, and therefore, we can get a good idea of what the spacetime conformal field theory might be, even when we are not required to be in any large radius limit. This should shed light on the $\frac 1N$ corrections to the conformal field theories. The detailed form of the string partition function is not known, and it is of course important to construct it and check modular invariance. This is not clear, as we can not go to a lightcone gauge where everything is manifestly unitary. Also, as we have constructed these theories with free fields, it is likely that everything is complicated by the screening charges, so that a very good knowledge of how to extract the real physical degrees of freedom is required. Although some progress has been made in \cite{DQ}, it is far from complete. Spaces with non-compact cosets \cite{GQ,BN} are also interesting, but their spacetime CFT description is bound to be complicated by the non compactness, as the meaning of the conformal boundary comes into question, and if we get a spacetime CFT it seems naturally to give rise to a continous spectrum of states which is certainly more difficult to analyze. \medskip \noindent {\bf Acknowledgments:} We wish to thank F. Larsen for discussions. Work supported in part by the United States Department of Energy grant DE-FG02-91ER40677 and an Outstanding Junior Investigator Award. \providecommand{\href}[2]{#2}\begingroup\raggedright
\section{Coulomb systems} \setcounter{page}{1} \label{chapter- Coulomb systems} \subsection{Introduction} \label{sec- introduction} The emergence of thermodynamics and of various physical laws from the more fundamental levels of atomic theory and statistical mechanics were high points in our education in physics. But by the time we reached textbook treatments of the complex world of the Coulomb interaction, for example, formation of atoms and molecules and equilibria between them, we learned to demand less from theory and to be more content with reasoning by analogy consistent with thermodynamics. We know experimentally that atoms and molecules form, that they can behave like mixtures of ideal gases, so phenomenologically they have chemical potentials and we no longer insist that this emerge in some limit from $N$-body quantum Coulomb systems. We also know that Coulomb systems can have screening and dielectric phases. But when screening is expected to occur, it is quite a common practice to rather bluntly replace the Coulomb potential by a screened potential inherited from mean-field theories without inquiring too deeply into the legitimacy of this change. Likewise, there does not exist a first principle theory of the dielectric constant that does not presuppose the existence of atoms and molecules. At best, assuming that atoms form, one uses several laws, such as the Clausius-Mosotti formula, whose microscopic foundations have to be elucidated. In all these cases, this is a sensible attitude since most of non-relativistic physics is reputed to be hidden within the $N$-particle Coulomb Hamiltonian. There are however basic facts that can be cleanly formulated as limiting theories and shown to persist near the limit as well. This review is devoted to these types of results on Coulomb systems at low density. At low density the most famous properties of Coulomb systems are related to screening. The classical Debye-H\"uckel theory and its quantum analogue the random phase approximation have been supplemented in recent years by additional information established both by rigorous proofs and by closer inspection of resummed perturbation theory consistent with aspects of the rigorous proofs. There is also a precise understanding how a dilute assembly of quantum nuclei and electrons can form gases of atoms and molecules in the Saha regime. These results are physically well understood but the challenge to derive them has shown that standard perturbation theories miss some effects; indeed it turns out that quantum Coulomb systems do not screen, in the strong sense that the classical system does. This is a principal focus of this review. The first rigorous results were at the level of thermodynamics, the celebrated stability of matter theorem of Dyson and Lenard and the existence of thermodynamic behavior established by Lieb and Lebowitz. They hold for all states of nonrelativistic matter. The importance of the stability result is that it shows that quantum mechanics and electrostatics is a consistent simplified statistical mechanics of nature that needs no intervention from more fundamental levels of description. The next level, which is the focus of this review, is to ask for information on correlations: here the variety of physical phenomena in question is so vast that we cannot hope for too much generality: thorough consideration of the low density behavior is the first and the most natural step in this direction. There are three parts. The first is devoted to classical systems. Here a significant achievement is the proof that the Debye-H\"uckel theory is a limiting theory and Debye screening holds close to this limit, that is, at sufficiently low density (or high temperature). This is the subject of chapter~\ref{chapter- debye-screening}. Frequently proofs of such results are very long with tedious parts devoid of physical insight, but there are also components where a physical insight becomes a precise argument or where a useful new identity such as a resummation of perturbation theory appears. The philosophy here and throughout the review is to state the theorems and to give only those ideas in the proofs that conform to this specification. Many of these arguments are founded on the Sine-Gordon transformation, described in chapter~\ref{chapter- debye-screening}. We hope that our applications will raise appreciation for this transformation. We hope to display enough arguments without overwhelming technicalities to satisfy the physicist curious to see the way between physical concepts and their full mathematical realization. In the second part we consider quantum mechanics. In between there is a natural transition chapter (chapter~\ref{chapter- dipole}) which revolves around the absence of screening in classical dipole systems. We present the simple proof of this lack of screening. This chapter prepares for the study of the quantum case because it turns out that the Feynman-Kac formula reduces quantum equilibrium statistical mechanics to a classical formalism in which both monopoles and higher order multipoles appear. The proper quantum mechanical effects on screening are exhibited in chapter~\ref{chapter- Semi-classical} in the simplified framework of the semi-classical approximation. The well known Feynman-Kac representation of quantum statistical mechanics is recalled and quantum statistics are neglected. Perturbing around the classical gas makes immediately clear why intrinsic quantum fluctuations always destroy Debye screening. This appears to have been first suggested by Federbush, see \cite[page 428, section 5]{BrFe81}. One also discovers that the remaining screening mechanisms are more subtle than in the classical case, as revealed by the different long distance behaviors of various types of correlations. This chapter is intended to be a gentle baptism before the confrontation with the full quantum mechanical setting in chapter~\ref{chapter-loops}. Here the mathematics is much harder and less has been proved: in chapters ~\ref{chapter-loops} and ~\ref{chapter-6} the style is that of formal perturbation theory. In these chapters, following mainly the work of Alastuey, Cornu and collaborators, one exploits the full power of the Feynman-Kac representation including the quantum statistics. In this language, the quantum Coulomb gas is similar to a certain classical gas of fluctuating multipoles and all the efficient techniques of classical statistical mechanics are at hand. In particular, one can perform the usual partial resummations needed to cure the Coulomb divergences and end up with well defined Mayer-like diagrammatic rules. This offers a new perturbative scheme for the many-body problem which is particularly suited for low density expansions, in contrast to the standard Feynman perturbation theory with respect to the coupling constant. The formalism is applied in chapter~\ref{chapter-6} to the determination of the long distance behavior of the correlations and to the equation of state at low density. The (formally) exact results presented here provide the most detailed information available now on these questions. Some of them are new, in particular the explicit formulae for the tails of the correlations at low density. Also one recovers and completes the equation of state already obtained by Ebeling and coworkers by different methods. In chapter~\ref{chapter-7} we report on a nice development, initiated by Fefferman, on atomic and molecular phases at low density (the Saha regime). One considers a joint limit, the atomic or molecular limit, of vanishing density and vanishing temperature by fixing the chemical potentials (negatives) and letting the temperature tend to zero. Lowering the temperature enhances the probabilities for quantum mechanical binding whereas lowering the density favors ionization. It turns out that in this limit each value of the chemical potentials determines in principle certain chemical species and the system behaves as a mixture of these ideal substances. Ionization equilibrium phases are obtained in this way. In this asymptotic sense, it gives a precise meaning to the notion of atoms and molecules in the many-body problem. The third part consists of a single chapter~\ref{chapter-convergent-expansions} that is more technical: rigorous results at low density require control of the convergence of various types of cluster expansions. The proof of Debye screening has stimulated the development of resummation techniques which may be useful in other situations. For example there are new combinatoric formulae for Mayer coefficients. Without aiming at completeness, this last chapter provides the dedicated reader with some arguments left out in the main course of the text and gives an updated summary of some of the available tools in this field. The considerations in this review are essentially concerned with our three dimensional world. There has of course been dramatic progress in the one and two dimensional solvable models of Coulomb systems, which we do not include in the review, apart from isolated comments included to demonstrate some contrast with three dimensions. Let us also mention that basic new results on the stability of Coulombic matter in presence of magnetic fields and quantized radiation field have been obtained recently (references in \ref{sec- quantum}). The history of proofs of fundamental results, for example that statistical mechanics predicts phase transitions, suggests that difficult results can simplify over time and serve as consistency checks on what comes next. Some but not all of the derivations in this review are far from this stage: nevertheless, the existence of the subtle tunneling corrections in the Debye asymptotic regime revealed by the Sine-Gordon representation is not seen by conventional Mayer graph summations. The suggestive view of the quantum gas as an assembly of random multipoles displayed by the Feynman-Kac representation and the related lack of screening are much easier to miss in the conventional many-body perturbation theory. The physically very natural molecular limit considered in the main theorem of chapter~\ref{chapter-7} should be introduced and discussed in any text book on quantum statistical mechanics. We conclude this introduction with a list of some questions that seem to us of interest, to be pursued in the spirit of this review and perhaps by similar techniques.\\[2mm] \noindent CLASSICAL: Debye regime\\[2mm] \hspace*{2mm} \begin{minipage}{15.5cm} A more detailed understanding of boundary conditions and charge expulsion.\\ Results on correlations along walls, surface charges.\\ Understanding the improved Debye-H\"uckel theories of Fisher et al. as limiting theories.\\[2mm] \end{minipage} \newpage \noindent Dipole systems\\[2mm] \hspace*{2mm} \begin{minipage}{15.5cm} More definitive results on the dielectric constant and clarification of the literature in the light of such results.\\[.2mm] \end{minipage} \noindent QUANTUM: Low density\\[2mm] \hspace*{2mm} \begin{minipage}{15.5cm} Prove absence of exponential screening or at least finiteness of each individual Mayer-like diagram of chapter~\ref{chapter-loops}.\\ Correlations in non-uniform systems, e.g. along walls.\\ Coherent treatment of para and diamagnetism, boundary currents.\\[.2mm] \end{minipage} \noindent Slightly imperfect molecular gases\\[2mm] \hspace*{2mm} \begin{minipage}{15.5cm} Lowest order corrections to the molecular limit, atomic correlations and van der Waals forces.\\ Possible existence of dielectric phases with true atomic dipoles.\\[.2mm] \end{minipage} \noindent Compatibility of macroscopic electrostatics and statistical mechanics of quantum charges.\\ \subsection{The classical Coulomb gas} \label{sec- classical} We consider several species, $\alpha = 1,\dots, {\cal S}$, of charged particles with charges $e_{\alpha }$ of both signs, confined to a box $\Lambda$ in ${\Bbb R}^3$. If there are $N$ particles in total sitting in an external electrostatic potential $\phi$ then the potential energy is given by \begin{equation}\label{1.1.1} U ({\bf r} _{1}, e_{\alpha _{1}},\dots ,{\bf r} _{N},e_{\alpha _{N}})+ \dots + \sum_{i=1}^N e_{\alpha_i}\phi({\bf r} _{i}) \end{equation} where \begin{equation}\label{1.1.1a} U ({\bf r} _{1}, e_{\alpha _{1}},\dots ,{\bf r} _{N},e_{\alpha _{N}} )= \sum_{i<j=1}^N e_{\alpha_i} e_{\alpha_j} V({\bf r} _i - {\bf r} _j ) \end{equation} $V$ is the two body Coulomb potential energy \begin{equation}\label{1.1.2} V({\bf r}) = \frac{1}{ \mid{\bf r}\mid}. \end{equation} and the ellipsis indicates potentials due to other forces, that are discussed below. Suppose that $N_{\alpha }$ denotes the number of particles of species $\alpha $. Then the equilibrium statistical mechanics is determined by the partition function \begin{eqnarray}\label{1.1.3} && \Xi_\Lambda = \Xi_\Lambda(\beta \phi) = \sum_{\{N_\alpha\}}^\infty \prod_{\alpha = 1}^{{\cal S} } \frac{(z_\alpha)^{N_\alpha}}{N_\alpha !} \int_{\Lambda} d{\bf r}_{1} \ldots \int_{\Lambda} d{\bf r}_{N} \exp\(-\beta U - \beta \sum_{i=1}^N e_{\alpha_i}\phi({\bf r} _{i})\) \end{eqnarray} and $z_{\alpha }$ is called the activity of species $\alpha$. In particular the pressure $P$ is given by \begin{equation}\label{1.1.3c} \beta P = \lim_{|\Lambda |\rightarrow \infty } \frac{1}{|\Lambda |}\ln \Xi_\Lambda . \end{equation} The activities in the partition function are related to the chemical potential $\mu _{\alpha }$ of species $\alpha $ by the standard formula \begin{equation}\label{1.1.3cc} z_{\alpha } = \frac{e^{\beta \mu_{\alpha }}} {(2\pi \lambda_{\alpha }^{2})^{3/2}} \end{equation} where $\lambda_{\alpha }$ is the de Broglie length given below in (\ref{1.1.3ccc}). This formula results from the classical limit of quantum mechanics. In three dimensions a purely Coulomb Boltzmann factor $\exp \( - \beta U \)$ is not integrable because of the Coulomb singularities at ${\bf r}_{i} = {\bf r}_{j}$: additional forces are necessary for stability. We suppose there is a length scale $\lambda $ where the Coulomb potential gives way to these other forces. We used $\lambda $ because it suggests a de~Broglie length \begin{equation}\label{1.1.3ccc} \lambda_{\alpha } = \hbar(\beta /m_{\alpha })^{1/2} \end{equation} where $m_{\alpha }$ is the mass of species $\alpha $. For example, for individual point particles, classical mechanics gives way to quantum mechanics when their wave functions overlap, which occurs at separation $O (\lambda _{\alpha })$. For ions, $\lambda $ would instead be the radius of an outer orbital. For Fermions, one attempts to model the Pauli principle together with electrostatic repulsion by classical effective short range forces. We will either represent these short range forces by a hard core of radius $\lambda $ or we will smooth the Coulombic singularity by replacing $|{\bf r}|^{-1}$ by \begin{equation}\label{1.1.3bc} V_{\infty ,\lambda }({\bf r} ) = |{\bf r}|^{-1}(1-\exp(-|{\bf r}|/\lambda )) \end{equation} Neither choice is likely to be a realistic description of the physics at this length scale, so these are only claimed to be useful when one can show that there are predictions which are not very dependent on how the short range force is chosen. The notation $V_{\infty ,\lambda }$ conveys the equation \begin{equation}\label{1.1.3be} V_{\infty ,\lambda } = V_{\infty } - V_{\lambda } \end{equation} where we set \begin{equation}\label{1.1.3bd} V_{\lambda }({\bf r} ) = |{\bf r}|^{-1}\exp(-|{\bf r}|/\lambda ); \ \ \ V_{\infty }({\bf r} ) = 1/|{\bf r} | \end{equation} The potential $V_{\infty ,\lambda }$ is the Green's function for $(-\Delta + \lambda ^{2}\Delta ^{2})/(4\pi ) $ with the boundary condition appropriate for a system confined to a box $\Lambda $ with insulating walls. We shall also consider systems confined in a box with conducting grounded walls. In these cases $V_{\lambda }({\bf r} - {\bf r} ')$ is replaced by $V_{\lambda }({\bf r} ,{\bf r} ')$ which then denotes the Greens' function with zero boundary conditions. Consider the case where the ellipsis in (\ref{1.1.1}) represent a hard core of radius $\lambda $. By Newton's theorem each point charge can then be spread out into an equivalent charge on a sphere of radius $\lambda $ without changing the Coulomb interaction energy. If self-energies of the spheres are added into the Coulomb energy then it is positive because it can be written as the integral of the square of the electric field. Since the self-energy of $N$ spheres is $\frac{1}{2}\sum _{\alpha } e_{\alpha}^{2}/\lambda$ this argument shows that \begin{equation}\label{1.1.3b} \mbox{Stability:} \ \ \ \ \ U \geq - B N; \ \ \ B = O(\frac{e_{\alpha }^{2}}{2 \lambda }) \end{equation} Thus the introduction of the hard cores has made $\Xi_\Lambda$ finite \cite{Ons39}. The same stability bound holds in the other case where the additional terms in the ellipsis in (\ref{1.1.1}) smooth out the Coulombic singularity. The total potential energy with self-energies added is \begin{equation} \label{1.1.3bb} \frac{1}{2}\sum _{i,j} e_{\alpha _{i}}e_{\alpha _{j}} \frac{1 - e^{-\mid{\bf r} _{i}-{\bf r} _{j}\mid /\lambda }}{\mid{\bf r} _{i}-{\bf r} _{j}\mid} \end{equation} which is positive because it equals, in Fourier space, \begin{equation}\label{1.1.3bbb} \frac{1}{2} \frac{4\pi }{(2\pi )^{3}} \int d {\bf k} \, \left( \frac{1}{|{\bf k}|^{2} } - \frac{1}{\lambda ^{-2} + |{\bf k}|^{2} } \right) |\sum _{i} e_{\alpha _{i}}e^{i{\bf k}\cdot {\bf r} _{i}}|^{2} \geq 0 \end{equation} The self energies are again $\frac{1}{2}\sum _{\alpha }e_{\alpha}^{2}/\lambda$ so we obtain (\ref{1.1.3b}). Analogous arguments using the appropriate eigenfunctions in place of $\exp (i{\bf k} \cdot {\bf r} )$ show that stability bounds hold also for particles in an grounded container. We use the summation identity \begin{equation}\label{1.1.4} \sum_{\{N_\alpha\}}^\infty\prod_{\alpha}\frac{1}{N_{\alpha}\,!}\cdots = \sum_{N=0}^\infty\frac{1}{N!} \sum_{\alpha_1,\ldots,\alpha_N}^{{\cal S}}\cdots \end{equation} to write the partition function (\ref{1.1.3}) in the form \begin{equation} \label{1.1.5} \Xi_\Lambda = \sum_{N=0}^\infty\frac{1}{N!} \int \prod_{k=1}^N d{\cal E}_k z({\cal E}_k) \exp\(-\beta U({\cal E}_1, \ldots, {\cal E}_k) -\beta \sum_{i=1}^N \phi({\cal E} _{i})\) \end{equation} where ${\cal E} = ({\bf r}, \alpha )$ unites the spatial and species coordinate so that $\phi({\cal E} ) = e_{\alpha }\phi({\bf r} )$ and the $d{\cal E}$ integration means \begin{equation}\label{1.1.6} \int d{\cal E}\cdots= \sum_{\alpha =1}^{{\cal S}}\int_{\Lambda}d{\bf r} \cdots \end{equation} and $z({\cal E}) = z_{\alpha }$. We denote by $\omega=({\cal E} _{1},\ldots,{\cal E} _N)$ a configuration of the particles, $U(\omega ) = U({\cal E} _{1},\ldots,{\cal E} _N)$ and $z(\omega ) = z({\cal E} _{1})\cdots z({\cal E} _{N})$. The grand canonical summation over these configurations is \begin{equation}\label{1.1.7} \int _{\Lambda } d\omega \cdots = \sum_{N=0}^\infty\frac{1}{N!} \int \prod_{k=1}^{N} d{\cal E}_k \cdots . \end{equation} and the grand canonical average is, for zero external potential, \begin{equation}\label{1.1.7a} \langle\cdots \rangle = \frac{1}{\Xi_\Lambda} \int _{\Lambda } d\omega z(\omega ) e^{-\beta U(\omega )} \cdots \end{equation} A system in which the species occur in pairs of equal but opposite charge with equal activities is said to be charge-symmetric. The observable that measures density of particles of species $\alpha $ at ${\bf r} $ is given by \begin{equation}\label{1.1.8a} \hat{\rho }({\cal E} ,\omega) = \sum_{j=1}^N \delta({\cal E} ,{\cal E} _j) = \sum_{j=1}^N \delta_{\alpha ,\alpha _{j}}\delta({\bf r} -{\bf r}_j). \end{equation} and $n$ - particle distributions for points ${\bf r} _{a}$, $a = 1,\dots , n$ are defined by \begin{eqnarray}\label{1.1.9a} \rho ({\cal E} _{1},\dots ,{\cal E} _{n}) = \langle \prod _{a=1}^{n} \hat{\rho }({\cal E} _{a})\rangle \end{eqnarray} These definitions coincide with the standard definitions of distribution functions in statistical mechanics when ${\cal E}_{1},\dots ,{\cal E} _{n}$ are distinct. If they are not all distinct (coincident points) then our definition includes squares or higher powers of delta functions. At various points in the text we will use the phrase ``does not include contributions from coincident points'' to indicate that such powers of delta functions are omitted. To illustrate what ``including contributions from coincident points'' means, consider $\int d{\cal E} \, f ({\cal E}) \rho ({\cal E} ,{\cal E} _{1})$. The integral smooths the delta function in (\ref{1.1.8a}) so that the diagonal terms $\sum _{j}\delta ({\cal E} ,{\cal E} _{j})\delta ({\cal E} _{1},{\cal E} _{j})$ in $\hat{\rho }({\cal E} ,\omega) \hat{\rho }({\cal E}_{1} ,\omega) $ give the ``coincident contribution'' $\sum _{j} f ({\cal E} _{j}) \delta ({\cal E}_{1} ,{\cal E} _{j})$. We also introduce the charge density observable \begin{equation}\label{1.1.8} \hat{c}({\bf r},\omega) = \sum_{\alpha } e_{\alpha }\hat{\rho }_{\alpha } ({\bf r} ,\alpha,\omega ) = \sum_{j=1}^N e_{\alpha_j} \delta({\bf r} -{\bf r}_j). \end{equation} and the charge distributions \begin{eqnarray}\label{1.1.9} c({\bf r} _{1},\dots ,{\bf r} _{n}) = \langle \prod _{a} \hat{c}({\bf r_{a}})\rangle \end{eqnarray} More generally one obtains mixed charge - particle distributions by taking the average of products of density observables and charge observables (\ref{1.1.8a}) and (\ref{1.1.8}). We shall also consider the corresponding truncated distributions (correlations) $\rho _{T}$ and $c_{T}$ defined in the usual way \begin{equation}\label{1.1.9b} \rho _{T}({\cal E} _{1},{\cal E} _{2}) = \rho ({\cal E} _{1},{\cal E} _{2}) - \rho ({\cal E} _{1}) \rho ({\cal E} _{2}) \end{equation} and so on. Correlation functions of charge densities can also be obtained as the response to an infinitesimal external potential by variational differentiation of the logarithm of the partition function $\Xi_{\Lambda }(\beta \phi)$ with respect to $-\beta \phi$. In particular, the charge density induced in the system by a potential perturbation at ${\bf r}_{1}$ is \begin{equation} c({\bf r}_{1})=\frac{\delta }{\delta (-\beta \phi({\bf r} _{1}))}\ln \Xi_{\Lambda }(\beta \phi) \label{1.1.9o} \end{equation} and the truncated distribution of two charge densities is \begin{equation}\label{1.1.10b} c_{T} ({\bf r} _{1},{\bf r} _{2}) = \frac{\delta } {\delta (-\beta \phi({\bf r} _{1}))} \frac{\delta } {\delta (-\beta \phi({\bf r} _{2}))} \ln \Xi_{\Lambda }(\beta \phi) \end{equation} By this definition, we find that \begin{equation} c_{T} ({\bf r} _{1},{\bf r} _{2}) = \langle \hat{c}({\bf r} _{1}) \hat{c}({\bf r} _{2}) \rangle - \langle \hat{c}({\bf r} _{1}) \rangle \langle \hat{c}({\bf r} _{2}) \rangle . \label{1.1.10c} \end{equation} The physical properties of an homogeneous infinitely extended charged fluid are conveniently described by its linear response to an external classical charge density $c^{{\rm ext}}({\bf r})$. One defines the static susceptibility $\chi({\bf r}_{2}-{\bf r}_{1})$ as the linear response at ${\bf r}_{2}$ of the charge density of the fluid to the external density at ${\bf r}_{1}$ \begin{equation} \chi({\bf r}_{2}-{\bf r}_{1})=\left[\frac{\delta } {\delta c^{{\rm ext}}({\bf r}_{1})}c({\bf r}_{2})\right]_{ c^{{\rm ext}}=0} \label{1.1.10d} \end{equation} The function $\chi({\bf r})$ describes the shape of the screening cloud around an infinitesimal point test charge at the origin. In Fourier space, \begin{equation}\label{1.1.fourier} \tilde{\chi}({\bf k}) = \int \chi({\bf r} )e^{-ik\cdot {\bf r} }d{\bf r} \end{equation} is related to the Fourier transform $S({\bf k})$ of the charge-charge correlations $c_{T}({\bf r},{\bf 0})$ (\ref{1.1.10c}) at zero potential (the structure factor of the fluid) by \begin{equation} \tilde{\chi}({\bf k})=-\frac{4\pi\beta}{|{\bf k}|^{2}}S({\bf k}) \label{1.1.10e} \end{equation} Finally, the total effective potential due to the test charge plus its screening cloud is \begin{equation} \tilde{V}^{{\rm eff}}({\bf k})=(1+\tilde{\chi}({\bf k}))\tilde{V}({\bf k})=\varepsilon^{-1}({\bf k})\tilde{V}({\bf k}) \label{1.1.10f} \end{equation} where we have defined the static dielectric function $\varepsilon({\bf k})$ by $\varepsilon^{-1}({\bf k}) =\tilde{\chi}({\bf k})+1$. Equivalently we immerse two test charges $({\bf r} _{a},e_{a})_{a=1,2}$ in the ensemble. Their (nonlinear) effective potential energy is the excess free energy due to the external charge distribution given by \begin{eqnarray}\label{1.1.eff1} && \exp \left(- \beta V^{{\bf eff}}(({\bf r} _{a},e_{a})_{a=1,2} \right) \nonumber \\ && \hspace{.2in}= \frac{\langle \exp \left(-\beta e_{1}V\ast \hat{c }({\bf r} _{1})- \beta e_{2}V\ast \hat{c }({\bf r} _{2}) - \beta e_{1}e_{2}V({\bf r} _{1} - {\bf r} _{2})\right) \rangle} {\langle\exp(-\beta e_{1}V\ast \hat{c }({\bf r} _{1}))\rangle \langle\exp(-\beta e_{2}V\ast \hat{c }({\bf r} _{2}))\rangle} \end{eqnarray} But then we take the test charges infinitesimal and find the leading term \begin{eqnarray}\label{1.1.eff} && V^{{\bf eff}}(({\bf r} _{a},e_{a})_{a=1,2} ) \sim e_{1}e_{2} V^{{\bf eff}}({\bf r} _{1} - {\bf r} _{2} ) \nonumber \\ && V^{{\bf eff}}({\bf r} _{1} - {\bf r} _{2} ) = V({\bf r} _{1}-{\bf r} _{2}) - \beta V\ast c_{T} \ast V({\bf r} _{1} - {\bf r} _{2}) \end{eqnarray} which is equivalent to our previous formula (\ref{1.1.10f}). A well known criterion for the plasma phase of the Coulomb system is the vanishing of $\varepsilon^{-1}({\bf k})$ as ${\bf k}\to 0$. This implies with (\ref{1.1.10e}) that $S({\bf k})\simeq (4\pi\beta)^{-1}|{\bf k}|^{2},\; {\bf k}\to 0$, which is equivalent with the second moment rule (Stillinger-Lovett rule) for the charge-charge correlation \begin{equation} \int d{\bf r} |{\bf r}|^{2} c_{T}({\bf r},{\bf 0})=-\frac{3}{2\pi\beta} \label{1.1.1sti} \end{equation} If the Coulomb interaction $U$ is set to zero the partition function is \begin{eqnarray}\label{1.1.11} \Xi_{{\rm ideal},\Lambda }(\beta \phi) &=& \sum_{N=0}^\infty\frac{1}{N!} \int \prod_{k=1}^{N} d{\cal E}_k z({\cal E}_k) \exp\(-\beta \sum _{j=1}^{N}e_{\alpha _{j}}\phi({\bf r} _{j})\) \nonumber \\ &=& \exp\( \int d{\cal E}\, z({\cal E}) e^{-\beta e_{\alpha }\phi({\bf r} )} \). \end{eqnarray} We will call the resulting system an ideal gas, because the external potential is just a multiplication of the activities by $\exp (-\beta e_{\alpha_{1} }\phi({\bf r}_{1}))$. In this case all distributions are computable, e.g., \begin{eqnarray}\label{1.1.12} \langle \hat{c}({\bf r_{1}}) \rangle_{{\rm ideal},\beta \phi} &=& \frac{\delta } {\delta (-\beta \phi({\bf r} _{1}))} \ln \Xi_{{\rm ideal},\Lambda }(\beta \phi) = \sum _{e_{\alpha _{1}}} e_{\alpha_{1} } z_{\alpha _{1}} e^{-\beta e_{\alpha_{1} }\phi({\bf r}_{1})} \end{eqnarray} where we used (\ref{1.1.9o}). \subsection{The quantum Coulomb gas} \label{sec- quantum} We consider ${\cal S} $ species of quantum point charges (electrons and nuclei) in a box $\Lambda $ in ${\Bbb R}^3$, with masses $m_{\alpha }$, charges $e_{a}$ and spins $s_{\alpha }$, $\alpha = 1,\dots, {\cal S}$. Each species obeys Fermi or Bose statistics and at least one species is Fermionic. The system is governed by the non-relativistic N-particle Hamiltonian \begin{equation}\label{1.2.1} H_{\Lambda, N} = - \sum _{i=1}^{N}\frac{\hbar^{2}}{2m_{\alpha _{i}}} \Delta _{i} + \sum _{i<j}e_{\alpha _{i}}e_{\alpha _{j}} V({\bf r} _{i} - {\bf r} _{j}) + \sum _{i=1}^{N} e_{\alpha _{i}}\phi({\bf r}_{i}) \end{equation} acting on the $N$-particle Hilbert space appropriately symmetrized according to the statistics of each species. The Laplacians $\Delta _{i}$ in (\ref{1.2.1}) have Dirichlet boundary conditions on the boundary of $\Lambda $. The external electrostatic potential is set to zero unless we warn the reader otherwise. We denote by $H_{N}$ the Hamiltonian for particles in infinite space. The stability of the Hamiltonian is the statement that if at least one of the particle species obeys Fermi statistics, then \begin{equation}\label{1.2.2} H_{N} \geq -B N \end{equation} with $B$ a positive constant independent of $N$ \cite{DyLe67,DyLe68}. The proof has been greatly simplified by Lieb and Thirring with the use of the Thomas-Fermi theory. This is reviewed in \cite{Lie76}. A new proof \cite{Gra97} is based on a remarkable electrostatic inequality for the classical Coulomb interaction (\ref{1.1.1}). For a review on recent developments on these questions, see \cite{Lie90}. We mention here that the problem of the stability of matter in presence of magnetic fields \cite{LiLoSo95,Fef95} and the quantized radiation field\footnote{The matter is non-relativistic and an ultraviolet cutoff is imposed.} \cite{BuFrGr96,FeFrGr98} has received much attention lately. Let \begin{equation} P_\alpha = \frac{1}{N_\alpha !} \sum_{p_\alpha} (\eta_\alpha)^{p_\alpha}{p_\alpha}, \;\;\;\; \eta_\alpha=\pm 1 \label{1.2.3} \end{equation} be the projection onto the symmetric ($\eta_\alpha=+1$) or antisymmetric ($\eta_\alpha=-1$) states of particles of species $\alpha$.The sum runs on all permutations $p_\alpha$ of the $N_\alpha$ particles and $(\eta_\alpha)^{p_\alpha}$ is the signature of $p_\alpha$; $p_\alpha$ acts both on the position ${\bf r}$ and the spin variable $\sigma$ of the particle. The projection onto the subspace of states of a many-body system with $N=\sum_{\alpha } N_\alpha$ particles having the appropriate statistics is given by \begin{equation} P = \prod_\alpha P_\alpha = \sum_p \( \prod_\alpha \frac{(\eta_\alpha)^{p_\alpha}}{N_\alpha !} \) p \label{1.2.4} \end{equation} where the sum runs now on all permutations of the $N$ particles that are compositions $p=p_{\alpha_{1}}\ldots p_{\alpha_{\cal S}}$ of permutations of particles of each species. Associating to each species a chemical potential $\mu _{a}$, the grand canonical partition function of the quantum gas is given by the sum \begin{eqnarray} \Xi_{\Lambda} &=& \sum_{\{N_\alpha\}}\mbox{Tr} P\exp\( -\beta\( H_{\Lambda,N} - \sum_{\alpha} \mu_\alpha N_\alpha \) \) \nonumber\\ &=& \sum_{\{N_\alpha\}} \sum_p \prod_\alpha \frac{(\eta_\alpha)^{p_\alpha} \exp(\beta \mu_\alpha N_\alpha)} {N_\alpha !} \int_\Lambda d{\bf r}_1\ldots{\bf r}_N \nonumber\\ &\times & \sum_{\{\sigma_{\alpha_{i}}\}} \langle\{{\bf r}_{p(i)},\sigma_{\alpha_{p(i)}}\} | \exp\(-\beta H_{\Lambda,N}\) | \{{\bf r}_i,\sigma_{\alpha_{i}}\}\rangle \label{1.2.5} \end{eqnarray} In (\ref{1.2.5}), $|\{{\bf r}_i,\sigma_i\}\rangle =\prod |{\bf r}_i,\sigma_i\rangle$ is a product of states of individual particles that also diagonalize the spin component along a fixed direction with quantum numbers $\sigma_\alpha$ taking the values $-s_\alpha,\;-s_\alpha+1\;,\ldots,\;s_\alpha$. The pressure is defined as in (\ref{1.1.3c}). Particle and charge distributions are defined as in the classical case, replacing the average (\ref{1.1.7a}) by \begin{equation}\label{1.2.6} \langle \cdots \rangle = \frac{1}{\Xi_\Lambda} {\rm Tr}P (e^{-\beta H_{\Lambda N}}\cdots ) \end{equation} Here the particle and charge distributions and correlations are defined as in (\ref{1.1.9a}) and (\ref{1.1.9}) where now the particle coordinates ${\bf r} _{j}$ in (\ref{1.1.8a}) are the quantum mechanical position operators\footnote{We use the same symbol ${\bf r}$ for the quantum mechanical position operator and for the argument of a correlation function. We do not consider spin correlations. All the quantum correlations in the review are purely positional with spin variables averaged out.}. In contrast to the classical situation, the response to an external electrostatic potential yields a new type of correlation functions, the imaginary time Green's functions. They are obtained by variational differentiation with respect to the external potential. Of special interest is the charge distribution at ${\bf r}_{2}$ induced by an external potential perturbation at ${\bf r}_{1}$ in the linear response regime \begin{eqnarray} \label{1.2.7b} c^{{\rm ind}}({\bf r}_{1}|{\bf r}_{2}) &=&\left[\frac{\delta } {\delta (- \beta\phi({\bf r} _{1}))}\langle \hat{c}({\bf r}_{2}) \rangle\right]_{\phi=0}\nonumber\\ &=& \left[\frac{\delta } {\delta (- \beta\phi({\bf r} _{1}))} \(\Xi_{\Lambda}^{-1}{\rm Tr} P\exp\( -\beta H_{\Lambda,N} \) \hat{c}({\bf r}_{2}) \)\right]_{\phi=0}\nonumber\\ &=& \frac{1}{\beta}\int _{0}^{\beta } d\tau \, c_{T}({\bf r}_{1},\tau,{\bf r}_{2}) \end{eqnarray} In (\ref{1.2.7b}) \begin{equation} c_{T}({\bf r}_{1},\tau,{\bf r}_{2})=\left\langle e^{\tau H_{\Lambda,N}}\hat{c}({\bf r}_{1})e^{-\tau H_{\Lambda,N}}\hat{c}({\bf r}_{2})\right\rangle-\langle \hat{c}({\bf r}_{1})\rangle \langle \hat{c}({\bf r}_{2})\rangle \label{1.2.7c} \end{equation} is the imaginary time displaced charge-charge correlation (or Duhamel function): it reduces to the static charge-charge truncated distribution at $\tau=0$. For an infinitely extended quantum charged fluid, we introduce the spatial Fourier transform $S({\bf k},\tau)$ of $c_{T}({\bf r},\tau,{\bf 0})$. The static structure factor of the fluid is then \begin{equation} S({\bf k})=S({\bf k},\tau=0) \label{1.2.7cc} \end{equation} and the generalization of the classical relations (\ref{1.1.10e}) and (\ref{1.1.10f}) to the quantum mechanical situation is \begin{equation} \tilde{\chi} ({\bf k}) =\varepsilon^{-1}({\bf k})-1=-\frac{4\pi}{|{\bf k}|^{2}}\int_{0}^{\beta}d\tau S({\bf k},\tau) \label{1.2.7d} \end{equation} Notice that in the latter case, the susceptibility $\tilde{\chi} ({\bf k})$ (defined as in (\ref{1.1.10d})) is no longer proportional to the static structure factor. Equivalence of ensembles and existence of thermodynamic functions was established in \cite{Gri69,LeLi69,LiLe72,LiNa75}. The analogous results for classical systems are corollaries. There are no results on existence of the infinite volume limit of distribution functions except for charge symmetric Bose systems with positive-definite interactions.\footnote{(\ref{1.1.3bc}) is a positive-definite interaction but hard cores plus Coulomb are not.} In this case at any density and temperature for which the system is stable\footnote{For sufficiently large activity the grand canonical ensemble diverges. For the ideal gas this happens at the Bose-Einstein condensation.} all particle distribution functions have unique infinite volume limits \cite{FrPa78,FrPa80}\footnote{Their proof only works for some choices of boundary conditions but this is probably not a physical limitation.}. These results include charge-symmetric classical systems with positive-definite potentials. They rely on the Sine-Gordon transformation described later. The physical content of these results on the existence of the infinite volume limit is that the limit exists independently of the way the container is enlarged. For example, we could consider two sequences of increasingly large containers related by a fixed translation. Both sequences would have the same infinite volume limit, so the limit must be translation invariant. In particular the limiting distributions could not describe a crystal with a low density of defects. On the other hand one would expect to be able to select such a state by choosing a sequence of containers whose boundaries match the natural planes of the crystal, unless the system never forms crystals. Therefore these results are evidence that stable charge symmetric systems with positive-definite potentials are always in a fluid phase. Further results on existence of the infinite volume limit of classical distribution functions at low density are included in the work on Debye screening \cite{BrFe78,Imb83} reviewed in the next chapter, but these are low density results. \subsection{Neutrality} \label{sec- neutrality} \def\underline{\bf e}{\underline{\bf e}} \def\underline{\mbox{\boldmath $\mu$\unboldmath}}{\underline{\mbox{\boldmath $\mu$\unboldmath}}} \def\mbox{\boldmath $\mu$\unboldmath}{\mbox{\boldmath $\mu$\unboldmath}} In systems with short range interactions the densities of different species can be varied independently, but Coulomb systems maintain charge neutrality \begin{equation}\label{2.1.1bb} \sum _{\alpha } \rho _{\alpha }e_{\alpha } = 0 \end{equation} by expelling excess charge to the boundary \cite{LiLe72,GrSc95b}. We introduce a vector notation \begin{eqnarray} \underline{\bf e} &=& (e_{1},\ldots,e_{{\cal S}}), \ \ \ \underline{\mbox{\boldmath $\mu$\unboldmath}} = (\mu_{1},\ldots,\mu_{{\cal S}}) \label{2.1.1cc} \end{eqnarray} for the set of charges and chemical potentials, and decompose $\underline{\mbox{\boldmath $\mu$\unboldmath}}$ \begin{equation} \underline{\mbox{\boldmath $\mu$\unboldmath}}=\mbox{\boldmath $\mu$\unboldmath}+\nu\underline{\bf e},\;\;\;\;\;\;\mbox{\boldmath $\mu$\unboldmath}\cdot\underline{\bf e} =0, \;\;\;\;\;\;\;\nu=\frac{\underline{\mbox{\boldmath $\mu$\unboldmath}}\cdot\underline{\bf e}}{|\underline{\bf e}|^{2}} \label{2.1.dd} \end{equation} into its components perpendicular and parallel to the charge vector. Then, the charge neutrality is equivalent with the fact that the infinite volume limit of grand canonical pressure does not depend on the component $\nu$ of $\underline{\mbox{\boldmath $\mu$\unboldmath}}$ along the charge vector \begin{equation} P(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}})=P(\beta,\mbox{\boldmath $\mu$\unboldmath}+\nu\underline{\bf e}) =P(\beta, \mbox{\boldmath $\mu$\unboldmath}) \label{2.1.1ee} \end{equation} This means that different choices of chemical potentials $\mu_{\alpha}$ or activities $z_{\alpha }$ do not necessarily lead to different densities. It is common to break this redundancy by imposing \begin{equation}\label{2.1.1b} \mbox{pseudo-neutrality:} \ \ \ \sum _{\alpha } e_{\alpha }z_{\alpha } = 0. \end{equation} The system will be neutral regardless of whether pseudo-neutrality is imposed or not, but in section~\ref{sec-charge-expulsion} we will argue that charges that are expelled to the boundary create a constant external electrostatic potential $\psi$ which renormalizes the activities $z_{\alpha } \rightarrow z_{\alpha } \exp [-\beta e_{\alpha} f]$ in such a way as to restore the condition for the renormalized activities. Thus imposing the condition amounts to working with the renormalized activities. \newpage \section{Debye screening}\label{chapter- debye-screening} \setcounter{page}{1} This chapter is devoted to screening in classical Coulomb systems. We begin with a review of the original theory of Debye and H\"uckel followed by an account of some rigorous theorems. Then there follow several sections that illustrate the main ideas used in the proofs with discussions concerning open problems. \subsection{Debye - H\"uckel Theory} \label{sec-debye-huckel} The mean field theory approximation for the dilute plasma phase \cite{DeHu23} is at the center of our discussion. We will briefly review their argument to prepare the way for further developments. For further background see \cite{McQ76,HaMc76}. The activity $z_{\alpha }$ has dimension ${\rm length}^{-3}$ and $\beta e_{\alpha }^{2}$ and $\lambda $ have dimensions ${\rm length}.$ Thus if $l_{D}$ is defined by \begin{equation}\label{2.1.1} l_{D} = (4\pi \sum_{\alpha } e_{\alpha }^{2} z_{\alpha } \beta )^{-\frac{1}{2}} \end{equation} then $l_{D}$ is a length, called the Debye length. The activities in this definition of $l_{D}$ are required to satisfy the pseudo-neutrality condition (\ref{2.1.1b}). It is standard to define the Debye length in terms of densities by the formula \begin{equation}\label{2.1.1a} {\kappa ^{-1}} = (4\pi \sum_{\alpha } e_{\alpha }^{2} \rho _{\alpha } \beta )^{-\frac{1}{2}} \end{equation} and we have used the notation $l_{D}$ instead of the standard $\kappa^{-1}$ because the two definitions are not equal, although their ratio tends to one in the Debye-H\"uckel limit described below. The conclusions of Debye-H\"uckel theory include that the pressure $P$ and charge-charge correlation obey \begin{eqnarray}\label{2.1.2} && \beta (P - P_{{\rm ideal}}) \sim \frac{1}{12 \pi l_{D} ^{3}}\nonumber \\ && c_{T}({\bf r} _{1},{\bf r} _{2}) \sim -\left( (4\pi )^{2} \beta l_{D}^{4} |{\bf r} _{1} - {\bf r} _{2}| \right)^{-1} \exp( - \frac{|{\bf r} _{1} - {\bf r} _{2}|}{l_{D}}) + (4\pi \beta l_{D}^{2})^{-1}\delta ({\bf r} _{1} - {\bf r} _{2}) \end{eqnarray} The Debye-H\"uckel argument suggests that these are really limiting laws, holding in a limit in which the number of particles in a volume $l_{D}^{3}$ tends to infinity, while the number in a volume $\lambda ^{3}$ tends to zero. Perturbation expansions \cite{May50,Mee61,LeSt68,HaMc76,MaMa77} are also likely to be asymptotic as opposed to convergent, but it is still natural to believe that the second result holds in a stronger sense: that without taking the limit there can be exponential decay of correlations, particularly if one is reasonably close to the limit. We shall refer to this stronger statement as Debye screening. The first law in (\ref{2.1.2}) can be expressed in terms of densities by using $\rho _{\alpha } = z_{\alpha } \partial \beta P/\partial z_{\alpha} $ and $\beta P_{{\rm ideal}} = \sum z_{\alpha }$. One finds that \begin{equation}\label{2.1.2b} \beta (P - \sum _{\alpha }\rho _{\alpha }) \sim -\frac{\kappa ^{3}}{24 \pi}\nonumber \\ \end{equation} The second result (\ref{2.1.2}) was based on an argument that can be paraphrased as follows. Consider a grand canonical ensemble of particles interacting with each other and also with a charge $e_{0}$ held fixed at the origin. This is the same as adding to $\beta U$ the external potential term $\beta\int e_{0}\hat{c}({\bf r} ,\omega ) V({\bf r} ) d{\bf r}$. Then the potential at another point ${\bf r} $ will be the sum of $V({\bf r}) $ and another potential $\phi$ created by the charges in the gas. Debye and H\"uckel approximate this $\phi$ by assuming that the gas reacts to the combined potential $f = e_{0}V + \phi$ as an ideal gas, which leads to the self-consistent scheme \begin{equation}\label{2.1.4} \frac{1}{4\pi }\Delta f({\bf r},\omega ) = - \hat{c}({\bf r} ,\omega ) - e_{0}\delta ({\bf r} ) \approx - \langle\hat{c}({\bf r})\rangle_{{\rm ideal},\beta f} - e_{0}\delta ({\bf r} ) \end{equation} We wrote $f({\bf r} ,\omega )$ to emphasize that the exact field is configuration dependent but having made this approximation we write $f ({\bf r} )$ from now on. From (\ref{1.1.12}), \begin{equation}\label{2.1.7} \frac{1}{4\pi }\Delta f({\bf r} ) = - \sum _{\alpha }z_{\alpha } e_{\alpha } \exp \left [ - \beta e_{\alpha }f({\bf r}) \right] - e_{0}\delta ({\bf r} ) \end{equation} If $\beta f({\bf r})$ is a weak potential then $\exp(x) \approx 1 + x$ and pseudo-neutrality (\ref{2.1.1b}) leads to \begin{equation}\label{2.1.10} [\Delta - l_{D} ^{-2}] f({\bf r} ) = - 4\pi e_{0} \delta ({\bf r} ) \end{equation} which has the solution \begin{equation}\label{2.1.9} f({\bf r} ) = e_{0} |{\bf r} | ^{-1} \exp \(-|{\bf r}|/l_{D} \) \end{equation} The exponential decay of $f$ is the origin of exponential decay of charge density correlations within this mean field approximation. Indeed noting that $f$ is related to the effective potential introduced in section~\ref{sec- classical} by $f = e_{0}V^{{\rm eff}}$ we can obtain the second Debye-H\"uckel law from (\ref{1.1.eff}) using the Fourier transform. In (\ref{2.1.4}) the $\langle\hat{c}({\bf r})\rangle_{{\rm ideal},\beta f}$ is the mean density of a spherically symmetric opposite charge ``screening'' cloud that forms around the positive charge at the origin. Note that within the linear approximation $\exp \left(x \right) = 1 +x$ our calculations imply \begin{equation}\label{2.1.13} \int d {\bf r} \, \langle \hat{c}({\bf r})\rangle _{{\rm ideal},\beta f} \sim - e_{0} \end{equation} By Newton's theorem, such a cloud is equivalent to an additional negative charge at the origin that neutralizes the fixed charge up to a remainder exponentially small in the radius to the point ${\bf r} $. Of course no single charge configuration of point particles can be spherically symmetric. The cloud is an ensemble average over configurations. This cancellation of charge is captured in ``sum rules'' which will be discussed in section~\ref{sec-Debye-screening}. Related to this point is the fact that the integral over ${\bf r} _{2}$ of the right hand side of (\ref{2.1.2}) vanishes. \subsection{Theorems on Debye screening} \label{sec-Debye-screening} In this section we will survey the rigorous results on Debye screening and the Debye-H\"uckel approximation concentrating on three dimensional systems but with some remarks about other dimensions. The one dimensional Coulomb system without hard core has been solved exactly \cite{Len61,EdLe62}. Lenard's solution has Debye screening for charge density observables, but one dimensional systems have exceptional screening properties, because the linear potential $-|x-y|$ has no multipoles. For example, the electric field is piecewise-constant with jumps at the positions of particles so that outside an interval containing any neutral configuration, one can achieve a zero electric field. On the other hand if a single fractional charge is placed in a system of integer charges then the resulting potential can never be screened because the ensemble of integer charges can do no more than cancel integer parts of the corresponding electric field. One dimensional Coulomb systems have been reviewed in \cite{CKMN81}. Since then, highly non-trivial results have been obtained on the correlations of classical charges confined to a circle and interacting with a logarithmic potential \cite{For92,For93a,For93b}. Some comments on two dimensional systems are given at the end of this section. For three dimensions the results are complicated to state completely and the complications are probably artifacts of the proofs. Therefore let us restrict ourselves to a simple case of a hard core plasma with two species, whose charges $e_{\alpha }$ have rational ratio. Recall that the combination $\beta e_{\alpha }^{2}$ has dimensions of length. By absorbing a fundamental unit of charge into $\beta $ we assume that $\beta $ carries the dimension of length while $e_{\alpha }$ is dimensionless. To put it another way, we state results regarding $\beta , z_{\alpha }, \lambda $ as parameters, and keeping $e_{\alpha }$ fixed. The results below are valid in a thermodynamic limit in which there are boundary conditions on the Coulomb potential. Details may be found on page 199 of \cite{BrFe80}; the boundary conditions are a little artificial but their essence is that the walls of the container are grounded, so that all electrostatic potentials vanish at the boundary $\partial \Lambda $. \begin{theorem}\label{th-1.3.1} Suppose that \begin{eqnarray} && \mbox{Pseudo-Neutrality: } \sum _{\alpha }z_{\alpha } e_{\alpha } =0 \label{2.2.1}\\ && \mbox{Debye Sphere Assumptions: } \left\{\begin{array}{l} z_{\alpha } \l_{D} ^{3} e^{-\frac{\beta}{2\lambda}} \gg 1\\ z_{\alpha } \lambda ^{3} \ll 1 \end{array} \right. \label{2.2.2} \end{eqnarray} then all (truncated) $n$-point particle and charge density correlations decay exponentially on length scale $l\approx l_{D}$. Furthermore, if $\Xi_{\Lambda }({\bf r} _{1},{\bf r} _{2})$ is the grand canonical charge symmetric ensemble of integer charges with two fractional charges held fixed at ${\bf r} _{1},{\bf r} _{2}$, then the two fractional charges are screened in the sense that \begin{equation}\label{2.2.3} \frac{\Xi_{\Lambda }({\bf r} _{1},{\bf r} _{2})}{\Xi_{\Lambda }} = \frac{\Xi_{\Lambda }({\bf r} _{1})}{\Xi_{\Lambda }} \frac{\Xi_{\Lambda }({\bf r} _{2})}{\Xi_{\Lambda }} + \ O \( \exp \(-|{\bf r}_{1} - {\bf r} _{2}|/l \) \). \end{equation} \end{theorem} The Debye sphere assumptions say that the number of particles in a volume of size $\lambda ^{3}$ is small, so that the hard core is out of play, and the number of particles inside a sphere of size $l_{D}$ is large. For comparing these statements with the ones in the references it is useful to realize that \begin{equation}\label{2.2.3b} 4\pi \sum _{\alpha }e_{\alpha }^{2} z_{\alpha } l_{D} ^{3} = \frac{l_{D}}{\beta } \end{equation} so the first Debye sphere condition is equivalent to $\beta/l_{D}$ being small. Theorem~\ref{th-1.3.1} is a combination of results from \cite{Bry78,BrFe80,Imb83a,BrKe87}. It is shown in \cite{Imb83} that there is Debye screening in Jellium and in systems with irrational charges. Jellium is a limit in which all species with positive charges are smeared out into a background charge density by letting their charge tend to zero as their density increases. The results of Brydges and Federbush were not proved with enough uniformity to persist as this limit is taken. The Debye-H\"uckel limiting laws should hold in the \begin{equation}\label{2.2.4} \mbox{Debye-H\"uckel Limit: } z_{\alpha}\lambda ^{3} \rightarrow 0; \ \ \ z_{\alpha}l_{D} ^{3} e^{-\frac{\beta}{2\lambda}} \rightarrow \infty \end{equation} but published proofs \cite{Ken83,Ken84} have the much more stringent conditions that $\beta /\lambda $ is kept bounded and the system be charge symmetric. Similar results for quantum systems have been established by \cite{Fon86}. The original Debye-H\"uckel results included correction terms containing a hard core\footnote{\label{footnote-dh} $\beta (P - \sum _{\alpha }\rho _{\alpha }) \sim -\frac{\kappa ^{3}}{24 \pi}\sigma (\kappa a)$ where $\sigma (\kappa a) = \frac{3}{(\kappa a)^{3}} \bigg ( 1 + \kappa a - \frac{1}{1+\kappa a} - 2\ln (1+\kappa a)\bigg)$. It may be better to consider $a$ to be a length scale that characterizes ``effective'' short range forces created by cooperation between Coulomb forces and hard cores length as opposed to a genuine hard core.} length scale $a$. More systematically, hard core corrections have been obtained by resummations of Mayer expansions \cite{Mee58,Mee61}\footnote{This is compared with extended Debye-H\"uckel theories in \cite{BeFi98}.} and by the Kac limit developed in \cite{LeSt68} which develop corrections in $\kappa a$. There is also a long history of intuitive extensions of the Debye-H\"uckel theory to include corrections from hard cores and dipolar effects. These are reviewed and advanced in \cite{FiLe93,Fis94,LeFi95,Fis96,LeFi96b,LeFi97,ZuFiLe97}\footnote{There is a competing program \cite{Ste95,YeZhSt96} based in part on the mean spherical approximation instead of Debye-H\"uckel theory.}. However consistency of these approximations is guesswork: is there an asymptotic expansion in $\kappa a$? What physical effects provide the next largest corrections to Debye-H\"uckel theory? In the proofs of screening complications at the boundary have been avoided by artificial boundary conditions, but only laziness stands in the way of proofs for the case where the particles are in a conducting grounded container so that the Coulomb potential vanishes at the boundary. Particles with $1/r$ forces, i.e., particles in an insulating container, have not been completely analyzed, but Debye screening for a simplified model is established \cite{FeKe85}. This case is discussed some more in section~\ref{sec-tunneling}. It is difficult because screening fails at the boundary; there are power law forces between charges on the boundary. It should be possible to omit the pseudo-neutrality condition (\ref{2.2.1}), but to do it rigorously may require insight into the boundary condition problem. In section \ref{sec-charge-expulsion} we will discuss these issues along with charge expulsion. We return to (\ref{2.1.13}). The $\sim$ involves approximations but the theorems known as sum rules say that if truncated distributions decay integrably or better then this is exact in the limit $e_{0} \rightarrow 0$: the screening cloud around a fixed infinitesimal charge neutralizes it. More generally any fixed set of charges surrounds itself with a screening cloud in such a way that all multipoles are canceled. In particular, the function of ${\bf r} $ defined by \begin{eqnarray}\label{2.2.5} && c_{T}({\bf r}|{\bf r}_{1},e_{\alpha_{1}},\ldots,{\bf r} _{n},e_{\alpha _{n}})= \nonumber \\ && \hspace{-1in} \langle \hat{c}({\bf r} ) \hat{\rho} ({\bf r} _{1},e_{\alpha _{1}}) \cdots \hat{\rho} ({\bf r} _{n},e_{\alpha _{n}}) \rangle - \langle \hat{c}({\bf r} ) \rangle \langle \hat{\rho} ({\bf r} _{1},e_{\alpha _{1}}) \cdots \hat{\rho} ({\bf r} _{n},e_{\alpha _{n}}) \rangle \end{eqnarray} which represents the excess charge density at ${\bf r}$ when particles in the system are fixed at ${\bf r}_{1},\ldots,{\bf r}_{n}$ has no multipole moments in the Debye regime \begin{equation}\label{2.2.5a} \int d{\bf r} g_{l}({\bf r})c_{T}({\bf r}|{\bf r}_{1},e_{\alpha_{1}},\ldots,{\bf r} _{n},e_{\alpha _{n}})=0 \end{equation}for all harmonic polynomials $g_{l}({\bf r})$ of order $l=0,1,\ldots $. In (\ref{2.2.5a}) the contribution of coincident points ${\bf r} ={\bf r}_{i}$ is included. The case $l=0,\; g_{0}({\bf r})=1$, called the charge sum rule, is expected to hold very generally in homogeneous phases of Coulomb systems (classical and quantum). These results are reviewed in \cite{Mar88}. Results on the effect of slow decay at boundaries are also covered in this review. In the next section, we present a derivation of the sum rules with the help of the Sine-Gordon formalism, extending the arguments of \cite{FoMa84} to charges with hard cores. There are also results \cite{AlMa85} that say that if truncated distributions are integrable and decay monotonically at infinity then they must decay faster than any inverse power. A different result of the same genre was obtained by \cite{Fed79}. \emph{Two dimensions, no short range forces}: in units where the Coulomb interaction between two unit charges is $-\ln r$ let $\beta _{1}= 2, \beta _{\infty } = 4$ and define the sequence of intermediate thresholds $\beta _{n} = \beta _{\infty }(2n-1)/(2n)$ with $n = 1, 2, \dots $. The Coulomb potential is not stable in the sense of (\ref{1.1.3b}) but the instability is weak enough to permit thermodynamic behavior anyway \cite{HaHe71,DeLa74,Fro76}, at least for $\beta < \beta _{1}$. At $\beta _{1}$ the Gibbs factor $\exp { -\beta _{1}V_{2}}$ for two oppositely charge particles is no longer integrable and the partition function diverges. The partition function for the Yukawa gas also diverges and, in the Yukawa gas Mayer expansion, the diagram with two vertices becomes infinite at $\beta = \beta _{1}=2$. However, all diagrams with more vertices remain finite. Similarly, at $\beta _{2}$, neutral diagrams with 2, 4 vertices diverge, all diagrams with more vertices remain finite, and so on. It has been proved \cite{BrKe87} that for $\beta < 4/3 \beta _{1}$ the Mayer expansion without the infinite two vertex term is convergent for small activity. The natural extension of this result to the higher thresholds almost certainly holds for $\beta < \beta _{\infty}$ \footnote{An almost equivalent result appeared in \cite{DiHu93}, but the authors have reported a serious error invalidating their proof \cite{Dim98}}. These results imply that there are natural infinite renormalizations of the partition function. Presumably, this enables some version of the Yukawa and Coulomb gases to be defined in the range $\beta \in [\beta _{1}, \beta _{\infty })$ as a renormalized limit of systems with regularized interaction, because dropping a divergent term in the Mayer expansion is the same as multiplying the partition function by an infinite factor. This has not been discussed clearly in the literature, but see \cite{GaNi85b,Spe86}. It has been proven \cite{Yan87} that the two species charge symmetric Coulomb gas of point particles in a grounded container has exponential screening if $\beta \ll \beta _{1}$ and the activity are small. The grand canonical ensemble with insulating boundary conditions is harder to analyze because it has to be restricted to neutral configurations, whereas image charges provide neutrality for free in the grounded container case. As the last paragraph suggests, boundary conditions are more important in two dimensions than in three. For example it is known \cite[Theorem 4.1]{FrSp81b} that a pair of oppositely charged fractional charges immersed in a two dimensional Coulomb gas are not screened from each other when the thermodynamic limit is taken with insulating container boundary conditions. On the other hand, although the details are not published anywhere, it ought to follow from the results of Yang that they are screened when the thermodynamic limit is taken with grounded containers. This influence of boundary conditions does not happen in three dimensions. We will discuss this further in section~\ref{sec-tunneling}. \emph{Two dimensions, with stabilizing short range forces}: The significance of $\beta _{\infty }$ is that the Kosterlitz-Thouless transition takes place at a critical $\beta_{c} \approx \beta _{\infty }$\footnote{$ \beta_{c} \rightarrow \beta _{\infty } $ as the density tends to zero}. By stabilizing the interaction with short range forces such as hard cores one can reach $\beta _{c}$ without infinite renormalizations and see the $\beta > \beta _{c}$ dipole phase. The original argument \cite{KoTh73} was given a complete proof \cite{FrSp81} for the Coulomb system on a lattice. A more detailed analysis of the Kosterlitz-Thouless phase, based on term by term analysis of the Mayer expansion, is given in \cite{AlCo92,AlCo97a,AlCo97b}. There is confusion over whether the thresholds $\beta _{n}$ should be interpreted as the successive collapse into neutral clusters of 2, 4,\dots particles. \cite{GaNi85b,Spe86} proposed such an interpretation based on successive divergences in the Mayer expansion. This is analogous to the Yukawa gas alluded to above, except that now infra-red divergences cause the thresholds. To complete their argument one must show that there are bulk observables whose correlations have associated singularities. On the other hand \cite{FLL95} argue that no singularities show up in thermodynamic functions and consequently the thresholds are not physical. The heuristic renormalization group argument \cite{KoTh73} predicts that exponential screening of particle distributions holds for all $\beta < \beta _{c}$ if the density is sufficiently small. A complete proof of screening at low activity for $\beta $ in this range would be a wonderful achievement. \emph{Exact formulas in two dimensions} for the excess free energy of a single charge fixed at the origin are conjectured in \cite{LuZa97}. See also \cite{Smi92} for other interesting formulas. These formulas in principle could be used to express the excess free energy of a fractional charge in terms of the density ( or the activity) and temperature. The authors study the Sine-Gordon functional integral with boundary conditions that correspond to a grounded conducting container so that violating neutrality by fixing a fractional charge makes sense. Their conjecture translates, via the Sine-Gordon transformation to the two component charge symmetric gas with a purely Coulomb potential at least for $\beta < \beta _{1}$. With limitations on the size of the fractional charge their formulas apply for $\beta \in [\beta _{1}, \beta _{\infty })$. In two dimensions, there are a number of exactly solvable models for the special value of the temperature that corresponds to $\beta_{1}$, starting with the two dimensional one component plasma \cite{Jan81}. These models \cite{Ala87,Jan90,Jan92} exhibit "super-Debye screening", namely Gaussian decay of the correlations. They have played an important role in testing various refined screening properties in presence of walls and of different types of inhomogeneities. Although a characteristic of Debye phases is the rapid decay of the particle correlations, one should be aware that it is never the case for potential and electric field fluctuations \cite{LeMa84,Mar88}. For instance, correlations of the components of the electric field $E_{\mu}$ always have an asymptotic dipolar behavior \begin{equation}\langle E_{\mu}({\bf r}) E_{\nu}({\bf 0})\rangle \simeq -\frac{1}{\beta} \nabla_{\mu}\nabla_{\nu}\left(\frac{1}{\mid{\bf r}\mid}\right) \label{2.2.2field} \end{equation}This observation has recently led to the development of the interesting view point that there are some universal properties of Debye phases that closely resemble those of a system with short range forces at its critical temperature \cite{JMP94,Jan96,FoJaT96}. Actually, potential and field fluctuations are critical in Debye phases, and this leads to universal finite size corrections to the free energy as in critical systems. \subsection{The Sine-Gordon transformation}\label{sec-sinegordon} \subsubsection{The Fourier transform of a Gaussian is a Gaussian} In chapter~\ref{chapter- Coulomb systems} we have defined a potential $$ V_{\infty,\lambda }({\bf r} ) = |{\bf r}|^{-1}(1-\exp(-|{\bf r}|/\lambda )) $$ which is a Coulomb potential smoothed at the origin. We saw that $V_{\infty,\lambda }$ has a positive Fourier transform. For such potentials there is an identity that expresses the interaction in terms of external potentials --- an exact version of the mean field idea of Debye-H\"uckel. This is the Sine-Gordon transformation \cite{Kac59,Sie60}. Kac and Siegert noticed that if self-energies are included, then the Gibbs factor \begin{equation}\label{2.3.3a} e^{ -\frac{\beta }{2} \sum_{i, j=1}^N e_{\alpha_i} e_{\alpha_j} V_{\infty,\lambda }({\bf r} _i ,{\bf r} _j ) } = e^{ -\frac{\beta }{2} \int d {\bf r} \, \int d{\bf r} ' \, \hat{c}({\bf r} ) V_{\infty ,\lambda }({\bf r} ,{\bf r}') \hat{c}({\bf r} ') } \end{equation} is Gaussian in $\hat{c}({\bf r})$. By a functional integral generalization \cite{Sim79} of ``the Fourier transform of a Gaussian is a Gaussian'' there exists a Gaussian probability measure, intuitively described by \begin{equation}\label{2.3.3} d\mu_{\lambda } (\phi) = {\cal D} \phi \, e^{ - \frac{1}{2} \int \phi({\bf r} ) V_{\infty,\lambda }^{-1} \phi({\bf r} ) d{\bf r} } \end{equation} such that \begin{equation}\label{2.3.1} e^{ -\frac{\beta }{2} \sum_{i, j=1}^N e_{\alpha_i} e_{\alpha_j} V_{\infty,\lambda }({\bf r} _i ,{\bf r} _j ) } = \int d\mu _{\lambda }(\phi) e^{ -i\beta ^{1/2} \int d {\bf r} \, \hat{c}({\bf r} ) \phi({\bf r} ) }. \end{equation} $V_{\infty,\lambda }^{-1}$ is the operator inverse of the operator whose kernel is $V_{\infty,\lambda }$. In our case we find from (\ref{1.1.3bd}) that it is the partial differential operator \begin{equation}\label{2.3.3b} V_{\infty,\lambda }^{-1} = \frac{1}{4\pi }\( -\Delta + \lambda ^{2} \Delta ^{2} \) \end{equation} possibly with boundary conditions on $\Delta $ . Equation (\ref{2.3.1}) says that the Gibbs factor is a superposition of Gibbs factors for external potentials $i\phi$. It follows that if an ideal gas of particles in an external potential is integrated over the external potentials then the result is a partition function for a gas with two-body interaction $V_{\infty,\lambda }$, including self-energies. Self-energies are equivalent to a shift in the activities. We define \begin{equation}\label{2.3.4a} z^{(\lambda)}({\cal E} ) = z({\cal E}) e^{ \frac{\beta }{2}e_{\alpha}^{2} V_{\infty,\lambda }({\bf r} ,{\bf r} ) } \end{equation} Then the partition function with potential energy $U(\omega )$ built out of $V_{\infty,\lambda }$ and an external potential $\psi $, \begin{equation}\label{2.3.4aa} \Xi_\Lambda (\beta ^{1/2}\psi ) = \int _{\Lambda } d\omega \, z(\omega ) e^{ -\beta U(\omega ) - \beta ^{1/2} \int \psi \hat{c}(\omega )d{\bf r} } \end{equation} is given by \begin{equation}\label{2.3.4c} \Xi_\Lambda (\beta ^{1/2}\psi ) = \int d\mu _{\lambda }(\phi) \, \Xi_{{\rm ideal},\Lambda,z^{(\lambda)}} (\beta ^{1/2}[i\phi + \psi]). \end{equation} where \begin{equation}\label{2.3.4b} \Xi_{{\rm ideal},\Lambda,z^{(\lambda)}}(f) = \int _{\Lambda } d\omega \, z^{(\lambda)}(\omega ) e^{-\int f\hat{c}d{\bf r} } \end{equation} is the ideal gas partition function (\ref{1.1.11}) with the renormalized activities (\ref{2.3.4a}). (\ref{2.3.4c}) is the Sine-Gordon representation. It says that the partition function for the interacting system exactly equals a (functional) integral over \emph{imaginary} external fields of the ideal gas partition function. Gaussian measures have been studied in the mathematical literature and it is known that for this one the typical potential $\phi$ is a continuous function \cite[Theorem~A.4.4]{GlJa87}. This result holds provided $\lambda \not =0$; when $\lambda =0$ the instability of the Coulomb potential returns to haunt us in the shape of the typical $\phi$ becoming a distribution instead of a function. The right hand side of formula (\ref{2.3.3}) is best understood as a mnemonic for the translation formula, also known as the Cameron-Martin formula, \begin{equation}\label{2.3.5} d\mu _{\lambda }(\phi+g) = d\mu _{\lambda }(\phi) \, e^{-\frac{1}{2}\int g V_{\infty,\lambda }^{-1} g d{\bf r} - \int \phi V_{\infty,\lambda }^{-1} g d{\bf r} } \end{equation} which is valid for any function $g$ for which $\int g V_{\infty,\lambda }^{-1}g$ exists\footnote{In the sense $\int (\mid \nabla g\mid ^{2} + \mid \lambda \Delta g\mid ^{2})\, d{\bf r} < \infty $.}. Translations $\phi \rightarrow - ig$ are legitimate as well whenever the integrand permits analytic continuation to $\alpha = i$ after the change of variable $\phi \rightarrow \phi + \alpha g$. The translation formula is the reason why computations with the intuitive formula (\ref{2.3.3}) are usually correct. It is not physically very reasonable to try to capture the very local fluctuations on length scales much smaller than $l_{D}$ in a mean field picture. For example, the activity $z^{(\lambda )}$ is very different from $z$ when $\lambda $ is small, but we shall see in section~\ref{sec-debye-huckel2} that $z$ is the correct effective activity, in the sense that an ideal gas with this activity is the best approximation in the Debye-H\"uckel limit. Also hard cores do not have positive Fourier transforms and cannot be represented by a Sine-Gordon potential. For these reasons we now consider a better class of representations in which the particle picture is retained on scales less than $O(l_{D})$ and the mean field or Sine-Gordon is used to represent interactions on larger scales. Define the Yukawa interaction with a decay length $L \geq \lambda $, \begin{equation}\label{2.3.6} V_{L}({\bf r} ) = |{\bf r} |^{-1}e^{-|{\bf r} |/L}; \ \ \ V_{L,\lambda }({\bf r} ) = |{\bf r} |^{-1} \left( e^{-|{\bf r} |/L} - e^{-|{\bf r} |/\lambda } \right) \end{equation} We split the interaction $V_{\infty,\lambda }$ according to \begin{equation} V_{\infty,\lambda } = V_{\infty,L} + V_{L,\lambda } \end{equation} and use the Sine-Gordon transformation to represent just the part $V_{\infty,L}$. The result is \begin{equation}\label{2.3.4} \Xi_\Lambda (\beta ^{1/2}\psi ) =\int d\mu _{L} (\phi) \, \Xi_{L,\Lambda,z^{(L)}} (\beta ^{1/2} [i\phi + \psi]) \end{equation} where \begin{equation}\label{2.3.7} \Xi_{L,\Lambda,z^{(L)}}(f) = \int _{\Lambda } d\omega \, z^{(L)}(\omega ) e^{-\beta U_{L}(\omega ) - \int f \hat{c}(\omega )d{\bf r} } \end{equation} is no longer ideal; $U_{L}$ is built out of two-body interactions $V_{L,\lambda }$. If hard cores are used in place of the cutoff $\lambda $ in $V_{L,\lambda }$ then $U_{L}$ is built out of $V_{L}$ and hard cores. The replacement of $V$ by $V_{\infty,L}$ is the same as smoothing the Coulomb interaction by introducing form factors for the charges so they are smeared out into spherical distributions of characteristic radius $L$. Note that \begin{equation}\label{2.3.formfactor} \tilde{V}_{\infty,L}({\bf k}) = \tilde{F}(L^{2}k^{2}) \frac{4\pi}{k^{2}} \tilde{F}(L^{2}k^{2}), \ \ \ \tilde{F}(L^{2} k^{2} ) = \left( \frac{1}{1+L^{2} k^{2}}\right)^{1/2} \end{equation} which shows that $\tilde{F}$ is the Fourier transform of the form factor. \subsubsection{Sum rules} \newcommand{\EE}[1]{\langle #1 \rangle } \newcommand{\Ephi}[1]{\langle #1 | \phi \rangle } \newcommand{\Ephibar}[1]{\langle #1 | \psi \rangle } \newcommand{\E}[1]{E\!\!\!\!E [ #1 ] } Sum rules provide a good illustration of aspects of the Sine-Gordon transformation that will reappear in the sequel. Here we sketch how the argument in \cite{FoMa84}\footnote{This paper uses the smooth short range regularization (\ref{1.1.3bc}).} applies to a Coulomb system with hard cores. We will show that sum rules result from an invariance of the Sine-Gordon measure under infinitesimal translation by harmonic functions. Formulas arising from infinitesimal symmetries are called Ward identities in quantum field theory. While it is interesting that sum rules are Ward identities, it is also easy to derive them without the Sine-Gordon transformation \cite{Mar88,BrMa99}. We consider charged particles with hard core interactions of radius $\lambda $ in a box $\Lambda $ whose boundary is grounded and impenetrable to the hard cores. We spread each charge in the ensemble uniformly onto the surface of its hard core sphere. By Newton's theorem the interaction energy $U$ is unchanged by the spreading out. However the self-energy of each particle becomes finite and of the order of $\frac{1}{\lambda }$. It can be added into $U$ and the change in $U$ is compensated by replacing $z_{\alpha }$ by $ z_{\alpha }^{(\lambda )} = z_{\alpha } \exp ( \beta e_{\alpha }^{2}/ (2\lambda) ) $. By the Sine-Gordon transformation \begin{eqnarray}\label{sum-rules.2} && \Xi_\Lambda = \int \Xi_\Lambda (i \beta^{1/2} \bar{\phi}_{\lambda} ) \, d\mu (\phi ) \nonumber \\ && d\mu (\phi ) = e^{-1/(8\pi ) \int_\Lambda (\partial \phi)^{2} } {\cal D} \phi \end{eqnarray} where $\bar{\phi}_{\lambda}({\bf r})$ is the average of $\phi $ over a sphere of radius $\lambda $ and center ${\bf r}$ and $\Xi_\Lambda (\phi )$ is the hard core gas in external field $\phi $: \begin{eqnarray}\label{sum-rules.3} && \Xi_\Lambda (\phi ) = \sum_{\{N_\alpha\}}^\infty \prod_{\alpha = 1}^{{\cal S} } \frac{(z_\alpha^{(\lambda )} )^{N_\alpha}}{N_\alpha !} \int_{\Lambda} d{\bf r}_{1} \ldots \int_{\Lambda} d{\bf r}_{N} \exp\( - \sum _{j} e_{\alpha_j} \phi ({\bf r} _{j}) - {\rm hard \ core} \) \end{eqnarray} $d\mu (\phi) $ is characterized by \begin{equation}\label{sum-rules.4} d\mu (\phi + g) = d\mu (\phi ) e^{1/(4\pi ) \int_\Lambda \phi \Delta g - 1/(8\pi ) \int_\Lambda (\partial g)^{2} }; \end{equation} where $g$ is any function which vanishes on the boundary with $\int_\Lambda (\partial g)^{2} < \infty $. The complete expectation \begin{eqnarray}\label{sum-rules.5} \langle F \rangle &=& \frac{1}{\Xi_\Lambda } \sum_{\{N_\alpha\}}^\infty \prod_{\alpha = 1}^{{\cal S} } \frac{(z_\alpha^{(\lambda )} )^{N_\alpha}}{N_\alpha !} \int_{\Lambda} d{\bf r}_{1} \ldots \int_{\Lambda} d{\bf r}_{N} \nonumber\\ &\times& \exp\( - \beta U - {\rm hard \ core} \) F \end{eqnarray} can be taken in two stages; the first stage is an expectation conditioned on $\phi $ \begin{eqnarray}\label{sum-rules.6} \Ephi{F} &=& \frac{1}{\Xi_\Lambda ( \phi ) } \sum_{\{N_\alpha\}}^\infty \prod_{\alpha = 1}^{{\cal S} } \frac{(z_\alpha^{(\lambda )} )^{N_\alpha}}{N_\alpha !} \int_{\Lambda} d{\bf r}_{1} \ldots \int_{\Lambda} d{\bf r}_{N} \nonumber\\ &\times& \exp\( - \sum _{j} e_{\alpha_j} \phi ({\bf r} _{j}) - {\rm hard \ core} \) F\nonumber \end{eqnarray} and the second is the remaining integration over $\phi $ \begin{eqnarray}\label{sum-rules.7} \E{F} &=& \int \ e^{\ln \Xi_\Lambda (i \beta^{1/2} \phi )} \ F \ d\mu (\phi ) \ / \int \ e^{\ln \Xi_\Lambda (i \beta^{1/2} \phi )} \ d\mu (\phi ) \end{eqnarray} and these fit together by \begin{equation}\label{sum-rules.8} \langle F \rangle = \E{ \Ephibar{F} } \mbox{ where } \psi = i \beta^{1/2} \bar{\phi}_{\lambda} \end{equation} To obtain a sum rule we make the change of variables $\phi \rightarrow \phi + t g$ in the $d\mu (\phi )$ integral on the right hand side and differentiate with respect to $t$ at $t = 0$. The derivative must vanish since the left hand side is independent of $t$. The derivative of $\ln \Xi_\Lambda (\phi )$ with respect to $\phi ({\bf r} )$ is $-\Ephi{\hat{c}({\bf r} )}$, which accounts for the first line in the next equation. The second line arises because the derivative of $\Ephi{F}$ with respect to $\phi ({\bf r})$ is the truncated expectation $ - \Ephi{F \hat{c}({\bf r} )} + \Ephi{F} \Ephi{\hat{c}({\bf r} )} $. \begin{eqnarray}\label{sum-rules.9} 0 &=& \E{ \Ephibar{F } \ \Ephibar{\hat{c}(\bar{g}_{\lambda })}} - \E{ \Ephibar{F } } \ \E{ \Ephibar{\hat{c}(\bar{g}_{\lambda } ) } } \nonumber \\ &+& \E{ \Ephibar{F \hat{c}(\bar{g}_{\lambda })} - \Ephibar{F } \ \Ephibar{\hat{c}(\bar{g}_{\lambda })} } + \mbox{contributions from } (\ref{sum-rules.4}) \end{eqnarray} The contributions from (\ref{sum-rules.4}) are proportional to \begin{equation}\label{sum-rules.9b} \E{\Ephibar{F} \int \phi \Delta g } - \E{\Ephibar{F}} \E{ \int \phi \Delta g } \end{equation} We would like to choose $g$ to be a harmonic polynomial because then these contributions vanish. However this is not a legal choice of $g$ because no harmonic polynomial can vanish on the boundary of $\Lambda $. Instead choose $g$ to be a harmonic polynomial multiplied by another function $h$ that is one in the interior of $\Lambda $ and which goes to zero near the boundary. Now there are non-vanishing contributions from $\int \phi g \Delta h$ and $\int \phi \partial g \partial h$ localized near the boundary. Suppose that the container $\Lambda $ is enlarged: there is competition between the growth of the harmonic polynomial $g ({\bf r} )$ for ${\bf r} $ near the boundary and the decay of the truncated expectation (\ref{sum-rules.9b}). If the system is in a screening phase where truncated expectations of local functions of $\phi $ decay exponentially then the polynomial growth of $g$ is crushed as the container is enlarged and the contributions in (\ref{sum-rules.9b}) vanish in the thermodynamic limit. It is true that the conditional expectations $\Ephibar{\hat{c}({\bf r} ) }$ are not exactly local in $\phi $ but if the Mayer expansion of $\Ephibar{\hat{c}({\bf r} ) }$ is convergent then $\Ephi{\hat{c}({\bf r} )}$ is almost (exponentially) local in $\phi $. This captures a physically natural condition that the system should not be so dense that there are long range correlations caused by the hard core. If the Mayer expansion converges at $\psi =0$ then it converges uniformly for all $\psi $ because $\psi \not = 0 $ just means that activities are multiplied by modulus one phase factors. Thus in the infinite volume limit (\ref{sum-rules.9}) simplifies to $ 0 = \E{ \Ephibar{F \hat{c}(\bar{g}_{\lambda })} } - \E{ \Ephibar{F } } \ \E{ \Ephibar{\hat{c}(\bar{g}_{\lambda }) } } $ where $g$ is a harmonic polynomial. Now we simplify further using (\ref{sum-rules.8} ) and Newton's theorem (spherical average of harmonic = value at center) $\bar{g}_{\lambda }({\bf r}) = g({\bf r})$ to reach \begin{eqnarray}\label{sum-rules.11} 0 &=& \EE { F \hat{c}(g) } - \EE{F} \EE{\hat{c}(g)} \ \ \ g \mbox{ any harmonic polynomial} \end{eqnarray} For example by choosing $F = \hat{\rho}({\bf r}_{1},\alpha_{1})\ldots \hat{\rho}({\bf r}_{n},\alpha_{n}) $ (\ref{sum-rules.11}) becomes the same as (\ref{2.2.5a}). In conclusion, the sum rules (\ref{2.2.5a}) for a system of classical charges with hard cores hold if the Mayer expansion for the hard core system without Coulomb interaction is convergent and if the system is in a screening phase. \subsection{Debye spheres}\label{sec-debyesphere} In the previous section we divided the interaction into the Yukawa part $V_{L,\lambda }$ whose decay length is $L$ and a remaining part, the slow part $V_{\infty,L}$, which carries the slowly varying long range piece of the Coulomb interaction. In this section, and in detail in section~\ref{sec-mayer2}, the Yukawa part is shown to leave the gas essentially ideal, provided its range $L$ is of order the Debye length $l_{D}$ and the Debye sphere assumptions (\ref{2.2.2}) are imposed. The underlying idea is that the effect of the $V_{L,\lambda }$ part of the interaction is to associate some of the monopoles into aggregates which can be regarded as new species so that the resulting multi-species gas is ideal. The Mayer expansion is used to express this. Instead of leaving out the long range part of the interaction, one can instead say that the gas inside a sphere of radius less than the Debye length is close to ideal. This illuminates the Debye-H\"uckel argument in which the charge density was computed by an ideal gas formula: the mean field captures the part of the interaction with range larger than $L$ and the interactions of shorter range are small enough to use an ideal gas calculation. Consider the partition function for a Yukawa gas in an external potential $\phi$, \begin{equation}\label{2.4.1.1a} \Xi_{L,\Lambda }(\beta ^{1/2} \phi) = \int d\omega \, z(\omega ) e^{-\beta U_{L}(\omega ) - \beta ^{1/2} \int \phi \hat{c}(\omega )d{\bf r} } \end{equation} The Mayer expansion is an expansion for the logarithm of the partition function or the pressure $P$ in powers of the activities $z_{\alpha }$ whose first term $\beta P V = \sum z_{\alpha } + \dots $ gives the ideal gas law when $\phi=0$: it has the form \begin{eqnarray} \label{2.4.1.2} && \Xi_{L,\Lambda}(\beta ^{1/2}\phi) = \nonumber \\ && \hspace{.2in} \exp \left(\sum _{m \geq 1} \frac{1}{m!} \int \prod _{k} d{\cal E}_{k} z({\cal E}_{k}) e^{-\beta ^{1/2} e_{\alpha _{k}}\phi({\bf r} _{k})} u_{m}({\cal E} _{1},\dots ,{\cal E} _{m}) \right) \end{eqnarray} where $u_{m}$ are given \cite{Rue69,HaMc76} as sums of contributions from connected ``Mayer'' graphs. In particular $u_{1}({\cal E}) = 1$ so that the $m=1$ contribution to the product is $$ \exp \left[\sum _{\alpha } \int d{\bf r} \, z_{\alpha} e^{-\beta ^{1/2} e_{\alpha}\phi}\right] $$ which equals the ideal gas partition function $\Xi_{{\rm ideal},\Lambda }(\beta ^{1/2} \phi) $ that appeared above in (\ref{1.1.11}). The higher $m$ terms in (\ref{2.4.1.2}) have the form of additional ideal gas species, e.g., the $m = 2$ term is an ideal gas partition function for a gas whose particles are the possible aggregates of two of original particles, with a cluster activity $\frac{1}{2!}z({\cal E}_{1}) z({\cal E}_{2})u_{2}({\cal E} _{1},{\cal E} _{2})$. Such a cluster describes two particles bound together because the functions $u_{m}$ decay exponentially in the separation of any pair of arguments, because the graphs that contribute to them are connected. Although these ``Mayer'' aggregates do capture effects of aggregates of particles they are not directly physical because their activities need not be positive. This decomposition of the interacting gas into species of ideal gases is only useful if the activities of large aggregates are small. In fact we need them to be summable which is the same as demanding that the Mayer expansion is convergent. The standard condition \cite{Rue69} for convergence for the activity expansion for a gas with activity $z({\cal E} )$ and two-body potential $v({\cal E} , {\cal E} ')$ is $$ \sup _{{\cal E} } \int |e^{-\beta v({\cal E} ,{\cal E} ')}-1| |z({\cal E} )e^{-\beta ^{1/2}\phi({\cal E} )}| d{\cal E} ' e^{2B\beta } < e^{-1} $$ where $B$ is the stability constant, i.e., the best constant such that $$ \sum _{1 \leq i<j \leq N}v({\cal E} _{i}, {\cal E} _{j}) \geq -B N. $$ Suppose for simplicity that there are two species with equal but opposite charges $e_{\alpha } = \pm 1$ and equal activities $$ z({\cal E} ) = z_{\alpha } = z $$ Using the stability estimate (\ref{1.1.3b}) we find the Mayer expansion converges if \begin{equation}\label{2.4.ruelle} |z| \beta L^{2} \exp(\beta/\lambda ) \ll 1 \end{equation} which we calculated using $\exp (\pm V_{L,\lambda }) - 1 \approx \pm V_{L,\lambda }$. This condition is sufficient for convergence of the Mayer expansion, but it is not a very accurate condition when $\lambda < \beta < L$. In section~\ref{sec-mayer2} we will show that there is a better condition, namely \begin{equation}\label{2.4.1.5} z \beta ^{3} e^{\frac{\beta }{ \lambda }} \ll 1; \ \ \ z \beta L ^{2} e^{\frac{\beta }{ \beta }} \ll 1. \end{equation} The $\beta /\beta $ emphasizes why the estimate is an improvement. For the Yukawa potential and hard core, one needs \begin{equation}\label{2.4.1.5b} z_{\alpha} \lambda ^{3} \ll 1 \end{equation} as well. Now observe that when $L = \gamma l _{D}$ the second condition in (\ref{2.4.1.5}) drops out, provided $\gamma $ is small, but independent of $z, \beta, \lambda $. The other two conditions are implied by our Debye sphere hypotheses (\ref{2.2.2}), when combined with the definition of $l_{D}$. Thus the Debye sphere hypotheses say that the gas inside a sphere of radius $L = \gamma l _{D}$ is a multi-species ideal gas whose activities are small according to \begin{equation}\label{2.4.1.5c} \frac{1}{m!} \int \prod _{k} d{\cal E}_{k} |z({\cal E}_{k})| |u_{m}({\cal E} _{1},\dots ,{\cal E} _{m})| \leq O\left(\frac{L}{l_{D}}\right)^{2(m-1)} |z| |\Lambda | \end{equation} where $z$ is the largest of the $z_{\alpha }$. For details on the derivation of this estimate see section ~\ref{sec-mayer2}. The conditions for convergence of the Mayer expansion when there is an external potential are the same except that the exponential of the potential is absorbed into the activity. Thus an imaginary potential has no effect in (\ref{2.4.1.5c}) because of the absolute value in $|z({\cal E}_{k})|$. \subsection{The Debye-H\"uckel limit}\label{sec-debye-huckel2} The standard derivation \cite{May50} of the thermodynamics, the first part of the Debye-H\"uckel law (\ref{2.1.2}), proceeded by resumming a class of ring diagrams in the Mayer expansion or low density expansion. A more systematic basis for this resummation is based on the Kac limit, see \cite{LeSt68}. In this section, following \cite{Ken82,Ken83,Ken84}, we sketch a proof of the Debye-H\"uckel law based on the fact that the representation (\ref{2.3.4}) becomes an explicitly integrable Gaussian integral in the Debye-H\"uckel limit (\ref{2.2.4}). The Debye-H\"uckel mean field emerges as the stationary point for the Sine-Gordon action. This line of argument has the advantage of not using any expansion which might be divergent. Our definition of the Debye-H\"uckel limit is low density in the extreme sense that the hard core becomes irrelevant and the effects of association into dipoles and other aggregates are out of play. The original Debye-H\"uckel theory (footnote~\ref{footnote-dh}) has terms in $\kappa a$ that would disappear in this limit if $a$ is taken to be the hard core radius. Debye-H\"uckel theories (see the discussion below (\ref{2.2.4})) have been advocated as a good description for some transitions in ionic fluids at higher densities in which phase separation occurs. These are extensions of Debye-H\"uckel theory in which hard core corrections remain and effective dielectric constants, which represent the effect of charges pairing to form dipoles, are used. For consistency considerations it is desirable to show that these theories are also limiting cases of the Coulomb system or obtainable in some systematic approximation. Let us work in units where $l_{D} = 1$. This means that in the right hand side of the representation (\ref{2.3.4}) we rescale lengths by ${\bf r} = l_{D}{\bf r} '$ so that the Debye length becomes ${l'}_{D} = 1$ and ${\bf r} '$ becomes dimensionless. Then we have exactly the same formulas but the parameters are replaced by dimensionless parameters: \begin{eqnarray}\label{2.5.1} \beta' = \beta /l_{D}, \ \ \ \lambda' = \lambda /l_{D}, \ \ \ {z'}_{\alpha } = {z}_{\alpha } l_{D} ^{3}, \ \ \ L' = L/l_{D}, \ \ \ \Lambda' = l_{D}^{-3}\Lambda \end{eqnarray} In terms of the primed parameters the Debye-H\"uckel limit becomes \begin{eqnarray}\label{2.5.2} {z'}_{\alpha } \rightarrow \infty , \ \ \ \beta ' \rightarrow 0 \mbox{ with } \left\{ \begin{array}{l} 4\pi \sum_{\alpha } e_{\alpha }^{2} {z'}_{\alpha } \beta' = 1 \\ {z'}_{\alpha }{\lambda '} ^{3} \rightarrow 0\\ {z'}_{\alpha } {l'}_{D}^{3} e^{-\beta' /2\lambda' } \rightarrow \infty \end{array}\right. \end{eqnarray} From now on in this section all formulas involve the primed quantities, so we will save on notation by dropping the primes. We write (\ref{2.3.4}) in the form \begin{eqnarray}\label{2.5.3} \Xi_{\Lambda }(\beta ^{1/2}\psi ) &=& \int d\mu_{L} (\phi) \, \exp \left( F(i\phi + \psi ) \right) \end{eqnarray} where \begin{equation}\label{2.5.3a} F(i\phi + \psi ) = \ln \Xi_{L , \Lambda, z^{(L)}} (i\beta ^{1/2} \phi + \beta ^{1/2} \psi ) \end{equation} We define the action functional $S(\psi,g)$ by \begin{equation}\label{2.5.3b} -S(\psi,g ) = \frac{1}{2} \int g V_{\infty,L }^{-1} g d{\bf r} + F(g + \psi ) \end{equation} The partition function $\Xi_{\Lambda }(\beta ^{1/2}\psi )$ does not depend on $L$ in the left hand side of (\ref{2.5.3}) so we may take the infinite volume and Debye-H\"uckel limits at fixed $L$ and then afterwards let $L \rightarrow 0$. The correct order of limits is infinite volume limit followed by Debye-H\"uckel limit, but we do it in the opposite order. A serious technical point in the papers we cited is to show that these limits can be interchanged. In the Debye-H\"uckel limit the action $S(\psi,i\phi )$ becomes a quadratic polynomial in $\phi$. We forestall the appearance of some terms linear in $\phi$ by using the translation formula (\ref{2.3.5}) with $\phi \rightarrow \phi - ig$ applied to (\ref{2.5.3}) \begin{eqnarray}\label{2.5.4} && \Xi_{\Lambda }(\beta ^{1/2}\psi ) = e^{\frac{1}{2}\int g V_{\infty,L }^{-1} g d{\bf r} } \int d\mu_{L} (\phi) \, \nonumber \\ &&\hspace{1in} \times \exp \left[ \int i\phi V_{\infty,L}^{-1} g d{\bf r} + F(i\phi + \psi + g) \right] \end{eqnarray} $g$ will be chosen later. By (\ref{2.4.1.5c}) the Mayer expansion for $F(i\phi + g + \psi )$ is convergent uniformly in the Debye-H\"uckel limit, provided $L \ll 1$: indeed the $O({z}^{N+1})$ term in the Mayer expansion is $O({L}^{N}) z |\Lambda |$ uniformly in the Debye-H\"uckel limit. The $z|\Lambda |$ is a common volume factor. Divide it out and fix $L \ll 1$. Then the first term in the Mayer expansion dominates the sum of all the others uniformly in the Debye-H\"uckel limit. Therefore the Debye-H\"uckel limit can be taken term by term under the sum in the Mayer expansion and under the $d\mu $ integral by standard theorems of analysis (dominated convergence). Every term becomes a quadratic polynomial in $\phi$ and only the first term in the Mayer expansion contributes in the limit $L \rightarrow 0$. To see this in more detail, note that the first term is \begin{equation}\label{2.5.6} \sum _{\alpha} \int_{\Lambda} d{\bf r} \, z^{(L)}_{\alpha} e^{-e_{\alpha}\beta ^{1/2}(i\phi + g + \psi) } \end{equation} By (\ref{2.5.2}) terms of order $O(z^{(L)}_{\alpha} \beta ^{3/2})$ vanish in the Debye-H\"uckel limit. The $O(z^{(L)}_{\alpha} \beta^{1/2} )$ term vanishes in the limit by pseudo-neutrality (\ref{2.2.1}). Therefore, in the Debye-H\"uckel limit, the non-vanishing part of $F(i\phi + g + \psi )$ is \begin{eqnarray}\label{2.5.7} && F_{{\rm DH}}(i\phi + g + \psi ) = \int_{\Lambda} d{\bf r} \( \sum _{\alpha } z_{\alpha} + \frac{1}{8\pi } (i\phi + g + \psi )^{2} + \frac{1}{8\pi } V_{\infty,L }(0 )\) \nonumber \\ && \hspace{1in} + R_{2}(i\phi+g+\psi) + R_{0} \end{eqnarray} where $V_{\infty,L }(0 )$ arose from the difference between $z_{\alpha }$ and $z^{(L)}_{\alpha}$; see (\ref{2.3.4a}). $R_{2}$ is a quadratic remainder term \begin{equation}\label{2.5.7a} R_{2}(f) = \int \int f({\bf r} _{1}) r({\bf r} _{1},{\bf r} _{2})f({\bf r}_{2})d{\bf r} _{1}d{\bf r} _{2} \end{equation} $R_{0}$ is independent of $f = i\phi + g + \psi $. $R_{0}$ and $R_{2}$ account for all the remaining terms in the Mayer expansion. There is no term linear in $f$: inspection of $O(f)$ contributions from the Mayer expansion is required to verify that the pseudo-neutrality condition makes them all vanish in the Debye-H\"uckel limit. Furthermore $R_{0}$ and $R_{2}$ are independent of $z, \beta $ and vanishing as $L \rightarrow 0$. If we substitute $F_{{\rm DH}}$ back into (\ref{2.5.4}) and regroup into terms independent of $\phi$, terms linear in $\phi$ and terms quadratic in $\phi$, the result is \begin{eqnarray}\label{2.5.8} \Xi_{\Lambda }/\Xi_{{\rm ideal},\Lambda } &\sim & e^{-S_{{\rm DH}}(\psi , g)} \int d\mu_{L } (\phi) \, e^{ - \frac{1}{8\pi } \int _{\Lambda } \(\phi^{2} - V_{\infty,L }(0)\) d{\bf r}} \ e^{-{S'}_{{\rm DH}}(\psi , g;i\phi)+R_{2}(i\phi)} \end{eqnarray} Note that the activity in the ideal gas $\Xi_{{\rm ideal},\Lambda }$ is $z$, whereas a naive calculation based on (\ref{2.3.4c}) might draw us into using $z^{(\lambda )}$. The $\sim $ means that we keep only the terms that are non-vanishing in the Debye-H\"uckel limit. We have defined the quadratic functional \begin{equation}\label{2.5.13} - S_{{\rm DH}}(\psi, g) = \frac{1}{2} \int g V_{\infty,L}^{-1} g \, d{\bf r} + \frac{1}{8\pi } \int (g + \psi )^{2} \, d{\bf r} + R_{2}(\psi +g) + R_{0} \end{equation} whose first variation in direction $\phi$ is \begin{equation}\label{2.5.13b} -{S'}_{{\rm DH}}(\psi , g;\phi) = \int \phi V_{\infty,L}^{-1} g \, d{\bf r} + \frac{1}{4\pi } \int \phi (g + \psi ) \, d{\bf r} + R_{2}'(\psi +g;\phi) \end{equation} $R_{2}'(\psi +g;\phi)$ is the first variation of $R_{2}(\psi +g)$ in direction $\phi$. It is natural to choose $g$ to make $S_{{\rm DH}}'$ vanish, but instead we settle for making it almost vanish by choosing $g$ to solve \begin{equation}\label{2.5.5} -\Delta g + g + \psi =0 \end{equation} because then by (\ref{2.3.3b}) \begin{equation}\label{2.5.13c} -{S'}_{{\rm DH}}(\psi , g;\phi) = \frac{1}{4\pi } L^{2} \int \phi \Delta ^{2} g \, d{\bf r} + R_{2}'(\psi +g; \phi) \end{equation} which means that ${S'}_{{\rm DH}}(\psi , g;\phi)$ disappears at the end when we take $L \rightarrow 0$. However we will keep these terms around for a few more lines because we want to derive some formulas that will be used in later sections. Define \begin{equation}\label{2.5.14} \Xi_{{\rm DH},\Lambda } = \int d\mu_{L } (\phi) \, e^{- \frac{1}{8\pi } \int _{\Lambda } \( \phi^{2} - V_{\infty,L }(0)\) d{\bf r} + R_{2}(i\phi)} \end{equation} which is the normalization for a new Gaussian probability measure \begin{equation}\label{2.5.14b} d\mu _{{\rm DH}}(\phi) = \Xi_{{\rm DH},\Lambda }^{-1} d\mu _{L}(\phi)\, e^{- \frac{1}{8\pi } \int _{\Lambda } \( \phi^{2} - V_{\infty,L }(0)\) d{\bf r} + R_{2}(i\phi)} \end{equation} Then (\ref{2.5.8}) is the same as \begin{eqnarray}\label{2.5.12} \Xi_{\Lambda }/\Xi_{{\rm ideal},\Lambda } &\sim & e^{-S_{{\rm DH}}(\psi , g)} \ \Xi_{{\rm DH},\Lambda } \ \int d\mu _{{\rm DH}}(\phi) \, e^{-{S'}_{{\rm DH}}(\psi , g; i\phi)} \end{eqnarray} As we have explained above, when $L \rightarrow 0$, we can drop the ${S'}_{{\rm DH}}$ so that the normalized $d\mu _{{\rm DH}}$ also disappears and we are down to the evaluation of $\Xi_{{\rm DH},\Lambda }$. The normalization $\Xi_{{\rm DH},\Lambda }$ can be expressed as a determinant using \begin{equation}\label{2.5.9} \int d\mu _{L}(\phi) \, e^{- \frac{\alpha }{8\pi } \int _{\Lambda } \phi^{2} d{\bf r} } = \det\(I + \frac{\alpha }{4\pi } V_{\infty,L }\)^{-\frac{1}{2}} \end{equation} We have dropped the $R_{2}$ term because (\ref{2.4.1.5c}) implies that \begin{equation}\label{2.5.9a} |R_{2}(i\phi)| \leq O(L^{2})\int_{\Lambda} \phi^{2}d{\bf r} \end{equation} and so it gives no contribution in the limit $L \rightarrow 0$. To obtain this estimate from (\ref{2.4.1.5c}) note that $R_{2}$ involves two variational derivatives with respect to $\phi $. Each derivative brings a factor $\beta ^{1/2}$. These factors combine with $z$ in (\ref{2.4.1.5c}) and become a Debye length $l_{D} = 1$. We want $\alpha = 1$, but we put it there because taking $\alpha =0$ shows that both sides are correctly normalized. $V_{\infty,L }$ is the operator whose kernel is the potential $V_{\infty,L }({\bf r} _{1}- {\bf r} _{2})$. The arguments ${\bf r} _{1},{\bf r}_{2}$ are confined to the finite volume $\Lambda $ either by boundary conditions on the Coulomb potential, if the particles are in a grounded conducting container, or by restriction on the range of integration, if the particles are in an insulating container. It can be shown that the thermodynamic limit washes out the difference, but the easier case is the first; then there are zero boundary conditions on the Laplacian in $V_{\infty,L}$ so that the eigenfunctions are trigonometric functions when $\Lambda $ is a box. Using $\det (I + X) = \exp [\sum _{\xi } \ln(1 + \xi )]$, where $\xi $ runs over the eigenvalues of the operator $X$, one can show that the leading term in the infinite volume limit is \begin{equation}\label{2.5.10} \Xi_{{\rm DH},\Lambda } \approx \exp \left[ -\frac{1}{2} (2\pi)^{-3} |\Lambda | \int d {\bf k} \,\( \ln (1 + \frac{1}{4\pi }\tilde{V}_{\infty,L}) - \frac{1}{4\pi }\tilde{V}_{\infty,L} \) \right] \end{equation} where the Fourier transform $\tilde{V}_{\infty,L}$ is given in (\ref{1.1.3bbb}). The limit as $L \rightarrow 0$ of this expression is \begin{equation}\label{2.5.11} \exp \left[ -\frac{1}{2} (2\pi)^{-3} |\Lambda | \int d {\bf k} \,\( \ln (1 + \frac{1}{{\bf k}^{2}}) - \frac{1}{{\bf k}^{2}} \) \right] = \exp [| \Lambda | /(12\pi )] \end{equation} Returning to (\ref{2.5.12}) we have shown that that the leading terms in the Debye-H\"uckel limit followed by infinite volume limit followed by $L \rightarrow 0$ are \begin{equation} \label{2.5.16} \ln \left( \Xi_{\Lambda }/\Xi_{{\rm ideal},\Lambda } \right) \sim \frac{|\Lambda |}{12 \pi} + \frac{1}{8\pi}\int g (-\Delta ) g d{\bf r} + \frac{1}{8\pi} \int (g + \psi )^{2} d{\bf r} \end{equation} where $g$ satisfies the stationarity condition (\ref{2.5.5}). When $\psi =0$ we recover the thermodynamic part of the Debye-H\"uckel law, (\ref{2.1.2}). Furthermore the stationarity condition coincides with the linearized Debye-H\"uckel equation (\ref{2.1.10}) in units where $l_{D} = 1$ when \begin{equation}\label{2.5.15} f = \beta ^{-1/2}(g + \psi), \ \psi ({\bf r} ) = e_{0}\beta ^{1/2} V({\bf r} ) \end{equation} The strange factors of $\beta ^{1/2}$ appear because of the way we normalized the Sine-Gordon transformation. The Debye-H\"uckel result (\ref{2.1.2}) on correlations, in units where $l_{D}=1$, is obtained by taking two functional derivatives of the approximation (\ref{2.5.16}) for $\ln\( \Xi_{\Lambda }/\Xi_{{\rm ideal},\Lambda }\)$ with respect to $-\beta ^{1/2}\psi $. This interchange of functional derivatives with Debye-H\"uckel limit might be hard to justify, but convexity helps in the case of the single charge density expectation \cite{Ken83}. There may be inequalities comparing the Debye-H\"uckel limit with the exact system. Some results follow from Jensen's inequality, but they are limited to the unrealistic case of charge-symmetric systems \cite{Ken82}. See also theorem~\ref{theorem-DHupperbound} in the next section. Questions: Are extended Debye-H\"uckel theories such as the ones considered by Fisher et al. limiting theories? One should no longer let $L \rightarrow 0$ and should choose $g$ to be the exact stationary configuration for $S_{{\rm DH}}(\psi , g)$, which contains quadratic remainder terms $R_{2}$ giving corrections to the dielectric constant. Also the activity may be renormalized, whereas in our case $L \rightarrow 0$ causes $z^{(L)}$ to coincide with $z$. Is the expansion \cite{LeSt68} asymptotic in $\kappa a$? \subsection{Charge expulsion} \label{sec-charge-expulsion} We continue to use units in which $l_{D} = 1$ introduced in section~\ref{sec-debye-huckel2}. In our preliminary discussion of the Debye-H\"uckel theory in section~\ref{sec-debye-huckel} we claimed that the activities should be constrained by the pseudo-neutrality condition (\ref{2.2.1}). If we start with a system that does not satisfy this condition, then even when $\psi =0$, the action $S(0,i\phi)$ defined in (\ref{2.5.3b}) is not stationary in $\phi$ at $\phi = 0$. However, a translation $\phi \rightarrow \phi - ig$ such that $S(0,g + i\phi)$ is stationary in $\phi$ at $\phi = 0$ reveals the physical effect of charge expulsion and an attendant renormalization of activities. This is a nonlinear effect, not small enough to fit within the approximation of the last section. The physical interpretation is that excess charge is expelled to the boundary where it forms a boundary layer of total charge proportional to the surface area. This layer modifies the electrostatic potential in the interior in such a way that the activities are renormalized to satisfy pseudo-neutrality, which is why the excess charge is of order surface area and not volume in this grand canonical ensemble. For simplicity we omit the complications that arose in the last section due to the length scale $\lambda $ by considering a system with interaction $V_{\infty ,L }$ with insulating boundary conditions and no other forces. In this case we can use the Sine-Gordon representation (\ref{2.3.4c}) to represent the entire interaction. We can also apply the translation formula (\ref{2.3.5}) for $\phi \rightarrow \phi - ig$. The result is \begin{eqnarray}\label{2.6.1} && \Xi_L (\beta ^{1/2}\psi ) = \int d\mu _{L }(\phi) \, e^{\frac{1}{2}\int g V_{\infty,L }^{-1} g d{\bf r} + \int i\phi V_{\infty,L }^{-1} g d{\bf r} } \nonumber \\ && \hspace{1in} \times \Xi_{{\rm ideal},z^{(L)}} (\beta ^{1/2}[i\phi + g + \psi]). \end{eqnarray} The translation $g$ is chosen so that the linear term $\int \phi V_{\infty,L }^{-1} g d{\bf r}$ is canceled by a corresponding term from $\Xi_{{\rm ideal},z^{(L)}}$. This is the same as choosing $g$ to be the minimizer $g$ of \begin{equation}\label{S-infty-lambda} - S_{\infty, L }(\psi ,g) = \frac{1}{2 } \int g V_{\infty,L }^{-1} g d{\bf r} \, + \ln \Xi_{{\rm ideal},z^{(L)}} (\beta ^{1/2}[g + \psi]) \end{equation} Insulating boundary conditions mean that $V_{\infty,L }^{-1}$ is given by (\ref{2.3.3b}). The $\int d{\bf r} $ integral extends over infinite volume but the $\Xi_{{\rm ideal},z^{(L)}}$ is an ideal gas in the finite volume $\Lambda $. There exists \cite{Ken84} a unique $g$ that minimizes this functional. Uniqueness is an easy consequence of it being convex. Let $\psi = 0$. Then in the deep interior of $\Lambda $ $g$ is close to the constant $g_{0}$ that minimizes \begin{equation}\label{2.6.2a} \ln \Xi_{{\rm ideal},z^{(L)}} (\beta ^{1/2} g_{0}) = \int_{\Lambda } d{\cal E} \, z^{(L)}({\cal E} ) e^{-\beta ^{1/2} e_{\alpha } g_{0}} \end{equation} Outside $\Lambda $, $g$ solves $V_{\infty,L }^{-1}g = 0$ and $g \rightarrow 0$ at infinity. Define renormalized activities $z^{(L,R)}$ by \begin{equation} z_{\alpha }^{(L,R)} = \exp(- e_{\alpha } \beta ^{1/2} g_{0}) z_{\alpha }^{(L)} \end{equation} and note that \begin{equation} \Xi_{{\rm ideal},z^{(L)}} (\beta ^{1/2} f) = \Xi_{{\rm ideal},z^{(L,R)}} (\beta ^{1/2} [f-g_{0}]) \end{equation} Since $g \approx g_{0}$ inside $\Lambda $ the bulk contribution to $S_{\infty ,L}(g)$ comes from the ideal gas term which has the renormalized activities $z^{(L,R)}$. But since $\Delta g$ is non-zero near the boundary there is a boundary contribution to the pressure in $\frac{1}{2 } \int g V_{\infty,L }^{-1} g d{\bf r} $ and also in the ideal gas term because $g \not = g_{0}$ at the boundary. These arise from the surface charge expelled from the interior and account for the capacity term in the free energy, discussed in the analysis of systems with net charge \cite[page 55]{LiLe72}, \cite{GrSc95b}\footnote{If $L=0$ and $g$ is chosen to be constant in the interior of $\Lambda $ and harmonic in the exterior, then the boundary contribution in (\ref{S-infty-lambda}) exactly equals the electrostatic energy stored in a perfect conductor with shape $\Lambda$.}. From (\ref{2.6.1}) \begin{eqnarray} \Xi_L (0) &=& e^{-S_{\infty ,L}(g)} \int d\mu _{L }(\phi) \, e^{\int i\phi V_{\infty,L }^{-1} g d{\bf r} } \ \frac{ \Xi_{{\rm ideal},z^{(L,R)}}(\beta ^{1/2}[i\phi + g - g_{0}]) }{ \Xi_{{\rm ideal},z^{(L,R)}}(\beta ^{1/2}[g - g_{0}]) }\nonumber \end{eqnarray} By our choice of translation $g$ we have arranged that the $i\phi$ fluctuates around zero so that $\exp \left(-S_{\infty ,L}(g) \right)$ should be the leading term in the Debye-H\"uckel limit. The next term in the approximation would be a Debye-H\"uckel term that comes from integrating over the $\phi$ fluctuations using a quadratic approximation as in section~\ref{sec-debye-huckel2}. \subsection{Symmetries and tunneling}\label{sec-tunneling} We continue to work in units in which $l_{D} = 1$. There is a loss of intuition entailed by the Sine-Gordon transformation because the integration extends over imaginary potentials, but there are also aspects one can see more easily in the Sine-Gordon language. In particular if all charges $e_{\alpha }$ are integral multiples of a fundamental unit of charge, which for simplicity we assume to be one, then the action $S(\psi,g)$ in (\ref{2.5.3b}) is invariant under the discrete symmetry \begin{equation}\label{2.7.1} S(\psi,g) = S(\psi,g + i \delta h), \ \ \ \delta h = 2\pi \beta^{-1/2} \end{equation} This section is concerned with the breaking of this symmetry, which plays a role in the proofs of screening. We will see also that there are tunneling corrections associated with the symmetry. Let us first note that the Coulomb system in an insulating container in three or more dimensions cannot enjoy this symmetry because the Gaussian measure $d\mu _{L}(\phi)$ satisfies \begin{equation}\label{2.7.1b} \int d\mu_{L} (\phi) \, \phi^{2}(0) = V_{\infty,L}(0) < \infty \end{equation} precluding any symmetry of the form $\phi \rightarrow \phi + c$. We have seen in section~\ref{sec-debye-huckel2} that the Debye-H\"uckel approximation is a quadratic approximation around the stationary configuration $\phi = 0$. The symmetry (\ref{2.7.1}) tells us that there are infinitely many stationary configurations related by the symmetry to $\phi = 0$. We claim that even though (\ref{2.7.1b}) implies that the stationary configuration $\phi= 0$ is favored, in the thermodynamic limit there are arbitrarily large regions where $\phi$ is trapped in the other stationary configurations. This is analogous to the Ising model at low temperature in which there occur arbitrarily large islands of $-$ spins even though the boundary is set at $+$ and the majority of spins are $+$. In the zero temperature limit in the Ising model, the islands of $-$ spins disappear, but if one wants to understand low temperature as opposed to zero temperature, then they are important. In the same way, to obtain screening near, as opposed to in the Debye H\"uckel limit, the presence of large islands where the $\phi = 0$ approximation breaks down has to be taken into account. To take into account the presence of many stationary configurations related by the symmetry we use a functional version of the Villain approximation \begin{equation}\label{2.7.1a} e^{\cos (x)} \approx \sum _{n} e^{-\frac{1}{2}(x-2\pi n)^{2}+1} \end{equation} on the term $\exp(F)$ in (\ref{2.5.3}). In (\ref{A.villain}) in section~\ref{sec-polymer-rep1} we explain in terms of this one dimensional analogue why taking into account the extra wells in this way improves the convergence of perturbation theory. $\exp \left(F \right)$ is invariant under a translation $\phi \rightarrow \phi + h$ where $h = h({\bf r} )$ is any function of ${\bf r} $ that takes values in the set $\{n\delta h\}$ of periods, because $\Xi_{L , \Lambda, z^{(L)}}$ clearly has this property. Also from (\ref{2.5.7}) it has the quadratic approximation \begin{eqnarray}\label{2.7.2} F_{{\rm DH}}(\psi + i\phi) &=& \int_{\Lambda} d{\bf r} \( \sum _{\alpha } z_{\alpha} + \frac{1}{8\pi } (i\phi + \psi )^{2} + \frac{1}{8\pi } V_{\infty,L }(0 )\) \nonumber \\ &&\hspace{.5in} + R_{2}(i\phi + \psi ) + R_{0} \end{eqnarray} Thus an analogue of (\ref{2.7.1a}) is \begin{equation}\label{2.7.villain} e^{ F(\psi + i\phi)} = e^{R_{3}(\psi + i\phi)} \sum _{h} e^{F_{{\rm DH}}(\psi - ih + i\phi)} \end{equation} where $R_{3}$ is an $O(\beta ^{1/2})$ error term representing cubic corrections to each well and put there to obtain exact equality. $h = h({\bf r} )$ is summed over all functions that vanish outside $\Lambda $, take values in the set $\{n\delta h\}$ of periods and are piecewise constant on a lattice of unit cubes filling $\Lambda $. If we think of a typical piecewise constant function $h({\bf r})$ as a height, then the resulting landscape splits into plateaus where $h({\bf r} )$ is constant but more important are the jumps from one plateau to another. In analogy with the Ising model we call the connected surfaces in $\Lambda $ where $h$ jumps ``contours.'' Our objective is to argue that near the Debye-H\"uckel limit the sum over $h$ is dominated by terms in which these contours are small and very dilute. The value of $h$ in a given plateau fixes the well in which $\phi $ is trapped and the contour is where $\phi $ tunnels from one well to another. We substitute (\ref{2.7.villain}) into the Sine-Gordon transformation (\ref{2.5.3}) and retrace the analysis of section~\ref{sec-debye-huckel2} leading up to the Debye-H\"uckel approximation (\ref{2.5.12}) with $\psi $ replaced by $\psi - ih$ and $g$ replaced by $ig$ \begin{eqnarray}\label{2.7.6} &&\hspace{-.5in} \Xi_{\Lambda } /\Xi_{{\rm ideal},\Lambda } = \Xi_{{\rm DH},\Lambda }\sum _{h} e^{- S_{{\rm DH}}(\psi - ih, ig)} \nonumber \\ && \times \int d\mu _{{\rm DH}}(\phi) \, e^{-{S'}_{{\rm DH}}(\psi -ih , ig; i\phi) + R_{3}(\psi+ig+i\phi)} \end{eqnarray} For each $h$ in the sum $g({\bf r})$ can be any function. We choose it to make $S_{{\rm DH}}(\psi -ih, ig)$ stationary so that ${S'}_{{\rm DH}}$ vanishes. For the moment we drop the term $R_{3}$, then the $d\mu _{{\rm DH}}$ integral drops out because it is normalized and we are left with an approximation \begin{equation}\label{2.7.6a} \Xi_{\Lambda } /\Xi_{{\rm ideal},\Lambda } \approx \Xi_{{\rm DH},\Lambda } \sum _{h} e^{- S_{{\rm DH}}(\psi - ih, ig)} \end{equation} where \begin{eqnarray} \label{2.7.Sdef} S_{{\rm DH}}(\psi - ih, ig) &=& \frac{1}{2 }\int g V_{\infty,L}^{-1} g d{\bf r} + \frac{1}{8\pi }\int (g - h -i\psi )^{2} d{\bf r} \nonumber \\ &&\hspace{1in} + R_{2} + R_{0} \end{eqnarray} When $\psi =0$ $g$ minimizes $S_{{\rm DH}}(-ih, ig)$ by trying to make $(g-h)^{2}$ vanish but cannot quite succeed because $\int g V_{\infty,L}^{-1} g d{\bf r}$ forces $g$ to be smooth whereas $h$ has jumps. Suppose $h({\bf r} ) = h_{\Gamma }({\bf r} )$ has only one contour $\Gamma $ and $h$ vanishes at the boundary of $\Lambda $. Then $\Gamma $ has an exterior ``sea-level'' plateau where $h = 0$ and an interior plateau where $h \not =0$. In the interiors of both plateaus $g({\bf r} ) \approx h({\bf r} ) $. The error in the $\approx$ is $O(\exp \left( - {\rm dist}({\bf r} ,\Gamma \right))$. Any $h$ with $h = 0$ at the boundary can be decomposed in a unique way \begin{equation}\label{2.7.hdecomp} h = \sum _{j} h_{j} \end{equation} where each $h_{j}$ has only one contour $\Gamma _{j}$ and $h_{j}$ vanishes at the boundary. Thus $h_{j}$ is defined on the whole box $\Lambda $ but it assumes just two values: zero outside the contour $\Gamma _{j}$ and $h_{j}({\rm int})$ at all points in the interior of the contour. Across the contour $\Gamma _{j}$ $h_{j}$ jumps by $h_{j}({\rm int})$ which is the same as the jump of $h$ across $\Gamma _{j}$. The minimizers are linear functions of $h$ so $g = \sum g_{j}$ where $g_{j}$ is the minimizer for the single contour $h_{j}$. If the contours are well separated then by $g({\bf r} ) = h({\bf r} ) + O(\exp \left({\rm dist}({\bf r} ,\Gamma \right))$. \begin{equation}\label{2.7.Sdecomp} S_{{\rm DH}}(\psi - ih, ig) \approx \sum _{j} S_{{\rm DH}}(\psi - ih_{j}, ig_{j}) \end{equation} Thus \begin{equation}\label{2.7.ising} \Xi_{\Lambda } /\Xi_{{\rm ideal},\Lambda } \approx \Xi_{{\rm DH},\Lambda } \sum _{h} \prod _{j} e^{- S_{{\rm DH}}(\psi - ih_{j}, ig_{j})} \end{equation} The approximation in these two equations was to leave out exponentially small cross terms in $S_{{\rm DH}}$ involving $(g_{i} -h_{i})(g_{j} - h_{j})$ and derivatives of $g_{i}$ times derivatives of $ g_{j}$. The sum over $h$ is equivalent to summing over the number $N$ of contours, the shapes $\Gamma _{j}$ of the contours and the jumps $h_{j}({\rm int})$ at each contour so that \begin{equation}\label{2.7.contour} \Xi_{\Lambda } /\Xi_{{\rm ideal},\Lambda } \approx \Xi_{{\rm DH},\Lambda } \sum _{N} \frac{1}{N!} \sum _{\Gamma _{1},\dots ,\Gamma _{N}} \prod _{j} \sum _{h_{j}({\rm int})} e^{- S_{{\rm DH}}(\psi - ih_{j}, ig_{j})} \end{equation} The exponential of $S_{{\rm DH}}(\psi - ih_{j}, ig_{j}) $ suppresses the jump $h_{j}({\rm int})$ at $\Gamma _{j}$ by a factor \begin{equation}\label{2.7.suppress} \exp[ - O(|\Gamma_{j} |) h_{j}({\rm int})^{2}] \end{equation} where $|\Gamma_{j} |$ is the area of the contour. The $h_{j}({\rm int})^{2}$ dependence is because $S_{{\rm DH}}$ is quadratic in $g,h$ when $\psi =0$, $g_{j}$ is linear in $h_{j}$ and $h_{j}$ is just $h_{j}({\rm int}) \times $ a unit jump configuration. In (\ref{2.7.contour}) the term with no contours, $N = 0$, is our original Debye-H\"uckel approximation (\ref{2.5.16}). From (\ref{2.7.contour},\ref{2.7.suppress}) we see that the contours should behave like an ideal gas with an exponentially tiny density near the Debye-H\"uckel limit. Nevertheless they represent a phenomenon that probably makes the standard perturbation series asymptotic but not convergent and incapable of proving that there is screening near as opposed to at the limit. In contrast (\ref{2.7.contour}) is the basis for a convergent expansion which does imply screening near the Debye-H\"uckel limit. The terms we left out in arriving at this representation are exponentially small interactions between the dilute contours. There are also constraints (hard core interactions) in the sum that forbid intersections of contours. Such interactions are within the purview of conventional convergent Mayer expansions, which is why it is possible to develop a convergent expansion. However this expansion is clumsy in comparison with standard perturbation theory. One has to show for example that the $R_{3}$ error term does not destroy the picture we have just developed. One of the principal accomplishments \cite{GlJa87} of the constructive quantum field theory program was to develop a calculus for the analysis of ``almost Gaussian'' integrals and this is what is used at this point. The reader who suffers from residual curiosity can turn to chapter~\ref{chapter-convergent-expansions} for a compressed and updated account of this calculus and its applications, or to \cite{BrFe81} for an older review. Note the role of grounded container boundary conditions is to anchor $g$ and therefore $h$ at zero on the boundary. The $1/r$ potential with no boundary conditions, i.e. particles in an insulating container, is more difficult than the case of grounded boundary because the anchoring is weaker, since it comes from (\ref{2.7.1b}). This has been partially investigated \cite{FeKe85}. Their argument is hard because there are long range forces at the boundary which make it difficult to control the size of boundary effects. In the case of two dimensions with insulating boundary conditions, there is no anchoring at the boundary because the Gaussian measure\footnote{defined by taking a limit outside the integral as a small mass tends to zero} $d\mu _{L}$ does not break the symmetry (\ref{2.7.1}). We mentioned in section~\ref{sec-Debye-screening} a result of Fr\"ohlich and Spencer \cite[Theorem 4.1]{FrSp81b} that fractional charges are not screened when insulating boundary conditions are imposed. This is because of this symmetry. Intuitively there is no preference for which potential well $\phi({\bf r} _{1})$ is trapped in, but once it has made up its mind, $\phi({\bf r} _{2})$ wants to choose the same well, which is a long range correlation. This argument predicts that the correlation decays to a constant. On the other hand, if the boundary is grounded, then the $0$ well is selected essentially everywhere, except for a very dilute gas of contours that enclose regions where $\phi$ is in other wells. In this case all local observables are expected to have exponential decay, but this is only proved for integer charge observables \cite{Yan87}. Recall in analogy that the Ising model at low temperature with plus boundary conditions has exponential decay but this decay is lost if the boundary spins are not tied down because the state is no longer pure. It is an open problem to prove that two dimensional systems with insulating boundary conditions have screening of integer charge observables. Returning to three dimensions when there are irrational charges the symmetry is replaced by a quasi-periodicity, nevertheless there is screening in these systems without exact symmetry \cite{Imb83}. The argument requires Pirogov-Sinai theory. There is an intriguing difference between fractional charge observables and integer charge observables. Suppose the gas consists of unit positive and negative charges but there is an additional half-integer charge fixed at the origin ${\bf r} = 0$, represented by a factor $\exp \left(-\beta ^{1/2}i\phi (0)/2 \right)$. The $\phi(0) $ fluctuates around $g(0) \approx h(0)$, so there is a factor $\exp \left(-\beta ^{1/2}ih(0)/2 \right)$ which is $\pm 1$ in a way that depends on all contours surrounding the origin, which is a more non-local (topological) effect than when an integer charge is placed at the origin. To put it another way: a phase factor $\exp \left(-\beta ^{1/2}ih_{j}(0)/2 \right)$ must be included in the sum over $h_{j}({\rm int})$ in (\ref{2.7.contour}) whenever $\Gamma _{j}$ encloses the origin. This does not destroy screening in three dimensions, but it might be worth further study. We will show, in appendix \ref{sec-polymer-rep1}, that the tunneling corrections are of size $\exp {( - O (\beta /l_{D}))}$ so they are not of any consequence in the Debye-H\"uckel asymptotic regime. However, they should be important if $\beta /l_{D}$ is not particularly small because then the period is small and the potential wells that trap $\phi$ are close together and shallow. There could be a phase transition in which the imaginary potential $i\phi$ ceases to be well localized in wells. If as a consequence there are more fluctuations in the typical $\phi$, then the gas will be organized into more tightly bound neutral aggregates because when the position ${\bf r} $ of a cluster or particle with charge $e_{\alpha }$ is integrated, the rapidly varying phase factor $\exp (-ie_{\alpha }\beta^{1/2} \phi({\bf r} ))$ will give cancellations. \footnote{We thank James Glimm for conversations on this point.} Note that the Kosterlitz-Thouless phase transition fits this description. Tunneling represents a phenomenon which arises because the system consists of discrete charges as opposed to infinitely divisible charge distributions, which would give a single well as in the Debye-H\"uckel approximation. \newpage \section{Dipoles} \setcounter{page}{1} \label{chapter- dipole} The Coulomb plasma is the main focus of this review. However in chapters~\ref{chapter-4} and \ref{chapter-loops} we will find that a system of point quantum charges has a close analogy to a certain system of dipoles and higher order multipoles. The resulting multipole forces are at the origin of a breakdown of exponential screening in quantum mechanics. It is therefore of interest to give here a short review concerning the lack of screening in classical dipole systems and estimates on the dielectric constant. Concerning the analogy an important caveat should be kept in mind. In a classical system exponential screening is restored as soon as free charges are added to classical dipoles. At any non-zero temperature, the quantum gas always has a proportion of free charges due to ionization processes, but these free \emph{quantum} charges are not able to restore exponential screening. The analogies and differences between the behavior of quantum point charges and classical dipoles are made more precise in section~\ref{sec-4.6} \subsection{The Dipole ensemble} \label{sec- DipoleDescription} A dipole is specified by the coordinate ${\cal E} = ({\bf r}, {\bf d})$ that unites the position ${\bf r} $ and the dipole moment ${\bf d}$. We suppose that in the absence of interaction a single dipole experiences no preferred directions so that \begin{equation}\label{3.1.1} \int d{\cal E} \dots = \int _{\Lambda } d{\bf r} \, \int d\Omega \dots \mbox{ with } \int d\Omega = 1, \end{equation} where $d\Omega$ is surface measure on the sphere of unit vectors $\hat{{\bf d}}$ and ${\bf d} = d\hat{{\bf d}}$ with $d$ a fixed dipole moment. The interaction between a pair of dipoles is \begin{equation}\label{3.1.2} V({\cal E} _{1}, {\cal E} _{2} ) = ({\bf d}_{1}\cdot\nabla_{1}) \ ({\bf d}_{2}\cdot\nabla_{2}) \ V({\bf r} _{1} - {\bf r} _{2}) \end{equation} where $\nabla_{i}$ is the gradient operator that acts on ${\bf r} _{i}$. $V({\bf r} _{1} - {\bf r} _{2})$ should be the Coulomb potential between two unit charges at ${\bf r} _{1}, {\bf r} _{2}$, but the singularity of the Coulomb potential at short distance will lead to instability for the dipole system. We choose a length scale $L$ and a form factor $F$ and set \begin{equation}\label{3.1.2a} \tilde{V}({\bf k} ) = (4\pi )|{\bf k}|^{-2}\tilde{F}^{2}(L^{2}k^{2}) \end{equation} The form factor satisfies $\int F ({\bf r}) d{\bf r} = 1$ and means that point charges are replaced by ``charge clouds.'' This should be regarded as an effective interaction which arises by integrating out forces on length scales less than $L$ in a more realistic description of the short range physics. An effective action of this form is not appropriate unless the gas is dilute so that interactions such as hard cores are not playing a strong role. See for example the use of a Mayer expansion as in section \ref{sec-debyesphere} which led to such an effective interaction. The partition function is \begin{equation}\label{3.1.3} \Xi_{\Lambda } = \sum _{N} \frac{z^{N}}{N!} \int d{\cal E} _{1} \, \dots \int d{\cal E} _{N} \, e^{-\beta U({\cal E} _{1},\dots ,{\cal E} _{N}) }, \end{equation} where the interaction {\it includes} self-energies, \begin{equation}\label{3.1.4} U({\cal E} _{1},\dots ,{\cal E} _{N}) = \frac{1}{2}\sum _{i,j = 1}^{N} V({\cal E} _{i}, {\cal E} _{j} ). \end{equation} If instead the interaction is $\sum f({\cal E} _{j})$ where $f$ is an external field, then we obtain an ideal gas partition function \begin{eqnarray}\label{3.1.3b} \Xi_{{\rm ideal},\Lambda}(f) &=& \sum _{N} \frac{z^{N}}{N!} \int d{\cal E} _{1} \, \dots \int d{\cal E} _{N} \, \exp\( - \sum _{j=1}^{N} f({\cal E} _{j}) \) \nonumber \\ &=& \exp \left(z\int d{\cal E} \,e^{-f({\cal E} )} \right)\nonumber \end{eqnarray} As in section~\ref{sec-sinegordon} we can reconstruct the interaction between the dipoles by integrating over an imaginary external field, $f({\cal E} ) = \beta ^{1/2}{\bf d}\cdot\nabla i\phi({\bf r})$, with respect to the measure $d\mu _{L}$ discussed in (\ref{2.3.3}): \begin{equation}\label{3.1.6} \Xi_\Lambda = \int d\mu _{L} (\phi) \, \Xi_{{\rm ideal},\Lambda} (\beta^{1/2}{\bf d}\cdot\nabla i\phi ). \end{equation} Formally this is $\int {\cal D} \phi \exp \left(-S(i\phi ) \right)$ where we define the action $S$ by \begin{eqnarray}\label{3.1.7} - S(g) &=& \frac{1}{2}\int g V^{-1}g + z \int d{\cal E} \, \exp (-\beta ^{1/2} {\bf d}\cdot\nabla g) \nonumber \\ &=& \frac{1}{2}\int g V^{-1} g + z \int d{\cal E} \, \cosh (\beta ^{1/2} {\bf d}\cdot\nabla g) \end{eqnarray} where the inverse operator $V^{-1}$ has a kernel whose Fourier transform is $\tilde{V}^{-1}({\bf k} )$. We shall consider the observable that measures the density of dipoles with coordinate ${\cal E} $ in a configuration $\omega = ({\cal E} _{1},\dots ,{\cal E} _{N})$, namely \begin{equation}\label{3.1.5} \hat{\rho }({\cal E} ,\omega ) = \sum _{i=1}^{N} \delta ({\cal E} , {\cal E} _{j}), \end{equation} Distribution functions such as the two point function $\langle \hat{c}({\cal E}_{a}) \hat{c}({\cal E}_{b}) \rangle$ are defined in analogy to section~\ref{sec- classical}. Distributions at non-coincident points of $\hat{c}({\cal E} ,\omega )$, where ${\cal E} = ({\bf r} , {\bf d} )$, become expectations of $z\exp (-i\beta^{1/2} {\bf d} \cdot\nabla \phi({\bf r} ) )$ in the Sine-Gordon language . Another useful quantity is \begin{equation}\label{3.1.5a} \langle e^{-\beta ^{1/2}\int i\phi f}\rangle = \Xi_\Lambda ^{-1} \int d\mu _{L} (\phi) \, \Xi_{{\rm ideal},\Lambda} (i\beta^{1/2}{\bf d}\cdot\nabla \phi ({\bf r})) e^{-i\beta ^{1/2}\int \phi f} \end{equation} which measures ($\beta \times$)energy of a charge distribution $f$ in the sea of dipoles. By unraveling the Sine-Gordon transformation one finds that the charge distribution $f$ has self-energy determined by the potential $V({\bf r} _{1} - {\bf r} _{2})$ corresponding to (\ref{3.1.2a}) and it interacts with system dipoles according to potential energy $({\bf d}_{2}\cdot\nabla_{2}) \ V({\bf r} _{1} - {\bf r} _{2})$. We are using the expected value $\langle \ \rangle $ to indicate either the finite volume or infinite volume limit(s). This infinite volume limit is known to exist for several choices of boundary conditions \cite{FrPa80,FrSp81b,FuSp97}. We shall also consider expectations of $\phi$. The two-point function measures the effective potential between two infinitesimal charges: as in (\ref{1.1.eff1}), but in the Sine-Gordon language, $\beta \times $ the effective potential between two infinitesimal test charges $({\bf r} _{a}, e_{a})_{a= 1,2}$ is given by \begin{equation}\label{3.1.8} e^{- \beta e_{1}e_{2}V^{{\rm eff}}({\bf r} _{1}-{\bf r} _{2})} = \frac{ \langle e^{-ie_{1}\beta^{1/2} \phi({\bf r} _{1})} e^{-ie_{2}\beta^{1/2} \phi({\bf r} _{2})} \rangle }{ \langle e^{-ie_{1}\beta^{1/2}\phi({\bf r} _{1})} \rangle \langle e^{-ie_{2}\beta^{1/2}\phi({\bf r} _{2})} \rangle } \end{equation} The numerator removes the self-energies of the test charges. Since $e_{1}, e_{2} \rightarrow 0$ we have \begin{equation}\label{3.1.9} V^{{\rm eff}}({\bf r} _{1}-{\bf r} _{2}) = \langle\phi({\bf r} _{1})\phi({\bf r} _{2})\rangle \end{equation} Furthermore the dielectric constant $\varepsilon$ is obtained from the large ${\bf r} $ asymptotics, \begin{equation}\label{3.1.10} V^{{\rm eff}}({\bf r} ) \sim \frac{1}{\varepsilon |{\bf r} |} \end{equation} \subsection{A no-screening theorem for dipoles}\label{sec- noscreening} The action $S(g)$ in (\ref{3.1.7}) is invariant under $g \rightarrow g + c$, where $c$ is any constant. This is a continuous symmetry and the action is local. Locality and continuous symmetry generally lead to power law correlations because fluctuations, $\phi \rightarrow \phi + f$, where $f({\bf r} )$ is slowly varying as a function of ${\bf r} $, are very weakly suppressed\footnote{The Coulomb system possessed a similar symmetry, but with the essential difference that $c$ could only have discrete values.}. No-screening theorems for dipoles have been obtained \cite{Par79}. Subsequently \cite{FrSp81b} gave the proof we present. This no-screening theorem is a little remarkable because it is valid for all activities and all temperatures, but its validity at high density is dependent on the way the short distance singularity in the dipole force is stabilized. If there are hard cores, then crystalline phases are likely. This is proven \cite{FrSp81b} for dipoles on a lattice. For models where the Coulomb potential is smoothed out such as our $V_{\infty ,\lambda }$ potential Fr\"ohlich and Spencer proved further results to the effect that the two point function has no oscillations, suggesting that this system is always in a fluid phase. \begin{theorem}\label{theorem- noscreening} For all activities $z \geq 0$ and all inverse temperatures $\beta \geq 0$ there is no screening in the sense that the two point function has a singularity at ${\bf k} =0$ in its Fourier transform and \begin{equation}\label{3.2.1} \tilde{V}^{{\rm eff}}({\bf k} ) \geq \tilde{{\cal V}}({\bf k} ), \end{equation} where \begin{equation}\label{3.2.2} \tilde{{\cal V}}({\bf k} ) = \(\tilde{V}^{-1}(k) + \frac{z\beta d^{2}}{3} {\bf k}^{2}\)^{-1} \end{equation} is the Fourier transform of the dipole potential with a dielectric correction. \end{theorem} $\tilde{{\cal V}}({\bf k} )$ is the Fourier transform of the kernel ${\cal V} ({\bf r} - {\bf r} ')$ which is the inverse of the Hessian of $S$ at $\phi =0$. A version of this theorem holds in any dimension and also for a large class of positive-definite interactions in place of dipole-dipole. The factor 3 is the dimension of space. We will remark in the proof the main property of the interaction being used. The reader will see that the argument is a very general mean field theory bound. From this theorem and discussion of the dielectric constant in section~\ref{sec- classical} we immediately obtain a mean field theory bound for the dielectric constant, assuming there is a dielectric constant: \begin{equation}\label{3.2.3} \varepsilon \leq 1 + \frac{z\beta d^{2}}{3} 4\pi \end{equation} There is a large literature on the molecular theory of the dielectric constant, with atoms and molecules modeled by preformed classical permanent dipoles, using various short range regularizations. See, for example, \cite{StPaHo81} and references therein. In particular, the Clausius-Mosotti law $(\varepsilon -1) (\varepsilon +2)^{-1} = 4\pi\rho \beta d^{2}/9$ was derived in \cite{HoSt74,HoSt76} by taking the Kac limit, in which $V({\bf r})$ is replaced by $\gamma ^{3}V(\gamma {\bf r})$ and $\gamma \rightarrow \infty$. However, this law is contradicted by the bound (\ref{3.2.3}) if one just sets $z = \rho$ at low density. The reason is that the result of the Kac limit depends on the treatment of dipoles in the vicinity of the test charge and the scaling of test charges. The literature would benefit from clarification of this point. See the further comment at the end of this section. Theorem~\ref{theorem- noscreening} is a consequence of \begin{theorem}\label{theorem-DHupperbound} (The mean field upper bound) \begin{equation}\label{3.2.4} \langle e^{-\int i\phi f}\rangle \ \leq \exp \left( \inf_{g} - S(g,f) \right) \end{equation} where \begin{equation}\label{3.2.5} - S(g,f) = \frac{1}{2}\int g V^{-1}g - \int gf + z \int d{\cal E} \, \( \cosh (\beta ^{1/2} {\bf d}\cdot\nabla g) - 1\) \end{equation} \end{theorem} \noindent {\bf Proof. }\hspace{2mm} We make a complex translation $\phi \rightarrow \phi - ig$. This is done using the translation formula (\ref{2.3.5}), but we arrive at the same place if we do it in the intuitive formula $\int {\cal D} \phi \, \exp(-S(i\phi) -\int i\phi f)$, which is the numerator of $\langle e^{-\int i\phi f}\rangle$. On expanding out the exponent we obtain an upper bound on $|\exp (\dots )|$ by dropping all the imaginary terms. Thus \begin{eqnarray}\label{3.2.6} &&\hspace{-.3in} -{\rm Re } S(i\phi + g) - {\rm Re } \ \int (i\phi+g)f \nonumber \\ &=& {\rm Re } \frac{1}{2}\int (i\phi+g) V^{-1}(i\phi+g) - \int gf + z {\rm Re } \int d{\cal E} \, \cosh (\beta ^{1/2} {\bf d}\cdot\nabla (i\phi+g))\nonumber \\ &=& -\frac{1}{2}\int \phi V^{-1}\phi + \frac{1}{2}\int g V^{-1}g\nonumber - \int gf + z \int d{\cal E} \, \cos (\beta ^{1/2} {\bf d}\cdot\nabla \phi) \cosh (\beta ^{1/2} {\bf d}\cdot\nabla g)\nonumber \\ &\leq& -S(i\phi) - S(g,f) \end{eqnarray} so that in the resulting bound on $\langle \exp(-\int i\phi f)\rangle$ the $\int {\cal D} \phi \exp \left(-S(i\phi ) \right) $ cancels in numerator and denominator and we obtain the mean field upper bound, since $g$ is arbitrary.\hspace*{\fill} QED \medskip \textbf{Proof of Theorem ~\ref{theorem- noscreening}.} We rewrite the mean field upper bound as \begin{equation}\label{3.2.7} 1 - \langle e^{-i\alpha \int \phi f}\rangle \ \geq 1 - \exp \left( - S(\alpha g,\alpha f) \right) \end{equation} which holds for any function $g$. Let $\alpha \rightarrow 0$ to obtain \begin{equation}\label{3.2.8} \frac{1}{2} \langle(\int \phi f)^{2}\rangle \ \geq -\frac{1}{2}\int g V^{-1}g + \int gf - \frac{1}{2} z \int d{\cal E} \, \( \beta ^{1/2} {\bf d}\cdot\nabla g\)^{2} \end{equation} This is valid provided $g$ has enough decay so that the terms of higher order than $\alpha ^{2}$ are convergent integrals. {\it This and positive-definiteness are the only properties of the dipole interaction used in the proof of (\ref{3.2.1})}. Now choose $g = {\cal V} f$. The result is \begin{equation}\label{3.2.9} \langle(\int \phi f)^{2}\rangle \ \geq \int f {\cal V} f. \end{equation} This is the same as \begin{equation}\label{3.2.10} \int f V ^{{\rm eff}} f \geq \int f {\cal V} f \end{equation} so that \begin{equation}\label{3.2.11} \int \tilde{V}^{{\rm eff}}({\bf k} )|\tilde{f}({\bf k} )|^{2} d{\bf k} \geq \int \tilde{{\cal V}}({\bf k} )|\tilde{f}({\bf k} )|^{2} d{\bf k} \end{equation} which, being true for a large class of $f$, is the result we claimed. \hspace*{\fill} QED \medskip A Kac limit can be defined by replacing $V({\bf r})$ by $\gamma ^{-1}V(\gamma {\bf r})$ and letting $\gamma \rightarrow 0$, which is the same as taking $L \rightarrow \infty $. By scaling one can show that it is equivalent to replacing $S(i\phi)$ by $\gamma ^{-3} S(i\phi)$ in the Sine-Gordon transformation (\ref{3.1.6},\ref{3.1.7}). This selects the stationary point of the action so that bounds such as Theorem~\ref{theorem- noscreening} are saturated in this limit, \emph{provided} the external charge density $f$ is also scaled suitably. \subsection{The scaling limit of the lattice dipole gas}\label{subsec- scalinglimit} The results of the last section show absence of screening but one would like to know that the system is in a phase characterized by correlations that are asymptotic to canonical power laws prefaced by a dielectric constant or tensor. In this section we will describe preliminary attempts to delineate when the system is actually in a dielectric phase in this more detailed sense. Unfortunately the most appealing result is established only for lattice systems, but the removal of the restriction to lattice models is a feasible mathematical problem. In a lattice model each site ${\bf r}$ in a simple cubic lattice in a $\nu \geq 1$ dimensional box $\Lambda $ can be occupied by zero, one or more dipoles, each of which can only point in a lattice direction. The Sine-Gordon transformation of such systems leads to a partition function of the form \begin{equation}\label{3.2.12} \Xi_{\Lambda } = \int {\cal D} \phi \, e^{-S(\phi)} \end{equation} where ${\cal D} \phi = \prod _{{\bf r} \in \lat } d\phi({\bf r} )$ is a finite dimensional integration and the action $S$ has the general form \begin{equation}\label{3.2.13} S(\phi) = \sum _{<{\bf r} {\bf r} '> } W(\phi({\bf r} ) - \phi({\bf r} ') ) . \end{equation} where $<{\bf r} {\bf r} '> = <{\bf r} ' {\bf r}>$ is a nearest neighbor pair of sites. For example, a dipole gas with no interactions other than the dipole interaction with lattice regularization is given by \begin{equation}\label{3.2.14} W(\phi({\bf r} ) - \phi({\bf r} ') ) = \frac{1}{8\pi}(\phi({\bf r} ) - \phi({\bf r} '))^{2} - \frac{z}{\nu} \cos \(d \beta ^{1/2} (\phi({\bf r}') - \phi({\bf r} ) ) \). \end{equation} The first term is the finite difference version of $(\nabla \phi)^{2}$ and the second term corresponds to the $\cosh(i \beta ^{1/2}{\bf d} \cdot \nabla \phi )$ we have seen before. If there is a hard core interaction preventing more than one dipole per lattice site then $(z/\nu) \cos$ is replaced by $\ln (1 + (z/\nu) \cos )$. Some of the results we are about to state were derived assuming periodic boundary conditions on the lattice. This partition function is also called the {\it anharmonic bedspring}, in which case the variables $\phi$ are considered to be displacements of the nodes of a bedspring (in a distinguished direction). General results \cite{NaSp97} have been obtained under the assumption that $S$ is convex and is a local (or nearly local) function of the gradient of $\phi$, namely assume there is a positive number $\delta >0$ independent of $\phi $ such that the matrix \begin{equation}\label{3.2.15} S'' (\phi; {\bf r} ,{\bf r} ') = \frac{\partial ^{2}S} {\partial \phi({\bf r} )\partial \phi({\bf r} ')} \end{equation} of second derivatives obeys \begin{equation}\label{3.2.16} \sum _{<{\bf r} {\bf r} '>} \zeta ({\bf r} ) S'' (\phi; {\bf r} ,{\bf r} ') \zeta ({\bf r} ') \geq \delta \sum _{<{\bf r} {\bf r} '> } \(\zeta ({\bf r} ) - \zeta ({\bf r} ')\)^{2} \end{equation} for all $\zeta $. The finite difference Laplacian $\Delta $ associated to the lattice is a matrix defined by \begin{equation}\label{3.2.17} \sum_{<{\bf r} {\bf r} '>} \(\zeta ({\bf r} ) - \zeta ({\bf r} ')\)^{2} = \sum_{<{\bf r} {\bf r} '>} \zeta ({\bf r} ) (-\Delta) ({\bf r} ,{\bf r} ') \zeta ({\bf r} '). \end{equation} Thus this convexity assumption is, by definition, the matrix inequality \begin{equation}\label{3.2.convexity} {\rm Convexity:} \ \ S''(\phi) \geq \delta (-\Delta) \end{equation} This type of assumption has a natural continuum analogue: a lower bound on the second variation of $S$ by $\delta \int (\nabla \phi )^{2}$ or perhaps a quadratic form in $-\Delta + L^{2} \Delta ^{2}$, but the next assumption brings in the lattice in a strong way. The assumption is that there is a constant $C$ such that \begin{equation}\label{3.2.upper} \sum _{<{\bf r} {\bf r} '> } \zeta ({\bf r} ) S'' (\phi; {\bf r} ,{\bf r} ') \zeta ({\bf r} ') \leq C \sum _{<{\bf r} {\bf r} '> } \zeta ({\bf r} )^{2}. \end{equation} This is not as natural as the convexity assumption and is false for a continuum model whose action contains $\delta \int (\nabla \phi )^{2}$ or higher order derivatives. Now we come to the {\it scaling limit} which concerns expectations of functions of fields averaged over a length scale $\ell$ that is taken to infinity. For any smooth compactly supported function $f$ with $\sum f({\bf r} ) = 0$ define \begin{equation}\label{3.2.20} \phi(f_{\ell}) = \sum _{{\bf r} } \phi({\bf r} ) \ell^{-\nu/2+1}f(\frac{{\bf r} }{\ell}). \end{equation} The $\sum f({\bf r} ) = 0$ means that $\phi(f_{\ell})$ is really an integral of $\nabla \phi $ against a function of compact support. The theorem will say that when $\ell$ is large $\phi(f_{\ell})$ becomes Gaussian. Gaussian random variables are characterized by their covariance, which, in the theorem, will be the continuum Green's function $C({\bf r} , {\bf r} ')$ solving \begin{equation}\label{3.2.21} - \sum _{i,j=1,\dots ,d} \varepsilon _{ij} \partial _{i} \partial _{j} C({\bf r} , {\bf r} ') = \delta ({\bf r} - {\bf r} '), \end{equation} where $\partial _{i} = \partial /\partial r_{i}$ and $\varepsilon _{ij}$ will be the dielectric tensor. For the theorem we need to define the Gaussian measure \begin{equation}\label{3.2.22} d\mu _{C}(\phi) = \frac{1}{Z} {\cal D} \phi\exp \left(-\frac{1}{2} \int \sum_{i,j} \varepsilon_{ij} \partial _{i}\phi({\bf r} ) \partial_{j}\phi({\bf r} ) d{\bf r} \right) \end{equation} Such a Gaussian measure exists (essentially the massless free field ) and is characterized by Wick's theorem which says \begin{theorem}\label{thm-wick} Define the functional Laplacian \begin{equation}\label{3.2.23} \DDelta = \int d{\bf r} \, \int d{\bf r} ' \, \frac{\delta}{\delta \phi({\bf r} )} C({\bf r} , {\bf r} ') \frac{\delta}{\delta \phi({\bf r} ')} \end{equation} Let $P$ be a polynomial in fields $\phi({\bf r} _{1}),\dots ,\phi({\bf r} _{n})$, then \begin{equation}\label{3.2.24} \int d\mu _{C}(\phi)\, P(\phi) = \exp \left( \frac{1}{2} \DDelta \right) P(\phi) \arrowvert_{\phi=0} \end{equation} where $\exp \left( \frac{1}{2} \DDelta \right)$ is a power series $I + \frac{1}{2} \DDelta + \cdots $ which terminates after a finite number of terms when applied to a polynomial $P$. \end{theorem} An immediate consequence is $\int d\mu _{C}(\phi) \phi({\bf r})\phi({\bf r} ') = C({\bf r} , {\bf r} ')$ which is why $C$ is called a covariance. The theorem \cite{NaSp97} is \begin{theorem}\label{theorem- naddaf-spencer} Under the convexity assumption (\ref{3.2.convexity}) and the uniform upper bound (\ref{3.2.upper}) the continuum scaling limit of the anharmonic bedspring is a continuum massless free field. In other words there exists a constant positive-definite dielectric tensor $\varepsilon _{ij}$ such that, for any $f^{1},\dots ,f^{n}$, as $\ell \rightarrow \infty $, \begin{equation}\label{3.2.25} \langle\prod_{i} \phi(f_{\ell}^{i}) \rangle \rightarrow \int d\mu _{C} \, \prod_{i} \phi(f^{i}) \end{equation} \end{theorem} Thus, in dimension $\nu =3$, the lattice dipole gas (\ref{3.2.14}) is in a dipolar phase characterized by a dielectric constant at least when $\frac{1}{3}z\beta d^{2}4\pi <1$ because this is the parameter range for which (\ref{3.2.14}) is convex. In mathematical terms this is a new type of central limit theorem. Central limit theorems are usually a statement about sums of independent random variables. In the statistical mechanics of lattice spin systems in the high temperature phase the spins are somewhat independent and one expects a central limit theorem that the sum of the spins in a large block of side $\ell$ normalized by $\ell ^{-\nu /2}$ is almost Gaussian and independent of similar sums over disjoint blocks. A beautiful theorem of this type has been proved \cite{New80,NeWr81,NeWr82}. The most gentle way to encode high temperature ( approximate independence) is to require only a finite susceptibility which remarkably is the assumption made by Newman, together with a type of ferromagnetism. In the case of the anharmonic bedspring the susceptibility is not finite; correlations decay with non-integrable power laws and the variables $\phi $ and their gradients are very far from independent. Now we no longer have a standard central limit theorem; indeed the right normalization for a block is now $\ell^{-\nu/2+1}$ but nevertheless these block sums become Gaussian. From the point of view of the renormalization group the central limit theorem is the case of a high temperature fixed point obtained by scaling keeping the mass fixed so that the renormalization group dynamics drives the system to a white noise Gaussian fixed point. Theorem~\ref{theorem- naddaf-spencer} instead describes a massless Gaussian fixed point. Notice that the convexity hypothesis permits interactions that are not small. The proof sharpens the estimates of \cite{BrLi75,BrLi76}, in particular the Brascamp-Lieb bounds: \begin{theorem}\label{theorem- brascamp-lieb} Let $F = F(\phi)$ be any continuously differentiable function of fields. Under the convexity assumption (\ref{3.2.convexity}), \begin{equation}\label{3.2.26} \langle F^{2} \rangle - \langle F \rangle^{2} \leq \sum _{{\bf r} ,{\bf r} '} \langle \frac{\partial F}{\partial \phi({\bf r} )} S''^{-1}(\phi; {\bf r} ,{\bf r} ') \frac{\partial F}{\partial \phi({\bf r}' )} \rangle \end{equation} and \begin{equation}\label{3.2.27} \langle e^{\phi(f)} \rangle \leq \exp \left( \frac{1}{2}\sum _{{\bf r} ,{\bf r} '} f({\bf r} ) (-\delta \Delta )^{-1}({\bf r} ,{\bf r} ') f({\bf r} ') \right) \end{equation} \end{theorem} It is a corollary of theorem~\ref{theorem- naddaf-spencer} that all reasonable functions of fields $\phi(f_{\ell})$ converge to massless free field expectations. Unfortunately the standard observables are not included in this class, being local functions of the unaveraged field $\phi ({\bf r} )$. For example the dipole density is \begin{equation}\label{3.2.28} \hat{\rho }({\cal E};\phi) = ze^{ -i\beta^{1/2} {\bf d} \cdot \nabla \phi({\bf r} ) } \end{equation} Theorem~\ref{theorem- naddaf-spencer} is more or less equivalent to \begin{equation}\label{3.2.29} \langle\prod \alpha ^{\nu/2 - 1}\phi(\alpha {\bf r} _{i}) \rangle \longrightarrow \int d\mu _{C} (\phi ) \prod \phi({\bf r} _{i}) \ \mbox{as } \alpha \rightarrow \infty \end{equation} but to get at the density observables we need information on asymptotics when some but not all of the points ${\bf r} _{1},\dots ,{\bf r} _{n}$ are driven apart by a scaling. Theorem~\ref{theorem- brascamp-lieb} provides an upper bound but lower bounds are not yet proven. There is no physical reason to doubt that they hold, but we are interested in a proof because if the Naddaf-Spencer argument could be extended to prove that more general functionals of $\phi $ have power law decay then a complete proof that there is {\it no} exponential screening in quantum Coulomb systems is within reach using the strategy in \cite{BrKe94}. An alternative route was started by \cite{BrKe94b} but it is presently unreasonably complicated. One further remark about dipole systems is that the pressure and dipole distributions are analytic in the activity in a small neighborhood of zero \cite{GaKu83,BrYa90}. Thus the Mayer expansion is convergent at small activities. Both proofs are complicated and despite many efforts, no one has yet obtained the result by direct attack on the size of the Mayer coefficients. \def\({\left(} \def\){\right)} \def{\bf r}{{\bf r}} \def{\cal F}{{\cal F}} \def\b\xi{\mbox{\boldmath $\xi$\unboldmath}} \def\underline{\mbox{\boldmath $\mu$\unboldmath}}{\underline{\mbox{\boldmath $\mu$\unboldmath}}} \def\underline{\bf N}{\underline{\bf N}} \newcommand{d^{[0,1]}\mu}{d^{[0,1]}\mu} \newpage \section{Semi-classical Coulomb gas} \setcounter{page}{1} \label{chapter- Semi-classical}\label{chapter-4} \subsection{The Feynman-Kac representation} \label{sec-4.1 } In this chapter, we give an introduction to some of the main effects produced by quantum mechanics in the Coulomb gas. These effects are most easily seen by perturbing around the classical gas, i.e. neglecting the quantum statistics and keeping the first relevant contributions to an expansion in powers of the Planck constant $\hbar$ (the Wigner-Kirkwood expansion)\footnote{Note that to have stability of the classical reference system, the semi-classical analysis requires a regularization of the Coulomb potential at the origin, except for the one component plasma which remains well behaved in the classical limit.}. We have no control on the possible convergence or asymptotic character of these $\hbar$-series, but the examination of the lowest order terms gives immediately a prediction to the main issue: Debye screening cannot survive if the quantum mechanical nature of the charges is taken into account. This serves as a gentle training for the reading of chapter~\ref{chapter-loops} where the fully quantum mechanical gas is considered, since most of the arguments and of the mathematical structure will be similar. The formalism that is best adapted for both a semi-classical and a low density analysis of the quantum Coulomb gas is the Feynman-Kac functional integral representation. This representation is of course not new and goes back to \cite{Gin65,Gin71} with his study of the low activity expansion of quantum fluids with short range forces. But, for quantum charges, the formalism shows its full capability and becomes particularly operational to deal with the long range of the Coulomb potential. In section \ref{sec-4.3} we discuss the simpler situation of only two quantum charges immersed in a classical plasma. The model can be studied without recourse to $\hbar$ expansions and illustrates clearly how quantum fluctuations will irremediably destroy Debye screening: one finds an $r^{-6}$ decay of the correlation between the two quantum charges. In section \ref{sec-4.4} we comment on the relation between this type of correlation between individual charges and the conventional van der Waals potential. Section \ref{sec-4.5} comes back to a system of infinitely many quantum charges and summarizes the results obtained by the semi-classical analysis. In the last section, we combine the Sine-Gordon and the Feynman-Kac functional integrals: this gives an alternative view on the lack of exponential screening. Again a simplified model that captures the essence of quantum fluctuations illustrates how the occurrence of dipole (and higher multipole) fields will spoil the proof of Debye screening. In contrast, mixtures of classical particles that carry both charges and dipoles still show strong screening properties. We start here and in section \ref{sec-4.2} by recalling familiar ideas about the representation of the quantum mechanical statistical operator by the Feynman-Kac functional integral. For the sake of simplicity, we consider first a single particle of mass $m$ in three dimensions moving in an external potential $V({\bf r})$. According to the original path integration \cite{FeHi65}, the configurational matrix elements of the statistical operator associated with the one-particle Hamiltonian \begin{equation} H=-\frac{\hbar^2}{2m}\Delta +V,\;\;\;\;\,\,\Delta=\mbox{Laplacian in}\; {\Bbb R}^3 \label{4.1} \end{equation} read \begin{equation} \langle{\bf r}_1 |\exp\left(-\beta\left(-\frac{\hbar^2}{2m}\Delta+V\right)\right) |{\bf r}_2\rangle= \sum_{paths}\exp\left(-\frac{1}{\hbar}S({\bf r}(\cdot))\right) \label{4.2} \end{equation} In (\ref{4.2}) $S({\bf r}(\cdot))$ is the classical action corresponding to the potential $-V$ (the Euclidean action) \begin{equation} S({\bf r}(\cdot))=\int_0^{\beta \hbar}dt\left(\frac{m}{2}\left |\frac{d{\bf r}(t)}{dt}\right |^2 +V({\bf r}(t))\right) \label{4.3} \end{equation} associated with the path ${\bf r}(t)$ starting from ${\bf r}_1$ at "time" $t=0$ and ending in $ {\bf r}_2$ at "time" $\beta\hbar$. The summation in (\ref{4.2}) runs over all such paths. It is very useful to parameterize the path ${\bf r}(t)$ by dimensionless variables, making the change \begin{eqnarray} s&=&\frac{t}{\beta\hbar}\;,\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;0\leq\:s\:\leq 1 \nonumber\\ {\bf r}_{1,2}(s)&=&(1-s){\bf r}_1 +s{\bf r}_2 +\lambda \b\xi(s)\;,\;\;\;\;\;\;\;\b\xi(0)=\b\xi(1)=0 \label{4.4} \end{eqnarray} where \begin{equation} \lambda=\hbar\sqrt{\frac{\beta}{m}} \label{4.5} \end{equation} is the de Broglie thermal wave length. In (\ref{4.4}) $\b\xi(s)$ represents now a closed path, called the Brownian bridge, starting and returning to the origin within the "time" $s=1$. In terms of these dimensionless variables, the matrix element (\ref{4.2}) can be written in the precise form \cite{Sim79,Roe94} \begin{eqnarray} \langle{\bf r}_1\mid\exp\left(-\beta\left(-\frac{\hbar^2}{2m}\Delta+V\right)\right)\mid{\bf r}_2\rangle &=& \left(\frac{1}{2\pi\lambda^2}\right)^{3/2}\exp\left(-\frac{|{\bf r}_1 -{\bf r}_2|^2}{2\lambda^2}\right) \nonumber \\ && \hspace{-2.25in} \times\int D(\b\xi)\exp\left(-\beta\int_{0}^{1} ds V((1-s){\bf r}_1 +s{\bf r}_2 +\lambda \b\xi(s))\right) \label{4.6} \end{eqnarray} In (\ref{4.6}) $ D(\b\xi)$ is a Gaussian measure (the Brownian bridge measure) with weight formally represented (up to normalization) by $\exp\left(-\frac{1}{2}\int_0^1\left|\frac{d\b\xi(s)}{ds}\right|^2\right)$. It has zero mean and is entirely defined by its covariance \begin{equation} \int D(\b\xi)\xi_{\mu}(s_1)\xi_{\nu}(s_2)=\delta_{\mu,\nu}(\min(s_1,\:s_2)-s_1 s_2) \label{4.7} \end{equation} where $\xi_{\mu}(s)$ are the Cartesian coordinates of $\b\xi(s)$. The representation (\ref{4.6}) has several advantages: physical parameters ${\bf r}_1,\;{\bf r}_2,\;\beta,\;\hbar,\;\lambda$ appear explicitly in the formula. Moreover the diagonal matrix elements \begin{eqnarray} && \hspace{-.5in} \langle{\bf r}\mid\exp\left(-\beta\left(-\frac{\hbar^2}{2m}\Delta+V\right)\right)\mid{\bf r}\rangle=\nonumber\\ && \left(\frac{1}{2\pi\lambda^2}\right)^{3/2}\int D(\b\xi)\exp\left(-\beta\int_0^1dsV({\bf r} +\lambda \b\xi(s))\right) \label{4.8} \end{eqnarray} have the form of an integral over a certain $\b\xi$-dependent Boltzmann factor $\exp(-\beta V({\bf r},\b\xi))$ with $V({\bf r},\b\xi)=\int_0^1dsV({\bf r} +\lambda\b\xi(s))$, as in classical statistical mechanics. In this view, one can think of a quantum point particle as a classical extended object, the closed filament $\b\xi(s)$ located at ${\bf r}$. This closed filament, which plays formally the same role as a classical internal degree of freedom, represents the intrinsic quantum fluctuation with size controlled by the de Broglie wave length $\lambda$. The statistical weight of a filament having extension $R$ behaves as $\exp(-R^2/\lambda^2)$; thus, in the semi-classical regime ($\lambda$ small), only short filaments contribute to the integral (\ref{4.8}). Obviously one recovers the genuine classical Boltzmann factor $\exp(-\beta V({\bf r}))$ in (\ref{4.8}) as $\lambda \rightarrow 0$. For these reasons, the Feynman-Kac representation is particularly suitable to study the semi-classical aspects of equilibrium quantities. \subsection{The gas of charged filaments} \label{sec-4.2} The extension of the representation (\ref{4.8}) to the many particle system is straightforward. In the sequel, $V({\bf r}_{1}-{\bf r}_{2})$ represents a two-body regularized Coulomb potential, for instance of the form (\ref{1.1.3bc}) . To each particle labeled by $i,\; i=1,\ldots, N$, we associate an independent Brownian bridge $\b\xi_i(s)$, distributed with the Gaussian measure $D(\b\xi_i)$, and the diagonal elements of the $N$- particles statistical operator read \begin{eqnarray} \langle\{{\bf r}_i\}\mid\exp\(-\beta H_{\Lambda,N}\)\mid\{{\bf r}_i\}\rangle=\prod_{i=1}^{N}\frac{1}{(2\pi\lambda_{\alpha_{i}}^2)^{3/2}} \int\prod_{i=1}^N D(\b\xi_i) \nonumber \\ \exp\(-\frac{\beta}{2}\sum_{i\neq j}^N e_{\alpha_i}e_{\alpha_j}\int _0^1 ds V({\bf r}_i + \lambda_{\alpha_i}\b\xi_i(s)-{\bf r}_j - \lambda_{\alpha_j}\b\xi_j(s))\) \label{4.9} \end{eqnarray} The Dirichlet boundary conditions on $H_{N,\Lambda}$ are implemented by the constraint that all paths in the integrand of (\ref{4.9}) have to remain inside $\Lambda$ for all $s,\; 0\leq\; s \; \leq 1$. We do not write this constraint explicitly since it will be trivially removed in the infinite volume limit. The grand-canonical Maxwell-Boltzmann partition function is defined by keeping in (\ref{1.2.5}) only the term corresponding to the trivial permutation. Denoting it still by $\Xi_{\Lambda}$ in this chapter, it is given by the sum \begin{eqnarray} \Xi_\Lambda&=&\sum_{\{N_\alpha\}}^\infty \prod \frac{((2\sigma_\alpha+1)\exp(\beta\mu_{\alpha}))^{N_\alpha}}{N_\alpha !}\nonumber\\&\times& \int_\Lambda d{\bf r}_1 \ldots d{\bf r}_N \langle\{{\bf r}_i\}\mid\exp\(-\beta H_{\Lambda,N}\)\mid\{{\bf r}_i\}\rangle \label{4.11} \end{eqnarray} The factor $2\sigma_\alpha +1$ is the spin degeneracy, since the Hamiltonian (\ref{1.2.1}) is independent of the spin. A look at the Boltzmann-like factor in (\ref{4.9}) suggests naturally the introduction of an enlarged classical phase space whose elements ${\cal F} $ are \begin{equation} {\cal F} = (\alpha,\;{\bf r},\;\b\xi) \label{4.12} \end{equation} where $\alpha$ is the species and $\b\xi$ is the filament shape of a particle at position ${\bf r}$. We shall call a point ${\cal F}$ in this phase space a "filament". Since the Hamiltonian is invariant under the permutations of the particles of the same species, we can use again the summation identity (\ref{1.1.4}) to write the partition function (\ref{4.11}) in the form \begin{equation} \Xi_\Lambda = \sum_{N=0}^\infty\frac{1}{N!}\int \prod_{k=1}^N d{\cal F}_k z({\cal F}_k)\exp\(-\beta U({\cal F}_1, \ldots, {\cal F}_N)\) \label{4.14} \end{equation} where the phase space integration means \begin{equation} \int d{\cal F}\cdots=\sum_{\alpha=1}^{\cal S}\int_{\Lambda}d{\bf r}\int D(\b\xi)\cdots \label{4.15} \end{equation} In (\ref{4.14}) we have set \begin{equation} z({\cal F})=(2\sigma_\alpha+1)z_\alpha,\;\;\;\;z_{\alpha}=\frac{\exp(\beta\mu_{\alpha})} {\(2\pi\lambda_{\alpha}^2\)^{3/2}} \label{4.16} \end{equation} \begin{equation} U({\cal F}_1,\ldots, {\cal F}_N)=\sum_{1=i<j}^N e_{\alpha_i} e_{\alpha_j}V({\cal F}_i,{\cal F}_j) \label{4.17} \end{equation} with \begin{equation} V({\cal F}_1,{\cal F}_2)=\int_0^1dsV({\bf r}_1+\lambda_{\alpha_1}\b\xi_1(s)-{\bf r}_2-\lambda_{\alpha_2}\b\xi_2(s)) \label{4.18} \end{equation} It is clear that the system defined by the relations (\ref{4.14})-(\ref{4.18}) has the structure of a classical statistical mechanical system of objects having internal degrees of freedom $\alpha,\; \b\xi$ (the species and the shape of a filament), with activity (\ref{4.16}) and two-body interaction (\ref{4.18}). We call it "the system of filaments" and can apply to it all the standard methods of classical statistical mechanics. In particular, one can define correlations between filaments in the usual way (see (\ref{1.1.8a}) and (\ref{1.1.9a})). Introducing the "$\delta$-function" which identifies the points ${\cal F}_1$ and ${\cal F}_2$ in phase space, i.e. \begin{equation} \int d{\cal F}_2\delta({\cal F}_1,{\cal F}_2)F({\cal F}_2)=F({\cal F}_1),\;\;\;\;\;\delta({\cal F}_1,{\cal F}_2)= \delta_{\alpha_1,\alpha_2}\delta(\b\xi_1,\b\xi_2)\delta({\bf r}_1-{\bf r}_2) \label{4.19} \end{equation} the density of filaments and the two filament distribution are \begin{equation} \rho({\cal F})=\langle\hat{\rho}({\cal F})\rangle,\;\;\;\;\rho({\cal F}_a,{\cal F}_b)=\langle\hat{\rho}({\cal F}_a)\hat{\rho} ({\cal F}_b)\rangle \label{4.20} \end{equation} where \begin{equation} \hat{\rho}({\cal F})=\sum_i\delta({\cal F}_i,{\cal F}) \label{4.21} \end{equation} and $\langle\cdots\rangle$ denotes the grand-canonical average in the system of filaments (the contribution of coincident filaments, included in (\ref{4.20}), will be specified later if needed). However, one should remember that the shapes of the filaments are not directly observable quantities: the physical particle distributions are obtained by integrating them out in (\ref{4.20}) and (\ref{4.21}). Thus the density of particles of type $\alpha$ is \begin{equation} \rho(\alpha,{\bf r})=\int D(\b\xi)\rho({\cal F}) \label{4.22} \end{equation} the two particle distribution is \begin{equation} \rho(\alpha_a,{\bf r}_a,\alpha_b,{\bf r}_b)=\int D(\b\xi_a)D(\b\xi_b)\rho({\cal F}_a,{\cal F}_b) \label{4.23} \end{equation} and in the same way for the truncated distributions, \begin{eqnarray} \rho_T({\cal F}_a,{\cal F}_b)&=&\rho({\cal F}_a,{{\cal F}}_b)-\rho({\cal F}_a)\rho({\cal F}_b) \nonumber\\ \rho_T(\alpha_a,{\bf r}_a,\alpha_b,{\bf r}_b)&=& \int D(\b\xi_a)D(\b\xi_b)\rho_T({\cal F}_a,{{\cal F}}_b) \label{4.24} \end{eqnarray} At this point, one has to make a fundamental observation. Although the system of filaments can be treated with the rules of classical statistical mechanics, it differs from a genuine assembly of classical charged random wires (distributed according to the Gaussian measure $D(\b\xi)$ ) on an important point: {\em the interaction (\ref{4.18}) inherited from the quantum mechanical nature of the charges is not the standard electrostatic potential between two closed wires of shapes $\b\xi_1$ and $\b\xi_2$}. The latter potential, denoted $V_{{\rm cl}}({\cal F}_1,{\cal F}_2)$, is \begin{equation} V_{{\rm cl}}({\cal F}_1,{\cal F}_2)=\int_0^1 ds_1\int_0^1 ds_2 V({\bf r}_1+\lambda_{\alpha_1}\b\xi_1(s_1)-{\bf r}_2-\lambda_{\alpha_2}\b\xi_2(s_2)) \label{4.25} \end{equation} since every element of charge $\b\xi_1(s_1)ds_1$ carried by the first filament has to interact with every element $\b\xi_2(s_2)ds_2$ of the other by the Coulomb law, and the corresponding total electrostatic energy is \begin{equation} U_{{\rm cl}}({\cal F}_1,\ldots, {\cal F}_k))=\sum_{i<j=1}^k e_{\alpha_i} e_{\alpha_j}V_{{\rm cl}}({\cal F}_i,{\cal F}_j) \label{4.26} \end{equation} \emph{Although the filaments are quantum mechanical in origin, we will use this subscript $_{{\rm cl}}$ to denote the filament system with interaction $U_{{\rm cl}}$}. Replacing the interaction (\ref{4.18}) by (\ref{4.25}) would indeed lead to a genuine system of charged filaments obeying exponential Debye screening in the appropriate low density or high temperature regime, according to the theorems quoted in chapter~\ref{chapter- debye-screening} and the discussion in section~\ref{sec-4.6}. Hence different behaviors between quantum and classical charges in thermal equilibrium must be traced back to the difference between the interactions (\ref{4.18}) and (\ref{4.25}) \begin{eqnarray} W({\cal F}_1,{\cal F}_2) &=& e_{\alpha_1} e_{\alpha_2}(V({\cal F}_1,{{\cal F}}_2)-V_{{\rm cl}}({\cal F}_1,{\cal F}_2) ) \nonumber\\ &=& e_{\alpha_1} e_{\alpha_2}\int_0^1 ds_1\int_0^1 ds_2(\delta(s_1 - s_2)-1) \nonumber \\ &&\hspace{-.5in}\times V({\bf r}_1+\lambda_{\alpha_1}\b\xi_1(s_1)-{\bf r}_2-\lambda_{\alpha_2} \b\xi_2(s_2)) \label{4.27} \end{eqnarray} The large distance behavior of $W({\cal F}_1,{\cal F}_2)$ is obtained from the multipolar expansion of the potential $V$ written in the form \begin{eqnarray} V({\bf r}_1+\lambda_{\alpha_1}\b\xi_1(s_1)-{\bf r}_2-\lambda_{\alpha_2}\b\xi_2(s_2))&=& V({\bf r}_1-{\bf r}_2)+M_1(s_1)V({\bf r}_1-{\bf r}_2)\nonumber\\ &&\hspace{-2in} +M_2(s_2)V({\bf r}_1-{\bf r}_2) + M_1(s_1)M_2(s_2)V({\bf r}_1-{\bf r}_2) \label{4.28} \end{eqnarray} where $M_{i}(s)$ are the multipolar differential operators \begin{equation} M_{i}(s)=\sum_{k=1}^{\infty}\frac{\(\lambda_{\alpha_{i}}\b\xi_{i}(s)\cdot \nabla_{{\bf r}_{i}}\)^k}{k!},\;\;\;\;\;i=1,2 \label{4.28a} \end{equation} In the decomposition (\ref{4.28}), the three first terms correspond to the charge-charge and charge-multipole interactions, whereas the last term represents the rest of the multipole-multipole interactions. Since \begin{equation} \int_0^1 ds_1 (\delta(s_1-s_2)-1)=\int_0^1 ds_2 (\delta(s_1-s_2)-1)=0 \label{4.29} \end{equation} we see that only this latter term contributes to $W$ \begin{equation} W({\cal F}_1,{\cal F}_2)=e_{\alpha_1} e_{\alpha_2}\int_0^1 ds_1\int_0^1 ds_2(\delta(s_1-s_2)-1) M_1(s_1)M_2(s_2)V({\bf r}_1-{\bf r}_2) \label{4.30} \end{equation} The dominant large distance behavior of $W$ \begin{eqnarray} W({\cal F}_1,{\cal F}_2) &\sim& e_{\alpha_1} e_{\alpha_2}\int_0^1 ds_1\int_0^1 ds_2(\delta(s_1-s_2)-1)\nonumber\\ &&\hspace{-1in} \times(\lambda_{\alpha_1}{\b\xi}_1(s_1)\cdot\nabla_{{\bf r}_1}) (\lambda_{\alpha_2}{\b\xi}_2(s_2)\cdot\nabla_{{\bf r}_2}) V({\bf r}_1-{\bf r}_2),\;\;\;\;\;\;\;\mid {\bf r}_1-{\bf r}_2\mid \rightarrow \infty \nonumber \\ \label{4.31} \end{eqnarray} is typically a dipole-dipole potential, decaying as $\mid {\bf r}_1-{\bf r}_2\mid^{-3}$, due to the dipole elements $e_{\alpha_{1}}\lambda_{\alpha_1}{\b\xi}_1(s_1)ds_1$ and $e_{\alpha_{2}}\lambda_{\alpha_2}{\b\xi}_1(s_2)ds_2$ associated with two filaments of arbitrary shapes. It will be shown (chapter ~\ref{chapter-loops}) that the charge-charge and charge-multipole interactions have a fast screening in the quantum gas, but not the multipole-multipole interaction. This long range term will precisely be the cause for the breakdown of exponential screening in the quantum system for all values of the density and of the temperature. For the sake of pedagogy, we illustrate the above assertions in a simplified model involving only two quantum mechanical charges. This model presents already all the basic mechanisms which will be at work in the general case. \subsection{Quantum fluctuations destroy exponential screening} \label{sec-4.3} We consider the simplified situation where all particles are classical but two. For this it suffices to single out two filaments ${\cal F}_a=(\alpha_a, \,{\bf r}_a, \, \b\xi_a) $, ${\cal F}_b=(\alpha_b, \,{\bf r}_b, \, \b\xi_b)$ corresponding to two particles with non vanishing de Broglie lengths $\lambda_a$, $\lambda_b$, and to set the de Broglie lengths of all the other particles equal to zero. Then the phase space variable of a classical particle reduces to $(\alpha,\; {\bf r})$ and its interaction with the quantum particle ${\cal F}_a$ to \begin{equation}V({\cal F}_a, {\bf r})=\int_0^1 dsV({\bf r}_a+\lambda_a\b\xi_a(s)-{\bf r})= V_{{\rm cl}}({\cal F}_a, {\bf r}) \label{4.32} \end{equation}which is identical to the classical potential (\ref{4.25}) with one $\lambda$ set equal to $0$, whereas the two quantum particles still interact with the potential (\ref{4.18}). Here the quantum mechanical nature of the two specified particles will be treated non-perturbatively. The model is semi-classical only in the sense that their surrounding medium is now a classical plasma at density $\rho$ and inverse temperature $\beta$. As in (\ref{1.1.7}) we denote by $\omega=(\alpha_1,{\bf r}_1,\ldots,\alpha_n,{\bf r}_n)$ a configuration of the classical particles and by $\hat{c}({\bf r},\omega)=\sum_{j=1}^n e_{\alpha_j}\delta({\bf r} -{\bf r}_j)$ the corresponding microscopic charge density. For this system, the total interaction (\ref{4.17}) reads \begin{eqnarray} U({\cal F}_a,{\cal F}_b,\omega)&=& e_{\alpha_a} e_{\alpha_b}V({\cal F}_a,{\cal F}_b)+ e_{\alpha_a}\int d{\bf r} V({\cal F}_a, {\bf r})\hat{c}({\bf r},\omega)\nonumber\\ &+&e_{\alpha_b}\int d{\bf r} V({\cal F}_b, {\bf r})\hat{c}({\bf r},\omega) +U_0(\omega) \label{4.34} \end{eqnarray} where $U_0(\omega)$ is the Coulomb energy of the classical particles. The distribution of two filaments ${\cal F}_a$ and ${\cal F}_b$ in thermal equilibrium with the classical Coulomb gas enclosed in a finite volume $\Lambda$ is defined by \begin{equation} \rho({\cal F}_a,{\cal F}_b)=\frac{1}{\Xi_0}\int_\Lambda d\omega\exp\(-\beta U({\cal F}_a,{\cal F}_b,\omega)\) \label{4.35} \end{equation} In (\ref{4.35}), $\int_\Lambda d\omega$ means (canonical or grand canonical) summation on the configurations $\omega$ of the classical particles in $\Lambda$ and $\Xi_0=\int_\Lambda d\omega\exp(-\beta U_0(\omega))$ is the corresponding partition function (in this definition, there is no contribution of coincident filaments). For a single filament we define in the same way\footnote{$-\beta ^{-1}\ln \rho ({\cal F} )$ and $-\beta ^{-1}\ln \rho ({\cal F}_{a},{\cal F} _{b} )$ are the excess free energies when one or two filaments are embedded in the classical plasma.} \begin{equation} \rho({\cal F})=\frac{1}{\Xi_0}\int_{\Lambda } d\omega\exp\(-\beta U({\cal F},\omega)\) \label{4.36} \end{equation} with \begin{equation} U({\cal F},\omega)=e_\alpha\int d{\bf r} \, V({\cal F},{\bf r})\hat{c}({\bf r},\omega)+U_0(\omega) \label{4.37} \end{equation} Let us also introduce the auxiliary genuine classical distributions of charged filaments $\rho_{{\rm cl}}({\cal F}_a,{\cal F}_b), \;\rho_{{\rm cl}}({\cal F})\;$ defined in the same way as (\ref{4.36}, \ref{4.37}), but with $V$ and $U$ replaced by $V_{{\rm cl}}$ and $U_{{\rm cl}}$. Comparing now the quantum and this classical system, we observe that in view of (\ref{4.32}) \begin{equation} U({\cal F},\omega)= U_{{\rm cl}}({\cal F},\omega)\;,\;\;\;\;\,\rho({\cal F})=\rho_{{\rm cl}}({\cal F}) \label{4.38} \end{equation} but taking into account the definitions (\ref{4.26}) and (\ref{4.27}) \begin{eqnarray} U({\cal F}_a,{\cal F}_b,\omega)&=&W({\cal F}_a,{\cal F}_b)+U_{{\rm cl}}({\cal F}_a,{\cal F}_b,\omega)\nonumber\\ \rho({\cal F}_a,{\cal F}_b)&=&\exp\(-\beta W({\cal F}_a,{\cal F}_b)\)\rho_{{\rm cl}}({\cal F}_a,{\cal F}_b) \label{4.39} \end{eqnarray} We consider now these distributions in the thermodynamic limit, keeping the same notations. Because of translational invariance, $\rho({\cal F})=\rho_{\alpha }(\b\xi)$ is independent of ${\bf r}$ and $\rho({\cal F}_a,{\cal F}_b)$ depends on $|{\bf r}_{a}-{\bf r}_{b}|$. In the Debye regime, we know that $\rho_{{\rm cl}}({\cal F}_a,{\cal F}_b)$ clusters exponentially fast as $|{\bf r}_a-{\bf r}_b| \rightarrow\infty$ (with $\b\xi_a,\; \b\xi_b$ fixed) \begin{eqnarray} \rho_{{\rm cl}}({\cal F}_a,{\cal F}_b)&=&\rho_{{\rm cl}}({\cal F}_a)\rho_{{\rm cl}}({\cal F}_b)\nonumber\\ &+&O\(\exp\(-C\inf_{0\leq s_1,s_2\leq 1}|{\bf r}_a+\lambda_{\alpha_a}\b\xi_a(s_1)- {\bf r}_b-\lambda_{\alpha_b}\b\xi_b(s_2)|\)\)\nonumber\\ \label{4.40} \end{eqnarray} Thus the asymptotic behavior of $\rho({\cal F}_a,{\cal F}_b)$ in (\ref{4.39}) is governed by that of $W({\cal F}_a,{\cal F}_b)$, i.e. according to (\ref{4.31}), $\rho({\cal F}_a,{\cal F}_b)$ clusters as $|{\bf r}_a-{\bf r}_b|^{-3}$. The two-particle distribution is obtained by averaging $\rho({\cal F}_a,{\cal F}_b)$ on the shapes of the filaments. Using (\ref{4.38}, \ref{4.40}) and expanding the factor $\exp\(-\beta W({\cal F}_a,{\cal F}_b)\)$ yields \begin{eqnarray} \rho({\bf r}_a,{\bf r}_b)&=&\int D(\b\xi_a)D(\b\xi_b) \exp\(-\beta W({\cal F}_a,{\cal F}_b)\)\rho_{{\rm cl}}({\cal F}_a,{\cal F}_b)\nonumber\\ &=&\rho_a\rho_b-\beta \int D(\b\xi_a)D(\b\xi_b)W({\cal F}_a,{\cal F}_b)\rho_a(\b\xi_a)\rho_b(\b\xi_b)\nonumber\\ &+&\frac{1}{2}\beta^2\int D(\b\xi_a)D(\b\xi_b)W^2({\cal F}_a,{\cal F}_b)\rho_a(\b\xi_a)\rho_b(\b\xi_b) + \cdots +O\(\exp(-C|{\bf r}_a-{\bf r}_b|)\) \label{4.41} \end{eqnarray} where $\rho _{a} = \rho (\alpha _{a},{\bf r} _{a})$. The remainder in (\ref{4.41}) still gives an exponentially decreasing contribution since the probability of long filaments is Gaussian small. Let us show that the term linear in $W$ in (\ref{4.41}) does not contribute to the asymptotic behavior of $\rho({\bf r}_a,{\bf r}_b)$. Indeed one notes that the measure $ D(\b\xi)$ is invariant under reflections and rotations of the filament in three-dimensional space (see (\ref{4.7})). The same is true for $\rho(\b\xi)$ in an homogeneous phase of the classical plasma. Hence introducing the multipolar expansion of $W$ with the help of (\ref{4.28a}), (\ref{4.30}) one sees the occurrence of the $\b\xi$-averages \begin{eqnarray} && \int D(\b\xi_a) D(\b\xi_b)\rho(\b\xi_a)\rho(\b\xi_b)\nonumber \\ &&\hspace{-1in} [\lambda_a\b\xi_a(s_1)\cdot\nabla_{{\bf r}_{a}}]^k [\lambda_b\b\xi_b(s_2)\cdot\nabla_{{\bf r}_{b}}]^lV({\bf r}_a-{\bf r}_b)\;\,\,\, k,\:l\geq 1 \label{4.42} \end{eqnarray} In view of these symmetry considerations, the terms with $k$ or $l$ odd vanish, and the terms with both $k$ and $l$ even are necessarily proportional to $\(\nabla_{{\bf r}}^2\)^{\frac{k+l}{2}}V({\bf r}), {\bf r}={\bf r}_a-{\bf r}_b$, which is a rapidly decreasing function since $\nabla_{{\bf r}}^2\frac{1}{|{\bf r}|}=0, {\bf r}\neq 0$. Since all moments of the Gaussian measure $ D(\b\xi)$ are finite, one concludes that the $W$-linear term in (\ref{4.41}) decays faster than any power of $|{\bf r}_a-{\bf r}_b|$. The leading behavior is therefore determined by the quadratic term. According to (\ref{4.41}) and (\ref{4.31}) one finds for the normalized truncated distribution \begin{eqnarray} \frac{\rho_{T}({\bf r}_{a},{\bf r}_{b})}{\rho_a\rho_b}&=&\frac{\rho({\bf r}_a,{\bf r}_b)}{\rho_a\rho_b}-1\nonumber\\ &=&\frac{\beta^{2}}{2\rho_{a}\rho_{b}}\int D(\b\xi_a)D(\b\xi_b)W^2({\cal F}_a,{\cal F}_b)\rho_a(\b\xi_a)\rho_b(\b\xi_b)+{\rm O}(W^{3})\nonumber\\ &=&\frac{B(\beta,\rho)}{|{\bf r}_a-{\bf r}_b|^6}+O\(\frac{1}{|{\bf r}_a-{\bf r}_b|^8}\) \label{4.43} \end{eqnarray} where $B(\beta,\rho)>0$ can be expressed, from (\ref{4.31}), as the Brownian integral of a positive functional, depending on the temperature and density of the surrounding plasma. The correction $\mid r_{a} - r_{b}\mid^{-8}$ in (\ref{4.43}) arises from the square of the dipole-quadrupole interaction. Note that the result (\ref{4.43}) holds in any screening phase of the classical plasma and is non-perturbative in $\hbar$. In the classical limit or in the high temperature limit, one finds that the dominant contribution to $B(\beta,\rho)$ is the same \cite{AlMa89} \begin{equation} B(\beta,\rho)\simeq\hbar^4\frac{\beta^4}{240}\frac{e_a^2e_b^2}{m_am_b},\;\;\;\hbar\rightarrow 0 \;\;\;\mbox{or}\;\;\;\beta\rightarrow 0 \label{4.44} \end{equation} The role of the quantum fluctuations is clearly demonstrated: the dipole-like interaction $W$, which originates from these intrinsic quantum fluctuations, can definitely not be shielded by the classical charges, even if the latter are in the Debye phase. Nevertheless a partial screening remains: the monopole is still screened and this, together with rotation invariance, reduces the bare Coulomb decay $r^{-1}$ to $r^{-6}$. It will be shown in section \ref{sec-4.5} as well in chapters~\ref{chapter-loops} and \ref{chapter-6} that these facts remain true in the full quantum gas. \subsection{Origin of the van der Waals forces} \label{sec-4.4} One can interpret the result (\ref{4.43}) by saying that there exists an effective potential $\Phi_{1}(r)$ between two individual quantum charges defined by $\frac{\rho_{T}({\bf r}_{a},{\bf r}_{b})}{\rho_a\rho_b}=e^{-\beta\Phi_{1}(|{\bf r}_{a}-{\bf r}_{b}|)}-1$. It is attractive at large distance and behaves as \begin{equation} \beta\Phi_{1}(r)\simeq -\frac{B(\beta,\rho)}{r^{6}},\;\;\;\; r\rightarrow\infty \label{4.44a} \end{equation} One may wonder if this potential between individual quantum charges is related to the van der Waals potential, also decaying as $r^{-6}$. The van der Waals forces are usually computed for preformed atoms and molecules in empty space, and are attributed to the dipole fluctuations in these bound entities. To allow for the possibility of quantum mechanical binding in simple terms, we can generalize the model of section~\ref{sec-4.3} by specifying four particles, say two electron-proton (e-p) pairs, in equilibrium with a classical plasma \cite{Mar96}. The information about the correlations between these four particles is contained in the fully truncated four point correlation function $\rho_{T}({\bf r}_{a},{\bf r}_{b},{\bf R}_{1}, {\bf R}_{2})$ where ${\bf r}_{a},{\bf r}_{b}$ designate the coordinates of the electrons and ${\bf R}_{1}, {\bf R}_{2}$ those of the protons. This correlation is defined by the natural generalization of the expressions (\ref{4.35})-(\ref{4.39}) to four particles immersed in the plasma. We are interested in the amount of correlation that exists between the two (e-p) pairs as their centers of mass are taken far apart, but ignoring the relative positions of the particles. Assuming for simplicity that the two protons are infinitely heavy (i.e. classical), the center of mass of the pairs coincide with ${\bf R}_{1}, {\bf R}_{2}$ and the desired normalized correlation is \begin{equation} F(R)=\frac{1}{2}\frac{\int d{\bf r}_{a}\,\int d{\bf r}_{b} \,\rho_{T}({\bf r}_{a},{\bf r}_{b},{\bf R}_{1}, {\bf R}_{2})} {\int d{\bf r}_{a}\,\rho_{T}({\bf r}_{a},{\bf R}_{1})\int d{\bf r}_{b} \,\rho_{T}({\bf r}_{b},{\bf R}_{2})}, \;\;\;\;R=|{\bf R}_{1}- {\bf R}_{2}| \label{4.44b} \end{equation}The factor $1/2$ takes into account that the electrons, treated here as identical particles, can be found in the neighborhood of either one of the protons (the contribution of coincident particles is omitted in (\ref{4.44b})\footnote{As before, exchange effects due to the Fermi statistics of the electrons are neglected. Exchange correlations have a short range.}). We define the effective (temperature and density dependent) potential $\Phi_{2}(R)$ between the two pairs by \begin{equation} F(R)=e^{-\beta\Phi_{2}(R)}-1\simeq -\beta \Phi_{2}(R),\;\,\;R\rightarrow\infty \label{4.44c} \end{equation}In contrast to the standard calculation of van der Waals forces, $\Phi_{2}$ incorporates now the effects of the thermal fluctuations of the electrons and of the screening provided by the medium. As in (\ref{4.41}), the asymptotic form of $\Phi_{2}(R)$ is determined by the quadratic term in the quantum potential $W$, namely $$ -\beta \Phi_{2}(R)\simeq \int \bar{D}(\b\xi_{a})\bar{D}(\b\xi_{b}) W^{2}({\cal F}_{a},{\cal F}_{b}) $$ where now $ \bar{D}(\b\xi_{a}),\;\bar{D}(\b\xi_{b})$ are dressed Brownian measures involving the single (e-p) pair correlation. One finds \begin{equation} \beta\Phi_{2}(R)\simeq -\frac{C(\beta,\rho)}{R^{6}},\;\,\,R \rightarrow\infty,\;\,\,C(\beta,\rho)>0 \label{4.44d} \end{equation} There are two situations of interest, the high temperature limit ($\beta\rightarrow 0, \;\rho$ fixed) and the atomic limit. The atomic limit is obtained by simultaneously lowering the temperature ($\beta\rightarrow\infty$) to favor binding and reducing the density at an exponential rate ($\rho\simeq e^{-\delta\beta}\rightarrow 0,\;\delta>0$) to get independent atoms. In this limit (e-p) pairs form noninteracting hydrogen atoms in their ground states (the atomic limit will be discussed in chapter~\ref{chapter-7} for the full quantum gas). In the high temperature limit, one recovers the result (\ref{4.44}), i.e. \begin{equation} C(\beta,\rho)\simeq \hbar^{4}\frac{\beta^{4}e^{4}}{240m^{2}},\;\;\;\beta\rightarrow 0 \label{4.44e} \end{equation} As $\beta\rightarrow 0$, the electrons tend to be fully ionized and the correlation becomes that found between individual charges. In the atomic limit, for $\delta>0$ not too large, one has \begin{equation} \lim_{\beta\rightarrow\infty,\rho\rightarrow 0}\beta^{-1}C(\beta,\rho)=-C_{{\rm van\; der\; Waals}} \label{4.44f} \end{equation} where $C_{{\rm van\; der\; Waals}}<0$ is the usual coefficient found in text books from a second order perturbation calculation of the residual Coulomb interaction between two hydrogen atoms in their ground states. A similar analysis has now been carried out in the full quantum mechanical electron-proton gas with the diagrammatic methods of chapter~\ref{chapter-loops}. One finds indeed that the coefficient of the $r^{-6}$ tail of the proton-proton correlation approaches the value $C_{{\rm van\; der\;Waals}}$ in the scaling limit where hydrogen atoms form \cite{AlCoMa98b}. In the medium, density and temperature effects modify the amplitude of the usual van der Waals potential, but not its $r^{-6}$ range. In particular, the free charges that are always present in the medium only partially screen the dipole interaction between atoms, contrary to a naive mean field description that would lead to an exponential screening of these interactions. The partial screening of protons is due in part to collective effects, and also to individual electrons that will bind with the proton. As the temperature goes to zero the former effects vanish, and the screening is eventually entirely due to the formation of a neutral bound state. Hence the atomic limit coincides with a zero density calculation with two preformed neutral atoms in empty space. In fact, the coefficient $C(\beta,\rho)$ interpolates continuously between the high temperature and atomic regimes. There is no qualitative difference between the mechanism producing the van der Waals forces between bound entities and that causing the $r^{-6}$ tail (\ref{4.43}) between individual charges. Both have their common origin in the same basic quantum fluctuations put into evidence in the preceding section. One sees that the difference is only quantitative. In the atomic regime, the coefficient $C(\beta,\rho)$ has a non-zero value close to $|C_{{\rm van\; der\; Waals}}|$ whereas it becomes vanishingly small at high temperature. \subsection{Semi-classical analysis of Coulombic correlations} \label{sec-4.5} We come now to the full quantum mechanical gas, but treated semi-classically as a whole. We use again the classical gas as a reference system by splitting the interaction between the filaments (\ref{4.17}) into \begin{equation} U({\cal F}_1,\ldots,{\cal F}_n)=U_0(\alpha_1,{\bf r}_1,\ldots,\alpha_n,{\bf r}_n)+ U_m({\cal F}_1,\ldots,{\cal F}_n) \label{4.45} \end{equation} where \begin{equation} U_0(\alpha_1,{\bf r}_1,\ldots,\alpha_n,{\bf r}_n)=\sum_{i<j}^n e_{\alpha_{i}}e_{\alpha_{j}}V({\bf r}_i-{\bf r}_j) \label{4.46} \end{equation} is the Coulomb energy of $n$ classical point particles and $U_m$ is the residual multipolar interaction. It is defined as the sum of two-body interactions $V_m({\cal F}_1,{\cal F}_2)=V({\cal F}_1,{\cal F}_2)-V({\bf r}_{1}-{\bf r}_{2})$ obtained from (\ref{4.18}) by removing the charge-charge potential: \begin{eqnarray} V_m({\cal F}_1,{\cal F}_2)&=&\int_0^1ds(M_1(s)+M_2(s))V({\bf r}_1-{\bf r}_2)\nonumber\\ &+&\int_0^1ds M_1(s) M_2(s)V({\bf r}_1-{\bf r}_2)\label{4.47}\\ U_m({\cal F}_1,\ldots,{\cal F}_n)&=&\sum_{i<j}^n e_{\alpha_{i}}e_{\alpha_{j}}V_m({\cal F}_i,{\cal F}_j)\label{4.48} \end{eqnarray} Since the Planck constant occurs only through the de Broglie lengths $\lambda_\alpha$, it is clear that $V_m$ and $U_m$ are power series in $\hbar$ starting with a linear term (see (\ref{4.5}) and (\ref{4.28a})). Thus considering $\hbar$ as a small parameter, we are left with the standard problem of perturbing the classical equilibrium state (in the framework of the system of filaments) by the small modification (\ref{4.48}) of the potential. This will generate formally power series in $\hbar$ for the distributions of filaments, and in turn, after averaging on their shapes, for the truncated distributions of the quantum charges \begin{equation} \rho_T(\alpha_1,{\bf r}_1,\ldots,\alpha_n,{\bf r}_n)=\sum_{k\geq 0} \hbar^k \rho_T^{(k)}(\alpha_1,{\bf r}_1,\ldots,\alpha_n,{\bf r}_n) \label{4.49} \end{equation} We shall not give here the rather complex algorithm for calculating the coefficients $\rho_T^{(k)}$ of this expansion, but only describe some qualitative aspects and state results. From now on we assume that the classical reference system is in the Debye phase: its correlations $\rho_{{\rm cl},T}(\alpha_1,{\bf r}_1,\ldots,\alpha_n,{\bf r}_n)$ cluster exponentially fast and obey the multipolar sum rule (\ref{2.2.5a}). By expanding around the classical gas, each $\rho_T^{(k)}$ involves combinations of convolutions of classical correlations with the multipole potentials coming from (\ref{4.47}), so that the $\rho_T^{(k)}$ are expressed in terms of the supposedly known correlations $\rho_{{\rm cl},T}$ of the reference system. The main point is that, although the classical correlations have an exponential fall-off, the law of formation of the $\rho_T^{(k)}$ is such that some of these convolutions decay only algebraically (there are two exceptions, namely if the potential is convoluted with the total charge density, then multipolar sum rules insure a fast decay, or if derivatives of the Coulomb potential are contracted to form a Laplacian $\nabla^2V({\bf r})$). The results are as follows \cite{AlMa88,AlMa89}. Obviously \begin{equation} \rho_T^{(0)}(\alpha_1,{\bf r}_1,\ldots,\alpha_n,{\bf r}_n)= \rho_{{\rm cl},T}(\alpha_1,{\bf r}_1,\ldots,\alpha_n,{\bf r}_n) \label{4.50} \end{equation} \begin{equation} \rho_T^{(k)}(\alpha_1,{\bf r}_1,\ldots,\alpha_n,{\bf r}_n)=0,\;\;\;\;k\; \mbox{odd} \label{4.51} \end{equation} This is because, according to (\ref{4.28a}), odd powers of $\hbar$ in the expansion (\ref{4.49}) correspond to odd moments of filaments. Such odd moments have zero Gaussian average. The coefficients $\rho_T^{(2)}(\alpha_1,{\bf r}_1,\ldots,\alpha_n,{\bf r}_n)$ are decreasing faster than any inverse power for large spatial separations of the arguments (for the two point function see \cite{Jan76}). At the order $\hbar^4$ the truncated two particle correlation behaves as $$ \langle \hat{\rho}(\alpha_{1},{\bf r}_{1})\hat{\rho}(\alpha_{2},{\bf r}_{2})\rangle^{(4)}_{T}= \rho_T^{(4)}(\alpha_1,{\bf r}_1,\alpha_2,{\bf r}_2)\phantom{---------} $$ \begin{equation} \sim \frac{2}{5}\(\frac{\beta^2}{24}\)^2\int d{\bf r} \,\int d{\bf r}^{\prime}\, f({\bf r}_1-{\bf r}_2+{\bf r}-{\bf r}^{\prime})G({\bf r},\alpha_1)G({\bf r}^{\prime},\alpha_2)\phantom{----} \label{4.51a} \end{equation} In (\ref{4.51a}) $f({\bf r})$ is the square of the dipole-dipole potential \begin{equation} f({\bf r})=\sum_{\mu,\nu=1}^3\(\partial_\mu\partial_\nu V({\bf r})\)^2 \label{4.52} \end{equation} and \begin{equation} G({\bf r},\alpha_i)=\sum_{\alpha} \frac{e_\alpha^2}{m_\alpha}\langle\hat{\rho}(\alpha,{\bf 0}) \hat{\rho}(\alpha_i,{\bf r})\rangle_{{\rm cl},T},\;\;\;\,i=1,2 \label{4.53} \end{equation} where $\langle\cdots\rangle_{{\rm cl}}$ means equilibrium average with respect to the classical reference system. Since $f({\bf r})\sim\frac{6}{|{\bf r}|^6}$, $|{\bf r}|\rightarrow\infty$, the leading contribution to (\ref{4.51a}) is \begin{equation} \langle \hat{\rho}(\alpha_{1},{\bf r}_{1})\hat{\rho}(\alpha_{2},{\bf r}_{2})\rangle^{(4)}_{T}\sim \frac{1}{|{\bf r}_1-{\bf r}_2|^6}\frac{\beta^4}{240} \(\int d{\bf r} \, G({\bf r},\alpha_1)\)\(\int d{\bf r} \, G({\bf r},\alpha_2)\) \label{4.54} \end{equation} If one examines the correlation of a particle with the total charge density $\hat{c}({\bf r})$, one must replace $\hat{\rho}(\alpha,{\bf r})$ in one factor $G$ by the charge $\hat{c}({\bf r})$. Then the coefficient of $|{\bf r}_1-{\bf r}_2|^{-6}$ vanishes in (\ref{4.54}) because of the charge sum rule (\ref{2.2.5a}), and one has to carry the asymptotic expansion of $f({\bf r})$ further on. For the particle-charge correlation, one finds \begin{eqnarray} \langle \hat{\rho}(\alpha_{1},{\bf r}_{1})\hat{c}(\alpha_{2},{\bf r}_{2})\rangle^{(4)}_{T}&=& \sum_{\alpha_{2}}e_{\alpha_{2}}\rho_T^{(4)}(\alpha_1,{\bf r}_1,\alpha_2,{\bf r}_2)\nonumber\\ &\sim&\frac{1}{|{\bf r}_1-{\bf r}_2|^8}\frac{\beta^4}{16}\(\int d{\bf r} \, G({\bf r},\alpha_1)\) \(\frac{1}{3}\int d{\bf r} \, |{\bf r}|^2K({\bf r})\)\nonumber\\ \label{4.55} \end{eqnarray} and for the charge-charge correlation \begin{eqnarray} \langle \hat{c}(\alpha_{1},{\bf r}_{1})\hat{c}(\alpha_{2},{\bf r}_{2})\rangle^{(4)}_{T}&=& \sum_{\alpha_1,\alpha_2}e_{\alpha_1}e_{\alpha_2}\rho_T^{(4)} (\alpha_1,{\bf r}_1,\alpha_2,{\bf r}_2)\nonumber\\&\sim&\frac{1}{|{\bf r}_1-{\bf r}_2|^{10}}\frac{7\beta^4} {4}\(\frac{1}{3}\int d{\bf r} \, |{\bf r}|^2K({\bf r})\)^2 \label{4.56} \end{eqnarray} with \begin{equation} K({\bf r})=\sum_\alpha\frac{e_\alpha^2}{m_\alpha} \langle\hat{\rho}(\alpha,{\bf 0})\hat{c}({\bf r})\rangle_{{\rm cl},T} \label{4.57} \end{equation} In fact, as discussed in sections~\ref{sec-6.1.2} and \ref{subsec-6.1.3}, the formulae (\ref{4.54}), (\ref{4.55}) and (\ref{4.56}) become asymptotically exact for the quantum Coulomb gas in the low density limit. In this limit the functions $G({\bf r})$ and $K({\bf r})$ are given by their Debye-H\"{u}ckel approximation\footnote{Note that the exchange contributions vanish exponentially fast with $\hbar$, so leading contributions in $\hbar$ can be obtained from Maxwell-Boltzmann statistics.}. The formulae simplify in the case of the jellium when one has only one species of charges. Then there is no distinction between particle and charge correlations up to a factor $e$ $(eG({\bf r})=K({\bf r})=\frac{e}{m}\langle\hat{c}({\bf}0)\hat{c}({\bf r})\rangle_{{\rm cl},T}$), so (\ref{4.56}) applies in all cases. Furthermore, in view of the second moment condition (\ref{1.1.1sti}), the correlations at order $\hbar^4$ in the jellium behaves as \begin{equation} \rho_T^{(4)}({\bf r}_1,{\bf r}_2)\sim\frac{7}{16\pi^2}\(\frac{\beta}{m}\)^2 \frac{1}{|{\bf r}_1-{\bf r}_2|^{10}},\;\;\;\;\;|{\bf r}_1-{\bf r}_2|\rightarrow\infty \label{4.58} \end{equation} For the jellium, \cite{GoLe94} have given the complete formula for $\rho_T^{(4)}({\bf r}_1,{\bf r}_2)$ at the order $\hbar^{4}$ including the short range contributions. Correlations involving more than two particles also have slow decays at order $\hbar^4$. For instance $\rho_T^{(4)}(\alpha_1,{\bf r}_1,\alpha_2,{\bf r}_2,\alpha_3,{\bf r}_3)$ decays also as $|{\bf r}|^{-6}$ when the particle 1 is sent to infinity, the other being kept fixed. If the three particles are taken simultaneously far apart (say, $|{\bf r}_1-{\bf r}_2|=r,\;|{\bf r}_1-{\bf r}_3|=ar,\;|{\bf r}_2-{\bf r}_3|=br,\;r \rightarrow\infty$), this function decays as $r^{-9}$. Moreover, the four point function $\rho_T^{(4)}(\alpha_1,{\bf r}_1,\alpha_2,{\bf r}_2,\alpha_3,{\bf r}_3, \alpha_4,{\bf r}_4)$ decays only as $r^{-3}$ when two pairs of particles are sent far away (say, $|{\bf r}_1-{\bf r}_2|,|{\bf r}_3-{\bf r}_4|$ fixed, $r=|{\bf r}_1-{\bf r}_3|\rightarrow\infty$). This slow decay is due to the breaking of rotational invariance around the points ${\bf r}_1$ and ${\bf r}_2$ when the directions ${\bf r}_1-{\bf r}_2$ and ${\bf r}_3-{\bf r}_4$ are fixed. It is precisely because of rotation invariance that the pure dipolar decay $r^{-3}$ is reduced to $r^{-6}$ in the two point function, as in the model of section \ref{sec-4.3}. We have reported on the semi-classical tails (\ref{4.54})-(\ref{4.56}) in some details with the purpose of illustrating the following points. There is definitely no exponential screening in the quantum gas, even in the high temperature phase and what ever the density may be. Moreover, in contrast with a fluid with short range forces, various types of decay laws have to be expected depending on what type of observables are concerned, and what cluster of particles are taken far apart. The occurrence of these different types of decay are due to the underlying multipolar forces displayed in the filament formalism in conjunction with the effect of screening sum rules (\ref{2.2.5a}). We finally discuss the possible observable implications of these findings. Since the calculations pertain to the semi-classical regime, we only consider real systems under conditions that the quantum effects are small. We determine the crossover distance $r_{0}$ above which quantum algebraic tails dominate the classical behavior by equating (\ref{4.58}) to the usual Debye law \begin{equation} \frac{\rho e^{- r_{0}/l_{D}}}{4\pi l_{D}^{2} r_{0}}=\frac{7}{16\pi^{2}}\left(\frac{\beta}{m}\right)^{2} \frac{\hbar^{4}}{r_{0}^{10}} \label{4.59} \end{equation} The classical exponential is normalized in order to obey the charge sum rule. For instance, for sodium chloride at room temperature, one finds $r_{0}\sim 60\;l_{D} $. For a white dwarf under prevailing conditions, one has $r_{0}\sim 40\;l_{D} $. In both cases, the quantum effects on the clustering turn out to be very small and the use of the Debye potential is still legitimate from a quantitative point of view. However if we apply crudely (\ref{4.59}) (disregarding its semi-classical nature) to the electrons in a metal (replacing $l_{D}$ by the Thomas-Fermi screening length $\lambda_{TF}$), one finds $r_{0}$ of the order of $\lambda_{TF}$. This indicates that the use of the exponential potential might be less reliable in this case. \subsection{Breakdown of exponential screening in the Quantum Sine-Gordon representation}\label{sec-4.6} In this section we show how the Sine-Gordon transformation of section~\ref{sec-sinegordon} extends to the quantum mechanical system \cite{FrPa78}. We saw in (\ref{2.3.4c}) that the classical partition function is a Gaussian superposition of ideal gas partition functions in external fields $i\phi ({\bf r} )$. The same holds for quantum mechanics except that one must use time dependent fields $i\phi ({\bf r},s)$. After the Sine-Gordon transformation is discussed we use it to illustrate the natural appearance of dipole forces in the quantum system and make some comparisons with classical systems of particles that simultaneously have charge and dipole moments. We consider a quantum Coulomb system as in (\ref{1.2.1}), but replace the Coulomb interaction (\ref{1.1.1}) by \begin{equation}\label{4.6.0} U_{\infty,L} = \frac{1}{2}\sum_{i,j}^N e_{\alpha_i} e_{\alpha_j} V_{\infty ,L}({\bf r} _i - {\bf r} _j ) \end{equation} where $V_{\infty ,L}$ is the potential defined in (\ref{1.1.3be}) whose singularity at the origin is smoothed out on length scale $L$. Note also that self-energies are included in $U_{\infty,L}$. This choice of interaction simplifies the Sine-Gordon transformation. It should be possible to reduce the $1/r$ to this case by a Mayer expansion as was done in the classical case in section~\ref{sec-debyesphere} because the Mayer expansion is known to converge \cite{BrFe76,BrFe77}. From (\ref{1.2.5}) with external field $\phi $ set to zero and using the Trotter product formula \begin{eqnarray} \Xi_{\Lambda} &=& \mbox{Tr} P\exp\( -\beta\( H_{\Lambda,N} - \underline{\mbox{\boldmath $\mu$\unboldmath}} \cdot \underline{\bf N} \) \) \nonumber\\ &=& \mbox{Tr} P \lim_{n \rightarrow \infty} \prod_1^n \( e^{-\frac{\beta}{n}[H_{\Lambda,N}^{0}- \underline{\mbox{\boldmath $\mu$\unboldmath}} \cdot \underline{\bf N}]} e^{-\frac{\beta}{n} U_{\infty ,L}} \) \label{4.6.1} \end{eqnarray} where the $H_{\Lambda,N}^{0}$ is the free Hamiltonian and $\underline{\mbox{\boldmath $\mu$\unboldmath}} \cdot \underline{\bf N} =\sum_{\alpha}\mu_{\alpha}N_{\alpha}$. We insert \begin{equation}\label{4.6.2} e^{-\frac{\beta}{n} U_{\infty ,L}} = \int d\mu _{L} (\phi) \, e^{ -i(\beta/n) ^{1/2} \int d {\bf r} \,\hat{c}({\bf r} ) \phi({\bf r} ) } \end{equation} Each factor $e^{-\frac{\beta}{n} U_{\infty ,L}}$ requires its own auxiliary field $\phi_i({\bf r})$, $i = 1, \ldots, n$, so that the right hand side of (\ref{4.6.1}) contains \begin{equation} \int d \mu_{L}(\phi_{1})\cdots d \mu_{L}(\phi_{n}) \, \mbox{Tr} P \prod_1^n \( e^{-\frac{\beta}{n} H_{\Lambda,N}^{0}} e^{ -(\beta /n)^{1/2} i \int d {\bf r} \,\hat{c}({\bf r} ) \phi_{j}({\bf r} ) } \) \end{equation} By scaling the fields $\phi_{j} \rightarrow n^{-1/2}\phi_{j}$ and uniting them into one time dependent field $\phi({\bf r},s) \equiv \phi_j({\bf r} )$ when $s \in [\frac{(j-1)}{n},\frac{j}{n})$, we obtain \begin{equation} \label{4.6.3} \Xi_{\Lambda} = \int d^{[0,1]}\mu_{L}(\phi) \, \Xi_{{\rm ideal},\Lambda}(\beta ^{1/2}i\phi) \end{equation} where $\int d^{[0,1]}\mu_{L}(\phi)$ is the Gaussian defined by \begin{equation} \label{4.6.4} \int d^{[0,1]}\mu_{L}(\phi) \, \phi({\bf r}, s ) \phi({\bf r} ',s ') = \delta (s - s ') V_{\infty ,L}({\bf r} - {\bf r} '); \ \ \ 0 \leq s \leq 1 \end{equation} and \begin{eqnarray}\label{4.6.5} && \Xi_{{\rm ideal},\Lambda}(\phi) = \mbox{Tr} P \exp\( - \int_{0}^{1} \( \beta H_{\Lambda,N}^{0} - \beta \underline{\mbox{\boldmath $\mu$\unboldmath}} \cdot \underline{\bf N} + \int d {\bf r} \,\hat{c}({\bf r} ) \phi({\bf r},s ) \) \, ds \) \end{eqnarray} is the ideal gas partition function for a time-dependent external field and the exponential is time-ordered. In the case of Boltzmann statistics, which means omitting the sum over permutations in the projection $P$, the trace factors over particles so that \begin{eqnarray} \label{4.6.6} && \Xi_{{\rm ideal, Bolt},\Lambda}(\beta\phi) = \sum_{\{N_\alpha\}} \prod _{\alpha }\frac{1}{N_{\alpha }!} \mbox{Tr} \nonumber \\ && \hspace{-1in} \times \exp\( - \beta\int_{0}^{1} \( H_{\Lambda,N}^{0} - \underline{\mbox{\boldmath $\mu$\unboldmath}} \cdot \underline{\bf N} + \int d {\bf r} \,\hat{c}({\bf r} ) \phi({\bf r},s ) \) \, ds \) \nonumber \\ && = \exp \left( \sum _{\alpha } \mbox{Tr} e^{ - \beta\int_{0}^{1} ds [h_{\alpha }(s)-\mu _{\alpha }] }\right) \end{eqnarray} where $h_{\alpha }(s) = \frac{\hbar^{2}}{2m_{\alpha }}(-\Delta) + e_{\alpha }\phi({\bf r}, s) $ is the one-particle Hamiltonian for species $\alpha $ in the time-dependent external field and the exponential is still time-ordered. Without making the simplification of Boltzmann statistics one has, instead, \begin{equation} \label{4.6.6b} \Xi_{{\rm ideal},\Lambda}(\beta\phi) = \exp \left( \sum _{\alpha } \sum _{q=1}^{\infty } \frac{ \eta_{\alpha }^{q-1}}{q} \mbox{Tr} e^{ - \beta\int_{0}^{q}ds [h_{\alpha }(s)-\mu _{\alpha }] }\right) \end{equation} where $\phi ({\bf r} ,s)$ is extended periodically in $s$. This formula will be proved in section~\ref{subsec-5.1.4}. Notice that this quantum ideal gas is invariant under \begin{equation}\label{4.6.6d} \phi({\bf r},\tau ) \rightarrow \phi({\bf r},\tau ) + f(\tau) \end{equation} where $f$ is any function independent of ${\bf r} $ such that $\int_0^1 d\tau f(\tau) = 0$. The Gaussian $d^{[0,1]}\mu$ is also formally invariant under this transformation. This ``Goldstone mode'' is a signal that there are long range correlations in the $\phi$ field, whereas in the classical system $\phi$ correlations are exponentially decaying. By the Feynman-Kac formula, \begin{eqnarray}\label{4.6.7} && \Xi_{{\rm ideal, Bolt},\Lambda}(\beta\phi) = \exp \left( \int d{\cal F} \, z({\cal F} ) e^{ - \beta e_{\alpha} \phi({\cal F} ) } \right)\nonumber \\ && \phi({\cal F} ) = \int_{0}^{1} \phi ({\bf r} + \lambda _{\alpha }\xi(s),s) \,ds \end{eqnarray} where ${\cal F}, d{\cal F} $ were defined in (\ref{4.12}, \ref{4.15}). There is a similar formula for the ideal gas (\ref{4.6.6b}) without the Boltzmann statistics simplification and it is given in section~\ref{subsec-5.1.4}. It could be used in the following considerations, but would not change the discussion in any important way. Furthermore we set the spins $s_{\alpha } = 0$. When $\lambda _{\alpha } = 0$ the phase space integration $d{\cal F} $ reduces to $\sum _{\alpha }\int d{\bf r} $ and (\ref{4.6.7}) reduces to the classical ideal gas. This suggests \cite{BrKe94} the following model. We replace the integration $\int d{\cal F} $ over all Wiener paths by integration concentrated on just one kind of path which oscillates about the initial point by a distance $O(\lambda _{\alpha })$ (the size of the wave packet) in a random direction: let $d\sigma(\vec{e})$ be a spherically symmetric integration on vectors $\vec{e}$. Then $\int d{\cal F} $ is replaced by \begin{equation} \sum _{\alpha } \int \, d{\bf r} \, \int \, d\sigma(\vec{e}) \label{4.6.8} \end{equation} and each $({\bf r} ,\vec{e})$ labels a path: ${\bf r} + \lambda _{\alpha } \xi (s)$ with $\xi(s) = \vec{e} \sin (2 \pi s)$. We chose $ \sin (2 \pi s)$ because it is orthogonal to $1$. The consequence of this and another minor approximation \cite{BrKe94} is that the dependence of (\ref{4.6.6}) on $\phi({\bf r} ,s) $ is only through two modes \[ \phi_1({\bf r} ) = \int_{0}^{1} \phi({\bf r} , s)\, ds; \ \ \ \phi_2({\bf r} ) = \sqrt{2} \int_{0}^{1} \phi({\bf r} ,s) \sin (2 \pi s) \, ds \] Since $\phi_1$, $\phi_2$ are Gaussian and $\int d^{[0,1]}\mu _{L} \, \phi_i({\bf r} ) \phi_j({\bf r} ') = V_{\infty ,L}({\bf r} -{\bf r} ') \delta_{ij}$, the integral $\int d^{[0,1]}\mu _{L} $ can be replaced by $\int d\mu _{L}(\phi _{1}) \, \int d\mu _{L} (\phi _{2})$ and the partition function becomes \begin{eqnarray} && \Xi_{\Lambda} = \int d\mu_{L}(\phi_2) \int d\mu_{L}(\phi_1) \nonumber \\ &&\hspace{-1in} \times \exp \( \sum _{\alpha } e^{\beta \mu_{\alpha } } \int d{\bf r} \, \int d\sigma(\vec{e}) \, e^{i\beta^{1/2} \phi_1 + i(\beta/2)^{1/2} \lambda _{\alpha }\vec{e}\cdot\nabla\phi_2}\) \label{4.6.9b} \end{eqnarray} Notice that if $d\sigma(\vec{e})$ is set to $\delta(\vec{e})$ we revert to the classical Coulomb gas. If $\phi_1$ is set to zero then by reversing the Sine-Gordon transformation we obtain the partition function of a classical dipole gas with dipole moments $\vec{e}$ distributed according to $d\sigma$. Furthermore part of the Goldstone mode (\ref{4.6.6b}) survives as \begin{equation}\label{4.6.6c} \phi _{2}({\bf r} ) \rightarrow \phi _{2}({\bf r} ) + {\rm const} \end{equation} If the fields $\phi = \phi _{1} = \phi _{2}$ are set equal, this becomes exactly the Sine-Gordon transform of a classical gas of particles that are simultaneously carrying charge and a dipole moment. {\it But in this case the symmetry (\ref{4.6.6c}) no longer holds and the methods of chapter~\ref{chapter- debye-screening} prove that correlations decay exponentially, near the Debye-H\"uckel limit.} Indeed consider the Debye-H\"uckel approximation wherein the exponent in (\ref{4.6.9b}) is replaced by the quadratic approximation \begin{eqnarray} && Q = -\frac{1}{2} \sum _{\alpha } e^{\beta \mu_{\alpha } } \int d{\bf r} \, \int d\sigma(\vec{e}) \,\( \beta^{1/2} \phi + (\beta/2)^{1/2} \lambda _{\alpha }\vec{e}\cdot\nabla\phi \) ^{2}\nonumber \\ =&& -\frac{1}{2} \sum _{\alpha } \beta e^{\beta \mu_{\alpha } } \int d{\bf r} \, \( \phi ^{2} + \frac{1}{6} (\lambda _{\alpha }\nabla\phi)^{2} \) \label{4.6.10} \end{eqnarray} The Gaussian $d\mu _{L}\exp \left(-Q \right)$ has a covariance whose Fourier transform is \begin{equation}\label{4.6.11} 4\pi \({\bf k} ^{2} + L^{2} {\bf k} ^{4} + l_{D}^{-2} + a {\bf k} ^{2}\)^{-1} \end{equation} where $l_{D}^{-2} = 4\pi \sum _{\alpha } \beta e^{\beta \mu_{\alpha }}$ and $a = 4\pi \sum _{\alpha } \beta \lambda _{\alpha }^{2} e^{\beta \mu_{\alpha }}$. There is no pole for any real ${\bf k} $ which indicates exponential decay of all correlations. It has been wrongly said \cite[section II G]{Mar88} that classical ions with structure might have weaker screening properties than pure charges. The present discussion shows that such systems have Debye screening near the Debye-H\"uckel limit. The model defined by (\ref{4.6.8}) and (\ref{4.6.9b}) exemplifies again that the breakdown of exponential screening is really due to quantum mechanics, manifested here by the existence of the two fields $\phi _{1}$ and $\phi _{2}$ (and more generally by the time-dependent field $\phi ({\bf r},s)$ in the Sine-Gordon representation). Reconsider indeed that $\phi _{1} \not = \phi _{2}$ are independent Gaussian fields. The quadratic approximation is misleading here because the cross-terms coupling $\phi _{1}, \phi _{2}$ are lost in this approximation but they are present at cubic and higher order in (\ref{4.6.9b}). In \cite{BrKe94} it is argued that these higher order couplings allow the long range correlations of $\phi _{2}$ to creep into the correlations of both $\phi _{1}$ and $\phi _{2}$. This gives an alternative but more qualitative derivation of the no-screening results. The massive zero frequency field $\phi_{1}$ mediates (parts of) the electrostatic potential $V_{{\rm cl}}$ defined in (\ref{4.25}) whereas the non-zero frequency field $\phi_{2}$ mediates (part of) the non-electrostatic part $W$ defined in (\ref{4.27}) that cannot be screened. All these discussions show that screening is destroyed because lack of commutativity of momentum and position bring time dependent phenomena in through the back door (e.g., the imaginary time ordering mentioned below (\ref{4.6.5})). It is in fact easy to prove that imaginary time dependent observables are not screened \cite{BrSe86}. We should also mention that another class of semiclassical models has been introduced, to deal with polarizability effects in dielectrics \cite{HoSt81b,BrHo88}. In these models, internal atomic degrees of freedom are treated quantum mechanically (e.g. harmonically bound pairs of charges) while the center of mass of atoms behave in a classical manner with pair-wise dipolar forces. These models capture physically interesting effects but do not incorporate all the effects stemming from the Coulomb potential (binding, ionization, collective screening) in an equally consistent and fundamental footing. Hopefully the loop formalism developed in chapter~\ref{chapter-loops} does provide a way to treat all these effects in a fully coherent way. \def\({\left(} \def\){\right)} \def{\bf r}{{\bf r}} \def\b\xi{\mbox{\boldmath $\xi$\unboldmath}} \def{\cal F}{{\cal F}} \def{\cal L}{{\cal L}} \def{\bf X}{{\bf X}} \def{\bf R}{{\bf R}} \newpage \section{The gas of charged loops} \setcounter{page}{1} \label{chapter-loops} \subsection{The statistical mechanics of loops} \label{sec-5.1} \subsubsection{The magic formula} \label{subsec-5.1.1} We are ready to develop the general Feynman-Kac representation of the Coulomb gas. The first step (section \ref{sec-5.1}) is to include the Fermi statistics of the electrons and the Bose or Fermi statistics of the nuclei. This can be done with some combinatorial effort by a rearrangement of the sums in the partition function (\ref{1.2.5}), leading to the "magic formula" (\ref{5.16}), that has a marvelous effect: as in the preceding section, the gas appears as a classical-like assembly of extended objects, the charged loops, and the diagrammatic techniques of classical statistical mechanics will apply. The magic formula has been known at least implicitly for a long time, starting from \cite{Gin65}, see also \cite{Gin71}. A combinatorial proof of it, following the lines of \cite{Gin65} can be found in \cite{Cor96a}. It has been also retrieved recently in \cite{HoSt94}. In subsection \ref{subsec-5.1.4} we give the main lines of a derivation that does not make use of combinatorial arguments. From now on, $V({\bf r}_{1}-{\bf r}_{2})=\frac{1}{|{\bf r}_{1}-{\bf r}_{2}|}$ is the exact Coulomb potential on all scales: a short range regularization is no more needed since the latter is henceforth provided by quantum mechanics that will be now fully taken into account. The rest of the chapter is devoted to the development of a Mayer diagrammatic for this system, following mainly \cite{Cor96a}. In section \ref{sec-5.2} the monopole and multipole interactions are singled out in close analogy to what was done in section~\ref{sec-4.2}, in order to isolate the part $W({\cal L}_1,{\cal L}_2)$ (\ref{5.34}) that will ultimately be responsible for the non-exponential decay. Then, following the classical treatment of Coulomb divergences, one sums the Coulomb chains, providing thus the screening length $\kappa^{-1}$ (\ref{5.47}) (a comparison of this screening length with that of the standard RPA theory is postponed to section \ref{subsec-6.1.4}). Finally in section \ref{subsec-5.2.3} the class of resummed diagrams is reorganized in prototype graphs involving four types of bonds. Prototype graphs are finite, decay at least as $r^{-3}$ at large separation and will serve as the starting point of low density calculations. The content of this chapter as well as the next one is mathematically formal since nothing is known about the asymptotic nature of the infinite series of prototype diagrams. We now come back to the partition function (\ref{1.2.5}). We first represent the matrix element in (\ref{1.2.5}) by a Feynman-Kac integral. Since $H_{\Lambda,N}$ is independent of the spins, the spin scalar product $\prod_{i=1}^N<\sigma_{\alpha _{p(i)}}|\sigma_{\alpha _{i}}>$ factorizes out in this matrix element, and according to (\ref{4.6}) its positional part is represented by the functional integral \begin{eqnarray} & &<\{{\bf r}_{p(i)}\}|\exp\(-\beta H_{\Lambda,N}\) |\{{\bf r}_i\}>= \prod_{i=1}^N\(\frac{1}{2\pi\lambda_{\alpha_{i}}^2}\)^{3/2}\exp\( -\frac{|{\bf r}_{p(i)}-{\bf r}_i|^2}{2\lambda_{\alpha_{i}}^2}\)\nonumber\\ &\times &\int\prod_{i=1}^N D(\b\xi_i)\exp\(-\frac{\beta}{2}\sum_{i\neq j}^N e_{\alpha_{i}}e_{\alpha_{j}}\int_0^1 dsV({\bf r}_{i,p(i)}(s)-{\bf r}_{j,p(j)}(s))\) \label{5.1} \end{eqnarray} We have kept the notation (\ref{4.4}) for a path with extremities ${\bf r}_{i,p(i)}(s=0)={\bf r}_i$ and ${\bf r}_{i,p(i)}(s=1)={\bf r}_{p(i)}$ \begin{equation} {\bf r}_{i,p(i)}(s)=(1-s){\bf r}_i +s{\bf r}_{p(i)}+\lambda_{\alpha_{i}}\b\xi_i(s) \label{5.2} \end{equation} The same remark made after (\ref{4.9}) concerning the implementation of the boundary conditions remains valid: all paths are constrained to stay inside $\Lambda$. The main observation to be made at this point is that by a rearrangement of the sums, the partition function (\ref{1.2.5}) can be formally written in the same classical form as (\ref{4.14}), provided that one introduces a suitably enlarged phase space. This rearrangement exploits the fact that any permutation $p_\alpha$ can uniquely be decomposed into a product of cycles. A cycle of length $q$, $1\:\leq q\:\leq N_\alpha$, is a subset of the $N_\alpha$ particle indices that are permuted among themselves in a cyclic way under $p_\alpha$, for instance \begin{equation} (1,\,2,\ldots,q-1,\,q)\longrightarrow (2,\,3,\ldots,q,\,1) \label{5.3} \end{equation}The objects constituting the enlarged phase space are precisely associated with such cycles in the following way. The permutations $p$ in (\ref{5.1}) involve now open paths (except for the trivial one). However the set of open paths belonging to a cycle of $q$ elements occurring in $p$ can be joined together to form a closed loop ${\cal L}$ (sometimes also called a polymer) with $q$ particle coordinates. The loop ${\cal L}$ corresponding to the cycle (\ref{5.3}) is made of the collection of paths $\{{\bf r}_{k,k+1}(s)\}_{k=1,\ldots,q}$, with $q+1$ identified to $1$ (see Figure~\ref{fig1}). \noindent Thus a loop is specified by the variables \begin{equation} \(\alpha,\;q,\;\{{\bf r}_k\}_{k=1,\ldots,q},\;\{\b\xi_k\}_{k=1,\ldots,q}\) \label{5.4} \end{equation} It is convenient to unite the $q$ filaments constituting the loop into a single path parameterized by a time parameter $s$ running from $0$ to $q$ setting \begin{equation} {\bf R}(s)={\bf r}_{k,k+1}(\tilde{s}),\;\;\;k=[s]+1,\;\,\;0\leq\,s\,\leq q \label{5.5} \end{equation} where $[s]$ = integer part of $s$ and $\tilde{s}=s\mbox{(mod 1)}=s-[s]$. One can then locate the loop at ${\bf R}$ by selecting the position of one of the particles, say ${\bf r}_1={\bf R}$, and write \begin{equation} {\bf R}(s)={\bf R}+\lambda_\alpha{\bf X}(s),\:\:\; {\bf X}(0)={\bf X}(q) =0 \label{5.6} \end{equation} Then the particles are located at the points \begin{equation} {\bf r}_{k}={\bf R}+\lambda_{\alpha}{\bf X}_{k},\;\;\,\;{\bf X}_{k}={\bf X}(k-1) \label{5.10} \end{equation} In (\ref{5.6}), ${\bf X}(s)$ is the shape of the loop ${\cal L}$ at ${\bf R}$, as $\b\xi (s)$ was describing the shape of a filament at ${\bf r}$: it is again a dimensionless Brownian bridge starting and returning to the origin within the time $q$. Its normalized Gaussian measure $D({\bf X})$ is the composition of that of the $q$ open filaments that constitute the loop (see (\ref{4.6})) \begin{equation} D({\bf X})= ( 2\pi q)^{3/2}\prod_{k=1}^{q}\frac{\exp\(-\frac{1}{2}({\bf X}_{k+1}-{\bf X}_{k})^{2}\)}{(2\pi)^{3/2}}d{\bf X}_{2}\ldots d{\bf X}_{q} D( \b\xi_1)\ldots D(\b\xi_q )\\ \label{5.7} \end{equation} where one sets ${\bf X}_{1}={\bf X}_{q+1}=0$. One can calculate its covariance from the definitions (\ref{5.5})-(\ref{5.6}) \begin{equation} \int D({\bf X }) X_{\mu} (s_{1}) X_{\nu}(s_{2})=\delta_{\mu\nu}q\left[\min\left(\frac{s_{1}}{q},\frac{s_{2}}{q}\right) -\frac{s_{1}}{q}\frac{s_{2}}{q}\right] \label{5.8} \end{equation} Thus one can think of a loop ${\cal L}$ either as a set of particles coordinates (as in (\ref{5.4})) or again as a single Brownian path, setting \begin{equation} {\cal L}=(\alpha,\;\;q,\;\;{\bf R},\;\;{\bf X}(s),\;\;0\leq\;s\;\leq q) \label{5.9} \end{equation} It is now possible, generalizing also (\ref{4.16})-(\ref{4.18}), to define activities and interactions of loops. The interaction between two loops ${\cal L}_i,i=1,2,$ is the sum of those of the particles constituting the two loops $$ V({\cal L}_1,{\cal L}_2)=\sum_{k_{1}=1}^{q_{1}}\sum_{k_{2}=1}^{q_{2}}\int_0^1ds V\({\bf r}_{1,k_{1},k_{1}+1}(s)-{\bf r}_{2,k_{2},k_{2}+1}(s)\) $$ \begin{equation} =\int_0^{q_{1}}ds_{1} \, \int_0^{q_{2}}ds_{2} \,\delta({\tilde s}_1-{\tilde s}_2)V({\bf R}_1+\lambda_{\alpha_1}{\bf X}_{1}(s_1)-{\bf R}_2-\lambda_{\alpha_2}{\bf X}_{2}(s_2)) \label{5.11} \end{equation} and for $n$ loops \begin{equation} U({\cal L}_1,\ldots,{\cal L}_n)=\sum_{1=i<j}^n e_{\alpha_{i}} e_{\alpha_{j}} V({\cal L}_i,{\cal L}_j) \label{5.12} \end{equation} The activity $z({\cal L})$ of a loop incorporates the effects of quantum statistics, the spin degeneracy as well as the internal interaction $U({\cal L})$ of the particles in the same loop \begin{equation} z({\cal L})=(2s_\alpha+1)\frac{(\eta_\alpha)^{q-1}}{q}\; \frac{\exp(\beta\mu_\alpha q)}{(2\pi q\lambda_\alpha^2)^{3/2}} \;\exp(-\beta U({\cal L})) \label{5.13} \end{equation} with \begin{eqnarray} U({\cal L})&=&\frac{e_\alpha^2}{2}\sum_{k\neq k^\prime}^q\int_0^q ds V\({\bf r}_{k,k+1}(s)-{\bf r}_{k^\prime,k^\prime+1}(s)\)\nonumber\\ &=&\frac{e_{\alpha}^{2}}{2}\int_0^{q} ds_1\int_o^{q}ds_2\(1-\delta_{[s_{1}],[s_{2}]}\)\delta({\tilde s}_1-{\tilde s}_2)V(\lambda_\alpha({\bf X}(s_1)-{\bf X}(s_2)))\nonumber\\ \label{5.14} \end{eqnarray} Clearly, all the above quantities are invariant under cyclic permutations of the particles in the same loop. Notice also that particles in the same loop have the same charge; thus their mutual Coulomb interactions are positive and this implies the bound \begin{equation} |z({\cal L})|\leq \frac{2s_{\alpha}+1}{q}\;\frac{\exp(\beta\mu_\alpha q)}{(2\pi q\lambda^2_\alpha )^{3/2}} \label{5.15} \end{equation} Considering now the phase space of loops together with the above definitions, it can be shown (see section~\ref{subsec-5.1.4}) that the partition function can be written in the same form as (\ref{4.14}) ({\em the magic formula}) \begin{equation} \Xi_\Lambda=\sum_{n=0}^\infty\frac{1}{n!}\int\prod_{i=1}^nd{\cal L}_iz({\cal L}_i) \exp(-\beta U({\cal L}_1,\ldots,{\cal L}_n)) \label{5.16} \end{equation} where the phase space integration means here \begin{equation} \int d{\cal L}\cdots=\sum_{\alpha=1}^{\cal S}\sum_{q=1}^\infty \int_\Lambda d{\bf R}\int D({\bf X})\cdots \label{5.16a} \end{equation} We call the system defined by the relations (\ref{5.8})-(\ref{5.16a}) the "system of loops"\footnote{Our definitions of the weight of a loop (\ref{5.7}), the activity (\ref{5.13}) and the self-energy (\ref{5.14}) differ slightly from those found in \cite{Cor96a}. It is natural to define the loop measure (\ref{5.7}) again as that of the Brownian bridge process in time $[0,q]$.}. It reduces to the system of filaments (i.e. charges with Maxwell-Boltzmann statistics) if all loops with more than one particle are omitted. This view of the quantum Coulomb system as a gas of loops is well-suited in the low density regime, when loops remain well localized objects. \subsubsection{Loop and particle correlations} \label{subsec-5.1.2} It is clear that the statistical mechanical system of loops defined by the above relations has a classical structure. Therefore, introducing the "$\delta$-function" which identifies two loops, \begin{equation} \delta({\cal L}_1,{\cal L}_2)=\delta_{\alpha_1,\alpha_2}\delta_{q_1,q_2}\delta({\bf R}_1-{\bf R}_2)\delta({\bf X}_1,{\bf X}_2) \label{5.17} \end{equation} one can define the density of loops $\rho({\cal L})$ and the two loop distribution function $\rho({\cal L}_1,{\cal L}_2)$ by the formulae (\ref{4.20})-(\ref{4.21}) with ${\cal L}$ replacing ${\cal F}$. In order to obtain the particle distributions from those of the loops, some care has to be exercised because loops are constituted of several particles. For a configuration of particles distributed on a set of loops ${\cal L}_i$, we can write the particle density and distributions as a summation on the loop index $i$ \begin{equation} \hat{\rho}(\alpha,{\bf r})=\sum_{i}\delta_{\alpha_{i},\alpha}\sum_{k=1}^{q_i}\delta({\bf r}_{(k,i)} -{\bf r}) \label{5.18} \end{equation} where, according to (\ref{5.10}), ${\bf r}_{(k,i)}={\bf R}_i+\lambda_{\alpha_i}{\bf X}_{i}(k-1),\;\;\;k=1,\ldots,q_i$, are the positions of the particles belonging to the loop $i$. For ${\bf r}_{a}\neq{\bf r}_{b}$ one writes also \begin{eqnarray} \hat{\rho}(\alpha_a,{\bf r}_a)\hat{\rho}(\alpha_b,{\bf r}_b)&=& \sum_{i\neq j}\delta_{\alpha_{i},\alpha_{a}} \delta_{\alpha_{j},\alpha_{b}}\sum_{k=1}^{q_i}\sum_{k^\prime=1}^{q_j}\delta({\bf r}_{(k,i)}-{\bf r}_a) \delta({\bf r}_{(k^\prime,j)}-{\bf r}_b)\nonumber\\ &&\hspace{-.75in}+\delta_{\alpha_a,\alpha_b}\sum_i\delta_{\alpha_{i},\alpha_a}\sum_{k\neq k^\prime}^{q_{i}} \delta({\bf r}_{(k,i)}-{\bf r}_a)\delta({\bf r}_{(k^\prime,i)}-{\bf r}_b) \label{5.19} \end{eqnarray} In (\ref{5.18}) and (\ref{5.19}), $k$ ($k^\prime$) runs on particles belonging to the same loop. The first term in (\ref{5.19}) refers to particles in different loops and the second term to particles within the same loop. Then a comparison between the density of loops $\hat{\rho}({\cal L})=\sum_i\delta({\cal L}_i,{\cal L})$ and the particle density (\ref{5.18}) shows that \begin{equation} \hat{\rho}(\alpha,{\bf r})=\sum_{q=1}^\infty\int d{\bf R} \int D({\bf X})\sum_{k=1}^q \delta({\bf r}_k-{\bf r}) \hat{\rho}({\cal L}) \label{5.20} \end{equation} Taking the average of the above relation on the ensemble of loops gives the number density \begin{equation} \rho(\alpha,{\bf r})=\sum_{q=1}^\infty q\int D({\bf X})\rho(\alpha,q,{\bf r},{\bf X}) \label{5.21} \end{equation} where $\rho (\alpha ,q,{\bf r} ,{\bf X} ) = \rho ({\cal L})$ with ${\cal L} = (\alpha,q,{\bf r} ,{\bf X} )$. We have used the fact that $\rho({\cal L})$ is invariant under cyclic permutations of the $q$ particles in the loop. The two particle distributions are obtained by averaging the relation (\ref{5.19}) where we distinguish again between pairs belonging to different loops or exchanged in the same loop (${\bf r}_{a}\neq{\bf r}_{b}$) \begin{equation} \rho(\alpha_a,{\bf r}_a,\alpha_b,{\bf r}_b)=\rho^{(nex)}(\alpha_a,{\bf r}_a,\alpha_b,{\bf r}_b)+ \delta_{\alpha_a,\alpha_b}\rho^{(ex)}(\alpha_a,{\bf r}_a,{\bf r}_b) \label{5.22} \end{equation} with \begin{eqnarray} \label{5.23} &&\rho^{(nex)}(\alpha_a,{\bf r}_a,\alpha_b,{\bf r}_b)=\nonumber \\ && \hspace{-1in}\sum_{q_{a}=1}^\infty\sum_{q_{b}=1}^\infty q_a q_b\int D({\bf X}_a)\int D({\bf X}_a)\rho(\alpha_a,q_a,{\bf r}_a,{\bf X}_a;\alpha_b,q_b,{\bf r}_b,{\bf X}_b) \end{eqnarray} and \begin{equation} \rho^{(ex)}(\alpha_a,{\bf r}_a,{\bf r}_b)=\sum_{q=2}^{\infty}\int d{\bf R} \int D({\bf X}) \sum_{k\neq k^\prime}^{q}\delta({\bf r}_k-{\bf r}_a)\delta({\bf r}_{k^{\prime}}-{\bf r}_b)\rho(\alpha_{a},q, {\bf R},{\bf X}) \label{5.23a} \end{equation} For a translation invariant state ($\rho(\alpha,q, {\bf R},{\bf X})$ independent of ${\bf R}$), one has, using (\ref{5.10}) \begin{eqnarray} \rho^{(ex)}(\alpha_a,{\bf r}_a,{\bf r}_b)&=&\sum_{q=2}^\infty\int D({\bf X})\sum_{k\neq k^\prime}^{q} \delta(\lambda_{\alpha_{a}}({\bf X}_{k}-{\bf X}_{k^{\prime}})+{\bf r}_{b}-{\bf r}_{a})\rho(\alpha_{a},q,{\bf X}) \nonumber\\ &=&\sum_{q=2}^\infty\int D({\bf X})q\sum_{k=2}^{q} \delta(\lambda_{\alpha_{a}}{\bf X}_{k}+{\bf r}_{b}-{\bf r}_{a})\rho(\alpha_{a},q,{\bf X}) \label{5.23b} \end{eqnarray} The second line follows again from the invariance of $\rho(\alpha_{a},q,{\bf X})$ under cyclic permutations of the particles in the loop. One can also write the corresponding similar representations for the truncated distributions. \subsubsection{Loop interactions} The same fundamental observation made for the system of filaments remains valid here: the interaction (\ref{5.11}) is not the standard electrostatic Coulomb energy of two charged loops, which would be \begin{equation} V_{{\rm cl}}({\cal L}_1,{\cal L}_2)=\int_0^{q_{1}}ds_1\int_0^{q_{2}}ds_2 V({\bf R}_1+\lambda_{\alpha_{1}}{\bf X}_1(s_1)-{\bf R}_2-\lambda_{\alpha_{2}}{\bf X}_2(s_2)) \label{5.25} \end{equation} and the discussion following (\ref{4.25}) can be reproduced word by word. In particular, the Coulomb potential has the multipolar expansion $$ V({\bf R}_1+\lambda_{\alpha_{1}}{\bf X}_1(s_1)-{\bf R}_2-\lambda_{\alpha_{2}}{\bf X}_2(s_2))= V({\bf R}_1-{\bf R}_2)+ $$ \begin{equation} M_1(s_1)V({\bf R}_1-{\bf R}_2)+ M_2(s_2)V({\bf R}_1-{\bf R}_2)+ M_1(s_1)M_2(s_2)V({\bf R}_1-{\bf R}_2) \label{5.26} \end{equation} with now \begin{equation} M_i(s)=\sum_{k=1}^{\infty}\frac{(\lambda_{\alpha_{i}}{\bf X}_{i}(s)\cdot\nabla_{{\bf R}_i})^k}{k!},\;\;\;\;i=1,2 \label{5.27} \end{equation} and hence the loop interaction can be decomposed into its charge-charge, charge-multipole and multipole-multipole components \begin{equation} V({\cal L}_1,{\cal L}_2)=V_{cc}({\cal L}_1,{\cal L}_2)+V_{cm}({\cal L}_1,{\cal L}_2)+V_{mc}({\cal L}_1,{\cal L}_2)+V_{mm}({\cal L}_1,{\cal L}_2) \label{5.28} \end{equation} \begin{eqnarray} V_{cc}({\cal L}_1,{\cal L}_2)&=&q_1 q_2 V({\bf R}_1-{\bf R}_2)\label{5.29}\\ V_{cm}({\cal L}_1,{\cal L}_2)&=&q_1\int_0^{q_{2}}ds_2 M_2(s_2)V({\bf R}_1-{\bf R}_2)\label{5.30}\\ V_{mc}({\cal L}_1,{\cal L}_2)&=&q_2\int_0^{q_{1}}ds_1 M_1(s_1)V( {\bf R}_1- {\bf R}_2)\label{5.31}\\ V_{mm}({\cal L}_1,{\cal L}_2)&=&\int_0^{q_{1}}ds_1\int_0^{q_{2}}ds_2\delta(\tilde{s}_1-\tilde{s_2}) M_1(s_1)M_2(s_2)V({\bf R}_1-{\bf R}_2)\nonumber\\ \label{5.32} \end{eqnarray} The relevant quantity which will again embody the long range effects is the difference between the quantum and classical interaction \begin{equation} W({\cal L}_1,{\cal L}_2)=e_{\alpha_{1}}e_{\alpha_{2}}(V({\cal L}_1,{\cal L}_2)-V_{{\rm cl}}({\cal L}_1,{\cal L}_2)) \label{5.33} \end{equation} It depends only on the multipole-multipole part of the interaction as in (\ref{4.30}) \begin{eqnarray} W({\cal L}_1,{\cal L}_2)&=&e_{\alpha_{1}}e_{\alpha_{2}} \int_0^{q_{1}}ds_1\int_0^{q_{2}}ds_2(\delta(\tilde{s}_1-\tilde{s_2})-1) \nonumber \\ &&\hspace{.5in}\times M_1(s_1)M_2(s_2)V({\bf R}_1-{\bf R}_2) \label{5.34} \end{eqnarray} and has the asymptotic dipolar character (\ref{4.31}) as $|{\bf R}_1-{\bf R}_2|\rightarrow\infty$\begin{eqnarray} &&\hspace{-1in}W({\cal L}_1,{\cal L}_2) \sim e_{\alpha_{1}}e_{\alpha_{2}} \int_0^{q_{1}}ds_1\int_0^{q_{2}}ds_2\nonumber \\ &&\hspace{-1in} \times (\delta(\tilde{s}_1-\tilde{s}_2)-1) (\lambda_{\alpha_{1}}{\bf X}_1(s_1)\cdot\nabla_{{\bf {\bf R}}_{1}})(\lambda_{\alpha_{2}} {\bf X}_{2}(s_{2})\cdot\nabla_{{\bf R}_{2}}) V({\bf R}_1-{\bf R}_2) \label{5.35} \end{eqnarray} \subsubsection{Derivation of the magic formula } \label{subsec-5.1.4} We derive first the formula (\ref{5.16}) in the special case where there is an external field $\phi ({\bf r})$ and no two-body interaction. In this case the magic formula is essentially the same as (\ref{4.6.6b}). It is enough to consider just one species with charge $e_{\alpha }$ and spin $s_{\alpha }$ because the partition function factors. In terms of annihilation and creation operators $\hat{a}_{\sigma}({\bf r} ), \hat{a}_{\sigma}^{\ast}({\bf r} )$ the many-body Hamiltonian for the external field case is \begin{equation}\label{5.1.4.0} H_{\alpha } = \sum _{\sigma} \int \hat{a}_{\sigma}^{\ast}({\bf r} ) h_{\alpha } \hat{a}_{\sigma}({\bf r} ) \,d{\bf r}; \ \ \ h_{\alpha } = \frac{\hbar^{2}}{2m_{\alpha }}(-\Delta) + e_{\alpha } \phi \end{equation} We begin with the averaged particle number \begin{eqnarray}\label{5.1.4.1} \langle N\rangle&=&\frac{d}{d(\beta \mu) } \ln \Xi_\Lambda = \frac{1}{ \Xi_\Lambda} \sum _{\sigma} \int \,d{\bf r} \, \mbox{Tr} \left(\hat{a}_{\sigma}^{\ast } ({\bf r} ) \hat{a}_{\sigma}({\bf r} ) e^{-\beta( H_{\alpha }-\mu N)} \right)\nonumber\\ N&=&\sum _{\sigma} \int \hat{a}_{\sigma}^{\ast}({\bf r} ) \hat{a}_{\sigma}({\bf r} ) \,d{\bf r};\;\;\;\;\Xi_\Lambda=\mbox{Tr} e^{-\beta( H_{\alpha }-\mu N)} \end{eqnarray} The canonical commutation relations for the annihilation and creation operators $\hat{a}_{\sigma}({\bf r} ), \hat{a}_{\sigma}^{\ast } ({\bf r} )$ imply that \begin{equation}\label{5.1.4.2} \hat{a}_{\sigma}({\bf r} ) e^{-\beta H_{\alpha }} = e^{-\beta H_{\alpha }} \int e^{-\beta h_{\alpha }} ({\bf r} ,{\bf r} ')\hat{a}_{\sigma}({\bf r} ') \, d{\bf r} '\nonumber \\ \end{equation} where $\exp (- \beta h_{\alpha } )({\bf r} ,{\bf r} ' )$ is the kernel of the operator $\exp (-\beta h_{\alpha })$. Inserting this relation in (\ref{5.1.4.1}), using the cyclicity of the trace and the canonical commutation relations gives \begin{eqnarray}\label{5.1.4.5a} \frac{d}{d(\beta \mu) } \ln \Xi_\Lambda &=& \sum _{q=1}^{\infty } \eta^{q-1} \mbox{Tr} e^{-q \beta [h_{\alpha } - \mu ]} \end{eqnarray} where $\mbox{Tr}$ is now the one-particle trace and it includes the sum over spin. In the case of zero external field this is the standard formula for the density of an ideal Bose or Fermi gas by summing the geometric series. Integrating (\ref{5.1.4.5a}) from $-\infty$ to $\beta\mu$ yields \begin{equation}\label{5.1.4.6} \Xi_\Lambda = \exp \left( \sum _{q=1}^{\infty } \frac{\eta^{q-1}}{q} \mbox{Tr} e^{- q \beta [h_{\alpha } - \mu ]} \right) \end{equation} If we replace $q$ in the exponent by $\int_{0}^{q}\dots ds$ and time-order the exponential then this formula and our derivation remain valid for a periodic time-dependent external field $\phi ({\bf r} ,s)$, and we obtain (\ref{4.6.6b}). The ``magic formula'' in the case of external field, no two-body interaction and one particle species follows by substituting in the Feynman-Kac representation \begin{eqnarray}\label{5.1.4.5b} \sum _{q=1}^{\infty } \frac{\eta^{q-1}}{q} \mbox{Tr} e^{- \beta\int_{0}^{q} [h_{\alpha }(s) - \mu ] \,ds} &=& \int d{\cal L} \, z({\cal L}) e^{-\beta e_{\alpha}\phi ({\cal L})} \nonumber \\ \phi ({\cal L}) = \int_{0}^{q} \phi({\bf R} (s),s) \,ds \end{eqnarray} where we use the notation described in subsection~\ref{subsec-5.1.1}. The ``magic formula'' (\ref{5.16}) in the case of two-body interactions follows by integrating over imaginary time-dependent fields $i\phi ({\bf r} ,s)$ with a Gaussian measure. According to our discussion of the quantum Sine-Gordon transform in section~\ref{sec-4.6} this reconstructs the two-body interaction with self-energies. The formulas (\ref{5.13},\ref{5.14}) follow by absorbing the self-energies into the activities. \subsection{The loop Mayer expansion} \label{sec-5.2} \subsubsection{The loop Ursell function} Since the system of loops has a classical structure, it can be treated by the well known methods of classical statistical mechanics. In this section we apply the standard techniques of the virial expansion to the two-point Ursell function $h({\cal L}_a,{\cal L}_b)$ together with the partial resummations needed to deal with the long range of the Coulomb potential. The Ursell function is defined in terms of the two-loops truncated distribution $\rho_T({\cal L}_a,{\cal L}_b)$ by \begin{equation} \rho_T({\cal L}_a,{\cal L}_b)=\rho({\cal L}_a)\rho({\cal L}_b)h({\cal L}_a,{\cal L}_b) \label{5.36} \end{equation} Although loops are extended objects, we still call them "points" and represent them graphically by points in the diagrammatic language. The non-integrated loop variables ${\cal L}_a$, ${\cal L}_b$ are called root points. We recall that the Ursell function has the following simple loop-density expansion \cite{HaMc76} \begin{equation} h({\cal L}_a,{\cal L}_b)=\sum_\Gamma\frac{1}{S_\Gamma}\int\prod_{n=1}^Nd{\cal L}_n\rho({\cal L}_n) [\prod f]_\Gamma \label{5.37} \end{equation} In (\ref{5.37}), the sum runs over all unlabeled topologically different connected diagrams $\Gamma$ with two root points ${\cal L}_a$, ${\cal L}_b$ and $N$ internal points $N=0,1,2,\ldots$, without articulations (an articulation point is such that, when removed, the diagram splits into two pieces, at least one of which is disconnected from any root point). To each bond $({\cal L}_i, {\cal L}_j)$ in the diagram is associated the Mayer factor (called f-bond) \begin{equation} f({\cal L}_i, {\cal L}_j)=\exp(-\beta_{i,j}V({\cal L}_i, {\cal L}_j))-1,\;\,\;\;\beta_{i,j}=\beta e_{\alpha_{i}}e_{\alpha_{j}} \label{5.38} \end{equation} and $[\prod f]_\Gamma$ is the product of all f-bonds in the $\Gamma$-diagram; the symmetry factor $S_\Gamma$ is the number of permutations of internal points ${\cal L}_n$ that leave this product invariant. It is clear from (\ref{5.28}) and (\ref{5.29}) that the loop potential $V({\cal L}_i, {\cal L}_j)$ is Coulombic at large distances, so that the bond (\ref{5.38}) is not integrable. In the classical case it is known that by resumming appropriate classes of graphs (the convolution chains leading to the Debye-H\"{u}ckel potential), one can introduce new bonds and diagrams (the prototype diagrams) that are free of divergences. One applies here the same procedure to the quantum gas in the loop formalism. The main guidance comes from the fact that the pure monopole part (\ref{5.29}) would lead to Debye screening. Therefore the subsequent operations aim at disentangling the monopole from the multipolar effects, and then extracting as much classical screening as possible in order to reduce the range of the interaction. When all these steps are completed, no bond will decay slower than $|{\bf r}|^{-3}$, thus converting the gas of Coulombic loops into a certain gas of dipoles (together with higher order multipoles). \subsubsection{Summing the Coulomb chains} \label{subsec-5.2.2} The basic f-bond (\ref{5.38}) is split into the sum of five contributions according to \begin{equation} f=f_T+f_{cc}+f_{cm}+f_{mc}+f_{mm} \label{5.39} \end{equation} with the following analytical and graphical definitions (Figure~\ref{fig2}) (abbreviating\\ $f({\cal L}_i,{\cal L}_j)=f(i,j)$) \begin{eqnarray} f_T(i,j)&=&f(i,j)+\beta_{i,j}V(i,j) \label{5.40}\\ f_{cc}(i,j)&=&-\beta_{i,j}V_{cc}(i,j) \label{5.41}\\ f_{cm}(i,j)&=&-\beta_{i,j}V_{cm}(i,j) \label{5.42}\\ f_{mc}(i,j)&=&-\beta_{i,j}V_{mc}(i,j) \label{5.43}\\ f_{mm}(i,j)&=&-\beta_{i,j}V_{mm}(i,j) \label{5.44} \end{eqnarray} \noindent Then, the Ursell function is given by the same formula as (\ref{5.37}), summing now diagrams $\tilde{\Gamma}$ with the same topological structure as the $\Gamma$-diagrams, but where the bonds can be any one of the quantities (\ref{5.40})-(\ref{5.44}). Points without arrows in a $\tilde{\Gamma}$-diagram and to which are attached only two dashed lines are called Coulomb points (Figure~\ref{fig3}). A Coulomb point involves only interactions with the total charge $qe_\alpha$ of the loop (considered as a point like object) without multipolar effects. A Coulomb chain is a convolution chain where all intermediate points are Coulomb points. There are four types of Coulomb chains associated with the four bonds (\ref{5.41})-(\ref{5.44}) (Coulomb chains without intermediate points are defined as the bonds themselves). As in the classical theory, we sum all Coulomb chains of a given type, and denote these sums $F_{cc},F_{cm}, F_{mc},F_{mm}$ (Figures~\ref{fig4} and \ref{fig5}). One has \begin{eqnarray} F_{cc}({\cal L}_a,{\cal L}_b)&=&f_{cc}({\cal L}_a,{\cal L}_b)\nonumber\\ &+&\sum_{N=1}^\infty\int\prod_{i=1}^{N}d{\cal L}_i\rho({\cal L}_i)f_{cc}({\cal L}_a,{\cal L}_1)f_{cc}({\cal L}_1,{\cal L}_2) \cdots f_{cc}({\cal L}_N,{\cal L}_b)\nonumber\\ \label{5.44a} \end{eqnarray} Since by (\ref{5.29}) $f_{cc}({\cal L}_i,{\cal L}_j)=-\beta_{i,j}q_iq_jV({\bf R}_i-{\bf R}_j)$ does not depend on the configurations ${\bf X}_i,{\bf X}_j$ of the loops, the summation on the internal variables can immediately be performed, giving a factor $\sum_{\alpha}\sum_{q}\int D({\bf X})q^{2}e_{\alpha}^{2}\rho(\alpha,q,{\bf X})$ for each intermediate point in (\ref{5.44a}). Because of translation invariance the loop-density $\rho({\cal L})=\rho(\alpha,q,{\bf X})$ does not depend on the location ${\bf R}$ of the loop, therefore the multiple convolution (\ref{5.44a}) in the ${\bf R}$ variables sum up as in the classical chains to \begin{equation} F_{cc}(a,b)=-\beta_{a,b}q_{a}q_{b}\Phi({\bf R}_a-{\bf R}_b) \label{5.45} \end{equation} where $\Phi({\bf R})$ is a Debye-like potential \begin{equation} \Phi({\bf R})=\frac{\exp(-\kappa |{\bf R}|)}{|{\bf R}|} \label{5.46} \end{equation} with inverse screening length $\kappa$ given by \begin{equation} \kappa^2=4\pi\beta \sum_{\alpha}\sum_{q}\int D({\bf X})q^{2}e_{\alpha}^{2}\rho(\alpha,q,{\bf X}) \label{5.47} \end{equation} From the definitions (\ref{5.30})-(\ref{5.31}) and (\ref{5.42})-(\ref{5.43}), the sum of Coulomb chains corresponding to the charge-multipole bonds are obtained from (\ref{5.45}) by application of the multipolar operator (\ref{5.27}) on $\Phi({\bf R}_1-{\bf R}_2)$ (Figure~\ref{fig4}) \begin{eqnarray} F_{cm}(1,2)=-\beta_{1,2}q_1\int_0^{q_{2}}ds_2M_2(s_2)\Phi ({\bf R}_1-{\bf R}_2)\label{5.48}\\ F_{mc}(1,2)=-\beta_{1,2}q_2\int_0^{q_{1}}ds_1M_1(s_1)\Phi ({\bf R}_1-{\bf R}_2) \label{5.48a} \end{eqnarray} Obviously the resummed bonds (\ref{5.48}) and (\ref{5.48a}) have a short range. Let us have a closer look at the multipole-multipole resummed bond (Figure~\ref{fig5}) \begin{eqnarray} F_{mm} (1,2)&=&f_{mm}(1,2)+f_{mc}\star f_{cm}(1,2)+ f_{mc}\star f_{cc}\star f_{cm}(1,2)+\cdots\nonumber\\ &&\hspace{-.5in}=-\beta_{1,2}V_{mm}(1,2)\label{5.49}\\ &&\hspace{-.5in}-\beta_{1,2}\int_0^{q_{1}}ds_1M_1(s_1) \int_0^{q_{2}}ds_2M_2(s_2)(\Phi({\bf R}_1-{\bf R}_2)-V({\bf R}_1-{\bf R}_2)) \nonumber\\ &&\hspace{-.5in}=-\beta W(1,2)-\beta_{1,2}\int_0^{q_{1}}ds_1M_1(s_1) \int_0^{q_{2}}ds_2M_2(s_2)\Phi({\bf R}_1-{\bf R}_2)\nonumber\\ \label{5.50} \end{eqnarray} The first term in (\ref{5.49}) is the multipole-multipole quantum potential that has only "equal time" contributions, whereas the rest of the chain can be obtained by an application of multipolar operators at both ends (omitting here the term with no intermediate Coulomb points). The expression (\ref{5.50}) results from (\ref{5.34}), showing the asymptotic dipolar character of $F_{mm}$ (see (\ref{5.35})). Notice that the sum of the resummed bonds equals \begin{eqnarray} F(1,2)&\equiv& F_{cc}(1,2)+ F_{cm}(1,2)+ F_{mc}(1,2)+ F_{mm}(1,2)\nonumber\\ &=&-\beta W(1,2)-\beta_{1,2}\Phi_{{\rm cl}}(1,2) \label{5.51} \end{eqnarray} where $\Phi_{{\rm cl}}(1,2)$ is a short-range interaction defined as in (\ref{5.25}) \begin{equation} \Phi_{{\rm cl}}({\cal L}_1,{\cal L}_2)=\int_0^{q_1}ds_1\int_0^{q_2}ds_2 \Phi({\bf R}_1+\lambda_{\alpha_{1}}{\bf X}_1(s_1)-{\bf R}_2-\lambda_{\alpha_{2}}{\bf X}_2(s_2)) \label{5.52} \end{equation} representing the screened classical interaction between loops. The screening length $\kappa^{-1}$ in (\ref{5.47}) can be written in a more familiar form by noting the identity coming from (\ref{5.23b}) \begin{equation} \int d{\bf r} \rho^{(ex)}_{T}(\alpha,{\bf r},{\bf 0})=\sum_{q=2}^{\infty}\int D({\bf X})q(q-1) \rho(\alpha,q,{\bf X}) \label{5.52a} \end{equation} hence, using also (\ref{5.21}) \begin{equation} \kappa^{2}=4\pi\beta\sum_\alpha e_\alpha^2\(\rho_\alpha+\int d{\bf r} \rho^{(ex)}_{T}(\alpha,{\bf r},{\bf 0})\) \label{5.52b} \end{equation} An alternative expression for $\kappa^{2}$ follows from the quantum charge sum rule (\ref{2.2.5a}) \begin{equation} \int d{\bf r} c_{T}({\bf r}, {\bf 0})=\sum_{\alpha_{1}\alpha_{2}} e_{\alpha_{1}}e_{\alpha_{2}}\int d{\bf r} \left(\rho_{T}(\alpha_{1},{\bf r},\alpha_{2},{\bf 0})+\rho_{\alpha_{1}}\delta_{\alpha_{1}\alpha_{2}}\delta ({\bf r})\right)=0 \label{5.52bb} \end{equation} which leads to the following relation, once expressed in terms of the loop distributions by (\ref{5.22}) \begin{equation} \sum_\alpha e_\alpha^2\(\rho_\alpha+\int d{\bf r} \rho^{(ex)}_{T}(\alpha,{\bf r},{\bf 0})\)=-\sum_{\alpha_{1},\alpha_{2}}e_{\alpha_{1}} e_{\alpha_{2}}\int d{\bf r}\rho_{T}^{(nex)}(\alpha_1,{\bf r},\alpha_2,{\bf 0}) \label{5.52c} \end{equation} and thus \begin{equation} \kappa^{2}=-4\pi\beta \sum_{\alpha_{1},\alpha_{2}}e_{\alpha_{1}} e_{\alpha_{2}}\int d{\bf r}\rho_{T}^{(nex)}(\alpha_1,{\bf r},\alpha_2,{\bf 0}) \label{5.52d} \end{equation} In the integrals in (\ref{5.52c}) and (\ref{5.52d}), the contribution of coincident points is not included. On physically reasonable grounds, $\kappa ^{2}$ is indeed a positive quantity. One sees from (\ref{5.52b}) that $\kappa ^{2}$ reduces to the classical Debye value when exchange effects can be neglected, and it will coincide with the RPA value at high density (section~\ref{subsec-6.1.4}), both quantities being positive. Moreover, the screening cloud $\sum_{\alpha_{2}} e_{\alpha_{2}}\int d{\bf r}\rho_{T}^{(nex)}(\alpha_1,{\bf r},\alpha_2,{\bf 0})$ of charges of different species around the charge $e_{\alpha_{1}}$ should have the sign opposite to $e_{\alpha_{1}}$, implying that the right hand side of (\ref{5.52d}) is positive. \subsubsection{The prototype diagrams} \label{subsec-5.2.3} The $\tilde{\Gamma}$-diagrams can be divided into classes such that all diagrams in a class lead to the same so-called prototype $\Pi$-diagram when all the Coulomb points are integrated out. Points in a $\tilde{\Gamma}$-diagram that survive as points in a $\Pi$-diagram are either convolutions which do not involve Coulomb points, or points linked to three or more other points of $\tilde{\Gamma}$. There will be four different bonds (called F-bonds) in prototype diagrams: \begin{description} \item[(i)] the screened charge-charge bond (\ref{5.45}) and the two charge-multipole bonds (\ref{5.48})-(\ref{5.48a}). Since Coulomb points cannot occur in prototype graphs, the use of these bonds is subjected to the excluded convolution rule: in a $\Pi$-diagram, convolutions $F_{cc}\star F_{cc},F_{mc}\star F_{cc}, F_{cc}\star F_{cm},F_{mc}\star F_{cm}$ are forbidden. \item[(ii)] a bound $F_l$ that incorporates all the bonds that are not taken into account in ({\bf i}). \end{description} Let us now construct the bond $F_l$. We distinguish two cases: \begin{description} \item[(a)] $f_T$ and $F_{mm}$ occur as individual bonds in $\Pi$-diagrams since they are never attached to Coulomb points in $\tilde{\Gamma}$-diagrams, \item[(b)] single bonds in a $\Pi$-diagram attached to points that had more than two links in a $\tilde{\Gamma}$-diagram, as the result of the suppression of Coulomb points. \end{description} The latter situation is obtained when two points in a $\tilde{\Gamma}$-diagram were linked \begin{description} \item[(b.1)] by any one of the f-bonds together with Coulomb chains in parallel (Figure~\ref{fig6}) \item[(b.2)] by several Coulomb chains (Figure~\ref{fig7}). \end{description} The Coulomb chains occurring in ({\bf b.1}) and ({\bf b.2}) must have intermediate points since there is at most one f-bond between two points in a $\tilde{\Gamma}$-diagram. The sum of such Coulomb chains,$F_{ch}$, is obtained by removing from $F$ (\ref{5.51}) the single bonds (\ref{5.41})-(\ref{5.44}). Thus, with (\ref{5.28}) \begin{equation} F_{ch}(i,j)=F(i,j)+\beta_{i,j} V(i,j)=-\beta W(i,j)+\beta_{i,j}(V(i,j)-\Phi_{{\rm cl}}(i,j)) \label{5.53} \end{equation} The contributions described in ({\bf a}) and ({\bf b}) sum up to \begin{equation} F_l=f_T +F_{mm}+f(\exp (F_{ch})-1)+\exp (F_{ch})-1-F_{ch} \label{5.54} \end{equation} The last two parts of (\ref{5.54}) correspond to the contributions ({\bf b.1}) and ({\bf b.2}) (for the combinatorial aspects, see \cite{Cor96a}. Combining the definitions (\ref{5.38}), (\ref{5.40}), (\ref{5.51}) and (\ref{5.53}) in (\ref{5.54}) leads to the final result (Figure~\ref{fig8}) \begin{equation} F_l(i,j)=\exp(-\beta W(i,j)-\beta_{i,j}\Phi_{{\rm cl}}(i,j) )-1 -F_{cc}(i,j)-F_{cm}(i,j)-F_{mc}(i,j) \label{5.55} \end{equation} Eventually the loop-Ursell function is given by a sum of prototype graphs $\Pi$ \begin{equation} h({\cal L}_1,{\cal L}_2)=\sum_\Pi\frac{1}{S_{\Pi}}\int\prod_{n=1}^Nd{\cal L}_n\rho({\cal L}_n)[\prod F]_{\Pi} \label{5.56} \end{equation} with F-bonds $F_{cc}, F_{mc},F_{cm}$ and $F_{l}$. The $\Pi$-diagrams have the same structure as the $\Gamma$-diagrams, with the additional excluded convolution rule formulated in ({\bf i}). All F-bonds are rapidly decreasing except $F_l$. Obviously the asymptotic behavior of $F_l$ is dominated by that of $W({\cal L}_1,{\cal L}_2)$ \begin{equation}F_l({\cal L}_1,{\cal L}_2) \sim\exp(-\beta W({\cal L}_1,{\cal L}_2))-1\sim -\beta W({\cal L}_1,{\cal L}_2),\;\;\;|{\bf R}_1-{\bf R}_2| \rightarrow\infty \label{5.57} \end{equation}which is of dipole type as shown by (\ref{5.35}). It is interesting to note that each $F_l$ bond has the same form (up to short range contributions) as that found in the model of section ~\ref{sec-4.3} when only two quantum mechanical charges are present. The long range part $W$ depends solely on the intrinsic quantum fluctuations of the two particles (or the two loops), and not on the temperature and density of the state. There are of course several ways of reorganizing the diagrams with bare Coulomb bonds (\ref{5.38}) into prototype graphs by partial resummations. Other resummations may be more adequate depending on the density and temperature regime and the physical quantity to be computed. The present reorganization appears to be optimal in the following sense. The remaining long range part has been isolated in the single bond $F_{l}$. The decay cannot be improved any more by chain resummations. By (\ref{5.57}) and (\ref{5.35}), the longest range part of such chains would arise from convolutions of $W$-bonds, but such convolutions keep their dipolar character \cite{Cor96b}; the same is true for classical dipoles, see for instance \cite{HoSt74}. Moreover, our prototype diagrams are the natural generalization of those used in the classical case. Indeed, if we set the de Broglie lengths equal to zero, the multipolar operator (\ref{5.27}) vanishes as well as $W$, $F_{cm}$ and $F_{mc}$. The effects of the quantum statistics disappear, and both $\Phi$ (\ref{5.46}) and $\Phi_{{\rm cl}}$ (\ref{5.52}) reduce to the usual Debye potential $\Phi_{D}$. Hence one finds that the only surviving bonds are $$ \lim_{\hbar\rightarrow 0}F_{cc}(1,2)=-\beta_{1,2}\Phi_{D}(1,2) $$ and from (\ref{5.55}) $$ \lim_{\hbar\rightarrow 0}F_{l}=e^{-\beta_{1,2}\Phi_{D}(1,2) }-1+\beta_{1,2}\Phi_{D}(1,2) $$ They are precisely the two bonds occurring in the Meeron theory of the classical plasma \cite{Mee58,Mee61}. \subsubsection{Integrability of prototype diagrams} \label{subsec-5.2.4} Let us present some crude arguments for the integrability of $\Pi$-diagrams; they can be substantiated by the more thorough analysis of \cite{Cor96a}, but a proof is still missing. Each internal point involves a density factor $\rho({\cal L})=\rho(\alpha,q,{\bf X})$ which depends only on the internal variables of the loop and not on its location in space because of translation invariance. Integration on an internal point ${\cal L}$ will be of the form \begin{equation}\int d{\bf R}\sum_\alpha\sum_{q=1}^\infty\int D({\bf X}) \rho(\alpha,q,{\bf X}) G(\alpha,q,{\bf R},{\bf X}\;;\ldots) \label{5.58} \end{equation}where $G({\cal L}\;;\ldots)$ denotes the ${\cal L}$-dependence of the bonds attached to this point. It can be argued that the loop-density $\rho({\cal L})$ is bounded in the shape variables ${\bf X}$ and inherits a prefactor $\exp(\beta\mu_\alpha q)$ from the activity of the loop (see (\ref{5.13}), (\ref{5.15}) and (\ref{5.60}) below). Thus for a sufficiently negative chemical potential this factor should insure convergence for the $q$ summation on the size of the loops. If these conditions are met, the integrability of $\Pi$-diagrams, when the distance between the loops becomes large, is determined by the the decay of the F-bonds. Because of the absence of articulation points, any cluster of points is linked by at least two bonds to the remaining points of the diagram. Hence the product $[\prod F]_{\Pi}$, once integrated on the internal variables, decays at least as $|{\bf R}|^{-6}$ in the loop distances, showing the integrability for the whole diagram. One should add that the splitting (\ref{5.39}) of the original f-bond may introduce spurious non-integrable singularities at the origin in some diagrams. These singularities will cancel out when suitably collecting together such dangerous diagrams (the f-bond (\ref{5.38}) gives a finite contribution at the origin because of the smoothening provided by the functional integration on the filaments). To summarize the situation, once the topological reduction to prototype diagrams has been performed, the two-particle correlation is reduced, in the sense of the formal diagrammatic cluster expansion (\ref{5.56}), to a system that resembles a classical gas of multipoles described in the loop formalism by the four types of bonds $F_{cc},F_{mc},F_{cm}$ and $F_{l}$. \subsubsection{The density expansion} \label{subsec-5.2.5} An important remark has to be made when one comes down to computations, in particular when one seeks to obtain explicit low density expansions of physical quantities. The diagrammatic representation (\ref{5.56}) of the loop Ursell function is a straightforward expansion in the density of loops, but not in the particle densities $\rho_\alpha$ themselves, for two reasons: \begin{itemize} \item the resummed F-bonds depend on the particle densities through the density dependent screened potential $\Phi$ (\ref{5.46}), \item the loop densities entering as weights of the internal points in $\Pi$-diagrams are not simply proportional to the particle densities. \end{itemize} The first point is not new and occurs of course already in the diagrammatic treatment of the classical gas. The second point is more delicate: the loop and particle densities are related by the formula (\ref{5.21}); hence the particle-particle distributions (\ref{5.22}) calculated from (\ref{5.56}) have a very implicit dependence on the $\rho_\alpha$. In principle this dependence can be extracted as follows. According to the standard rules of the Mayer diagrammatics, the loop density is represented by the activity expansion \begin{equation} \rho({\cal L})=z({\cal L})\sum_G\frac{1}{S_G}\int\prod_{n=1}^{N}d{\cal L}_nz({\cal L}_n)[\prod f]_G \label{5.60} \end{equation} The $G$-diagrams are defined as the $\Gamma$-diagrams (with one root point) except that articulation points are allowed and weights are given by activities. The integrals diverge in (\ref{5.60}) because of the long range of the $f$-bonds, but after a suitable topological reduction (resumming Coulomb chains), the representation of $\rho({\cal L})$ becomes a series of prototype graphs, each of them being finite. The procedure is similar to that given above for the Ursell function and will not be described here in detail, see \cite{Cor96a}. Once this is done, formula (\ref{5.21}) yields in principle the particle densities as functions of the activities \begin{equation} \rho_\alpha=\rho_\alpha(z_1,\ldots,z_{\cal S}),\;\;\;\,\alpha=1,\ldots,{\cal S} \label{5.61} \end{equation} Because of the overall neutrality $\sum_{\alpha=1}^{{\cal S}}e_\alpha\rho_\alpha=0$ the system (\ref{5.61}) consists only of ${\cal S}-1$ independent relations, and so does not determine the $z_\alpha$ uniquely. A unique determination of the $z_a$ is usually made by imposing the convenient additional relation (already discussed in section~\ref{sec- neutrality}) \begin{equation} \sum_{\alpha=1}^{{\cal S}}e_\alpha z_\alpha=0 \label{5.62} \end{equation} Thus by inverting the system (\ref{5.61})-(\ref{5.62}) one can express first $\rho({\cal L})$ and then the particle-particle correlations in terms of the $\rho_\alpha$. Explicit calculations in this formalism can of course only be performed in the low density regime. Applications to the determination of the asymptotic behavior of the correlations and to the virial equation of state will be presented in the next chapter. \def\({\left(} \def\){\right)} \def{\bf r}{{\bf r}} \def{\bf \xi}{\mbox{\boldmath $\xi$\unboldmath}} \def{\cal F}{{\cal F}} \def{\cal L}{{\cal L}} \def{\bf k}{{\bf k}} \def{\bf X}{{\bf X}} \newpage \section{Correlations and the equation of state} \setcounter{page}{1} \label{chapter-6} \subsection{Asymptotic behavior of particle and charge correlations} \label{sec-6.1} The first application of the general formalism concerns the determination of the large distance behavior of the correlations without recourse to $\hbar$- expansions, including the quantum statistics and using the exact Coulomb potential on all scales. The calculation remains perturbative in the sense that we examine the large distance behavior of each $\Pi$-diagram in the series (\ref{5.56}) representing the Ursell function, but, as before, without control on the sum of these series. The semi-classical analysis of section~\ref{sec-4.3} has revealed that the two particle correlation, at order $\hbar^{4}$, decay as $r^{-6}$ at large distance. We show that the same result holds for each of the $\Pi$-diagrams contributing to the particle- particle correlation $\rho_{T}(\alpha_{a},{\bf r}_{a},\alpha_{b},{\bf r}_{b})$, i.e. $$ \rho_{T}(\alpha_{a},{\bf r}_{a},\alpha_{b},{\bf r}_{b})\sim \frac{A_{ab}(\beta,\{\rho_{\alpha}\})}{|{\bf r}_{a}-{\bf r}_{b}|^{6}} $$ where the coefficient $A_{ab}(\beta,\{\rho_{\alpha}\})$ is the formal sum of the asymptotic contributions of the $\Pi$-diagrams as $|{\bf r}_{a}-{\bf r}_{b}|\to\infty$ at densities $\rho_{\alpha}$ and inverse temperature $\beta$. It is then possible to calculate the lowest order contributions in the densities $\rho_{\alpha}$ to $A_{ab}(\beta,\{\rho_{\alpha}\})$ leading to the explicit formula (\ref{6.23}) below. We will describe the general strategy and illustrate some key points by explicit calculations, details can be found in \cite{Cor96b}. A short review can be found in \cite{AlCo97}. \subsubsection{Large distance behavior of the loop Ursell function} \label{sec-6.1.1} The first task is to study the behavior of the Ursell function (\ref{5.56}) in the space of loops as the loops are separated. This large distance behavior will be determined by that of the four bonds $F_{cc},F_{cm},F_{mc}$ and $F_{l}$, once integration on the internal loop variables has been performed in prototype graphs. Among these bonds, only $F_{l}$ has the long range multipolar character (\ref{5.57}) that may induce a slow decay of the graph as the two loops are separated. Therefore one has to keep track of the effects of $F_{l}$-bonds (and more specifically of the basic quantum interaction $W$ (\ref{5.33})) in the convolutions occurring in prototype diagrams. For this it is appropriate to recall some facts on the asymptotic form of convolutions in three dimensional space. The forthcoming analysis depends on the assumption (made throughout the whole section \ref{sec-6.1}) that the leading asymptotic term has a monotone decay, whose behavior is governed by that of the small wave numbers ${\bf k}\sim 0$ of the Fourier transform. Possible decay with oscillation, due to singularities at non-zero values of ${\bf k}$, is not considered. Such oscillating decays are not expected to occur at positive temperature and low density. Consider a (non-oscillating) function $g({\bf r})$ which decays as $r=|{\bf r}|\rightarrow\infty$ as (up to a multiplicative constant) \begin{equation} g({\bf r})\sim \frac{(\ln r)^{m}}{r^{n}},\;\;\;n,m\;\mbox{integers},\;\,\,n\geq 3,\;\;\;m\geq 0 \label{6.1b} \end{equation} This type of decay is reflected by the existence of a non-analytic part $\tilde{g}_{sing}({\bf k})$ in its Fourier transform $\tilde{g}({\bf k})$ as ${\bf k}\rightarrow 0$. One can split the low ${\bf k}$-expansion of $\tilde{g}({\bf k})$ into $\tilde{g}({\bf k})=\tilde{g}_{reg}({\bf k})+\tilde{g}_{sing}({\bf k})$ where $\tilde{g}_{reg}({\bf k}) =a_{0}+\sum_{\mu=1}^{3}a_{\mu}k_{\mu} +\sum_{\mu,\nu=1}^{3}a_{\mu\nu}k_{\mu}k_{\nu} +\cdots$ has a Taylor expansion around ${\bf k}={\bf 0}$ and $\tilde{g}_{sing}({\bf k})$ behaves (up to multiplicative constants) as \cite{GeSh64} \begin{equation} \tilde{g}_{sing}({\bf k})\sim \left\{ \begin{array}{ll} |{\bf k}|^{n-3}(\ln|{\bf k}|)^{m+1}, & n\;\mbox{odd}\\ |{\bf k}|^{n-3}(\ln|{\bf k}|)^{m}, & n\;\mbox{even}\end{array}\right.,\;\;\,\,{\bf k}\rightarrow 0 \label{6.1a} \end{equation} The singular part of Fourier transforms of derivatives $\partial_{\mu_{1}}\cdots \partial_{\mu_{j}}g({\bf r}) $ is\\ ${ k}_{\mu_{1}}\cdots { k}_{\mu_{j}}\tilde{g}_{sing}({\bf k})$. Since the Fourier transform of the convolution $(g_{1}\star g_{2})({\bf r})$ is the product $\tilde{g}_{1}({\bf k})\tilde{g}_{2}({\bf k})$, one deduces from the above correspondence rules that for $n_{1},n_{2}>3$ \begin{equation} (g_{1}\star g_{2})({\bf r}) \sim g_{1}({\bf r})\int d{\bf r}^{\prime}g_{2}({\bf r}^{\prime})+ g_{2}({\bf r})\int d{\bf r}^{\prime} g_{1}({\bf r}^{\prime}),\,\;\,r\rightarrow\infty \label{6.3} \end{equation} In particular, if $n_{2}>n_{1}>3$, the convolution decays as the slowest of the two functions \begin{equation} (g_{1}\star g_{2})({\bf r}) \sim g_{1}({\bf r})\int d{\bf r}^{\prime} g_{2}({{\bf r}^{\prime}),\,\;\,r\rightarrow\infty } \label{6.2} \end{equation} At the border $n_{1}=n_{2}=3$ of the integrability domain one has \begin{equation} (g_{1}\star g_{2})({\bf r})\sim \frac{(\ln r)^{m_{1}+m_{2}+1}}{r^{3}} \label{6.4} \end{equation} One sees from (\ref{6.4}) that convolutions of $r^{-3}$ functions can generate in principle logarithmic factors in the power law decays. To simplify the discussion we abbreviate (\ref{6.1b}) by $g({\bf r})\approx r^{-n}$, omitting from the notation the possible logarithmic dependence. Since $F_{l}$ decays as $r^{-3}$, it is clear from the above convolution rules that no $\Pi $-diagram (or subdiagram) decays slower than $\approx r^{-3}$. The slowest possible decay comes then from diagrams where the two root points can be disconnected by cutting a single $F_{l}$-bond: otherwise chains containing a $F_{l}$-bond will be multiplied by another chain (or subdiagram) decaying at least as $\approx r^{-3}$ so the decay of the whole diagram will be at least as $\approx r^{-6}$. To put this idea at work, one splits the $F_{l}$-bond into two pieces \begin{equation} F_{l}=-\beta W+\tilde{F}_{l},\,\;\;\,\,\,\tilde{F}_{l}\equiv F_{l}+\beta W \label{6.6} \end{equation} and enlarge accordingly the set of $\Pi$-diagrams to $\tilde{\Pi}$-diagrams equipped with the new bonds $-\beta W$ and $\tilde{F}_{l}$. One has from (\ref{5.57}) \begin{eqnarray} \tilde{F}_{l}(1,2)&\sim& \exp(-\beta W(1,2))-1+\beta W(1,2)\sim\frac{1}{2}(\beta W(1,2))^{2}\nonumber\\ \mbox{and}\;\;\;\;\, (W(1,2))^{2}&\sim&\frac{C}{|{\bf R}_{1}-{\bf R}_{2}|^{6}},\;\;\;\;|{\bf R}_{1}-{\bf R}_{2}|\rightarrow\infty \label{6.7} \end{eqnarray} thus $\tilde{F}_{l}$ is integrable and $W$ is the slowest bond in the new set of $\tilde{\Pi}$-diagrams. Introduce now the subclass of $\tilde{\Pi}_{C}$-diagrams that remain connected when one removes a $W$-bond, and denote $H({\cal L}_{a},{\cal L}_{b})$ the value of the sum of these diagrams. With this definition, the loop truncated distribution function (\ref{5.36}) can be expressed as an exact Dyson equation in terms of convolution chains involving $H$ and $W$ \begin{eqnarray} &&\rho_{T}({\cal L}_{a},{\cal L}_{b}) = \rho({\cal L}_{a})\rho({\cal L}_{b})H({\cal L}_{a},{\cal L}_{b}) -\beta \int d{\cal L}_{1}d{\cal L}_{1}^{\prime}G({\cal L}_{a},{\cal L}_{1}) W({\cal L}_{1},{\cal L}_{1}^{\prime})G({\cal L}_{1}^{\prime},{\cal L}_{b})\label{6.9}\\ &&\hspace*{.25in}+\beta^{2}\int d{\cal L}_{1}d{\cal L}_{1}^{\prime}\int d{\cal L}_{2}d{\cal L}_{2}^{\prime}G({\cal L}_{a},{\cal L}_{1}) W({\cal L}_{1},{\cal L}_{1}^{\prime})G({\cal L}_{1}^{\prime},{\cal L}_{2}) W({\cal L}_{2},{\cal L}_{2}^{\prime})G({\cal L}_{2}^{\prime},{\cal L}_{b}) + \cdots \label{6.10} \end{eqnarray} In (\ref{6.9})-(\ref{6.10}) we have set \begin{equation} G({\cal L}_{1},{\cal L}_{2})=\rho({\cal L}_{1})\rho({\cal L}_{2})H({\cal L}_{1},{\cal L}_{2})+\delta({\cal L}_{1},{\cal L}_{2}) \rho({\cal L}_{1}) \label{6.11} \end{equation} where the term $\delta({\cal L}_{1},{\cal L}_{2})\rho({\cal L}_{1})$ takes into account the case when $W$-bonds are in direct convolution with themselves. By construction, no diagram in $H({\cal L}_{a},{\cal L}_{b})$ decays slower than $\approx |{\bf R}_{a}-{\bf R}_{b}|^{-6}$. Indeed if a $\tilde{\Pi}_{C}$-diagram contains no $W$-bonds, the slowest possible decay is determined by that of $\tilde{F}_{l}$ according to (\ref{6.7}). If it contains a $W$-bond, this $W$ bond (or the chain where it belongs to) is necessarily multiplied by another chain that decays not slower than $\approx r^{-3}$, hence the overall decay of the diagram is at least $\approx r^{-6}$. In this way we have isolated the origin of decays that are slower than $\approx r^{-6}$ in the chains of $W$-bonds (\ref{6.9})-(\ref{6.10}). Since (\ref{6.9}) contains in particular the term $\rho({\cal L}_{a})\rho({\cal L}_{b})W({\cal L}_{a},{\cal L}_{b})$ the loop correlation decays indeed as $\approx r^{-3}$. \subsubsection{Large distance behavior of the two-particle correlation} \label{sec-6.1.2} In the loop formalism, the two-particle distribution consists of the two terms displayed in (\ref{5.22})-(\ref{5.23a}). The dominant part at large distances will come from $\rho^{(nex)}$ involving particles belonging to different loops. The contribution $\rho^{(ex)}$ of particles exchanged in the same loop will be discussed at the end of this subsection. To obtain $\rho^{(nex)}$ we have to integrate $\rho_{T}({\cal L}_{a},{\cal L}_{b})=\rho({\cal L}_{a})\rho({\cal L}_{b})h({\cal L}_{a},{\cal L}_{b})$ over the internal variables of the root loops ${\cal L}_{a},{\cal L}_{b}$ according to (\ref{5.23}): \begin{equation} \rho^{(nex)}(\alpha_{a},{\bf r}_{a},\alpha_{b},{\bf r}_{b})=\sum_{q_{a}q_{b}}q_{a}q_{b}\int D({\bf X_{a}}) D({\bf X_{b}}) \rho({\cal L}_{a})\rho({\cal L}_{b})h({\cal L}_{a},{\cal L}_{b}) \label{6.11a} \end{equation} The point to be made below is that, {\em once these integrations have been performed}, the asymptotic contribution of the chains of $W$-bonds (\ref{6.9})-(\ref{6.10}) is strictly faster than $\approx |{\bf R}_{a}-{\bf R}_{b}|^{-6}$. Hence the decay of $\rho^{(nex)}$ is entirely determined by that of the $\tilde{\Pi}_{C}$-diagrams constituting the function $H$, i.e. $\approx {\bf r}^{-6}$ according to the analysis of the preceding subsection. Consider for simplicity the diagram consisting of a single $W$-bond in (\ref{6.9}): its contribution to $\rho^{(nex)}({\bf r}_{a},{\bf r}_{b})$ is \begin{equation} -\beta\sum_{q_{a}=1}^{\infty}\sum_{q_{b}=1}^{\infty}q_{a}q_{b}\int D({\bf X}_{a})\int D({\bf X}_{b}) \rho(\alpha_{a},q_{a},{\bf X}_{a})\rho(\alpha_{b},q_{b},{\bf X}_{b})W({\cal L}_{a},{\cal L}_{b}) \label{6.12} \end{equation} The same mechanism that was put forward in the model of section~\ref{sec-4.3} to show that the term linear in $W$ was in fact short-ranged works also in the present case. Since in a homogeneous phase of the Coulomb gas the loop-density $\rho(\alpha,q,{\bf X})$ as well as the measure $D({\bf X})$ are invariant under global rotations of the shape ${\bf X}$ of the loop, the expressions \begin{equation} \int D({\bf X})\rho(\alpha,q,{\bf X})\frac{({\bf X}(s)\cdot{\bf \nabla_{\bf R}})^{k}}{k!},\;\,\,k=1,2,\ldots \label{6.13} \end{equation} occurring in (\ref{6.12}) according to the multipole expansion (\ref{5.34}) of $W$ have to vanish for $k$ odd and are necessarily proportional to $(\nabla^{2}_{\bf R})^{k/2}$ for $k$ even. Then the contribution (\ref{6.13}) to (\ref{6.12}) is strictly local since $\nabla^{2}_{\bf R}\(\frac{1}{|{\bf R}|}\)=0,{\bf R} \neq{\bf 0}$: thus (\ref{6.12}) decays faster than any inverse power. More generally, it can be shown that as a consequence of the rotational symmetry and of the harmonicity of the Coulomb potential, all chains (\ref{6.10}) involving more than one $W$-bond decay also strictly faster than $\approx |{\bf r}_{a}-{\bf r}_{b}|^{-6}$. As an example, we give the argument for the term (\ref{6.10}) involving a two $W$-bond convolution and keeping only the slowest part of these $W$-bonds, i.e. the dipole-dipole interaction (\ref{5.35}). Once (\ref{5.35}) has been inserted in (\ref{6.10}) one obtains convolutions relative to the loop positional variables ${\bf R}$ of the form \begin{equation} (G_{\nu_{1}}\star \partial_{\nu_{1}}\partial_{\mu_{1}}V\star G_{\mu_{1}\nu_{2}}\star \partial_{\nu_{2}}\partial_{\mu_{2}}V\star G_{\mu_{2}})({\bf R}_{a}-{\bf R}_{b}) \label{6.14} \end{equation} with \begin{eqnarray} G_{\nu}({\bf R})&=&\int D({\bf X}_{a})\int D({\bf X}){ X}_{\nu}(s) G({\bf R},{\bf X}_{a},{\bf X}) \label{6.15}\\ G_{\mu\nu}({\bf R})&=&\int D({\bf X}^{\prime})\int D({\bf X}){ X}^{\prime}_{\mu}(s^{\prime}) { X}_{\nu}(s)G({\bf R},{\bf X}^{\prime},{\bf X}) \label{6.16} \end{eqnarray} where we write \begin{equation} G({\cal L}_{1},{\cal L}_{2})=G({\bf R}_{1}-{\bf R}_{2},{\bf X}_{1},{\bf X}_{2}) \label{6.17} \end{equation} because of translation invariance. In (\ref{6.14})-(\ref{6.17}) we have omitted the loop indices $\alpha,q$ and the times $s$ as well as the corresponding summations. The loop internal variables have been integrated and sums on repeated Cartesian indices are understood. To analyze the long distance behavior of (\ref{6.14}) it is convenient to first consider its Fourier transform \begin{equation} \tilde{G}_{\nu_{1}}({\bf k})\frac{{ k}_{\nu_{1}}{ k}_{\mu_{1}}}{|{\bf k}|^{2}}\tilde{G}_{\mu_{1}\nu_{2}}({\bf k}) \frac{{ k}_{\nu_{2}}{ k}_{\mu_{2}}}{|{\bf k}|^{2}}\tilde{G}_{\mu_{2}}({\bf k}) \label{6.18} \end{equation} Since the measure $D({\bf X})$ as well as $G({\bf R},{\bf X}_{1},{\bf X}_{2})$ are invariant under spatial rotations , $\tilde{G}_{\nu}({\bf k})$ and $\tilde{G}_{\mu\nu}({\bf k})$ transform as tensors: hence they are necessarily of the form \begin{eqnarray} \tilde{G}_{\nu}({\bf k})&=&{ k}_{\nu}a(|{\bf k}|)\nonumber\\ \tilde{G}_{\mu\nu}({\bf k})&=&\delta_{\mu\nu}b(|{\bf k}|)+{ k}_{\mu}{ k}_{\nu}c(|{\bf k}|) \label{6.19} \end{eqnarray} for some functions $a, b$ and $c$ of $|{\bf k}|$. We know that $G({\bf R},{\bf X}_{1},{\bf X}_{2})$ and hence $G_{\nu}({\bf R}),\;\; G_{\mu\nu}({\bf R})$ do not decay slower than $\approx |{\bf R}|^{-6}$. Taking the correspondence (\ref{6.1b})-(\ref{6.1a}) into account, this implies that the small $|{\bf k}|$ expansion of $\tilde{G}_{\nu}({\bf k})$ and $\tilde{G}_{\mu\nu}({\bf k})$ must be of the form \begin{eqnarray} \tilde{G}_{\nu}({\bf k})&=&{ k}_{\nu}(a+O\(|{\bf k}|^{2}(\ln|{\bf k}|)^{m_{1}+1}\)\nonumber\\ \tilde{G}_{\mu\nu}({\bf k})&=&\delta_{\mu\nu}(b_{1}+b_{2}|{\bf k}|^{2}+O(|{\bf k}|^{3}(\ln |{\bf k}|)^{m_{2}})\nonumber\\ &+&{ k}_{\mu}{ k}_{\nu} \(c+O(|{\bf k}|(\ln|{\bf k}|)^{m_{3}}\) \label{6.20} \end{eqnarray} for some $m_{1},m_{2},m_{3}\geq 0$. Upon inserting (\ref{6.20}) into (\ref{6.18}) one sees that the Coulomb singularities $|{\bf k}|^{-2}$ are compensated and (\ref{6.18}) behaves as $C|{\bf k}|^{2}+O(|{\bf k}|^{4}(\ln|{\bf k}|)^{m+1})$ as $|{\bf k}|\rightarrow 0$, hence (\ref{6.14}) decays strictly faster than $\approx |{\bf R}|^{-6}$ (possibly as $\approx |{\bf R}|^{-7}$). The same argument extends to all higher order multipoles and chains in (\ref{6.10}) with arbitrary numbers of $W$-bonds, showing that $\rho^{(nex)}(\alpha_{a},{\bf r}_{a},\alpha_{b},{\bf r}_{b})$ decays at least as $\approx |{\bf r}_{a}-{\bf r}_{b}|^{-6}$. We briefly comment about the exchange part $\rho^{(ex)}(\alpha_{a},{\bf r}_{a},{\bf r}_{b})$ of the particle-particle distribution. One sees from (\ref{5.23a}) that this exchange part involves the loop density $\rho({\cal L})$ with the position of two particles fixed at ${\bf r}_{a}$ and ${\bf r}_{b}$. The Mayer expansion of the loop density (\ref{5.60}) has an activity factor $z({\cal L})$ that has the bound (\ref{5.15}). One can show that after the needed resummation of Coulomb chains, all prototype graphs representing $\rho({\cal L})$ inherit the same activity factor $z({\cal L})$. Furthermore prototype graphs are bounded functions of the loop variables ${\bf r}_{1},\ldots,{\bf r}_{q}$. Taking these facts into account in (\ref{5.23b}), the contribution of a prototype graph to $\rho^{(ex)}(\alpha_{a},{\bf r}_{a},{\bf r}_{b})$ will be majorized by $$ \sum_{q=2}^{\infty}q\int D({\bf X})|z({\cal L})|\sum_{k=2}^{q}\delta(\lambda_{\alpha_{a}}{\bf X}_{k}+{\bf r}_{a}-{\bf r}_{b}) \leq \frac{2s_{\alpha_{a}}+1}{\lambda_{\alpha_{a}}^{3}} $$ $$ \sum_{q=2}^{\infty}e^{\beta\mu_{\alpha_{a}} q} \int d{\bf X}_{2}\ldots d{\bf X}_{q}\prod_{k=1}^{q} \frac{\exp(-\frac{1}{2}|{\bf X}_{k+1}-{\bf X}_{k}|^{2})}{(2\pi)^{3/2}} \sum_{k=2}^{q}\delta(\lambda_{\alpha_{a}}{\bf X}_{k}+{\bf r}_{a}-{\bf r}_{b}) $$ \begin{equation} =(2s_{\alpha_{a}}+1)\left[\frac{1}{(2\pi)^{3}}\int d{\bf k}\exp(-i{\bf k}\cdot({\bf r}_{a}-{\bf r}_{b})) \frac{1}{\exp\(\beta\(\frac{\hbar^{2}}{2m}|{\bf k}|^{2}-\mu_{\alpha_{a}}\)\)-1}\right]^{2} \label{6.22} \end{equation} The integral in (\ref{6.22}) is nothing else than the off-diagonal part of the one-body density matrix of a free Bose gas (the Bose distribution occurs because we have replaced the fermionic sign $\eta_{\alpha_{a}}$ by $1$ in the bound). At sufficiently low density ($\exp(\beta\mu_{\alpha_{a}}<1$), this function, the Fourier transform of a infinitely differentiable function, decays faster than any inverse power of $|{\bf r}_{a}-{\bf r}_{b}|$. Under the assumption that the same property remains true for the sum of all prototype graphs, we conclude that the asymptotic behavior of the particle-particle distribution is dominated by its non-exchange part previously discussed. The diagrams contributing to the coefficient of the $r^{-6}$ tail can also be analyzed in the low density regime \cite{Cor97a}. At the lowest order in the densities, one finds that the tail has the same form (\ref{4.54}) as in the semi-classical gas \begin{equation} \rho_{T}(\alpha_{a},{\bf r}_{a},\alpha_{b},{\bf r}_{b})\sim \frac{1}{|{\bf r}_{a}-{\bf r}_{b}|^{6}} \frac{\hbar^{4}\beta^{4}}{240}\int d{\bf r} G^{D.H.}({\bf r},\alpha_{a}) \int d{\bf r} G^{D.H.}({\bf r},\alpha_{b}) \label{6.23} \end{equation} where $G^{D.H.}({\bf r},\alpha)$ is defined as in (\ref{4.53}), but calculated in the Debye-H\"{u}ckel approximation \begin{equation} \int d{\bf r} G^{D.H.}({\bf r},\alpha)=\rho_{\alpha}\(\frac{e_{\alpha}^{2}}{m_{\alpha}} -\frac{4\pi\beta e_{\alpha}}{\kappa^{2}}\sum_{\gamma}\frac{e_{\gamma}^{3}\rho_{\gamma}}{m_{\gamma}}\) \label{6.23a} \end{equation} This is to be expected, since at lowest order in the densities, the particles surrounding the two fixed charges at ${\bf r} _{a}$ and ${\bf r} _{b}$ are fully ionized and will behave as a classical plasma. It is interesting to specialize (\ref{6.23}) to the electron-proton system. In this case one finds from (\ref{6.23}) and (\ref{6.23a}) with $\alpha$=electron or proton, $e_{{\rm el}}=-e_{{\rm pr}}=e,\;\rho_{{\rm el}} =\rho_{{\rm pr}}=\rho$ \begin{equation} \rho_{T}(\alpha_{a},{\bf r}_{a},\alpha_{b},{\bf r}_{b})\sim \frac{\rho^{2}}{|{\bf r}_{a}-{\bf r}_{b}|^{6}} \frac{\hbar^{4}\beta^{4}}{240}\frac{e^{4}}{4}\(\frac{1}{m_{{\rm el}}}+\frac{1}{m_{{\rm pr}}}\)^{2} \end{equation} showing that in this system all tails have exactly the same decay rate at low density. The tail of the particle correlation in presence of a uniform magnetic field ${\bf B}$ has also been calculated \cite{Cor97,Cor98a}. Because of breaking of the full rotation invariance, it has the slower decay $\sim\frac{P_{4}(\cos\theta)}{|{\bf r}_{a}-{\bf r}_{b}|^{5}}$ where $P_{4}(\cos\theta)$ is a Legendre polynomial and $\theta$ is the angle between ${\bf B}$ and ${\bf r}_{a}-{\bf r}_{b}$. \subsubsection{ Correlations of the total charge density} \label{subsec-6.1.3} As in the semi-classical case, the total charge density enjoys better screening properties than individual particles. This is reflected by the fact that correlations involving the total charge density $\hat{c}({\bf r})=\sum_{\alpha}e_{\alpha}\hat{\rho}(\alpha,{\bf r})$ such as the charge-particle correlation \begin{equation} <\hat{c}({\bf r}_{a})\hat{\rho}(\alpha_{b},{\bf r}_{b}))>_{T}=\sum_{\alpha_{a}}e_{\alpha_{a}} \rho_{T}(\alpha_{a},{\bf r}_{a},\alpha_{b},{\bf r}_{b}) \label{6.24} \end{equation} will have a faster decay than the particle-particle correlation. To illustrate the origin of this additional screening mechanism, it is convenient to construct a certain "dressing" of the root point ${\cal L}_{a}$ in the Ursell function. For this we call ${\cal L}_{a}$ a Coulomb root point in a $\Pi$-diagram if ${\cal L}_{a}$ is linked to the rest of the diagram by exactly one $F_{cc}$ bond or one $F_{cm}$ bond, and split the Ursell function accordingly \begin{equation} h({\cal L}_{a},{\cal L}_{b})=h^{c}({\cal L}_{a},{\cal L}_{b})+h^{nc}({\cal L}_{a},{\cal L}_{b}) \label{6.25} \end{equation} $h^{c}({\cal L}_{a},{\cal L}_{b})$ $(h^{nc}({\cal L}_{a},{\cal L}_{b}))$ is the sum of $\Pi$-diagrams where ${\cal L}_{a}$ is (is not) a Coulomb point. This definition implies immediately the following integral equation \begin{equation} h^{c}=F_{cc} + F_{cc}\rho\star h^{nc}+F_{cm}+F_{cm}\rho\star h \label{6.26} \end{equation} and thus \begin{equation} h=F_{cc}+F_{cm}+(F_{cc}\rho+1)\star h^{nc} +F_{cm}\rho\star h \label{6.27} \end{equation} where here the symbol $\star$ means both integration on the internal variables and convolution with respect to the positions of the loops; $\rho=\rho({\cal L})$ is the weight of intermediate points. Note that $h^{nc}$ occurs in the second term of (\ref{6.26}) because convolutions of $F_{cc}$ bonds are forbidden in $\Pi$-diagrams. As we know the two first terms of (\ref{6.27}) are rapidly decreasing; the third term contributes to $\rho_{T}({\cal L}_{a},{\cal L}_{b})$ as $$ \rho({\cal L}_{a})\int d{\cal L} \(F_{cc}({\cal L}_{a},{\cal L})\rho({\cal L}) +\delta({\cal L}_{a},{\cal L})\)h^{nc}({\cal L},{\cal L}_{b})\rho({\cal L}_{b}) $$ \begin{equation} =\int d{\cal L} S({\cal L}_{a},{\cal L})\rho({\cal L})h^{nc}({\cal L},{\cal L}_{b})\rho({\cal L}_{b}) \label{6.28} \end{equation} with "dressing" factor of the point ${\cal L}_{a}$ defined by \begin{equation} S({\cal L}_{a},{\cal L})=\rho({\cal L}_{a})F_{cc}({\cal L}_{a},{\cal L})+\delta({\cal L}_{a},{\cal L}) \label{6.29} \end{equation} When summed on the internal variables of the loop ${\cal L}_{a}$ and the charges, it is immediately seen, using (\ref{5.45}) and (\ref{5.47}) that (\ref{6.29}) gives \begin{equation} \sum_{\alpha_{a}}e_{\alpha_{a}}\sum_{q_{a}}q_{a}\int D({\bf X}_{a})S({\cal L}_{a},{\cal L})= e_{\alpha}q_{\alpha}\(\delta({\bf r})-\kappa^{2}\frac{\exp(-\kappa r)}{4\pi r}\),\;\;\;{\bf r}={\bf R}_{a}-{\bf R} \label{6.30} \end{equation} with Fourier transform \begin{equation} e_{\alpha}q_{\alpha}\(1-\frac{\kappa^{2}}{\kappa^{2}+|{\bf k}|^{2}}\)= e_{\alpha}q_{\alpha}\frac{|{\bf k}|^{2}}{\kappa^{2}+|{\bf k}|^{2}} \label{6.31} \end{equation} Hence the convolution (\ref{6.28}) will contribute to the Fourier transform of the charge-particle correlation as \begin{equation} \frac{|{\bf k}|^{2}}{\kappa^{2}+|{\bf k}|^{2}}\times\left\{\begin{array}{ll}\mbox{Fourier transform of $\Pi$-diagrams constituting $h^{nc}$}\\ \mbox{integrated on the internal variables of their root points}\end{array}\right\} \label{6.32} \end{equation} Since the singular term of these diagrams is $|{\bf k}|^{3}(\ln|{\bf k}|)^{m}$ (corresponding to a $\approx r^{-6}$ decay as seen in section~\ref{sec-6.1.2}), the singular term in (\ref{6.32}) is $|{\bf k}|^{5}\ln|{\bf k}|^{m}$ implying a $\approx r^{-8}$ decay. The same conclusion can be drawn for the contribution of the last term of (\ref{6.27}) on the grounds of rotational invariance and of the short range of $F_{cm}$. The final result is that the charge-particle has the following large distance behavior in the low density regime \cite{Cor97a} \begin{equation} \langle\hat{\rho}(\alpha_{a},{\bf r}_{a})\hat{c}({\bf r}_{b})\rangle_{T} \sim\frac{1}{|{\bf r}_{a}-{\bf r}_{b}|^{8}} \frac{\hbar^{4}\beta^{4}}{16}\int d{\bf r} G^{D.H.}({\bf r}, \alpha_{a}) \(\frac{1}{3}\int d{\bf r} |{\bf r}|^{2}K^{D.H.}({\bf r})\) \label{6.33} \end{equation} where the last factor is the expression (\ref{4.57}) calculated in the Debye-H\"{u}ckel approximation \begin{equation} \frac{1}{3}\int d{\bf r} |{\bf r}|^{2}K^{D.H.}({\bf r})=-\frac{2}{\kappa^{2}}\sum_{\alpha}\frac{e_{\alpha}^{3}\rho_{\alpha}}{m_{\alpha}} \label{633a} \end{equation} For the charge-charge correlation, both root points ${\cal L}_{a}$ and ${\cal L}_{b}$ can be dressed as before, leading to a $\approx r^{-10}$ decay with \begin{eqnarray} \langle\hat{c}({\bf r}_{a})\hat{c}({\bf r}_{b})\rangle_{T}&\sim&\frac{1}{|{\bf r}_{a}-{\bf r}_{b}|^{10}} \frac{7\hbar^{4}\beta^{4}}{4} \(\frac{1}{3}\int d{\bf r} |{\bf r}|^{2}K^{D.H.}({\bf r})\)^{2}\nonumber\\ &\sim&\frac{1}{|{\bf r}_{a}-{\bf r}_{b}|^{10}}\frac{7\hbar^{4}\beta^{4}}{\kappa^{4}} \(\sum_{\alpha}\frac{e_{\alpha}^{3}\rho_{\alpha}}{m_{\alpha}}\)^{2} \label{6.34} \end{eqnarray} The fact that the space integral of (\ref{6.30}) vanishes (or equivalently (\ref{6.31}) vanishes at ${\bf k}=0$) is the elementary form of the classical charge sum rule in the Debye-H\"{u}ckel theory (see (\ref{2.1.13})). In the quantum gas, this vestige of classical screening can still operate to improve the decay of the charge-charge correlation from $\approx r^{-6}$ to $\approx r^{-10}$, but not beyond. Finally the susceptibility (\ref{1.2.7d}) (i.e. the response to an external classical charge) decays as $r^{-8}$. Its tail at low density is given by \begin{equation} \chi(r)\sim\frac{4\pi}{r^{8}}\frac{\hbar^{4}\beta^{5}}{8\kappa^{4}}\(\sum_{\alpha} \frac{e_{\alpha}^{3}\rho_{\alpha}}{m_{\alpha}}\)^{2} \label{6.34a} \end{equation} We emphasize that (\ref{6.23}), (\ref{6.33}), (\ref{6.34}) and (\ref{6.34a}) are exact asymptotic results for the multicomponent quantum gas. \subsubsection{A comparison with the standard many-body perturbation theory} \label{subsec-6.1.4} The standard finite temperature many-body perturbation theory is an expansion in powers of the coupling constant (the charge $e^{2}$). Here we adopt the definitions and normalizations of \cite{FeWa71} and \cite{CoMa91}\footnote{There is a minus sign missing in front of the right hand side of \cite[formula (2.4)]{CoMa91}.}. We recall a few facts. The terms of the expansion are represented by Feynman graphs in which bonds $(\alpha_{1},{\bf r}_{1},s_{1} ;\alpha_{2},{\bf r}_{2},s_{2})$ are joined either by free propagators $\delta_{\alpha_{1},\alpha_{2}}G^{0}_{\alpha_{1}}({\bf r}_{2}-{\bf r}_{1},s_{1}-s_{2})$ or by interaction lines $U_{\alpha_{1},\alpha_{2}}({\bf r}_{2}-{\bf r}_{1},s_{1}-s_{2})$ (Figure~\ref{fig9}). At each vertex $(\alpha,{\bf r},s)$ are attached two propagator lines and one interaction line. A basic link between physical quantities and perturbation theory can be established through the imaginary time displaced charge correlation (\ref{1.2.7c}) by \begin{equation}c_{T}({\bf r}, s, {\bf 0}) =-\sum_{\alpha\gamma}e_{\alpha}e_{\gamma} {\cal P}_{\alpha\gamma}({\bf r}, s) \label{6a1} \end{equation} The total polarization ${\cal P}_{\alpha\gamma}({\bf r}, \tau)$ consists of all connected Feynman graphs in which the points $(\alpha,{\bf r},s)$ and $(\gamma,{\bf 0},0)$ are joined by internal lines. ${\cal P}_{\alpha\gamma}({\bf r}, \tau)$ is a time periodic function of period $\beta$. Introducing its space and time Fourier transform ${\cal P}({\bf k},n)$, the structure factor (\ref{1.2.7cc}) and the susceptibility (\ref{1.2.7d}) are given by ($n$ indexes even Matsubara frequencies $\omega_{n}=\frac{\pi n}{\beta}$, $n$ even)\footnote {In general, $n$ is odd for fermionic propagators and even for bosonic propagators, but the charge correlation (\ref{1.2.7c}) involves only quadratic expressions of the Fermi or Bose fields and only even frequencies occur in its time Fourier series.} \begin{equation} S({\bf k})=-\sum_{n=-\infty}^{\infty}\sum_{\alpha\gamma}e_{\alpha}e_{\gamma} {\cal P}_{\alpha\gamma}({\bf k},n) \label{6a2} \end{equation}\begin{equation}\chi({\bf k})=\beta V({\bf k})\sum_{\alpha\gamma}e_{\alpha}e_{\gamma} {\cal P}_{\alpha\gamma}({\bf k}, n=0), \;\;\;\;\;\; V({\bf k})=\frac{4\pi}{|{\bf k}|^{2}} \label{6a3} \end{equation}It is convenient to consider the proper polarization ${\cal P}_{\alpha\gamma}^{\ast}({\bf k},n)$, the set of all polarization parts that cannot be separated into two polarization parts by cutting a single interaction line. ${\cal P}_{\alpha\gamma}({\bf k},n)$ and ${\cal P}_{\alpha\gamma}^{\ast}({\bf k},n)$ are related by a Dyson equation \begin{equation} {\cal P}_{\alpha\gamma}({\bf k},n)=\frac{{\cal P}_{\alpha\gamma}^{\ast}({\bf k},n)} {1-\beta V({\bf k})\sum_{\alpha\gamma}e_{\alpha}e_{\gamma}{\cal P}_{\alpha\gamma}^{\ast}({\bf k},n)} \label{6a4} \end{equation}An effective frequency dependent interaction can be defined by \begin{equation} U^{eff}({\bf k},n)=\frac{V({\bf k})}{1-\beta V({\bf k})\sum_{\alpha\gamma}e_{\alpha}e_{\gamma}{\cal P}_{\alpha\gamma}^{\ast}({\bf k},n)} \label{6a5} \end{equation}This definition is justified by the fact that its zero frequency component coincides with the static effective potential (\ref{1.1.10f}) (with $\varepsilon ({\bf k} )$ given by (\ref{1.2.7d})) \begin{equation} U^{eff}({\bf k},n=0)=\varepsilon^{-1}({\bf k})V({\bf k}) \label{6a6} \end{equation}One has also the relation \begin{equation}{\cal P}_{\alpha\gamma}({\bf k}, n)={\cal P}_{\alpha\gamma}^{\ast}({\bf k},n)+\beta\sum_{\delta\eta} {\cal P}_{\alpha\delta}^{\ast}({\bf k},n)e_{\delta}e_{\eta}U^{eff}({\bf k},n) {\cal P}_{\eta\gamma}^{\ast}({\bf k},n) \label{6a6a} \end{equation} The well known random phase approximation (RPA) amounts to replacing ${\cal P}_{\alpha\gamma}^{\ast}({\bf r},\tau)$ by the lowest order proper polarization term, i.e. the product of two free propagators \begin{equation} {\cal P}_{RPA,\alpha\gamma}^{\ast}({\bf r},s)= -\eta_{\alpha}\delta_{\alpha\gamma}(2s_{\alpha}+1) G^{0}_{\alpha}({\bf r},s)G^{0}_{\alpha}(-{\bf r},-s)=\delta_{\alpha\gamma} {\cal P}_{RPA,\alpha}^{\ast}({\bf r},s) \label{6a7} \end{equation} Hence from (\ref{6a4}) and (\ref{6a5}) (see Figure~\ref{fig10}) \begin{equation} {\cal P}_{RPA,\alpha\gamma}({\bf k},n)=\frac{\delta_{\alpha\gamma} {\cal P}_{RPA,\alpha}^{\ast}({\bf k},n)} {1-\beta V({\bf k})\sum_{\alpha}e_{\alpha}^{2}{\cal P}_{RPA,\alpha}^{\ast}({\bf k},n)} \label{6a8} \end{equation} \begin{equation} U^{eff}_{RPA}({\bf k},n)=\frac{V({\bf k})}{1-\beta V({\bf k})\sum_{\alpha}e_{\alpha}^{2}{\cal P}_{RPA,\alpha}^{\ast}({\bf k},n)} \label{6a9} \end{equation} and with (\ref{6a2}) and (\ref{6a3}), this defines also $S_{RPA}({\bf k})$ and $\chi_{RPA}({\bf k})$. Let us show that the RPA approximation fails to capture the algebraic tails in the correlations\footnote{The present considerations apply only if the temperature is different from zero. At zero temperature, the effective potential shows the long range Friedel oscillations $\cos(2k_{F}r)/r^{3}$ due to the sharpness of the Fermi surface.}. The free propagator has the form \begin{equation}G^{0}_{\alpha}({\bf k},s)=\exp\(-\beta s\(\frac{\hbar^{2}|{\bf k}|^{2}}{2m_{\alpha}}- \mu_{\alpha}\)\) [n^{0}_{\alpha}({\bf k})-\theta(s)] \label{6a10} \end{equation} where $n^{0}_{\alpha}({\bf k})=\(\exp\(\frac{\hbar^{2}\beta |{\bf k}|^{2}}{2m_{\alpha}}- \beta\mu_{\alpha}\)\pm 1\)^{-1}$ is the free Fermi or Bose distribution and $\theta(s)$ is the Heaviside function. It is clearly an infinitely differentiable function of ${\bf k}$ (if $\mu_{\alpha}<0$ for Bose particles). One can deduce from (\ref{6a7}) that the same is true for ${\cal P}_{RPA,\alpha}^{\ast}({\bf r},n)$ and one finds the small ${\bf k}$ behaviour \begin{eqnarray} {\cal P}_{RPA,\alpha}^{\ast}({\bf k},n=0)&=& -\frac{\partial \rho^{0}_{\alpha}}{\partial(\beta\mu_{\alpha})} +O(|{\bf k}|^{2})\neq 0\nonumber\\ {\cal P}_{RPA,\alpha}^{\ast}({\bf k},n\neq 0)&=&O(|{\bf k}|^{2}),\;\,\,{\bf k}\rightarrow 0 \label{6a11} \end{eqnarray} where $\rho^{0}_{\alpha}=(2s_{a}+1)\int d{\bf k} n_{\alpha}^{0}({\bf k}) $ is the density of the free gas. With (\ref{6a8}) this implies that the ${\cal P}_{RPA,\alpha}^{\ast}({\bf k},n)$ are infinitely differentiable at ${\bf k} =0$ and by (\ref{6a2}) and (\ref{6a3}) $S_{RPA}({\bf k})$ and $\chi_{RPA}({\bf k})$ inherit the same property. Thus, in space, these functions decay faster than any inverse power of the distance. From (\ref{6a8}) and (\ref{6a11}) one sees that $\chi_{RPA}({\bf k})$ can be approximated for small ${\bf k}$ by the Debye-like form \begin{equation} \chi_{RPA}({\bf k})\simeq -\frac{\kappa_{RPA}^{2}}{|{\bf k}|^{2}+\kappa_{RPA}^{2}} \label{6a12} \end{equation} with \begin{equation} \kappa_{RPA}^{2}=4\pi \beta\sum_{\alpha}e_{\alpha}^{2} \frac{\partial \rho^{0}_{\alpha}}{\partial(\beta\mu_{\alpha})} \label{6a13} \end{equation} so defining the RPA screening length $\kappa_{RPA}^{-1}$. It is instructive to compare it with the loop screening length $\kappa^{-1}$ (\ref{5.52b}). For this, we introduce the general relation \begin{eqnarray} \frac{\partial}{\partial(\beta\mu_{\alpha})}\rho_{\alpha}&=&\rho_{\alpha}+\int d{\bf r}\rho_{T} (\alpha,{\bf r},\alpha,{\bf 0})\nonumber\\ &=&\rho_{\alpha}+\int d{\bf r}\rho_{T}^{(ex)}(\alpha,{\bf r},{\bf 0})+ \int d{\bf r}\rho_{T}^{(nex)}(\alpha,{\bf r},\alpha,{\bf 0}) \label{6a14} \end{eqnarray} in (\ref{5.52b}), so that $\kappa^{2} $ can also be written as \begin{equation} \kappa^{2}=4\pi\beta\(\sum_{\alpha}e_{\alpha}^{2} \frac{\partial \rho_{\alpha}}{\partial(\beta\mu_{\alpha})}-\sum_{\alpha}e_{\alpha}^{2} \int d{\bf r} \rho_{T}^{(nex)}(\alpha,{\bf r},\alpha,{\bf 0})\) \label{6a15} \end{equation} In the integrals in (\ref{6a14}) and (\ref{6a15}), there are no contributions of coincident points. The loop screening length interpolates between the classical Debye length at low density (as seen from (\ref{5.52b}) when exchange effects can be neglected) and the RPA length at high density. Indeed, in the high density regime, the kinetic energy dominates the Coulomb interaction energy: the gas becomes free so the non-exchange part of the correlation can now be neglected in (\ref{6a15}) and $\kappa^{2}$ tends to the RPA value. As far as the effective potential (\ref{6a5}) is concerned, one finds in the RPA approximation that its zero frequency component $U^{eff}_{RPA}({\bf k},n=0)$ is also Debye-like for small ${\bf k}$ with a screening length slightly different from $\kappa_{RPA}^{-1}$ (for explicit formulae see \cite{CoMa91,Cor96b}). However its non-zero frequency components $U^{eff}_{RPA}({\bf k},n\neq 0)$ are not shielded and keep a pure Coulombic singularity $|{\bf k}|^{-2}$, but within the RPA theory these singularities do not induce any algebraic tails in the particle and charge correlations. For instance, if ${\cal P}_{\alpha\gamma}^{\ast}({\bf k},n)$ is approximated by ${\cal P}_{RPA,\alpha\gamma}^{\ast}({\bf k},n)$ in (\ref{6a6a}), the vanishing of the non-zero frequency components as $|{\bf k}|\rightarrow 0$ kills the $|{\bf k}|^{-2}$ singularity coming from $U^{eff}_{RPA}({\bf k},n\neq 0)$. One may however wonder how these tails emerge in the framework of the many-body perturbation theory. The role played by $U^{eff}_{RPA}$ in Feynman graphs is analogous to that of the resummed bond $F_{l}$ (\ref{5.55}) in loop prototype diagrams. $F_{l}$ has a long range part $W$ (see (\ref{6.6})), but this long range part generates tails in the correlations only if it occurs at least quadratically in prototype graphs (see section \ref{sec-6.1.2}). In the same way, $U^{eff}_{RPA}$ must enter quadratically in Feynman graphs to generate tails in the correlations. This leads to including as a first correction to the RPA approximation the proper polarization insertion ${\cal P}^{\ast}_{1}$ made of two propagator loops linked by two RPA lines of Figure~\ref{fig11}. The effect of this correction has been studied in the one component system \cite{CoMa91}. The Coulombic singularities of $U^{eff}_{RPA}({\bf k},n\neq 0)$ come now non-trivially into the game and one finds that ${\cal P}_{1}^{\ast}({\bf k}, n=0)$ and ${\cal P}_{1}^{\ast}({\bf k}, n\neq0)$ have respectively a $|{\bf k}|^{3}$ and a $|{\bf k}|^{7}$ singular term in their small ${\bf k}$ expansion. One can then infer through the Dyson equation (\ref{6a4}) that these singular terms induce $r^{-10}$ and $r^{-8}$ tails in $S({\bf r})$ and $\chi({\bf r})$ respectively, in accordance with the findings of section~\ref{subsec-6.1.3}. The graphs of Figure~\ref{fig11} have been considered previously in \cite{MaAs87,LaVo87} in connection with calculating the cohesive energy of metals and inter-atomic potentials at zero temperature. It is not possible to establish in general strict correspondence rules between the many-body perturbation algorithm and that of the loop formalism. Since the reference system of the many-body perturbation scheme is the noninteracting gas, any approximation there gives expressions in terms of free quantities. In the loop expansion,weights and bonds in Mayer diagrams contain already the coupling constant $e^{2}$ and the statistics in a non-perturbative way, and they involve the exact particle densities. This difference is clearly seen in the two screening lengths $\kappa^{-1}_{RPA}$ and $\kappa^{-1}$ that depend respectively on the free densities $\rho^{0}_{\alpha}$ and exact densities $\rho_{\alpha}$. One can however establish a precise relation between the loop chain potential and the RPA effective potential in the Boltzmann limit when statistics are neglected (see \cite[section IV]{CoMa91}). The two formalisms play complementary roles. The many-body formalism is suitable for performing low temperature and ground state calculations. By analytic continuation in time it can also be used to compute real time displaced correlations. Information on ground state quantities in the loop formalism would necessitate applying the Feynman-Kac formula backwards at some point before taking the $\beta\rightarrow\infty$ limit. However the loop formalism provides a better insight into Coulombic correlations and screening. For instance the reduction to a gas of multipoles is not visible in the many-body framework where the effective potential $U_{RPA}^{eff}$ keeps an unscreened bare Coulomb part. Low density behaviors are much more conveniently obtained from the loop expansion: it would not be straightforward to recover the exact results of section \ref{sec-6.1} from an analysis of Feynman graphs. The next section offers another example of the usefulness of the loop formalism. \subsection{The virial equation of state} \label{6.1} The derivation of the equation of state of a quantum mechanical plasma is an old problem of practical and theoretical interest. In particular, a precise knowledge of this equation plays an important role in astrophysical conditions, and it is not possible to give credit here to the large literature devoted to this question. We quote the thorough study of the quantum Coulomb gas undertaken by Ebeling and collaborators by the usual methods of the many-body problem. This work is reported in \cite{KKER86} and some comments will be given at the end of this section. In the sequel, we put forward that the loop representation is also a convenient tool for calculating the equation of state at low density and fixed temperature. In particular the formalism enables simultaneously keeping track, in a coherent and systematic way, of the quantum effects and of the long range of the Coulomb potential, a notoriously difficult problem. A detailed presentation of all calculations can be found in \cite{AlPe92,AlCoPe94,AlCoPe95,AlPe96} and a review on this subject is presented in \cite{Ala94}. Here we outline the general computational scheme and illustrate some salient points. To evaluate the density expansion of the thermodynamic functions, it is appropriate to start from a standard identity relating the free energy $f(\beta, \{\rho_\alpha\})$ per unit volume to the two--particle correlation with a variable coupling parameter \begin{equation} \beta f =\beta f_{{\rm ideal}} +\frac{\beta}{2} \sum_{\alpha_{a}\alpha_{b}}e_{\alpha_a} e_{\alpha_b}\int_0^1 dg\int d{\bf r}\rho_{T,g}(\alpha_a,{\bf r},\alpha_b,0)\frac{1}{|{\bf r}|} \label{6.35} \end{equation} Here $\rho_{T,g}(\alpha_a,{\bf r},\alpha_b,{\bf 0})$ is the truncated two-particle correlation (without the contribution of coincident points) for a system of charges interacting by the Coulomb potential $g/|{\bf r}|$ with a dimensionless coupling parameter $g,\;\;0\leq\;g\;\leq 1$. For each $g$, $\rho_{T,g}(\alpha_a,{\bf r},\alpha_b,0)$ must be evaluated at the same values of the densities and temperature that occur as arguments of $f(\beta, \{\rho_\alpha\})$. The particle correlation $\rho_{T,g}(\alpha_a,{\bf r}_a,\alpha_b,{\bf r}_b)$ is in turn related to the loop correlations through (\ref{5.22}), (\ref{5.23}) and (\ref{5.23a}) and the loop correlations have the diagrammatic expansion discussed in section \ref{sec-5.2}. Then one has to organize the contributions of this expansion in terms that are of increasing order in the particle densities and insert them in (\ref{6.35}). This will yield the desired expansion for the free energy. Finally the pressure $P(\beta, \{\rho_\alpha\})$ follows from the usual formula \begin{equation} P=\sum_\alpha \rho_\alpha \frac{\partial}{\partial\rho_\alpha }f\;-\;f \label{6.36} \end{equation} The calculation has been carried out up to order $\rho^{5/2}.$ \subsubsection{The Maxwell-Boltzmann pressure} \label{subsec-6.2.1)} At this point one can make the following comment about the treatment of exchange effects. It is clear from (\ref{5.13}) that $q$-particle exchanges are of decreasing order $\exp(\beta\mu_\alpha q)$ as $q$ grows ($\mu_\alpha$ negative); hence they can also be treated perturbatively at low density. This is the view point adopted in \cite{AlCoPe95} and that is reported here. One considers first the Maxwell-Boltzmann (MB) Coulomb gas, and one includes the exchange effects in a second perturbative stage. The MB system is simpler since loops reduce to filaments. Also the exchange part (\ref{5.23a}) of the correlation can be ignored in the first stage. Dropping here the Fermi statistics should not alarm the reader. All terms of the low density expansion remain well defined, since these terms involve only finitely many body Hamiltonians, and thus the stability of matter plays no role at any finite order. As already noted, the local singularity of the Coulomb potential is smoothed by the usual laws of quantum mechanics (the uncertainty principle) as in atomic physics. In the works \cite{AlCoPe94,AlCoPe95}, prototype diagrams are constructed with a slightly different definition of resummed bonds than in section \ref{subsec-5.2.3}. One has here, keeping similar notations as in section \ref{subsec-5.2.2} and \ref{subsec-5.2.3} \begin{eqnarray}\label{6.36a} &&\text{the charge-charge bond } F_{cc}, \text{ the charge-dipole bonds } F_{cd},\;F_{dc}\nonumber \\ &&\text{the dipole-dipole bond } F_{dd}, \text{ the residual bond } F_{r} \end{eqnarray} The bond $F_{cc}$ is identical to (\ref{5.45}) (with $q_{a}=q_{b}=1$); $F_{cd}$ and $F_{dc}$ are defined as in (\ref{5.48}) and (\ref{5.48a}) but retaining only the dipole part \begin{equation} F_{cd}(1,2)=-\beta e_{1}e_{2}\int_{0}^{1}ds_{2}(\lambda_{2}{\bf \xi}_{2}(s_{2})\cdot{\bf \nabla}_{2}) \frac{\exp(-\kappa |{\bf R}_{1}-{\bf R}_{2}|)}{|{\bf R}_{1}-{\bf R}_{2}|} \label{6.37} \end{equation} The bond $F_{dd}$ is equal to the dipole part of the second term in (\ref{5.50}), i.e \begin{eqnarray} && F_{dd}(1,2)=-\beta e_{1}e_{2} \nonumber \\ &&\hspace{-1in}\times \int_{0}^{1}ds_{1}\int_{0}^{1}ds_{2} (\lambda_{1}{\bf \xi}_{1}(s_{1})\cdot{\bf \nabla}_{1})(\lambda_{2}{\bf \xi}_{2}(s_{2})\cdot{\bf \nabla}_{2}) \bigg (\frac{\exp(-\kappa |{\bf R}_{1}-{\bf R}_{2}|)-1}{|{\bf R}_{1}-{\bf R}_{2}|}\bigg) \label{6.38} \end{eqnarray} while $F_{r}$ incorporates all the other contributions (higher order multipoles and quantum effects). Note that in this decomposition we have the two types of long range $ r^{-3}$-bonds $F_{dd}$ and $F_{r}$ (in contrast to the prototype graphs of subsection \ref{subsec-5.2.3} where only $F_{l}$ is dipolar). As already explained in subsection \ref{subsec-5.2.5}, there are two sources for the density dependence: the weight $\rho(\alpha, {\bf \xi})$, that has itself an expansion in the particle densities $\rho_{\alpha}$, and the screening length $\kappa$ occurring in the resummed bonds. One performs first the activity expansion of $\rho_{\alpha}$ with the help of appropriate prototype graphs (here the bonds are as in (\ref{6.36a}) but weights are activities), and one eliminates activities in favor of densities with the help of the relation $\rho_{\alpha}=\int D({\bf \xi}) \rho(\alpha,{\bf \xi})$. The analysis shows that $\rho(\alpha,{\bf \xi})$ is given as double power series in $\rho^{1/2}$ and $\ln\rho$ with first terms \begin{eqnarray} \rho(\alpha,{\bf \xi})&=&\rho_{\alpha}+\sum_{\gamma}\rho_{\alpha}\rho_{\gamma}\int d{\bf r}\int D({\bf \xi}_{1})\label{6.39}\\ &\times &\left[\exp\(-\beta e_{\alpha}e_{\gamma}\int_{0}^{1}dsV(|{\bf r}+\lambda_{\gamma}{\bf \xi}_{1}(s)- \lambda_{\alpha}{\bf \xi}(s)|)\)\right.\nonumber\\ &-&\left.\int D({\bf \xi}_{2})\exp\(-\beta e_{\alpha}e_{\gamma}\int_{0}^{1}dsV(|{\bf r}+\lambda_{\gamma}{\bf \xi}_{1}(s)- \lambda_{\alpha}{\bf \xi}_{2}(s)|)\)\right]\nonumber \\ &+& O(\rho^{5/2})\nonumber \end{eqnarray} These series are introduced in the statistical weights of the expansion of the Ursell function. Finally one obtains also the free energy (\ref{6.35}) and the pressure (\ref{6.36}) as double power series in $\rho^{1/2}$ and $\ln\rho$. Half powers of $\rho$ come from the Debye length $\kappa^{-1}\sim\rho^{-1/2}$ in the screened potential, while $\ln\rho$ terms occur because the $r^{-3}$-bonds $F_{dd}$ and $F_{r}$ are at the border line of integrability in three dimensions. The density dependence contained in the Debye length can be extracted by a scaling analysis. Introducing the dimensionless variable $x=\kappa|{\bf R}_{1}-{\bf R}_{2}|$ one sees that $F_{cc}$ scales as $\kappa$, $F_{cd}\;(F_{dc})$ as $\kappa^{2}$, and $F_{dd}$ as $\kappa^{3}$. More generally a higher order multipole term of the form $$ ({\bf \xi}_{1}(s_{1})\cdot{\bf \nabla}_{1})^{k_{1}}({\bf \xi}_{2}(s_{2})\cdot{\bf \nabla}_{2})^{k_{2}} \frac{\exp(-\kappa |{\bf R}_{1}-{\bf R}_{2}|)}{|{\bf R}_{1}-{\bf R}_{2}|} $$ scales as $\kappa^{k_{1}+k_{2}+1}$. In $F_{r}$ the quantum effects (associated with the de Broglie length $\lambda_{\alpha}$) and classical screening (associated with the Debye length $\kappa^{-1}$) can be disentangled by a decomposition of the form \begin{equation} F_{r}=F_{q}(1+\kappa H_{1}+\kappa^{2}H_{2}+\cdots)+\kappa^{2}G_{2}+\kappa^{3}G_{3}+\cdots \label{6.40} \end{equation} where $F_{q}$ is the quantum (truncated) bond \begin{eqnarray} F_{q}(1,2)&=&\exp\left(-\beta_{12}\int_{0}^{1}dsV(|{\bf r}+\lambda_{\alpha_{1}}{\bf \xi}_{1}(s)- \lambda_{\alpha_{2}}{\bf \xi}_{2}(s)|)\right)\nonumber\\ &-&1+\frac{\beta_{12}}{r}-\frac{\beta_{12}^{2}}{2r^{2}}+\beta_{12}\int_{0}^{1}ds [\lambda_{\alpha_{1}}{\bf \xi}_{1}(s)\cdot{\bf \nabla}- \lambda_{\alpha_{2}}{\bf \xi}_{2}(s)\cdot{\bf \nabla}]\(\frac{1}{r}\)\nonumber\\ \label{6.41} \end{eqnarray} and the $H_{i}$ and $G_{i}$ are scaled functions involving contributions of higher order multipoles. We distinguish three different types of effects in this game: \begin{itemize} \item classical screening manifested by the occurrence of the Debye length in the scaled bonds \item quantum diffraction effects due to the coupling of quantum fluctuations ${\bf \xi}$ to the classical Debye potential as in $F_{cc}, \;F_{cd}$ and $F_{dd}$ \item quantum bound and scattering states appearing through $F_{r}$. \end{itemize} As an illustration the table in Figure~\ref{fig12} gives a (non exhaustive) list of examples of graphs contributing to different effects in the MB free energy. The calculations are tractable up to order $\rho^{5/2}.$ There are two simplifications up to this order: internal weights $\rho(\alpha,{\bf \xi})$ can always be replaced by their lowest order $\rho_{\alpha}$ and only low order multipoles contribute in view of the above mentioned scaling arguments. The MB pressure is given by the terms (\ref{6.48a}) to (\ref{6.51}) in the final result stated in the next subsection. \subsubsection{The exchange contributions and the final result} \label{subsec-6.2.1} To include the exchange contributions perturbatively, one has to come back to the expression (\ref{1.2.5}) of the grand-partition function with statistics and collect together all the terms involving the exchange of exactly n particles \begin{equation} \Xi_{\Lambda}=\Xi_{\Lambda}^{{\rm MB}}+\sum_{n=2}^{\infty}\Xi_{\Lambda}^{(n)}=\Xi_{\Lambda}^{{\rm MB}} \(1+\sum_{n=2}^{\infty}\frac{\Xi_{\Lambda}^{(n)}}{\Xi_{\Lambda}^{{\rm MB}}}\) \label{6.44} \end{equation} where $\Xi_{\Lambda}^{{\rm MB}}$ is the Maxwell-Boltzmann partition function and the contribution with $N_{\alpha}$ exchanged particles has necessarily an activity factor $\exp(\beta\mu_{\alpha}N_{\alpha})$. The expansion (\ref{6.44}) generates in turn an expansion of the grand-canonical pressure \begin{equation} P=\beta^{-1}\lim_{|\Lambda|\rightarrow\infty}\frac{1}{ |\Lambda|}\ln\Xi_{\Lambda}=P^{{\rm MB}}+\sum_{n=2}^{\infty}P^{(n)} \label{6.45} \end{equation} with $P^{{\rm MB}}$ the Maxwell-Boltzmann pressure and $P^{(n)}$ the contribution of $n$ exchanged particles. The matrix element in (\ref{1.2.5}) corresponding to a two particle exchange, say $\langle{\bf r}_{2},{\bf r}_{1},{\bf r}_{3},\ldots,{\bf r}_{N}|\exp(-\beta H_{\Lambda,N})|{\bf r}_{1},{\bf r}_{2},{\bf r}_{3},\ldots,{\bf r}_{N}\rangle$, will involve $N-2$ closed filaments ${\bf r}_{i,i}(s),\;i=3,\ldots,N$ together with two open filaments ${\bf r}_{1,2}(s)$ and ${\bf r}_{2,1}(s)$. Keeping these two open filaments fixed, but performing the configuration integrals and grand-canonical sums on all the remaining closed filaments leads to the consideration of the inhomogeneous Maxwell-Boltzmann partition function $\Xi_{\Lambda}^{(2)(inhom)}$. It is the grand-partition function of a system of (closed) filaments in presence of the potential due to these two open filaments. Similarly, higher order exchange terms can be viewed as partition functions of closed filaments in presence of an external potential created by the open ones. The density of filaments $\rho^{(inhom)}({\cal F})$ in the inhomogeneous system can be expanded in terms of the correlations $\rho({\cal F}), \;\rho_{T}({\cal F}_{1},{\cal F}_{2})$ of the homogeneous gas by standard methods. Finally, one can apply to the latter correlations all the diagrammatic techniques previously described to obtain the density expansion of the MB gas. There is however one caveat. The particle densities are given in principle by the usual relation \begin{equation} \rho_{\alpha}=\frac{\partial}{\partial \mu_{\alpha}}P(\{\mu_{\gamma}\}) \label{6.46} \end{equation} where $P$ is the total grand-canonical pressure (\ref{6.45}) including the statistics. But $P^{{\rm MB}}$ has been computed directly as function of densities that we call now $\rho^{{\rm MB}}_{\alpha}$ within the MB approximation. Then the $\rho^{{\rm MB}}_{\alpha}$ can be eliminated in favor of the true densities $\rho_{\alpha}$ by noting that both $P$ and $P^{{\rm MB}}$ are evaluated at the same values of the chemical potentials $\mu_{\alpha}$. Hence we have the thermodynamic relation \begin{equation} \mu_{\alpha}=\frac{\partial}{\partial \rho_{\alpha}^{{\rm MB}}}f^{{\rm MB}}(\{\rho_{\gamma}^{{\rm MB}}\}) \label{6.47} \end{equation} with $f^{{\rm MB}}$ the free energy of the MB gas. Using (\ref{6.46}) together with (\ref{6.47}) enables us to express the $\mu_{\alpha}$ and $\rho^{{\rm MB}}_{\alpha}$ in terms of the desired particle densities $\rho_{\alpha}$. The final result is \cite{AlPe96} \begin{eqnarray} \beta P&=&\sum_{\alpha}\rho_{\alpha}-\frac{\kappa^{3}}{24\pi}\label{6.48a}\\ &+&\pi\beta^{3}\sum_{\alpha\gamma}\rho_{\alpha}\rho_{\gamma}e_{\alpha}^{3} e_{\gamma}^{3}\left[ \frac{(\ln 2-1)}{6}+\(\frac{1}{3}-\frac{3}{4}\ln 2+\frac{1}{2}\ln 3\)\beta \kappa e_{\alpha}e_{\gamma}\right]\nonumber\\ &+&C_{1}\beta^{5}\kappa^{-1}\sum_{\alpha\gamma\delta} \rho_{\alpha}\rho_{\gamma}\rho_{\delta}e_{\alpha}^{3}e_{\gamma}^{4}e_{\delta}^{3} +C_{2}\beta^{6}\kappa^{-3}\sum_{\alpha\gamma\delta\eta} \rho_{\alpha}\rho_{\gamma}\rho_{\delta}\rho_{\eta}e_{\alpha}^{3}e_{\gamma}^{3} e_{\delta}^{3}e_{\eta}^{3}\nonumber\\ & &\label{6.49}\\ &+&\frac{1}{16}\beta^{2}\hbar^{2}\kappa^{3}\sum_{\alpha}\rho_{\alpha} \frac{e_{\alpha}^{2}}{m_{\alpha}}\label{6.50}\\ &-&\frac{\pi}{\sqrt{2}}\sum_{\alpha\gamma}\rho_{\alpha}\rho_{\gamma}(1+\frac{3}{2}\beta\kappa e_{\alpha }e_{\gamma})\left[\lambda_{\alpha\gamma}^{3} Q(x_{\alpha\gamma})+\frac{\sqrt{2}}{3}\beta^{3}e_{\alpha}^{3}e_{\gamma}^{3} \ln(\kappa\lambda_{\alpha\gamma})\right]\nonumber\\ & &\label{6.51}\\ &+&\frac{\pi}{\sqrt{2}}\sum_{\alpha}\frac{(-)^{2\sigma_{\alpha}+1}}{2\sigma_{\alpha}+1} \rho_{\alpha}^{2}\lambda_{\alpha\alpha}^{3}\(1+\frac{3}{2}\beta\kappa e_{\alpha}^{2}\)E(x_{\alpha\alpha})\label{6.51a}\\ &+&O(\rho^{3}\ln \rho)\nonumber \end{eqnarray} with \begin{eqnarray*} && \kappa= (4\pi\beta\sum_{\alpha}e_{\alpha}^{2}\rho_{\alpha})^{1/2}, \;\lambda_{\alpha\gamma} =(\beta\hbar^{2}/m_{\alpha\gamma})^{1/2},\;\\ && m_{\alpha\gamma}= m_{\alpha}m_{\gamma}/(m_{\alpha}+m_{\gamma}),\; C_{1}=15.205\pm.001,\;C_{2}=-14.733\pm.001 \end{eqnarray*} In (\ref{6.51}), $x_{\alpha\gamma}=-\sqrt{2}\beta e_{\alpha}e_{\gamma}/\lambda_{\alpha\gamma}$ and $Q(x_{\alpha\gamma})$ is the quantum second order virial coefficient first introduced in \cite{Ebe67,Ebe68} \begin{eqnarray} Q(x_{\alpha\gamma})&=&\frac{1}{\sqrt{2}\pi\lambda_{\alpha\gamma}^{3}}\lim_{R\rightarrow\infty} \left\{\int_{r<R}d{\bf r}\left[(2\pi\lambda_{\alpha\gamma}^{2})^{3/2}\langle{\bf r}|\exp(-\beta H_{\alpha\gamma})|{\bf r} \rangle \right.\right.\nonumber\\ &-&\left.\left.1+\frac{\beta e_{\alpha}e_{\gamma}}{r}-\frac{\beta^{2} e_{\alpha}^{2}e_{\gamma}^{2}}{2r^{2}}\right] +\frac{2\pi}{3}\beta^{3}e_{\alpha}^{3}e_{\gamma}^{3} \(\ln\(\frac{3\sqrt{2}R}{\lambda_{\alpha\gamma}}\)+C\)\right\}\nonumber\\ \label{6.52} \end{eqnarray} while $E(x_{\alpha\alpha})$ is the exchange integral \begin{equation} E(x_{\alpha\alpha})=2\sqrt{\pi}\int d{\bf r} \langle-{\bf r}|\exp(-\beta H_{\alpha\alpha})|{\bf r}\rangle \label{6.53} \end{equation} In (\ref{6.52}) and (\ref{6.53}) $H_{\alpha\gamma}$ is the relative one-body Coulomb Hamiltonian for a particle of mass $m_{\alpha\gamma}$ submitted to the potential $\frac{e_{\alpha}e_{\gamma}}{r}$, and $C=.577216\ldots$ is the Euler-Mascheroni's constant. The formula (\ref{6.48a})-(\ref{6.51a}) incorporates and generalizes several earlier results. The perfect gas contribution $\sum_{\alpha}\rho_{\alpha}$ (the dominant one) has first been established rigorously in \cite{LePe73}, completed by \cite{Hug85} \footnote{Hughes studies the continuity of the free energy at vanishing density in relation to the techniques of Fefferman discussed in the next chapter.}. The next term is the lowest order correction to the free gas: one recognizes the familiar Debye-H\"{u}ckel contribution (\ref{2.1.2b}). The quantum diffraction effects appear in the term (\ref{6.50}) at the order $\rho^{5/2}$. The term (\ref{6.51}) incorporates the total contribution of bound and scattering states of the two-body Coulomb Hamiltonian\footnote{The $\ln (\kappa\lambda_{\alpha\gamma})$ term can be recombined with $Q(x_{\alpha\gamma})$. The splitting is introduced to have $Q(x_{\alpha\gamma})$ defined as in \cite{KKER86}. Note also that our de Broglie length differs from that in \cite{KKER86} by a factor $\sqrt{2}$.}. The regularization of the integral in (\ref{6.52}), which insures that $Q(x_{\alpha\gamma})$ is finite, is not arbitrary but comes from the structure of the bond $F_{q}$ (\ref{6.41}). One has here a $\rho^{2}$ contribution, the proper second order virial term, and a $\rho^{5/2}$ multiplicative correction that arises as a many-body mean field effect. Finally, the exchange term (\ref{6.51a}) has a similar structure. The terms up to order $\rho ^{5/2}$, including the exact $\rho ^{2}$-virial contribution (\ref{6.51}) are found in the work of Ebeling \cite{Ebe67,Ebe68} and related references \cite{KKER86}. The full formula (\ref{6.48a}-\ref{6.51a}), up to the diffraction term (\ref{6.50}) appears in \cite{KKER86}. These works use the method of the effective potential (or method of Morita). In this method, one associates to the quantum Coulomb gas an equivalent classical system of point particles interacting with an effective many-body interaction. In the case of the MB gas, this effective many-body potential $\Psi_{N}$ is defined by interpreting the diagonal element $\langle\{{\bf r}_{i}\}|\exp(-\beta H_{N})|\{{\bf r}_{i}\}\rangle$ as a classical Boltzmann factor \begin{equation}\langle\{{\bf r}_{i}\}|\exp(-\beta H_{N})|\{{\bf r}_{i}\}\rangle=\prod_{i=1}^{N}\(\frac{1}{2\pi\lambda_{\alpha_{i}}^{2}}\)^{3/2} \exp(-\beta\Psi_{N}({\bf r}_{1},\ldots,{\bf r}_{N})) \label{6.54} \end{equation}One can decompose $\Psi_{N}$ into a sum of two-body, three-body and higher order interactions, the two-body potential being \begin{equation} \Psi_{2}({\bf r}_{1},{\bf r}_{2})=-\beta^{-1} \ln \left[\prod_{i=1}^{2}\(2\pi\lambda_{\alpha_{i}}^{2}\)^{3/2} \langle{\bf r}_{1},{\bf r}_{2}|\exp(-\beta H_{2})|{\bf r}_{1},{\bf r}_{2}\rangle\right] \label{6.55} \end{equation} Calculations up to order $\rho^{5/2}$ have been performed by keeping only the two-body potential (\ref{6.55}), see references in \cite{KKER86}. The long range Coulombic part of (\ref{6.55}) can again be eliminated by summing the convolution chains and one can use the Abe-Meeron diagrammatics. This procedure gives the above formula for the pressure except (\ref{6.50}). It turns out that one has to take the three-body interaction into account to recover the diffraction term. One can specialize the equation of state to the one component plasma and take the classical limit\footnote{This can be achieved by considering an asymmetrical two component system and letting $m_{2}\rightarrow\infty,\; e_{2}\rightarrow 0,\rho_{2}\rightarrow\infty $ with $\rho_{2}e_{2}$ finite and $\rho_{2}e_{2}+\rho_{1}e_{1}=0$. The one component plasma remains well behaved in the classical limit because of the spreading of the positive charge.}: these classical terms coincide with those calculated in \cite{CoMu69}. One retrieves also the $\hbar$-correction $\pi\beta^{2}e^{2}\rho^{2}\hbar^{2}/6m$ to the classical equation of state \cite{PoHa73}. This correction has to be of order $\rho^{2}$: indeed one checks that the $\hbar^{2}\rho^{5/2}$ contributions coming from (\ref{6.50}) and (\ref{6.51}) compensate exactly. Finally high temperature series are recovered by expanding $Q(x)$ and $E(x)$ in powers of $\sqrt{\beta}$ (such expansions are derived in \cite{KKER86}, and these series agree with the expressions obtained by the methods of \cite{Dew62,Dew66,DSSK95}. Explicit formulae for various thermodynamic functions as well as their specialization to these particular cases can be found in \cite{AlPe96}. The generalization of the low density equation of state to a uniformly magnetized plasma is given in \cite{Cor97,Cor98b}. \def\({\left(} \def\){\right)} \def{\bf r}{{\bf r}} \def{\bf \xi}{{\bf \xi}} \def{\cal F}{{\cal F}} \def{\cal L}{{\cal L}} \def\underline{\bf m}{\underline{\bf m}} \def\underline{\bf e}{\underline{\bf e}} \def\underline{\bf N}{\underline{\bf N}} \def\underline{\mbox{\boldmath $\mu$\unboldmath}}{\underline{\mbox{\boldmath $\mu$\unboldmath}}} \def\mbox{\boldmath $\mu$\unboldmath}{\mbox{\boldmath $\mu$\unboldmath}} \def\mbox{\boldmath $\lambda$\unboldmath}{\mbox{\boldmath $\lambda$\unboldmath}} \newpage \section{The atomic and molecular limit} \setcounter{page}{1} \label{chapter-7} \subsection{The electron-proton gas in the Saha regime} \label{sec-7.1} The diagrammatic technique developed in chapter~\ref{chapter-loops} has provided a systematic low density expansion of the equation of state at fixed temperature. In particular, the lowest order term $\beta P=\sum_{\alpha=1}^{\cal S}\rho_{\alpha}$ in (\ref{6.48a}) represents a mixture of perfect Maxwell-Boltzmann gases constituted by the ${\cal S}$ species of fully dissociated charges: it is a free plasma state. Possible two-body bound states appear only as corrections in the second order virial coefficient. In this chapter, we will treat another regime where the basic constituents are now chemically bound entities behaving as ideal substances. This regime ( called the Saha regime) is characterized by a joint limit where both the density and the temperature go to zero in a coupled way. The (exponentially fast) rate at which the density is reduced as $T \rightarrow 0$ determines a certain energy-entropy balance, selecting in turn the formation of some specific chemical species. Indeed low temperature favors binding over ionization, whereas low density, by increasing the available phase space per particle, favors dissociation. This limit is called the molecular (or atomic) limit. It was first formulated in precise terms by \cite{Fef85} and further studied by \cite{CLY89,GrSc95a}. Equilibrium ionization phases in this context are discussed in \cite{MaMa90}. The Saha regime is also described in the usual language of many-body perturbation theory in \cite{EKK76,KKER86}. In order to understand the issues involved in a simple setting, we consider the electron-proton (e-p) system and the possible formation of hydrogen atoms (see also \cite{Fef86,Mar93} for general discussion and background). Let us adopt for a moment the thermodynamic view point that a chemist would take in presence of three preformed species, the electrons (e), the protons (p) and the hydrogen atoms (a) in their ground state. As a first approximation he takes into account the binding energy of each chemical species, but otherwise treats them as perfect gases of point particles. In this approximation, the grand-canonical densities are \footnote{This density $\rho_{j}=\rho_{{\rm ideal},j}$ as well as the densities (\ref{7.32}) are those of ideal gases to which the interacting system will be eventually compared. For brevity, we drop the mention ideal here and in (\ref{7.32}).} \begin{equation}\rho_{j}=\(\frac{m_j}{2\pi \beta \hbar^2}\)^{3/2}\exp({-\beta(E_j-\mu_j))},\;\;\,\;\;\;j=e,\:p,\:a \label{7.1} \end{equation}where $\mu_j$ are the respective chemical potentials, $m_a=m_e+m_p,\;\; E_e=E_p=0$ and $E_a<0$ is the ground state energy of the hydrogen atom. The law of chemical equilibrium for the dissociation reaction $e+p\leftrightarrow a$ requires \begin{equation} \mu_a=\mu_e+\mu_p \label{7.2} \end{equation}and one also must have the neutrality $\rho_e=\rho_p $. Taking (\ref{7.2}) into account and introducing the combinations \begin{equation} \mu=\frac{\mu_e+\mu_p}{2},\;\;\,\nu=\frac{\mu_e-\mu_p}{2} \label{7.4} \end{equation}it is easily seen that the neutrality condition imposes the choice \begin{equation}\nu=\nu_{0}(\beta)=\frac{3}{4\beta}\ln\frac{m_p}{m_e} \label{7.5} \end{equation}and hence (\ref{7.1}) becomes \begin{equation} \rho_e = \rho_p=\(\frac{\sqrt{m_em_p}}{2\pi\beta\hbar^2}\)^{3/2}\exp(\beta\mu), \ \ \ \rho_a = \(\frac{m_e+m_p}{2\pi\beta\hbar^2}\)^{3/2}\exp(-\beta(E_a-2\mu)) \label{7.6} \end{equation} Finally, since the species are treated as non-interacting, the pressure obeys the law of perfect gases \begin{eqnarray} \beta P &=& \rho_e+\rho_p+\rho_a =\(\frac{1+\gamma}{2}\)\rho\label{7.7} \end{eqnarray} In (\ref{7.7}), the pressure is expressed in terms of the total number density $\rho $ of protons and electrons $\rho=\rho_e+\rho_p+2\rho_a$ and we have introduced the Saha coefficient for the degree of ionization \begin{equation} \gamma=\frac{\rho_e}{\rho_e+\rho_a},\;\;\;\;\;\;\,\;\,0\:\leq\:\gamma\:\leq\:1 \label{7.10} \end{equation} A low density-low temperature regime can clearly be obtained by choosing $\mu$ negative and letting $\beta\rightarrow \infty$. Let us examine various cases noting from (\ref{7.6}) that $\rho_a/\rho_e\sim \exp(-\beta(E_a-\mu)$. \begin{description} \item[(i)] Fix $\mu<E_a$. The atomic density is exponentially small compared to that of the electrons and protons; there is full dissociation ($\gamma\rightarrow 1$) and the equation of state reduces to $\beta P=\rho$ as in the low density expansion (\ref{6.48a}). \item[(ii)] Fix $\mu>E_a$ (but $\mu<E_a/2$). One obtains the opposite situation where the pure atomic phase dominates; now $\gamma\rightarrow 0$ and $\beta P=\frac{1}{2}\rho$ as $\beta\rightarrow \infty$. \item[(iii)] At the borderline $\mu=E_a$, $\rho_e$ and $\rho_a$ are of the same order. More precisely, replace $\mu$ by \begin{equation} \tilde{\mu}(\beta)=E_a +\lambda\beta^{-1}+o(\beta^{-1}) \label{7.11} \end{equation} for some $\lambda,-\infty<\lambda<\infty$, i.e. approach the point $(E_a,0)$ in the $(\mu,T)$ plane along a direction having a finite slope $\lambda$. Then one obtains ionization equilibrium phases that interpolate between the fully ionized and the atomic ones. Their degree of ionization found from (\ref{7.6}), (\ref{7.10}) and (\ref{7.11}) is \begin{equation} \gamma=\(\(\frac{m_e+m_p}{\sqrt{m_em_p}}\)^{3/2}\exp (\lambda) +1\)^{-1} \label{7.12} \end{equation} and they obey the equation of state (\ref{7.7}). Written in terms of the pressure, the Saha coefficient reads after a short computation \begin{equation} \gamma=\(1+\beta P \(\frac{2\pi \beta \hbar^2 (m_e+m_p)}{m_em_p}\)^{3/2}\exp(-\beta E_a)\)^{-1/2} \label{7.13} \end{equation} \end{description} This discussion makes clear that to an increase of $\mu$ corresponds an increase of the densities that favors the electron-proton binding when $\mu$ crosses the value $E_a$. How can we justify this picture from statistical mechanics? Consider now the exact infinite volume pressure $P(\beta,\mu_e,\mu_p)$ of the e-p system in the grand-canonical ensemble, and recall from section~\ref{sec- neutrality} that it depends only on $\mu=\frac{\mu_e+\mu_p}{2}$ \begin{equation} P(\beta,\mu_e,\mu_p)=P(\beta,\mu) \label{7.14} \end{equation} One may expect that the equation of state associated with $P(\beta,\mu)$ becomes close to (\ref{7.7}) as $\beta\rightarrow \infty$ for appropriate values of $\mu$ as in cases (i)-(iii). The theorem on the atomic and molecular limit will gives a precise formulation of this assertion. In the elementary considerations just presented, we have not taken into account, among other things, that electrons and protons can form other complexes as, for instance, hydrogen molecules. Let us examine this issue from the view point of statistical mechanics. If typical configurations of charges are indeed those of a dilute gas of hydrogen atoms (case (ii)), these configurations will be mainly formed of e-p pairs, the extension of a pair being of the order of the Bohr radius $a_B$, whereas two different pairs are at distance $\rho_a^{-1/3} \;(a_B\ll \rho_a^{-1/3}) $. If we choose at random in this configuration a region $D$ of linear extension $R$ with \begin{equation}a_B\:\ll \:R\:\ll\:\rho_a^{-1/3} \label{7.15} \end{equation}we will observe that this region is empty most of the time, but if it contains something it is exactly one hydrogen atom, except on rare occasions. In the grand-canonical formalism, it means that one must be able to find chemical potentials $\mu_e,\mu_p$ such that for $\beta\:\gg\:1$ and all $(N_e,N_p)\neq (0,0),(1,1)$ \begin{equation}p_{00}(\beta,\mu_e,\mu_p,D)\gg p_{11}(\beta,\mu_e,\mu_p,D)\gg p_{N_eN_p}(\beta,\mu_e,\mu_p,D) \label{7.16} \end{equation}where \begin{equation} p_{N_eN_p}(\beta,\mu_e,\mu_p,D)=\frac{1}{\Xi_{D}}\exp(\beta(\mu_eN_e+\mu_pN_p)) \mbox{Tr}\exp(-\beta H_{D,N_eN_p}) \label{7.17} \end{equation}is the probability to find exactly $N_e$ electrons and $N_p$ protons in $D$. Thus the condition (\ref{7.16}) means indeed that if $D$ is not void, it is more probable to find an e-p pair than anything else. The dominant term in $\mbox{Tr}\exp(-\beta H_{D,N_eN_p})$ as $\beta\to\infty$ will be $\exp(-\beta E_{D,N_{e}N_{p}})$ where $E_{D,N_{e}N_{p}}$ is the ground state energy of $H_{D, N_{e}N_{p}}$. Hence to obtain the inequalities (\ref{7.16}), one must be able to find values of $\mu_{e}$ and $\mu_{p}$ such that \begin{equation} E_{D,N_{e}N_{p}}-(\mu_{e}N_{e}+\mu_{p}N_{p})>E_{D,11}-(\mu_{e}+\mu_{p})>0 \;\;\mbox{for all}\;(N_e,N_p)\neq (0,0),\:(1,1) \label{7.18} \end{equation}Since $\rho_{a}$ tends to zero exponentially fast as $\beta\to\infty$ we can also let the region $D$ grow with $\beta$ (but maintaining the inequalities (\ref{7.15})) so that $E_{D,N_{e}N_{p}}$ will differ from the bottom of the spectrum of $H_{N_{e}N_{p}}$ in infinite space \begin{equation} E_{N_eN_p}=\mbox{inf spectrum}(H_{N_eN_p}) \label{7.19} \end{equation}by a vanishingly small error. Moreover, introducing $\mu,\nu$ as in (\ref{7.4}) \begin{equation}\mu_eN_e+\mu_pN_p=\mu(N_e+N_p)+\nu(N_e-N_p) \label{7.20} \end{equation}we remark that the part $\mu$ of the chemical potential controls the total number density and $\nu$ the total charge density. We can take advantage of the fact that the infinite volume limit of the pressure will not depend on $\nu$ to make a convenient choice. As it was done in (\ref{7.5}), we fix $\nu$ to ensure neutrality in the ideal gas when the lowest energy state corresponds to charged complexes. Then $\nu$ will be of order $0(\beta^{-1})$ as in the dissociated e-p system and can be dropped from (\ref{7.20}) as $\beta$ gets large. Therefore, in view of these remarks (that will be made mathematically more precise in section~\ref{sec-7.3}) the relevant inequalities to be satisfied to produce the situation (\ref{7.16}) as $\beta\to\infty$ are (with $E_{11}=E_{a}$) \begin{equation} E_{N_eN_p}-\mu(N_e+N_p)>E_{a}-2\mu>0\;\;\mbox{for all}\;(N_e,N_p)\neq (0,0),\:(1,1) \label{7.21} \end{equation}for some values of $\mu$. Let us examine some implications of (\ref{7.21}). If $(N_e,N_p)=(1,0)$ (single electron) $E_{10}=0$ implies in (\ref{7.21}) \begin{equation} E_{a}<\mu<0 \label{7.22} \end{equation} If $(N_e,N_p)=(2,2)$, (\ref{7.21}) gives \begin{equation} \mu<\frac{1}{2}(E_{22}-E_{a}) \label{7.23} \end{equation} where $E_{22}$ is the ground state energy of the hydrogen molecule. One can find $\mu$ satisfying (\ref{7.22}) and (\ref{7.23}) only if $E_{a}<\frac{1}{2}(E_{22}-E_{a})$ or equivalently if $|E_{22}-2E_{a}|<|E_{a}|$. The latter inequality means that the binding energy gained by the formation of an hydrogen molecule must be less than the binding energy of the atom itself, a well known fact. It allows the entropy to win over molecular binding at low density. However the numerical values of $(E_{22}-E_{a})/2 \sim - 9 \text{ev}$ and of $\mu \sim -8 \text{ev}$, $k_{\rm B}T \sim 0.026 \text{ev}$ under prevailing Earth conditions give an exceedingly small value of the ratio $\rho _{a}/\rho _{{\rm H}_{2}} \sim \exp (\beta (E_{22}-E_{a} - 2\mu ))$. Only molecular hydrogen appears on Earth, while the formation of the atomic gas discussed here can only occur in very dilute extraterrestrial medium. These considerations on molecular hydrogen do not suffice to guarantee the validity of the full set of inequalities (\ref{7.21}): this necessitates a simultaneous examination of all ground state energies $E_{N_{e}N_{p}}$, a highly nontrivial problem. We shall discuss it further at the end of the next section. A geometric illustration of the inequalities (\ref{7.21}) is given in Figure~\ref{fig13}. Notice that the binding energies of the ion $H^{-}$ and the ionized molecule $H_{2}^{+}$ are too weak: they do not meet the inequalities analogous to (\ref{7.21}), appropriate to these cases, for any value of $\mu$. Coming back to the e-p pressure, we may conjecture that if $\mu$ is increased just beyond $\frac{1}{2}(E_{22}-E_{a})$, the chances for binding also increase, and the pressure $P(\beta,\mu)$ will become close to that a free gas of hydrogen molecules as $\beta\rightarrow \infty$. More generally, if one fixes $\mu <0$, one expects that $P(\beta,\mu)$ approaches the pressure of free gases of some complexes that are selected at this particular value of the chemical potential. \subsection{The main theorem} \label{sec-7.2} With these preliminaries in mind, we are ready to formulate the main theorem about the Saha regime for electrons and an arbitrary number of nuclei with masses $m_{\alpha}$, charges $e_{\alpha}$, particle numbers $N_{\alpha}$ and chemical potentials $\mu_{\alpha},\: \alpha=1,\ldots,{\cal S}$ ($\alpha=1$ refers to electrons). As in section~\ref{sec- neutrality} we adopt a vector notation for these. Set \begin{eqnarray} \underline{\bf m} &=& (m_{1},\ldots,m_{{\cal S}}), \ \ \ \underline{\bf e} = (e_{1},\ldots,e_{{\cal S}})\nonumber\\ \underline{\bf N} &=& (N_{1},\ldots,N_{{\cal S}}), \ \ \ N=\sum_{\alpha}^{{\cal S}}N_{\alpha}, \ \ \ \underline{\mbox{\boldmath $\mu$\unboldmath}} = (\mu_{1},\ldots,\mu_{{\cal S}}) \label{7.24} \end{eqnarray} As we know (see (\ref{2.1.1ee})), the infinite volume pressure depends only on the component $\mbox{\boldmath $\mu$\unboldmath}$ of the chemical potential orthogonal to the charge vector. As a consequence, the kind of complexes that will occur for a certain value of the chemical potentials $\underline{\mbox{\boldmath $\mu$\unboldmath}}$ depends only on the value of $\mbox{\boldmath $\mu$\unboldmath}$. We will use this freedom to make a convenient choice of $\nu$ along the proof. The study of the molecular limit of $P(\beta, \mbox{\boldmath $\mu$\unboldmath})$ (i.e. $\beta\rightarrow \infty,\;\mbox{\boldmath $\mu$\unboldmath}$ fixed) will be carried out under the basic assumption on $\mbox{\boldmath $\mu$\unboldmath}$: \begin{description} \item[(A)] there exists $\kappa>0$ ($\kappa$ independent of $\underline{\bf N}$) such that \end{description} \begin{equation} H_{\underline{\bf N}}-\mbox{\boldmath $\mu$\unboldmath}\cdot\underline{\bf N} \:\geq\: \kappa N\;\;\;\;\;\;\;\;\;\; \mbox{for all} \;\underline{\bf N} \label{7.27} \end{equation} The complexes (atoms, ions, molecules) that will be formed in this limit are defined as follows. Consider the lowest energy state of the $N$-body Hamiltonian $H_{\underline{\bf N}}$ \begin{equation} E_{\underline{\bf N}}=\mbox{inf spectrum}(H_{\underline{\bf N}})=\inf\{(\Phi,H_{\underline{\bf N}}\Phi):\:\Phi\in {\cal H}_{N},\:||\Phi||=1\} \label{7.28} \end{equation} and set \begin{equation} E(\mbox{\boldmath $\mu$\unboldmath})=\inf_{\underline{\bf N}\neq \underline{{\bf 0}}}(E_{\underline{\bf N}}-\mbox{\boldmath $\mu$\unboldmath}\cdot\underline{\bf N}) \label{7.29} \end{equation} A value $\underline{\bf N}_{j}$ for which this infimum is taken, i.e. \begin{equation} E_{\underline{\bf N}_{j}}-\mbox{\boldmath $\mu$\unboldmath}\cdot \underline{\bf N}_{j}=E(\mbox{\boldmath $\mu$\unboldmath}) \label{7.30} \end{equation} is called the composition of the complex (j) having ground state energy $E_{j}=E_{\underline{\bf N}_{j}}$. Notice that (\ref{7.27}) imposes the bound $N_{j,\alpha}\:\leq\:\kappa^{-1}E(\mbox{\boldmath $\mu$\unboldmath})$ for all $\alpha$, so complexes are made out of a finite number of constituents, hence their number g is also finite. The complex (j), made of $N_{j,\alpha},\;\alpha=1,\ldots,{\cal S}$, elementary constituents, has total particle number, mass and charge \begin{equation} N_{j}=\sum_{\alpha}^{\cal S}N_{j,\alpha},\,\;\;\;\;\,\,m_{j}=\underline{\bf m}\cdot\underline{\bf N}_{j}, \;\;\,\,\;\;\;e_{j}=\underline{\bf e}\cdot\underline{\bf N}_{j} \label{7.31} \end{equation} Strictly speaking, $H_{\underline{\bf N}}$ has no ground state because of translation invariance. However for $N \geq 2$ one can decompose $H_{\underline{\bf N}}=K_{\underline{\bf N}}^{{\rm cm}}+H_{\underline{\bf N}}^{{\rm rel}}$ into the sum of the kinetic energy of the center of mass and the relative Hamiltonian, expressed in the coordinates relative to the center of mass. Since the spectrum of $K_{\underline{\bf N}}^{{\rm cm}}$ starts from zero, one has also $E_{\underline{\bf N}}=\mbox{inf spectrum}(H_{\underline{\bf N}}^{{\rm rel}})$. Consider now the threshold $\sigma_{c}(H_{\underline{\bf N}}^{{\rm rel}})$ of the continuous spectrum of $H_{\underline{\bf N}}^{{\rm rel}}$. We know from the HVZ theorem \cite{ReSi79} that \begin{equation} \sigma_{c}(H_{\underline{\bf N}}^{{\rm rel}})=\inf _{a,b;N_{a}+N_{b}=N}(E_{\underline{\bf N}_{a}}+E_{\underline{\bf N}_{b}}),\;\;\;N\geq 2 \label{7.31a} \end{equation} where the infimum is taken over all partitions of the $N$ particles into two clusters $a$ and $b$ having $N_{a}$ and $N_{b}$ particles, and $N_{a}=1,\ldots,N-1$. From the very definition of $E(\mbox{\boldmath $\mu$\unboldmath})$, one concludes from (\ref{7.31a}) that\linebreak[3] $\sigma_{c}(H_{\underline{\bf N}}^{{\rm rel}}-\mbox{\boldmath $\mu$\unboldmath}\cdot \underline{\bf N})\geq 2E(\mbox{\boldmath $\mu$\unboldmath}), N\geq 2$. This means that the part of the spectrum of $H_{\underline{\bf N}}^{{\rm rel}}-\mbox{\boldmath $\mu$\unboldmath}\cdot \underline{\bf N}$ located in $[E(\mbox{\boldmath $\mu$\unboldmath}), \;2E(\mbox{\boldmath $\mu$\unboldmath}))$ consists of isolated eigenvalues ($E(\mbox{\boldmath $\mu$\unboldmath})$ is strictly positive). In particular $E_{\underline{\bf N}_{j}}$ has to be an eigenvalue of $H_{\underline{\bf N}_{j}}^{\rm rel}$ and the corresponding eigenstate is the the wave function of the bound complex $(j)$. When $N=1$, complexes reduce to the individual electrons or nuclei themselves with pure kinetic energy and we take then $E_{\underline{\bf N}}=0$. When there are charged complexes, we add to (A) the assumption \begin{description} \item[(B)] if a complex has a charge $e_{j}\neq 0$, there is another complex having a charge of sign opposite to $e_{j}$ (but not necessarily equal to $-e_{j}$) . \end {description} insuring that an overall neutral gas can be formed. We describe now the mixture of free gases made of these complexes. To each complex (j) one associates the density of a perfect gas of point particles \begin{equation} \rho_{j}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}})=\(\frac{m_{j}}{2\pi\beta\hbar^{2}}\)^{3/2} \exp[-\beta(E_{j}-\underline{\mbox{\boldmath $\mu$\unboldmath}}\cdot\underline{\bf N}_{j})] \label{7.32} \end{equation}and define the pressure of an ideal mixture of such gases by \begin{equation}\beta P_{{\rm ideal}}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}})=\sum_{j=1}^{g}\rho_{j}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}}) \label{7.33} \end{equation}The corresponding number density of electrons and nuclei is \begin{equation} \rho_{{\rm ideal}}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}})=\sum_{\alpha=1}^{{\cal S}}\frac{\partial}{\partial\mu_{\alpha}} P_{{\rm ideal}}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}})=\sum_{j=1}^{g}N_{j}\rho_{j}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}}) \label{7.34} \end{equation}If some complexes are charged, $ P_{{\rm ideal}}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}})$ for a general $\underline{\mbox{\boldmath $\mu$\unboldmath}}$ may belong to a non-neutral ensemble. Then, under assumption (B), the neutrality $\sum_{j}e_{j}\rho_{j}=0$ is implemented by setting \begin{equation}P_{{\rm ideal, neutral}}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}})=\inf_{\nu} P_{{\rm ideal}}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}}+\nu\underline{\bf e}) \label{7.35} \end{equation}Then $P_{{\rm ideal, neutral}}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}})=P_{{\rm ideal, neutral}}(\beta,\mbox{\boldmath $\mu$\unboldmath})$ verifies (\ref{2.1.1ee}), and one can formulate the theorem on the molecular limit as \begin{description} \item[Theorem] {\em Let $\mbox{\boldmath $\mu$\unboldmath}$ be a chemical potential satisfying the condition (A) and determining complexes (j), $j=1,\ldots,g$, as described above, fulfilling the condition (B). Then there exists $\varepsilon>0$ such that \begin{equation} P(\beta,\tilde{\mbox{\boldmath $\mu$\unboldmath}})=P_{{\rm ideal, neutral}}(\beta,\tilde{\mbox{\boldmath $\mu$\unboldmath}})(1+O(\exp(-\beta\varepsilon))) \label{7.36} \end{equation} as $\beta\rightarrow\infty$ and $\tilde{\mbox{\boldmath $\mu$\unboldmath}}\rightarrow\mbox{\boldmath $\mu$\unboldmath}$.} \end{description} The theorem deserves several comments that will be again illustrated in the e-p system. One should first appreciate that all the residual effects due to thermal excitations and interactions between complexes when the temperature and the densities are not strictly equal to zero are controlled by the $O(\exp(-\beta\varepsilon))$ correction. The issue of the theorem is precisely about obtaining this control. Concerning the assumption (A), it is a difficult task to find what are the complexes associated with a chemical potential $\mbox{\boldmath $\mu$\unboldmath}$ and to verify (\ref{7.27}). Conversely, if we ask for the occurrence of a certain chemical species, it is still beyond the present possibilities to rigorously determine the range of $\mbox{\boldmath $\mu$\unboldmath}$ (if any) allowing for the formation of this molecule. Suppose that we ask for the formation of hydrogen atoms in the e-p system. This means that there should be an interval $\mu\in (\mu_{1},\mu_{2})$ such that (A) holds and $\inf_{(N_e,N_p)\neq (0,0)}(E_{N_eN_p}-\mu(N_e+N_p))$ is taken at $(N_e,N_p)=(1,1)$ with $E(\mu)=E_{a}-2\mu$. This is equivalent to the fact that the inequalities (\ref{7.21}) hold when $\mu\in (\mu_{1},\mu_{2})$ (if (\ref{7.21}) is true for such a $\mu$, one can pick $\kappa>0$ small enough so that the same inequalities are true for $\mu+\kappa$; then $\kappa$ provides the constant in the lower bound (\ref{7.27})). A little bit of thinking reveals that the validity of the inequalities (\ref{7.21}) for a range of $\mu$ ($\mu$ slightly above $E_{a}$) is equivalent in turn with the following statement about the stability of matter: there exists a constant $B$ with $0<B<|E_{a}|$ such that \begin{equation}H_{N_eN_p}\geq -B(N_e+N_p-1) \;\;\,\mbox{for all }(N_e,N_p)\neq (0,0),(1,1) \label{7.37} \end{equation} The point in (\ref{7.37}) is that the constant $B$ can be chosen strictly less than $|E_{a}|$ for all cases except of course for the hydrogen atom itself. Therefore a complete proof of the existence of the atomic phase relies on exhibiting a sufficiently small stability constant. Although (\ref{7.37}) must hold on experimental and numerical grounds, a rigorous proof has not yet been provided (see \cite{Fef86} for a more detailed discussion). Taking from now on (\ref{7.37}) as a working hypothesis, we recover easily the previous cases (i)-(iii) when applying the theorem to the e-p system. If $\mu$ is slightly above $E_{a}$, there is a single complex, the hydrogen atom (a) with $E(\mu)=E_{a}-2\mu$ (case (ii)). If $\mu=E_{a}$, we have three complexes (e,p,a) and $E(\mu)=|E_{a}|$ (case (iii)). If $\mu<E_{a}$, the complexes are (e,p) and $E(\mu)=|\mu|$ (case (i)). When there are charged complexes, the minimizer $\nu$ of $P_{ideal}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}}+\nu\underline{\bf e})$ has the value (\ref{7.5}) so that the result of the theorem is precisely the equation of state (\ref{7.7}) up to an exponentially small correction. For arbitrary nuclei, the theorem describes generically similar situations involving more complicated atoms, ions and molecules. If $\mbox{\boldmath $\mu$\unboldmath}$ selects a single complex, we have a free gas of that chemical species. If $\mbox{\boldmath $\mu$\unboldmath}$ selects several complexes, we can have coexisting free gases as $T\rightarrow 0$, generalizing the situation (iii) of the section \ref{sec-7.1}. As a corollary of the theorem, we can specify the relative proportion of these gases by taking $\tilde{\mbox{\boldmath $\mu$\unboldmath}}$ a function of the temperature that approaches $\mbox{\boldmath $\mu$\unboldmath}$ linearly as $T\rightarrow 0$ (as in (\ref{7.11})) \begin{equation} \tilde{\mbox{\boldmath $\mu$\unboldmath}}(\beta)=\mbox{\boldmath $\mu$\unboldmath}+{\mbox{\boldmath $\lambda$\unboldmath}}\beta^{-1}+o(\beta^{-1}) \label{7.38} \end{equation} The densities of the gases will have weights proportional to $\exp({\mbox{\boldmath $\lambda$\unboldmath}}\cdot\underline{\bf N}_{j})$, leading then to general Saha formulae. Although the determination of the kind of complexes occurring for a given value of $\mbox{\boldmath $\mu$\unboldmath}$ remains unsolved, it is pleasing to see how the thermodynamics of ideal substances emerges in principle from the basic statistical mechanics of electrons and nuclei in the molecular limit. \subsection{Elements of proof} \label{sec-7.3} The theorem was first proven \cite{Fef85} for the e-p system under the assumption (\ref{7.37}) (with a weaker control ${\rm O}(\beta^{-1})$ of the error term). All the main ideas and subsequent developments summarized here have their roots in this work. The general strategy of the proof does not rely on low density expansions, but on an analysis of typical equilibrium configurations of dilute gases of atoms and molecules. It proceeds along the following qualitative ideas. Let $\mbox{\boldmath $\mu$\unboldmath}$ be fixed and satisfy the condition (\ref{7.27}), and take $\beta$ very large. Then, in such configurations, if we look into a region of linear extension $R$ in the range (\ref{7.15}), according to the same discussion as that following (\ref{7.15}), this region is empty with high probability, but if it contains something, this will essentially be one of the complexes determined by $\mbox{\boldmath $\mu$\unboldmath}$. The idea is now to decompose the total domain $\Lambda=\bigcup_{r}D_{r}$ into subdomains $D_{r}$ of diameter $R$ and to let $R$ grow with $\beta$ in a suitable way. One has to solve the two following problems. \begin{description} \item[Problem 1] Determine the size of $D_{r}$ such that the partition function $\Xi_{D_{r}}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}})$ of a single subdomain is that of a system having at most complexes of the type $(j),\; j=1, \ldots, g$, in $D_{r}$. \item[Problem 2] Show that the residual interaction between the different regions $D_{r}$ is negligible as $\beta\rightarrow\infty$ so that the total partition function $\Xi_{\Lambda}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}})\simeq \prod_{r}\Xi_{D_{r}}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}})$ is close to that of independent subdomains $D_{r}$. \end{description} We make the preliminary decision (valid throughout the whole proof) to choose $\nu=\nu_{0}(\beta)$ as the minimizer of (\ref{7.35}) insuring the neutrality of the ideal gas. This is equivalent to the pseudo-neutrality condition met earlier on several occasions. If some complex is not neutral and the condition (B) holds, $\nu_{0}(\beta)$ exists, is unique and $\nu_{0}(\beta)={\rm O} (\beta^{-1})$ as can be checked from the explicit formulae (\ref{7.33}) and (\ref{7.34}) (see (\ref{7.5}) in the e-p system). If all complexes are neutral one takes $\nu=0$. Thus in the sequel we keep in mind that \begin{equation} \underline{\mbox{\boldmath $\mu$\unboldmath}}=\mbox{\boldmath $\mu$\unboldmath} +{\rm O}(\beta^{-1})\underline{\bf e} \label{7.38a} \end{equation} \subsubsection{Towards a solution of problem 1} We envisage now problem 1, following mainly \cite{CLY89}. We write the grand-canonical partition function for a finite region $D\subset{\Bbb R}^3$ of volume $|D|$ as a sum on all particle numbers \begin{equation}\Xi_{D}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}})=1+\sum_{j=1}^{g}\Xi_{D}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}},\underline{\bf N}_{j}) +\sum_{\underline{\bf N} \neq \underline{\bf N}_{j}}\Xi_{D}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}},\underline{\bf N}) \label{7.39} \end{equation}where we have singled out the contribution of the $g$ complexes that correspond to the chosen value of $\mbox{\boldmath $\mu$\unboldmath}$ and \begin{eqnarray} \Xi_{D}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}},\underline{\bf N}) &=&\mbox{Tr}_{D}\exp\(-\beta\( H_{D,\underline{\bf N}}-\underline{\mbox{\boldmath $\mu$\unboldmath}}\cdot\underline{\bf N}\)\)\nonumber\\ &=&\mbox{Tr}_{D}\exp\(-\beta K_{D,\underline{\bf N}}^{{\rm cm}}\) \mbox{Tr}_{D}\exp\(-\beta\( H_{D,\underline{\bf N}}^{{\rm rel}}-\underline{\mbox{\boldmath $\mu$\unboldmath}}\cdot\underline{\bf N}\)\)\nonumber\\ \label{7.40} \end{eqnarray} In (\ref{7.40}), $\mbox{Tr}_{D}$ means that the trace is taken on states having the Dirichlet conditions appropriate to the finite volume Hamiltonians $H_{D}$ as well as the appropriate particle statistics. For $R$ large enough, one expects that the center of mass contribution is close to that of a free particle in space, and that the finite volume ground state energy $E_{D,j}$ of the complex $j$ is close to its value $E_{j}$ in infinite space. Indeed one can establish \begin{equation} \mbox{Tr}_{D}\exp\(-\beta K_{D,\underline{\bf N}_{j}}^{{\rm cm}}\)=\(\frac{m_{j}}{2\pi\beta\hbar^{2}}\)^{3/2} |D|\(1+{\rm O}\(\frac{1}{R}\)\) \label{7.41} \end{equation} and \begin{equation} E_{D,j}=E_{j}+{\rm O}\(\frac{1}{R^{2}}\) \label{7.42} \end{equation} Now let \begin{equation} |D|=|D_{\beta}|\simeq e^{c\beta},\;\;\;\,c>0 \label{7.42a} \end{equation} grow exponentially fast as the temperature tends to zero, and consider the contribution of the ground state of $ H_{D,\underline{\bf N}_{j}}^{{\rm rel}}$ to the partition function (\ref{7.40}) for $\underline{\bf N}=\underline{\bf N}_{j}$. Taking (\ref{7.41}), (\ref{7.42}) and (\ref{7.42a}) into account, this contribution is \begin{eqnarray} & &\(\frac{m_{j}}{2\pi\beta\hbar^{2}}\)^{3/2} e^{-\beta(E_{j}-\underline{\mbox{\boldmath $\mu$\unboldmath}}\cdot\underline{\bf N}_{j})}|D_{\beta}|\(1+{\rm O}\(e^{-\frac{c}{3}\beta}\)\) \nonumber\\&=&\rho_{j}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}})|D_{\beta}|\(1+{\rm O}\(e^{-\frac{c}{3}\beta}\)\) \label{7.43} \end{eqnarray} By (\ref{7.30}) and (\ref{7.38a}), the density of the ideal gas $\rho_{j}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}})$ (\ref{7.32}) decreases as $\beta^{-3/2}e^{-\beta E(\mbox{\boldmath $\mu$\unboldmath})},\;E(\mbox{\boldmath $\mu$\unboldmath})>0$; thus one can choose $c>0$ small enough so that (\ref{7.43}) is still exponentially small for $\beta$ large. We like to show that, under the condition (\ref{7.27}), all the other terms in the sum (\ref{7.39}) are negligible compared to (\ref{7.43}), provided that the rate of growth of $|D_{\beta}|$ is suitably chosen. As $|D|\to\infty$, part of the spectrum of $H_{D,\underline{\bf N}}^{{\rm rel}}$ becomes continuous, corresponding to the scattering states of all the sub-complexes belonging to $H_{D,\underline{\bf N}}^{{\rm rel}}$. The basic idea is to ensure the convergence of the traces by the pure kinetic energy: one will dispense with the full Hamiltonian with the help of (\ref{7.27}) and the ensuing gap property. Then the strict positivity of the constant $\kappa$ in (\ref{7.27}) and of the gap will be used to control the volume dependence of these traces. For this one borrows a small fraction of the kinetic energy $K_{\underline{\bf N}}$ writing \begin{equation} H_{\underline{\bf N}}=\eta K_{\underline{\bf N}}+(1-\eta)K_{\underline{\bf N}}+V_{\underline{\bf N}},\;\;\;\,\eta>0 \label{7.45} \end{equation} where $ K_{\underline{\bf N}}=-\sum_{i=1}^{N}\frac{\hbar^{2}\Delta_{i}}{2m_{i}}$ and $V_{\underline{\bf N}}$ is the total Coulomb interaction. Clearly the Hamiltonians $(1-\eta)K_{\underline{\bf N}}+V_{\underline{\bf N}}$ and $(1-\eta)^{-1}(K_{\underline{\bf N}}+V_{\underline{\bf N}})=(1-\eta)^{-1}H_{\underline{\bf N}}$ are related by the scaling transformation $x_{i}\longrightarrow (1-\eta)x_{i}$. Hence they are unitarily equivalent. This implies, using (\ref{7.27}) and(\ref{7.38a}) \begin{eqnarray} (1-\eta)K_{\underline{\bf N}} +V_{\underline{\bf N}}-\underline{\mbox{\boldmath $\mu$\unboldmath}}\cdot\underline{\bf N}&\geq& (1-\eta)^{-1}\mbox{inf spectrum}(H_{\underline{\bf N}}-\mbox{\boldmath $\mu$\unboldmath}\cdot\underline{\bf N})\phantom{-------}\nonumber\\ &+&(1-\eta)^{-1}\eta \mbox{\boldmath $\mu$\unboldmath}\cdot\underline{\bf N}+ {\rm O}(\beta^{-1})N\nonumber\\ &\geq&(1-\eta)^{-1}(\kappa-\eta\max_{i}|\mu_{i}|)N+ {\rm O}(\beta^{-1})N\equiv c_{1}N\nonumber\\ \label{7.46} \end{eqnarray} Thus \begin{equation} H_{\underline{\bf N}}-\underline{\mbox{\boldmath $\mu$\unboldmath}}\cdot\underline{\bf N}\geq \eta K_{\underline{\bf N}}+c_{1}N \label{7.46a} \end{equation} where $c_{1}>0$ provided that $\eta$ is small enough and $\beta$ is large enough. The same inequality holds when $H_{\underline{\bf N}}$ is replaced by $H_{D,\underline{\bf N}}$ since Dirichlet boundary conditions increase the kinetic energy; hence one concludes that \begin{equation} \Xi_{D}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}},\underline{\bf N})\leq \exp(-\beta c_{1}N) \mbox{Tr}_{D}e^{-\beta\eta K_{\underline{\bf N}}}\leq |D|^{N}\prod_{i=1}^{N} \(\frac{m_{i}}{2\pi\eta\beta\hbar^{2}}\)^{3/2}e^{-\beta c_{1}N} \label{7.47} \end{equation} If now $|D|=|D_{\beta}|$ grows at the rate $e^{c\beta}$ with $0<c<c_{1}$, $\Xi_{D_{\beta}}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}},\underline{\bf N})$ is bounded by $e^{-c_{2}\beta N}$ for some $c_{2}>0$. Thus by choosing $N_{0}$ large, one can make the exponential decay of the whole sum of terms with $N\geq N_{0}$ in (\ref{7.39}) fast enough to have \begin{equation} \sum_{\underline{\bf N},N\geq N_{0}}\Xi_{D_{\beta}}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}},\underline{\bf N})=\rho_{j}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}})|D_{\beta}|{\rm O}\(e^{-\beta\varepsilon}\),\;\,\;\varepsilon>0 \label{7.48} \end{equation} For the remaining terms, i.e. the terms with $\underline{\bf N}\neq\underline{\bf N}_{j},\;N<N_{0}$ and the contribution of the excited states to $\Xi_{D}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}},\underline{\bf N}_{j})$, one uses essentially the fact that $E(\mbox{\boldmath $\mu$\unboldmath})$ is separated from the rest of the spectrum by a gap, according to the discussion following (\ref{7.31a}). On the subspace orthogonal to the ground states of $H_{\underline{\bf N}_{j}}^{{\rm rel}}-\mbox{\boldmath $\mu$\unboldmath}\cdot\underline{\bf N}_{j}$ one has, arguing as in (\ref{7.46}) \begin{eqnarray} (1-\eta)K_{\underline{\bf N}} +V_{\underline{\bf N}}-\underline{\mbox{\boldmath $\mu$\unboldmath}}\cdot\underline{\bf N}&\geq& (1-\eta)^{-1}(E(\mbox{\boldmath $\mu$\unboldmath}) + 2\delta)\phantom{-------}\nonumber\\ &+&(1-\eta)^{-1}\eta \mbox{\boldmath $\mu$\unboldmath}\cdot\underline{\bf N}+ {\rm O}(\beta^{-1})N\nonumber\\ &\geq&(1-\eta)^{-1}(E(\mbox{\boldmath $\mu$\unboldmath})+2\delta-\eta\max_{i}|\mu_{i}|N)+{\rm O}(\beta^{-1})N\nonumber\\ \label{7.48a} \end{eqnarray} for some $\delta>0$. Since now $N<N_{0}$, one can choose $\eta$ sufficiently small and $\beta$ sufficiently large to have on this subspace \begin{equation} H_{\underline{\bf N}}-\underline{\mbox{\boldmath $\mu$\unboldmath}}\cdot\underline{\bf N}\geq \eta K_{\underline{\bf N}}+E(\mbox{\boldmath $\mu$\unboldmath})+\delta,\;\,\,\delta>0,\;\;N<N_{0} \label{7.48b} \end{equation} Hence for $\underline{\bf N}\neq\underline{\bf N}_{j},\;N<N_{0}$, one finds as in (\ref{7.47}) \begin{equation} \Xi_{D}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}},\underline{\bf N})\leq e^{-\beta (E(\mbox{\boldmath $\mu$\unboldmath})+\delta)} \mbox{Tr}_{D}e^{-\beta\eta K_{\underline{\bf N}}}\leq |D|^{N}\prod_{i=1}^{N} \(\frac{m_{i}}{2\pi\eta\beta\hbar^{2}}\)^{3/2}e^{-\beta E(\mbox{\boldmath $\mu$\unboldmath})}e^{-\beta\delta } \label{7.49} \end{equation} Since $|D_{\beta}|^{N}$ grows at most as $e^{c\beta N_{0}}$ we can again have \begin{equation} \Xi_{D_{\beta}}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}},\underline{\bf N})= \rho_{j}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}})|D_{\beta}|{\rm O}(e^{-\beta\varepsilon}),\;\;\underline{\bf N}\neq\underline{\bf N}_{j},\;N<N_{0}, \;\;\varepsilon>0 \label{7.50} \end{equation} by taking $c$ small. Notice that a volume factor $|D_{\beta}|$ corresponding to the free motion of the overall center of mass has been explicitly maintained in the estimates (\ref{7.48}) and (\ref{7.50}) (at the expense of a possibly smaller constant $\varepsilon$). Finally one deals with the excited states of $H_{\underline{\bf N}_{j}}^{{\rm rel}}-\mbox{\boldmath $\mu$\unboldmath}\cdot\underline{\bf N}_{j}$ in the same way. Combining this with (\ref{7.43}), (\ref{7.48}) and (\ref{7.50}) in the grand canonical sum (\ref{7.39}), one obtains \begin{equation} \Xi_{D_{\beta}}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}})=1+\sum_{j=1}^{g}\rho_{j}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}})|D_{\beta}| \(1+{\rm O}\(e^{-\varepsilon\beta}\)\) \label{7.50a} \end{equation} Since the second term in the right hand side is exponentially small as $\beta\rightarrow \infty$ (see the comment after (\ref{7.43})) one has also \begin{equation} \ln\Xi_{D_{\beta}}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}})=\sum_{j=1}^{g}\rho_{j}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}})|D_{\beta}| \(1+{\rm O}\(e^{-\varepsilon\beta}\)\) \label{7.50b} \end{equation} Taking into account the definition (\ref{7.33}) together with our choice of $\nu$, this says that the pressure $P_{D_{\beta}}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}})$ in $D_{\beta}$ is given by \begin{equation} \beta P_{D_{\beta}}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}})=\frac{1}{|D_{\beta}|}\ln \Xi_{D_{\beta}}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}}) =\beta P_{{\rm ideal}}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}} )\(1+{\rm O}\(e^{-\varepsilon\beta}\)\) \label{7.51} \end{equation} This solves problem 1: in a volume $|D_{\beta}|\simeq e^{c\beta}$ with $c>0$ sufficiently small (depending only on $\mbox{\boldmath $\mu$\unboldmath}$), the pressure is exclusively due to the non-interacting complexes $(j),\; j=1,\ldots,g$. However this does not prove the theorem yet. In (\ref{7.51}), $D_{\beta}$ is constrained by the growth condition (\ref{7.42a}) and not an independent thermodynamic variable as it should be in the grand canonical ensemble. One needs to solve problem 2. \subsubsection{Towards a solution of problem 2} Problem 2 is technically considerably more sophisticated, so let us give first some very heuristic ideas. Decompose ${\Bbb R}^3=\bigcup_{r}D_{r}$ into a disjoint union of domains $D_{r}$ with characteristic functions $\chi_{D_{r}}$, $\chi_{D_{r}}(x) =1$ if $x$ is in $D_{r}$ and $\chi_{D_{r}}(x) =0$ otherwise. One wants to show that the interactions between the different domains $D_{r}$ become negligible as $\beta\rightarrow\infty$. To this end, define an uncorrelated Hamiltonian by \begin{equation}H^{{\rm uncor}}_{\Lambda}=H_{\Lambda}-U^{{ \rm cor}} \label{7.52} \end{equation} where $U^{{ \rm cor}}$ is the potential energy between the different domains \begin{equation} U^{{ \rm cor}}=\sum_{i<j}e_{\alpha_{i}}e_{\alpha_{j}}V^{{ \rm cor}}(x_{i},x_{j}) \label{7.52a} \end{equation} \begin{equation}V^{{ \rm cor}}(x_{1},x_{2})=\frac{1-\sum_{r}\chi_{D_{r}(x_{1})} \chi_{D_{r}}(x_{2})}{|x_{1}-x_{2}|} \label{7.53} \end{equation}$U^{{ \rm cor}}$ (as well as $H^{{\rm un\rm cor}}_{\Lambda}$) will be an effective temperature dependent energy via the condition $ |D_{r}|\simeq e^{c\beta}$ determined in problem 1. Note that if $|x_{1}-x_{2}|<R$ and both $x_{1}$ and $x_{2}$ belong to the same subdomain, then $V^{{\rm \rm cor}}(x_{1},x_{2})=0$; also $V^{{\rm \rm cor}}(x_{1},x_{2})$ is Coulombic at large distances. This resembles the properties of a regularized Coulomb interaction. Suppose that we are able to define the uncorrelated Hamiltonian in a less crude way than in (\ref{7.52}) and (\ref{7.53}) with the effect of replacing (\ref{7.53}) by a smooth positive definite regularized potential, finite everywhere for $|x_{1}-x_{2}|<R$ and asymptotically Coulombic . A possible candidate is \begin{equation}V^{{ \rm cor}}_{{\rm reg}}(x_{1},x_{2})=\int dy_{1}\int dy_{2}\frac{\phi_{R}(y_{1})\phi_{R}(y_{2})} {|x_{1}-y_{1}-x_{2}+y_{2}|} \label{7.54} \end{equation}where $\phi_{R}(y)=R^{-3}\phi(\frac{y}{R})$ with $\phi(y)$ a smooth function, ($\phi(y)=0,\;|y|\geq 1,\;\int dy\phi(y)=1$) representing a charge density supported in a region of extension $R$ with total charge equal to $1$. Other candidates closer to the form (\ref{7.53}) are \begin{equation}V^{{\rm \rm cor}}_{{\rm reg}}(x_{1},x_{2})=\frac{1-h\(\frac{x_{1}-x_{2}}{R}\)}{|x_{1}-x_{2}|} \label{7.55} \end{equation}with $h(x)$ a smooth short range function, $h(0)=1$, and $\frac{1-h(x)}{|x|}$ positive definite. Then we can use the basic positivity argument already introduced in section~\ref{sec- classical} \begin{eqnarray} U^{{\rm \rm cor}}_{{\rm reg}}&=&\frac{1}{2}\sum_{ij}^{N}e_{\alpha_{i}}e_{\alpha_{j}}V^{{\rm \rm cor}}_{{\rm reg}}(x_{i},x_{j}) -\frac{1}{2}V^{{\rm \rm cor}}_{{\rm reg}}(0)\sum_{i}^{N}e_{\alpha_{i}}^{2}\nonumber\\ &\geq& -\frac{1}{2}V^{{\rm \rm cor}}_{{\rm reg}}(0)\sum_{i=1}^{N}e_{\alpha_{i}}^{2} \label{7.56} \end{eqnarray} because of the positive definiteness of $V^{{\rm \rm cor}}_{{\rm reg}}(x_{1},x_{2})$. Note that by scaling, in both forms (\ref{7.54}) and (\ref{7.55}), one has $V^{{\rm \rm cor}}_{{\rm reg}}(0) = {\rm O}\(\frac{1}{R}\)$. Therefore if one can find such a regularized correlation energy, one will have (being still very sketchy) that $ H^{{\rm un\rm cor}}_{\Lambda}$ provides a lower bound to the full Hamiltonian \begin{eqnarray} H_{\Lambda}-\underline{\mbox{\boldmath $\mu$\unboldmath}}\cdot \underline{\bf N} &=& H^{{\rm uncor}}_{\Lambda}+U^{{\rm cor}}_{{\rm reg}}-\underline{\mbox{\boldmath $\mu$\unboldmath}}\cdot \underline{\bf N} \geq H^{{\rm uncor}}_{\Lambda}-(1+{\rm O}(R^{-1}))\underline{\mbox{\boldmath $\mu$\unboldmath}}\cdot\underline{\bf N} \label{7.57} \end{eqnarray} Thus, up to a small correction to the chemical potential, one finds that the partition function is dominated by that of the uncorrelated domains \begin{equation} \Xi_{\Lambda}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}})\leq \prod_{D_{r}\cap \Lambda \neq \emptyset } \Xi_{D_{r} }(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}}(1+{\rm O}(R^{-1}))) \label{7.58} \end{equation} We set now $|D_{r}|=|D_{r,\beta}|\simeq e^{c\beta}$, $\beta$ sufficiently large, and apply to each of the $D_{r}$ the result (\ref{7.50b}) of problem 1. Taking the logarithm of (\ref{7.58}) and absorbing the error in the chemical potential in the ${\rm O}(e^{-\varepsilon \beta)})$ correction, one obtains in the thermodynamic limit \begin{eqnarray} \beta P(\beta, \underline{\mbox{\boldmath $\mu$\unboldmath}})&\leq& \lim_{|\Lambda|\rightarrow\infty}\left[\frac{1}{|\Lambda|}\sum_{D_{r,\beta}\cap \Lambda\neq\emptyset}|D_{r,\beta}|\right] \beta P_{{\rm ideal}}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}} )\(1+{\rm O}\(e^{-\varepsilon\beta}\)\)\nonumber\\ &=&\beta P_{{\rm ideal}}(\beta,\underline{\mbox{\boldmath $\mu$\unboldmath}} )\(1+{\rm O}\(e^{-\varepsilon\beta}\)\) \label{7.59} \end{eqnarray} The technically most elaborate parts of the works \cite{Fef85,CLY89,GrSc95a} consists in constructing an uncorrelated Hamiltonian having the requested positivity and smoothness properties, and comparing it with the true Hamiltonian $H_{\Lambda}$. In its main lines the construction of \cite{CLY89,GrSc95a} goes as follows. The decomposition ${\Bbb R}^3=\bigcup_{r}D_{r}$ is realized with regions $D_{r}$ that are translates and dilations (with the scale parameter $R$) of a fundamental domain $D_{0}$. Let ${\cal F}_{D_{r}}$ be the Fock space for a Coulomb system localized in $D_{r}$, having Hamiltonian $H_{D_{r}}$ with Dirichlet boundary conditions on $\partial D_{r}$. The most obvious way to define states of independent Coulomb systems, each of them confined in a region $D_{r}$, is to introduce the product space ${\cal F}_{\Lambda}^{\rm uncor}=\prod_{D_{r}\cap \Lambda \neq \emptyset}^{\bigotimes} {\cal F}_{D_{r}}$ and the Hamiltonian $\sum_{D_{r}\cap \Lambda \neq \emptyset}H_{D_{r}}$ on it. One wants to compare this obviously strictly uncorrelated Hamiltonian with the true $H_{\Lambda}$ acting on ${\cal F}_{\Lambda}$ when the scale parameter $R\to \infty$. For this, one introduces a partition of the unity on ${\Bbb R}^3$ of the form $\sum_{r}\chi^{2}_{r,R}(x)=1$ for all $x\in {\Bbb R}^3$ where $\chi_{r,R}$ are suitably chosen smooth functions essentially supported in the region $D_{r}$ of linear size $R$. With the help of this partition, one defines an isometry $J : {\cal F}_{\Lambda}\to {\cal F}_{\Lambda}^{\rm uncor}$ which maps the states of the system onto the uncorrelated ones. Then one compares the image $J^{-1}(\sum_{D_{r}\cap \Lambda \neq \emptyset}H_{D_{r}})J$ of the strictly uncorrelated Hamiltonian with $H_{\Lambda}$ on ${\cal F}_{\Lambda}$. The calculation gives \begin{equation} H_{\Lambda}-J^{-1}(\sum_{D_{r}\cap \Lambda \neq \emptyset}H_{D_{r}})J=U^{\rm cor}_{{\rm reg}}+O\left(\frac{1}{R^{2}} \right) \label{7.60} \end{equation} where $U^{{\rm cor}}_{{\rm reg}}$ has the form (\ref{7.52a}) and (\ref{7.55}), and \begin{equation} h(x)=\int dy \chi^{2}(x+y)\chi^{2}(y) \label{7.60a} \end{equation} is a smooth function localized in the neighborhood of $D_{0}$: (\ref{7.60}) is a more precise version of (\ref{7.52}). The ${\rm O}\left(\frac{1}{R^{2}}\right)$ term is the price of the increase in kinetic energy paid for the additional Dirichlet conditions imposed at the boundary of each $D_{r}$ in the strictly uncorrelated Hamiltonian $\sum_{r}H_{D_{r}}$. This price can be made small as $R\to\infty$ because of the differentiability of the localization functions $\chi_{r,R}$. The main issue is now the lower bound (\ref{7.56}), i.e. to have $V^{{\rm \rm cor}}_{{\rm reg}}(x_{1},x_{2})$ positive definite. In \cite{CLY89}, the regions $D_{r}$ are cubes, and this positivity is obtained by replacing the Coulomb potential in $\sum_{D_{r}\cap \Lambda \neq \emptyset}H_{D_{r}}$ by a Debye potential $\frac{e^{-\varepsilon|x|}}{|x|}$ with small $\varepsilon$ ($\varepsilon\simeq e^{-c\beta}$). Then it turns out that $$ V^{{\rm \rm cor}}_{{\rm reg}}(x_{1},x_{2})= \frac{1}{|x_{1}-x_{2}|}\(1-\frac{h((x_{1}-x_{2})/R)}{h(0)} e^{-\varepsilon |x_{1}-x_{2}|}\) $$ is similar to $ V_{\infty ,\lambda }$ (\ref{1.1.3bc}). Its Fourier transform can be shown to be positive for $\varepsilon$ and $R$ in appropriate ranges by using properties of $h(x)$ (\ref{7.60a}), and the arguments (\ref{7.56})-(\ref{7.59}) can eventually be cast in a rigorous form. However, to complete the proof, one has to show that the assumption (A) in the main theorem as well as the results obtained in problem 1 are stable when the Coulomb potential is approximated by the Debye potential. In particular, if (A) holds with a constant $\kappa$, then one has also for the Debye Hamiltonian $H^{\varepsilon}_{\underline{\bf N}}$ \begin{equation} H^{\varepsilon}_{\underline{\bf N}}-\mbox{\boldmath $\mu$\unboldmath}\cdot\underline{\bf N}\geq\kappa(\varepsilon)N\;\;\,{\rm with}\;\,\kappa(\varepsilon)\to\kappa, \;\;\varepsilon\to 0 \label{7.61} \end{equation}This theorem is proven in \cite{CLY89} with the same localization techniques In \cite{GrSc95a}, the decomposition ${\Bbb R}^3=\bigcup_{r}D_{r}$ is realized with simplices. Then the stability (\ref{7.56}) follows remarkably from the geometrical properties of a simplex. Namely for the spherical average $\bar{h}(|x|)$ of $h(x)$ (\ref{7.60a}) (here $\chi=\chi_{D_{0}}$ is the characteristic function of a simplex), the function $\frac{\bar{h}(0)-\bar{h}(|x|)}{|x|}$ has a positive Fourier transform. The original proof \cite{Fef85} uses certain coverings of ${\Bbb R}^{3}$ (called Swiss cheese) made of small cubes and large balls of different radii, and a positive definite potential of the form (\ref{7.54}). The delicate point is to estimate the difference $V^{{\rm error}}= V^{{\rm cor}}- V^{{\rm cor}}_{{\rm reg}}$ between (\ref{7.53}) and (\ref{7.54}). $V^{{\rm error}}$ still contains a local singularity as well as non-positive definite Coulomb energy terms. By averaging on the different radii of the balls, Fefferman shows that these dangerous contributions in $V^{{\rm error}}$ are majorized by a Coulomb potential $\frac{{\rm O}(\beta^{-1})}{|x|}$ with a small amplitude, and that these dangerous terms become eventually negligible. Then the stability (\ref{7.56}) follows from the positive definiteness of $ V^{{\rm cor}}_{{\rm reg}}$ (\ref{7.54}). In all cases the proofs involve fixing a number of important details not mentioned here. Finally, (\ref{7.59}) must be completed by the converse inequality; the job is fortunately easier and relies on an application of the variational principle. We refer to the original papers for this final step. \newpage \section{Convergent expansions} \setcounter{page}{1} \label{chapter-convergent-expansions} In previous chapters we have alluded to the Mayer expansion and results on the convergence of Mayer expansions. This is an old subject reviewed in numerous places including \cite{Bry84} but since that review some beautiful new combinatoric formulas have been discovered. The main objective of this summary is to update \cite{Bry84} to bring these to the attention of our readers, together with some applications and ramifications. This is done in sections~\ref{sec-mayer1},~\ref{sec-mayer2}. Related ideas convert the considerations of section~\ref{sec-tunneling} into convergent expansions and thereby prove there is screening. We outline these arguments in sections~\ref{sec-truncated-expectations}, \ref{sec-polymer-rep2}, \ref{sec-polymer-rep1}. This updates an older review \cite{BrFe81}. We are accustomed to expansions in theoretical physics where the terms are labeled by connected graphs. It is usually the case that the number of connected graphs at order $N$ grows more rapidly than $N!$ so that such expansions appear to be divergent at least until cancellations between graphs have been taken into account. The older review \cite{Bry84} was centered on rigorous bounds that say that the sum over all connected graphs at order $N$ is actually smaller than the sum over just the least connected graphs, i.e. tree graphs. The new part of this review is theorem~\ref{thm-forest} which makes the relation between all graphs and tree graphs much clearer. \subsection{Tree graph formulas and the Mayer expansion} \label{sec-mayer1} We begin with a little review of the theory of activity expansions \cite{Rue69,HaMc76,MaMa77}. Suppose $N$ particles labeled $1, 2, \dots , N$ have two-body potentials $V_{ij}$. The total potential of a subset $X$ of these particles is \begin{equation}\label{C.Gibbs1} U_X = \sum_{ij \in X} V_{ij} \end{equation} where $ij = ji$ is an unordered pair of distinct particles and $ij \in X$ means $i,j \in X$. Consider the formulas \begin{eqnarray}\label{C.truncated2} e^{-U_X} &=& \sum _{G} \prod _{ij \in G} \left(e^{-V_{ij}}-1\right)\nonumber \\ \left(e^{-U_X}\right)_{c} &=& \sum _{G \ {\rm connected}} \prod _{ij \in G} \left(e^{-V_{ij}}-1\right) \end{eqnarray} In the first the sum is over all graphs $G$, connected or disconnected, on vertices in $X$. It is an immediate consequence of expanding the product over $ij$ of $\exp \left(-V_{ij} \right) - 1 + 1$, remembering that a graph on $X$ is, by definition, an arbitrary subset of $\{ij:ij\in X \}$. The second formula is the usual way to define the {\it connected part \/} $\exp \left(-U_X \right)_{c}$, but there is another equivalent definition, namely the connected part is the (unique) recursive solution of \begin{eqnarray}\label{C.truncated1} && e^{-U_X} = \sum_{M \geq 1} \frac{1}{M!} \sum _{\stackrel{X_{1},\dots ,X_{M}}{\cup X_{j} = X}} \prod_{j=1 }^{M} \left(e^{-U_{X_{j}}}\right)_{c} \mbox{ if } |X| > 1\nonumber \\ && \left(e^{-U_X}\right)_{c} = 1 \mbox{ if } |X| = 1 \end{eqnarray} where $X_{1},\dots ,X_{M}$ are disjoint. The equivalence of the two definitions is not a hard exercise and is the main step in obtaining the Mayer expansion, which is \begin{theorem}\label{thm-mayer1} Let $U_{N} = U_{\{1,\dots ,N \}}$ where $V_{ij} = V({\bf r} _{i},{\bf r} _{j})$ is the pair interaction between particles in a grand canonical ensemble $\Xi$. Then \begin{equation}\label{C.2} \ln \Xi = \sum _{N \geq 1} \frac{1}{N!} \int \prod _{j=1}^{N} d{\cal E}_{j}\, z({\cal E}_{j} ) \left(e^{-\beta U_{N}}\right)_{c} \end{equation} \end{theorem} The proof that this formal expansion results from (\ref{C.truncated1}) is in most textbooks in statistical mechanics. The classical proofs that this expansion is convergent for small activity were given in \cite{Gro62,Pen63,Rue63}. Our line of development begins with \cite{Pen66} who showed that when the interactions are repulsive the sum over {\it all} connected graphs at order $N$ is dominated by the sum over tree graphs at the same order. By Cayley's theorem the number of tree graphs on $N$ vertices is $N^{N-2}$ which is comparable to $N!$. The convergence of the Mayer expansion at small activity will be seen to be an immediate consequence. {\it Notation:\/} We consider the class ${\cal F}$ of graphs called \underline{f}orests: A graph $G$ is a forest if it has no closed loops or equivalently if each connected component of $G$ is a tree graph. A single isolated vertex which is not in any bond is considered to be an empty tree graph. The empty graph which has no lines, so that every vertex is isolated, is therefore a forest. A forest $G$ has the property that for any pair of vertices $ij$, either there is a unique path that joins $i$ to $j$ consisting of bonds in $G$ or there is no path at all. To every bond $ij \in \{1,2,\dots ,N \}$ is associated a non-negative parameter $s_{ij}$. The vector whose components are all these parameters is denoted by $\bs $ and $\openone $ is the vector where each $s_{ij} = 1$. Let $G$ be a forest. For each bond $ij$ we set \begin{equation} \label{C.1} \sigma_{ij}(G, \bs ) = \left\{ \begin{array}{l} \inf\{s_{b}:b \in \mbox{ path in G joining } i \mbox{ and } j \}\\ 0 \mbox{ if no path} \end{array} \right. \end{equation} Thus if $ij$ belongs to the forest $G$ then $\sigma_{ij} = s_{ij}$. Also, $\sigma_{ij}(G,\bs )$ depends only on the $s_{ij}$ parameters assigned to bonds in $G$. We let $\bbox{\sigma} (G,\bs )$ denote the vector whose components are $\sigma_{ij}(G,\bs )$. We set $\partial ^{G} = \prod _{b \in G} \partial /\partial s_{b}$. The symbol $\int ^{t} d^{G}s$ means that for each $b \in G$ $s_{b}$ is integrated over the interval $[0,t]$. We write $f \circ \bbox{\sigma} (G,\bs ) = f(\bbox{\sigma} (G,\bs ))$ so that the notation $(\partial ^{G} F) \circ \bbox{\sigma}(G,\bs )$ will mean: first do the derivatives with respect to $s_{b}$ with $b \in G$ and then evaluate at $\bbox{\sigma} (G,\bs)$. The following theorem is a type of fundamental theorem of calculus. It reached this simple form through the series of papers \cite{GJS74,BrFe78,BaFe84,BrKe87,AbRi95}. \begin{theorem}\label{thm-forest} Let $F(\bs )$ be any continuously differentiable function of parameters $s_{ij}$ where $ij=ji$ and $i \not = j \in \{1,2,\dots ,N\}$ and $N \geq 2$. Let $t \geq 0$. Then \begin{equation}\label{C.4} F({t \openone} ) = \sum _{G \in {\cal F}} \int^{t} d^{G}s \, (\partial ^{G} F) \circ \bbox{\sigma}(G,\bs ) \end{equation} where \[ \int^{t} d^{G}s \, (\partial ^{G} F) \circ \bbox{\sigma}(G,\bs ) = F (0\openone ) \] when $G$ is the empty forest. \end{theorem} Note that the theorem is valid when $t=0$ because all terms in the sum over $G$ vanish except when $G$ is the empty forest and in this case the right hand side is $F({0 \openone} )$, by definition. Also, if $N=2$, the theorem reduces to $F(t) = F(0) + \int^{t} \partial_{12} F(s_{12}) ds_{12}$, which is the fundamental theorem of calculus. A proof of this theorem is given in section~\ref{sec-proofs}. To study the Gibbs factor $\exp \left(- \beta \sum V_{ij} \right)$ by this theorem we introduce parameters $s_{ij}$ in such a way that the Gibbs factor is the value of \begin{equation}\label{C.3} F(\bs ) = e^{-\beta U_{N}(\bs )} = \exp \left( -\beta \sum _{ij} V_{ij}(s_{ij} )\right) \end{equation} when $\bs = \openone$. We choose $V_{ij}(s_{ij})$ to vanish at $s_{ij}=0$ so that $\bs = 0$ corresponds to all interactions being switched off. We will call such a choice of dependence on $s_{ij}$ an interpolation. Assume we have such an interpolation. Given a forest $G$ on vertices $\{1,2,\dots ,N \}$, we can decompose it into connected components. These connected components are subgraphs, trees, on subsets of vertices $ X_{1},\dots ,X_{M}$. Thus $G$ determines a partition of the set of vertices into ``clusters'' $X_{1},\dots ,X_{M}$. Vertices label particles and from the definition of $\bbox{\sigma} \equiv \bbox{\sigma}(G,\bs )$ there are no interactions in $ U_{N}(\bbox{\sigma})$ between particles in different trees so that $(\partial ^{G} F) \circ \bbox{\sigma}$ factors across the partition $X_{1},\dots ,X_{M}$. By comparing the definition of connected part (\ref{C.truncated1}) with the result of the theorem one obtains \begin{theorem}\label{thm-mayer2} For $N \geq 2$ \begin{equation}\label{C.2b} \left(e^{-\beta U_{N}}\right)_{c} = \sum _{T} \int^{1} d^{T}s \, \left( \partial ^{T} e^{-\beta U_{N}} \right) \circ \bbox{\sigma}(T,\bs ) \end{equation} where $T$ is summed over all connected trees $T$ on $N$ vertices. \end{theorem} \noindent {\em Tree Graph domination:\/} Suppose that the potential is repulsive, $V_{ij} \geq 0$. Then let $V_{ij}(s_{ij}) = s_{ij}V_{ij}$ or indeed any other interpolation with $V_{ij}'(s) \geq 0$. By \begin{equation}\label{C.tree-lower} U_{N}(\bbox{\sigma}(T,\bs )) \geq \sum_{ij \in T} V_{ij}(\sigma_{ij}(T,\bs )) = \sum_{ij \in T} V_{ij}(s_{ij}) \end{equation} and theorem~\ref{thm-mayer2} \begin{equation} \label{C.2c} |\left(e^{-\beta U_{N}}\right)_{c}| \leq \sum _{T} \int^{1} d^{T}s \, \prod _{ij \in T} \beta V_{ij}'(s_{ij})e^{-\beta V_{ij}(s_{ij})} = \sum _{T} \prod _{ij \in T} \left( 1 - e^{-\beta V_{ij}} \right) \end{equation} which says that the connected part --- the sum over all connected graphs --- is dominated by the sum just over tree graphs, which is the remarkable result \cite{Pen66}. However there are also tree graph domination results when the interaction is not purely repulsive. $\bbox{\sigma} (G,\bs )$ has the following unobvious property to be proved in section~\ref{sec-proofs}. \begin{theorem}\label{thm-C.stability.a} If \begin{equation}\label{C.stability1} \sum _{ij} V_{ij} \geq -B N \end{equation} then \begin{equation}\label{C.stability2} \sum _{ij} \sigma_{ij}(G,\bs )V_{ij} \geq -B N \end{equation} The theorem remains valid when self-energies $1/2 \sum _{ii} V_{ii}$ are added to both left hand sides. \end{theorem} \noindent {\it Tree Graph Domination} (\ref{C.2c}) now becomes \begin{equation} \label{C.2da} |\left(e^{-\beta U_{N}}\right)_{c}| \leq e^{\beta BN} \sum _{T} \prod _{ij \in T} \beta |V_{ij}| \end{equation} \noindent {\em Tree Graph limit:\/} Consider the limit \begin{equation}\label{C.9a2} z \rightarrow \infty, \ \ \ \beta \rightarrow 0 \mbox{ with } z\beta \mbox{ fixed} \end{equation} By theorems~\ref{thm-mayer1} and \ref{thm-mayer2} the leading terms are \begin{equation}\label{C.9a} \ln \Xi \sim \sum _{N} \frac{1}{N!} \sum _{T} \int \prod_{k=1}^N z({\cal E}_{k}) \, d{\cal E}_{k} \, \prod _{ij \in T} \left( -\beta V({\cal E} _{i} , {\cal E} _{j}) \right) \end{equation} where $T$ is summed over all connected trees $T$ on $N$ vertices. There are $N-1$ lines in a connected tree graph on $N$ vertices so every term in the sum is $O (z^{N}\beta ^{N-1}) = O (z)$. The estimates in section~\ref{sec-mayer2} show that the Mayer expansion converges uniformly [after dividing out an overall factor of $z \times \mbox{volume}$] as this tree graph limit is taken so the limit can be taken under the sum over $N$. This series can even be summed exactly. It is identical with the series in powers of $z$ for \begin{equation}\label{C.9b} \frac{1}{2}\int h ({\cal E} ) V^{-1} h ({\cal E} ) \, d{\cal E} + \int z({\cal E}) \, d{\cal E} \, e^{-\beta h({\cal E} )} \end{equation} evaluated at the $h$ that makes it stationary. $V^{-1}$ is the operator inverse of of the operator whose kernel is $V$. This is the well known folk theorem that Feynman tree graphs sum to the classical action. More general interpolations besides $s_{ij}V_{ij}$ are actually useful because sometimes one has interactions which are really the sum of several interactions at different scales and then it is useful to use more complicated interpolations that turn them off one after the other. Since there will be examples of this in the next section we also record a more general stability result \begin{theorem}\label{thm-C.stability.b} If the interpolation satisfies \begin{equation}\label{C.stability3} \sum _{ij} V_{ij}'(t) \geq -B(t)N \end{equation} then \begin{equation}\label{C.12} U_{N}(\bbox{\sigma} (G,\bs )) \geq -2\sum _{ij\in T}\int _{0}^{s_{ij}}B(s)ds \end{equation} \end{theorem} \noindent {\it Tree Graph Domination} (\ref{C.2c}) now becomes \begin{equation} \label{C.2db} |\left(e^{-\beta U_{N}}\right)_{c}| \leq \sum _{T} \int^{1} d^{T}s \, \prod _{ij \in T} \beta |V_{ij}'(s_{ij})| e^{2\beta \int _{0}^{s_{ij}}B(s)ds } \end{equation} \subsection{Convergence of Mayer expansions}\label{sec-mayer2} We demonstrate some uses of tree graph domination by giving estimates on the radius of convergence for the Mayer expansion for a variety of systems including ones discussed earlier in this review. We will use the following two principles: A tree graph on $N$ vertices whose lines represent factors $f({\bf r} - {\bf r} ')$ contributes exactly \begin{equation}\label{E.tree1} \left( \int f({\bf r} ) d{\bf r} \, \right)^{{N-1}} \times {\rm Volume} \end{equation} because tree graphs, being connected, have $N-1$ lines. The integrals over vertices can be evaluated in order starting with ones on lines that are outermost branches (Trim the tree!). Furthermore the number of tree graphs at order $N$ is $N^{N-2}$ (Cayley's theorem). \subsubsection{Repulsive potentials}\label{subsubsec-repulsive potentials} As an instructive exercise we recover the result in \cite{Pen66}. Consider for simplicity one species. By tree graph domination (\ref{C.2c}) the Mayer expansion is bounded by \begin{eqnarray} \label{E.tree1a} && \sum _{N} \frac{1}{N!} \int |\left(e^{-\beta U_{N}}\right)_{c} z^{N}| d^{N}{\bf r}\nonumber \\ && \leq \sum _{N} \frac{1}{N!}\sum _{T} \left(|z|\int (1 - e^{-\beta V({\bf r} )}) d{\bf r} \, \right)^{N-1} |z| |\Lambda |\nonumber \\ && \leq \sum _{N} \frac{N^{N-2}}{N!} \left(|z|\int (1 - e^{-\beta V({\bf r} )}) d{\bf r} \, \right)^{N-1} |z| |\Lambda | =\sum _{N} \frac{N^{N-2}}{N!} Q^{N-1} |z| |\Lambda | \end{eqnarray} which is convergent provided \begin{equation} \label{E.tree1b} Q := |z|\int d{\bf r} \, (1- e^{-\beta V({\bf r} ) }) < e^{-1} \end{equation} This estimate is also used in the next two examples \subsubsection{Debye spheres}\label{subsubsec-debye spheres} In section~\ref{sec-debyesphere} we encountered conditions (\ref{2.4.1.5}). The system consisted of charged particles interacting either by two body potentials \begin{equation}\label{E.9-1} V_{L,\lambda }({\bf r} ) = |{\bf r} |^{-1} \left( e^{-|{\bf r} |/L} - e^{-|{\bf r} |/\lambda } \right) \end{equation} or by the Yukawa potential $V_{L}({\bf r} )$ and hard cores of radius $\lambda $. First we consider two species interacting by $(\ref{E.9-1})$. Suppose we can find an interpolation as in theorem~\ref{thm-C.stability.b}, then we can repeat the argument given above for repulsive potentials, but now using the tree graph domination (\ref{C.2db}) instead and find convergence when \begin{equation}\label{C.8} Q: = 2|z| \int _{0}^{1} ds\, \int d{\bf r} \, |\beta V'({\bf r},s )| \exp \left( 2\beta\int_{0}^{s}B(s^{\prime}) ds^{\prime} \right) < e^{-1} \end{equation} The factor $2$ is there because there are two species. The interpolation is chosen to be $V({\bf r} ,t) = V_{L,\ell(t)}({\bf r})$ where $\ell(t)$ is a length scale which decreases from $L$ when $t=0$ to $\lambda $ when $t=1$. Then we show below that \begin{equation}\label{E.10} V'({\bf r},t) = - \frac{\partial }{\partial t} \frac{e^{- |{\bf r}|/\ell(t)}}{|{\bf r} |} \end{equation} obeys a stability estimate (\ref{C.stability3}) with \begin{equation} \label{E.9-2} B(t) = \frac{\partial }{\partial t} \frac{1}{2\ell(t)} \end{equation} Also \begin{equation} \int |V'({\bf r},s )|d{\bf r} = -\frac{\partial }{\partial s} \int \frac{e^{- |{\bf r}|/\ell(s)}}{|{\bf r} |}d{\bf r} = -\frac{\partial }{\partial s} 4\pi \ell(s)^{2} \end{equation} because $|V'({\bf r},s )| = V'({\bf r},s )$. Therefore (\ref{C.8}) reads \begin{eqnarray} Q &=& - 2 |z|\beta \int_{0}^{1} ds \, \frac{\partial }{\partial s} 4\pi \ell(s)^{2} \exp \left(\beta \int _{0}^{s} \frac{\partial }{\partial s^{\prime}} \frac{1}{\ell(s^{\prime})}ds^{\prime}\, \right) \nonumber \\ &=& 16\pi |z| \beta \int_{\lambda }^{L} d\ell \, \ell \exp \left( \frac{\beta }{\ell} - \frac{\beta }{L} \right) < e^{-1} \end{eqnarray} When $\lambda < \beta < L$ we break up the range of the $d\ell$ integration into $[\lambda ,\beta ]$ and $[\beta ,L]$ and find that this criterion is satisfied if \begin{equation} |z|\beta L ^{2} \ll 1, \ \ \ |z|\beta^{3} e^{\beta /\lambda } \ll 1 \end{equation} which is the condition (\ref{2.4.1.5}). The Ruelle estimate (\ref{2.4.ruelle}) is not as good because it does not distinguish the length scales and essentially associates the bad stability factor $\exp(\beta /\lambda )$ to all scales. In the Debye-H\"uckel limit $Q = O (L^{2}/l_{D}^{-2})$ is determined by the first criterion because $|z|\beta^{3} e^{\beta /\lambda } \rightarrow 0$. By this remark and (\ref{E.tree1a}) estimate (\ref{2.4.1.5c}) follows. Results related to (\ref{2.4.1.5}) were given in \cite{GoMa81,Imb83a,BrKe87}. To obtain the stability estimate (\ref{E.9-2}), note that by the argument given near (\ref{1.1.3bb}) the stability constant $B (\ell)$ for $V_{L,\ell }$ is half the self-energy, which can be calculated using the Fourier transform to be \begin{equation}\label{E.9} B = \left(\frac{1}{2\ell } - \frac{1}{2L}\right) \end{equation} Now take $L$ in (\ref{E.9}) equal to $\ell + d\ell$ . It is not a hard exercise to rework this calculation for two dimensions. One recovers a result of \cite{Ben85} that if $\beta < 2$ [ $4\pi$ using the units in \cite{Ben85}] that the Mayer expansion is convergent for $|z|$ small uniformly in the short distance cutoff $\lambda $. In other words for $\beta $ smaller than the first threshold the Mayer expansion converges for the two dimensional Yukawa gas without any cutoff. This is notable because this interaction is singular at ${\bf r} =0$ and is not stable in the Ruelle sense. The reader will find this exercise done in \cite{BrKe87}. The Yukawa gas with hard core is convergent under the conditions \cite{BrKe87}, see also \cite{Imb83a}, \begin{equation} |z|\lambda ^{3} \ll 1; \ \ \ |z|\beta L ^{2} \ll 1; \ \ \ |z|\beta^{3} e^{\beta /\lambda } \ll 1 \end{equation} Briefly, this is derived by combining the hard core example with the Yukawa by letting $s$ vary over $[0,2]$ so that $V({\bf r} ,s) $ has no interactions at $\bs = 0$ and has the complete interaction when $s = 2\openone $. For $s \in [0,1]$ we choose $V'$ as in the repulsive example and for $s \in [1,2]$ we have $V'$ given by (\ref{E.10}) with $\ell(s) = L$ at $s=1$ and $\ell(s) = 0$ at $s=2$. The main point is that the hard core is already there when the Yukawa is turned on so that $2B(s)$ for $s$ corresponding to $\ell(s) \approx 0$ is the stability constant including the hard cores, which prevents a divergence. Tree graph formulas were applied iteratively by \cite{GoMa81} to expand successive scales in the interaction in a renormalization group analysis. They showed how to adapt stability estimates to scales and particular clusters of particles. They were motivated to look for better estimates on the convergence of Mayer expansions in order to prove confinement at all parameters for three dimensional Euclidean lattice quantum electromagnetism by exploiting a connection with screening. The iterated tree graph formulas of G\"opfert and Mack were then simplified \cite{BrKe87} by passing to a limit in which the scales are replaced by a continuous parameter (the $t$ in Theorem~\ref{thm-forest}). This was partly motivated by the study of renormalization \cite{Pol84}. \cite{AbRi95} then found Theorem~\ref{thm-forest} as a corollary of the results in \cite{BrKe87}. The ideas in \cite{GoMa81} were independently developed by Gallavotti and Nicolo and applied by Benfatto to the two dimensional Yukawa gas \cite{Gal85,GaNi85,Ben85,GaNi86}. Note that the trees in these papers express the hierarchical structure of the clusters on different scales and should not be confused with our trees which are Feynman or Mayer graphs. \subsubsection{The Mayer expansion for a polymer gas}\label{subsubsec-polymer-gas} This application is technical and should be omitted on first reading. It will be used later in this chapter. Consider the grand canonical partition function \begin{equation}\label{E.poly1} Z(\Lambda) = \sum_{N} \frac{1}{N!} \sum _{\stackrel{X_{1},\dots ,X_{N} \subset \Lambda}{\cup X_{j} = \Lambda } } \prod _{j} A(X_{j}) \end{equation} where $\Lambda $ is a finite set. In applications to lattice models $\Lambda $ is all lattice points in a big box. A typical element of $\Lambda $ is denoted by $\eta $. Swimming around in this big box are the ``polymers'' $X_{1}, \dots ,X_{N}$, which are disjoint subsets of $\Lambda $. For lattice models they are often nearest neighbor connected but instead of making this assumption we assume that the activity $A (X)$ of polymer $X$ obeys a bound of the form \begin{equation}\label{E.conv1} |A(X)| < a^{|X|} \sum _{G}\prod _{\eta \eta ' \in G} w(\eta, \eta ') \end{equation} where $|X|$ is the number of elements $\eta $ in $X$, $G$ is summed over all tree graphs on $X$. For example when $\Lambda $ is a box in a lattice, we could take $w(\eta, \eta ')$ to vanish whenever $\eta, \eta '$ are not nearest neighbors and to be one otherwise. In this case $A ( X)= 0$ unless $X$ is connected. Now we show that the Mayer expansion is convergent when \begin{equation}\label{E.poly2} Q:= a \sum _{\eta '} \left( \delta_{ \eta ,\eta '} + w(\eta,\eta ') \right) < e^{-1} \end{equation} where $\delta_{ \eta ,\eta '}$ is the Kronecker delta function on the elements of the set $\Lambda $. Suppose there is a notion of distance between elements in $\Lambda $ and $w(\eta,\eta ')$ decays exponentially as ${\rm dist} (\eta, \eta ' )\rightarrow \infty $. If $eQ <1$ then correlations obtained by differentiating $\ln Z(\Lambda) $ with respect to external fields $\psi $ with $A (X) = A (X,\psi )$ decay exponentially. For example $\eta$ could be a point in a lattice in ${\Bbb R}^{3}$ which actually represents a unit box centered on $\eta$ so that $X$ represents a union of these boxes. If $A(X,\psi )$ is independent of variations of $\psi ({\bf r})$ at points ${\bf r}$ outside $X$ then \begin{equation}\label{E.poly3} \frac{\delta ^{2}}{\delta \psi ({\bf r} ) \delta \psi ({\bf r} ')} \ln Z (\Lambda ) \end{equation} and higher variational derivatives will decay exponentially , uniformly in the size of $\Lambda $ as the points ${\bf r} , {\bf r} '$ are separated. We drop the constraint that polymers are disjoint from the sum in (\ref{E.poly1}) and impose it instead by a hard core Gibbs factor $\exp (-U_{N} )$. Let \begin{equation}\label{E.log4} J(X_{1},\dots ,X_{N}) = A(X_{1})\cdots A(X_{N}) \left( e^{- U_{N}(X_{1},\dots ,X_{N})} \right )_{c} \end{equation} In the Mayer expansion \begin{equation}\label{E.log1} \sum_{N \geq 1} \frac{1}{N!} \sum _{X_{1},\dots ,X_{N} \subset \Lambda } J(X_{1},\dots ,X_{N}) \end{equation} we write \begin{equation}\label{E.log1c} \sum _{X_{i}} = \sum \frac{1}{n_{i}!} \sum _{\eta _{1},\dots ,\eta _{n_{i}}} \dots \end{equation} Then we set $M = \sum_{i} n_{i}$ and note that $M!/(\prod n_{i}!)$ is the number of ways to partition $\{1,\dots ,M \}$ into subsets $\gamma _{1},\dots ,\gamma _{N}$. Therefore (\ref{E.log1}) is the same as \begin{eqnarray}\label{E.log1b} && \sum _{M\geq 1}\frac{1}{M!} \sum_{N\geq 1} \frac{1}{N!} \sum _{\gamma _{1},\dots ,\gamma _{N}} \sum _{\eta _{1},\dots ,\eta _{M}} J(X_{1},\dots ,X_{N}) \nonumber \\ && X_{i} = \{\eta _{j}: j \in \gamma _{i}\} \end{eqnarray} with the constraint on the sum over $\eta _{1},\dots ,\eta _{M}$ that $\eta _{j}$ with $j$ in the same partition are distinct. We substitute in the tree graph domination bound (\ref{C.2c}) \begin{eqnarray}\label{E.log2} |\left( e^{- U_{N}(X_{1},\dots ,X_{N})} \right )_{c}| &\leq& \sum _{T}\prod _{ij \in T} \bigg | e^{-V(X_{i}, X_{j})} - 1 \bigg | \nonumber \\ &\leq& \sum _{T}\prod _{ij \in T} \sum _{\eta \in X_{i}, \eta ' \in X _{j}} \bigg |e^{-V(\eta, \eta')} - 1 \bigg | \end{eqnarray} and (\ref{E.conv1}). Note that the hard core interaction $V(\eta, \eta')$ means that $\exp \left(-V(\eta, \eta') \right) - 1$ is $-\delta_{ \eta ,\eta '}$. The trees with $w$ bonds and the tree with $\delta_{ \eta ,\eta '}$ bonds link into one connected tree $G$ on $\{1,\dots M \}$ and the respective sums are equivalent to a sum over all possible connected trees on $\{1,\dots M \}$. The partition $\{\gamma _{1},\dots ,\gamma _{N} \}$ determines uniquely which bonds in this tree are $V$ bonds and which are $w$ bonds. Indeed the sum over partitions, including the $1/N!$ is equivalent to summing over the choice $w$ bond or $V$ bond for each bond in the tree. Therefore the sum over partitions is the same as assigning to each bond $ij$ \begin{equation}\label{E.log3} \delta_{\eta_{i},\eta_{j}} + w(\eta_{i} , \eta_{j}) \end{equation} Therefore the expansion (\ref{E.log1b}) is bounded by \begin{equation}\label{E.log5} \leq \sum _{M\geq 1}\frac{a^{M}}{M!} \sum _{\eta _{1},\dots ,\eta _{M}} \sum _{G}\prod _{ij \in G} (\delta_{\eta_{i},\eta_{j}} + w(\eta_{i}, \eta_{j})) \end{equation} dropping the constraints on the distinctness of $\eta _{1},\dots ,\eta _{M}$. As above, (\ref{E.tree1}--\ref{E.tree1b}), this is convergent if (\ref{E.poly2}) holds. The additional claim we made concerning exponential decay of correlations is an immediate consequence: suppose we perform variational derivatives such as \[ \frac{\delta }{\delta \psi (\eta _{a})} \frac{\delta }{\delta \psi (\eta _{b})} \] on $\ln Z$ where $\psi $ is an external field that enters through dependence $A(X) = A (X,\psi )$. Then all terms in the Mayer expansion (\ref{E.log1}) vanish except those where the union of the sets $X_{j}$ contains both points $\eta _{a},\eta _{b}$, because $\delta A (X,\psi )/ (\delta \psi (\eta ) ) = 0$ when $\eta \not \in X$. When $\eta _{a}, \eta _{b}$ are far apart the surviving terms are exponentially small since they are connected graphs with exponentially decaying propagators. This argument that is being made term by term in the Mayer expansion is valid when all the terms are summed because we established that the expansion is convergent, provided the convergence has some uniformity in $\psi $. \subsection{Gaussian integrals and truncated expectations}\label{sec-truncated-expectations} In this section, following \cite{Bry84,BrKe87} we derive tree graph formulas for truncated Gaussian expectations. The point of the formulas is that all the loops in the associated Feynman diagrams are resummed back into a positive Gaussian measure, leaving only tree diagrams, which carry the connectedness information. Truncated expectations for arbitrary observables will be defined later in (\ref{B.14}) but for products of polynomials they can equivalently be defined by applying Wick's theorem and then discarding all but the connected graphs. Suppose $\av{ \ }$ is a Gaussian average over variables $\phi _{i}, \ i = 1,\dots N$, \begin{eqnarray}\label{D.4} && \av{P} = Z^{-1} \int P e^{-1/2 \sum _{i,j} \phi _{i} A_{ij} \phi _{j}} \prod _{i}d\phi _{i} \nonumber \\ && \av{ \phi_{i} \phi_{j}} = A^{-1}_{ij} \equiv C_{ij} \end{eqnarray} $A$ is a matrix with positive eigenvalues. Suppose $P_{i} = P_{i}(\phi _{i})$ are polynomials each of which depends only on one of the variables, then by Wick's theorem \begin{equation}\label{D.laplacian1} \av{\prod _{l}P_{l}} = e^{\frac{1}{2}\sum _{i,j}\DDelta_{ij}} \prod _{l}P_{l}\arrowvert_{\phi=0} \nonumber \\ \end{equation} where the exponential is defined by power series and \begin{equation}\label{D.laplacian2} \DDelta _{ij} = C_{ij} \frac{\partial }{\partial \phi _{i}} \frac{\partial }{\partial \phi _{j}} \end{equation} We apply theorem~\ref{thm-forest} with \begin{equation}\label{D.laplacian3} F(\bs ) = \exp \left( \frac{1}{2} \sum _{i,j} s_{ij}\DDelta_{ij} \right) P \end{equation} with $P = \prod _{l}P_{l}$ and the conventions $s_{ij} = s_{ji}$ and $s_{ii}=1$. The right hand side of theorem~\ref{thm-forest} will then contain \begin{equation}\label{D.laplacian4} \exp \left( \frac{1}{2}\sum _{i,j} \sigma_{ij}(G,\bs )\DDelta_{ij} \right) \DDelta ^{G} P \arrowvert_{\phi=0} \end{equation} where $\DDelta ^{G} = \prod _{ij \in G} \DDelta _{ij}$. This is a new Gaussian expectation $\av{ \ }^{G,\bs }$ characterized by \begin{equation}\label{D.4b} \av{ \phi_{i} \phi_{j}}^{G,\bs } = \sigma_{ij}(G,\bs ) C_{i,j} \end{equation} The key point is that it truly is a Gaussian expectation: we shall show below that $C_{ij}\sigma_{ij}$ has positive eigenvalues, so that it has an inverse which defines the expectation $\av{ \phi_{i} \phi_{j}}^{G,\bs }$ in parallel to (\ref{D.4}). We define \begin{equation}\label{D.expectation} \av{P}^{G} := \int^{1} d^{G}s \, \av{ P }^{G,\bs } \end{equation} and conclude from theorem~\ref{thm-forest} that \begin{equation}\label{D.truncated1} \av{\prod _{i}P_{i}} = \sum _{G} \av{\DDelta ^{G} \prod _{i}P_{i}}^{G} \end{equation} where $G$ is summed over all forests. Notice that the previous equations contained exponentials of $\DDelta$ which only make immediate sense on polynomials, but now that we have got rid of these we can use this formula when the $P_{i}$ are arbitrary smooth functions. By the discussion above theorem~\ref{thm-mayer2} there is a factorization $\av{ \ }^{G} = \prod_{T\subset G} \av{ \ }^{T}$ where $T$ runs over trees in the forest $G$. Therefore by definition of the truncated expectation \footnote{A superscript $T$ refers to a tree graph and a subscript $T$ means the expectation is truncated. We apologize for the notational collision.} $\av{ \ }_{T}$ given in (\ref{B.14}) we obtain \begin{equation}\label{D.truncated2} \av{\prod _{i=1}^{N}P_{i}}_{T} = \sum _{G} \av{\DDelta ^{G}\prod _{i}P_{i}}^{G} \end{equation} where $G$ runs only over connected trees on $N$ vertices. This argument easily generalizes to allow each $P_{i}$ to depend on more variables $\phi _{x}, \ x \in X_{i}$. This formula expresses a truncated Gaussian expectation as a sum only over tree graphs. The loops that would appear in a standard application of Wick's theorem are resummed into the expectation $\av{ - }^{G}$ defined in terms of the Gaussian expectation with the altered propagator $C_{ij}\sigma _{ij}$. To see that the matrix $C_{ij}\sigma _{ij}$ has positive eigenvalues: since $A_{ij}$ has positive eigenvalues, $C_{ij}$ has positive eigenvalues which is equivalent to \begin{equation}\label{D.4a} \sum _{i,j} \eta _{i}C_{ij}\eta _{j} \geq \lambda _{\rm min} \sum _{j}\eta _{j}^{2} \end{equation} By theorem~\ref{thm-C.stability.a} with $V_{ij} = \eta _{i}[C_{i,j} - \lambda _{\rm min}\delta _{ij}] \eta _{j}$, $B=0$ and self energies included, \begin{equation}\label{D.4c} \sum _{i,j} \eta _{i}C_{ij}\sigma_{ij}\eta _{j} \geq \lambda _{\rm min} \sum _{j}\eta _{j}^{2} \end{equation} which implies that the matrix $A (\bbox{\sigma} )^{-1} = C_{ij}\sigma_{ij}$ has positive eigenvalues. We define the expectation $\av{ \ }^{G,\bs }$ by replacing the matrix $A$ by $A (\bbox{\sigma} )$ in (\ref{D.4}). \subsection{Polymer representations and exponential decay}\label{sec-polymer-rep2} In this section we describe how to obtain a convergent expansion for the logarithm of functional integrals of the form \begin{equation}\label{8.pol-rep-1} \av{Z(\Lambda )e^{i\int \phi f}} = \int d \mu (\phi ) Z(\Lambda,\phi) e^{i\int \phi f} \end{equation} where $d\mu $ is a Gaussian measure with a propagator [covariance] $C ({\bf r} ,{\bf r} ')$ that has both infra-red and ultraviolet cutoffs so that it is smooth on the diagonal and has exponential decay as ${\bf r} - {\bf r} ' \rightarrow \infty $. The $\Lambda $ in the functional $Z(\Lambda ,\phi )$ means that variational derivatives at points outside $\Lambda $ vanish: \begin{equation}\label{8.locality1} Z_{1}(\Lambda ,\phi ; f) \equiv \frac{\partial }{\partial \alpha } Z(X,\phi+\alpha f)_{|_{\alpha =0}} = 0 \end{equation} if $f = 0$ in $\Lambda $ and the same is true for all higher derivatives. Functionals such as \begin{equation}\label{8.ex1} Z(\Lambda,\phi) = \exp ( z \int_{\Lambda } \cos \phi ({\bf r} ) \, d{\bf r} ) \end{equation} factorize \begin{equation}\label{8.locality2} Z(X_{1}\cup X_{2},\phi) = Z(X_{1},\phi) Z( X_{2},\phi) \end{equation} but we will work with a less restrictive property, which is that $Z(\Lambda ,\phi )$ has a {\it polymer representation} \begin{equation}\label{8.locality3} Z(\Lambda ,\phi ) = \sum_{N} \frac{1}{N!} \sum _{\stackrel{X_{1},\dots ,X_{N} \subset \Lambda}{\cup X_{j} = \Lambda } } \prod _{j} K(X_{j},\phi) \end{equation} where the interiors of the sets $X_{1},\dots ,X_{N}$ are disjoint. In other words $Z(\Lambda ,\phi )$ is at least a sum over contributions that factorize. This type of representation is not useful if the sets $X_{j}$ are very complex so we insist that all sets are finite unions of unit cubes in ${\Bbb R} ^{3}$, centered on points with integral coordinates. Recall the well known combinatoric miracle that when the partition function has a graphical expansion, the logarithm has the same expansion except that only the connected graphs appear. This comes about as a result of the following easily verified relation (cumulant expansion) between expectations and exponentials \begin{equation}\label{B.connected} \av{e^{F}} = \exp \left(\sum_{N\geq 1} \frac{1}{N!}\av{F^{N}}_{T} \right) \end{equation} where the \underline{t}runcated expectation $\av{F_{1} \cdots F_{N}}_{T}$ is recursively defined by solving \begin{eqnarray}\label{B.14} && \av{F_{1} \cdots F_{N}} = \sum _{\pi }\prod _{\gamma \in \pi } \av{F^{\gamma } }_{T} \mbox{ if } N > 1 \nonumber \\ && \av{F_{1}}_{T} = \av{F_{1}} \end{eqnarray} where $\pi $ is summed over all partitions of $\{1,2,\dots ,N \}$ and $F^{\gamma } = \prod _{j\in \gamma }F_{j}$. For the case (\ref{8.ex1}) the expansion for the logarithm of (\ref{8.pol-rep-1}) based on these relations is convergent at small $z$, essentially because the cosine has the special property that it and all of its derivatives are bounded uniformly in $\phi $ \footnote{By reversing the Sine-Gordon transformation the expansion becomes the Mayer expansion for $\pm 1$ charges interacting by two body potential $C ({\bf r} ,{\bf r} ')$.}. The next steps can be viewed as a reduction of more general functionals such as $\exp (-\lambda \int \phi ^{4} )$ to this bounded case. Relations similar to (\ref{B.connected},\ref{B.14}) hold for the polymer representation (\ref{8.locality3}). This is plausible once one realizes that the polymer representation is an exponential in a different commutative product $\circ$ defined by \begin{equation}\label{B.17} A \circ B(X) = \sum_{Y\subset X} A(Y)B(X \setminus Y). \end{equation} As always $Y, X$ are finite unions unit cubes. $Y, X$ are permitted to be the empty set. We do not distinguish between open and closed cubes, that is we identify two cubes if they have the same interior and regard them as disjoint if their interiors are disjoint. Define \begin{equation}\label{B.16b} {\cal E} xp(K) = {\cal I} + K + \frac{1}{2!}K \circ K + \cdots \end{equation} ${\cal I}(X) = 0$ unless $X$ is the empty set, in which case it is one. ${\cal I}$ is the identity for this product. We only define ${\cal E} xp(K)$ for functionals $K$ that vanish when applied to the empty set. It is easy to verify that \begin{equation}\label{B.16} {\cal E} xp(K)(X ,\phi ) = \left\{ \begin{array}{l} \sum_{N \geq 0} \frac{1}{N!} \sum _{\stackrel{X_{1},\dots ,X_{N} \subset X}{\cup X_{j} = X} } \prod _{j} K(X_{j},\phi)\\ 1 \mbox{ if } X = \emptyset \end{array} \right. \end{equation} Thus $Z (\Lambda ,\phi ) = {\cal E} xp (K) (\Lambda ,\phi )$. Furthermore ${\cal E} xp(A + B) = {\cal E} xp(A) \circ {\cal E} xp(B)$. As an example, consider the function $\Box (X)$ that equals one if $X$ is a cube and vanishes in all other cases. Then by (\ref{B.16}), for any set $X$, \begin{equation}\label{B.expbox} {\cal E} xp(\Box )(X) =1 \end{equation} Returning to our functional integral (\ref{8.pol-rep-1}), let us define \begin{equation}\label{B.19} K_{T}(X) = \sum_{N\geq 1} \frac{1}{N!} \av{ K^{\circ N}(X)}_{T}, \ \ \ = 0 \mbox{ if } X = \emptyset \end{equation} More explicitly \begin{equation}\label{B.ncirc} \av{ K^{\circ N}(X)}_{T} = \av{K\circ \dots \circ K(X) }_{T} = \sum _{\stackrel{X_{1},\dots ,X_{N} \subset X} {\cup X_{j} = X } } \av{\prod _{j} K(X_{j}) }_{T} \end{equation} where the truncated expectation $\av{ \ }_{T}$ is defined as above. Then $K_{T}$ is the logarithm in the sense \begin{equation}\label{B.18} \av{{\cal E} xp(K)(X )} = {\cal E} xp(K_{T})(X ) \end{equation} which is the $\circ$ equivalent of (\ref{B.connected}). However these series are terminating after a finite number of terms that depends on the set $X$ because there are at most a finite number of partitions of any given $X$. As an extreme example, if $X = \Delta $ is a single cube, then there are no partitions of $X$ into proper subsets so only the $N= 1$ term in (\ref{B.19}) survives and accordingly $K_{T}(\Delta ) = \av{K(\Delta )} $. This analysis is unaffected by the external field term $\exp (\int \phi f )$ in the partition function (\ref{E.poly1}) because it can be absorbed into the polymer activities $K$. From the definition of ${\cal E} xp$ \begin{equation}\label{8.pol-rep-2} {\cal E} xp (K (\phi)) (\Lambda ) e^{\int_{\Lambda }\phi f} = {\cal E} xp (K (\phi ,f)) (\Lambda ), \ \ \ K (X,\phi ,f) := K (X,\phi ) e^{\int_{X}\phi f} \end{equation} We want an expansion for the standard logarithm of the partition function. Comparing the definition of ${\cal E} xp$ in (\ref{B.16}) with the polymer gas partition function (\ref{E.poly1}), we see that the standard logarithm is given by the Mayer expansion described in section~\ref{subsubsec-polymer-gas}, provided we can weaken the constraint $\cup X_{j} = X$ that occurs in ${\cal E} xp(K_{T})(X )$ to $\cup X_{j} \subset X$. There is a simple way to achieve this: for any function of sets $\tilde{A} (X)$ declare that $\tilde{A}$ is normalized if $\tilde{A} (X) = 1$ whenever $X$ is a unit box, then \begin{eqnarray} {\cal E} xp(\tilde{A})(X) &=& {\cal E} xp(\Box) \circ {\cal E} xp(\tilde{A} - \Box) (X) \nonumber \\ &=& \sum_{N} \frac{1}{N!} \sum _{X_{1},\dots ,X_{N} \subset X} {\cal E} xp(\Box) (X \setminus \cup X_{j}) \prod _{j} A(X_{j}) \nonumber \\ &=& \sum_{N} \frac{1}{N!} \sum _{X_{1},\dots ,X_{N} \subset X} \prod _{j} A(X_{j}), \ \ \ A := \tilde{A} - \Box \end{eqnarray} The constraint $\cup X_{j} = X$ has been lifted because by (\ref{B.expbox}) ${\cal E} xp(\Box) = 1$ on $X \setminus \cup X_{j}$. If $\tilde{A}$ is \begin{equation}\label{B.hat} \tilde{A}(X) = K_{T}(X)/\prod _{\Delta \subset X}K_{T}(\Delta ), \end{equation} where $/$ has its usual division meaning, then it is normalized. Also, by the same principle as was used in (\ref{8.pol-rep-2}) \begin{equation}\label{B.norm1} {\cal E} xp(K_{T})(X) = \left( \prod _{\Delta \subset X}K_{T}(\Delta ) \right) {\cal E} xp(\tilde{A})(X) \end{equation} so \begin{equation}\label{B.poly} {\cal E} xp(K_{T})(X) = \left( \prod _{\Delta \subset X}K_{T}(\Delta ) \right) \sum_{N} \frac{1}{N!} \sum _{X_{1},\dots ,X_{N} \subset X} \prod _{j} A(X_{j}) \end{equation} In section~\ref{subsubsec-polymer-gas} we gave a criterion (\ref{E.conv1},\ref{E.poly2}) for how small the activity $A(X)$ must be for convergence of the Mayer expansion and exponential decay of correlations. $A$ is built out of $K_{T}$ so we need to see that $K_{T}$ obeys a bound similar to (\ref{E.conv1},\ref{E.poly2}). This is where section~\ref{sec-truncated-expectations} comes into play. By the tree graph formula (\ref{D.truncated2}) $K_{T}$ is small provided variational derivatives with respect to $\phi $ of the initial polymer activity $K (X,\phi)$ in the partition function (\ref{8.pol-rep-1}) are small. We will give a partial explanation of this through the following example. Example: suppose \begin{equation}\label{B.poly2} Z (\Lambda) = e^{-V (\Lambda )}, \ \ \ V (\Lambda,\phi ) = \lambda \int_{\Lambda }\phi ^{4} \end{equation} Because this factors into a product over boxes in $\Lambda $, by the same argument that shows that ${\cal E} xp (\Box) = 1$ this equals ${\cal E} xp (K)$ with \[ K (X,\phi ) = \Box (X) e^{-V (X,\phi )} \] When $\lambda $ is small variational derivatives of $K (X,\phi )$ are small, uniformly in $\phi $, because \[ \frac{d}{d\alpha } e^{-V (\Delta,\phi +\alpha g)} = -4\lambda \int_{\Delta } \phi ^{3} g \, d{\bf r} \, e^{-\lambda \int_{\Delta } \phi ^{4} } \sim O (\lambda ^{1/4}) \] because $\phi ^{4} $ controls $\int \phi ^{3}g$ when $\phi $ is large. This is somewhat of an oversimplification because the dependence of the estimate on the test function $g$ is not uniform. Unfortunately a gentle reference to cover these specific missing details is not available, but \cite{Lem95} is detailed and close in spirit to the discussion here. More general background references are \cite{GlJa87,Riv91,BrFe92}. The ability to work with the bounded $\exp (-\lambda \int \phi ^{4} )$ in place of $\int \phi^{4}$ is half the reason why we obtain a convergent expansion when $\lambda $ is small. The other half is that all formulas involve at most tree graphs whose numbers are compensated by $N!$ factors, see the discussion at the beginning of section~\ref{sec-mayer2}. Exponential decay of correlations for functional integrals of the form (\ref{8.pol-rep-1}) follows as in the discussion of the polymer gas in section~\ref{sec-mayer2}. It is possible that polymer representations could be useful in numerical renormalization group calculations. This has been considered \cite{MaPo85,MaPo89}. Their program can be viewed as a natural step beyond methods that do work well for hierarchical models \cite{KoWi91,KoWi94}. \subsection{The polymer representation}\label{sec-polymer-rep1} In this section we indicate how exponential decay of correlations for the Coulomb system is deduced from the considerations of section~\ref{sec-polymer-rep2}. The missing step, outlined in this section, is to show that the Sine-Gordon representation is an integral of the form (\ref{8.pol-rep-1}) with $Z (\Lambda ,\phi ) = {\cal E} xp (K)(\Lambda ,\phi )$ for polymer activity $K (X,\phi )$ which becomes very small as the Debye-H\"uckel limit is approached. As introduced in section~\ref{sec-debye-huckel2} we use units where $l_{D} = 1$. We also simplify the presentation by discussing a Coulomb system with two species with equal but opposite charges and equal activities $z$ interacting by a smoothed Coulomb potential of the form $V_{\infty ,L}$ with $L = l_{D}$. Then there is no need for the Mayer expansion of section~\ref{sec-debyesphere}, which complicates the original proofs. With these simplifications the left hand side $\exp \left(F \right)$ in the ``Villain'' approximation (\ref{2.7.villain}) becomes the ideal gas partition function \begin{equation}\label{A.local1} \exp \left(\int_{\Lambda } d{\cal E} \, z^{(L)}({\cal E} ) e^{-\beta ^{1/2} e_{\alpha } i\phi} \right) \end{equation} We make the dependence on $\Lambda $ explicit by writing $\exp \left(F(\Lambda ,i\phi ) \right)$. The left hand side of the Villain approximation (\ref{2.7.villain}) factors \begin{equation} e^{F(X \cup Y,i\phi )} = e^{F(X ,i\phi )} e^{F(Y,i\phi )} \end{equation} whenever $X$ and $Y$ are disjoint. The sum over $h$ in the right hand side of the Villain approximation also factors, if $X,Y$ are unions of unions of cubes. From this there follows the same property \begin{equation}\label{A.local2} e^{R_{3}(X \cup Y,i\phi )} = e^{R_{3}(X ,i\phi )} e^{R_{3}(Y,i\phi )} \end{equation} for $R_{3}$. From (\ref{2.7.6}) \begin{eqnarray}\label{A.1} &&\hspace{-.5in} \Xi_{\Lambda } /(\Xi_{{\rm ideal},\Lambda }\Xi_{{\rm DH},\Lambda }) = \int d\mu _{{\rm DH}}(\phi) \, \sum _{h} e^{- S_{{\rm DH}}(- ih, ig)} \nonumber \\ && \times e^{-{S'}_{{\rm DH}}(-ih , ig; i\phi) + R_{3}(ig + i\phi)} \end{eqnarray} For simplicity we set the external field $\psi = 0$. In the formula (\ref{2.7.Sdef}) for $S_{{\rm DH}}(- ih, ig)$, $R_{2}, R_{0}$ vanish for this simplified model. The propagator for $d\mu _{{\rm DH}}(\phi) $ decays exponentially which is the property we need in order to use the expansion from section~\ref{sec-polymer-rep2}. Now we write the $d\mu $ integrand in the form \begin{equation}\label{A.2} {\cal E} xp (K) (\Lambda ,\phi ) = \sum_{N} \frac{1}{N!} \sum _{\stackrel{X_{1},\dots ,X_{N} \subset \Lambda}{\cup X_{j} = \Lambda } } \prod _{j} K(X_{j},i\phi) \end{equation} The rough idea will be that each $X_{j}$ is a contour $\Gamma _{j}$ or several conglomerated contours together with a collar of width $M$ which is thick enough so that the effect of the contour is locked inside $X_{j}$. To take care of regions without contours there will also be the dominant possibility that $X$ is a single cube. The outcome of our manipulations will be a polymer activity $K (X,\phi )$ whose $p-$th variational derivative with respect to $\phi $ is of size \begin{eqnarray}\label{A.K1} \frac{p!}{a^{p}} \exp \left( \kappa \|\phi \|^{2}_{X} - O(a)|X| \right ), \ \ \ a = [z l_{D}^{3}]^{1/2}, \ \ \ |X| > 1 \end{eqnarray} $a^{2}$ is the number of particles in a Debye sphere and is very large near the Debye-H\"uckel limit. The $\exp (- O(a)|X|)$ is the tunneling factor promised in section~\ref{sec-tunneling}. $|X|$ is the number of unit boxes in $X$. $\| \ \|_{X}$ is a norm involving $\int _{X}|\phi |^{2}$ and integrals of derivatives of $\phi $. The constant $\kappa $ is small enough so that $\exp \left(\kappa \|\phi \|^{2}_{X} \right)$ can be integrated with respect to $d\mu _{{\rm DH}}$. Since polymers never overlap this is enough for a proof that the expansion of section~\ref{sec-polymer-rep2} is convergent. The non-overlapping of polymers is the reason why they are useful for obtaining convergent expansions. To understand why this crucial limitation on $\kappa $ cannot be achieved unless the tunneling effects represented by contours are taken into consideration consider the one dimensional version of the Villain approximation. \begin{equation}\label{A.villain} e^{a ^{2} \cos (\phi/a )} = e^{R_{3}(\phi)} \sum _{n} e^{-\frac{1}{2}(\phi-2\pi a n)^{2}+a ^{2}} \end{equation} If we restrict $R_{3}(\phi)$ to the period $[-\pi a ,\pi a ]$ and note that it is largest at $\phi = a \pi $ we find, by throwing away all but the $n = 0$ term, an upper bound \begin{equation}\label{A.Rsub3-1} e^{R_{3}(\phi)} \leq e^{\frac{1}{2} \kappa \phi^{2}}; \ \ \ \kappa = 1 - \frac{4}{\pi ^{2}} < 1 \end{equation} Since $R_{3}(\phi)$ is a periodic function the bound holds for all $\phi$. $R_{3}$ is analytic with more or less the same Gaussian bound in the strip $\phi + iy$ with $|y| / a \ll 1$ so \begin{equation}\label{A.Rsub3-3} \frac{d^{p}}{d\phi ^{p}} e^{R_{3}(\phi )} \sim \frac{p!}{a ^{p}} e^{\frac{1}{2} \kappa \phi^{2}} \end{equation} $\kappa < 1$ means that $\exp \left(R_{3} \right)$ and all derivatives can be integrated against the one dimensional analogue of $d\mu _{{\rm DH}}$, namely the Gaussian $\exp \left(-\phi^{2}/2 \right)$ \begin{equation}\label{A.Rsub3-2} \frac{1}{\sqrt{2\pi }} \int e^{-\phi^{2}/2} \frac{d^{p}}{dx^{p}}e^{R_{3}(\phi)}dx = \left\{ \begin{array}{ll} O(a ^{-p})& \mbox{if } p \not = 0\\ \rightarrow 1& \mbox{if } p = 0 \end{array} \right. \end{equation} as $a \rightarrow \infty $. This fails if $R_{3}$ is instead defined as the error in the single well approximation $a ^{2}\cos (\phi/a) \approx 1 - \phi^{2}/2 + R_{3}$. The tunneling phenomena contained in the Villain approximation are invisible in perturbation theory but they are the reason why perturbation theory will not converge. The appearance of $\exp (-a |X| )$ in (\ref{A.K1}) quantifies the physical idea that the tunneling effects are very rare in the Debye-H\"uckel limit when $a$ is large. This bound (\ref{A.K1}) together with the considerations of section~\ref{sec-polymer-rep2} implies the exponential decay of correlations. The first task on the way to (\ref{A.2}) is to obtain an exact version of (\ref{2.7.Sdecomp}) so that the sum over $h$ will factor as in (\ref{2.7.contour}). Let $M$ be a (large) integer. Given $h$ we choose the translation $g$ to minimize $S_{{\rm DH}}(- ih, ig)$ subject to the constraint that $g({\bf r} ) = h({\bf r} )$ at every point that is distant by more than $M-1$ from any contour in $h$. The price of imposing this constraint is that in (\ref{A.1}) we no longer have ${S'}_{{\rm DH}}(-ih , ig; i\phi) = 0$ as was the case in section~\ref{sec-tunneling}. Now ${S'}_{{\rm DH}}(-ih , g; i\phi)$ is an integral of $\phi$ and derivatives of $\phi$ times derivatives of $g$ over the surface of the region where the constraint holds. These derivatives are of order $O(\exp(-M))$ because the constraint is almost irrelevant since the minimizer without the constraint satisfies $g({\bf r} ) = h({\bf r} ) + O(\exp \left(-{\rm dist}({\bf r} ,\Gamma \right))$. We choose $M$ large so that ${S'}_{{\rm DH}}(-ih , g; i\phi)$ is as small as the error term $R_{3}(ig + i\phi)$. With this choice of $g$ we have an exact version of the additivity property of $S_{{\rm DH}}(- ih, ig)$ in (\ref{2.7.Sdecomp}): whenever $h = h_{1} + h_{2}$, \begin{equation}\label{A.local3} S_{{\rm DH}}(- ih, ig) = \sum _{j=1,2} S_{{\rm DH}}(- ih_{j}, ig_{j}) \end{equation} provided every contour in $h_{1}$ is separated by a distance of $2M$ or more from every contour in $h_{2}$. See the discussion below (\ref{2.7.ising}) and note that our constraint forces the cross terms to vanish. Under the same conditions we also have \begin{equation}\label{A.local4} {S'}_{{\rm DH}}(-ih , g; i\phi) = \sum _{j} {S'}_{{\rm DH}}(-ih_{j} , g_{j}; i\phi) \end{equation} Define $K(X,i\phi )$ to vanish if $X$ is the empty set or not a connected union of cubes, otherwise \begin{equation}\label{A.5} K(X,i\phi) = e^{R_{3}(X,ig + i\phi)} \left\{ \begin{array}{l} \sum_{h: X(h) = X} e^{- S_{{\rm DH}}(- ih, ig)} e^{-{S'}_{{\rm DH}}(-ih , ig; i\phi) } \\[.1in] 1 \end{array} \right. \end{equation} The second line applies when $X$ is a single cube. The constraint $X(h) = X$ is defined below. The constraint forces $X$ to hold contours when it is not a single cube and contours are responsible for the $\exp (-O (a)|X| )$ in (\ref{A.K1}). This definition leads to the desired representation \begin{equation}\label{A.6} \Xi_{\Lambda } /(\Xi_{{\rm ideal},\Lambda }\Xi_{{\rm DH},\Lambda }) = \int d\mu _{{\rm DH}}(\phi) \, \sum_{N} \frac{1}{N!}\sum _{\stackrel{X_{1},\dots ,X_{N} \subset \Lambda}{\cup X_{j} = \Lambda } } \prod _{j} K(X_{j},i\phi) \end{equation} It is convenient to measure distance using \begin{equation}\label{A.dist} {\rm dist} ({\bf r} ,{\bf r} ') = \max _{i} |r_{i} - r_{i}'| \end{equation} so that the unit cube centered on the origin is $\{{\bf r} : {\rm dist} (0,{\bf r} ) < 1/2\}$. $X(h)$ is defined to be the open set consisting of all points closer than $M$ to some contour in $h$. To derive (\ref{A.6}) note that any $h$ occurring in the sum (\ref{A.1}) determines $X(h)$ which decomposes uniquely into disjoint open connected subsets $X_{j}$. Each $X_{j}$ determines a unique $h_{j}$ such that \begin{eqnarray}\label{A.hdecomp} &&X(h_{j}) = X_{j}; \ \ \ h_{j} = 0 \mbox{ outside }\Lambda \nonumber \\ &&h_{j} = h + \mbox{ a constant when restricted to } X_{j}\nonumber \\ &&h = \sum _{j}h_{j} \end{eqnarray} The first two conditions determine $h_{j}$ and the last is implied by them. Therefore the sum over $h$ in (\ref{A.1}) is equivalent to summing over $N$, $X_{1},\dots ,X_{N}$ followed by $h_{1},\dots ,h_{N}$. By (\ref{A.local3}) and the symmetry \begin{equation}\label{A.symmetry} e^{R_{3}(X_{k},i\sum _{j}g_{j} + i\phi)} = e^{R_{3}(X_{k},ig_{k} + i\phi)} \end{equation} we can factor the summands. This symmetry is the one we remarked in (\ref{2.7.1}) and it holds in this form because for $j \not = k$ $g_{j} = h_{j}$ is constant on $X_{k}$. Incidentally a fractional charge immersed in the system destroys this symmetry. See the remarks on fractional charges in section~\ref{sec-tunneling}. \subsection{Proofs for theorems in chapter VIII } \label{sec-proofs} \bigskip \noindent {\em Proof of theorems~\ref{thm-C.stability.a} and \ref{thm-C.stability.b}:\/} Following \cite{BrKe87}. Let $\chi (a \geq \tau )$ denote a function of $\tau $ which is $1$ when $a \geq \tau$ and zero otherwise. Then \begin{eqnarray} U_{N}\circ \bbox{\sigma} (T,\bs ) &=& \sum _{ij} \int _{0}^{\sigma_{ij}(T,\bs )} d\tau \, V_{ij}'(\tau ) \nonumber \\ &=& \int _{0}^{1} d\tau \, \sum _{ij} V_{ij}'(\tau ) \chi (\sigma_{ij}(T,\bs ) \geq \tau ) \end{eqnarray} A value of $\tau $ and a tree $T$ determine a partition of the particles $=$ vertices of $T$ into clusters: simply erase all bonds $ij \in T$ for which $s_{ij} < \tau $. The rest of the tree falls into connected subtrees. Each cluster consists of the vertices in each subtree. The bonds $ij \in T$ we erased are in fact eliminated by the $\chi$ in the above equation. The key point is that if $lk$ is another bond not in $T$ with $k$ and $l$ is separate clusters that used to be connected by $ij$ then it too is eliminated by $\chi$ because $s_{ij} < \tau $ implies $\sigma_{kl}(T,\bs) < \tau $. Therefore at each value of $\tau $ we have a $\tau $ dependent partition of the vertices into clusters $C_{\alpha }$ such that \begin{equation} \sum _{ij} V_{ij}'(\tau ) \chi (\sigma_{ij}(T,\bs ) \geq \tau ) = \sum _{\alpha }\sum_{ij \in C_{\alpha }} V_{ij}'(\tau ) \end{equation} Thus a stability estimate $\sum V_{ij}'(\tau ) \geq -B(\tau )N$ is inherited by $U_{N}\circ \bbox{\sigma} (T,\bs )$, because we can apply it to each cluster in the last equation and then integrate over $\tau $. Theorem~\ref{thm-C.stability.a} follows easily. {\it Theorem~\ref{thm-C.stability.b}:\/} Following \cite{GoMa81} there is a systematic way to improve stability estimates by using numerically calculated constants for small clusters. The simplest instance is to note that for clusters of single particles we have no interaction to bound so that \begin{equation} \sum _{ij} V_{ij}'(\tau ) \chi (\sigma_{ij}(T,\bs ) \geq \tau ) \geq -B(\tau ) N_{1} \end{equation} where $N_{1}$ is the number of particles not in single clusters, \begin{eqnarray} N_{1} &=& |\{i: i \in \mbox{ some } C_{j} \mbox{ with } |C_{j}| > 1 \}| \nonumber \\ &\leq & |\{i: s_{ij} \geq \tau \mbox{ some } ij \mbox{ in } T \}| \nonumber \\ &\leq & \sum _{i,j} \chi (s_{ij} \geq \tau ) = 2 \sum _{ij} \chi (s_{ij} \geq \tau ) \end{eqnarray} Therefore \begin{equation} U_{N}\circ \bbox{\sigma} (T,\bs ) \geq -\int _{0}^{1} d\tau \, 2 B(\tau ) \sum _{ij} \chi (s_{ij} \geq \tau ) \end{equation} which is (\ref{C.12}). \hspace*{\fill} QED \medskip {\em Proof of Theorem ~\ref{thm-forest}:\/} We know the case $N = 2$. We assume $N > 2$ and make the inductive hypothesis that the theorem holds for all cases with less than $N$ vertices. It is sufficient to prove that both sides have the same $t$ derivative and agree at $t=0$. The $t$ derivative of the right hand side is \begin{equation} \sum _{G \in {\cal F}}\sum _{b \in G} \int^{t} d^{G\setminus b}s \, \partial ^{G} F \circ \bbox{\sigma}(G,\bs )_{s_{b}=t} \end{equation} We interchange the sums and obtain $\sum X_{b}$ where \begin{equation}\label{F.5} X_{b} = \sum _{G \in {\cal F}, G \ni b} \int^{t} d^{G\setminus b}s \, \partial ^{G} F \circ \bbox{\sigma}(G,\bs )_{s_{b}=t} \end{equation} Now we fix $b$. To use the inductive hypothesis, we reduce $N$ by one by rewriting $X_{b}$ as a sum over all forests $\bar{G}$ on $N-1$ vertices. The $N-1$ vertices are obtained from the original vertices $\{1,2,\dots ,N \}$ by identifying the vertices at either end of $b$. The possible bonds $\bar{c}$ are viewed as sets of identified bonds on $\{1,2,\dots ,N \}$. For example, if $b = 12$, then $1i$ and $2i$ become the same bond which we would denote by $\bar{1i}$ or $\bar{2i}$. $\bar{\bs }$ denotes a vector of parameters $\bar{s}_{\bar{c}}$ assigned to bonds $\bar{c}$. Given such parameters, we determine parameters $\bs $ by assigning $\bar{s}_{\bar{c}}$ to all bonds $c \in \bar{c}$ and $t$ to bond $b$. Thus define $\bs = {\bf j}(\bar{\bs})$ where the $c$ component is given by \begin{equation} j_{c}(\bar{\bs}) = \left \{ \begin{array}{ll} \bar{s}_{\bar{c}} & \mbox{ if } c \not = b\\ t & \mbox{ if } c = b \end{array} \right. \end{equation} To each forest $G \ni b$ on $N$ vertices we associate a unique forest $G/b$ on $N-1$ vertices by identifying the vertices in the pair $b$. Thus \begin{equation} G/b = \{\bar{c}: c \in G \} \end{equation} \bigskip \noindent For any function $f$ of parameters $\bs $, \begin{equation}\label{F.6} \sum _{G: G/b = \bar{G}} (\partial ^{G \setminus b} f) \circ {\bf j}(\bar{\bs} ) = \partial ^{\bar{G}} (f \circ {\bf j})(\bar{\bs }) \end{equation} To understand this, consider $$ \left( \frac{\partial f}{\partial u} \right)_{u=s} + \left( \frac{\partial f}{\partial v} \right)_{v=s} = \frac{\partial }{\partial s} \left(f(u,v)_{u=v=s}\right). $$ \bigskip \noindent Also, for any function $f$ of parameters $\bs $ and $G \ni b$, \begin{equation}\label{F.7} \int ^{t} d^{G\setminus b}s \, f \circ \bbox{\sigma} (G,\bs )_{s_{b}=t} = \int ^{t} d^{G/b}\bar{s} \, f \circ {\bf j} \circ \bar{\bbox{\sigma}}(G/b,\bar{\bs }) \end{equation} The key point here is that $\sigma _{c} (G,\bs )_{s_{b}=t}$ depends only on $\bar{c}$. \bigskip \noindent We return to (\ref{F.5}). By summing over $G$ with $G/b$ held fixed \begin{equation} X_{b} = \sum_{\bar{G}}\sum _{G:G/b=\bar{G}} \int^{t} d^{G\setminus b}s \, \partial ^{G} F \circ \bbox{\sigma}(G,\bs )_{s_{b}=t}. \end{equation} By (\ref{F.7}) this equals \begin{equation} \sum_{\bar{G}}\sum _{G:G/b=\bar{G}} \int^{t} d^{\bar{G}} \bar{s} \, \partial ^{G} F \circ {\bf j} \circ \bar{\bbox{\sigma}}(\bar{G},\bar{\bs } ) \end{equation} which by (\ref{F.6}), with $f = \partial _{b}F$ is \begin{equation} \sum_{\bar{G}} \int^{t} d^{\bar{G}} \bar{s} \, (\partial ^{\bar{G}} \partial _{b} F \circ {\bf j}) \circ \bar{\bbox{\sigma}}(\bar{G},\bar{\bs } ) \end{equation} which by the inductive hypothesis is \begin{equation} (\partial _{b}F )\circ {\bf j}(t\bar{\openone}) = \partial _{b}F(t\openone) \end{equation} Therefore the $t$ derivative of the right hand side of (\ref{C.4}) is $\sum _{b}\partial _{b}F(t\openone)$ which is the $t$ derivative of the left hand side of (\ref{C.4}). \hspace*{\fill} QED \medskip \section*{Acknowledgments} D. C. Brydges thanks the National Science Foundation for partial support under grants DMS-9401028 and DMS-9706166 and P. Federbush, T. Kennedy and G. Keller for collaborations on which some of this review is based. Ph. Martin acknowledges several informative discussions with A. Alastuey, F. Cornu, and B. Jancovici. \bibliographystyle{agsm}
\section{Introduction} The globular clusters (GCs) are among the oldest known objects in the Universe and can be considered as fossils of the formation of galaxies. GCs are simple coeval stellar systems which formed on a very short timescale during phases of intense star formation in galaxies. They are therefore more easily understood than mixed stellar field populations in a galaxy. The presence of globular cluster systems (GCSs) around most of the galaxies as well as the well known correlations found between some of their characteristics and those of the parent galaxies suggest that the formation of a galaxy and of its GCS are closely related (e.g. Harris \cite{harris}, Djorgovski \& Santiago \cite{djorgovski}). The alternative scenarios for the formation of early-type galaxies can be divided into two classes: the classical, long-lived scenario of a monolithic collapse which occurred in the early times of the Universe (e.g. Larson \cite{larson}) and the scenario issued from the hierarchical models of galaxy formation where the galaxies are formed through galaxy mergers (Kauffmann \cite{kauffmann}, Baugh et al. \cite{baugh}). During the unique collapse or for each major merging involving gas rich components, an intense star formation is expected with the likely formation of GCs. Indeed, it has been verified by the discovery of numerous proto-globular clusters by the Hubble Space Telescope in recently merging galaxies (and starbursting galaxies) in addition to the old ones from the progenitors (Schweizer \cite{schweizer} for a review): it demonstrates that GCs trace the major events in the star formation of the parent galaxy. The consequences of the latter scenario on the GCSs have been extensively investigated by Ashman and Zepf (1992). To date the case of the {\it brightest} elliptical galaxies located in rich clusters is by far the best studied. A simple monolithic formation of these galaxies seems now to be ruled out by the observation of several sub-populations of GCs around these objects, indicating a more complex formation (e.g. Geisler et al. \cite{geisler}, Forbes et al. \cite{forbes}). Traces of merging is very frequent in these systems and merging has probably played a role in the galaxy formation and/or evolution. Recent spectroscopic studies of GCs around NGC 1399 (Kissler-Patig et al. \cite{kissbrod}) and M 87 (Cohen et al. \cite{cohen}) show that they probably formed massively in the early phases of the formation of these galaxies with a small contribution of recent events. However the case of cluster ellipticals may well not be appropriate to disentangle the models of galaxy formation since the hierarchical models also predict that the cluster ellipticals formed at relatively high redshift (e.g. Kauffmann \cite{kauffmann}). In fact the case of field galaxies seems more interesting since the predictions of each model differ significantly for this class of objects. The hierarchical models predict their formation in recent dissipational mergers of gas-rich structures (e.g. Baugh et al. \cite{baugh}) which differ a lot from an early unique collapse. Early-type field galaxies with intermediate or low-luminosity are even more interesting. They might be the product of disk galaxies smaller than those involved in the formation of bright ellipticals (Kauffmann \& Charlot \cite{kauffcharl}). Indeed they generally exhibit a disky structure which is consistent with a formation in a dissipational merger of gas rich systems (Faber et al. \cite{faber}, Bender \cite{bender}) nevertheless the presence of gas during the merging is crucial to form new stars. Unfortunately, the low-luminosity isolated early-type galaxies are by far less studied than the bright cluster ellipticals. In general they show no obvious hint of a past merger event (Kissler-Patig \cite{kissler}). Until now, the few photometric studies of the GCSs around low-luminosity early-type galaxies are compatible with a single population and no complex formation (Kissler-Patig et al. \cite{kissfor}). But in most cases the low sensitivity of the colors used to estimate parameters like metallicity or age prevents a definitive conclusion. In this paper we present the study of the GCS of the isolated, low-luminosity S0 galaxy NGC 7457. It is a disky galaxy (Michard \& Marchal \cite{michard}) and its central light distribution is fitted by a power-law without evidence for the presence of a core (Lauer et al. \cite{lauer}). This galaxy does not show any trace of past interactions or merging (Schweizer \& Seitzer \cite{schweizseitz}). Nevertheless a steep central power-law in a disky galaxy can be the consequence of dissipative merging of gaseous galaxies (Bender \cite{bender}). Therefore the GCS around NGC 7457 may well have kept some traces of some merger events. It is in this context that we have undertaken the study of GCS around this galaxy. The main characteristics of NGC 7457 are presented in table 1. Deep B, V and I photometry will allow us to investigate the luminosity function of the GCS around this galaxy, to estimate the total number of GCs and the specific frequency and to study the color distribution of the GCs. \begin{table*} \caption[]{General data for the target galaxy, NGC 7457 from the LEDA database (http://www-obs.univ-lyon1.fr/leda/). The B and V magnitudes are the total magnitudes $\rm B_T$ and $\rm V_T$ from the LEDA database corrected by us for Galactic extinction with $\rm A_G(B)=4.1\times E(B-V)$ and $\rm A_G(V)=3.1\times E(B-V)$ and E(B-V)=0.0525 from Burstein and Heiles (\cite{burstein}). V is the heliocentric velocity. The distance modulus is calculated with the heliocentric velocity corrected for the Local Group infall onto the Virgo cluster and assuming $\rm H_0 = 75$km/s/Mpc (Paturel et al. \cite{paturel}).} \begin{flushleft} \begin{tabular}{lllllllllll} \hline Name & RA(2000) & DEC(2000) & {\sl l} & {\sl b} & Type & $\rm D_{25}$(arcmin) & V(km/s) & $\rm B$ & $\rm V$ & distance modulus \\ \hline NGC 7457 & 23 00 59.9 & 30 08 39 & 96.22 & -26.9& E-SO & 3.98 & 813 & 11.87 & 11.04 & 30.59\\ \hline \end{tabular} \end{flushleft} \end{table*} \section{Observations and data reduction} \subsection{Description of the data} The B, V and I images were obtained at CFHT with the MOS instrument. A summary of the exposures is given in table 2. The size of the field is $\rm 10\times 10~ arcmin^{2}$ with a pixel size of 0.44$\rm\arcsec$. The equivalent exposure times are 10800 seconds in B, 4800 seconds in V and 2400 seconds in I with a seeing of 1.25$\rm\arcsec$, 1.04$\rm\arcsec$ and 1.11 $\rm\arcsec$ respectively. \begin{table} \caption []{Summary of the observations of the globular cluster system of NGC 7457 at CFHT with the MOS instrument.} \begin{flushleft} \begin{tabular}{llll} \hline Date & Filter & Exposure time & Number of exposures\\ \hline 09/04/97 & B & 1200s& 5 \\ 09/04/97 & I & 300s & 3 \\ 09/04/97 & V & 600s& 4 \\ 09/05/97 & I & 300s & 4 \\ 09/05/97 & V & 600s& 4 \\ 09/05/97 & B & 1200s & 4 \\ \hline \end{tabular} \end{flushleft} \end{table} After having corrected all the images from bias and flat field, we combined the frames for each filter using a median algorithm to remove the cosmic rays and reduce the background noise. The galaxy profile has been next subtracted to perform a more reliable photometry near the central parts of the galaxy. The profile has been flattened by an iterative median-filter operation with a $\rm 15\times 15$ pixels window. This smoothed frame was subtracted from the original picture. Figure 1 is the final V picture on which photometry has been performed. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{chapfig1.ps}} \vspace{2cm} \caption[]{The globular cluster system around NGC 7457 in the V band. The galaxy profile has been subtracted. The 89 globular cluster candidates are marked with a circle. North is up and East on the left of the figure.} \end{figure} The detection has been performed in the V band since the observations are far deeper than in the I band and somewhat deeper than in the B band. This has been done using the DAOPHOT software in IRAF. 846 objects have been detected at a $\rm 6\sigma$ level above the local background. Such a high level of detection has been chosen since we wanted to keep only objects with accurate photometry measurements. A standard selection has been applied on the parameters "sharpness" and "roundness" given by the software. Finally we discarded some objects by visual inspection. The detection in the B and I frames was made by searching by position for objects matching the V detections. Photometry was done in the three bands with the PHOT software and only measurements with a photometric error lower than 0.2 mag have been considered. From the 846 objects detected in V, 844 have a V magnitude ($\rm <\sigma (V)>\ =0.04$ mag), 763 have been measured in I ($\rm <\sigma (I)>\ =0.03$ mag) and 698 in B ($\rm <\sigma (B)>\ =0.07$ mag). \subsection{The photometric zero-point} The calibration in the three bands was done using 2 fields (including 11 standard stars) with photometric standards (SA 110 and SA 113, Landolt \cite{landolt}). We determined the photometric zero-point by performing aperture photometry. The error in magnitude is estimated to be 0.02 mag in V, 0.05 mag in I and 0.05 mag in B. A Galactic color excess E(B-V)=0.0525 has been taken in the direction of NGC 7457 from Burstein \& Heiles (\cite{burstein}) and each magnitude has been corrected using the values $\rm A_B=0.22$, $\rm A_V=0.16$ and $\rm A_I = 0.10$ mag calculated using the Galactic reddening curve of Pei (\cite{pei}). \subsection{The completeness} The knowledge of the completeness is essential to derive a density profile and the total number of GCs around the galaxy. For each filter, we computed this completeness by adding artificial stars with the ADDSTAR routine of the DAOPHOT software. The magnitude was changed per bins of 0.25 mag or 0.5 mag. For each magnitude, 500 extra stars were added (over five runs in order to avoid crowding). The same procedure of detection as presented in section 2.1 has been applied. The 50\% completeness limit of a $\rm 6 \sigma$ detection is found to be 23.95 mag in V, 22.45 mag in I and 23.9 mag in B. The results are summarized in table 3. \begin{table} \caption[]{Completeness percentages for the B, V and I observations as a function of the magnitude of the point sources.} \begin{flushleft} \begin{tabular}{llll} \hline Magnitude & V band & I band & B band\\ \hline 21.0 & & 94.2 & \\ 21.5 & & 95.6 & \\ 22.0 & & 92.6 & \\ 22.25 & & 81.6 & \\ 22.5 & 95.2 & 34.0 & \\ 23.0 & 93.8 & & 94.8 \\ 23.5 & 92.2 & & 91.6 \\ 23.75 & 85.8 & & 78.6 \\ 24.0 & 37.4 & & 35.2 \\ \hline \end{tabular} \end{flushleft} \end{table} \section{The selection of globular clusters candidates} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{chapfig2.ps}} \caption[]{Color-color plots for the globular cluster candidates around NGC 7457. The point sources objects detected in the three photometric bands are reported with crosses. The typical internal photometric error as given by DAOPHOT is 0.3 mag for each color ($\rm 2\sigma$). The solid line is the envelope found by plotting synthesis models of stellar populations (Worthey \cite{worthey}) for a large range of metallicity (0.002 $\rm Z\odot$ to 2.5 $\rm Z\odot$) and age (1 Gyr to 18 Gyr). The objects whose colors are compatible with the models when accounting for the photometric error given for each measurement by DAOPHOT are marked with a filled circle. We keep only the objects selected in both color-color diagrams.} \end{figure} The GCs are selected on the basis of their expected range in magnitude and on their colors. Therefore, to perform such a selection, a good photometry in at least two bands is necessary. We start from the list of point-like objects detected in the V frame and for which the V and I magnitudes are measured with a photometric error lower than 0.2 mag (see section 2.1). We exclude the central region of the galaxy (galactocentric distance lower than 15 arcsec) as well as the very external part of the image (galactocentric distance larger than 226 arcsec) in order to avoid too large a background contamination (see section 4.1 and figure 3). We are left with 412 objects. First of all, we make a selection in magnitude, keeping only objects fainter than $\rm V>19$ mag. This corresponds to $\rm 3\sigma$ above the peak of the globular cluster luminosity function (GCLF) at the distance of NGC 7457 (distance modulus of 30.59, see table 1), for a standard GCLF with $\rm M_V=-7.4$ mag and $\rm \sigma (M_V)=1.2$ mag (Harris \cite{harris}). The selections in colors are made by comparing the V-I, B-V and B-I colors with the predictions of stellar population synthesis models. In figures 2a and b are plotted these colors together with the envelope of points generated from the models of Worthey (\cite{worthey}) for single stellar populations with a large range of parameters (a metallicity ranging from $\rm 0.002~ Z\odot$ to $\rm 2.5~Z\odot$ for ages comprised between 1 Gyr and 18 Gyr and three different initial mass functions (Salpeter (exponent x=-2.35 from $\rm M_l=0.21~ M\odot$ to $\rm M_u = 10~ M\odot$), Miller-Scalo (Miller \& Scalo, \cite{miller}, with $\rm M_l=0.1~ M\odot$ to $\rm M_u = 10~ M\odot$) and with an exponent x=-1.35 (from $\rm M_l=0.33~ M\odot$ to $\rm M_u = 10~ M\odot$)). We select all objects whose colors are consistent with the envelope of the models adopting a $\rm 2\sigma$ uncertainty on each color given by the DAOPHOT software. We use the models of Bertelli et al. (\cite{bertelli}): they lead to a similar selection of GC candidates. The resulting range of acceptable colors is $\rm 0.80<V-I<1.50,~ 1.3<B-I<2.7,~ 0.5<B-V<1.6$ but our selection is more refined than adopting simple color ranges since we check the compatibility of the three colors by their locus in color-color diagrams. For the objects which are not detected in B (5\% of the sample) a range of plausible values for their B magnitude is estimated from the models and their V-I colors. Then we reject the objects which must be detected in B (the limiting B magnitude is taken at 23.5 mag (90\% of completeness)). After these selections we are left with 89 GC candidates, out of which only 3 have no B measurement. \section{The number of globular clusters and the specific frequency} \subsection{The density profile} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{chapfig3.ps}} \caption[]{Upper panel: the surface density profile (open circles) of point-like objects detected around NGC 7457 in the V frame from the values quoted in table 4. The profile flattens at about 225 $\rm\arcsec$ to a constant level of $\rm 8.5\pm 0.9~ objects/arcmin^{2}$ (line). Lower panel: the surface density profile of the globular cluster system (open circles) compared with the isophotal galaxy light in V plotted in arbitrary units with filled squares.} \end{figure} We study the density profile of GCs using the {\it unselected} V detections down to V=23.9 mag, our 80\% completeness limit. To compute this density profile we bin the sample in elliptical rings having the same inclination and ellipticity as NGC 7457 ($\rm \epsilon = 0.41$ and $\rm p.a. = 130\degr$, LEDA database). The major axis step is about 25 $\rm\arcsec$ and the minor 15 $\rm\arcsec$. The counts are then corrected for magnitude incompleteness. No correction for geometrical incompleteness is necessary except for the last ring (major axis from 300 to 325 $\rm\arcsec$ from the center) for which we lack 10\% of the surface because of the limited size of the observed field. We do not consider the most central part of the galaxy which is saturated (for radii smaller than 15 $\rm\arcsec$). Inside the geometrical bins, we group the data in magnitude bins of 0.5 mag to correct for incompleteness at different magnitudes. The counts are shown in table 4. For the first ring, we could not estimate the Poissonian error because we are too close from the center of the galaxy. \begin{table*} \caption[]{The globular cluster counts around NGC 7457. The counts are performed in elliptical rings down to a magnitude $\rm V =23.9$ mag. In column 2 are reported the counts without any correction. In column 4 and 5 the counts are corrected for completeness, the quoted errors are Poissonian. In column 6 the density of GCs is corrected for the background contamination, the quoted error accounts for the error on the counts (column 4) and on the background ( $\rm 8.5\pm 0.9~ objects/arcmin^{2}$, see text).} \begin{flushleft} \begin{tabular}{llllll} \hline Ring major axis & Objects counts & Ring surface & Corrected counts & Surface density & Density profile\\ arcsec & & arcmin$^{2}$ & & objects/armin$^{2}$ & GCs/armin$^{2}$\\ \hline 15-25 & 8 & 0.21 & 8.5 $\pm$ ? & 41.3 $\pm$ ? & 32.7 $\pm$ ?\\ 25-50 & 34 & 0.97 & 37.2 $\pm$ 6.1 & 38.5 $\pm$ 6.3 & 30.0 $\pm$ 6.4\\ 50-75 & 44 & 1.61 & 47.9 $\pm$ 6.9 & 29.8 $\pm$ 4.3 & 21.2 $\pm$ 4.4\\ 75-100 & 37 & 2.25 & 40.1 $\pm$ 6.3 & 17.8 $\pm$ 2.8 & 9.3 $\pm$ 3.0\\ 100-125 & 34 & 2.90 & 36.4 $\pm$ 6.0 & 12.6 $\pm$ 2.1 & 4.0 $\pm$ 2.3\\ 125-150 & 36 & 3.54 & 38.6 $\pm$ 6.2 & 10.9 $\pm$ 1.8 & 2.4 $\pm$ 2.0\\ 150-175 & 54 & 4.18 & 58.6 $\pm$ 7.7 & 14.0 $\pm$ 1.8 & 5.5 $\pm$ 2.0\\ 175-200 & 46 & 4.83 & 50.0 $\pm$ 7.1 & 10.4 $\pm$ 1.5 & 1.8 $\pm$ 1.7\\ 200-225 & 48 & 5.47 & 51.3 $\pm$ 7.2 & 9.4 $\pm$ 1.3 & 0.8$\pm$ 1.6\\ 225-250 & 41 & 6.11 & 44.3 $\pm$ 6.7 & 7.2 $\pm$ 1.1 & -1.3 $\pm$ 1.4\\ 250-275 & 60 & 6.76 & 64.3 $\pm$ 8.0 & 9.5 $\pm$ 1.2 & 1.0 $\pm$ 1.5\\ 275-300 & 56 & 7.40 & 59.7 $\pm$ 7.2 & 8.1 $\pm$ 1.0 & -0.5$\pm$ 1.4\\ 300-325 & 58 & 7.27 & 62.1 $\pm$ 7.9 & 8.5 $\pm$ 1.2 & 0.0$\pm$ 1.4\\ \hline \end{tabular} \end{flushleft} \end{table*} In figure 3 are plotted the radial variation of the surface density, distribution for our point-like V detections together with the galactic V light profile. The background contamination can been estimated when the density profile reaches a constant value. Practically this occurs for galactocentric distances larger than 200 arcsec. The average of the counts in the five last rings (table 4, column 5) gives $\rm 8.5\pm 0.9~ objects/arcmin^{2}$ in agreement with the deep counts down to $\rm V = 23.9$ mag, our 80\% limit completeness (Peterson et al. \cite{peterson}, Tyson \cite{tyson}). After applying corrections for incompleteness, the total number of objects down to this magnitude is estimated to be $\rm 537\pm 38$. The background is subtracted statistically over an area of $\rm 46.34~arcmin^{2}$ (thus $\rm 394\pm 46$ objects) and we are left with a total of $\rm 143\pm 60$ GC candidates. The density profile of the GCs follows the galaxy light profile. The fit of a power-law on the profiles ($\rm \propto r^{-\alpha}$) gives $\rm \alpha=2.10\pm 0.26$ for the GC density profile and $\rm \alpha=2.02\pm 0.23$ for the galaxy light profile. We used all the data points except for the first bin for which we do not estimate the error. Such a slope ($\rm \alpha\simeq -2$) for the GC profile of NGC 7457 whose absolute V magnitude is -19.55 mag is in agreement with the correlation found between this slope and the absolute magnitude of the parent galaxy, fainter galaxies having a steeper distribution of GCs (e.g. Harris \cite{harris}, Kissler-Patig \cite{kissler}). A similar profile for the galaxy light and the GCS seems not to be the rule for bright ellipticals (e.g. Ashman \& Zepf \cite{ashzepf}) but the case of low-luminosity early-type galaxies is less clear since very few data are available for them. \subsection{The globular cluster luminosity function} \begin{figure} \vspace{2cm} \resizebox{\hsize}{!}{\includegraphics{chapfig4.ps}} \caption{The globular cluster luminosity distribution in the V band fitted by a gaussian function with mean and $\rm \sigma$ equal respectively to 23.1 mag and 0.9 mag.} \end{figure} To compute the globular cluster luminosity function we would need to consider only genuine GCs. To this aim, we use the GC candidates selected in section 3 from V and I detections. We have to account for incompleteness effects due to the magnitude limits in V and I. The range of V-I colors for the GC candidates is estimated from the models of Worthey namely $\rm 0.80<V-I<1.5$. Since the GC candidates are first detected in the V band, we apply a selection which ensures the full range of possible V-I color for each detected candidate. With a completeness limit (50\%) at 22.45 mag for I and 23.95 mag for V we are limited by the I detections and we have to truncate the sample at V=23.25 mag which is reduced to 65 GC candidates. The GCLF is shown for V$<$ 23.25 mag in figure 4. A gaussian distribution has been fitted to the bins not affected by incompleteness. The parameters of the gaussian are $\rm M_V=23.1\pm 0.2$ mag for the mean and a dispersion of $\rm \sigma_V=0.9\pm 0.3$ mag. The mean value that we find is fully consistent with a mean absolute magnitude $\rm M_V=-7.4$ mag for the distance modulus of NGC 7457 (table 1). The dispersion we find for the GCLF of NGC 7457 is marginally consistent with those found in galaxies like M 31 or the Milky Way with $\rm \sigma_V=1.2\pm 0.1$ mag whereas for bright ellipticals the dispersion seems larger with $\rm \sigma_V \simeq 1.4$ mag (e.g. Ashman \& Zepf \cite{ashzepf}). \subsection{The specific frequency} The number of GCs in an early-type galaxy has been found to scale with the parent galaxy luminosity (e.g Djorgovski \& Santiago 1992). Therefore following Harris \& van den Bergh (1981) the GC counts are normalized per absolute visual magnitude of $\rm M_V = 15$ mag and a specific frequency is defined as $\rm S_N = N\times 10^{0.4\cdot (M_V~+~15)}$ where N is the total number of GCs in the galaxy and $\rm M_V$ its absolute visual magnitude. $\rm S_N$ is a measure of the GC efficiency. $\rm S_N$ is found to be $\rm \sim 5$ for most early-type galaxies with $\rm -19 > M_V > -22$ although with a large dispersion. Brighter early-type galaxies exhibit in average a larger $\rm S_N$, some giant ellipticals in rich clusters having a very large specific frequency $\rm S_N>10 $ (Kissler-Patig \cite{kissler}). In the section 4.1 we estimated the total number of GCs down to $\rm M_V = 23.9$ mag. Then integrating over the entire luminosity distribution of the section 4.2 we deduce a total number of $\rm 178\pm 75$ clusters. This translates to a specific frequency $\rm S_N = 2.7\pm 1.1$ assuming an absolute V magnitude $\rm M_V =-19.55$ mag for NGC 7457. We can compute in the same way the specific frequency from the number of detected GC down to V = 23.1 mag (the turn-over of the GCLF) and, after multiplying by two, we are left with $132 \pm 58$ objects thus $\rm S_N = 2.0\pm 0.9$. \section{The color distributions of the globular clusters} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{chapfig5.ps}} \caption[]{B-I color versus metallicity of the MW globular clusters (open circles) together with the the results of Worthey's models for old homogeneous stellar populations (crosses). The solid line is the result of the linear regression.} \end{figure} Now we study the V-I and B-I color distributions of the GCs around NGC 7457 in order to search, if any, the presence of sub-populations with different metallicities. Indeed the B-I color and in a lesser extent the V-I color are sensitive to the metallicity provided that the GCs are old ($> 5$ Gyr). In case of old stellar populations these two colors are poor tracers of the age of the stellar populations (Worthey \cite{worthey}, Kissler-Patig et al. \cite{kissbrod}). First we need to calibrate the V-I and B-I colors as a function of the metallicity. Empirical relations based on the observation of the Milky Way (MW) GCs are commonly used (e.g. Couture et al. \cite{couture}) but these relations become very insecure outside the metallicity range of the MW GCs ($\rm -2.0\le [Fe/H] \le -0.5$). Kissler-Patig et al. (1998) have extended the comparison between V-I and the metallicity deduced from spectroscopic data to higher metallicities thanks to their observations of NGC 1399 and found a flatter relation of $\rm [Fe/H]$ as a function of V-I than when only the MW GCs are considered. In this paper, we will adopt the relation of Kissler-Patig et al. (\cite{kissbrod}): $$\rm [Fe/H] = (-4.50\pm 0.30) + (3.27\pm 0.32) (V-I)$$ For the relation between the B-I color and the metallicity of GCs we re-analyse the problem by using the data on the MW clusters (Harris 1996, McMaster database) together with the synthesis models of Worthey (1994) for old stellar populations ($> 5$ Gyr) with a large range of metallicities ($\rm -2.0\le [Fe/H] \le 0.5$). We consider only old ages for the models in order to be consistent with the age of the MW GCs. Therefore, the calibration relation will be valid for old stellar populations only ($> 5$ Gyr). When comparing the colors listed in Worthey (1994) with the observed MW GCs, a small shift in color is needed. Assuming an average age of 12 Gyr for the MW GCs, we get a slight correction (-0.09 in V-I, -0.14 in B-I) to apply to the models. These corrections are in reasonable agreement with the color shifts suggested by Worthey. The data are presented in figure 5. A linear regression gives: $$\rm [Fe/H] = (-5.34\pm 0.42) + (2.44 \pm 0.30) (B-I)$$ Given the uncertainties about these calibration relations, we have checked that our results do not depend on the exact choice of the conversion formulae. Indeed we have used different relations like those of Couture et al. for both V-I and B-I as well as the relation we find between [Fe/H] and V-I following the same method as described in the last paragraph (MW data and synthesis models). \subsection{The V-I color distribution} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{chapfig6.ps}} \caption[]{The histogram of the V-I color distribution for the globular clusters of NGC 7457. The typical internal photometric error is 0.10 mag as the size of the bins.} \end{figure} The V-I color histogram is reported in figure 6 for the sample of 65 GCs with $\rm V<23.25$ mag (see section 4.2). The distribution is found unimodal using the KMM mixture-modeling algorithm (Ashman et al. \cite{ashbird}). The gaussian distribution fitted on the data has a mean $\rm <V-I>\ =1.04\pm 0.02$ mag and a standard deviation $\rm \sigma(V-I)=0.15\pm 0.04$ mag. We can also estimate the mean and standard deviation using maximum likelihood estimators (Pryor \& Meylan \cite{pryor}), this method has the advantage of being performed without any hypothesis on the shape of the distribution. We find very similar results i.e. $\rm <V-I>=1.06 \pm 0.02$ mag and $\rm \sigma(V-I)=\ 0.16\pm 0.02$ mag. Therefore the observed standard deviation is only slightly larger than the mean internal photometric error for the sample ($\rm \sigma_{err}(V-I)=0.1$ mag).\\ A mean $\rm <V-I>\ =1.04$ mag translates to $\rm <[Fe/H]>\ =-1.1$ dex. With an absolute V magnitude of -19.55 for NGC 7457 such a mean is consistent with the correlation found between the absolute magnitude of the parent galaxy and the mean metallicity of the GCSs (e.g. Ashman \& Zepf \cite{ashzepf}). \subsection{The B-I color distribution} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{chapfig7.ps}} \caption[]{The histogram of the B-I color distribution for the globular clusters of NGC 7457. The typical internal photometric error is 0.1 mag of the order of the size of the bins (0.1 mag).} \end{figure} To compute the B-I color distribution we have to select GC candidates detected in V and with a measured magnitude in I and B. Given the deepness of the V detection as compared to the I and B ones, the resulted sample is equivalent to a selection based only on B and I detections. Once again we only keep objects with $\rm V>19$ mag. We have also to be as complete as possible in the B and I bands for the analysis of the B-I color distribution. The range of B-I colors is estimated from the models of Worthey used for the selection of the GCs candidates i.e. $\rm 1.3<B-I<2.7$. As before, we adopt a completeness limit at 50\% which corresponds to 22.45 mag in I and 23.9 mag in B. To ensure the full range of possible colors for each candidate detected in B we must truncate the sample at B=23.7 mag. The sample is reduced to 41 objects. The B-I color histogram is reported in figure 7. The distribution is larger than for the V-I color but it is also found unimodal with the KMM test. Such a result is expected since Ashman et al. (\cite{ashbird}) have shown that a bimodality cannot be easily detected with the KMM algorithm when the number of data is lower than 50 as it is the case here. The fit with a gaussian distribution gives a mean $\rm <B-I>\ =1.87\pm 0.02$ mag leading to a $\rm [Fe/H]=-0.8$ dex consistent with the mean metallicity deduced from the V-I distribution given the uncertainties on the calibration relations; the observed standard deviation is $\rm \sigma(B-I)=0.25\pm 0.06$ mag. Using maximum likelihood estimators we find $\rm <B-I>\ =1.91\pm 0.04$ mag and $\rm \sigma(B-I)=0.28\pm 0.03$ mag. Therefore the observed standard deviation of the B-I distribution is larger than the mean internal error $\rm \sigma_{err}(B-I)=0.1$ mag. \subsection{A hidden bimodality?} The V-I and B-I distributions of the GCs around NGC 7457 are found unimodal. Before discussing the implications of this result we discuss in what conditions a bimodality would be really detected. First, as already discussed it is imperative to detect more GCs and therefore to perform a deeper photometry in at least two bands. Moreover Kissler-Patig et al. (\cite{kissfor}) have outlined that, even if present, different populations of GCs are probably hard to discriminate in low-luminosity galaxies. A conspiracy between age and metallicity of the two populations could mimic a unimodal V-I distribution, the V-I color being very sensitive neither to age nor to metallicity. Another difficulty is that even if the metallicity distribution of the GCS is bimodal the two peaks might be too close to be disentangled. For example Forbes et al. (\cite{forbes}) have found a correlation between the mean metallicity of the metal-rich population of GCs and the luminosity of bright ellipticals whereas there is almost no correlation between the metallicity of the metal-poor GCs and the parent galaxy luminosity. If such a correlation also holds for fainter galaxies, their redder GC population will be less metal-rich and its color peak closer to that of their metal-poor GC population than it is the case in the bright ellipticals. Therefore, the two peaks in the metallicity distribution, if they exist, will be hardly detected in these galaxies. In order to investigate more quantitatively this effect, we can try to estimate what sort of bimodality is detectable with our data. Ashman et al. (\cite{ashbird}) have tested the bimodal detectability as a function of $\rm \Delta \mu$, the difference between the two means divided by the standard deviation of the two sub-populations supposed to have the same dispersion. For a total number of $ \rm \sim 50$ GCs, a bimodality is detected as soon as $\rm \Delta \mu \ge 3$. If we adopt a typical $\rm \sigma ([Fe/H])=0.3$ dex for each sub-population we find $\rm \Delta [Fe/H]\ge 0.9$ dex. However, we must note that this estimation is based on the assumption that the two sub-populations have the same number of objects. As noted by Kissler-Patig et al. (\cite{kissfor}, see also section 6.2), if there is not an equal number of GCs in each sub-population, they cannot be separated when $\rm \Delta [Fe/H]\sim 1$ dex. $\rm \Delta [Fe/H]=0.9$ dex corresponds to $\rm \Delta (V-I)= 0.3$ and $\rm \Delta (B-I)= 0.4$ mag according to the calibration formulae given above. Thus, under these hypotheses, the two peaks might be separated. A 1-dex separation seems to a rough order of magnitude of what we may find in GCSs. Indeed, in bright ellipticals the difference between the two metallicity peaks approximately or slightly larger than 1 dex (e.g. Ashman \& Zepf \cite{ashzepf}). Also, in the Milky Way, a similar gap has been measured between the metallicity of the halo and disk/bulge GCs (Armandroff \& Zinn \cite{armandroff}). The bottomline is that with our data we can hope to separate two peaks of a bimodal distribution similar to those found around bright ellipticals or the Milky Way provided that the two sub-populations are roughly equally represented. Any bimodal distribution with two metallicity peaks separated by less than 1 dex could not be separated. \section{Discussion} NGC 7457 is a S0 low-luminosity field galaxy. S0 galaxies are thought to be the continuation from disky ellipticals to early-type spirals in the Hubble sequence (Kormendy \cite{kormendy}). We can expect that S0 and disky ellipticals have been formed through similar processes. Therefore this type of galaxy is particularly interesting for testing the scenarii of galaxy formation, in particular whether these galaxies formed in a dissipational merger event (see section 1). The simple model of a spiral--spiral merger predicts several properties for the GCS of the remnant galaxy (Ashman \& Zepf \cite{ashman}): the newly produced GCs should rise the specific frequency by a factor of two, the new clusters and stars will be more concentrated than the GCS of the progenitors, and the new clusters should appear as a second population in the color distribution. At face value NGC 7457 does not seem to confirm the hypothesis of a formation by spiral--spiral merging, at least when compared to the predictions of the simple model of Ashman \& Zepf (\cite{ashman}). NGC 7457 is found to have a rather low number of GCs steeply distributed around the galaxy, and the color distribution of these GCs appears unimodal. Indeed, the specific frequency $\rm S_N=2.7\pm 1.1$ is compatible with the range observed in the four studied Sa and Sab galaxies (ranging from 0.7 to 3.5 with a mean of 2.0 and a dispersion of 1.0, taken from Ashman \& Zepf \cite{ashzepf}). The surface density profile of the GCs follows the one of the stars (see section 4.1). The absence of a recent merging event is suggested by the lack of any fine structure as defined by Schweizer \& Seitzer (\cite{schweizer}) for this galaxy. However two results might constrain formation theories. First the color distributions of the GCs appear broader than for a single population such as, e.g., the Milky Way halo clusters. No clear bimodality has been detected. Nevertheless given the poor statistics it has been shown in section 5.3 that it does not exclude necessarily the presence of distinct sub-populations. And, second, the peak of the color distribution appears relatively red. \subsection{A broad color distribution} Does the unimodality of the color distribution imply the absence of two distinct GC populations? In spite of the difficulties discussed in the section 5.3, the narrow V-I distributions found in low-luminosity early-type galaxies (Kissler-Patig et al. \cite{kisskohl}), together with a narrow luminosity function, can be used to exclude large difference in both metallicity and age within each galaxy, and therefore a recent gas-rich merger (z$<$1) (Kissler-Patig et al. \cite{kissfor}). The availability of the B-I colors for the GCs of NGC 7457 allows us to go further in the analysis. Indeed the B-I color is roughly twice as sensitive to metallicity than the V-I color (Couture et al. \cite{couture}). But an intrinsic difficulty with low-luminosity galaxies is their small number of GCs. Bimodality was shown to be undetectable in a dataset containing less than 50 objects (see section 5.3). Nevertheless, as we will see below, we can expect to observe some differences between the widths of a unimodal and a bimodal distributions. The dispersion in color of a ``single'' population of GCs can be estimated from the halo GCs of the MW. On the one hand, we can convert the metallicity dispersion into a color dispersion. With $\rm \sigma ([Fe/H])=0.3$ dex (Armandroff \& Zinn \cite{armandroff}) we expect a $\rm \sigma (V-I) \sim 0.05$ mag and $\rm \sigma (B-I)=0.1$ mag from the calibration relations of section 5. On the other hand, we can measure the dispersion from the V-I and B-I data of the $\sim 80$ halo GCs in the McMaster catalog (Harris 1996), and obtain $\rm \sigma (V-I)=0.05\pm 0.01$ mag and $\rm \sigma (B-I)=0.09\pm 0.01$ mag in excellent agreement with the first values. In V-I, the genuine dispersion in metallicity is extremely difficult to derive from the V-I colors, the expected dispersion being lower than the typical photometric errors of 0.1 mag. In B-I, however, typical photometric errors and intrinsic dispersion of a single population are comparable, so that several populations would broaden the color distribution to a detectable level. In NGC 7457 the dispersion in V-I is only slightly larger than the internal error (0.15 mag against 0.10 mag) and we estimate that the true dispersion is $\rm \sim 0.11\pm 0.11$ mag according to the relation $\rm \sigma^2_{obs} = \sigma^2_{err}+\sigma^2_{true}$. This standard deviation in V-I translates to a $\rm \sigma([Fe/H])\sim 0.4\pm 0.4$ dex when using the calibration relation given in the precedent section. The error is estimated by accounting for the uncertainties on the determination of the standard deviations and on the calibration formula given in section 5. The estimate is very insecure due to the large error on $\rm \sigma(V-I)$.\\ For the B-I color we find a dispersion (0.25 mag) clearly broader than the combination of a single population and photometric errors. The above relation leads to $\rm \sigma_{true}=0.23 \pm 0.07$ mag or $\rm \sigma([Fe/H])=0.6\pm 0.2$ dex. This value is compatible with the value tentatively deduced from the V-I distribution. Such a dispersion in metallicity seems intermediate between the one of a single population and the total dispersion of the GC populations in bright ellipticals. Indeed single populations such as the Galactic halo GCs (Armandroff \& Zinn \cite{armandroff}), or the GCs around M 81 or M 31 (Perelmuter \& Racine \cite{perelmuter}) have $\rm \sigma ([Fe/H])\sim 0.3$ dex; the individual components of the bimodal distribution in NGC 4472 have similar dispersions of $\sim 0.38$ dex in $\rm [Fe/H]$ (Geisler et al. \cite{geisler}). In contrast, the total dispersion of the system in M 87 is $\rm \sigma([Fe/H])=0.65$ dex (Lee \& Geisler \cite{lee}), in NGC 4472 $\rm \sigma([Fe/H])=0.7$ dex (Geisler et al. \cite{geisler}). Therefore, while the distribution in metallicity of the GCs around NGC 7457 is found to be unimodal, the width of the GC metallicity distribution is compatible with the presence of different populations probably less separated in metallicity than in the giant clusters ellipticals. This suggests a significantly different chemical enrichment of the GCs in NGC 7457 than, e.g., the halo population of the Galaxy. \subsection{The mean colors of the globular clusters: implications on the scenarios of formation} With $\rm M_V=-19.55$ and a mean metallicity of [Fe/H]$\ \simeq -1$ dex for its GCs, NGC 7457 follows the general trend found between the absolute magnitude of the galaxies (spirals $+$ ellipticals) and the [Fe/H] value of their GCs (e.g. Brodie \& Huchra \cite{brodie}, Ashman \& Zepf \cite{ashzepf}). Nevertheless, the spirals seem to have a lower GC metallicity as compared to ellipticals of similar luminosity and the mean metallicity of [Fe/H]$\ \simeq -1$ dex for the GCs around NGC 7457 is consistent with the mean values found for the metallicities of the GCs around the bright elliptical galaxies ($\rm M_V\le -20$, e.g. Ashman \& Zepf \cite{ashzepf}). We can also compare more quantitatively the color distribution of the GCs around NGC 7457 with that of the Galactic GCs. In addition to its broad dispersion, the mean B-I color found for the GCs of NGC 7457 (B-I$\ \simeq1.9$ mag, see section 5.2) is comparable to the mean of the Galactic disc/bulge GCs (B-I$\ \simeq 1.9$ mag, as derived from the McMaster catalog).\\ Moreover, Monte Carlo simulations of B-I color distributions similar to ours show that any metal-poor (B-I=1.55 mag, the mean color of the metal-poor clusters in the Galaxy) population as large as 20\% to 30\% of the metal-rich (B-I$=1.92$ mag) one would be detected. Therefore we can conclude to the absence of any significant population of metal-poor clusters similar to that of the MW halo. It is likely that such blue globular clusters were never present in NGC 7457 since NGC 7457 is an isolated galaxy ($\rm \rho=0.13~~ galaxies/Mpc^{3}$) and shows no signs of any interaction. Thus the loss of blue GCs loss through stripping seems excluded.\\ The formation in a spiral--spiral merger would imply the presence of blue GCs from the progenitor spirals unless the latter did not host blue GCs like the MW. In situ formation models usually explain blue GCs as formed in the early stage of the galaxy. To fit the absence of blue clusters in such scenarios an early epoch of star formation (to enrich the gas) without any formation of GCs would be required.\\ We can also explore the possibility that the halo population of NGC 7457 and therefore of its progenitors in case of merging would be slightly redder than the MW one. For both the merger and accretion models, no variation of the color of the blue cluster population as a function of the parent galaxy luminosity is expected as it seems to be confirmed by Forbes et al (\cite{forbes}). Indeed their mean metal-poor peak $\rm <[Fe/H]>\ = -1.2\pm 0.3 ~$ dex obtained for bright ellipticals translates to $\rm B-I=1.7$ and V-I = 1.0 (see also Neilsen et al. \cite{neilsen}). An extrapolation of the trend that Forbes et al. found between the metal-rich peak and the galaxy luminosity gives $\rm [Fe/H]\simeq -0.6 $ dex and thus $\rm B-I\simeq 1.95$, $\rm V-I\simeq 1.2 $. Such a bimodal distribution is consistent with our data since the two peaks could not be separated neither in V-I nor in B-I given the proximity of their colors (see section 5.3).\\ Finally, the absence of blue GCs around NGC 7457 could be explained in a scenario as the one advanced by C\^ot\'e et al. (\cite{cote}) where a galaxy forms GCs with a mean metallicity proportional to its luminosity and gains its metal-poor GCs by accreting smaller galaxies surrounding it. In this case, the absence of a significant number of metal-poor GCs in NGC 7457 could then be due to a lack of dwarf galaxies around this isolated galaxy. Two galaxies already were reported to lack blue GCs: NGC 3923 and NGC 3311. Zepf et al. (\cite{zepf}) have discussed various scenarios to explain the high mean metallicity ([Fe/H] = -0.56 dex) and the very few clusters with [Fe/H] $<$ -1 dex around NGC 3923, a luminous field elliptical. They propose several explanations around the merger picture, in which the number and metallicity of the clusters formed during the merging as well as those of pre-existing clusters around the progenitors might vary. Finally more or less accretion of satellite galaxies and their metal-poor GCs can also modify the final color distribution of the system.\\ In case of NGC 3311, the central cD in the Hydra cluster, Secker et al. (\cite{secker}) attribute the almost complete lack of blue GCs (less than 10\% of the GCs appear to have $\rm [Fe/H]<-1$ dex) to different history of pre-enrichment of the giant galaxy in the frame of an in situ formation of clusters. \section{Conclusions} We have studied the GCS around the low-luminosity ($\rm M_V=-19.55$ mag) early-type galaxy NGC 7457. Several characteristics of the GCS are found typical for this type of galaxy: the specific frequency is found low with $\rm S_N \sim 2.7$ and the GC density profile is steep and follows the stellar light of the parent galaxy. The mean metallicity of $\rm <[Fe/H]>\ \simeq -1$ dex is compatible with the correlation existing between the metallicity of the GCSs and the luminosity of the parent galaxies. The B-I and V-I color distributions are found unimodal but the B-I distribution is much wider than expected from a typical homogeneous population of GCs and consistent with what is expected for a bimodal distribution. The poor statistics (66 globular clusters in V-I and 41 in B-I) prevents from a detection of two peaks in the metallicity distribution of the GC system if they are closer than 1 dex in [Fe/H] and/or if the two sub-populations have a very different number of objects. Therefore our data are consistent with the presence of more than one population of GCs around NGC 7457. An alternative scenario would be a single GC population with a larger range in metallicity than found for homogeneous populations of GCs around the MW, nearby galaxies or bright ellipticals. No significant population of metal-poor GCs similar to the halo GCs of the Galaxy is detected around NGC 7457. Several possibilities are discussed to explain this absence of GCs with $\rm [Fe/H]\simeq -1.6 $ dex. Deeper photometry would be necessary to detect more GCs and to be able to identify two peaks in the metallicity distribution of the GCS. Another test of the hierarchical models of galaxy formation will be to estimate the age of the GCs since these models predict a rather recent merging for field early-type galaxies with $\rm z<1$. Such a determination requires spectroscopic observations of the GCs but will be difficult if the merging occurred more than 5 Gyr ago since spectroscopic indices as $\rm H_{\beta}$ are not sensitive to higher ages. \begin{acknowledgements} MK-P thanks the Alexander von Humboldt Foundation for its support through a Feodor Lynen Fellowship. \end{acknowledgements}
\section{Introduction} The dynamics of fluids in the expanding universe is of great importance in the cosmology. For investigating such a dynamics, the Newtonian treatment is often used as a good approximation for the region $l/L \ll 1$, where $l$ is the scale of fluctuations of fluids and $L$ corresponds to the Hubble radius \cite{Peebles}. In the Newtonian cosmology, the Lagrangian perturbation theory in the dust universe has been developed for non-linear density fluctuations, up to the caustic formation \cite{Zeldovich,Buchert,BE,SK}. In order to take into account the relativistic correction to the Newtonian dynamics, the cosmological post-Newtonian approximation has been formulated \cite{Futamase,SA}. Furthermore, in the cosmological post-Newtonian approximation, the Lagrangian perturbation theory has been also discussed \cite{MT,TF}. Such a treatment beyond the Newtonian approximation may become important because of not only the theoretical interest but also the recent progress in the observational cosmology. For example, the Sloan Digital Sky Survey (SDSS) aims at the deep survey over several hundred Mpc \cite{SDSS}, in which it is not clear whether the Newtonian treatment is sufficient for such a large area. Relativistic theories of linear perturbations have been developed \cite{Peebles,Bardeen,KS}. The second-order extension of relativistic perturbation theory has been also developed \cite{Tomita}. However, they still depend on the assumption that the density fluctuation is small. In order to overcome the drawback, several perturbative approaches for the nonlinear dynamics have been proposed, which are based on the fluid flow approach \cite{MPS}, the gradient expansion method \cite{PSS,CPSS,SSC}, or the relativistic Lagrangian approach \cite{Kasai,RMKB}. It is shown that the relativistic post-Zel'dovich approximation, a second-order extension of the relativistic Lagrangian approach, successfully express the non-linear evolution of the density contrast with higher accuracy than the conventional second-order theory \cite{MNK}. So far, however, these relativistic approaches are all restricted within the irrotational dust. There has been no established way of calculating non-linear evolution of the density fluctuation without such limitation. Our paper is aimed to add to knowledge in literature by presenting a relativistic framework of such calculation. This paper is organized as follows. In section II, we show the integrals of the density and the vorticity in the general relativity, in the simple manner using the Lagrangian condition. This fact suggests strongly that the Lagrangian condition allows us to formulate the Lagrangian description in the relativistic cosmology. Under this condition, section III presents a perturbative Lagrangian approach. Summary and Discussion are given in section VI. For comparison, the vorticity in the Newtonian cosmology is discussed in the appendix A. Residual gauge freedoms in the Lagrangian condition are clarified in the appendix B. Greek indices run from $0$ to $3$, and Latin indices from $1$ to $3$. We use the unit, $c = 1$. \section{Lagrangian description of dust fluid} Let us consider a simple dust universe, in which the energy momentum tensor is written as \begin{equation} T^{\mu\nu}=\rho u^{\mu} u^{\nu} . \end{equation} The conservation law $T^{\mu\nu}_{\ \ ;\nu}=0$ gives \begin{eqnarray} \rho_{;\mu} u^{\mu} + \rho u^{\mu}_{\ ;\mu} &=& 0 , \label{continuity} \\ u^{\mu}{}_{;\nu}u^{\nu} &=& 0 , \label{geodesics} \end{eqnarray} which are called as the continuity equation and the geodesic equation, respectively. The vorticity $\omega^{\mu}$ of the fluid flow is defined by \cite{Ehlers} \begin{equation} \omega^{\alpha}\equiv {1 \over 2} \epsilon^{\alpha\mu\nu\rho} u_{\mu}u_{\nu;\rho} , \label{vorticity} \end{equation} where $\epsilon^{\alpha\mu\nu\rho}$ denotes the complete anti-symmetric tensor with $\epsilon^{0123} = 1/\sqrt{- g}$ and $g\equiv\det(g_{\mu\nu})$. From the geodesic equation Eq. (\ref{geodesics}), we obtain the propagation equation for the vorticity \cite{Ehlers}: \begin{equation} \omega^{\mu}{}_{;\nu} u^{\nu}+u^{\nu}{}_{;\nu} \omega^{\mu} = u^{\mu}{}_{;\nu} \omega^{\nu} . \label{eqomegaprop} \end{equation} Using Eq. ($\ref{continuity}$), we have \begin{eqnarray} \Bigl( {\omega^{\mu} \over \rho} \Bigr){}_{;\nu} u^{\nu} =u^{\mu}{}_{;\nu} \Bigl( {\omega^{\nu} \over \rho} \Bigr) , \label{beltrami} \end{eqnarray} which may be called as the relativistic Beltrami equation (cf. Eq. (\ref{eqBeltrami}) in Appendix A). The Einstein equations are decomposed with respect to the fluid flow as follows: \begin{eqnarray} G_{\mu\nu}u^{\mu}u^{\nu}&=&8\pi G \rho , \label{H} \\ G_{\mu\nu}u^{\mu}P^{\nu}_{\ \alpha}&=&0 , \label{M} \\ G_{\mu\nu}P^{\mu}_{\ \alpha}P^{\nu}_{\ \beta}&=&0 , \label{E} \end{eqnarray} where $P^{\mu}_{\ \nu}$ is the projection tensor \begin{equation} P^{\mu}_{\ \nu}\equiv \delta^{\mu}_{\ \nu}+u^{\mu}u_{\nu} . \end{equation} So far the treatment is fully covariant. In the following, we adopt the Lagrangian condition (e.g. \cite{Friedrich}), in which the components of the matter 4-velocity take the values of \begin{equation} u^{\mu}=(1,0,0,0) . \label{lagrangecon} \end{equation} Under this condition, we immediately have $g_{00}=-1$ and $u_{\mu} = (-1, g_{0i})$. Furthermore, since the geodesic equation Eq. (\ref{geodesics}) under the Lagrangian condition simply tells $u_{\mu, 0} = 0$, $u_i (= g_{0i})$ are functions of spatial coordinates only: \begin{equation} u_i = u_i(\bbox{x}), \quad \mbox{and}\quad g_{0i} = g_{0i}(\bbox{x}). \end{equation} In the Lagrangian description, the continuity equation (\ref{continuity}) is simply \begin{equation} (\rho \sqrt{-g})_{,0} = 0 . \label{continuity2} \end{equation} Therefore, \begin{equation} \rho(\bbox{x},t) = \sqrt{ \frac{g(\bbox{x},t_0)}{g(\bbox{x},t)} } \rho(\bbox{x},t_0) . \label{masscon} \end{equation} Under the Lagrangian condition, the determinant of the metric tensor can be expressed as \begin{equation} g = -(1+\gamma^{ij}g_{0i}g_{0j}) \det (g_{ij}) , \end{equation} where $\gamma^{ij}$ is the inverse of the spatial metric $g_{ij}$. The relativistic Beltrami equation (\ref{beltrami}) also becomes simply \begin{equation} \Bigl( {\omega^{\mu} \over \rho} \Bigr){}_{,0}=0 , \end{equation} which is integrated to give \begin{equation} \frac{\omega^{i}}{\rho} = \left.\frac{\omega^{i}}{\rho}\right|_{t_0} . \label{cauchy} \end{equation} This is also expressed as \begin{equation} \omega^i(\bbox{x},t) = \sqrt{ \frac{g(\bbox{x},t_0)}{g(\bbox{x},t)} } \omega^i(\bbox{x},t_0) . \label{cauchy2} \end{equation} The $\omega^0$ component is not independent of $\omega^i$. Using the relation $ \omega^{\mu} u_{\mu}=0 $, we obtain \begin{equation} \omega^0=g_{0i} \omega^i , \end{equation} The result Eq. (\ref{cauchy}) tells us that the vorticity is coupled to the density enhancement and vice versa. In particular, if the vorticity does not vanish exactly at an initial time, the vorticity will blow up as the density grows larger and larger (i.e. in the collapsing region), even if it has only the decaying mode in the linear perturbation theory. It should also be emphasized that our results Eqs. (\ref{masscon}) and (\ref{cauchy}) in the fully general relativistic treatment precisely correspond to those in the Newtonian case (see Eqs. (\ref{Nmasscon}) and (\ref{Ncauchy}) in the Appendix). \section{Perturbative Lagrangian approach} In the previous section, we solved exactly the equations for the density and the vorticity. The results Eqs. (\ref{masscon}) and (\ref{cauchy2}) show that $\rho$ and $\omega^i$ are completely written in terms of the determinant of the metric tensor and their initial values. In this section, we solve the metric perturbatively. We assume that the background is spatially flat Friedmann-Lema\^itre-Robertson-Walker (FLRW) universe. The extension to the spatially non-flat case must be a straightforward task. The perturbed metric is decomposed into \begin{eqnarray}\label{permet} g_{0i}&=& B_{,i}(\bbox{x}) + b_{i}(\bbox{x}) , \\ g_{ij}&=& a^2 \left(\delta_{ij} + 2 H_L \delta_{ij} + 2H_{T}{}_{,ij}+ (h_{i,j}+h_{j,i}) + 2 H_{ij} \right) , \nonumber \end{eqnarray} where $B$, $H_L$, and $H_T$ are scalar mode quantities, $b_i$ and $h_i$ are the vector (transverse) mode, and $H_{ij}$ is the tensor (transverse-traceless) mode satisfying \begin{eqnarray} b^i_{\ ,i}&=&0 , \\ h^i_{\ ,i}&=&0 , \\ H^i_{\ i}&=&0 , \\ H^{ij}_{\ \ ,j} &=&0 . \end{eqnarray} Raising and lowering indices of the perturbed quantities are done by $\delta^{ij}$ and $\delta_{ij}$. Using the perturbed metric Eq. (\ref{permet}), the general expression for the perturbed Einstein tensor up to the linear order is as follows: \begin{eqnarray} G_{\mu\nu} u^{\mu} u^{\nu} &=& 3\Bigr(\frac{\dot{a}}{a}\Bigr)^2 + 2 \frac{\dot{a}}{a}(3\dot{H}_L + \nabla^2 \dot{H}_T) - \frac{2}{a^2}\Bigl(\frac{\dot{a}}{a}\nabla^2 B + \nabla^2 H_L\Bigr) , \label{eqG00} \end{eqnarray} \begin{eqnarray} G_{\mu\nu} u^{\mu} P^{\nu}_{\ i} &=& -2 \Bigl(H_{L} + \frac{\dot{a}}{a}B \Bigr)^{\bbox{\cdot}}{}_{,i} + \frac{1}{2}\nabla^2 \Bigl(\dot{h}_i - \frac{1}{a^2} b_i \Bigr) - 2 \Bigl(\frac{\dot{a}}{a}\Bigr)^{\bbox{\cdot}} b_i , \label{eqG0i} \end{eqnarray} \begin{eqnarray}\label{eqGij} G_{\mu\nu} P^{\mu}_{\ i} P^{\nu}_{\ j} &=& -a^2 \left[ \nabla^2 \Bigl(\ddot{H}_T + 3\frac{\dot{a}}{a} \dot{H}_T - \frac{1}{a^2} H_L - \frac{\dot{a}}{a^3} B \Bigr) + 2 \Bigl( \ddot{H}_L + 3 \frac{\dot{a}}{a}\dot{H}_L \Bigr) \right] \delta_{ij} \\ && + a^2 \Bigl( \ddot{H}_T + 3 \frac{\dot{a}}{a}\dot{H}_T - \frac{1}{a^2}H_L - \frac{\dot{a}}{a^3}B\Bigr){}_{,ij} \nonumber\\ &&+ \frac{a^2}{2} \left[ \Bigl(\ddot{h}_i + 3 \frac{\dot{a}}{a}\dot{h}_i-\frac{\dot{a}}{a^3}b_i\Bigr){}_{,j} + \Bigl(\ddot{h}_j + 3 \frac{\dot{a}}{a}\dot{h}_j-\frac{\dot{a}}{a^3}b_j\Bigr){}_{,i}\right] \nonumber\\ &&+ a^2 \Bigl( \ddot{H}_{ij} + 3\frac{\dot{a}}{a}\dot{H}_{ij} -\frac{1}{a^2}\nabla^2 H_{ij} \Bigr), \nonumber \end{eqnarray} where an overdot ($\dot{}$) denotes $\partial/\partial t$, $\nabla^2=\delta^{ij}\partial_i \partial_j$, and we have used the fact $a(t) = t^{2/3}$ so that $G_{\mu\nu} P^{\mu}_{\ i} P^{\nu}_{\ j}$ does not have the background quantity. \subsection{Scalar perturbations} Using Eqs. (\ref{eqG0i}) and (\ref{eqGij}), the Einstein equations for the scalar perturbations are \begin{eqnarray} \Bigl(H_L + \frac{\dot{a}}{a} B \Bigr)^{\bbox{\cdot}} &=& 0 , \label{eqscalar1} \\ \ddot{H}_L + 3 \frac{\dot{a}}{a} \dot{H}_L &=& 0, \label{eqscalar2} \\ \ddot{H}_T + 3 \frac{\dot{a}}{a} \dot{H}_T &=& \frac{1}{a^2} \Bigl( H_L + \frac{\dot{a}}{a} B\Bigr) . \label{eqscalar3} \end{eqnarray} In order to solve the above equations, we use the residual gauge freedom (see Eq. (\ref{gaugeB}) in Appendix B) to set $B = 0$. Then, from Eq. (\ref{eqscalar1}) we have \begin{equation} H_L = H_L(\bbox{x}) \equiv \frac{10}{9} \Psi(\bbox{x}) . \end{equation} Next, from Eq. (\ref{eqscalar3}), we obtain two independent solutions for $H_T$: one is proportional to $t^{2/3}$ and the other to $t^{-1}$. According to Eq. (\ref{gaugeHT}), we may also use the residual gauge freedom for $H_T$ to add any function which does not depend on $t$. One may choose it as \begin{equation} H_T= \left( t^{2/3}-t_0^{2/3}\right) \Psi(\bbox{x}) + \left(t^{-1} - t_0^{-1}\right) \Phi(\bbox{x}) , \end{equation} so that $H_T$ vanishes at the initial time $t=t_0$. The initial density field $\rho(\bbox{x}, t_0)$ is also expressed by the metric. If the density contrast $\delta \equiv (\rho-\rho_b)/\rho_b$ is sufficiently small, we obtain from Eqs. (\ref{H}) and (\ref{eqG00}) \begin{equation} \delta = - \nabla^2 \left( t^{2/3} \Psi + t^{-1} \Phi \right) . \end{equation} Therefore, the initial density can be related with the initial metric perturbation as \begin{equation} \rho(\bbox{x}, t_0) = \rho_b(t_0) \left[ 1 - \nabla^2 \left(t_0^{2/3} \Psi + t_0^{-1} \Phi \right) \right] . \end{equation} The above expression is used only to relate the initial density fluctuation to the initial metric perturbations in linear regime. Once the initial seed is given, the later non-linear evolution is evaluated by the non-perturbative expression Eq.~(\ref{masscon}). \subsection{Vector perturbations} The Einstein equations for the vector perturbations are \begin{eqnarray} \nabla^2 \Bigl( \dot{h}_i - \frac{1}{a^2}b_i\Bigr) &=& 4 \Bigl(\frac{\dot{a}}{a}\Bigr)^{\bbox{\cdot}} b_i , \label{eqvector1}\\ \ddot{h}_i + 3 \frac{\dot{a}}{a} \dot{h}_i - \frac{\dot{a}}{a^3} b_i &=& 0 . \label{eqvector2} \end{eqnarray} Introducing $\beta_i$ as \begin{equation} b_i(\bbox{x})=\nabla^2 \beta_i(\bbox{x}) , \end{equation} Eq. (\ref{eqvector1}) is solved to give \begin{equation} h_i=-3 \left(t^{-1/3}- t_0^{-1/3}\right) \nabla^2 \beta_i(\bbox{x}) +{8 \over 3} \left(t^{-1}-t_0^{-1}\right) \beta_i(\bbox{x}) , \label{vector} \end{equation} where we again used the residual gauge freedom (cf. Eq. (\ref{gaugehi}) in the Appendix B) to set $h_i(\bbox{x},t_0) = 0$. The $\beta_i(\bbox{x})$ is directly related to the initial value of the vorticity field. From Eq. (\ref{cauchy}), we have \begin{equation} \omega^i(\bbox{x},t_0) = \frac{1}{2\sqrt{-g(\bbox{x},t_0)}} \varepsilon^{ijk} \nabla^2 \beta_{j,k}, \end{equation} where $\varepsilon^{ijk}$ is the 3-dimensional Levi-Civita symbol with $\varepsilon^{123} = 1$. Particularly when the deviation from the background is sufficiently small, we obtain the first-order expression \begin{equation} \omega^i(\bbox{x},t_0) \simeq \frac{1}{2 a^3(t_0)} \varepsilon^{ijk} \nabla^2 \beta_{j,k}. \end{equation} \subsection{Tensor perturbations} The equation for the tensor perturbations is \begin{equation} \ddot{H}_{ij} + 3\frac{\dot{a}}{a}\dot{H}_{ij} -\frac{1}{a^2}\nabla^2 H_{ij} = 0 . \end{equation} This is a homogeneous wave equation in the expanding universe. The solutions are well known and we will not discuss the detail here. See, e.g., \cite{Weinberg}. Before closing this section, we emphasize the following point: As for the metric, our result is identical to that of the linear perturbation theory (e.g. \cite{Bardeen}). However, there is an important difference. In Bardeen's paper, it is essential to linearize the density contrast. On the other hand, our Lagrangian approach does not rely on the assumption that the density contrast should be small. It is actually an important advantage of the Lagrangian approach that it uses (or extrapolates) the well-known solutions of the linear theory to express the non-linear density contrast. \section{Summary and discussion} Motivated by the fact that the Lagrangian condition enables us to obtain simply the integrals of the density and the vorticity along the fluid flow, we have developed the Lagrangian perturbation theory in the general relativistic cosmology, fully using this condition. The main advantage of the present Lagrangian theory is to be amenable to the vorticity, while previous works are limited within the irrotational fluid. In this approach, only the metric is expanded and its behaviors are determined perturbatively through the evolution equations, while the density and the vorticity are calculated non-perturbatively from the integrals along the flow. Seeds of the density and vorticity fluctuations can be related with the metric perturbation at an initial time, only when the constraint equation is solved as the Poisson-like equation. Hence, this simple approach is the first that can describe the dynamics of the dust fluid with the rotational motion as well as the non-linear density fluctuations, up to the caustic formation. As an illustration, the first order Lagrangian perturbation has been solved explicitly. It is natural that at the first order of the metric perturbation, our result agrees with that of the usual linear theory \cite{Bardeen} when we assume the small density contrast. Contrary to the standard linear perturbation theory, it is explicitly shown in our approach that the vorticity is coupled to the density contrast and is amplified in high density (i.e., collapsing) region. In this paper, we have presented a relativistic framework to describe non-linear evolution of the density fluctuation which is not restricted to the irrotational case. Once the framework is given, it should be straightforward to apply it to higher-order computations. Actually, an extension to the second order is being investigated by Morita \cite{Morita}. More comprehensive study of non-linear couplings will be of the subject of future investigation. \section*{Acknowledgment} H.A. is indebted to K. Tomita for hospitality at Yukawa Institute for the Theoretical Physics, where a part of this work was done. MK would like to thank Gerhard B\"orner for hospitality at Max-Planck-Institut f\"ur Astrophysik, where a part of this work was done.
\section{Introduction} Most methods used in the field of linear and nonlinear time series analysis assume stationarity of the considered data. Non--stationarity is very likely to lead to wrong results. This is especially true for tests for nonlinearity. A common approach is to split the time series into segments which can be considered nearly stationary and perform individual tests. But for short time series or not too slowly varying non--stationarities these segments have to be made too short to meaningfully calculate a test statistic on them. \section{Surrogate Data} To generate surrogate data that has the same time variation of autocorrelations than the original data one can in principle follow the general approach of constrained randomization.\cite{constraint} The corresponding cost function would be the discrepancy between the autocorrelations of the original and the surrogate data in sliding windows. Autocorrelations should be calculated up to sufficiently high lags. The most striking disadvantage of this method is the extreme computational burden. Therefore, we use an alternative method:\\ \begin{tabular}{ll} 1. & split up the time series into segments\\ 2. & \label{zwei} generate a surrogate for each segment\\ 3. & join the segments to one new surrogate\\ \end{tabular}\\ For step \ref{zwei} we use an iterative algorithm.\cite{surrowe,tisean} The size of the segments is typically too small to perform individual tests for nonlinearity on them, because most test statistics need sufficiently large data sets. A disadvantage of the method is the loss of correlations {\em between} the segments which can lead to a bias. \section{Cyclostationary Processes} The first class of processes we consider are cyclostationary, having periodically varying parameters. These processes have been found to yield rejections of the null hypothesis\cite{timmer} when using a test statistic derived from the correlation sum. Here we use the maximum likelihood estimator of the correlation dimension\cite{takens,theiler88,tisean} \begin{equation} \label{eq:takens} D_2^{\rm ML}={C(\epsilon_0) \over \int\limits_0^{\epsilon_0}{C(\epsilon) / \epsilon} \; d\epsilon} \, . \end{equation} Following Timmer\cite{timmer}, the considered time series is generated by an AR(2)--process \begin{equation} \label{eq:AR2} x_{n} = a_1 x_{n-1} + a_2 x_{n-2} + \eta_{n}, \quad n = 1,\ldots,2000 \end{equation} with coefficients chosen to be \begin{eqnarray} \label{eq:a1a2} a_1 & = & 2\cos(2\pi/T)\exp(-1/\tau)\\ a_2 & = & -\exp(-2/\tau) \nonumber \end{eqnarray} where $T$ is the period and $\tau$ the relaxation time of a damped oscillator. We will use $T=7$ and $\tau=50$. Non--stationarity is introduced by a modulation of the period $T$ \begin{equation} \label{eq:modula} T(n)=T_{\rm mean}+A\sin(2\pi/T_{\rm mod} n) \, . \end{equation} By varying $A$ and $T_{\rm mod}$ the degree and time scale of the non--stationarity can be changed which is reflected in non--constant running variances and autocorrelations. \subsection{Testing the test size} The fraction of tests that falsely reject the null hypothesis, when it is in fact true, is called the {\em size} of the test. Since a single test\cite{timmer} says not much about the size, we performed $1000$ one--sided tests with $9$ surrogates for each process. A correct test with proper surrogates should yield the nominal size $\alpha=0.1$. \begin{figure}[t] \begin{center} \epsfbox{size_2000.ps} \end{center} \caption{Size of nonlinearity tests with varying segment sizes. The modulation period of the non--stationarity is fixed to $T_{\rm mod}=2000$. From upper to lower curve the values $A=$2.0, 1.5, 1.0 and 0.0 are used.} \label{fig:s2000} \end{figure} For the first tests we used $T_{\rm mod}=2000$ for the modulation period. As shown in Fig.~\ref{fig:s2000} the new method does indeed yield the correct size of the test for a segment size of about $T_{\rm mod}/2=1000$. Only for relatively strong non--stationarity ($A=2.0$) smaller segments seem to be necessary. \begin{figure}[t] \begin{center} \epsfbox{size_1000.ps} \end{center} \caption{Size of nonlinearity tests with varying segment sizes. Now $T_{\rm mod}=1000$ and $A=2.0$ are used.} \label{fig:s1000} \end{figure} If we reduce the time scale of the non--stationarity by setting $T_{\rm mod}=1000$ we get sizes shown in Fig.~\ref{fig:s1000}. As expected, only segment sizes which are smaller than $T_{\rm mod}$ can decrease the number of false rejections. \section{AR(1)--Processes} We found cases in which non--stationarity does not yield a higher rate of rejections of the null hypothesis but just the opposite. Consider an AR(1)-process \begin{equation} \label{eq:AR1} x_{n}=a_1 x_{n-1} + \eta_{n} \end{equation} with a nonlinear measurement function $M(\cdot)$ \begin{equation} \label{eq:Mx} M(x_{n})=x_n\sqrt{|x_n|} \, . \end{equation} Non--stationarity is introduced by varying $a_1$ from $0.5$ to $0.9$ along the time series. We now choose a {\em prediction error} as test statistic.\cite{tisean} It is based on locally constant predictions in an $m$--dimensional embedding space. Again we performed 1000 one--sided tests with 9 surrogates to estimate the size. Additionally we calculated the rank of the original data within the surrogates going from 1 if the prediction error of the original data set was lower than for all surrogates up to 10 if the original data yields the largest prediction error. For usual stationary surrogates we get the distribution presented on the left side of Fig.~\ref{fig:histall}. In 35\% of the tests the original data has a {\em larger} prediction error than all the surrogates and the null is rejected in only 2.3\% of the tests. This is the opposite of what we expect for non--{\em linear} signals, which should be better predictable. Here, predictions are easier for the stationary surrogates with constant autocorrelations than for the original non--stationary data. Prediction error is not a useful discriminating statistic in this case. \begin{figure}[t] \epsfbox{histo05.ps} \epsfbox{histo_all.ps} \caption{Distributions of the rank of the original data within the surrogates for stationary surrogates (left) and surrogates with segment sizes 1000 (solid) and 250 (dashed).} \label{fig:histall} \end{figure} On the right side of Fig.~\ref{fig:histall} the results for tests with the new surrogates and different segment sizes are shown. For a segment size of $1000$ we get the correct size and therefore a uniform distribution. With decreasing segment sizes the prediction error of the surrogates gets larger than for the original data. The predictions seem to be very sensitive to the autocorrelations that get lost between the segments. \section{IV. Summary} Dealing with non--stationary signals is still a difficult business. We have given numerical evidence that in a test for nonlinearity of a data set with possible non--stationarity in form of slow variation it may be feasible to generate surrogates on separate segments. These segments can be made smaller than would be necessary for individual nonlinearity tests. But the segment size and the test statistic should still be chosen with great care, since no general theory guarantees the correctness of the test. As a final remark we want to stress that the reasoning in this work is not applicable to sudden changes like jumps or spikes.
\section{Introduction} The rich and high precision data on deep inelastic $e p $ scattering at HERA \cite{H1} \cite{ZEUS}, covering both low and high $Q^2$ regions, lead to a theoretical problem of matching the non-perturbative (``soft") and perturbative (``hard ") QCD domains. This challenging problem has been under close investigation over the past two decades, starting from the pioneering paper of Gribov \cite{GRIBOV} ( see also \cite{GVD} ). Based on Gribov's general approach, one can interpret two time sequence of the $\gamma^* p$ interaction (see \fig{fig1}): \begin{enumerate} \item\,\,\,First the $\gamma^*$ fluctuates into a hadron system (quark-antiquark pair to the lowest order) well before the interaction with the target; \item\,\,\,Then the converted quark-antiquark pair (or hadron system) interacts with the target. \end{enumerate} These two stages are expressed explicitly in the double dispersion relation suggested in Ref. \cite{GRIBOV}: \begin{equation}\label{GRIBOVgen} \sigma(\gamma^* N)=\frac{\alpha_{em}}{3\pi}\int \frac{\Gamma(M^2) dM^2}{(Q^2+M^2)} \sigma(M^2,M'^2,s) \frac{\Gamma(M'^2) dM'^2}{(Q^2+M'^2)}, \end{equation} where $M$ and $M'$ are the invariant masses of the incoming and outgoing quark-antiquark pairs, $\sigma(M^2,M'^2,s)$ is the cross section of a $q \bar q $ interaction with the target, and the vertices $\Gamma^2(M^2)$ and $\Gamma^2(M'^2)$ are given by $\Gamma^2(M^2)=R(M^2)$, where $R(M^2)$ is the ratio: \begin{equation}\label{Rdef} R(M^2)=\frac{\sigma(e^+e^-\rightarrow \mbox{hadrons})}{\sigma(e^+e^-\rightarrow \mu^+\mu^-)}, \end{equation} which has beem measured experimentally. For large masses ($M,\, M' >> M_0$) we have $\,\Gamma(M^2\,\geq\,M^2_0)\times \Gamma(M'^2\,\geq\,M^2_0 )\,\, \longrightarrow \,\,R( M^2 \,\geq\,\,M^2_0 ) \,=\, 2$, where we assume that that the number of colours $N_c = 3 $. $M_0$ is a typical mass (a separation parameter), which is of the order of $M_0 = 1 GeV $, determined directly from the experimental data \cite{TABLE}. \begin{figure}[tbp] \begin{center} \epsfig{file=fig1sx.eps,width=80mm} \caption[]{\parbox[t]{ 0.80\textwidth}{\small \it The generalized Gribov's formula.}} \end{center} \label{fig1} \end{figure} The key problem in all approaches utilizing \eq{GRIBOVgen} is the description of the cross section $\sigma(M^2,M'^2,s)$. In this paper we follow the approach suggested in Ref. \cite{GLMepj98}, which is based on the ideas of Badelek and Kwiecinski \cite{BK}. Below we list the mains steps of this approach. \begin{enumerate} \item\,\,\,We introduce $M_0 \,\,\approx \,\,1\,GeV$ in the integrals over $M$ and $M'$ which plays the role of a separation parameter. For $M,\, M' \,>\,M_0$ the quark - antiquark pair are produced at short distances ( $r_{\perp}\,\,\propto\,\frac{1}{M}\,\,<\,\,\frac{1}{M_0}$ ), while for $M,\, M'\,<\,M_0$ the distance between quark and antiquark is too long ($r_{\perp}\,\,\propto\,\frac{1}{M}\,\,>\,\,\frac{1}{M_0} $), and we cannot treat this $q \bar q $ - pair in pQCD. Actually, we cannot even describe the produced hadron state as a $q \bar q $ - pair; \item\,\,\,For $M,\,M'\,<\,M_0$ we use the Additive Quark Model \cite{AQM} in which \begin{eqnarray} \sigma(M^2,M'^2,s) &=&\sigma^{soft}_N (M^2,s) \,\delta(M^2\,-\,M'^2) \nonumber\\ &=& \left(\,\sigma_{tot}(qN )\,+\,\sigma_{tot}( \bar q N )\,\right)\,\delta(M^2\,-\,M'^2) \label{SOFT} \end{eqnarray} The above assumption allows us to simplify the Gribov formula of \eq{GRIBOVgen}: \begin{equation}\label{GRIBOVeq} \sigma(\gamma^* N)=\frac{\alpha_{em}}{3\pi}\int \frac{R(M^2) M^2 dM^2}{\left( Q^2+M^2\right)^2}\sigma_{N}(M^2,s)\,\,; \end{equation} \item\,\,\,For $M,\, M' \,>\,M_0$ we consider the system with mass $M$ and/or $M'$ as a short distance quark - antiquark pair, and describe its interaction with the target in pQCD (see \fig{twogluons}). The exact formulae for $\sigma^{hard}(M^2,M'^2,s)$ is given below (see also Ref. \cite{GLMepj98}), but one can see from \fig{twogluons}, that this interaction can be expressed through the gluon structure function, and it is {\em not} diagonal with respect to the masses, contrary to the ``soft" interaction of a hadron system with small mass. \item\,\,\,In principle, the short distance between the quark and antiquark leads to short distances in the gluon-nucleon interaction. However, the typical distance of the gluon interaction $r_G\,\,\propto\,1/l_{\perp}$ (see Figs 1 and 2) is larger than the size of the quark - antiquark pair and $l_{\perp} \,\,<\,\,{k_{\perp}}$. It turns out that the calculation with $M_0\,\, \sim\,\,1\,GeV$, demands a new scale in the gluon-nucleon interaction. We introduce it, assuming that the gluon structure function $xG(x,l^2_{\perp})$ behaves as $xG(x,\mu^2)\,\times\,\frac{l^2_{\perp}}{\mu^2}$ for $l^2_{\perp}\,\leq\,\mu^2$. This means that we assume that the gluon-hadron total cross section is not equal to zero for long distances (in the ``soft" kinematic region). It should be stressed that we introduce two scales for soft nonperturbative interactions, distinguishing between quark and gluon interactions. The two scales appear naturally in our models for ``soft", nonperturbative interaction. For example, in the AQM which we use for $\sigma^{\scrbox{soft}}$, these two scales are the size of a hadron (distance between the quark and antiquark), and the size of the constituent quark which is related to gluon interaction scale. \item\,\,\, As has been discussed, we use the AQM [see \eq{SOFT}] to calculate $\sigma^{\scrbox{soft}}$. For the vector meson cross section the AQM leads to \begin{equation}\label{RCRSEC} \sigma_{\scrbox{R}}\,\,=\frac{1}{2}\left(\,\sigma(\pi^+ p)\,\,+\,\,\sigma(\pi^-p)\right). \end{equation} For the pion-proton cross section, we use the Donnachie - Landshoff Reggeon parameterization \cite{DL}, with an energy variable which is appropriate for the interaction of a hadronic system of mass $M$ with the target (see \cite{GLMepj98} for details). \begin{equation} \label{DL} \sigma_{M^2N}(s) = \dlformula, \end{equation} with \begin{eqnarray} \label{alphapom} {\alpha_{\pom}} &=& 1.079 \nonumber \\ {\alpha_{\regg}} &=& 0.55 \nonumber \\ x_M &=& \frac{M^2}{s} \nonumber \\ A &=& 13.1\, \mbox{mb} \nonumber \\ B &=& 41.08\, \mbox{mb}\, . \end{eqnarray} The values of $A$ and $B$ are chosen so that (\ref{DL}) is valid for the $\rho$-proton interaction. Using (\ref{DL}) in Gribov formula (\ref{GRIBOVeq}) we derive the soft transverse contribution to $\sigma(\gamma^* p)$: \begin{equation}\label{soft2} \sigma^{\scrbox{soft}}_{\scrbox{T}} = \softformula{M^2}\left\{\dlformula\right\}. \end{equation} \end{enumerate} \begin{figure}[tbp] \begin{center} \epsfig{file=fig2sx.eps,width=90mm} \caption[]{\parbox[t]{ 0.80\textwidth}{\small \it ``Hard" contribution to $\sigma(M^2,M'^2,s)$ in perturbative QCD .}} \end{center} \label{twogluons} \end{figure} It has been shown in \cite{GLMepj98} that even though the simple model discussed above reproduces the main features of the experimental data, it has several deficiencies which we wish to address: \begin{itemize} \item The calculations of the ``soft'' cross section appeared to overestimate the data, and to overcome this, the soft contribution was multiplied by a factor $\kappa =0.6$. There is no physical justification for such a small value. \item The contribution of a longitudinal polarized virtual photon to $\sigma(\gamma^* p)$ was neglected. \item An old GRV'94 parameterization was used for the gluons distribution inside the proton. This gave an energy dependence of $\sigma(\gamma^* p)$ which is steeper than the published data. \end{itemize} In the present paper we reexamine these points developing a formalism which also takes into account $\sigma_{\scrbox{L}}$, the longitudinal part of $\sigma(\gamma^* p)$, for both the soft and the hard components. We find that the contribution of $\sigma_{\scrbox{L}}$ significantly improves the energy dependence of $\sigma(\gamma^* p)$, in a good agreement with the data for $x < 10^{-2}$. A similar approach to photon - proton interaction was also developed in Ref.\cite{MRS} where quite different scales have been introduced. The goal of such an approach (see Refs.\cite{GLMepj98},\cite{BK} and \cite{MRS}) is to find parameters that separate the long distance (non-perturbative) and short distance (perturbative) interactions in QCD. A similar philosophy to ours is used in Ref.\cite{BW}, where an attempt was made to describe photon - proton interactions, assuming that the main contribution stems from short distances, and the long distance interaction appears as a result of the shadowing corrections. The description of $\sigma( \gamma^* p )$ can be achieved in quite a different way using the generalized vector dominance model\cite{GVD} or a Regge motivated fit of the experimental data\cite{ALLM}\cite{ALLM97}. In these approaches one looses the explicit connection with the microscopic theory, which makes futher theoretical interpretation of the results rather difficult. \section{Description of the Model \label{secMODEL}} \subsection{The Contribution of a Transverse Polarized Photon \label{secT}} \begin{figure}[tbp] \begin{center} \epsfig{file=fig3.ps,width=140mm} \caption[]{\parbox[t]{ 0.80\textwidth}{\small \it Diagrams which cntribute to the scattering of a $\bar{q}q$ pair off the target proton.} \label{fig3}} \end{center} \end{figure} As we have discussed previously, the main assumption of our model is that $\sigma(M^2,M'^2,s)$ in \eq{GRIBOVgen} can be calculated as \begin{equation} \label{S+H} \sigma(M^2,M'^2,s)=\sigma^{soft}_N(M^2\,<\,M^2_0,s)\,\, \delta( M^2 - M'^2 )+ \sigma^{hard}(M^2\,>\,M^2_0,M'^2\,>\,M^2_0,s)\,\,. \end{equation} In the previous section we have discussed the calculation of $\sigma^{\scrbox{soft}}$. Now we present, for completeness, our formulae for $\sigma^{\scrbox{hard}}$ which have been given in Ref.~\cite{GLMepj98}. The pQCD contribution for $\sigma_{\scrbox{T}}$, within the framework of the two gluon model, is illustrated in the four diagrams shown in Fig.~\ref{fig3}. The production amplitude ${\cal M}_{{\lambda\lambda'}}$ of the $\bar{q}q$, can be factorized into the wave function $\psi_{\lambda\lambda'}$ of the $\bar{q}q$ system inside the virtual photon, and the amplitude ${\cal T}_{\lambda\lambda'}$ for the scattering of the $\bar{q}q$ pair off the proton: \begin{equation}\label{mlamlam1} {\cal M}_{\lambda\lambda'} ({k_{\perp}},z) = \sqrt{N_c}\int d^2 {k_{\perp}}'\int_0^1 dz' \psi_{\lambda\lambda'} ({k_{\perp}}',z'){\cal T}_{\lambda\lambda'} ({k_{\perp}}',z';{k_{\perp}},z)\, . \end{equation} The transition ampiltude ${\cal T}_{\lambda\lambda'}$ for the four contributions is: \begin{eqnarray}\label{tlamlam} \lefteqn{{\cal T}_{\lambda\lambda'} ({k_{\perp}}', {k_{\perp}}) =} & & \\ & & i \frac{4\pi s}{2N_c} \int \frac{d^2 \ell_{\perp}}{\ell_{\perp}^4} \biggl [ 2 \delta (\mbox{\boldmath $k$}_{\perp}' - \mbox{\boldmath $k$}_{\perp}) - \delta (\mbox{\boldmath $k$}_{\perp}' - \mbox{\boldmath $k$}_{\perp} - \mbox{\boldmath $\ell$}_{\perp}) - \delta (\mbox{\boldmath $k$}_{\perp}' - \mbox{\boldmath $k$}_{\perp} + \mbox{\boldmath $\ell$}_{\perp} ) \biggr ] \alpha_{\rm S} (\ell^2_{\perp}) f(x, \ell^2_{\perp})\, , \nonumber \end{eqnarray} where, \begin{equation}\label{ddxg} f(x,\ell^2)=\frac{\partial xG(x,\ell^2)}{\partial\ln\ell^2} \end{equation} (see \cite{LMRT} for details). Substituting (\ref{tlamlam}) in (\ref{mlamlam1}) we have, \begin{equation}\label{mlamlam2} {\cal M}_{\lambda\lambda'} = i\frac{2\pi^2 s}{\sqrt{N_c}}\int\frac{d\ell^2_{\perp}}{\ell_{\perp}^4} \alpha_{\rm S}(\ell^2_{\perp}) f(x,\ell^2_{\perp})\Delta\psi_{\lambda\lambda'} \, , \end{equation}where the definition of $\Delta\psi$ is to be understood from~ \eqs{mlamlam1}{mlamlam2}: \begin{equation}\label{Deltapsi} \Delta\psi = 2\psi(\mbox{\boldmath $k$}_{\perp},z) - \psi(\mbox{\boldmath $k$}_{\perp} - \mbox{\boldmath $\ell$}_{\perp},z) - \psi(\mbox{\boldmath $k$}_{\perp} + \mbox{\boldmath $\ell$}_{\perp},z)\, . \end{equation} $\psi_{\lambda\lambda'} $ has been calculated in \cite{LMRT} both for a transverse and for a longitudinal polarised incoming photon. The wave function for a transverse polarized photon and the amplitude for producing $\bar{q}q$ with helicities $\lambda$ and $\lambda'$ read: \newcommand{\lint}[2]{\int\frac{d #1^2}{#1^4} \alpha_{\rm S}(#2 ^2) f(x,#2 ^2)} \newcommand{\lintarg}[3]{\left\{ \frac{1}{#3^2+#2^2} - \frac{1}{\sqrt{\left( #3^2 + #2^2 + #1^2\right)^2- 4 #2^2 #1^2}} \right\}} \newcommand{\lintargb}[3]{\left\{ \frac{1}{#3^2+#2^2} - \frac{1}{2 #2^2} + \frac{#3^2-#2^2+#1^2}{2 #2^2\sqrt{\left( #3^2 + #2^2 + #1^2\right)^2- 4 #2^2 #1^2}} \right\}} \newcommand{\lintargc}[3]{\left\{ \frac{#2^2 - #3^2}{#2^2 + #3^2 } + \frac{#3^2-#2^2+#1^2}{\sqrt{\left( #3^2 + #2^2 + #1^2\right)^2- 4 #2^2 #1^2}} \right\}} \begin{eqnarray} \psi^{\pm}_{\lambda\lambda'}\left({\mathbf {k_{\perp}}} ,z\right) &=& -\delta_{\lambda , -\lambda'} Z_f e \left[(1-2z)\lambda\mp 1\right] \frac{{2\mathbf\epsilon_{\pm}\cdot {k_{\perp}}}}{\bar{Q}^2+{$k^2_{\perp}$}} , \label{psilamlamt} \\ {\cal M}^{\pm}_{\lambda\lambda'} &=& -2 Z_f e \left( i\frac{2\pi^2 s}{\sqrt{N_c}}\right) \delta_{\lambda , -\lambda'} \left[(1-2z)\lambda\mp 1\right] {\mathbf\epsilon_{\pm}\cdot {k_{\perp}}} \lint{\ell_{\perp}}{\ell_{\perp}}\,\times \nonumber \\ & & \hspace{1cm}\lintargb{\ell_{\perp}}{{k_{\perp}}}{\bar{Q}} . \label{mlamlamt} \end{eqnarray} In (\ref{psilamlamt}) and (\ref{mlamlamt}), $Z_f$ is the charge of the quark with flavour $f$ in units of the electron charge $-e$, $\bar{Q}^2\equiv z(1-z)Q^2$ and $\epsilon_{\pm}$ denotes the photon polarization vector presented in a circular basis, { \begin{equation} \mathbf{\epsilon_{\pm}}=\frac{1}{\sqrt{2}}\left(0,0,1,\pm i\right). \end{equation} To evaluate the cross section one should, first, sum over the quark helicities $\lambda$ and $\lambda'$, and average over the two transverse polarisation states $(\pm)$ of the photon \begin{eqnarray} |{\cal M}^T|^2 &=& \frac{32\pi^4 s^2}{N_c} Z^2_f e^2 \left[z^2+(1-z)^2\right] \lint{\ell_{\perp}}{\ell_{\perp}}\,\times \nonumber \\ & & \hspace{1cm}\lintargc{\ell_{\perp}}{{k_{\perp}}}{\bar{Q}} . \end{eqnarray} The cross section is obtained by integration over $z$ and over ${k_{\perp}}$, \begin{eqnarray} \sigma_{\scrbox{T}} &=& \frac{\alpha_{\scrbox{em}}}{4\pi^2 N_c}\int^1_0 dz\left[z^2+(1-z)^2\right] \int\frac{d {$k^2_{\perp}$}}{\bar{Q}^2+{$k^2_{\perp}$}} \lint{\ell_{\perp}}{\ell_{\perp}}\,\times \nonumber \\ & & \hspace{3cm}\lintargc{\ell_{\perp}}{{k_{\perp}}}{\bar{Q}} . \end{eqnarray} Following \cite{GLMepj98}, in order to perform the integration over $z$, we introduce the variables $M$ and $\tilde M$: \begin{eqnarray} M^2 &=& \frac{{$k^2_{\perp}$}}{z(1-z)} \nonumber \\ \tilde M^2 &=& \frac{\ell^2_{\perp}}{z(1-z)} \label{mmtdef}\, , \end{eqnarray} and rewrite the integrals of $\sigma_{\scrbox{T}}$ in terms the variables $M$, $\mathbf{\tilde M}$ and { $\mathbf{\ell_{\perp}}$,} \newcommand{\sqrt{1-4\frac{\lts}{\mt^2}}}{\sqrt{1-4\frac{\ell^2_{\perp}}{\tilde M^2}}} \newcommand{1-2\frac{\lts}{\mt^2}}{1-2\frac{\ell^2_{\perp}}{\tilde M^2}} \begin{eqnarray} \sigma_{\scrbox{T}} &=& \frac{\alpha_{\scrbox{em}}}{4\pi^2 N_c}\int\frac{d M^2}{Q^2+M^2}\int\frac{d \tilde M^2}{\tilde M^2} \int\frac{d \ell^2_{\perp}}{\ell^2_{\perp}}\frac{1-2\frac{\lts}{\mt^2}}{\sqrt{1-4\frac{\lts}{\mt^2}}} \alpha_{\rm S}(\ell^2_{\perp}) f(x,\ell^2_{\perp})\,\times \nonumber \\ & & \hspace{3cm}\lintargc{\tilde M}{M}{Q} . \label{EQsight2} \end{eqnarray} Using a generalization of the Gribov approach, we take a cutoff on the integration over $M^2$ and insert the ratio (\ref{Rdef}) inside the integration sign, \begin{eqnarray} \sigma^{\scrbox{hard}}_{\scrbox{T}} &=& \frac{\alpha_{\scrbox{em}}}{2\pi^2 N_c}\int^{\infty}_{M^2_0} \frac{d M^2\,\,R(M^2)}{Q^2+M^2}\int\frac{d \tilde M^2}{\tilde M^2} \int\frac{d \ell^2_{\perp}}{\ell^2_{\perp}}\frac{1-2\frac{\lts}{\mt^2}}{\sqrt{1-4\frac{\lts}{\mt^2}}} \alpha_{\rm S}(\ell^2_{\perp}) f(x,\ell^2_{\perp})\,\times \nonumber \\ & & \hspace{1cm}\lintargc{\tilde M}{M}{Q} . \label{EQsight3} \end{eqnarray} The last integral in (\ref{EQsight3}) can be integrated by parts, using (\ref{ddxg}). In the limit $4\ell^2_{\perp}\ll\tilde M^2$, the value of the integral is dominated by the upper limit of the integration, therefore we replace $\ell^2_{\perp}$ in $\alpha_{\rm S}(\ell^2_{\perp})$ and $xG(x,\ell^2_{\perp})$ with $\tilde M^2/4$, and obtain, for three colors \begin{eqnarray} \label{hardT} \sigma^{\scrbox{hard}}_{\scrbox{T}} &=& \frac{2\pi\, \alpha_{em}}{3} \int^{\infty}_{M^2_0} \frac{d M^2 \,\,R(M^2)}{Q^2\,+\,M^2}\,\, \int^{\infty}_{0}\frac{d{\tilde M}^2}{{\tilde M}^4} \alpha_{\rm S}(\frac{{\tilde M}^2}{4})\,x\,G(x,\frac{{\tilde M}^2}{4}) \times \nonumber \\ & & \hspace{2cm} \left\{\,\,\frac{M^2\,-\,Q^2}{M^2\,+\,Q^2}\,\,+\,\,\frac{Q^2\,+\,{\tilde M}^2\,-\,M^2} {\sqrt{(\,Q^2\,+\,M^2\,+\,{\tilde M}^2\,)^2\,-\,4\,M^2\,{\tilde M}^2}}\,\,\right\}\,\,. \end{eqnarray} In the case of heavy quarks, we assume that the quark is heavy enough so that any contribution to the soft cross-section can be neglected. The heavy quarks contribution is then written with the replacements: $ 4\,M^2\,{\tilde M}^2 \rightarrow 4\,(M^2-4m_Q^2)\,{\tilde M}^2$, and $R(Q^2) \rightarrow R^{QQ}(M^2)$. $R^{QQ}$ is the heavy quark contribution to the ratio (\ref{Rdef}) and $m_Q$ is the mass of the heavy quark, \begin{eqnarray} \label{hardTQQ} \sigma^{\scrbox{hard}}_{\scrbox{T},\bar{Q}Q} &=& \frac{2\pi\, \alpha_{em}}{3} \int^{\infty}_{M^2_0} \frac{d M^2 \,\,R^{QQ}(M^2)}{Q^2\,+\,M^2}\,\, \int^{\infty}_{0}\frac{d{\tilde M}^2}{{\tilde M}^4} \alpha_{\rm S}(\frac{{\tilde M}^2}{4})\,x\,G(x,\frac{{\tilde M}^2}{4}) \nonumber \\ & & \mbox{\hfill} \left\{\,\,\frac{M^2\,-\,Q^2}{M^2\,+\,Q^2}\,\,+\,\,\frac{Q^2\,+\,{\tilde M}^2\,-\,M^2} {\sqrt{(\,Q^2\,+\,M^2\,+\,{\tilde M}^2\,)^2\,-\,4\,(M^2-4m_Q^2)\,{\tilde M}^2}}\,\,\right\}\,\,. \end{eqnarray} In the above formulae $x=x(M^2)=(Q^2+M^2)/W^2$ with $W$ being the center of mass energy of the photon-nucleon interaction. \subsection{The Contribution of a Longitudinal Polarized Photon \label{secSIGL}} \subsubsection{``Soft'' Contribution to $\sigma_L$ \label{subsecSIGSL}} A priori, it is straight forward to write the formula for the ``soft'' component of $\sigma_{\scrbox{L}}$ in AQM. The result is similar to our model for $\sigma_{\scrbox{T}}$, except that $M^2$ in the numerator should be replaced with the photon virtuality. This factor causes the longitudinal cross section to vanish for $Q^2 \rightarrow 0$ (see (\ref{alphapom}) for the values of the parameters). \begin{equation}\label{softl} \sigma^{\scrbox{soft}}_{\scrbox{L}} = \softformula{Q^2}\left\{\dlformula\right\}. \end{equation} However, it turns out that the AQM over estimates the experimental data, and it has to be reduced with some phenomenological procedure, such as a numerical factor \cite{SchSpi} or a phenomenological function which decreases with $Q^2$~\cite{MRS}. In our formalism the ratio between the AQM and pQCD contributions depends on the separation parameter $M_0$. Thus, by lowering the value of $M_0$ the ``soft'' component is suppressed. We have found that for the longitudinal part of the cross section, taking any value of $M_0$ which is below the resonances mass fits the experimental data, as opposed to the transverse cross section where we used $0.7 < M_0^2 < 0.9 {\,\mbox{GeV}^2}$ . \subsubsection{``Hard'' Contribution to $\sigma_L$} We calculate the pQCD contribution for $\sigma_{\scrbox{L}}$, from the diagrams of Fig.~\ref{fig3}, using \eqs{mlamlam1}{Deltapsi}. For the case of a longitudinal polarized incoming photon, we use the expression derived in \cite{LMRT} for the wave function \begin{equation} \psi^{\scrbox{L}}_{\lambda\lambda'} = -2\delta_{\lambda , -\lambda'} Z_f e Q z(1-z) \frac{1}{\bar{Q}^2+{$k^2_{\perp}$}} , \label{psiL} \end{equation} and substitute it in (\ref{mlamlam2}) to obtain the amplitude for a longitudinal photon to produce $\bar{q}q$ with helicities $\lambda$ and $\lambda'$ \begin{eqnarray} {\cal M}_{\lambda\lambda'}^{\scrbox{L}} &=& -4\delta_{\lambda , -\lambda'} Z_f e Q z(1-z) i\frac{2\pi^2 s}{\sqrt{N_c}} \lint{\ell_{\perp}}{\ell_{\perp}}\,\times \nonumber \\ & & \hspace{4cm}\lintarg{\ell_{\perp}}{{k_{\perp}}}{\bar{Q}} . \label{mlamlaml} \end{eqnarray} The cross section is then obtained by squaring, summing over the final helicities and integrating over $z$ and ${k_{\perp}}$, \begin{eqnarray} \sigma_{\scrbox{L}} &=& \frac{16 \pi\alpha_{\scrbox{em}}}{N_c}\sum Z_f^2 Q^2 \int_0^1 dz \left[ z(1-z) \right]^2 \int \frac{d{$k^2_{\perp}$}}{\bar{Q}^2+{$k^2_{\perp}$}} \lint{\ell_{\perp}}{\ell_{\perp}} \times \nonumber \\ & & \hspace{4cm}\lintarg{\ell_{\perp}}{{k_{\perp}}}{\bar{Q}}. \label{EQsighl1} \end{eqnarray} Changing the integration variables ${$k^2_{\perp}$}$ and $\ell^2_{\perp}$ to $M$ and $\tilde M$, respectively, using (\ref{mmtdef}), we have \begin{eqnarray} \sigma_{\scrbox{L}} &=& \frac{16 \pi\alpha_{\scrbox{em}}}{N_c}\sum Z_f^2 Q^2 \int_0^1 dz \int \frac{dM^2}{Q^2+M^2} \lint{\tilde M}{z(1-z)\tilde M} \times \nonumber \\ & & \hspace{4.5cm} \lintarg{\tilde M}{M}{Q} \, . \label{EQsighl2} \end{eqnarray} We now write the formula for $\sigma^{\scrbox{hard}}_{\scrbox{L}}$, in the same way that we did in section \ref{secT}, by taking a cutoff on the integration over $M^2$, and insert the ratio (\ref{Rdef}) inside the integration sign, \begin{eqnarray} \sigma^{\scrbox{hard}}_{\scrbox{L}} &=& \frac{8 \pi\alpha_{\scrbox{em}}}{N_c}Q^2\int_{M_0^2}^\infty \frac{R(M^2)dM^2}{Q^2+M^2}\int_0^\infty\frac{d\tilde M^2}{\tilde M^4} \int_0^1 dz f\left(x,z(1-z)\tilde M^2\right) \alpha_{\rm S}\left(z(1-z)\tilde M^2\right) \times \nonumber \\ & & \hspace{4.5cm}\lintarg{\tilde M}{M}{Q}\, . \label{EQsighl3} \end{eqnarray} Recalling (\ref{ddxg}), we can integrate \eq{EQsighl3} by parts and obtain (for $N_c=3$), \newcommand{\sqrt{(Q^2+M^2+\mt^2)^2-4 M^2 \mt^2}}{\sqrt{(Q^2+M^2+\tilde M^2)^2-4 M^2 \tilde M^2}} \begin{eqnarray} \sigma^{\scrbox{hard}}_{\scrbox{L}} &=& \frac{8\pi\alpha_{\scrbox{em}}}{3}Q^2\int_{M_0^2}^\infty \frac{R(M^2)dM^2}{Q^2+M^2}\int_0^\infty\frac{d\tilde M^2}{\tilde M^4} \, \overline{xG}\left(x,\tilde M^2\right)\times \nonumber \\ & & \hspace{2cm} \left\{\frac{1}{Q^2+M^2} - \frac{1}{\sqrt{(Q^2+M^2+\mt^2)^2-4 M^2 \mt^2}} \,- \nonumber \right.\\ & & \hspace{5.5cm}\left. \frac{\tilde M^2\left(Q^2+\tilde M^2-M^2\right)} {\left(\sqrt{(Q^2+M^2+\mt^2)^2-4 M^2 \mt^2}\right)^3}\right\}, \end{eqnarray} where, \begin{equation} \overline{xG}\left(x,\tilde M^2\right) \equiv \int_0^1 xG\left(x,z(1-z)\tilde M^2\right) \alpha_{\rm S}\left(z(1-z)\tilde M^2\right) dz\, . \end{equation} For $\sigma^{\scrbox{hard}}_{\scrbox{L},\bar{Q}Q}$, the contribution of heavy quarks to the longitudinal component of $\sigma(\gamma^* p)$, we replace $ 4M^2{\tilde M}^2 $ by $4(M^2-4m_Q^2){\tilde M}^2$ and $R(Q^2)$ with $R^{QQ}(M^2)$, as we did for the transverse part: \newcommand{\sqrt{(Q^2+M^2+\mt^2)^2-4(M^2-4m_Q^2) \mt^2}}{\sqrt{(Q^2+M^2+\tilde M^2)^2-4(M^2-4m_Q^2) \tilde M^2}} \begin{eqnarray} \sigma^{\scrbox{hard}}_{\scrbox{L},\bar{Q}Q} &=& \frac{8\pi\alpha_{\scrbox{em}}}{3}Q^2\int_{M_0^2}^\infty \frac{R^{QQ}(M^2)dM^2}{Q^2+M^2}\int_0^\infty\frac{d\tilde M^2}{\tilde M^4} \, \overline{xG}\left(x,\tilde M^2\right)\times \nonumber \\ & & \hspace{2cm} \left\{\frac{1}{Q^2+M^2} - \frac{1}{\sqrt{(Q^2+M^2+\mt^2)^2-4(M^2-4m_Q^2) \mt^2}} \,- \nonumber \right.\\ & & \hspace{4cm}\left. \frac{\tilde M^2\left(Q^2+\tilde M^2-M^2\right)} {\left(\sqrt{(Q^2+M^2+\mt^2)^2-4(M^2-4m_Q^2) \mt^2}\right)^3}\right\}\, . \label{hardLQQ} \end{eqnarray} \section{Comparison With The Experimental Data} We now present the results for our calculation of $\sigma(\gamma^* p)$ from the master equation \begin{equation}\label{sigmatotal} \sigma(\gamma^* p) = \sigma^{\scrbox{soft}}_{\scrbox{T}} + \sigma^{\scrbox{hard}}_{\scrbox{T}} + \sigma^{\scrbox{hard}}_{\scrbox{T},\bar{Q}Q} + \sigma^{\scrbox{soft}}_{\scrbox{L}} + \sigma^{\scrbox{hard}}_{\scrbox{L}} + \sigma^{\scrbox{hard}}_{\scrbox{L},\bar{Q}Q} \, . \end{equation} As stated, our model has three parameters, namely, $M_{0,L}, M_{0,T}$ and $\mu$. In \cite{GLMepj98} the relatively high value of $M_0=\sqrt{5}$ GeV was taken, in order to give an energy dependence that was in a reasonable agreement with the published experimental results. We found that if the hard longitudinal contributions are not neglected, the transverse separation parameter can be reduced to the value of $0.7 < M_{0,T}^2 < 0.9 \,{\,\mbox{GeV}^2}$, while the longitudinal one can be taken to be $M_{0,L}^2 \;\raisebox{-.4ex}{\rlap{$\sim$}} \raisebox{.4ex}{$<$}\; 0.4 \,{\,\mbox{GeV}^2}$. The results of the calculation are stable if the separation parameters are chosen within these bounds. \begin{figure}[tbp] \begin{center} \epsfig{file=grvmrst.ps,width=140mm} \caption[]{\parbox[t]{ 0.80\textwidth}{\small \it Comparison of the gluon distribution parametrizaions which were used in Eq.\ \mbox{(\ref{sigmatotal})} to calculate $\sigma(\gamma^* p)$. The full triangles correspond to data points which have been extracted in \cite{ZEUS98} from experimental data.} \label{comparexg}} \end{center} \end{figure} In the calculation of the hard components of $\sigma(\gamma^* p)$, we need to specify the input gluon distribution $xG(x,Q^2)$ which appears in the formulae. We have tried several options: GRV'94 \cite{GRV94}, GRV'98 \cite{GRV98} and MRST'98 \cite{MRST}. We compare the results of each input distribution for the calculation of $\sigma(\gamma p)$ in \fig{comparexg}, where it is obvious that the best parameterization for our purpose is MRST'98 which yields, together with our formalism, a good description of the data. A difficulty in our calculation comes from the integration at very low $M^2$, where the published parameterizations for $xG$ are not valid. Naturally one has two possible options: the first is to impose a low cutoff on the integration below $M^2=\mu^2$. \, The second option is to use the general property of the gluon structure function which is linear in $Q^2$ at very small values of $Q^2$, and to rely on this property for the low mass region: \begin{equation}\label{mu} xG(x,l^2<\mu^2) = \frac{l^2}{\mu^2}xG(x,\mu^2)\, . \end{equation} In our calculation for the low mass integration we used~(\ref{mu}) for the gluon distribution. The coupling $\alpha_{\rm S}(Q^2)$ was also kept fixed for $Q^2<\mu^2$. The value of $\mu^2$ was taken as the minimal value which is allowed in the input parameterization that had been used: $\mu^2=0.4,0.8,1.25\,{\,\mbox{GeV}^2}$ for GRV'94, GRV'98 and MRST'98 respectively. Notice that for the MRST parameterization $\mu^2=1.25{\,\mbox{GeV}^2} > M_0^2$. This means that the transition from soft to hard is not sharp, and we still have (\ref{mu}) as a soft signature inside our hard formalism at $M_0^2<M^2<\mu^2$. \begin{figure}[tbp] \begin{center} \epsfig{file=sigmatotw2.ps,width=140mm} \caption[]{\parbox[t]{ 0.80\textwidth}{\small \it $\sigma(\gamma^* p)$ as a function of $W^2$ together with the experimental data. The full triangles correspond to data points which have been extracted in \cite{ZEUS98} from experimental data.} \label{stotw2}} \end{center} \end{figure} In \fig{stotw2} we show the results of our calculation for $\sigma(\gamma^* p)$ as a function of $W^2$ for fixed values of $Q^2$, together with the published experimental results. The calculated results are shown as a band which corresponds to the two limits on $M_{0,T}$. It is clear from the figure that the width of the band decreases with increasing $Q^2$. As for the longitudinal separation parameter, the results are not sensitive to the choice of $M_{0,L}$, and in all our calculations we simply used $M_{0,L} = 2 m_\pi$. The thick line marks the region of small $x$: to the left of it, $x$ is not small enough to justify using our model. One can see that for $x < 10^{-2}$, our results reproduce the experimental results both in value and in energy dependence. \begin{figure}[tbp] \begin{center} \epsfig{file=sigmatotfixedw.ps,width=140mm} \caption[]{\parbox[t]{ 0.80\textwidth}{\small \it $\sigma(\gamma^* p)$ as a function of $Q^2$ together with the experimental data.} \label{stotq2}} \end{center} \end{figure} In \fig{stotq2} we show our calculations for $\sigma(\gamma^* p)$ for fixed values of $W^2$, as a function of $Q^2$, scaled by factors of $1-128$. The small vertical thick lines delineate the boundary $x = 10^{-2}$, to the right of the marks our results are not reliable. \begin{figure}[tbp] \begin{center} \epsfig{file=longtranfixedw.ps,width=140mm} \caption[]{\parbox[t]{ 0.80\textwidth}{\small \it The ratio of $\displaystyle{\frac{\sigma_{\scrbox{L}}}{\sigma_{\scrbox{T}}}}$ as a function of $x$.} \label{longtran}} \end{center} \end{figure} The longitudinal component of $\sigma(\gamma^* p)$, according to our calculation, reaches a maximum value of 25{\%\ } of the total cross section at $x\approx 10^{-2}$. For small $x$, this result is closer to the published data \cite{NMC} than the results of Ref.\ \cite{MRS}, where they obtained $\sigma_{\scrbox{L}}$ for $x\approx 10^{-2}$ to be 15{\%\ } of the total cross section. The ratio of the longitudinal cross section to the transverse cross section, as a function of $x$, is shown in \fig{longtran} together with the (small $x$) data of Ref.\ \cite{NMC}. \begin{figure}[tbp] \begin{center} \epsfig{file=heavylight.ps,width=140mm} \caption[]{\parbox[t]{ 0.80\textwidth}{\small \it The ratio of $\frac{\sigma^{\scrbox{heavy}}}{\sigma^{\scrbox{light}}}$ at fixed $Q^2$. } \label{heavylight}} \end{center} \end{figure} \fig{heavylight} illustrates the importance of the heavy quarks contributions $\sigma^{\scrbox{hard}}_{\scrbox{T},\bar{Q}Q}$ and $\sigma^{\scrbox{hard}}_{\scrbox{L},\bar{Q}Q}$ [see Eqs.\ (\ref{hardTQQ}) and (\ref{hardLQQ})]. The contribution has a mild $W^2$-dependence but it is $Q^2$-dependent, and for large $Q^2$, the ratio $\sigma^{\scrbox{heavy}}/\sigma^{\scrbox{light}}$ gets as high as $0.3 - 0.4$ . Another ratio which we present is the ratio, of the hard to the soft contributions. Since $M_{0,L}$ is smaller than the lightest resonance mass, the AQM contribution to the longitudinal component is supressed, and therefore what we call ``soft'' contribution is actually the transverse AQM calculation. \begin{figure}[tbp] \begin{center} \epsfig{file=softhard.ps,width=140mm} \caption[]{\parbox[t]{ 0.80\textwidth}{\small \it The ratio of $\frac{\sigma^{\scrbox{soft}}}{\sigma^{\scrbox{hard}}}$ at fixed $Q^2$. } \label{softhard}} \end{center} \end{figure} In \fig{softhard} we present the ratio $\sigma^{\scrbox{soft}}/\sigma^{\scrbox{hard}}$ at fixed values of $Q^2$, where a clear power law behaviour as a function of $W^2$ can be seen. In soft processes, the hard contribution is minor, but is still present even for relatively small values of $Q^2$. The soft signature in hard processes is also present, but is smaller. For high virtualities it decreases from a few precent at low energy, to less than 1{\%\ } at high energies. \section{Conclusions} In this paper, we have shown that an approach based on Gribov's proposal, provides a successful description of the experimental data on photon-proton interaction, over a wide range of photon virtualities $0\le Q^2\le 100\,{\,\mbox{GeV}^2}$, and energies $3\le\sqrt{s}\, (= W) \le 300$ GeV\@. The key assumption on which our aproach is based, is that the non-perturbative and the perturbative QCD conributions in the Gribov formula can be separated by the parameter $M_0$. Oue successful reproduction of the experimental data (see Figs.\ \ref{stotw2} and~\ref{stotq2}) shows that this assumption appears to be valid. It further lends credence to using the additive quark model (AQM) to describe the non-perturbative contribution. The successful use of the AQM leads to an improved result presented in this paper, compared to our previous result\cite{GLMepj98}, where we found it necessary to introduce an damping factor for the AQM. The second important by-product of our calculation is the simple method we have used to separate between non-perturbative (``soft") and perturbative (``hard") contributions. We have used three separation parameters: $M_{0,T}$, $M_{0,L}$ and $\mu$. The values of these parameters, $M^2_{0,T} = 0.7 - 0.9 \,{\,\mbox{GeV}^2}$, $M^2_{0,L}\leq 0.4\,{\,\mbox{GeV}^2} $ and $\mu^2\approx 1\,{\,\mbox{GeV}^2}$, were determined by fitting to the experimental data. We believe that these values may be useful in the future, for more theoretical description of the matching between ``soft" and ``hard" processes in QCD. The third result which we find interesting, is that the GRV parameterizations of the structure functions cannnot successfully describe the experimental energy dependence of the total cross section at small values of $Q^2$ (see \fig{comparexg}), while the MRST parameterizations can. This result shows the interdependence of deep inelastic scattering data, and the theoretical description of the mathching of the ``soft" and ``hard" contributions. In addition we obtain: \begin{enumerate} \item A description of the ratio $\sigma_{\scrbox{L}}/\sigma_{\scrbox{T}}$ (see \fig{longtran}) which is in good agreement with the experimental data and other approaches. The fact that $\sigma_L$ appears to have only a ``hard" contribution should be stressed, as this could be a possible window for particular features of non-pertutbative QCD; \item A considerable contribution of the ``soft" processes at rather large values of photon virtualities $Q^2$. For example at $Q^2 = 15\,{\,\mbox{GeV}^2}$, $\sigma^{\scrbox{soft}}/\sigma^{\scrbox{hard}} \,\approx\,\,4\%$ at $W= 10 \,GeV$ (see \fig{softhard}). We also find, at low $Q^2$, a contamination of the ``soft" processes by the ``hard" ones. For high energy real photoproduction this contamination amounts to about 5{\%\ }. We believe that this fact should be taken into account when interpreting experimental data, especially as far as their energy dependance is concerned; \item The prediction of the cross section for the heavy quark production (see \fig{heavylight}) which is almost $W$ independent, while showing a steep $Q^2$ decrease. \end{enumerate} We propose a simple and successful phenomenological model for the photon-proton interaction, which provides a method for matching the ``soft" and ``hard" interactions at high energy, and which can be a guide for future theoretical approaches. It is worthwhile mentioning, that even in the present form our model can be useful for determining the initial parton distributions at leading twist, for the DGLAP evolution equations. {\bf Acknowledgments:} We thank A. Martin and A. Stasto for useful discussions of the results of this paper, and for providing us the kumac file for Fig.6. This research was supporteed in part by the Israel Science Foundation, founded by the Israel Academy of Science and Humanities.
\section{Introduction} For some considerable time, N-Body simulations have predic\-ted the existence of a vertical instability in galactic bars (Com\-bes \& Sanders \cite{cs81}, Combes et al.\ \cite{cdfp90}, Raha et al.\ \cite{rea91}). These simulations show that a bar forming in a flat disk will not remain thin, but will quickly buckle and form a thickened structure perpendicular to the plane of the disk. Viewed edge-on, such fat bars have a characteristic shape with very boxy isopho\-tes. In the most extreme cases, the structure appears double-lobed, like a peanut in its shell. These simulations are intriguing because observations of real edge-on galaxies reveal that a significant fraction have central bulges with boxy or peanut-shaped isophotes (Jarvis \cite{j86}, Shaw \cite{s87}, de Souza \& dos Anjos \cite{dd87}). It is therefore tempting to associate such bulge morphologies with edge-on bars, and even draw more general conclusions regarding the formation of all bulges. Unfortunately, establishing the link between bars and boxy bul\-ges observationally has proved difficult. Bars are recognizable only in fairly face-on galaxies, while the vertical structure of a bulge can only be deduced from an edge-on view. Thus, it has proved impossible to connect the two phenomena unequivocally using photometric data. A few years ago, we suggested that bars could be detected in edge-on galaxies from their kinematic signature (Kuijken \& Merrifield \cite{km95}). The orbits followed by material in a barred potential will be non-circular, and the arrangement of the orbits changes abruptly near resonances. As a consequence of this complexity, the observable line-of-sight velocities of material as a function of position in an edge-on barred galaxy will also display complex structure, with multiple components and gaps associated with the resonances. The existence of this complexity, which we originally investigated using simple perturbation theory applied to closed orbits, has subsequently been confirmed by full hydrodynamical gas simulations (Athanassoula \& Bureau \cite{ba99}). In a pilot spectral study, we found such kinematic structure in both the stellar and gaseous components of two edge-on disk galaxies (Kuijken \& Merrifield \cite{km95}). Since these galaxies were selected because they contained peanut-shaped bulges, this discovery provided some evidence that bars and boxy bul\-ges are the same phenomenon viewed from different directions. However, the sample size was very small, and lacked comparison data from galaxies without boxy bulges. We have therefore now carried out a larger spectral survey of edge-on galaxies. The sample was taken from the largest early-type disk galaxies observable from the northern hemisphere in the RC3 catalog (de Vaucouleurs et al.\ \cite{dvea91}), from which we selected a subset of 10 galaxies designed to span a complete range in bulge morphology, from elliptical to peanut-shaped. Section~\ref{shapesec} describes how the shapes of these bulges have been quantified, while Sect.~\ref{specsec} presents the spectral data. The results and their implications are discussed in Sect.~\ref{concsec}. \section{Bulge shapes}\label{shapesec} Previous studies of bulge shapes (e.g.\ Jarvis \cite{j86}, Shaw \cite{s87}, de Souza \& dos Anjos \cite{dd87}) have been based on the visual impression of galaxies in sky survey plates. Unfortunately, this subjective approach cannot be used reliably when comparing the bulge shapes to other properties of galaxies such as their kinematics. If, for example, we were to detect complex kinematics in a galaxy, there is a significant risk that we would then reinforce out initial prejudice by convincing ourselves that there were signs of boxiness in the galaxy's isophotes. What we require, therefore, is some more objective approach. In the case of elliptical galaxies, shapes are relatively easy to classify by measuring the minor departures of the isophotes from ellipses (e.g.\ Bender, D\"obereiner \& M\"ollenhoff \cite{bdm88}). However, the analysis of an edge-on disk galaxy is less straightforward, as the contribution to the total light from the disk, as well as dust absorption in the disk plane somewhat confuse the bulge isophotal shapes. Nevertheless, after some experiments we have found that it is possible to obtain robust measures of the bulge isophotes using the techniques developed for elliptical galaxies. Images of most of our sample galaxies' bulges were obtained in the I band at the William Herschel Telescope along with the spectral data described below. Where this was not possible, we have searched the La Palma archive, or failing this, used the Digitized Sky Survey (DSS). We have measured the bulge isophotes with the ELLIPSE task in the STSDAS analysis package of IRAF (Jedrjezewski \cite{j87}). We masked out a wedge-shaped region of each image within 12 degrees of the disk major axis to minimize the disk influence, and, in order to minimize the effects of extinction by the disk, only fitted on the side of the galaxy where the disk projects behind the bulge. The masking process leaves less than half of the isophote available for fitting, but by fixing the centroid and position angle of the isophotes to coincide with those of the disk stable results can still be obtained. Having measured the isophote shapes at a range of surface brightnesses, we then classified a galaxy's boxiness on the basis of the most extreme value (positive or negative) of the $a_4$ isophote shape parameter. Images of the half of the bulges to which the fit was applied, ordered by the value of this parameter, are presented in Fig.~\ref{nutfig}. Reassuringly, this objective ordering process arranges the galaxies in almost exactly the same sequence as one would have done by eye, but without the dangers of {\it a posteriori} bias that are inherent in any subjective classification. \begin{figure*} \resizebox{\hsize}{!}{\includegraphics{nutkinfig2grey.ps}} \caption{Montage of the sample of 10 galaxies sorted by deviation of their bulge isophotes from disky (top) to boxy (bottom). The right panels show I-band CCD images of the galaxies, except for those marked with an asterisk which were taken from the Digitized Sky Survey (DSS). The left panels show the two-dimensional [\ion{N}{ii}] emission line spectra, with position along major axis as the $x$-axis and wavelength as the $y$-axis. All panels are 80 arcsecs wide. The $a_4$ isophote shape parameter for the least elliptical bulge isophote is listed next to the galaxy name.} \label{nutfig} \end{figure*} \section{Spectral data}\label{specsec} Using the ISIS spectrograph on the William Herschel $4.2\,{\rm m}$ telescope, we obtained long-slit spectra for all the galaxies in the sample. In each case, the spectrograph slit was aligned a\-long the major axis of the galaxy (if necessary just avoiding the dust lane). The spectra were obtained using a 1200-line grating, giving a resolution of 1\AA\ (FWHM); the spectral range was centred on H$\alpha$ (6563\AA). We have previously found that the clearest signature of bar-induced peculiar kinematics comes from the emission lines in the spectra (Kuijken \& Merrifield \cite{km95}). The strongest emission line in this spectral region is H$\alpha$ itself, but it lies on top of the H$\alpha$ absorption line associated with the stellar continuum, so its signature is somewhat confused. A clearer signal comes from the neighbouring [\ion{N}{ii}] line at 6584\AA, which does not lie on top of any significant absorption features. We have therefore analyzed the spectra in the vicinity of this line by first subtracting a low-order fit to the stellar continuum, then subtracting sky spectra from the ends of the slit so as to remove any night sky lines. The resulting two-dimensional spectra -- intensity of the [\ion{N}{ii}] line as a function of position along the slit and wavelength -- are shown in Fig.~\ref{nutfig}. The wavelength scales in this figure have been adjusted so as to present all the galaxies with similar integrated line widths, thus rendering the kinematics from different galaxies more directly comparable. \section{Discussion}\label{concsec} From inspection of Fig.~\ref{nutfig}, it is apparent that there is, indeed, a link between bulge morphology and the complexity of the gas kinematics -- generally speaking, galaxies with non-boxy bulges have a simple kinematic structure, while boxier bulges seem to be associated with complex multiple-component emission line kinematics. Bureau and Freeman (\cite{bf99}, \cite{b98}) reach a similar conclusion on the basis of a sample of southern galaxies. The structure apparent in the more complex emission-line kinematics shown in Fig.~\ref{nutfig} -- X-shaped two-component systems, distorted central parallelogram structures, and skewed figures-of-eight -- are exactly the classes of feature that are generated by a non-axisymmetric barred potential when viewed from a variety of angles (Merrifield \cite{m96}, Bureau \& Athanassoula \cite{ba99}, Athanassoula \& Bureau \cite{ab99}). It is also notable that the radial scale over which the complex kinematics occurs is comparable in extent to the bulges of the host systems, again suggesting that the two phenomena are linked. However, before we conclude from the association of boxy bulges with complex kinematics that these systems are bars viewed edge-on, we should consider other possible causes of complexity in the observed kinematics. One such possibility is that the gaps in the emission-line profiles result from differential extinction by dust down the line of sight. However, in axisymmetric galaxies, the line-of-sight velocity of the gas must be a smooth function of distance down the line of sight, so it is difficult to see how partial obscuration in an unbarred galaxy could result in multiple-components in the kinematics. A second possibility is that the structure in the kinematics arises from rings of gas in the galaxy, as are seen in a number of more face-on disk systems (Buta \& Combes \cite{bc96}). In a two-dimensional spectrum of an edge-on galaxy, an axisymmetric ring of emission will appear as an inclined straight line. Superficially, some of the structures in the spectra in Fig.~\ref{nutfig} conform to this pattern. However, there are several crucial differences. First, rings are edge-brightened when seen in projection, yet most of the linear features in Fig.~\ref{nutfig} do not get brighter towards their ends. Second, axisymmetric rings project to straight lines that pass through the systemic velocity of the galaxy at its centre, whereas many of the linear features in Fig.~\ref{nutfig} are not quite straight, and have non-zero velocities at the centres of their galaxies. Such features cannot occur in an axisymmetric potential, so we are once again forced to conclude that these galaxies are barred. A third possibility, finally, is that we are seeing variations in ionization structure of the gas, and not inhomogeneities in the distribution of the gas itself. In fact, in barred galaxies we also expect such variations, induced by the shocks; and they are indeed observed as systematic variations in N[II]/H$\alpha$ ratio over the $(R,v)$ diagrams (see also Bureau \cite{b98}). A detailed analysis of the line ratio lies beyond the scope of this paper. However, as far as can be ascertained, the H$\alpha$ emission line shows identical structure to the [NII] line, providing further evidence that the structure cannot be attributed to the details of the gas' ionization state. In summary, this spectral study of edge-on galaxies quite firmly establishes the link between boxy bulges and galactic bars. However, we have only just begun to tap into the wealth of information that the spectra provide. Modelling the full complexity of spectral data such as those shown in Fig.~\ref{nutfig} should yield a wealth of information about barred galaxies, allowing us to map out their complete three-dimensional structure for the first time. \begin{acknowledgements} The Digitized Sky Survey was produced by STScI under U.S. Government grant NAG W-2166, and is subject to a variety of copyrights ({\tt http://{\rm \-}stdatu.stsci.edu/{\rm \-}dss/dss\_copy{\rm \-}right.html}). The William Herschel Telescope is operated on the island of La Palma by the Isaac Newton Group in the Spanish Observatorio del Roque de los Muchachos of the Instituto de Astrofisica de Canarias. Much of the analysis in this paper was performed using {\sc iraf}, which is distributed by NOAO, operated by AURA under cooperative agreement with the NSF{}. MRM has been supported by a PPARC Advanced Fellowship (B/94/AF/1840) \end{acknowledgements}
\section{Introduction} Since 1988, a non-extensive formalism of statistical mechanics \cite{tsallis1} has been developed and up to recent years, many works have been devoted to show the robustness and usefulness of this approach. We believe it is robust in the sense it allows generalizations of a variety of fundamental concepts of statistical thermodynamics \cite{concepts}, such that it avoid to enter in severe contradiction with well established facts. In addition, we believe it is useful in the sense it provides a theoretical basis and relevant explanation of some experimental and observational situations \cite{verify}, where Boltzmann-Gibbs statistics fails. That is to say, where Boltzmann-Gibbs statistical functions diverge and do not yield any physical prediction. However, this formalism, unlike Boltzmann-Gibbs', has a non-zero set of free parameters, here represented by $q$ ($q\in\Re$). This unique parameter controls the degree of the nonextensivity of the system in consideration. (At this point, it is worth noting that the formalism includes the standard, extensive, statistics as a special case for the value of $q=1$ and all expressions derived within this non-extensive framework give the results of Boltzmann-Gibbs statistics in the $q\rightarrow 1$ limit.) Only after 1995, some works started to address the long standing puzzle of understanding the physical meaning of $q$. Amongst the works related to this topic, two main streams started to become more apparent. On one side, there are attempts on the study of conservative \cite{hamiltonian} and dynamical systems, more precisely, dissipative systems with both, low \cite{low} and high \cite{high} dimensions. On the other, there have been some efforts of estimating bounds upon $q$ in measurable physical systems. The physical applications studied so far can be enumerated as follows: the microwave background radiation \cite{t-radiation,u-radiation}, the Stefan-Boltzmann constant \cite{p-stefan,u-stefan}, the early Universe \cite{d-early,u-early} and the primordial neutron to baryon ratio in a cosmological expanding background \cite{diego}. As a final note, a recent letter by Alemany \cite{ALEMANY} explored the definition of a new {\it fractal canonical ensemble}, associated with the parameter $q$. However, first estimations for the universe as a whole yielded to values of $q$ bigger than those allowed by nucleosynthesis \cite{diego}. In all these works, two different approaches for the quantal distribution functions have been used: a closed analytical form for them is still lacking. The asymptotic approach of the kind $\beta (1-q)\rightarrow 0$ of Tsallis et al. \cite{t-radiation} was used in Refs. \cite{t-radiation,p-stefan,d-early,diego} and also in some other applications such as the study of Bose-Einstein condensation in a fractal space \cite{curilef}, the specific heat of $^4\!$He \cite{curilefpapa} and the thermalization of an electron-phonon system \cite{koponen}. The second approach for the generalized distribution functions has been proposed by B\"{u}y\"{u}kk{\i}l{\i}\c{c} et al. \cite{buyukkilic}, which we refer to as {\it Factorization Approach}. This last term is justified in that it is generically based on the factorization of the generalized grand canonical distribution and the concomitant generalized partition function as if they were extensive quantities. However, although the generalized distribution functions within factorization approach have been derived before the asymptotic approach, they have not been prefered to be used in physical applications, mainly due to a work by Pennini et al. \cite{pennini}. There, the authors claimed that the quality of the results of the factorization approach deteriorates as the number of the particles of the system increases. But very recently, contrary to the previous belief, new results by Wang and L\'{e} M\'{e}haut\'{e} \cite{wang}, favoured the factorization approach and showed clearly that there exists a temperature interval where the ignorance of the approximation is significant, but otherwise, the results of this approach can be used with confidence no matter the number of particles. In fact, they showed that, for a macroscopic system (i.e., with $\sim 10^{23}$ particles) having two states with a small energy interval of about $10^{10}$ Hz, this forbidden temperature zone is very narrow and situated at extremely high temperatures ($\sim 10^{20}$ K) and as the number of levels increases the forbidden temperature zone shifts to higher temperatures. Therefore the results of the factorization approach can be used at low temperatures up to $10^{20}$ K without any ignorance for any physical system under consideration \cite{wang,wang2,wang3}. As it was expected, the study of Wang and L\'{e} M\'{e}haut\'{e} started to accelerate new attempts of applying these distribution functions to physical systems in order to estimate some alternative bounds upon $q$ \cite{u-stefan,u-early}; being this approach more handable than the one advanced by Tsallis et al. Although, after Wang and L\'{e} M\'{e}haut\'{e} \cite{wang}, some physical systems have been worked out with the help of the generalized distribution functions of the factorization approach, general expressions for magnitudes of fermions and bosons are still lacking. Indeed, the possible need for a nonextensive formalism of thermostatistics is clear for a long time in gravitation \cite{gravity}, magnetic systems \cite{magnet}, L\'{e}vy-like diffusions \cite{levy-like}, some surface-tension problems \cite{surface}, etc., since Boltzmann-Gibbs formalism is known to fail whenever the physical system under consideration includes (i) long-range interactions and/or (ii) long-memory effects and/or (iii) the system evolves in a multifractal-like space-time. The last property can be understood as the system evolves in a porous-like medium where the properties of the multifractal structures govern it. In the light of the facts stated above, we address the study of the generalized quantal distribution functions of fermions and bosons in a non-Euclidean, fractal-like space-time and to obtain general results for magnitudes, such as the number of particles and the internal energy that could be used in general for any application to generalized systems. Once these general formulae be established, we proceed further investigating how some simple constructs of the statistical theory are affected by the change of distribution functions, for instance, the Chandrasekhar mass and the Bose-Einstein condensation. A similar work studying the Bose-Einstein condensation was recently done, as commented above, by Curilef \cite{curilef} and our results are compatible in this case. Finally, it is worth noting that, for non-interacting fermions and bosons, the information about the fractal structure is kept in the nonextensivity parameter $q$, which is by now verified to be connected with the fractal dimension and the multifractal singularity spectrum \cite{low}. In a sense, we shall study if any correction to the usual number of particles and energy arises due to the change of the distribution functions. \section{The fermion case} Let us start by writing down the generalized distribution function for fermions, up to $(1-q)$ order. It is given by \cite{buyukkilic}, \begin{equation} n_{q[fermions]} = \frac{1}{e^{\beta(\epsilon-\mu)} + 1} + \frac{q-1}{2} \frac{\left[\beta(\epsilon-\mu)\right]^2 e^{\beta(\epsilon-\mu)}}{\left(e^{\beta(\epsilon-\mu)} + 1\right)^2} , \end{equation} where $\beta=1/kT$, $\epsilon$ is the energy level and $\mu$ is the chemical potential. We are interested in solving the following integrals. For the number of particles, \begin{equation} \label{total} <N>=\frac {4\pi V (2m)^{3/2}}{h^3} \left[ \int_0^\infty d\epsilon \frac{\epsilon^{1/2}}{e^{\beta (\epsilon - \mu)} +1} + \frac {q-1}{2} \int_0^\infty d\epsilon \frac{ (\beta(\epsilon - \mu ))^2 e^{\beta (\epsilon - \mu)} \epsilon^{1/2}}{(e^{\beta (\epsilon - \mu)} +1)^2} \right] \end{equation} and a similar one, replacing the factors $\epsilon^{1/2}$ for $\epsilon^{3/2}$, for the energy $<U>$. Note that we have assumed the same pre-factor for both terms. This can be justified if one thinks of this factor as coming from microscopic quantum considerations and a change of variables, from momentum to energy, inside the integrals. See for instance, section 2.4 of Ref. \cite{PATHRIA}. Then, we can split $<N>$ into two parts. Firstly we have the usual expression, given by the first term in the rhs of (\ref{total}), \begin{eqnarray} Usual=\frac {4\pi V (2m)^{3/2}}{h^3} \int_0^\infty d\epsilon \frac{\epsilon^{1/2}}{e^{\beta (\epsilon - \mu)} +1} =\frac {8\pi V (2m\mu)^{3/2}}{3h^3} \nonumber \\ \hspace{3cm}\left[ 1 + \frac 18 \left( \frac{\pi^2}{\beta \mu} \right)^2 + \frac{7}{640} \left(\frac{\pi}{\beta \mu} \right)^4 + \ldots \right]. \end{eqnarray} And we also have a second term, \begin{equation} I_q=\frac {4\pi V (2m)^{3/2}}{h^3} \frac {q-1}{2} \int_0^\infty d\epsilon \frac{ (\beta(\epsilon - \mu ))^2 e^{\beta (\epsilon - \mu)} \epsilon^{1/2}}{(e^{\beta (\epsilon - \mu)} +1)^2} = \frac {4\pi V (2m)^{3/2}}{h^3}\frac {q-1}{2} I . \end{equation} $I$ can be recasted using the following dimensionless variables: \begin{equation} x=\beta \epsilon, \hspace{2cm} \psi=\beta \mu. \end{equation} This enables us to write \begin{equation} I= \beta^{-3/2} \int_0^\infty \frac{(x-\psi)^2 e^{x-\psi} x^{1/2}} {(e^{x-\psi} +1)^2} dx = \beta^{-3/2} I_x. \end{equation} In order to solve this last $I_x$ we apply the following trick. We know that integrals of the form \begin{equation} \label{int} \int_0^\infty \frac{f(x)}{e^{x-\psi} +1 } dx \end{equation} have a solution, \begin{equation} \label{serie} \int_0^\psi f(x) dx + \frac{\pi^2}{6} \frac {df}{dx} + \frac {7 \pi^4}{360} \frac{d^3 f}{dx^3} + \frac {31 \pi^6}{15120} \frac{d^5f}{dx^5} + \ldots , \end{equation} where the derivatives have to be evaluated in $x=\psi$ \cite{PATHRIA}. In fact, this was already used to write down the solution for the usual term of (\ref{total}). The idea is then to introduce an extra parameter, $l$, in the integral (\ref{int}) and consider, \begin{equation} \label{int-l} \int_0^\infty \frac{f(x)}{e^{x-l\psi} +1 } dx . \end{equation} Here, it is important to stress that $f(x)$ does not depend on $l$. If we derive inside this integral with respect to $l$, the result will be \begin{equation} \label{int-f} \int_0^\infty \frac{f(x) \psi e^{x-l\psi} }{(e^{x-l\psi} +1)^2 } dx , \end{equation} and if $l=1$, it can reproduce our integral $I_x$ iff the function $f(x)$ is given by, \begin{equation} \label{def} f(x)= \frac{(x-\psi)^2 }{\psi} x^{1/2} . \end{equation} Thus, using (\ref{def}), the solution to $I_x$ is, \begin{equation} I_x= \left[ \frac {d}{dl} \int_0^\infty \frac{f(x)}{(e^{x-l\psi} +1) } dx \right]_{l=1}. \end{equation} We have to be very careful in using the correct formula to get the integral in $f(x)$: we should redefine $\tilde \psi=l\psi$ and apply then (\ref{serie}). So, the mechanism is, considering $f(x)$ as in (\ref{def}), compute the integral (\ref{int-l}), derive the result with respect to $l$ and finally evaluate in $l=1$. That yields, as we have seen, the integral we need.\footnote{A more transparent way of solving these integrals (see acknowledgments) is to take the integral (6) as \begin{equation} \frac{\partial}{\partial \psi} \int_0^{\infty} \frac{x^{5/2}} {e^{x-\psi} +1} dx - 2 \psi \frac{\partial}{\partial \psi} \int_0^{\infty} \frac{x^{3/2}}{e^{x-\psi} +1} dx + \psi^2 \frac{\partial}{\partial\psi} \int_0^{\infty} \frac{x^{1/2}} {e^{x-\psi} +1} dx \end{equation} and apply expansion (8) to each part. As we see, in fact there is no need to introduce such an extra parameter $l$, and it only stands as a mathematical trick in a particular form of solving the integrals. Numerical results for the coefficients obtained with both methods are the same. We warn the reader not to be distracted for such a trick.} Let treat each of the term in (\ref{serie}) separately. The first term is obtained inmediately: \begin{equation} \int_0^{\tilde \psi} \frac{(x-\psi)^2 }{\psi} x^{1/2} dx = \frac 1\psi \int_0^{\tilde \psi} (x^2 -2 x \psi + \psi^2) x^{1/2} dx, \end{equation} which finally leads to, \begin{equation} 1^{st}\;\;term\;=\;\frac 1\psi \left[ \frac{\tilde \psi^{7/2}}{7/2} - 2 \psi \frac{\tilde \psi^{5/2}}{5/2} + \psi^2 \frac{\tilde \psi^{3/2}}{3/2} \right]. \end{equation} Now we have to derive with respect to $l$ (recall that $\tilde \psi = l \psi$), obtaining \begin{equation} \frac{d\; 1^{st}\;\;term}{dl} = \frac 1\psi \left[ (7/2) l^{5/2} \frac{ \psi^{7/2}}{(7/2)} - 2 (5/2) l^{3/2} \frac{ \psi^{7/2}}{(5/2)} + (3/2) l^{1/2} \frac{\psi^{7/2}}{(3/2)} \right] , \end{equation} and evaluate at $l=1$. The result is, \begin{equation} \left[ \frac{d\; 1^{st}\;\;term}{dl} \right]_{l=1}=0. \end{equation} The second term in (\ref{serie}) is obtained as, \begin{equation} 2^{nd}\;\;term\;= \; \frac{\pi^2}{6} \left[ 2\frac{(x-\psi)}{\psi} x^{1/2} + \frac{(x-\psi)^2}{\psi} \frac 12 x^{-1/2} \right]_{\tilde \psi}, \end{equation} which finally yields to, \begin{equation} 2^{nd}\;\;term\;= \; \frac{\pi^2}{6} \left[ 2\frac{(l\psi-\psi)}{\psi} (l\psi)^{1/2} + \frac{(l\psi-\psi)^2}{\psi} \frac 12 (l\psi)^{-1/2} \right]. \end{equation} The derivative with respect to $l$ is, \begin{eqnarray} \frac{6}{\pi^2} \frac{d\; 2^{nd}\;\;term}{dl} = 2 (l\psi)^{1/2} + 2(l\psi-\psi) \frac 12 (l\psi)^{-1/2} + 2 (l \psi -\psi) \frac 12 (l\psi)^{-1/2} \nonumber \\ \;\;\;\;\;\;+ (l\psi-\psi)^2 \frac 12 l^{-3/2} \frac{-1}{2} \psi^{-3/2} . \end{eqnarray} Evaluating in $l=1$ we obtain, \begin{equation} \left[ \frac{d\; 2^{nd}\;\;term}{dl} \right]_{l=1}= \frac{\pi^2}{3} \psi^{1/2}. \end{equation} One can do the same with the third term. In this case we need to compute the third derivate. It is given by, \begin{equation} f^{iii}= \frac 3\psi x^{-1/2} - \frac{3}{2} (x-\psi) \frac{x^{-3/2}}{\psi} + (x- \psi)^2 \frac 38 \frac {x^{-5/2}}{\psi}. \end{equation} Computing this last for $x=\tilde \psi$ and deriving with respect to $l$ we get, \begin{equation} \frac{df^{iii}}{dl} = -3 (l\psi)^{-3/2} + 3 (l\psi- \psi) (l\psi )^{-5/2} + (l\psi- \psi)^2 \left(\frac{-15}{16}\right) (l\psi)^{-7/2}. \end{equation} And finally making $l=1$, the correction is found to be, \begin{equation} \left[ \frac{df^{iii}}{dl} \right]_{l=1} = -3 (\psi)^{-3/2}. \end{equation} Now we can go back and collect the results for $I$, which results, \begin{equation} I=\beta^{-3/2} I_x=\beta^{-3/2} \left( 0+ \frac{\pi^2}{3} \psi^{1/2} - \frac{7 \pi^4}{120} \psi^{-3/2} + \ldots \right), \end{equation} and the final expression for $<N>$ is then, \begin{eqnarray} <N> \frac{3h^3}{8\pi V (2m\mu)^{3/2}}= \left( 1 + \frac 18 \left( \frac{\pi^2}{\beta \mu} \right)^2 + \frac{7}{640} \left(\frac{\pi}{\beta \mu} \right)^4 + \ldots \right) \nonumber \\ +\frac {q-1}{2} \left( 0 + \frac{\pi^2}{2} \frac{1}{\beta \mu} - \frac{21 \pi^4}{240} \frac{1}{(\beta \mu)^3} + \ldots \right). \end{eqnarray} This final expression represents the correction terms due to non-extensivity that arise for $<N>$. We can do exactly the same for the energy $<U>$. $I_q$ is given now by, \begin{equation} I_q=\frac {4\pi V (2m)^{3/2}}{h^3}\frac {q-1}{2} \beta^{-5/2} I_x , \end{equation} where $I_x$ is, \begin{equation} I_x= \int_0^\infty \frac{ (x-\psi)^2 e^{x-\psi} x^{3/2}}{(e^{x-\psi} +1 )^2} dx . \end{equation} We apply again the same trick, in this case, with a function $g(x)= (1/\psi) (x-\psi)^2 x^{3/2}$. As before, \begin{equation} I_x= \left[ \frac {d}{dl} \int_0^\infty \frac{g(x)}{(e^{x-l\psi} +1) } dx \right]_{l=1}. \end{equation} The first term in the serie (\ref{serie}), is again equal to zero, as can be directly verified by computing the integral for $g$, deriving the result with respect to $l$ and evaluating at $l=1$. The second term needs \begin{equation} \frac{dg}{dx}= \frac{1}{\psi} \left[ 2(x-\psi) x^{3/2} + (x-\psi)^2 \frac 32 x^{1/2} \right]. \end{equation} Derivating this with respect to $l$ and evaluating at $l=1$ we obtain for the correction, \begin{equation} \frac{\pi^2}{3} \psi^{3/2}. \end{equation} The third term needs the third derivative of $g$. This is found to be, \begin{equation} \frac 1\psi \left[ 9 x^{1/2} + x^{-1/2} (x-\psi) \frac 92 - (x-\psi)^2 \frac 38 x^{-3/2} \right]. \end{equation} Again deriving with respect to $l$ and evaluating at $l=1$, it yields the following correction \begin{equation} \frac {7\pi^4}{360} 9 \psi^{-1/2}. \end{equation} Collecting the results, we get for $<U>$, \begin{eqnarray} <U> \frac{5h^3}{8\pi V (2m)^{3/2} \mu^{5/2}} = \left( 1 + \frac 58 \left( \frac{\pi^2}{\beta \mu} \right)^2 - \frac{7}{384} \left(\frac{\pi}{\beta \mu} \right)^4 + \ldots \right) \nonumber \\ \;\;\;\;\;\;+ \frac {q-1}{2} \left( 0 + \frac{5\pi^2}{6} \frac{1}{\beta \mu} + \frac{315 \pi^4}{720} \frac{1}{(\beta \mu)^3} + \ldots \right). \end{eqnarray} As above, this expression represents the correction due to non-extensivity that arise for the energy. It is important to note that the {\it $T \rightarrow 0$ limit is attained without any correction} thus suggesting that {\it non-extensivity can hardly have any role for very low temperatures.} One can now study particular physical systems. For instance, one can inmediately realize that the Chandrasekhar mass will be not modified at all in passing to a non-extensive context, at least up to order $(1-q)$. This is simple because the Chandresekhar mass is a construct that arises at $T=0$ where both corrections, to $<N>$ and to $<U>$, are exactly zero. However, if non-extensive statistics is concerned in the study of stellar structures, corrections will unavoidably arise. \section{The boson case} Now we analize the boson generalized distribution function, namely, \begin{equation} \label{bos-n} n_{q[bosons]} = \frac{1}{e^{\beta(\epsilon-\mu)} - 1} + \frac{q-1}{2} \frac{\left[\beta(\epsilon-\mu)\right]^2 e^{\beta(\epsilon-\mu)}}{\left(e^{\beta(\epsilon-\mu)} - 1\right)^2} , \end{equation} where again $\beta=1/kT$, $\epsilon$ is the energy level, $\mu$ is the chemical potential. Then, using equation (\ref{bos-n}), one can obtain the average number of particles as \begin{equation} \label{bos-int} \frac{\left(N\right)_q}{V} = \frac{2\pi (2m)^{3/2}}{h^3} \left[\int_0^{\infty}\frac{\epsilon^{1/2} d\epsilon} {e^{\beta(\epsilon-\mu)} -1} + \frac{q-1}{2} \int_0^{\infty} \frac{\beta^2(\epsilon-\mu)^2 e^{\beta(\epsilon-\mu)} \epsilon^{1/2} d\epsilon}{\left(e^{\beta(\epsilon-\mu)} - 1\right)^2} \right]. \end{equation} As in the standard case, we have to separate the state with $\epsilon =0$, which has zero weight in the integral (\ref{bos-int}). For this level of energy, the distribution function yields, \begin{equation} N_q(\epsilon=0)=\frac{z}{1-z}\left[1+\frac{q-1}{2}\frac{(\ln z)^2}{1-z}\right] \end{equation} where $z=e^{\beta\mu}$ is the fugacity of the gas. In the case $z \ll 1$, the correction term goes to zero as $z(\ln z)^2$ does. If $z\rightarrow 1$, the correction goes to 1 as $z$, and results are neglectable in comparison with the first, diverging, term. This can be seen in Fig. 1. Substitution of $x=\beta\epsilon$ and $\psi=\beta\mu$ in equation (\ref{bos-int}) yields, \begin{equation} \label{bos-int-2} \frac{\left(N_e\right)_q}{V} = \frac{2\pi (2mk)^{3/2}}{h^3} T^{3/2} \left[ I_1 +\frac{q-1}{2} I_q \right] \end{equation} where $\left(N_e\right)_q$ stands for the number of particles in the excited states ($\epsilon\neq0$), and $I_1$ and $I_q$ are defined to be \begin{equation} I_1 = \int_0^{\infty}\frac{x^{1/2} dx}{e^{x-\psi} -1}, \;\;\;\;\; \;\;\;\;\; I_q = \int_0^{\infty}\frac{(x-\psi)^2 e^{x-\psi} x^{1/2} dx} {\left(e^{x-\psi} -1\right)^2} . \end{equation} The integral $I_1$ is the one which appears in the standard, $q=1$ case of the boson gas and therefore the solution of it is the known one \cite{PATHRIA}, \begin{equation} I_1 = \Gamma(3/2) g_{3/2}(z) , \end{equation} where $g_{n}(z)=\frac{1}{\Gamma (n)} \int_0^\infty \frac{x^{n-1}} {z^{-1}e^x -1} dx \simeq \sum_{s=1}^{\infty}\left(z^s/s^n\right)$ for small $z$ and $\Gamma$ is the usual Gamma function, $\Gamma(n)=\int_0^\infty e^{-x} x^{n-1} dx$. For $0\leq z \leq 1$ and $\forall n, n>1$, the functions $g_n(z)$ are bounded by the Riemann zeta functions, which yields for all $z$ values of interest, \begin{equation} I_1 \leq \Gamma(3/2) \zeta(3/2) . \end{equation} On the other hand, the integral $I_q$ is again somewhat cumbersome to solve, but after some algebra, similar to what we did in the previous section, we managed to find the analytical solution as \begin{equation} I_q = \Gamma(7/2) g_{5/2}(z) - 2\psi \Gamma(5/2) g_{3/2}(z) + \Gamma(3/2) \psi^2 g_{1/2}(z) . \end{equation} To get the previous result one has to take into account the known relationship between the $g_n(z)$ and its derivatives, \begin{equation} g_{n-1}(z)=z\frac{\partial}{\partial z} g_n(z)= \frac{\partial}{\partial (\ln z)}g_n(z). \end{equation} Recalling again the definitions $\psi$ and $z$, we can write $\psi=\ln z$. Putting these solutions of the integrals in equation (\ref{bos-int-2}), one can obtain \begin{eqnarray} \frac{\left(N_e\right)_q}{V} = \frac{2\pi (2mk)^{3/2}}{h^3} T^{3/2}\times \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\; \nonumber \\ \left\{\Gamma(3/2) g_{3/2}(z) + \frac{q-1}{2} \left[ \Gamma(7/2) g_{5/2}(z) - 2 \Gamma(5/2) \ln z g_{3/2}(z) + \Gamma(3/2) (\ln z)^2 g_{1/2}(z) \right] \right\}, \end{eqnarray} which is the $q$-dependent solution for the number of particles in the excited states of boson systems, including the standard case as a special one if $q=1$. Like in the standard case, let us study each term of the correction in a separate fashion in order to find the bounded form of them. The first term is similar to $I_1$ and hence it is bounded by $\Gamma(7/2)\zeta(5/2)$. The second and the third terms are different from the first one in the sense that they do not take their largest values at $z=1$, but instead, both of them tend to $0$ when $z\rightarrow 0$ and $z\rightarrow 1$. This behavior can be seen in Fig. 2. The largest values are situated at $z_{max1}\simeq 0.447$ for the second term and $z_{max2}\simeq 0.175$ for the third. Consequently, if all these bounds are put together, then the total number of particles in all excited states is also bounded by, \begin{eqnarray} \left(N_e\right)_q \leq V \frac{2\pi (2mk)^{3/2}}{h^3} T^{3/2} \left\{ \Gamma(3/2)\zeta(3/2) + \frac{q-1}{2} \times \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\; \right. \nonumber\\ \left. \left[\Gamma(7/2)\zeta(5/2) - 2\Gamma(5/2) \ln (0.447) g_{3/2}(0.447) + \Gamma(3/2) (\ln (0.175))^2 g_{1/2}(0.175) \right] \right\} \end{eqnarray} which gives \begin{equation} \left(N_e\right)_q \leq V \frac{2\pi (2mk)^{3/2}}{h^3} T^{3/2} \left\{2.315 + (q-1) 3.079 \right\} . \end{equation} If we now concentrate on Bose-Einstein condensation, then the condition for the appearance of it can be expressed as \begin{equation} N > \left(N_e\right)_q . \end{equation} Alternatively, with constant $N$ and $V$, this condition can be recasted in the form, \begin{equation} T < \left(T_c\right)_q =\frac{h^2}{\left(2\pi\right)^{2/3} 2mk} \left\{\frac{N}{V\left[2.315+(q-1)3.079\right]}\right\}^{2/3}, \end{equation} or up to order 1 in $(q-1)$, \begin{equation} T < \left(T_c\right)_q= \frac{h^2}{\left(2\pi\right)^{2/3} 2mk} \left(\frac{N}{2.315 V} \right)^{2/3} \left( 1 + (q-1) 0.886 \right), \end{equation} where $\left(T_c\right)_q$ is the $q$-dependent characteristic temperature of the Bose-Einstein condensation. It is easily seen that this result shows that the critical temperature decreases when $q<1$, which is consistent with the previous result of Curilef \cite{curilef}. Note that the standard $\left(T_c\right)_1$ case can easily be obtained for $q=1$ value in the above expression. Any accurate simultaneous determination of $N$, $V$ and $T$ can yield, in principle, a bound upon $q$. However in practice, this could be well below any practical possibility, due to the smallness of the correction term. \section{Conclusions} Our main results can be summarized as follows. We have been able to solve, using the generalized quantal distribution functions in the factorization approach, the values for the average number of particles and energy in the case of system of non-interacting particles, either fermions or bosons. These results are expected to be useful in any analysis of statistical phenomena within the context of non-extensive scenarios. We could explicitly see that all the terms coming from the non-extensive part of the integrals go to zero when the temperature goes to zero, something that could be expected due to the form of $n_q$. Non-extensivity can not play a role for very low temperatures, at least up to order $(1-q)$. As a consequence, for instance, the well known Chandrasekhar limit for white dwarfs stars is not affected by a change of the statistical framework. However, it is to be explicitly stated that any other model of star, that happens with $T \neq 0$, will be affected by such change. This is why, for example, this statistical scenario could be useful to tackle the solar neutrino problem \cite{solar}. In the boson case, we have seen that a small correction appears to the Bose-Einstein condensation, this happening not exactly at $T=0$. If this correction is enough to work out a possible bound upon $q$ in this system remains to be studied. In the paper by Curilef it was argued that due to a possible fractility of the universe, distribution functions could be modified in the sense described here, and that this could be useful to study the behavior of diffuse gas clouds. This could be an example of where the $q \neq 1$ statistics may arise, although due to the form in which equilibrium distributions of boson stars arise it is unlikely that this could be applied for those cases. \subsection*{Acknowledgments} During the course of this research, D.F.T. was a Chevening Scholar and has been supported by the British Council (UK) and CONICET (ARGENTINA). He would like to thank R. Borzi for valuable advice. 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. We thank an anonymous referee for kindly suggesting the alternative method to solve the integrals commented above, among other useful insights.
\section{Introduction} It has been known for some time that the renormalization of light-front field theories is an extremely challenging task, although systematic procedures have been developed over the years to handle these problems consistently \cite{wilson,wegner,wilson2,pinsky,perry,perryboost,stan}, including boost-invariant regularization schemes \cite{perry,perryboost,stan}. Relatively recently, it was observed that a different boost-invariant regularization scheme may be employed in a treatment of light-front matrix field theories with $\phi^3$ interactions \cite{fa}, following earlier work on the small-$x$ behavior of light-cone wave functions \cite{abd,dalley}. In this work, we follow the approach suggested in \cite{fa} to determine the effective two-loop mass renormalization of the $\phi^4$ light-front Hamiltonian. In Section \ref{themodel}, we begin by presenting the light-front formulation of scalar $\phi^4$ field theory in $D \geq 2$ dimensions, and its subsequent quantization via commutation relations. In Section \ref{boundstate}, we employ a manifestly boost-invariant procedure to extract the effective two-loop mass renormalization of the light-front Hamiltonian. This result will be compared to a Feynman diagram calculation that can be performed straightforwardly in two dimensions. A brief discussion of our results will appear in Section \ref{the-end}. \section{Scalar $\phi^4$ Theory in Light-Cone Coordinates} \label{themodel} Consider the $D$-dimensional field theory described by the action \begin{equation} S= \int d^D{\bf x} \hspace{1mm}\left[ \frac{1}{2} \partial_{\mu} \phi \cdot \partial^{\mu} \phi - \frac{1}{2}m^2 \phi^2 - \frac{\lambda} {4!} \phi^4 \right], \label{phi4model} \end{equation} where $\phi({\bf x})$ is a scalar field defined on the $D$-dimensional Minkowski space ${\bf x} \equiv (x^0,x^1,\dots,x^{D-1})$. Working in the light-cone coordinate frame \begin{eqnarray} x^+ & = & \frac{1}{\sqrt{2}}(x^0 + x^{D-1}), \hspace{10mm} \mbox{``time coordinate''} \\ x^- & = & \frac{1}{\sqrt{2}}(x^0 - x^{D-1}), \hspace{10mm} \mbox{``longitudinal space coordinate''} \\ {\bf x}^{\perp} & = & (x^1,\dots,x^{D-2}), \hspace{10mm} \mbox{``transverse coordinates''} \end{eqnarray} one may derive from the light-cone energy-momentum tensor expressions for the light-cone Hamiltonian $P^-$ and conserved total momenta $(P^+, {\bf P}^{\perp})$: \begin{eqnarray} P^+ & = & \int dx^- d{\bf x}^{\perp} \hspace{1mm} (\partial_- \phi)^2, \hspace{43mm} \mbox{``longitudinal momentum''} \label{Pplus} \\ {\bf P}^{\perp} & = & \int dx^- d{\bf x}^{\perp} \hspace{1mm} \partial_- \phi \cdot \mbox{\boldmath$\partial$}^{\perp} \phi, \hspace{39mm} \mbox{``transverse momentum''} \label{Ptrans} \\ P^- & = & \int dx^- d{\bf x}^{\perp} \hspace{1mm} \left[ \frac{1}{2} m^2 \phi^2 - \frac{1}{2} \mbox{ \boldmath $\partial$}_{\perp} \phi \cdot \mbox{ \boldmath $\partial$}^{\perp} \phi + \frac{\lambda}{4!} \phi^4 \right], \hspace{1mm} \mbox{``light-cone Hamiltonian''} \label{Pminus} \end{eqnarray} where we adopt the notation $\mbox{ \boldmath $\partial$}_{\perp} \phi \equiv (\partial_1 \phi,\dots , \partial_{D-2} \phi)$. \medskip Light-cone quantization of the $\phi^4$ theory is performed in the usual way -- namely, we impose commutation relations at some fixed light-cone time ($x^+ = 0$, say): \begin{equation} [\phi(x^-,{\bf x}^{\perp}), \partial_- \phi(y^-,{\bf y}^{\perp}) ] = \frac{{\rm i}}{2} \delta (x^- - y^-) \delta ( {\bf x}^{\perp} - {\bf y}^{\perp} ). \label{commphi3} \end{equation} The light-cone Hamiltonian $P^-$ propagates a given field configuration in light-cone time $x^+$ while preserving this quantization condition. At fixed $x^+ = 0$, the Fourier representation \begin{eqnarray} \lefteqn{\phi(x^-,{\bf x}^{\perp}) = \frac{1}{(\sqrt{2\pi})^{D-1}} \int_0^{\infty} \frac{dk^+}{\sqrt{2 k^+}} \int d{\bf k}^{\perp} \times } \nonumber \\ & & \left[ a(k^+,{\bf k}^{\perp})e^{-{\rm i}(k^+ x^- - {\bf k}^{\perp} \cdot {\bf x}^{\perp})} + a^{\dagger}(k^+,{\bf k}^{\perp})e^{+{\rm i} (k^+ x^- - {\bf k}^{\perp} \cdot {\bf x}^{\perp})} \right], \label{phi} \end{eqnarray} together with the quantization condition (\ref{commphi3}), imply the relation \begin{equation} [ a(k^+,{\bf k}^{\perp}), a^{\dagger}({\tilde k}^+,{\tilde {\bf k}}^{\perp})] = \delta (k^+ - {\tilde k}^+) \delta ( {\bf k}^{\perp} - {\tilde {\bf k}}^{\perp} ). \end{equation} It is now a matter of substituting the Fourier representation (\ref{phi}) for the quantized matrix field $\phi$ into definitions (\ref{Pminus}), (\ref{Ptrans}) and (\ref{Pplus}), to obtain the following quantized expressions for the light-cone Hamiltonian and conserved total momenta: \begin{eqnarray} \lefteqn{ : P^+ : \hspace{3mm} = \int_0^{\infty} dk^+ \int d{\bf k}^{\perp} \hspace{1mm} k^+ \cdot a^{\dagger}(k^+,{\bf k}^{\perp}) a(k^+,{\bf k}^{\perp}), } \\ \lefteqn{ : {\bf P}^{\perp} : \hspace{3mm} = \int_0^{\infty} dk^+ \int d{\bf k}^{\perp} \hspace{1mm} {\bf k}^{\perp} \cdot a^{\dagger}(k^+,{\bf k}^{\perp}) a(k^+,{\bf k}^{\perp}), } \\ \lefteqn{ :P^-: \hspace{3mm} = \int_0^{\infty} dk^+ \int d{\bf k}^{\perp} \left( \frac{m^2 + |{\bf k}^{\perp}|^2}{2 k^+} \right) a^{\dagger}(k^+,{\bf k}^{\perp}) a(k^+,{\bf k}^{\perp}) } \nonumber \\ & + & \frac{\lambda}{ 4! \cdot 4(2\pi)^{D-1}} \int_0^{\infty} \frac{ dk_1^+ dk_2^+ dk_3^+ dk_4^+}{\sqrt{k_1^+ k_2^+ k_3^+ k_4^+ }} \int d{\bf k}_1^{\perp} d{\bf k}_2^{\perp} d{\bf k}_3^{\perp} d{\bf k}_4^{\perp} \times \nonumber \\ & & \left\{ \frac{}{} 4 \cdot \delta ({\bf k}_1 + {\bf k}_2 + {\bf k}_3 - {\bf k}_4) \cdot a^{\dagger}({\bf k}_1)a^{\dagger}({\bf k}_2)a^{\dagger}({\bf k}_3) a({\bf k}_4) \right. \nonumber \\ & + & 6 \cdot \delta ({\bf k}_1 + {\bf k}_2 - {\bf k}_3 - {\bf k}_4) \cdot a^{\dagger}({\bf k}_3)a^{\dagger}({\bf k}_4)a({\bf k}_1) a({\bf k}_2) \nonumber \\ & + & 4 \cdot \delta ({\bf k}_1 + {\bf k}_2 + {\bf k}_3 - {\bf k}_4) \cdot a^{\dagger}({\bf k}_4)a({\bf k}_1)a({\bf k}_2) a({\bf k}_3) \left. \frac{}{} \right\}, \end{eqnarray} where ${\bf k} \equiv (k^+,{\bf k}^{\perp})$. \section{Two-Loop Mass Renormalization} \label{boundstate} We are now interested in calculating the effective two-loop contribution to the mass term in the light-cone Hamiltonian, which is represented diagrammatically in Fig \ref{twoloop}. One advantage of working in light-cone coordinates is the absence of tadpole diagrams. i.e. the only two-loop contribution to the mass is given by the `setting sun' diagram below. \begin{figure}[h] \begin{center} \mbox{\psfig{figure=scalarIII.ps}} \end{center} \caption{`Setting-sun' diagram in scalar $\phi^4$ theory. \label{twoloop} } \end{figure} A convenient strategy is to compute this diagram for vanishing (external) longitudinal momentum $k^+ \rightarrow 0$. The details of this method appear in an earlier article \cite{fa}, so we will mention only the key ideas here. First, note that the two-loop process can be obtained by two applications of the light-cone Hamiltonian; one application corresponds to the creation of three partons from a single parton via the interaction $a^{\dagger}a^{\dagger}a^{\dagger}a$, while the subsequent interaction $a^{\dagger}aaa$ will absorb these same three partons to leave one parton left over. During the intermediate step, where there are three partons in flight, the sum of the light-cone momenta must be equal to the initial parton momentum, which we write as $(k^+,{\bf k}^{\perp})$. We now consider what happens in Fig \ref{twoloop}. when the incoming longitudinal momentum $k^+$ is made vanishingly small by taking the limit $k^+ \rightarrow 0$. It is important to stress that in this limit, we never set $k^+$ identically to zero; we simply investigate the limiting contribution of the diagram as $k^+$ -- which is always positive -- is made arbitrarily small. A detailed calculation paralleling the one carried out in \cite{fa} -- in which one studies the bound state integral equations governing the light-cone wave functions in the limit of vanishing $k^+$ -- may be used to calculate the limiting contribution of this diagram to the mass renormalization. The result is \begin{eqnarray} \lefteqn{\Gamma ({\bf k}^{\perp}) = -\frac{\lambda^2}{4! (2 \pi)^{2 D-2}} \int_0^1 d\alpha d\beta d\gamma \hspace{1mm} \delta(\alpha+\beta+\gamma - 1) \times } & & \nonumber \\ & & \int d{\bf k}_1^{\perp}d{\bf k}_2^{\perp}d{\bf k}_3^{\perp} \hspace{1mm} \delta({\bf k}_1^{\perp} + {\bf k}_2^{\perp} + {\bf k}_3^{\perp} - {\bf k}^{\perp}) \times \nonumber \\ & & \frac{1}{\beta \gamma (m^2+|{\bf k}_1^{\perp}|^2) + \alpha \gamma (m^2+|{\bf k}_2^{\perp}|^2) + \alpha \beta (m^2+|{\bf k}_3^{\perp}|^2)}. \label{gamma} \end{eqnarray} This quantity depends on ${\bf k}^\perp$ only, since any dependence on $k^+$ was eliminated after taking the limit $k^+ \rightarrow 0$. In order to find the renormalized mass, we set ${\bf k}^{\perp} = {\bf 0}$. Some remarks are in order. Firstly, note that a remnant of the longitudinal integration survives through the variables $\alpha, \beta$ and $\gamma$. These quantities represent the {\em fractions} of the external longitudinal momentum $k^+$, and are thus integrated over the interval $(0,1)$. They are not eliminated in the limit $k^+ \rightarrow 0$. To further illuminate the significance of these variables, we consider the special case of $1+1$ dimensions (i.e. $D=2$). In this case, the mass renormalization is \begin{equation} -\frac{\lambda^2}{4! (2 \pi)^2} \int_0^1 d\alpha d\beta d\gamma \hspace{1mm} \delta(\alpha+\beta+\gamma - 1) \frac{1}{ m^2(\beta \gamma+\alpha \gamma+\alpha \beta)}. \label{mass2D} \end{equation} Now consider the Feynman integral for the `setting sun' diagram of scalar $\phi^4$ theory in $1+1$ dimensions, with zero external momenta \cite{jens}. In this case, the mass renormalization turns out to be \begin{equation} -\frac{\lambda^2}{96 \pi^4} \int \frac{d^2 {\bf p}_1 d^2 {\bf p}_2} {(p_1^2+m^2)(p_2^2+m^2)[({\bf p}_1+{\bf p}_2)^2+m^2]}. \label{feynman} \end{equation} If we use Feynman parameters to rewrite this last integral \cite{collins}, and perform the momentum integration, we end up with expression (\ref{mass2D}). i.e. the integrals (\ref{gamma}) and (\ref{feynman}) are equal in $1+1$ dimensions. It would be interesting to test for equality in $D \geq 3$ dimensions, and we leave this for future work. Interestingly, the parameters $\alpha,\beta,\gamma$ above -- which correspond to fractions of longitudinal light-cone momentum in the light-cone approach -- represent the familiar Feynman parameters in the usual covariant Feynman diagram approach. We have therefore discovered a physical basis for the Feynman parameters. The actual value of the integral may be computed \cite{jens,grad}, and is given by \begin{equation} \frac{\lambda^2}{144 \pi^2 m^2} \left[ \frac{2 \pi^2}{3} - \psi'(\frac{1}{3}) \right], \end{equation} where $\psi(x)$ is Euler's psi function. The terms in the parenthesis has the numerical value $-3.51586\dots$. We leave an analysis of the self-energy contribution $\Gamma({\bf k}^{\perp})$ in $D > 2$ dimensions for future work. \section{Discussion} \label{the-end} We have calculated the two-loop mass renormalization for the effective light-cone Hamiltonian of scalar $\phi^4$ field theory. An explicit evaluation of the corresponding integral in two dimensions was given, and shown to be precisely equal to the corresponding two-loop `setting sun' Feynman diagram. Moreover, we were able to attribute physical significance to the familiar Feynman parameters; namely, as fractions of vanishingly small external light-cone momenta. Evidently, it would be interesting to study the integral (\ref{gamma}) in more detail for higher space-time dimensions. In particular, ultraviolet divergences are expected in dimensions $D\geq 3$, and so we may invoke dimensional regularization to regulate the integration over transverse momentum space. This would facilitate a straightforward comparison with existing Feynman diagram calculations in $\phi^4$ theory \cite{collins}. \medskip \vspace{10mm} {\bf Acknowledgments:} F.A. is grateful for fruitful discussions with Armen Ezekelian. J.O.A was supported in part by a Faculty Development Grant from the Physics Department at the Ohio State University, and by a NATO Science Fellowship from the Norwegian Research Council (project 124282/410).
\section{Introduction} \subsection{Inspiral of Compact Object Binaries} A pair of compact objects (black holes or neutron stars) in binary orbit about one another is stable in Newtonian gravity. In general relativity, however, the system will emit gravitational radiation, causing the bodies to spiral in towards one another. The gravitational radiation given off by this system is a prime candidate for detection by upcoming gravitational wave telescopes such as VIRGO and LIGO. When the compact objects are far apart, all gravitational effects are weak and one can use the Post-Newtonian approximation, expanding in powers of $GM/Rc^2$. The final ``plunge'', where the objects cease to orbit and collide rapidly, can be modelled by full general relativistic three-plus-one supercomputer simulations. An intermediate late inspiral phase, where strong gravitational effects are important, but the fraction of total energy lost to gravitational radiation each orbit is small, cannot be handled by supercomputer evolutions, which become unstable after several orbits. To handle strong gravity and slow inspiral, we need to use an approximation scheme. Modelling this intermediate phase will allow us to determine the early gravitational wave signal as well as provide initial data for the final plunge simulations. \subsection{Quasi-Stationary Approximation} The idea, initially proposed by Steven Detweiler \cite{det}, is that if the inspiral is slow, the system is nearly periodic: after one orbit, the objects have returned almost to their original locations, and radiation which has moved out has been replaced with new radiation of approximately the same shape. If the objects' orbits are circular rather than elliptical, the spacetime is nearly stationary. If the approximate orbital frequency is $\Omega$, moving forward in time by $\delta t$ and rotating the resulting spatial slice by $-\Omega\,\delta t$ will not change the picture very much. Our approach is to replace the true spacetime, which has this approximate continuous symmetry, with another solution to Einstein's equations in which the symmetry is exact. This ``stationary quasi-solution'', which must somehow replace the energy lost in radiation to prevent inspiral, should approximate the physical ``quasi-stationary solution'' over some time interval. This replaces a three-plus-one-dimensional numerical evolution with a three-dimensional instantaneous solution, reducing greatly the computing power required and hopefully escaping some of the numerical instabilities associated with evolution. To further simplify the problem, we will initially look for a spacetime with an additional translational symmetry orthogonal to the orbital plane. This toy model of co-orbiting cosmic strings further reduces the numerical problem to two dimensions while hopefully retaining some of the qualitative features. The rest of this paper consists of summaries of three techniques being developed for this project, but which may prove useful in other settings as well. Section~\ref{sec:2KV} describes research done in collaboration with Joseph D.~Romano \cite{2KV} on the refinement of a formalism developed by Geroch \cite{geroch2} to simplify the Einstein equations when the spacetime admits two commuting Killing vectors, as is the case with the co-orbiting cosmic string model. Section~\ref{sec:radbal} discusses the construction of a stationary radiating solution to a wave equation without the use of external forces by finding a preferred solution containing a balance of incoming and outgoing radiation, the subject of a collaboration with William Krivan and Richard H.~Price \cite{radbal}. Section~\ref{sec:cyl} briefly touches on the concept of describing small-amplitude gravitational radiation as a perturbation not to Minkowski spacetime but to the Levi-Civita solution \cite{cs} for a single cosmic string. \section{QSBI I: Spacetimes with Two Killing Vectors \\ (Whelan and Romano \cite{2KV})} \label{sec:2KV} \subsection{Lack of a Block-Diagonal {Co\-\"{o}r\-di\-nate} Basis} The spacetime of two co-orbiting cosmic strings has two continuous symmetries, which are described by two Killing vector fields. One of these, which we call $K_0$, represents the combination of time translation and rotation about the orbital axis which leaves the spacetime unchanged. It can be thought of roughly in terms of traditional cylindrical {co\-\"{o}r\-di\-nate}s as $\partial_t+\Omega\partial_\phi$. The other Killing vector simply corresponds to translation along the strings and can be thought of as $K_1\sim\partial_z$. In a numerical determination of the spacetime geometry, one seeks to fix the {co\-\"{o}r\-di\-nate} (i.e., gauge) information completely, and thus calculate the minimum number of quantities necessary to define the geometry. It is desirable, of course, to choose a gauge which takes advantage of the symmetries of the problem. Following from the example of static, axisymmetric spacetimes \cite{wald}, we might wish to define two Killing {co\-\"{o}r\-di\-nate}s $\{x^A\}$ ($x^0\sim t$, $x^1\sim z$), supplemented by two other {co\-\"{o}r\-di\-nate}s $\{x^i\}$ ($x^2\sim \rho$, $x^3\sim\varphi\sim\phi-\Omega t$) on a subspace orthogonal to the two Killing vectors, and bring the line element into a block-diagonal form \begin{equation} \lambda_{AB}(\{x^k\})dx^A dx^B+\gamma_{ij}(\{x^k\})dx^i dx^j. \end{equation} For our set of Killing symmetries, however, this fails because the symmetry group is not \emph{orthogonally transitive}, which means that it is impossible to construct two-surfaces everywhere orthogonal to both Killing vector trajectories. The measure of this failure is a pair of scalar fields $\{c_A\}$ given by \begin{equation} c_A=\dual{K_0\wedge K_1\wedge\,d K_A} . \end{equation} The co-rotational Killing vector $K_0$ is not surface-forming ($K_0\wedge\,d K_0\ne 0$), which leads to the non-vanishing of $c_0$, indicating a lack of orthogonal transitivity. \subsection{Manifold of Killing Vector Orbits} Although the metric components in any {co\-\"{o}r\-di\-nate} basis will have non-vanishing ``cross terms'' $\{g_{Ai}\}$ between the Killing and non-Killing directions, it still possible to describe the geometry in terms of two matrices $\{\lambda_{AB}\}$ and $\{\gamma_{ij}\}$ along with the two scalars $c_0$ and $c_1$. This done using a construction due to Geroch \cite{geroch2} which defines a two-manifold $\mathcal{S}$ of Killing vector orbits. The {co\-\"{o}r\-di\-nate}s $\{x^i|i=2,3\}$ are {co\-\"{o}r\-di\-nate}s on this two-manifold, and any tensor on the original four-dimensional spacetime manifold $\mathcal{M}$ which has vanishing inner products with, and Lie derivatives along, the Killing vectors corresponds to a tensor on the two-manifold $\mathcal{S}$. In particular, the tensors and scalars on $\mathcal{S}$ which describe the spacetime geometry are \begin{itemize} \item The symmetric matrix of inner products $\lambda_{AB}=K_A\cdot K_B$ \item The metric on $\mathcal{S}$, which has components $\{\gamma_{ij}\}$ and corresponds to the projection tensor $g_{\mu\nu}-\lambda^{AB}K_{A\mu}K_{B\nu}$ on $\mathcal{M}$ (where $\{\lambda^{AB}\}$ is the inverse of $\{\lambda_{AB}\}$) \item The two scalars $c_A$ which define the lack of orthogonal transitivity. \end{itemize} Given these objects, it is possible to define a non-{co\-\"{o}r\-di\-nate} basis on the four-dimensional spacetime by $e_A=K_A$ and $e_{i}=\gamma_{ij}dx^j$. In this basis the metric has components \begin{equation} g_{AB}=\lambda_{AB}\, , \qquad g_{Ai}=0\, , \qquad g_{ij}=\gamma_{ij}\, ; \end{equation} the $\{c_A\}$ give the commutation {co\-\"{e}f\-fi\-cient}s of the non-commuting basis vectors: \begin{equation} [e_2,e_3]\propto \lambda^{AB}c_A K_B\,. \end{equation} \subsection{Einstein Equations} Geroch's original paper contained four partial differential equations involving $\{\gamma_{ij}\}$, $\lambda_{00}$, $\lambda_{01}$, $\lambda_{11}$, $c_0$, and $c_1$ which were equivalent to the vacuum Einstein equations $G_{\mu\nu}=0$. We have found a streamlined derivation which treats the indexed quantities $\lambda_{AB}$ and $c_A$ as single entities, rather than dealing with each component individually. We have also derived explicit expressions (given in \cite{2KV}) for the components $G_{AB}$, $G_{Ai}$, and $G_{ij}$ which can be used even when the stress-energy tensor is non-vanishing. In the case where the off-block-diagonal components $\{T_{Ai}\}$ of the stress-energy tensor vanish, the Einstein equations $G_{Ai}=0$ simply say that the scalars $c_A$ are constants. A convenient choice of gauge in which to solve the remaining six Einstein equations is a basis in which \begin{eqnarray} c_0\equiv 2\Omega \, , \qquad c_1\equiv 0 \, , \\ \lambda_{00}=(\lambda+X(\rho,\varphi)^2)Z(\rho,\varphi)^{-1} \, , \quad \lambda_{01}=X(\rho,\varphi) \, , \quad \lambda_{11}=Z(\rho,\varphi) \, , \\ \gamma_{22}=1 \, , \qquad \gamma_{23}=0 \, , \qquad \gamma_{33}=-\lambda(\rho,\varphi)^{-1}\rho^2 F(\rho,\varphi) \,. \end{eqnarray} These four functions $X$, $Z$, $\lambda$, and $F$ of the two {co\-\"{o}r\-di\-nate}s $x^2=\rho$ and $x^3=\varphi$ provide a fixed gauge in which the problem can be solved numerically. There are six block-diagonal Einstein equations (for $\{G_{AB}\}$ and $\{G_{ij}\}$), but only four of them are independent because of the contracted Bianchi identities $(dx^i)_\mu \nabla_\nu G^{\mu\nu}=0$. These equations will be combined with a treatment of the cosmic string sources and the boundary conditions at infinity and the origin to produce a numerically determined spacetime. \section{QSBI II: Radiation-Balanced Boundary Conditions \\ (Whelan, Krivan and Price \cite{radbal})} \label{sec:radbal} \subsection{Radiative Boundary Conditions} The true physical spacetime which we are ultimately modelling contains gravitational radiation at infinity which is outgoing. This loss of energy leads to decay of the orbits and inspiral of the compact objects in a non-stationary solution. In many radiative problems, a solution with outgoing radiation can be constrained to be stationary by an external force whose agent does not couple to the radiation. However, this is at odds with the idea that in General Relativity all matter gravitates, so we should look instead for a solution where there is no net energy loss to infinity due to the radiation. A na\"{\i}ve replacement for the outgoing radiation boundary conditions would be a standing wave condition. However, that turns out to be inappropriate, as standing waves require a node (Dirichlet) or extremum (Neumann) at a particular location, and a standing wave condition thus fails to converge to a well-defined limit as that location is moved out to infinity. We have been investigating the question of how to implement a sensible boundary condition leading to a balance of radiation in the context of a simple theory: a nonlinear scalar field $\psi(t,\rho,\phi)$ in two-plus-one dimensions, where the source $\sigma$ and field $\psi$ are required to be co-rotating (i.e., depend only on $\rho$ and $\varphi=\phi-\Omega t$). This causes the field equation to take the form \begin{equation} \label{coroteqn} \Box^2\psi= \frac{\partial^2\psi}{\partial\rho^2} +\rho^{-1}\frac{\partial\psi}{\partial\rho} +\left( \rho^{-2}-\Omega^2 c^{-2} \right) \frac{\partial^2\psi}{\partial\varphi^2} = \sigma + \lambda \psi^3 . \end{equation} In the numerical solution for the non-linear equation \eqref{coroteqn}, one uses a finite {co\-\"{o}r\-di\-nate} grid and specifies a set of boundary conditions at the large finite radius $\rho=R$ which marks the end of the grid. Table \ref{tab:1} \begin{table}[tbp] \label{tab:1} \begin{tabular}{r|c|c} &BCs at $\rho=R$ & $R\longrightarrow\infty$ limit \\ \hline Outgoing & $(\partial_\rho+c^{-1}\partial_t)\psi^{\text{out}}_R=0$ & $\psi^{\text{out}}\sim e^{im(\Omega c^{-1}\rho+\varphi)}\sim e^{im\Omega(c^{-1}\rho-t)}$ \\ Ingoing & $(\partial_\rho-c^{-1}\partial_t)\psi^{\text{in}}_R=0$ & $\psi^{\text{in}}\sim e^{im(\Omega c^{-1}\rho-\varphi)}\sim e^{im\Omega(c^{-1}\rho+t)}$ \\ Neumann SW & $\partial_\rho\psi^{\text{N}}_R=0$ & {\bf N/A} \\ Dirichlet SW & $\psi^{\text{D}}_R=0$ & {\bf N/A} \\ \end{tabular} \caption{Some large-distance boundary conditions. Purely ingoing or outgoing radiation has a well-defined limit as the radius $R$ at which it is applied is taken to infinity, but standing wave conditions (e.g., Neumann or Dirichlet) do not.} \end{table} shows several possible boundary conditions. We call, e.g., the solution resulting from the application of the outgoing boundary condition at a particular $R$, $\psi^{\text{out}}_R$. The local conditions defining outgoing or ingoing radiation each produce well-defined limits (which we call simply $\psi^{\text{out}}$ and $\psi^{\text{in}}$) as $R$ is taken to infinity, but the standing wave solutions do not converge to any limit. Taking the average of the $R\rightarrow\infty$ limits $\psi^{\text{out}}$ and $\psi^{\text{in}}$ defines a function \begin{equation} \psi^{\text{avg}}=\frac{1}{2}(\psi^{\text{in}}+\psi^{\text{out}}) \end{equation} which has a balance of ingoing and outgoing radiation without reference to any particular radius $R$. In the linear ($\lambda=0$) theory, where the principle of superposition holds, it is a solution to the field equations, but in the non-linear theory it is not. It is also not defined as the limit of any local boundary conditions (in particular, the is average of out- and ingoing solutions is not produced by an average of out- and ingoing boundary conditions). \subsection{Green's Function Solution} The tasks of defining a solution analogous to $\psi^{\text{avg}}$ and numerically determining that solution in the absence of a local boundary condition can both be accomplished by a Green's function method. If we define a Green's function $G(\rho,\varphi|\rho_0,\varphi_0)$ such that \begin{equation} \label{GFeqn} \Box^2 G = \rho_0^{-1}\delta(\rho-\rho_0)\,\delta(\varphi-\varphi_0) \, , \end{equation} the differential equation \eqref{coroteqn} can be converted into an integral equation \begin{equation} \label{NLinteqn} \psi(\rho,\varphi)=\int\rho_0\,d\rho_0\,d\varphi_0\, G(\rho,\varphi|\rho_0,\varphi_0) \, [\sigma(\rho_0,\varphi_0)+\lambda\psi^3(\rho,\varphi)] \, . \end{equation} Solving to the Green's function equation \eqref{GFeqn} requires the application of boundary conditions at infinity, so there are actually a family of Green's functions with boundary conditions ``built into'' them. For example, using the retarded Green's function $G^{\text{out}}$ in \eqref{NLinteqn} gives the outgoing solution $\psi^{\text{out}}$, using the advanced Green's function $G^{\text{in}}$ gives the ingoing solution $\psi^{\text{in}}$, and similarly for any of the boundary conditions defined at finite radii in Table~\ref{tab:1}. Since the Green's function equation \eqref{GFeqn} is linear, even when the wave equation \eqref{coroteqn} is not, we can always average the retarded and advanced \emph{Green's functions} to give a time-symmetric Green's function \begin{equation} G^{\text{sym}}=\frac{1}{2}(G^{\text{in}}+G^{\text{out}}) \end{equation} which defines a radiation-balanced solution $\psi^{\text{sym}}$. In the linear theory, this solution $\psi^{\text{sym}}$ will be equal to the average $\psi^{\text{avg}}$ of the ingoing and outgoing solutions. In the non-linear theory, we can calculate the two numerically (by iterating \eqref{NLinteqn}) and compare them. We are ultimately trying to use the outgoing piece of $\psi^{\text{sym}}$ as an approximation for $\psi^{\text{out}}$ (and thus as a further approximation for the actual inspiralling solution), and we expect that approximation to be good when $\psi^{\text{sym}}\approx\psi^{\text{avg}}$. \subsection{Numerical Results} Numerical calculations \cite{radbal} for a strongly non-linear ($\lambda=20$) theory have shown that the approximation becomes good as the orbital velocity of the sources becomes non-relativistic. Figures \ref{fig:.25c} and \ref{fig:.125c} \begin{figure}[tbp] \epsfig{file=07_JWHELAN1_1.ps,width=15pc,clip=,angle=90,% bbllx=430,bblly=160,bburx=160,bbury=610} \begin{center} \caption{Comparison between $\psi^{\text{sym}}$ and $\psi^{\text{avg}}$ in the wave zone for a source of two particles of opposite charge at $\rho/2=.25$, $\varphi=(\pi\pm\pi)/2$, with a rotational frequency of $\Omega=1/2$ and an orbital velocity of one-fourth the speed of light. The fields are plotted versus $\rho/2$ for the range of angles $\varphi\in[1,\pi/2]$. The amplitude of $\psi^{\text{avg}}$ is around 20\% smaller and the phase is shifted 40 degrees relative to $\psi^{\text{sym}}$. This agreement is only middling because the sources are still somewhat relativistic.} \label{fig:.25c} \end{center} \end{figure} \begin{figure}[tbp] \epsfig{file=07_JWHELAN1_2.ps,width=15pc,clip=,angle=90,% bbllx=430,bblly=160,bburx=160,bbury=610} \begin{center} \caption{Comparison between $\psi^{\text{sym}}$ and $\psi^{\text{avg}}$ in the wave zone for a source of two particles of opposite charge at $\rho/2=.125$, $\varphi=(\pi\pm\pi)/2$, with a rotational frequency of $\Omega=1/2$ and an orbital velocity of one-eighth the speed of light. The fields are plotted versus $\rho/2$ for the range of angles $\varphi\in[1,\pi/2]$. With these less relativistic sources, the agreement is better than in Figure~\ref{fig:.25c}, with approximately a 9\% smaller amplitude for $\psi^{\text{avg}}$ and 7 degree phase shift.} \label{fig:.125c} \end{center} \end{figure} show the comparison between $\psi^{\text{sym}}$ and $\psi^{\text{avg}}$ for orbital velocities of one-fourth and one-eighth the speed of light, respectively. This agreement occurs even when the theory is highly nonlinear because, even though $\psi^{\text{in}}$ and $\psi^{\text{out}}$ differ in the wave zone, they are approximately the same in the inner, strong-field region, as illustrated in Figure \ref{fig:try}. \begin{figure}[tbp] \epsfig{file=07_JWHELAN1_3.ps,width=20pc,clip=,angle=-90,% bbllx=75,bblly=90,bburx=540,bbury=770} \begin{center} \caption{Illustration of the agreement between $\psi^{\text{out}}$ and $\psi^{\text{in}}$ in the strong-field region. The parameters are the same as in Figure \ref{fig:.125c}, i.e., oppositely-charged source particles at $\rho/2=.125$, $\varphi=(\pi\pm\pi)/2$, a rotational frequency of $\Omega=1/2$ and an orbital velocity of one-eighth the speed of light. The fields are plotted against $\rho/2$ for $\varphi=3\pi/8$. Note that while the two solutions differ in the outer, wave zone, they are nearly equal in the inner regions which are the source of the most of the nonlinear effects, leading to the approximate agreement of $\psi^{\text{sym}}$ and $\psi^{\text{avg}}$ (which are identical in the linear theory) illustrated in Figure \ref{fig:.125c}.} \label{fig:try} \end{center} \end{figure} The final step in the scalar field theory analysis will be to numerically extract the ``outgoing part'' from the time-symmetric solution $\psi^{\text{sym}}$ and compare it directly to $\psi^{\text{out}}$; we expect that the two will agree when $\psi^{\text{sym}}$ is approximately equal to $\psi^{\text{avg}}$. \section{QSBI III: Gravitational Waves on a Cosmic String Background} \label{sec:cyl} \subsection{Schwarzschild vs.\ Levi-Civita Backgrounds} In the three-plus-one binary inspiral problem, with non-extended sources, gravitational waves at infinity can be treated as perturbations to Minkowski spacetime. This is because at large distances, where higher multipoles of the non-radiative field can be ignored, the non-radiative part of the geometry is associated with the Schwarzschild metric \begin{equation} -(1-r_s/r)\,c^2dt^2+(1-r_s/r)^{-1}dr^2+r^2d\Omega^2 \, , \end{equation} which contains a length scale $r_s=2GM/c^2$; taking $r/r_s\rightarrow\infty$ reduces the Schwarzschild line element to the Minkowski one, which is just another way of saying that the Schwarzschild solution is asymptotically flat. With infinitely extended sources, this is not possible. Far from the strings, where the internal structure is not felt, the non-radiative effects are approximated by the Levi-Civita solution \cite{cs} of a single cosmic string at the origin with mass-per-unit length $\Lambda$, whose line element is given for small $\Lambda$ by \begin{equation} - \rho^{2C} c^2 dt^2 + \rho^{-2C} dz^2 + d\rho^2 + \rho^2 d\phi^2 \, . \end{equation} The mass parameter $C=2G\Lambda/c^2$ is now dimensionless; this lack of a length scale means that Levi-Civita spacetime is \emph{not} asymptotically flat and we must use Levi-Civita rather than Minkowski as our background metric. \subsection{Perturbations to Levi-Civita} Suppose the metric is given by the Levi-Civita metric plus a small perturbation $h_{\mu\nu}(t,\rho,\phi)$. (In our problem the $t$ and $\phi$ dependence will further be restricted to a function of $\phi-\Omega t$, but we are thinking here of more general considerations.) We can divide the components $\{h_{\mu\nu}\}$ into those which are odd under inversion of the $z$ {co\-\"{o}r\-di\-nate} ($h_{zt}$, $h_{z\rho}$, and $h_{z\phi}$) and those which are even ($h_{zz}$, $h_{tt}$, $h_{t\rho}$, $h_{t\phi}$, $h_{\rho\rho}$, $h_{\rho\phi}$, and $h_{\phi\phi}$). In the linearized theory, the odd and even sectors are uncoupled. To count the physical degrees of freedom, we consider the effects of a gauge transformation \begin{equation} h_{\mu\nu}\rightarrow h_{\mu\nu}-2\,\xi_{(\mu;\nu)} \end{equation} by a gauge parameter $\xi_\mu(t,\rho,\phi)$, as well as the Hamiltonian constraint $H$ and the three components $\{H_z,H_\rho,H_\phi\}$ of the momentum constraint. The three odd components of the metric perturbation consist of one gauge degree of freedom $\xi_{z}$, one constraint $H_{z}$, and one physical degree of freedom, which can be thought of as the ``cross polarization'' (since it involves $h_{z\phi}$). The seven even components include three gauge degrees of freedom $\xi_{t}$, $\xi_{\rho}$, and $\xi_{\phi}$, three constraints $H$, $H_{\rho}$, and $H_{\phi}$, and one physical ``plus polarization'' involving $h_{zz}$ and $h_{\phi\phi}$. The next step in a treatment akin to that done by Moncrief \cite{moncrief} for waves on a Schwarzschild background will be to derive explicit wave equations for these two gauge-invariant physical degrees of freedom. For the purposes of the quasi-stationary binary inspiral program, it may suffice to work with linearized versions of the gauge-fixed equations derived in \cite{2KV}, but the picture of general perturbations to Levi-Civita will at least be useful in interpreting the quantities involved. \section{Future Outlook} The work described in Section~\ref{sec:2KV} has produced non-linear differential equations which can be numerically implemented given a suitable treatment of the sources and the boundary conditions on the radiation at the outer boundary. The study of radiation balance in the context of non-linear scalar field theory in Section~\ref{sec:radbal} will be completed by the extraction of the outgoing part of the time-symmetric Green's function solution and the comparison between that and the actual outgoing solution (which we expect to be unavailable in the gravitational case). Application of these methods to General Relativity will be complicated by the fact that it is probably not possible to formulate a (linear) Green's Function solution to the non-linear wave equation in that case. The analysis of perturbations to Levi-Civita spacetime described in Section~\ref{sec:cyl} is still in the early stages, and its full development may or may not be necessary to the understanding of our problem, which possesses the additional co-rotational Killing vector. Finally, should the technique of finding a stationary solution with a balance of ingoing and outgoing radiation prove successful in the essentially two-dimensional problem of orbiting cosmic strings, it can be applied to the computationally more intensive three-dimensional problem. \section*{Acknowledgments} The author would like to thank his collaborators J.~D.~Romano, W.~Krivan, R.~H.~Price, as well as C.~Torre, K.~Thorne, J.~Creighton, J.~Friedmann, S.~Morsink, A.~Held, and the Relativity Group at the University of Utah, where this work was begun, and acknowledge the financial support of NSF Grant PHY-9734871, Swiss Nationalfonds, and the Tomalla Foundation Z\"{u}rich. \section*{References}
\subsection{Numerical viscosity in Marquina's flux formula} {\it M and MM- flux formulae}. The numerical flux across a given interface can be written like equation (\ref{numfluxroe}) with \begin{equation} {\bf q} = {\bf q}^{L} - {\bf q}^{R} \end{equation} \begin{equation} {\bf q}^{L,R} = {\bf \cal Q}^{L,R} {\bf u}^{L,R} \end{equation} Omitting the superscripts $L,R$ and taken into account the expressions in Table I for MM and the results in \cite{DFIM98}, the viscosity vector in the $x$-splitting is: \begin{eqnarray} q^{L,R}_1 &=& \frac{h^2}{\Delta} \left\{ M \left[ {\cal A}_{-} {\Omega}_{+} - {\cal A}_{+} {\Omega}_{-} \right] + p(c_{+}\aleph_{+}-c_{-}\aleph_{-}) \right\} + \nonumber \\ & & c_{0}p\frac{W}{h}\left\{ \frac{{\cal K}}{{\cal K} - 1} + \frac{v_y^2+v_{z}^2} {1-v_x^2} \right \} \nonumber \\ q^{L,R}_2 &=& \frac{h^2W}{\Delta} \left\{ M{\cal A}_{+}{\cal A}_{-} \left[ {\Omega}_{+}\lambda_{+}-{\Omega}_{-}\lambda_{-} \right] + p(c_{+}\lambda_{+}{\cal A}_{+}\aleph_{+}- c_{-}\lambda_{-}{\cal A}_{-}\aleph_{-}) \right\} + \nonumber \\ & & c_{0}pW^2v_x \left\{ \frac{1}{{\cal K} - 1} + 2\frac{v_y^2+v_{z}^2}{1-v_x^2} \right \} \nonumber \\ q^{L,R}_3 &=& \frac{h^2W}{\Delta}v_y \left\{ M \left[ {\Omega}_{+}{\cal A}_{-} - {\Omega}_{-}{\cal A}_{+} \right] + p(c_{+}\aleph_{+}-c_{-}\aleph_{-}) \right\} + \nonumber \\ & & c_{0}p \left\{ \frac{W^2}{{\cal K} - 1} + \frac{1+2W^2(v_y^2+v_{z}^2)}{1-v_x^2} \right \} \nonumber \\ q^{L,R}_4 &=& \frac{h^2W}{\Delta}v_{z} \left\{ M \left[ {\Omega}_{+}{\cal A}_{-}- {\Omega}_{-}{\cal A}_{+} \right] + p(c_{+}\aleph_{+}-c_{-}\aleph_{-}) \right\} + \nonumber \\ & & c_{0}p \left\{ \frac{W^2}{{\cal K} - 1} + \frac{1+2W^2(v_y^2+v_{z}^2)}{1-v_x^2} \right \} \nonumber \\ q^{L,R}_5 &=& \frac{h^2}{\Delta} \left\{ M \left[ {\cal A}_{-} {\Omega}_{+}{\cal D}_{+}- {\cal A}_{+} {\Omega}_{-}{\cal D}_{-} \right] + p[c_{+}\aleph_{+}{\cal D}_{+} - c_{-}\aleph_{-}{\cal D}_{-}] \right\} + \nonumber \\ & & c_{0}p\frac{W}{h}\left\{ \frac{hW-{\cal K}}{{\cal K} - 1} + \frac{(2hW-1)(v_y^2+v_{z}^2)}{1-v_x^2} \right \}, \label{viscos} \end{eqnarray} with the following auxiliary quantities \begin{eqnarray} M=\rho hW^2({\cal K}-1), \quad {\Omega}_{\pm}= c_{\pm}(v_x - \lambda_{\mp}), \quad {\cal D}_{\pm}= hW{\cal A}_{\pm}-1, \nonumber \\ {\cal K} \equiv {\displaystyle{\frac{\tilde{\kappa}} {\tilde{\kappa}-c_s^2}}, \:\:\:\: {\tilde \kappa}= \frac{1}{\rho} \left.\frac{\partial p}{\partial \varepsilon}\right|_{\rho}}, \:\:\:\: {\cal A}_{\pm} \equiv {\displaystyle{\frac{1 - v_x v_x}{1 - v_x {\lambda}_{\pm}}}} \nonumber\\ \Delta = h^3 W ({\cal K} - 1) (1 - v_x v_x) ({\cal A}_{+} {\lambda}_{+} - {\cal A}_{-} {\lambda}_{-}) \nonumber \\ \aleph_{\pm} = {\pm}\left \{ -v_x -W^2({\rm v}^{2}-v_x v_x)(2{\cal{K}}-1)(v_x - {\cal{A}_{\pm}}\lambda_{\pm}) +{\cal{K} \cal{A}_{\pm}} \lambda_{\pm} \right\} \end{eqnarray} where quantities $c_{\pm,0}$ are given in Table I and $\Delta$ is the determinant of the matrix of right-eigenvectors. The corresponding viscosity vectors in the other directions are trivially obtained by a cyclic permutation of subindices ${x,y,z}$. We have tested the efficiency of our numerical proposal, for Roe's and MM's flux formulae, running GENESIS (a 3D special relativistic hydro-code \cite{AIMM99}), without any optimization at compilation level, in a SGI Origin 2000. A standard initial value problem has been chosen: $\rho_L=10$, $\epsilon_L=2$, $v_L=0$, $\gamma_L=5/3$, $\rho_R=1$, $\epsilon_R=10^{-6}$, $v_R=0$ and $\gamma_R=5/3$, where the subscript $L$ ($R$) denotes the state to the left (right) of the initial discontinuity. This test problem has been considered by several authors in the past (see \cite{AIMM99} for details in 1D, 2D and 3D). We have compared the performance of two different implementations of the numerical flux subroutine: i) {\it Case A}, stands for the results obtained using our analytical prescription. This means to write down, in the numerical flux routine, just the expressions derived here for the viscosity vector ${\bf q}$. ii) {\it Case B}, stands for the results obtained running the code with a standard high-efficiency subroutine for inverting matrices (we use a LU decomposition plus an implicit pivoting which is, for general matrices, $O(N^3)$). This subroutine is called to get the left eigenvectors from the matrix of right eigenvectors and is adapted to the particular dimensions of the matrices ($3 \times 3$, in 1D, $4 \times 4$, in 2D and $5 \times 5$, in 3D). Hence, unlike case A, now we have to calculate numerically the following quantities: the matrix of left eigenvectors, the characteristic variables and, finally, the components of the viscosity vector ${\bf q}$. \begin{table} \centerline{\begin{tabular}{|cccccc|} \multicolumn{6}{c}{\bf TABLE II \rm} \\ \multicolumn{6}{c}{\protect \small CPU time (in microseconds).} \\ \multicolumn{6}{c}{}\\ \hline \hline & & \multicolumn{4}{c|}{TCI $(\mu s)$} \\ & & \multicolumn{2}{c}{Roe} & \multicolumn{2}{c|}{MM} \\ Case & \# Zones & Case $A$ & Case $B$ & Case $A$ & Case $B$ \\ \hline 1D & $100 \times 1 \times 1 $ & $12.2$ & $53.8$ & $23.8$ & $118.9$ \\ 2D & $20 \times 20 \times 1 $ & $25.5$ & $181.8$ & $49.0$ & $373.5$ \\ 3D & $14 \times 14 \times 14$ & $39.4$ & $431.9$ & $75.7$ & $879.0$ \\ \hline \hline \end{tabular}} \caption{\small Time per numerical cell and iteration (TCI) in microseconds employed by the numerical flux routine in our test-bed problem, for three different grids.} \end{table} Table II summarizes the results: the direct implementation of our numerical viscosity formulae leads to an improvement of the efficiency (in terms of CPU time) of the numerical fluxes subroutine in a factor which, in 3D calculations, ranges between about {\it eleven} and {\it twelve} depending on the particular flux formula used. When comparing Roe's and MM's cases a factor two --in favour of Roe-- arises due to the fact that MM's flux formulae needs to compute two viscosity vectors (one per each side of a given interface), unlike Roe's flux formula which needs only one viscosity vector evaluated at the average state. As it must be, the efficiency increases with the number of spatial dimensions involved in the problem due to the computationally expensive matrix inverting operations performed at each interface to get the numerical fluxes. Since the numerical flux routine is, typically, one of the most time-consuming, it translates into a speed up factor between {\it two} and {\it four} in the total execution time, depending on the specific weight of the flux formulae routine in each particular application. Our formulation also gives a unified description of the numerical fluxes (\ref{FF}), permitting a unique implementation with the possibility of switching in cases when the utilisation of a specific flux formula is more appropriate. In addition, due to the fact that we have eliminated, in the generalized MM's flux formula, all the {\it conditional clauses}, the efficiency is ensured either for scalar or vectorial processors. Another worthy by-products of our algebraic pre-processing concerns with the significant reduction of round-off errors, as a consequence of the number of operations suppressed and factorization. One of the important issues in designing a multidimensional hydro-code is the accurate preservation of any symmetries present in a physical problem. A numerical violation of these symmetries could arise as a consequence of accumulation of round-off errors in the calculation of the numerical fluxes, as we have explained in a previous paper \cite{AIMM99}. The algebraic simplifications, shown in the present paper, reduce the number of operations and cure such problem. Two last additional consequences arise from our work. First is that similar expressions can be worked out for any non-linear hyperbolic system of conservation laws for which the full spectral decomposition is known. In particular, when some of the vectorial subspaces has multiple degeneracy, a similar algebraic preprocessing is very convenient. The other important consequence is that an appropriate combination of a simplified formulation of the numerical viscosity together with the use of special relativistic Riemann solvers in General Relativistic Hydrodynamics \cite{PFIMM98}, should allow a very easy and efficient extension to General Relativistic Hydrodynamics.
\section{Introduction} Sonoluminescence (SL) is the phenomenon of light emission by a sound-driven gas bubble in fluid~\cite{Physics-Reports}. The intensity of a standing sound wave can be increased until the pulsations of a bubble of gas trapped at a velocity node have sufficient amplitude to emit picosecond flashes of light. The basic mechanism of light production is still controversial. We shall start by presenting a brief summary of the main experimental data (as currently understood) and their sensitivities to external and internal conditions. The most common situation is that of an air bubble in water. SL experiments usually deal with bubbles of ambient radius $R_{\mathrm ambient} \approx 4.5 \; \mu {\rm m}$. The bubble is driven by a sound wave of frequency of $20 - 30$ kHz. During the expansion phase, the bubble radius reaches a maximum of order $R_{\mathrm max}\approx 45 \; \mu {\rm m}$, followed by a rapid collapse down to a minimum radius of order $R_{\mathrm max}\approx 0.5 \; \mu {\rm m}$. The photons emitted by such a pulsating bubble have typical wavelengths of the order of visible light. The minimum observed wavelengths range between $200 \;{\rm nm}$ and $100 \; {\rm nm}$. This light appears distributed with a power-law spectrum (with exponent depending on the noble gas admixture entrained in the bubble) with a cutoff in the extreme ultraviolet. If one fits the data to a Planck black-body spectrum the corresponding temperature is several tens of thousands of Kelvin (typically $70,000\; {\rm K}$, though estimates varying from $40,000 \; {\rm K}$ to $100,000 \; {\rm K}$ are common). There is considerable doubt as to whether or not this temperature parameter corresponds to any real physical temperature. There are about one million photons emitted per flash, and the average total power released is $30 \;\hbox{ mW} \leq W \leq 100 \;\hbox{ mW}$. The photons appear to be created in a very tiny spatio-temporal region: of order $10^{-1} \mu {\rm m}$ and on timescales $\tau \leq 50 \;\hbox{\rm ps}$ (there have been claims that flash duration is less than $100\; {\rm fs}$, though more recent claims place flash duration in the range $50 - 250$ ps~\cite{Flash1,Flash2}). A truly successful theory of SL must also explain a whole series of characteristic sensitivities to different external and internal conditions. Among these dependencies the main one is surely the mysterious catalytic role of noble gas admixtures. Other external conditions that influence SL are magnetic fields and the temperature of the water (see~\cite{Physics-Reports}). These are only the most salient experimental dependencies of the SL phenomenon. In explaining such detailed and specific behaviour the Casimir approach (the QED vacuum approach) encounters the same problems as other approaches have. Nevertheless we shall argue that SL explanations using a Casimir-like framework are viable. \subsection{Quasi-static Casimir models: Schwinger's approach}\label{subsec:schw} The idea of a ``Casimir route'' to SL is due to Schwinger who several years ago wrote a series of papers~\cite{Sc1,Sc2,Sc3,Sc4,Sc5,Sc6,Sc7} regarding the so-called dynamical Casimir effect. Considerable confusion has been caused by Schwinger's choice of the phrase ``dynamical Casimir effect'' to describe his model. In fact, his original model is not dynamical and is at best quasi-static as the heart of the model lies in comparing two static Casimir energy calculations: that for an expanded bubble with that for a collapsed bubble. Schwinger estimated the energy emitted during this collapse as being approximately equal to the change in the static Casimir energy. The static Casimir energy of a dielectric bubble (of dielectric constant $\epsilon_{{\mathrm inside}}$) in a dielectric background (of dielectric constant $\epsilon_{{\mathrm outside}}$) is \begin{equation} E_{cavity} =+\frac{1}{6 \pi} \hbar c R^3 K^4 \left( {1\over\sqrt{\epsilon_{\mathrm {\mathrm inside}}}}- {1\over\sqrt{\epsilon_{\mathrm {\mathrm outside}}}} \right) +\cdots. \label{Esch} \end{equation} Here the dots stand for additional sub-dominant finite volume effects~\cite{CMMV1,CMMV2,MV}. The quantity $K$ is a high-wavenumber cutoff that characterizes the wavenumber at which the real parts of the refractive indices drop to their vacuum values. Hence $K$ is a physical cutoff given by condensed matter physics, not a regularization parameter to be renormalized away. This cutoff can be interpreted as the typical length scale beyond which the notion of a continuous dielectric medium is no longer meaningful. The result (\ref{Esch}) can also be rephrased in the clearer and more general form as~\cite{CMMV1,CMMV2,MV}: \begin{equation} E_{ cavity} = + 2 V \int \frac{d^3\vec{k}}{(2 \pi)^3} \; \frac{1}{2} \hbar \left[ \omega_{\mathrm {\mathrm inside}}(k) - \omega_{\mathrm {\mathrm outside}}(k) \right] + \cdots \end{equation} where it is evident that the Casimir energy can be interpreted as a difference in zero point energies due to the different dispersion relations inside and outside the bubble. In the case of SL $\omega_{\mathrm inside}(k)\approx ck$, $\omega_ {\mathrm outside}(k)=ck/n$ for $k < K$, and $\omega_ {\mathrm outside}(k)\approx ck$ for $k > K$. In Schwinger's original model he took $n\approxn_{\mathrm liquid}\approx 1.33$, $R\approx R_{\mathrm max}$, and $K\approx 2\pi/400 \;{\rm nm^{-1}}$, leading to about three million emitted photons~\cite{Sc4}. The three main strengths of models based on zero point fluctuations are:\\ 1) The vacuum production of photon pairs allows for the very short timescales that one requires to fit data. Typically one expects these timescales to be of the order of the time that the zero point modes of the EM field takes to be correlated on the bubble scale. (Roughly the light-crossing time for the bubble.) For a bubble of radius $0.5$ microns this time scale is about $1.6$ femtoseconds, which is certainly sufficiently rapid to be compatible with observed flash duration.\\ 2) One does {\em not} need to achieve ``real'' temperatures of thousands of Kelvin inside the bubble. Quasi-thermal behaviour is generated in quantum vacuum models by the squeezed nature of the two photon states created, and the ``temperature'' parameter is a measure of the squeezing, not a measure of any real physical temperature~\footnote{This ``false thermality'' must not be confused with the very specific phenomenon of Unruh temperature. In that case, valid only for uniformly accelerated observers in flat space, the temperature is related to the constant value of the acceleration. Instead, in the case of squeezed states, the apparent temperature can be related to the degree of squeezing~\cite{BLVS} of the real photon pairs generated via the dynamical Casimir effect.}.\\ 3) There is no actual production of far ultraviolet photons (because the refractive index goes to unity in the far ultraviolet) so one does not expect the dissociation effects in water that other models imply. Models based on the quantum vacuum automatically provide a cutoff in the far ultraviolet from the behaviour of the refractive index. Moreover this cutoff appears to be sensitive to the water temperature in such a way to explain the former described experimental dependencies---this observation going back to Schwinger's first papers on the subject~\cite{Sc4}. Thus one key issue in Schwinger's model is simply that of calculating static Casimir energies for dielectric spheres. It must be stressed that there is still considerable disagreement on this calculation. Milton, and Milton and Ng~\cite{KM} strongly criticize Schwinger's result. These points have been discussed extensively in~\cite{CMMV1,CMMV2,MV} where it is emphasized that one has to compare two different geometrical configurations, and different quantum states, of the same spacetime regions. In a situation like Schwinger's model for SL one has to subtract from the zero point energy (ZPE) for a vacuum bubble in water the ZPE for water filling all space. It is clear that in this case the bulk term is physical and {\em must} be taken into account. In the situation pertinent to sonoluminescence, the total volume occupied by the gas is not at all conserved (the gas is truly compressed), and it is far too naive to simply view the ingoing water as flowing coherently from infinity (leaving voids filled with air or vacuum somewhere in the apparatus). Since the density of water is approximately but not exactly constant, the influx of water will instead generate an outgoing density wave which will be rapidly damped by the viscosity of the fluid. The few phonons generated in this way are surely negligible. \subsection{Eberlein's dynamical model for SL} The quantum-vacuum approach to SL was extended in the work of Eberlein~\cite{Eberlein}. The basic mechanism in Eberlein's approach is a dynamical Casimir effect: Photons are produced due to a change of the refractive index in the portion of space between the minimum and the maximal bubble radius (a related discussion for time-varying refractive index is due to Yablonovitch~\cite{Yablonovitch}). This physical framework is actually implemented via a boundary between two dielectric media which accelerates with respect to the rest frame of the quantum vacuum state. The adiabatic change in the zero-point modes of the fields reflects in a non-zero radiation flux. In the Eberlein analysis the motion of the bubble boundary is taken into account by introducing a velocity-dependent perturbation to the usual EM Hamiltonian: \begin{eqnarray} H_{\epsilon} &\!=&\! \frac{1}{2} \int{\rm d}^3{\bf r} \left( {{\bf D}^2\over\epsilon} + {\bf B}^2 \right)\;,\:\:\: \Delta H \!=\! \beta \int{\rm d}^3{\bf r} \frac{\epsilon -1}{\epsilon}\; ({\bf D}\wedge{\bf B})\cdot{\bf\hat r}\;. \end{eqnarray} This is an approximate low-velocity result coming from a power series expansion in the speed of the bubble wall $\beta= \dot R/c$. (The bubble wall is known to collapse with supersonic velocity, values of Mach 4 are often quoted, but this is still completely non-relativistic with $\beta\approx 10^{-5}$.) Eberlein's final result for the energy radiated over one acoustic cycle is: \begin{equation} {\cal W} = 1.16\:\frac{(n^2-1)^2}{n^2}\,\frac{1}{480\pi} \left[{\hbar\over c^3}\right] \int_{0}^{T} {\rm d}\tau\; \frac{\partial^5 R^2(\tau)}{\partial \tau^5}\,R(\tau)\beta(\tau)\;. \end{equation} (Eberlein approximates $n_{\mathrm {\mathrm inside}} \approx n_{\mathrm air} \approx 1$ and sets $n_{\mathrm {\mathrm outside}} = n_{\mathrm water} \to n$. The $1.16$ is the result of an integral is estimated numerically.) In this mechanism the massive burst of photons is produced at and near the turn-around at the minimum radius of the bubble. There the velocity rapidly changes sign, from collapse to re-expansion. This means that the acceleration is peaked at this moment, and so are higher derivatives of the velocity. The main points of strength of the Eberlein model are the same as previously listed for the Schwinger model. However, Eberlein's model exhibits a significant weakness (which does not apply to the Schwinger model):\\ The calculation is based on an adiabatic approximation which does not seem consistent with results~\footnote{The adiabatic approximation is actually justified in the case of a model based on the bubble collapse case by the fact that the frequency $\Omega$ of the driving sound (and hence the timescale of the bubble collapse) is of the order of tens of kHz, while that of the emitted light is of the order of $10^{16}$ Hz. The problem we stress here is instead related to the ``self-consistency'' of Eberlein's model.}. In order to fit the experimental values the model requires, as an external input, the bubble radius time dependence. This is expressed as a function of a parameter $\gamma$ which describes the time scale of the collapse and re-expansion process. In order to fit the experimental values for $\cal W$ one has to fix $\gamma \approx 10 \;{\rm fs}$. This is far too short a time to be compatible with the adiabatic approximation. Although one might claim that this number can ultimately be modified by the eventual inclusion of resonances it would seem reasonable to take this ten femtosecond figure as a first self-consistent approximation for the characteristic timescale of the driving system (the pulsating bubble). Unfortunately, the characteristic timescale of the collapsing bubble then comes out to be of the same order of the characteristic period of the emitted photons. This shows that attempts at bootstrapping the calculation into self-consistency instead bring it to a regime where the adiabatic approximation underlying the scheme cannot be trusted. This discussion has lead us to discover a quite intricate situation. We have on the one hand simple estimates of the vacuum energy that can be involved in SL, estimates that are still the object of heated debate, and on the other hand we have a dynamical approach to the problem that seems to be partially self contradictory. In order to resolve the first issue and to understand the proper framework to deal with the second we shall now consider what we can best view as a ``toy model''. In spite of its simplicity this toy model will allow us to capture some basic results that we hope will guide future research on the ``Casimir route'' to sonoluminescence. \section{Bogolubov approach on a single oscillation} Let us consider a single pulsation of the bubble. At this stage of development, we are not concerned with the dynamics of the bubble surface. In analogy with the subtraction procedure of the static calculations of Schwinger~\cite{Sc1,Sc2,Sc3,Sc4,Sc5,Sc6,Sc7} or of Carlson {\em et al.}~\cite{CMMV1,CMMV2,MV} we shall consider two different configurations of space. An ``in'' configuration with a bubble of dielectric constant $\epsilon_{\mathrm {\mathrm inside}}$ (typically vacuum) in a medium of dielectric constant $\epsilon_{\mathrm {\mathrm outside}}$, and an ``out'' one in which one has just the latter medium (dielectric constant $\epsilon_{\mathrm {\mathrm outside}}$) filling all space. These two configurations will correspond to two different bases for the quantization of the field. (For the sake of simplicity we take, as Schwinger did, only the electric part of QED, reducing the problem to a type of scalar electrodynamics). The two bases will be related by Bogolubov coefficients in the usual way. Once we determine these coefficients we easily get the number of created particles per mode and from this the spectrum. We shall also make a consistency check by a direct confrontation between the change in static Casimir energy and the sum, $E=\sum_{k} \omega_{k} n_{k}$, of the energies of the emitted photons. \subsection{Bogolubov coefficients} We use the Schwinger framework. In spherical coordinates and with a time independent dielectric constant \begin{equation} \epsilon {\partial^2\over\partial t^2} E-\nabla^{2} E=0. \end{equation} Solutions are of the form \begin{equation} E=e^{i\omega t} {G(r)\over \sqrt{r}} Y_{lm}(\Omega). \end{equation} Then one finds \begin{equation} G^{''}+{1\over r}G^{'}+\left(\lambda^{2}-{(l+1/2)^2 \over r^{2}} \right)G=0. \end{equation} where $\lambda^2 = \epsilon \omega^2$. This is the standard Bessel equation, it admits as solutions Bessel functions of the first kind, $J_{\nu}(\lambda r)$, and Neumann functions, $N_{\nu}(\lambda r)$ (Bessel functions of the second kind), with $\nu=l+1/2$. For the ``in'' QED vacuum we have to take into account that the dielectric constant changes at the bubble wall. In fact we have \begin{equation} \epsilon=\left \{ \begin{array}{llll} \epsilon_{1} & = & \mbox{dielectric constant of air} & \mbox{if $r< R$},\\ \epsilon_{2} & = & \mbox{dielectric constant of water} & \mbox{if $r >R$.} \end{array} \right. \end{equation} We now use the fact that the dielectric constant of air is approximately equal 1 and shall deal only with the constant of water ($n= \sqrt{\epsilon_{2}} \approx1.3$) For the eigenmodes of the ``in'' state one has \begin{equation} G^{\mathrm in}_{\nu}(n,\omega,r)=\left \{ \begin{array}{ll} A_{\nu} J_{\nu}(\omega_{\mathrm in} r) & \mbox{if $r< R$,}\\ B_{\nu} J_{\nu}(n \omega_{\mathrm in} r)+C_{\nu} N_{\nu}(n \omega_{\mathrm in} r)& \mbox{if $r > R$.} \end{array} \right. \end{equation} The coefficients $A_{\nu}$, $B_{\nu}$ and $C_{\nu}$ are determined by the matching conditions \begin{equation} \begin{array}{lll} A_{\nu} J_{\nu}(\omega_{\mathrm in} R)&=& B_{\nu} J_{\nu}(n \omega_{\mathrm in} R)+C_{\nu}N_{\nu}(n \omega_{\mathrm in} R),\\ A_{\lambda} J_{\nu}{'}(\omega_{\mathrm in} R)&=& B_{\nu} J_{\nu}{'}(n \omega_{\mathrm in} R)+C_{\nu} N_{\nu}{'}(n \omega_{\mathrm in} R). \end{array} \label{coef} \end{equation} The eigenmodes for the ``out'' QED vacuum are easily obtained solving the same equation but for a space filled with an homogeneous dielectric~\footnote{ Keeping $R_{\mathrm min}$ finite significantly complicates the calculation but does not give much more physical information.}. \begin{equation} G^{\mathrm out}_{\nu}(n,\omega_{\mathrm out},r)=J_{\nu}(n \omega_{\mathrm out} r). \end{equation} The Bogolubov coefficients are defined as \begin{eqnarray} \alpha_{ij} &=& ({E_{i}^{\mathrm out}},{E_{j}^{\mathrm in}}),\:\:\: \beta_{ij}=({E_{i}^{\mathrm out}}^{*}, {E_{j}^{\mathrm in}}) \end{eqnarray} where the naive scalar product is as usual~\footnote{There are subtleties in the definition of scalar product which we shall deal with more fully in~\cite{letter,LBVS}. The naive scalar product adopted here is good enough for a qualitative discussion.} \begin{equation} (\phi_{1},\phi_{2})=+i \int_{\Sigma}\phi_{1}^* \stackrel{\leftrightarrow}{\partial}_{0}\phi_{2} \: d^{3}x. \end{equation} We are mainly interested in the coefficient $\beta$ since $|\beta|^{2}$ is linked to the total number of particles created. By a direct substitution it is easy to find \begin{eqnarray} \beta &=& (\omega_{\mathrm in}-\omega_{\mathrm out}) e^{i(\omega_{\mathrm out}+\omega_{\mathrm in})t} \delta_{l l^{\prime}}\delta_{m,-m^{\prime}} \nonumber\\ && \qquad \int_{0}^{\infty} G^{\mathrm out}_{l}(n,\omega_{\mathrm out},r) \; G^{\mathrm in}_{l^{\prime}}(n,\omega_{\mathrm in},r) \: r dr. \label{beta1} \end{eqnarray} After some work, the squared $\beta$ coefficients summed over $l$ and $m$ can be shown to be~\cite{letter,LBVS} \begin{eqnarray} \left|\beta(\omega_{\mathrm in},\omega_{\mathrm out})\right|^{2} &=& \left( \frac{n^2-1}{n^2} \frac{\omega_{\mathrm in}^2 R}{\omega_{\mathrm out}+\omega_{\mathrm in}} \right)^2 \sum_{\nu}(2\nu) \left| A_{\nu} \right|^{2} \nonumber\\ && \times \left[ \frac{ W[J_{\nu}(n\omega_{\mathrm out}r), J_{\nu}(\omega_{\mathrm in}r)]_{R}} {(n\omega_{\mathrm out})^2-\omega_{\mathrm in}^2} \right]^2. \label{E-beta-squared} \end{eqnarray} The number spectrum and total energy content of the emitted photons are \begin{eqnarray} \label{E:E} {dN(\omega_{\mathrm out})\over d\omega_{\mathrm out}} &=&\left(\int|\beta(\omega_{\mathrm in},\omega_{\mathrm out})|^{2} d\omega_{\mathrm in} \right),\\ E&=&\hbar \int {dN(\omega_{\mathrm out})\over d\omega_{\mathrm out}} \; \omega_{\mathrm out} \; d\omega_{\mathrm out}. \end{eqnarray} These expressions are too complex to allow an analytical resolution of the problem (except for the $R\to\infty$ limit). \subsection{Large-volume analytic limit} In this limit the total energy emitted should be approximately equal to the leading contribution in $R$ of the Casimir energy in the ``in'' state (a volume term if Schwinger was qualitatively correct, a surface or curvature one otherwise). Technically, if $R$ is very large (but finite in order to avoid infra-red divergences) then the ``in'' and the ``out'' modes can both be approximated by ordinary Bessel functions: $G^{\mathrm in}(n,\omega,r) \approx J_{\nu}(\omega_{\mathrm in} r)$, $G^{\mathrm out}(n,\omega,r)\approx J_{\nu}(n\omega_{\mathrm out} r)$. The Bogolubov coefficients simplify \begin{eqnarray} \beta_{ij} =({E^{\mathrm out}_{i}}^*,E^{\mathrm in}_{j}) \approx \left(1-\frac{1}{n}\right) e^{i\omega_{\mathrm in}(1/n +1)t} \delta_{ll^{\prime}}\delta_{m,-m^{\prime}} \delta(\omega_{\mathrm in}-n\omega_{\mathrm out}). \label{bogse} \end{eqnarray} This implies \begin{eqnarray} |\beta(\omega_{\mathrm in},\omega_{\mathrm out})|^{2} &\approx& \left(1-\frac{1}{n}\right)^{2} \sum_{l}(2l+1){R\over2\pi c}\delta(\omega_{\mathrm in} -n\omega_{\mathrm out}), \end{eqnarray} where we have invoked the standard scattering theory result that $(\delta^{3}(k))^{2} = V \delta^{3}(k) /(2\pi)^3$, specialized to the fact that we have a 1-dimensional delta function. The summation over angular momenta can be estimated as~\cite{LBVS} \begin{eqnarray} \sum_{l=0}^{l_{\mathrm max}}(2l+1) \approx l_{\mathrm max}^{2}(\omega_{\mathrm out}) &\approx& \left(R \, n \, \frac{\omega_{\mathrm out}}{c}\right)^2. \end{eqnarray} This finally gives \begin{eqnarray} |\beta(\omega_{\mathrm in},\omega_{\mathrm out})|^{2}&\approx &\left(n-1\right)^{2} {R^3\over2\pi c^{3}} \omega_{\mathrm out}^{2} \; \delta(\omega_{\mathrm in}-n\;\omega_{\mathrm out}). \label{Ebog} \end{eqnarray} We can now compute the spectrum and the total energy of the emitted photons \begin{equation} {dN(\omega_{\mathrm out})\over d\omega_{\mathrm out}}\approx n^{2} \; \left({n-1}\over{n}\right)^{2} \; \frac{R^{3}\omega_{\mathrm out}^{2}}{2\pi c^{3}}\; \Theta(K-n \,\omega_{\mathrm out}/c), \end{equation} and \begin{equation} E \approx \frac{1}{8\pi\, n^{2}} \; \left({n-1}\over{n}\right)^{2} \hbar c K \; (R K)^{3}. \label{energy} \end{equation} Hence, inserting our results (\ref{Ebog}) into Eqs.~(\ref{E:E}) for $dN(\omega)/d\omega$ and $E$, we deduce a spectrum that is proportional to phase space (and hence is a power law), up to the cutoff frequency where $n\to 1$. We interpret this as definitive proof that indeed Schwinger was qualitatively right: The main contribution to the Casimir energy of a (large) dielectric bubble is a bulk effect. The total energy radiated in photons balances the change in the Casimir energy up to factors of order one which the present analysis is too crude to detect. (For infinite volume the whole calculation can be re-phrased in terms of plane waves to accurately fix the last few prefactors.) It is important to stress that Eq.~(\ref{Esch}) and Eq.~(\ref{energy}) are not identical (even if in the large $R$ limit the leading term of Casimir energy of the ``in'' state and the total photon energy coincide). One can easy see that the volume term we just found [Eq.~(\ref{energy})] is of second order in $(n-1)$ and not of first order like Eq.~(\ref{Esch}). This is ultimately due to the fact that the interaction term responsible for converting the initial energy in photons is a pairwise squeezing operator~\cite{BLVS}. Equation (\ref{energy}) demonstrates that any argument that attempts to deny the relevance of volume terms to sonoluminescence due to their dependence on $(n-1)$ has to be carefully reassessed. In fact what you measure when the refractive index in a given volume of space changes is {\em not} directly the static Casimir energy of the ``in'' state, but rather the fraction of this static Casimir energy that is converted into photons. We have just seen that once conversion efficiencies are taken into account, the volume dependence is conserved, but not the power in the difference of the refractive index~\footnote{Indeed the dependence of $|\beta|^2$ on $(n-1)^2$ and the symmetry of the former under the interchange of ``in'' and ``out'' state also proves that it is the amount of change in the refractive index and not its ``direction'' (from ``in'' to ``out'') that governs particle production. This also implies that any argument using static Casimir energy balance over a full cycle has to be used very carefully. Actually the total change of the Casimir energy of the bubble over a cycle would be zero (if the final refractive index of the gas is again 1). Nevertheless in the dynamical calculation one gets photon production in both collapse as well expansion phases. (Although some destructive interferences between the photons produced in collapse and in expansion are conceivable, these will not be really effective in depleting photon production because of the substantial dynamical difference between the two phases and because of the, easy to check, fact that most of the photons created in the collapse will be far away from the emission zone by the time the expansion photons would be created.) This apparent paradox is easily solved by taking into account that the main source of energy is the sound field and that the amount of this energy actually converted in photons during each cycle is a very tiny amount of the total power.}. \subsection{Finite-volume numerical estimates:} For finite volume one can no longer rely on analytic results. Fortunately we know that for the total Casimir energy the next subdominant term is a surface area term that is suppressed by a factor of the cutoff wavelength divided by the bubble radius~\cite{MV}. Canonical estimates are: $\lambda_{\mathrm cutoff}/R_{\mathrm max} \approx 0.3 \; \mu {\rm m}/45 \; \mu {\rm m} \approx 1/150$. This suggests that the effects of finite bubble size will not be too dramatic ($1\%$ in total energy?). Applying a mixture of semi-analytic and numerical techniques~\footnote{For details, interested parties are referred to~\cite{LBVS}.} to formula~(\ref{E-beta-squared}) we numerically derive the spectrum $dN/d\omega$ given in Fig.~\ref{fig:thsp}. For comparison we have also plotted the large volume analytic approximation ({\em i.e.}, the leading bulk term by itself). \bfg{4} \begin{picture}(,) \centerline{\psfig{figure=fi.eps,height=8cm,angle=270}} \end{picture} \caption{Spectrum obtained by numerical estimate for finite volume. We have made $R\approx R_{ambient}\approx 4.5 \mu m$ and $\omega_{\mathrm cutoff}\approx 10^{15} Hz$. The sharply peaked curve is that appropriate to the (re-scaled) infinite volume limit (Schwinger limit).} \label{fig:thsp} \efg \subsection{Comment on the calculation} The lessons we have learned from this test calculation are: (1) The model proves (in an indirect way) that the Casimir energy liberated via the bubble collapse includes (in the large $R$ limit) a term proportional to the volume (actually to the volume over which the refractive index changes). In the case of a truly dynamical model one expects that the energy of the photons so created will be provided by other sources of energy ({\em e.g.}, the sound wave), nevertheless the presence of a volume contribution appears unavoidable. (2) In spite of its simplicity (remember that the model is still semi-static), the present calculation is already able to fit some of the experimental requirements, like the shape of the spectrum and the number of emitted photons in the case of $R=R_{\mathrm max}$. Of course the present model is still unable to fully fit other experimental features. For example it provides (like the original Schwinger model) maximal photon release at maximum expansion, and it is able to accommodate only a few arguments to explain the experimental dependencies. This means that a fully dynamical calculation is required in order to deal with these issues, and it is in order to understand what sort of model will ultimately be required that we shall now discuss in detail some basic features of sonoluminescence. \section{Hints towards a truly dynamical model} One of the key features of photon production by a space-dependent and time-dependent refractive index is that for a change occurring on a timescale $\tau$, the amount of photon production is exponentially suppressed by an amount $\exp(-\omega\tau)$. In an Appendix of~\cite{LBVS} we provided a specific toy model that exhibits this behaviour, and argued that the result is in fact generic. The importance for SL is that the experimental spectrum is {\em not\,} exponentially suppressed at least out to the far ultraviolet. Therefore any mechanism of Casimir-induced photon production based on an adiabatic approximation is destined to failure: Since the exponential suppression is not visible out to $\omega\approx10^{15} \; \hbox{Hz}$, it follows that {\em if\,} SL is to be attributed to photon production from a time-dependent refractive index ({\em i.e.}, the dynamical Casimir effect), {\em then} the timescale for change in the refractive index must be of order of a {\em femtosecond}~\footnote{It would be far too naive to assume that femtosecond changes in the refractive index lead to pulse widths limited to the femtosecond range. There are many condensed matter processes that can broaden the pulse width however rapidly it is generated. Indeed, the very experiments that seek to measure the pulse width \cite{Flash1,Flash2} also prove that when calibrated with laser pulses that are known to be of femtosecond timescale, the SL system responds with light pulses on the picosecond timescale.}. Thus any Casimir--based model has to take into account that {\em the change in the refractive index cannot be due just to the change in the bubble radius}. This means that one has to divorce the change in refractive index from direct coupling to the bubble wall motion, and instead ask for a rapid change in the refractive index of the entrained gases as they are compressed down to their van der Waals hard core. Yablonovitch~\cite{Yablonovitch} has emphasized that there are a number of physical processes that can lead to significant changes in the refractive index on a sub-picosecond timescale. In particular, a sudden ionization of the gas compressed in the bubble would lead to an abrupt change, from $1$ to $\approx 0$, of the dielectric constant. Now to get femtosecond changes in refractive index over a distance of about 100 nm (which is the typical length scale of the emission zone), the change in refractive index has to propagate at about $10^8$ metres/sec, about 1/3 lightspeed. To achieve this, one has to adjust basic aspects of the model: we feel that we must move away from the original Schwinger suggestion, in that it is no longer the collapse from $R_{\mathrm max}$ to $R_{\mathrm min}$ that is important. Instead we postulate a rapid (femtosecond) change in refractive index of the gas bubble when it hits the van der Waals hard core~\cite{LBVS}. We stress that this conclusion, though it moves slightly away from the original Schwinger proposal, is still firmly within the realm of the dynamical Casimir effect approach to sonoluminescence. The fact is that the present work shows clearly that a viable Casimir ``route'' to SL cannot avoid a ``fierce marriage'' between QFT and features related to condensed matter physics. It is thus crucial to look for possible unequivocal signatures of the dynamical Casimir effect. To this end it is theoretically possible to have a sharp distinction between any Casimir-like mechanism and other proposals implying a thermal spectrum by looking at the variance of carefully chosen two-photon observables~\cite{BLVS}. As a short example of how this can be done I shall give a brief description of a way to discriminate between thermal photons and two-mode squeezed-state photons (for a more detailed discussion see~\cite{BLVS}). Define the observable \begin{equation} N_{ab} \equiv N_{a}-N_{b}, \end{equation} and its variance \begin{equation} \Delta (N_{ab})^2= \Delta N_{a}^{2}+\Delta N_{b}^{2} -2 \langle N_{a} N_{b}\rangle +2 \langle N_{a}\rangle \langle N_{b} \rangle. \end{equation} The number operators $N_{a},N_{b}$ are intended to denote two photon modes, {\em e.g.} back to back photons. In the case of true thermal light we get \begin{equation} \Delta N_{a}^{2} = \langle N_{a}\rangle(\langle N_{a} \rangle +1), \end{equation} \begin{equation} \langle N_{a} N_{b}\rangle = \langle N_{a}\rangle \langle N_{b}\rangle, \end{equation} so that \begin{equation} \Delta (N_{ab})^2_{\mathrm thermal\ light} =\langle N_{a}\rangle(\langle N_{a}\rangle+1) +\langle N_{b}\rangle(\langle N_{b}\rangle+1). \end{equation} For a two-mode squeezed-state is easy to see~\cite{barnett} \begin{equation} \Delta (N_{ab})^2_{\mathrm two\ mode\ squeezed\ light}=0. \end{equation} In fact due to correlations, $\langle N_{a} N_{b}\rangle \neq \langle N_{a}\rangle \langle N_{b}\rangle$. Note also that, if you measure only a single photon in the pair, you get, as expected, a thermal variance $\Delta N_{a}^{2} = \langle N_{a}\rangle(\langle N_{a}\rangle+1)$. Therefore a measurement of the variance $\Delta (N_{ab})^2$ can be decisive in discriminating if the photons are really thermal or if nonclassical correlations between the photons occur~\cite{bk2}. In~\cite{BLVS,letter,LBVS} it is shown that the arguments just discussed push dynamical Casimir effect models for SL into a rather constrained region of parameter space and predict some typical ``signatures'' for it. This allows to hope that these ideas will become experimentally testable in the near future. \section{Discussion and Conclusions} The present calculation unambiguously verifies that a change of the refractive index in a given volume of space is, as predicted by Schwinger~\cite{Sc4,CMMV1,CMMV2,MV}, converted into real photons with a phase space spectrum. We have also explained why such a change must be sudden in order to fit the experimental data. This leads us to propose a somewhat different model of SL based on the dynamical Casimir effect, a model focussed this time on the actual dynamics of the refractive index (as a function of space and time) and not just of the bubble boundary (in Schwinger's original approach the refractive index changes only due to motion of the bubble wall). This proposal shares the generic points of strength attributable to the Casimir route but it is now in principle able to implement the required sudden change in the refractive index. In summary, provided the sudden approximation is valid, changes in the refractive index will lead to efficient conversion of zero point fluctuations into real photons. Trying to fit the details of the observed spectrum in sonoluminescence then becomes an issue of building a robust model of the refractive index of both the ambient water and the entrained gases as functions of frequency, density, and composition. Only after this prerequisite is satisfied will we be in a position to develop a more complex dynamical model endowed with adequate predictive power. In light of these observations we think that one can also derive a general conclusion about the long standing debate on the actual value of the static Casimir energy and its relevance to sonoluminescence: Sonoluminescence is not directly related to the {\em static} Casimir effect. The static Casimir energy is at best capable of giving a crude estimate for the energy budget in SL. We hope that this work will convince everyone that only models dealing with the actual mechanism of particle creation (a mechanism which must have the general qualities discussed in this article) will be able to eventually prove, or disprove, the pertinence of the physics of the quantum vacuum to Sonoluminescence. \section*{References}
\section{Introduction} The investigation of conservation laws in general relativity has a long history. From its very beginning much of this research was based on pseudo--tensors instead of fully covariant methods. The aim of many of these works was to find differential laws which, once transformed into integral ones, were interpreted as energy balances in such a way that expressions for the energy and momentum densities of the gravitational field could be identified. The covariant approach to this problem is mainly based on the analogy of the Bel--Robinson tensor \cite{BEL1} with the energy--momentum tensor of the electromagnetic field (there are other approaches based on this analogy, see for instance~\cite{MMCQ}). The Bel--Robinson tensor is conserved {\em in vacuum}, completely symmetric and traceless. Moreover, the completely timelike component referred to any observer (described by a timelike unit vector field) is non--negative, and its vanishing implies that the space--time is conformally flat (flat in vacuum). This is a desirable positivity property for any candidate to gravitational energy density. In spite of these good properties, the Bel--Robinson tensor has dimensions of energy density square and this fact makes its interpretation somewhat unclear. Nevertheless, it has revealed as a very useful tool in many kinds of studies, which has led to some efforts in finding extensions of the Bel--Robinson tensor for more general cases than vacuum. Therefore, the question arises whether generalizations of the Bel--Robinson tensor exist for space--times not necessarily empty. The Bel tensor \cite{BEL2} was the first attempt on this problem. It is a tensor whose completely timelike component is positive and zero only when the space--time is Minkowski. In vacuum, it reduces to the Bel--Robinson tensor, but in the general case it is no longer conserved. For general space--times, Sachs \cite{SACH} found a divergence--free tensor that coincides with the Bel--Robinson tensor in vacuum. Unfortunately, this tensor does not satisfy any positivity property and it is neither completely symmetric nor traceless. The systematic treatment of this problem was made by Collinson \cite{COLL}, who found all four--index divergence--free tensors with terms either quadratic in the Riemann tensor or linear in its second derivatives. The result found out is that any such tensor can be derived from only {\em one} tensor, namely $T_{10}^{\alpha\beta\lambda\mu}$ (see (\ref{colli}) in the appendix A), whose divergence with respect to the first index vanishes. However, there is not enough freedom to construct a tensor with its time component positive. In this paper we show that, unlike the general case, for space--times with zero scalar curvature ($R=0$) it is possible to construct a unique generalization of the Bel--Robinson tensor. We begin in Sect. 2 by proving that, when $R=0$, there exists another conserved tensor which cannot be derived from the Collinson one. In Sect. 3 we show that, demanding symmetry in the three free indices, there is not any other tensor independent from these two. This new tensor allows us to construct (Sect. 4) a divergence--free tensor which has the completely timelike component non--negative and zero only when the space--time is flat (excluding some cases that, via Einstein's field equations, have an unphysical matter content). This tensor is completely symmetric in its three last indices, but it is impossible to get a similar tensor symmetric in all their indices. We remark that it is not possible to construct any other tensor with such characteristics. In order to illustrate this development, we study in Sect.~5 some examples in which we can define this tensor, namely: the Friedmann--Lema\^{\i}tre--Robertson--Walker (FLRW) models with a energy--momentum content of (incoherent) radiation ($p=\varrho/3$), the plane--fronted gravitational waves with parallel rays ({\em pp waves}), and the Vaidya radiating space--time. Finally, we recall that for purely electromagnetic space--times, and supposing that the Einstein field equations hold, Penrose and Rindler \cite{PERI} also gave a generalization of the Bel--Robinson tensor by using spinor methods. This tensor is conserved, completely symmetric and traceless in its three last indices. In the appendix B we find its tensorial expression and a new (up to our knowledge) positivity property. \section{Deduction of the new conserved tensor for $R=0$} Otherwise said, throughout this paper we will consider the metric tensor $g_{\alpha\beta}$ to have signature $(-,+,+,+)$. The convention for indices on the Riemann tensor that will be used is defined through the Ricci identities: \begin{eqnarray} \left(\nabla_{\alpha}\nabla_{\beta}- \nabla_{\beta}\nabla_{\alpha}\right)v_{\lambda}= -R^{\sigma}_{\hspace{1mm}\lambda\alpha\beta}v_{\sigma} \, , \label{ricci} \end{eqnarray} where $v_{\alpha}$ is an arbitrary 1--form. The Ricci tensor and the scalar curvature are defined as usual: $R_{\alpha\beta}\equiv R^{\sigma}_{\hspace{1mm}\alpha\sigma\beta}$ and $R\equiv R^{\sigma}_{\sigma}$. We also recall the Riemann symmetries and the first and second Bianchi identities: \begin{eqnarray} & R_{\alpha\beta\lambda\mu}=R_{[\alpha\beta][\lambda\mu]} =R_{\lambda\mu\alpha\beta}\, , \nonumber \\ & \hspace{1cm} R_{[\alpha\beta\lambda]\mu}=0 \, , \nonumber\\ & \nabla_{[\nu}R_{\alpha\beta]\lambda\mu}=0 \, . \label{bianchi} \end{eqnarray} The procedure we are going to use to find the conserved tensor starts from the expression for the divergence of the Bel tensor \cite{BEL1}: \begin{eqnarray} \nabla_{\alpha}T^{\alpha\beta\lambda\mu}= R^{\beta\hspace{1.5mm}\lambda}_{\hspace{1mm}\rho \hspace{2mm}\sigma}J^{\mu\sigma\rho}+ R^{\beta\hspace{1.5mm}\mu}_{\hspace{1mm}\rho \hspace{2mm}\sigma}J^{\lambda\sigma\rho} -\frac{1}{2}g^{\lambda\mu}R^{\beta}_{\hspace{1mm} \rho\sigma\gamma}J^{\sigma\gamma\rho} \, , \label{divbel} \end{eqnarray} where the Bel tensor $T^{\alpha\beta\lambda\mu}$ and $J^{\alpha\beta\lambda}$ are defined as follows: \begin{eqnarray} T^{\alpha\beta\lambda\mu}&\equiv& \frac{1}{2}\left(R^{\alpha\rho\lambda\sigma} R^{\beta\hspace{1.5mm}\mu}_{\hspace{1mm} \rho\hspace{2mm}\sigma}+ {*R*}^{\alpha\rho\lambda\sigma} {*R*}^{\beta\hspace{1.5mm}\mu}_{\hspace{1mm} \rho\hspace{2mm}\sigma}+ {*R}^{\alpha\rho\lambda\sigma} {*R}^{\beta\hspace{1.5mm}\mu}_{\hspace{1mm} \rho\hspace{2mm}\sigma}+ {R*}^{\alpha\rho\lambda\sigma} {R*}^{\beta\hspace{1.5mm}\mu}_{\hspace{1mm} \rho\hspace{2mm}\sigma}\right) \, , \nonumber\\ J^{\lambda\mu\beta}&\equiv& \nabla^{\lambda}R^{\mu\beta}- \nabla^{\mu}R^{\lambda\beta}= \nabla_{\sigma}R^{\mu\lambda\beta\sigma} \, , \label{J} \end{eqnarray} being ``$\ast$'' the usual dual operator acting over any pair of antisymmetric indices: \begin{eqnarray*} & {*R}_{\alpha\beta\lambda\mu}\equiv \frac{1}{2} \eta_{\alpha\beta\sigma\rho} R^{\sigma\rho}_{\hspace{3mm}\lambda\mu} \, , \;\; {R*}_{\alpha\beta\lambda\mu}\equiv \frac{1}{2} \eta_{\lambda\mu\sigma\rho} R_{\alpha\beta}^{\hspace{3mm}\sigma\rho} \, , \\ & {*R*}_{\alpha\beta\lambda\mu}\equiv \frac{1}{4} \eta_{\alpha\beta\gamma\delta} \eta_{\lambda\mu\sigma\rho} R^{\gamma\delta\sigma\rho} \, , \end{eqnarray*} and $\eta_{\alpha\beta\lambda\mu}$ is the canonical volume 4--form. Our purpose now is to work out the right hand side (r.h.s.) of equation (\ref{divbel}) in order to convert it into a global divergence. To that end, we will repeatedly integrate by parts and make use of equations (\ref{ricci},\ref{bianchi}). To begin with, the last term in equation (\ref{divbel}) can be easily transformed into a divergence, by means of the Ricci and Bianchi identities (\ref{ricci},\ref{bianchi}): \begin{eqnarray} -\frac{1}{2}g^{\lambda\mu}R^{\beta}_{\hspace{1mm} \rho\sigma\gamma}J^{\sigma\gamma\rho}= -g^{\lambda\mu}\nabla_{\alpha}\nabla_{\rho} J^{\alpha\rho\beta} \, . \label{1term} \end{eqnarray} Next we expand the leading terms of the r.h.s. of (\ref{divbel}) by using the definition of $J^{\alpha\beta\lambda}$ (\ref{J}) and integrating by parts: \begin{eqnarray} R^{\beta\hspace{1.5mm}\lambda}_{\hspace{1mm}\rho \hspace{2mm}\sigma}\nabla^{\mu}R^{\sigma\rho}- \nabla_{\alpha}\left(R^{\mu\rho} R^{\beta\hspace{2mm}\lambda\alpha}_{ \hspace{1mm}\rho}\right)+ R^{\mu\rho}\nabla_{\alpha} R^{\beta\hspace{2mm}\lambda\alpha}_{ \hspace{1mm}\rho}+ \left[\lambda \longleftrightarrow \mu\right] \, . \label{2term} \end{eqnarray} The first term in the previous expression can be rewritten by means of the Ricci identities (\ref{ricci}) as follows: \begin{eqnarray} R^{\beta\hspace{1.5mm}\lambda}_{\hspace{1mm}\rho \hspace{2mm}\sigma}\nabla^{\mu}R^{\sigma\rho}= \left(\nabla^{\lambda}\nabla_{\sigma}- \nabla_{\sigma}\nabla^{\lambda}\right) \nabla^{\mu}R^{\sigma\beta}+ \nonumber\\ +\left[-\nabla^{\alpha}\left( R^{\mu\hspace{1.5mm}\lambda}_{\hspace{1mm}\alpha \hspace{2mm}\sigma}R^{\sigma\beta}\right)+ R^{\sigma\beta}\nabla^{\alpha} R^{\lambda\hspace{1.5mm}\mu}_{\hspace{1mm}\sigma \hspace{2mm}\alpha}\right] +R_{\sigma}^{\lambda}\nabla^{\mu}R^{\sigma\beta} \, . \label{3term} \end{eqnarray} Thus, we have converted the r.h.s. of equation (\ref{divbel}) into a divergence plus the following terms: \begin{eqnarray} R^{\sigma\beta}\nabla^{\alpha} R^{\lambda\hspace{1.5mm}\mu}_{\hspace{1mm}\sigma \hspace{2mm}\alpha}+ R^{\sigma\lambda}\nabla^{\mu}R_{\sigma}^{\beta}+ R^{\mu\rho}\nabla_{\alpha} R^{\beta\hspace{2mm}\lambda\alpha}_{ \hspace{1mm}\rho}+ \left[\lambda \longleftrightarrow \mu\right]= \nonumber \\ =R^{\sigma\beta}\left(\nabla_{\sigma}R^{\lambda\mu} -\nabla^{\lambda}R_{\sigma}^{\mu}\right) +R^{\sigma\lambda}\nabla^{\mu}R_{\sigma}^{\beta}+ R^{\mu\rho}\left(\nabla_{\rho}R^{\beta\lambda} -\nabla^{\beta}R_{\rho}^{\lambda}\right) +\left[\lambda \longleftrightarrow \mu\right] \, . \label{4term} \end{eqnarray} Now, taking into account the contracted Bianchi identities ($\nabla_{\mu}R^{\mu\nu}= \frac{1}{2}\nabla^{\nu}R$), these terms can be transformed into the following expression: \begin{eqnarray} -\nabla^{\beta}\left(R_{\sigma}^{\lambda} R^{\sigma\mu}\right)+ \nabla^{\mu}\left(R_{\sigma}^{\lambda} R^{\sigma\beta}\right)+ \nabla^{\lambda}\left(R_{\sigma}^{\mu} R^{\sigma\beta}\right) -2 R^{\beta\sigma}\left(\nabla^{\mu}R_{\sigma}^{\lambda}+ \nabla^{\lambda}R_{\sigma}^{\mu}\right)+ \nonumber\\ +\nabla_{\sigma}\left(R^{\beta\lambda}R^{\mu\sigma}+ R^{\beta\mu}R^{\lambda\rho}+ 2 R^{\beta\sigma}R^{\lambda\mu}\right)- \frac{1}{2}\left(R^{\beta\lambda}\nabla^{\mu}R+ R^{\beta\mu}\nabla^{\lambda}R+ 2 R^{\lambda\mu}\nabla^{\beta}R\right) \, . \label{5term} \end{eqnarray} The last three terms of this expression vanish when $R$ is constant, so we are finally left with $-2R^{\beta\sigma} \left(\nabla^{\mu}R_{\sigma}^{\lambda}+ \nabla^{\lambda}R_{\sigma}^{\mu}\right)$. Nevertheless, notice that our final purpose is to find a conserved tensor ${T''}^{\alpha\beta\lambda\mu}$ whose completely timelike component referred to an observer $\vec{u}$, ${T''}^{\alpha\beta\lambda\mu} u_{\alpha}u_{\beta}u_{\lambda}u_{\mu}$, is positive, which means that we are only interested in the symmetric part. Therefore, without lost of generality, we can symmetrize the whole expression and, as a consequence, the remaining terms transform themselves into a divergence: \begin{eqnarray} -4R^{\sigma(\beta} \nabla^{\lambda}R_{\sigma}^{\mu)}= -2\nabla^{(\beta}\left(R_{\sigma}^{\lambda} R^{\mu)\sigma}\right) \, . \label{6term} \end{eqnarray} So, we have finally achieved a conserved tensor if the scalar curvature vanishes (in fact, if it is constant). Collecting all the previous terms (\ref{1term}-\ref{6term}) we get the final result: \begin{eqnarray*} \nabla_{\alpha}{T''}^{\alpha\beta\lambda\mu}= -2 R^{(\beta\lambda}\nabla^{\mu)}R \, , \end{eqnarray*} where we have defined \begin{eqnarray} {T''}^{\alpha\beta\lambda\mu}&\equiv& {T}^{\alpha(\beta\lambda\mu)}- 4 R^{\alpha(\beta}R^{\lambda\mu)}+ g^{\alpha(\beta}R^{\lambda}_{\sigma}R^{\mu)\sigma}- \nonumber\\ &-& 2\nabla^{(\beta}\nabla^{\lambda}R^{\mu)\alpha}+ 2\nabla^{(\beta}\nabla^{|\alpha|}R^{\lambda\mu)}- 2\nabla^{\alpha}\nabla^{(\beta}R^{\lambda\mu)}- \nonumber\\ &-& 2 g^{\alpha(\beta}\nabla_{\sigma}\nabla^{\lambda} R^{\mu)\sigma}- \nabla_{\sigma}\nabla^{\sigma}R^{\alpha(\beta} g^{\lambda\mu)}+ \nabla_{\sigma}\nabla^{\alpha}R^{\sigma(\beta} g^{\lambda\mu)} \, . \label{T''} \end{eqnarray} It is a matter of checking that this tensor cannot be obtained from Collinson tensor $T_{10}^{\alpha\beta\lambda\mu}$ when it is restricted to the case $R=0$. Now, we are left with the question of its uniqueness. \section{Uniqueness} In this section we will prove that, in the case we are concerned with ($R=0$), it does not exist any other conserved tensor, symmetric in its three last indices, independent from ${T''}^{\alpha\beta\lambda\mu}$ and the Collinson tensor, ${T}_{10}^{\alpha\beta\lambda\mu}$. The reasoning is the following. Suppose you are given a tensor ${T'''}^{\alpha\beta\lambda\mu}$ which is conserved when $R=0$. The divergence computed in the general case will be a combination of the following type (taking into account symmetries and unit dimensions): \begin{eqnarray*} \nabla_{\alpha}{T'''}^{\alpha\beta\lambda\mu}= a R^{(\beta\lambda}\nabla^{\mu)}R+ b R\nabla^{(\beta}R^{\lambda\mu)}+ c R^{\sigma(\beta}\nabla_{\sigma}R g^{\lambda\mu)}+ d R g^{(\beta\lambda}\nabla^{\mu)}R+ \\ + e \nabla^{(\beta}\nabla^{\lambda}\nabla^{\mu)}R+ f \nabla^{\sigma}\nabla_{\sigma}\nabla^{(\beta}R g^{\lambda\mu)}+ h \nabla^{\sigma}\nabla^{(\beta}\nabla_{\sigma}R g^{\lambda\mu)}+ i \nabla^{(\beta}\nabla^{|\sigma|}\nabla_{\sigma}R g^{\lambda\mu)} \, , \end{eqnarray*} $a$, $b$, $c$, $d$, $e$, $f$, $h$ and $i$ being constants. This can be immediately cast in the following form: \begin{eqnarray*} \nabla_{\alpha}{T'''}^{\alpha\beta\lambda\mu}= (a-b)R^{(\beta\lambda}\nabla^{\mu)}R+ \nabla_{\alpha}\tau^{\alpha\beta\lambda\mu} \, , \end{eqnarray*} where $\tau^{\alpha\beta\lambda\mu}$ stands for: \begin{eqnarray*} \tau^{\alpha\beta\lambda\mu}\equiv b R g^{\alpha(\beta}R^{\lambda\mu)}+ c \left(R R^{\alpha(\beta}g^{\lambda\mu)}- \frac{1}{4} g^{\alpha(\beta}g^{\lambda\mu)}R^{2}\right)+ d \frac{1}{2} g^{(\beta\lambda}g^{\mu)\alpha}R^{2}+ \\ + e g^{\alpha(\beta}\nabla^{\lambda}\nabla^{\mu)}R+ f g^{\alpha\sigma}\nabla_{\sigma}\nabla^{(\beta}R g^{\lambda\mu)}+ h g^{\alpha\sigma}\nabla^{(\beta}\nabla_{\sigma}R g^{\lambda\mu)}+ i g^{\alpha(\beta}\nabla^{|\sigma|}\nabla_{\sigma}R g^{\lambda\mu)} \, , \end{eqnarray*} so clearly it is a tensor that vanishes when $R$ does. That is, if a tensor of the kind we are considering is divergence--free when $R=0$, in the general case its divergence should be a multiple of $R^{(\beta\lambda}\nabla^{\mu)}R$ plus the divergence of a tensor of the type $\tau^{\alpha\beta\lambda\mu}$. If we had two such tensors, a suitable combination of them removing the term $R^{(\beta\lambda}\nabla^{\mu)}R$ would give a conserved tensor for the general case and, therefore, due to the Collinson result \cite{COLL}, it could be constructed from $T_{10}^{\alpha\beta\lambda\mu}$. Given that the tensor $\tau^{\alpha\beta\lambda\mu}$ vanishes when $R=0$, the three tensors would not be independent in that case. On the other hand, this reasoning shows that, from the very beginning, we were able to know that in the $R=0$ case at most one more conserved tensor could exist apart from Collinson's one, as finally has been the case. \section{Positivity} As it has been pointed out above, in the general case all the conserved tensors can be constructed from $T_{10}^{\alpha\beta\lambda\mu}$. This construction is based in two procedures. First, it is clear that if we perform any permutation on the three last indices we will still have a conserved tensor. Second, by taking the two traces $T_{10\hspace{2.5mm}\rho}^{\alpha\beta\rho}$, $T_{10\hspace{2.5mm}\rho}^{\alpha\rho\beta}$ and multiplying them by $g^{\lambda\mu}$ we obtain new conserved tensors. Actually, there is not any other two--index divergence--free tensor independent from them. These two tensors can be taken to be (as usually obtained by hamiltonian differentiation): \begin{eqnarray*} t_{1}^{\alpha\beta}&=&2\nabla^{\alpha}\nabla^{\beta}R- 2g^{\alpha\beta}\nabla_{\mu}\nabla^{\mu}R+ \frac{1}{2}g^{\alpha\beta}R^{2}- 2 R R^{\alpha\beta} \, , \\ t_{2}^{\alpha\beta}&=& 2\nabla_{\sigma}\nabla^{\alpha}R^{\sigma\beta}- \nabla_{\sigma}\nabla^{\sigma}R^{\alpha\beta}- 2R^{\alpha\sigma}R^{\beta}_{\sigma}+ \frac{1}{2}R_{\sigma\rho}R^{\sigma\rho}g^{\alpha\beta}- \frac{1}{2}g^{\alpha\beta}\nabla_{\mu}\nabla^{\mu}R \, . \end{eqnarray*} This is all the freedom we have in the general case. In the $R=0$ case, there exists only one two--index conserved tensor, which is that obtained from $t_{2}^{\alpha\beta}$. On the other hand, as we consider tensors which are symmetric in the three last indices, from the Collinson tensor we will only have one independent tensor (apart from the two--index tensor), namely $T_{10}^{\alpha(\beta\lambda\mu)}$ or, equivalently, the Sachs tensor $T'^{\alpha\beta\lambda\mu}$ \cite{SACH} which is symmetric in its three last indices. Therefore, to construct the super--energy tensor in the $R=0$ case, we are led with three tensors: {\em i)} the Sachs tensor $T'^{\alpha\beta\lambda\mu}$ (restricted to the $R=0$ case): \begin{eqnarray*} {T'}^{\alpha\beta\lambda\mu}&\equiv& {T}^{\alpha(\beta\lambda\mu)}+ 2 R^{\alpha\sigma}R^{(\beta}_{\sigma}g^{\lambda\mu)} -\frac{1}{3} g^{\alpha(\beta}R^{\lambda}_{\sigma}R^{\mu)\sigma} -\frac{1}{2} R_{\sigma\rho}R^{\sigma\rho} g^{\alpha(\beta}g^{\lambda\mu)}+\\ &+&2\nabla^{(\beta}\nabla^{\lambda}R^{\mu)\alpha}- \frac{2}{3}\nabla^{(\beta}\nabla^{|\alpha|}R^{\lambda\mu)}- \frac{2}{3}\nabla^{\alpha}\nabla^{(\beta}R^{\lambda\mu)}+\\ &+&\frac{4}{3}\nabla_{\sigma}\nabla^{\sigma}R^{(\beta\lambda} g^{\mu)\alpha} -2 g^{\alpha(\beta} \nabla_{\sigma}\nabla^{\lambda}R^{\mu)\sigma}- \nabla_{\sigma}\nabla^{\alpha}R^{\sigma(\beta} g^{\lambda\mu)} \, , \end{eqnarray*} {\em ii)} the two--index tensor $t^{\alpha\beta}$: \begin{eqnarray*} t^{\alpha\beta}=2\nabla_{\sigma}\nabla^{\alpha}R^{\sigma\beta}- \nabla_{\sigma}\nabla^{\sigma}R^{\alpha\beta}- 2R^{\alpha\sigma}R^{\beta}_{\sigma}+ \frac{1}{2}R_{\sigma\rho}R^{\sigma\rho}g^{\alpha\beta} \, , \end{eqnarray*} and {\em iii)} the tensor $T''^{\alpha\beta\lambda\mu}$ previously found in (\ref{T''}). Now, we have to combine these three tensors in such a way that any observer measures a positive quantity. Moreover, we would like that this completely timelike component vanishes if and only if the space--time is flat. First of all, we have to take into account that terms made of derivatives of the Ricci tensor do not have a definite sign, so it would be necessary to eliminate their contributions. This aim can only be achieved by means of the following combination: \begin{eqnarray*} A^{\alpha\beta\lambda\mu}\equiv \frac{1}{2}\left( 3{T'}^{\alpha\beta\lambda\mu}- {T''}^{\alpha\beta\lambda\mu}\right)+ \frac{5}{2}t^{\alpha(\beta} g^{\lambda\mu)}\, , \end{eqnarray*} which more explicitly reads: \begin{eqnarray} A^{\alpha\beta\lambda\mu}&=&T^{\alpha(\beta\lambda\mu)}+ \nonumber \\ &+&2R^{\alpha(\beta}R^{\lambda\mu)}- 2R^{\alpha\sigma}R^{(\beta}_{\sigma}g^{\lambda\mu)}- g^{\alpha(\beta}R^{\lambda}_{\sigma}R^{\mu)\sigma}+ \frac{1}{2}R_{\sigma\rho}R^{\sigma\rho} g^{\alpha(\beta}g^{\lambda\mu)}+ \nonumber \\ &+&2\nabla^{(\beta}\nabla^{\lambda}R^{\mu)\alpha}- 2\nabla^{(\beta}\nabla^{|\alpha|}R^{\lambda\mu)} +3\nabla_{\sigma}\nabla^{\alpha}R^{\sigma(\beta}g^{\lambda\mu)}- \nonumber \\ &-&2g^{\alpha(\beta}\nabla_{\sigma}\nabla^{\lambda}R^{\mu)\sigma}+ 2\nabla_{\sigma}\nabla^{\sigma}R^{(\beta\lambda}g^{\mu)\alpha}- 2\nabla_{\sigma}\nabla^{\sigma}R^{\alpha(\beta}g^{\lambda\mu)} \,. \label{tena} \end{eqnarray} Since we have already exhausted all the freedom, we finally examine the completely timelike component referred to any timelike unit vector $\vec{u}$: \begin{eqnarray} A\left(\vec{u}\right)&\equiv& A_{\alpha\beta\lambda\mu}u^{\alpha}u^{\beta}u^{\lambda}u^{\mu}= T_{\alpha\beta\lambda\mu}u^{\alpha}u^{\beta}u^{\lambda}u^{\mu}+ \frac{1}{2}R_{\sigma\rho}R^{\sigma\rho}+ \nonumber \\ &+&2\left(R_{\alpha\beta}u^{\alpha}u^{\beta}\right)^{2}+ 3\left(R^{\alpha\sigma}R^{\beta}_{\sigma}\right) u_{\alpha}u_{\beta}- \left(\nabla_{\sigma}\nabla^{\alpha}R^{\sigma\beta}\right) u_{\alpha}u_{\beta} \, . \label{a} \end{eqnarray} To check the positivity of $A\left(\vec{u}\right)$ it is convenient to write out the last term of the previous expression in the following form: \begin{eqnarray*} \nabla_{\sigma}\nabla^{\alpha}R^{\sigma\beta}= C^{\alpha\hspace{3mm}\beta}_{\hspace{2mm}\sigma\rho} R^{\sigma\rho}+ 2R^{\alpha\sigma}R^{\beta}_{\sigma}- \frac{1}{2}g^{\alpha\beta}R_{\sigma\rho}R^{\sigma\rho} \, . \end{eqnarray*} At this point, we introduce four spatial tensors, namely $E_{\alpha\beta}(\vec{u})$, $H_{\alpha\beta}(\vec{u})$, $M_{\alpha\beta}(\vec{u})$ and $N_{\alpha\beta}(\vec{u})$, that (together with $R$) wholly characterize the Riemann tensor \cite{BOSE}. Their definitions, properties and some useful formulae are given in the appendix A. Introducing the previous definitions in (\ref{a}) we obtain, after some calculations: \begin{eqnarray} A\left(\vec{u}\right)= \left(E_{\sigma\rho}-M_{\sigma\rho}\right) \left(E^{\sigma\rho}-M^{\sigma\rho}\right)+ H_{\sigma\rho}H^{\sigma\rho}+ 3N_{\sigma\rho}N^{\sigma\rho}+ \left(R_{\alpha\beta}u^{\alpha}u^{\beta}\right)^{2} \,, \label{comt} \end{eqnarray} which is a sum of square terms (all the tensors appearing here are spatial). Therefore it is manifestly positive and its vanishing implies: \begin{eqnarray*} H_{\alpha\beta}=0\, , \; \; N_{\alpha\beta}=0 \, ,\\ R_{\alpha\beta}u^{\alpha}u^{\beta}=0 \, , \; \; E_{\alpha\beta}=M_{\alpha\beta} \, . \end{eqnarray*} The previous expressions lead to $R_{\alpha\beta}u^{\alpha}=0$ (see appendix A). This condition, when considering the Einstein field equations, immediately drives to an unphysical energy--momentum tensor. Hence, if we eliminate these unphysical space--times (for instance, adding any energy condition), the vanishing of $A\left(\vec{u}\right)$ finally implies the Minkowski space--time. \section{Some examples} In this section we are going to study the tensor $A^{\alpha\beta\lambda\mu}$ (\ref{tena}) in some space--times with vanishing scalar curvature. In particular, we are going to consider the following examples: (i) the radiation FLRW cosmological models, (ii) the {\em pp waves}, and (iii) the Vaidya radiating metric. In examples (i) and (ii) we will give the expression for $A^{\alpha\beta\lambda\mu}$ and its completely timelike component $A(\vec{u})$ (\ref{comt}) for an arbitrary observer $\vec{u}$. In the example (iii), for the sake of brevity we will give only the expression of $A(\vec{u})$, also for an arbitrary observer. \vspace{4mm} In the first example we study the case of the FLRW models (see for instance \cite{KSHM}) with vanishing scalar curvature, the radiation models, whose energy--momentum tensor (of perfect--fluid type) is given by the following expression (throughout this section we will use units in which $8\pi G=c=1$) \[ T_{\alpha\beta} = \varrho \,U_\alpha U_\beta + p\, h_{\alpha\beta}\,, \hspace{14mm} p=\frac{1}{3}\varrho \,,\] where $\vec{U}$ is the fluid velocity ($U^\alpha U_\alpha=-1$), $\varrho$ the energy density, $p$ the pressure, and $h_{\alpha\beta}=g_{\alpha\beta}+U_\alpha U_\beta$ the orthogonal projector to the fluid velocity. The line element of these conformally--flat models can be written as \[ ds^2=-dt^2+a^2(t)\left\{d\chi^2+\Sigma^2(\epsilon,\chi) \,(d\theta^2+\sin^2\theta d\varphi^2)\right\}\,, \label{lerw}\] where $\Sigma(\epsilon,\chi)$ is given by \[ \Sigma(\epsilon,\chi) = \left\{ \begin{array}{ll} \sin\chi & \mbox{if $\epsilon = 1$\, ,} \\ \chi & \mbox{if $\epsilon = 0$\, ,} \\ \sinh\chi & \mbox{if $\epsilon = -1$\,.}\end{array}\right.\] The fluid velocity $\vec{U}$, the scale factor $a(t)$ and the energy density $\varrho(t)$ are \[\vec{U}=\frac{\partial}{\partial t}\,, \hspace{8mm} a^2(t)= (t-t_o)\left[2A-\epsilon(t-t_o)\right] \,, \hspace{8mm} \varrho(t) = \frac{3A^2}{a^4(t)} \,,\] respectively, and $A$ and $t_o$ are arbitrary constants. After some straightforward calculations, and using the special properties of these space--times, we arrive at the following expression for $A^{\alpha\beta\lambda\mu}$ \begin{equation} A^{\alpha\beta\lambda\mu} = \frac{4}{3}\varrho^2\left\{ U^\alpha U^\beta U^\lambda U^\mu + U^\alpha U^{(\beta} h^{\lambda\mu)} + \frac{5}{3}h^{\alpha(\beta}U^\lambda U^{\mu)}+ \frac{1}{3}h^{\alpha(\beta}h^{\lambda\mu)}\right\}\,. \label{afrw} \end{equation} As we can see, it is proportional to the energy density squared. We can also check that it is indeed divergence--free. Now, let us compute the completely timelike component (\ref{comt}) of this tensor with respect to an arbitrary observer $\vec{u}$. To that end, we decompose $\vec{u}$ in the next way \[ \vec{u} = \gamma(\vec{U}+\vec{v})\,, \hspace{6mm} v^\alpha U_\alpha =0\,, \hspace{4mm} v^\alpha v_\alpha = v^2 \geq 0\,, \hspace{4mm} \gamma\equiv (1-v^2)^{-1/2} \,, \] where the case $\vec{v}=0$ corresponds to an observer comoving with the fluid ($\vec{u}=\vec{U}$). Then, from (\ref{comt},\ref{afrw}) we find that $A(\vec{u})$ is given by \[ A(\vec{u}) = \frac{4}{3}\varrho^2\gamma^4\left\{ 1+\frac{8}{3}v^2+\frac{1}{3}v^4\right\} \,. \] That is, it is function of $\varrho$ and $v$ only. Moreover, it increases monotonically as $v$ increases and its minimum corresponds to the case $v=0$, in which the observer is comoving with the fluid. \vspace{4mm} Now, we are going to study the tensor $A^{\alpha\beta\lambda\mu}$ in the case of the {\em pp waves} space--times. The corresponding line element can be written in null coordinates $\{u,v,\zeta,\bar{\zeta}\}$ as follows (see \cite{KSHM} for details) \[ ds^2 = -2 du dv + 2 d\zeta d\bar{\zeta} -2 H du^2\,,\] where $H$ is an arbitrary function which does not depend on $v$ [$H=H(u,\zeta,\bar{\zeta})$]. Using the following Newman--Penrose basis $\{\ell,k,m,\bar{m}\}$ \[ \vec{\ell}=\frac{\partial}{\partial v}\,,\hspace{5mm} \vec{k} = \frac{\partial}{\partial u}-H \frac{\partial}{\partial v}\,,\hspace{5mm} \vec{m}= \frac{\partial}{\partial \zeta}\,, \] the Ricci and self--dual Weyl tensors are \begin{equation} R_{\alpha\beta}=2\Phi\ell_\alpha\ell_\beta\,,\hspace{1cm} \hat{C}_{\alpha\beta\lambda\mu}\equiv C_{\alpha\beta\lambda\mu}+i\stackrel{*}{C}_{\alpha\beta\lambda\mu} = 2\Psi_4 V_{\alpha\beta} V_{\lambda\mu} \,, \label{riwe} \end{equation} respectively. Where the quantities $\Phi$, $\Psi_4$ and $V_{\alpha\beta}$ are given by \[ \Phi\equiv H_{,\zeta\bar{\zeta}}\,,\hspace{6mm} \Psi_4 = H_{,\bar{\zeta}\bar{\zeta}}\,,\hspace{6mm} V_{\alpha\beta}\equiv 2\ell_{[\alpha}m_{\beta]} \,.\] From (\ref{riwe}) we can see that the energy--momentum content can correspond with vacuum, Einstein--Maxwell or pure radiation fields. Moreover, the Petrov type is N, being $\vec{\ell}$ the repeated principal direction of the Weyl tensor, which in fact is a constant vector field ($\nabla_\alpha\ell_\beta=0$). After some calculations, we have found that the Bel tensor $T^{\alpha\beta\lambda\mu}$ and our tensor $A^{\alpha\beta\lambda\mu}$ are \[ T^{\alpha\beta\lambda\mu}=4(\Phi^2+ \Psi_4\bar{\Psi}_4)\ell^\alpha\ell^\beta\ell^\lambda \ell^\mu \,, \] \[ \frac{1}{4}A^{\alpha\beta\lambda\mu}= (3\Phi^2+\Psi_4\bar{\Psi}_4)\ell^\alpha\ell^\beta\ell^\lambda \ell^\mu+ \ell^\alpha\ell^{(\beta}\nabla^\lambda\nabla^{\mu)}\Phi- \ell^{(\beta}\ell^\lambda\nabla^{\mu)}\nabla^\alpha\Phi \] \[\hspace{1cm} +\left[g^{\alpha(\beta}\ell^\lambda\ell^{\mu)}- \ell^\alpha\ell^{(\beta}g^{\lambda\mu)}\right]\nabla^\sigma \nabla_\sigma\Phi \,.\] From this expression, the completely timelike component is given by \[ A(\vec{u}) = 4(3\Phi^2+\Psi_4\bar{\Psi}_4) (\ell^\alpha u_\alpha)^4 \,. \] Then, the vanishing of $A(\vec{u})$ implies the Minkowski space--time. \vspace{4mm} Finally, we are going to consider the Vaidya radiating space--time (see for instance \cite{KSHM}). For the sake of brevity we only give here the completely timelike component (\ref{comt}) of the tensor $A^{\alpha\beta\lambda\mu}$. The line element of this spherically symmetric metric can be written as follows \[ ds^2=-2 F^2(u,v) du dv + r^2(u,v)(d\theta^2+\sin^2\theta d\varphi^2) \,,\] where \[ F^2(u,v)=f(u)\frac{\partial r}{\partial v} \,, \hspace{8mm} \frac{\partial r}{\partial u}=\frac{1}{2} f(u)\left(\frac{2m(u)}{r}-1\right) \,. \] Here, $m(u)$ is the invariantly defined mass function. As is well--known, the Petrov type of this metric is D. Then, taking the following Newman--Penrose adapted basis \[ \vec{\ell}=\frac{-1}{F}\frac{\partial}{\partial v}\,, \hspace{5mm} \vec{k}=\frac{-1}{F}\frac{\partial}{\partial u}\,, \hspace{5mm} \vec{m}=\frac{1}{\sqrt{2}r}\left(\frac{\partial} {\partial \theta}+\frac{i}{\sin\theta}\frac{\partial} {\partial \varphi}\right) \,, \] where $\vec{\ell}$ and $\vec{k}$ are aligned with the principal directions of the Weyl tensor, the Ricci tensor is given by \[ R_{\alpha\beta}=2\Phi\ell_\alpha\ell_\beta\,, \hspace{8mm} \Phi\equiv -\frac{m_{,u}}{r^2 r_{,u}} \,.\] And the only non--zero component of the Weyl tensor in this basis (see \cite{KSHM}) is \[ \Psi_2 = -\frac{m(u)}{r^3(u,v)} \,, \] which in this case is real. In terms of these quantities, we have found the following expression for $A(\vec{u})$ \[ A(\vec{u}) = 2\left[\Psi_2+\Phi(\ell^\alpha u_\alpha)^2 \right]^2 + \] \[ +\,4\Psi^2_2\left[36(\ell^\alpha u_\alpha)^2 (k^\beta u_\beta)^2-18(\ell^\alpha u_\alpha)(k^\beta u_\beta) +1\right]+10\Phi^2(\ell^\alpha u_\alpha)^4 \,,\] and again, taking into account that \[2(\ell^\alpha u_\alpha)(k^\beta u_\beta)\geq 1 \hspace{5mm} \Longrightarrow\hspace{5mm} 36(\ell^\alpha u_\alpha)^2 (k^\beta u_\beta)^2-18(\ell^\alpha u_\alpha)(k^\beta u_\beta) +1\geq 1\,, \] $A(\vec{u})$ vanishes if and only if the space--time is the Minkowski space--time. When we restrict ourselves to observers lying on the 2--planes generated by the principal directions [$2(\ell^\alpha u_\alpha)(k^\beta u_\beta)=1$], which are precisely the observers that minimize $A(\vec{u})$, the result is \[ A(\vec{u}) = 2\left[\Psi_2+\Phi(\ell^\alpha u_\alpha)^2 \right]^2+ 4\Psi^2_2+10\Phi^2(\ell^\alpha u_\alpha)^4 \,.\] \section*{Acknowledgments} M.\'A.G.B. gratefully acknowledges the {\it Comissionat per a Universitats i Recerca de la Generalitat de Catalunya} for financial support. C.F.S. gratefully acknowledges financial support in the form of a fellowship from the Alexander von Humboldt Foundation.
\section{Introduction} Arkani-Hamed, Dimopoulos and Dvali(ADD)~{\cite {nima}} have recently proposed a uniquely interesting solution to the hierarchy problem. ADD hypothesize the existence of $n$ additional large spatial dimensions in which gravity (and perhaps Standard Model singlet fields) can live, called `the bulk', whereas all of the fields of the Standard Model(SM) are constrained to lie on `the wall', which is our conventional 4-dimensional world. Gravity thus appears to us as weak in ordinary 4-dimensional space-time since we merely observe it's action on the wall. It has been shown~{\cite {nima}} that such a scenario can emerge in string models where the effective Planck scale in the bulk is identified with the string scale. In such a theory the hierarchy is removed by postulating that the string or effective Planck scale in the bulk, $M_s$, is not far above the weak scale, {\it e.g.}, a few TeV. Gauss' Law then provides a link between the values of $M_s$, the conventional Planck scale $M_{pl}$, and the size of the compactified extra dimensions, $R$, \begin{equation} M_{pl}^2 \sim R^nM_s^{n+2}\,, \end{equation} where the constant of proportionality depends not only on the value of $n$ but upon the geometry of the compactified dimensions. If $M_s$ is near a TeV then $R\sim 10^{30/n-19}$ meters; for separations between two masses less than $R$ the gravitational force law becomes $1/r^{2+n}$. For $n=1$, $R\sim 10^{11}$ meters and is thus obviously excluded, but, for $n=2$ one obtains $R \sim 1$~mm, which is at the edge of the sensitivity for existing experiments{\cite {test}}. For $2<n$, the value of $R$ is only further reduced and thus we may conclude that the range $2\leq n$ is of phenomenological interest. Astrophysical arguments based on supernova cooling{\cite {astro}} seem to require that $M_s>50$ TeV for $n=2$, but allow $M_s\sim 1$ TeV for $n>2$ while cosmological arguments suggest{\cite {lhall} an even stronger constraint, $M_s>110$ TeV, for $n=2$. The Feynman rules for this scenario are obtained by considering a linearized theory of gravity in the bulk, decomposing it into the more familiar 4-dimensional states and recalling the existence of Kaluza-Klein towers for each of the conventionally massless fields. The entire set of fields in the K-K tower couples in an identical fashion to the particles of the SM. By considering the forms of the $4+n$ symmetric conserved stress-energy tensor for the various SM fields and by remembering that such fields live only on the wall one may derive all of the necessary couplings. An important result of these considerations is that only the massive spin-2 K-K towers (which couple to the 4-dimensional stress-energy tensor, $T^{\mu\nu}$) and spin-0 K-K towers (which couple proportional to the trace of $T^{\mu\nu}$) are of phenomenological relevance as all the spin-1 fields can be shown to decouple from the particles of the SM. For processes that involve massless fields at at least one vertex, as will be the case below, the contributions of the spin-0 fields can also be ignored. The detailed phenomenology of the ADD model has begun to be explored for a wide ranging set of processes in a rapidly growing series of recent papers~{\cite {pheno}}. Given the Feynman rules as developed by Guidice, Rattazzi and Wells and by Han, Lykken and Zhang{\cite {pheno}}, it appears that the ADD scenario has two basic classes of collider tests. In the first class, type-$i$, a K-K tower of gravitons can be emitted during a decay or scattering process leading to a final state with missing energy. The rate for such processes is strongly dependent on the number of extra dimensions as well as the exact value of $M_s$. However, in this case the value of $M_s$ is directly probed. In the second class, type-$ii$, which we consider here, the exchange of a K-K graviton tower between SM or MSSM fields can lead to almost $n$-independent modifications to conventional cross sections and distributions or they can possibly lead to new interactions. The exchange of the graviton K-K tower leads to a set of effective color and flavor singlet contact interaction operator of dimension-eight with the overall scale set by the cut-off in the tower summation, $\Lambda$. Naively $\Lambda$ and $M_s$ should be of comparable magnitude and so one introduces a universal overall order one coefficient for these operators, $\lambda$ ( whose value is unknown but can be approximated by a constant which has conventionally been set to $\pm 1$) with $\Lambda$ being replaced by $M_s$. The fact that $\Lambda$ can be smaller than $M_s$ is thus particularly important when considering the case of 2 extra dimensions due to the rather strong astrophysical and cosmological constraints that then apply. We note that $\lambda$ can in principle have either sign since the unknown physics above the cut-off can make an additional universal contribution to the coefficient of the relevant amplitude of indeterminate sign. Given the kinematic structure of these operators the modifications in the relevant cross sections and distributions can be directly calculated. In what follows we will consider the production of massive gauge boson pairs in $\gamma \gamma$ collisions via the exchange of a K-K tower of gravitons (the process $\gamma\gamma \rightarrow \gamma \gamma$ having been considered{\cite {pheno}} elsewhere). Such reactions can be examined in detail at future linear colliders via the Compton back-scattering of laser photons off of high energy colliding $e^+e^-$ or $e^-e^-$ beams{\cite {telnov}}. As we note below, the role of polarization, in both the initial state as well as the final state, is crucial in separating the graviton signal from the SM background and extending the search reach. We then will demonstrate that the discovery reach for graviton tower exchange in the $WW$ channel is greater than any other process so far examined. The corresponding reach in the $ZZ$ case will be shown to be rather modest and comparable to that found for the $\gamma \gamma \rightarrow \gamma \gamma$ process. We also note the added bonus feature associated with $\gamma\gamma$ collisions. Whereas, {\it e.g.} , $e^+e^-\rightarrow f\bar f$ can receive contributions from $Z$ and $\gamma$ Kaluza-Klein towers in some models, the processes $\gamma \gamma \rightarrow X\bar X$ do not at tree level and are thus can provide very clean signatures for graviton exchange. \section{$\gamma \gamma \rightarrow VV$ via Graviton Exchange} $\gamma \gamma$ collisions offer a unique and distinct window on the possibility of new physics in a particularly clean environment. At tree level the cross section for particle pair production, if it is allowed by the gauge symmetries, depends only gauge couplings. Unlike gauge boson pair production in $e^+e^-$ collisions, however, $P$, $C$, plus the Bose symmetry of the initial state photons forbids the existence of non-zero values, at the tree level, for either forward-backward angular asymmetries or left-right forward-backward polarization asymmetries. These were both found to be powerful tools in probing for K-K graviton tower exchanges in the $e^+e^-$ initiated channels{\cite {pheno}}. In the case of $\gamma\gamma$ collisions our remaining tools are the angular distributions of the produced pairs of vector bosons, their resulting states of polarization, and the cross section's sensitivity to the polarization of the initial state photons. The exchange of a tower of K-K gravitons leads to a new tree level contribution to the process $\gamma(k_1,\epsilon_1)\gamma(k_2,\epsilon_2)\rightarrow V(p_1,V_1)V(p_2,V_2)$, where $k_{1,2}$ are incoming and $p_{1,2}$ are outgoing momenta and $\epsilon_{1,2}(V_{1,2})$ represent the polarization vectors for the photons(vector bosons), respectively. Following the Feynman rules of either Guidice, Rattazzi and Wells or Han, Lykken and Zhang{\cite {pheno}} we can immediately write down the relevant matrix element which has the form of a dimension-8 operator: \begin{equation} {\cal M}_{grav}={8\lambda K\over M_s^4}\left[M_1+M_2+M_3+M_4+M_5\right]\,, \end{equation} where \begin{eqnarray} M_1 &=& {s^2\over {2}}\bigg[\epsilon_1\cdot V_1^* \epsilon_2\cdot V_2^*+ \epsilon_1\cdot V_2^* \epsilon_2\cdot V_1^* \bigg]+s\epsilon_1 \cdot \epsilon_2\bigg[k_1\cdot V_1^* k_2\cdot V_2^*+k_1\cdot V_2^* k_2\cdot V_1^*\nonumber \\ &-& k_1\cdot k_2 V_1^* \cdot V_2^* \bigg] \nonumber \\ M_2 &=& sV_1^* \cdot V_2^* \bigg[p_1\cdot \epsilon_1 p_2\cdot \epsilon_2+ p_1\cdot \epsilon_2 p_2\cdot \epsilon_1 \bigg]+ 2\epsilon_1 \cdot \epsilon_2 V_1^* \cdot V_2^* \bigg[p_1 \cdot k_1 p_2 \cdot k_2+p_1 \cdot k_2p_2 \cdot k_1\nonumber \\ &-& p_1 \cdot p_2 k_1 \cdot k_2\bigg]\nonumber \\ M_3 &=& -sp_1\cdot V_2^*\bigg[\epsilon_1 \cdot V_1^* \epsilon_2 \cdot p_2 +\epsilon_2 \cdot V_1^* \epsilon_1 \cdot p_2\bigg]-sp_2\cdot V_1^* \bigg[\epsilon_1 \cdot V_2^* \epsilon_2 \cdot p_1+ \epsilon_2 \cdot V_2^* \epsilon_1 \cdot p_1\bigg]\nonumber \\ M_4 &=& -2p_1\cdot V_2^* \epsilon_1 \cdot \epsilon_2 \bigg[k_1\cdot V_1^* k_2\cdot p_2+k_2\cdot V_1^* k_1\cdot p_2-p_2\cdot V_1^* k_1\cdot k_2\bigg]\\ M_5 &=& -2p_2\cdot V_2^* \epsilon_1 \cdot \epsilon_2 \bigg[k_1\cdot V_2^* k_2\cdot p_1+k_2\cdot V_2^* k_1\cdot p_1-p_1\cdot V_2^* k_1 \cdot k_2\bigg]\nonumber \,, \end{eqnarray} with $s$ being the usual Mandelstam variable. We note that in the expression above $M_s$ should actually be the cut-off scale used in performing the summation over the K-K tower in the $s$-channel. In principle, as discussed above, these two scales may differ by by some factor of order unity which we thus incorporate into the parameter $\lambda$. For $K=1(\pi/2)$ we recover the normalization convention employed by Hewett(Guidice, Rattazzi and Wells){\cite {pheno}}; we will take $K=1$ in the numerical analysis that follows but keep the factor in our analytical expressions. We recall from the original Hewett analysis that $\lambda$ is to be treated as a parameter of order unity whose sign is undetermined and that, given the scaling relationship between $\lambda$ and $M_s$, experiments in the case of processes of type-$ii$ actually probe only the combination $M_s/|K\lambda|^{1/4}$. For simplicity in what follows we will numerically set $|\lambda|=1$ and employ $K=1$ but we caution the reader about this technicality and quote our sensitivity to $M_s$ for $\lambda=\pm 1$ as is now the standard tradition. In the center of mass frame we imagine that the initial photons are coming in along the $z$-axis with the outgoing vector bosons making an angle $\theta$ with respect to this axis. Using $\beta^2=1-4m^2/s$, $m$ being the mass of the field $V$, and recalling that the transverse initial state photons are assumed to be circularly polarized, we can describe the complete kinematics for the process in terms of the following four-vectors: \begin{equation} \begin{array}{rcl} k_{1\mu} &=& {{\sqrt s}\over {2}}(1,0,0,1)\\ p_{1\mu} &=& {{\sqrt s}\over {2}}(1,\beta s_\theta,0,\beta c_\theta)\\ \epsilon_1^\mu &=& -{1\over {\sqrt 2}}(0,\lambda_1,i,0)\\ V_{1L}^{\mu *} &=& {{\sqrt s}\over {2m}}(-\beta,s_\theta,0,c_\theta)\\ V_{1T}^{\mu *} &=& {1\over {\sqrt 2}}(0,-\lambda_1^Vc_\theta,i,\lambda_1^V s_\theta) \end{array}\qquad \begin{array}{rcl} k_{2\mu} &=& {{\sqrt s}\over {2}}(1,0,0,-1)\\ p_{2\mu} &=& {{\sqrt s}\over {2}}(1,-\beta s_\theta,0,-\beta c_\theta)\\ \epsilon_2^\mu &=& {1\over {\sqrt 2}}(0,-\lambda_2,i,0)\\ V_{2L}^{\mu *} &=& -{{\sqrt s}\over {2m}}(\beta,s_\theta,0,c_\theta)\\ V_{2T}^{\mu *} &=& {1\over {\sqrt 2}}(0,-\lambda_2^Vc_\theta,-i,\lambda_2^V s_\theta) \end{array} \end{equation} with $s_\theta=\sin \theta$ and $z=c_\theta=\cos \theta$. The use of covariant and contravariant indices in these expressions should be noted by the reader. Here the indices $T,L$ on the polarization vectors $V_{1,2}$ denote states of transverse and longitudinal polarization. The quantities $\lambda_{1,2}$ and $\lambda_{1,2}^V$ parameterize the two transverse polarization states of the photons and vector bosons, respectively, and take on the values $\pm 1$. Given these kinematical expressions we can immediately evaluate all of the dot products appearing in the matrix element for any given choice of photon and/or $V$ polarizations. The resulting differential cross section for any particularly chosen set of helicities can then be written as \begin{equation} {d\sigma \over {dz}}={1\over {1+\delta_{VZ}}}{\beta \over {32\pi s}} |{\cal M}_{SM}+{\cal M}_{grav}|^2\,, \end{equation} where ${\cal M}_{SM}$ symbolically represents the more conventional contribution to these matrix elements arising in the SM or the MSSM, {\it etc}. Note the Kronecker delta factor in the denominator in the case of the identical particle $ZZ$ final state. Unfortunately $\gamma\gamma$ collisions at future linear colliders will not be as straightforward as the description above would indicate since the photons in the collision will not be either monoenergetic or in a unique state of polarization. Polarized $\gamma\gamma$ collisions may be possible at future $e^+e^-$ colliders through the use of Compton backscattering of polarized low energy laser beams off of polarized high energy electrons{\cite {telnov}}. The resulting backscattered photon distribution, $f_\gamma(x=E_\gamma/E_e)$, is as stated above far from monoenergetic and is cut off above $x_{max}\simeq 0.83$ implying that the colliding photons are significantly softer than the parent lepton beam energy. As one sees, this cutoff at large $x$, $x_{max}$, implies that the $\gamma \gamma$ center of mass energy never exceeds $\simeq 0.83$ of the parent collider and this has resulted in a significantly degraded $M_s$ sensitivity for final states involving fermion or scalar pairs{\cite {pheno}}. In addition, both the shape of the function $f_\gamma$ and the average helicity of the produced $\gamma$'s are quite sensitive to the polarization state of both the initial laser ($P_l$) and electron ($P_e$) whose values fix the specific distribution. While it is anticipated that the initial laser polarization will be near $100\%$, {\it i.e.}, $|P_l|=1$, the electron beam polarization is expected to be be near $90\%$, {\it i.e.}, $|P_e|=0.9$. We will assume these values in the analysis that follows. With two photon `beams' and the choices $P_l=\pm 1$ and $P_e=\pm 0.9$ to be made for each beam it would appear that 16 distinct polarization-dependent cross sections need to be examined. However, due to the exchange symmetry between the two photons and the fact that a simultaneous flip in the signs of all the polarizations leaves the product of the fluxes, mean helicities and the cross sections invariant, we find that there only six physically distinct initial polarization combinations{\cite {chak}}. In what follows we will label these possibilities by the corresponding signs of the electron and laser polarizations as $(P_{e1},P_{l1},P_{e2},P_{l2})$, For example, the configuration $(-++-)$ corresponds to $P_{e1}=-0.9$, $P_{l1}=+1$, $P_{e2}=0.9$ and $P_{l2}=-1$. Clearly some of these polarization combinations will be more sensitive to the effects of K-K towers of gravitons than will others so our analysis can be used to pick out those particular cases. Within this framework we must view the above differential cross section as that of a partonic subprocess in analogy with hadronic collisions, {\it i.e.}, $d \sigma \rightarrow d \hat \sigma$ and we identify $s \rightarrow \hat s=s_{e^+e^-}x_1x_2$. For any given choice of the initial state laser and electron polarizations, labelled by $(a,b)$ below, we can immediately write down the appropriate cross section by folding in the corresponding photon fluxes and integrating: \begin{equation} {d\sigma^{ab}\over {dz}}=\int ~dx_1~\int~dx_2~f^a_\gamma(x_1,\xi_1) f^b_\gamma(x_2,\xi_2)\Bigg[{{1+\xi_1\xi_2}\over {2}} {d\hat \sigma_{++} \over {dz}}+{{1-\xi_1\xi_2}\over {2}} {d\hat \sigma_{+-} \over {dz}}\Bigg]\,, \end{equation} where we explicitly note the dependence of the fluxes on the mean helicities $\xi_{1,2}$ which are known functions of energy as well as both the initial state laser and electron beam polarization. The $++$ and $+-$ labels on the subprocess cross sections indicate the appropriate values of $\lambda_{1,2}$ to chose in their evaluation. In order to obtain ${d\sigma \over {dz}}$ the polarizations of the vector bosons in the final state must be either specified or summed over. Similarly, we can obtain the unpolarized cross section by averaging over the initial state photon polarizations. Given the fluxes{\cite {telnov}} these integrals are easily evaluated numerically. The upper limit of both integrals is just $x_{max}$. In the present case the kinematics require the photon energies to satisfy the constraint $\tau=\hat s/s=x_1x_2\geq 4m^2/s=\tau_{min}$ which, together with the value of $x_{max}$, then determines the lower bounds on both $x_{1,2}$: $x_1^{min}=\tau_{min}/x_{max}$ and $x_2^{min}=\tau_{min}/x_1$. \begin{figure}[htbp] \centerline{ \psfig{figure=gamgravv.res2.ps,height=8.9cm,width=9.1cm,angle=90} \hspace*{5mm} \psfig{figure=gamgravv.res3.ps,height=8.9cm,width=9.1cm,angle=90}} \vspace*{0.15cm} \centerline{ \psfig{figure=gamgravv.res4.ps,height=8.9cm,width=9.1cm,angle=90}} \vspace*{0.05cm} \caption{Differential cross section for $\gamma\gamma \rightarrow W^+W^-$ at a 1 TeV $e^+e^-$ collider for (a)the SM and with $M_s=2.5$ TeV with (b)$\lambda=1$ or (c)$\lambda=-1$. In (a) from top to bottom in the center of the figure the helicities are $(++++)$, $(+++-)$, $(-++-)$, $(++--)$, $(+---)$, and $(+-+-)$; in (b) they are $(-++-)$, $(+-+-)$, $(+++-)$, $(+---)$, $(++++)$, and $(++--)$; in (c) they are $(-++-)$, $(+-+-)$, $(+---)$, $(+++-)$, $(++++)$, and $(++--)$.} \label{fig1} \end{figure} \section{Results} Now that we have specified all of the relevant machinery we can now turn to some results. We consider the two cases $V=W$ and $V=Z$ separately. \subsection{$\gamma \gamma \rightarrow W^+W^-$} The tree-level SM helicity amplitudes for the process $\gamma \gamma \rightarrow W^+W^-$ have been known for some time{\cite {wsm}} and even the complete one-loop corrections within the SM are also known{\cite {loop1}. In what follows we will ignore these loop corrections but remind the reader that they must be employed in a complete analysis that also includes detector effects {\it etc}. Given these amplitudes, we can immediately calculate the relevant differential cross section combining SM and graviton tower exchanges. To be definitive we will assume $\sqrt s=\sqrt s_{e^+e^-}$=1 TeV with $M_s$=2.5 TeV for purposes of demonstration. The results are shown in Fig.\ref{fig1} for the SM case as well as when the K-K tower is turned on with either sign of $\lambda$ for all six initial helicity combinations. In the SM case the shape of the angular distribution is easily understood by recalling that the $\gamma \gamma \rightarrow W^+W^-$ reaction takes place through $t$- and $u$-channel $W$ exchanges as well as a $\gamma \gamma W^+W^-$ four-point interaction. The $t$- and $u$-channel exchanges thus lead to a sharply rising cross section in both the forward and backward directions. Note that in the SM there is no dramatically strong sensitivity to the initial state lepton and laser polarizations in this case and all of the curves have roughly the same shape. In all cases the total cross section, even after generous angular cuts, is quite enormous, of order $~\sim 100$ pb, providing huge statistics to look for new physics influences. When the graviton terms are turned on there are several effects. First, all of differential cross section distributions become somewhat more shallow, particularly in the case of $\lambda=1$, but there is little change in the forward and backward directions due to the dominance of the SM poles. Second, there is now a clear and distinct sensitivity to the initial state polarization selections. In some cases, particularly for the $(-++-)$ and $(+-+-)$ helicity choices, the differential cross section increases significantly for angles near $90^o$ taking on an m-like shape. This shape is, in fact, symptomatic of the spin-2 nature of the K-K graviton tower exchange since a spin-0 exchange leads only to a flattened distribution. Given the very large statistics available with a typical integrated luminosity of 100 $fb^{-1}$, even after angular cuts are applied to remove the forward and backward SM poles, it is clear that the $\gamma \gamma \rightarrow W^+W^-$ differential cross section is quite sensitive to $M_s$ especially for the two initial state helicities specified above. \vspace*{-0.5cm} \noindent \begin{figure}[htbp] \centerline{ \psfig{figure=helicity.res1.ps,height=9.0cm,width=11.5cm,angle=90}} \vspace*{5mm} \centerline{ \psfig{figure=helicity.res2.ps,height=9.0cm,width=11.5cm,angle=90}} \vspace*{0.1cm} \caption{Fraction of LL(solid), TL+LT(dashed) and TT(dotted) $W^+W^-$ final states after angular cuts for the process $\gamma \gamma \rightarrow W^+W^-$ at a 1 TeV $e^+e^-$ collider as a function of $M_s$ for either sign of $\lambda$. The initial state polarization in (a) is $(-++-)$ whereas in (b) it is $(+-+-)$.} \label{fig2} \end{figure} \vspace*{0.4mm} In addition to a significant modification to the angular distribution, the K-K graviton tower exchange leads to another important effect through its influence on the polarizations of the two $W$'s in the final state. In the SM, independent of the initial electron and laser polarizations, the final state $W$'s are dominantly transversely polarized. Due to the nature of the spin-2 graviton exchange, the K-K tower leads to a final state where both $W$'s are completely longitudinally polarized. Thus we might expect that a measurement of the $W$ polarization will probe $M_s$. To see this, we show in Fig.\ref{fig2}} the polarization fractions of the two $W$'s as a function of $M_s$ at a 1 TeV collider assuming two different choices of the initial state polarizations. In the results presented in this figure, an angular cut of $|z|<0.8$ has been applied to remove the SM poles in the forward and backward directions. Here we see that the fraction of final states where both $W$'s are longitudinal, denoted by $LL$, starts out near unity but falls significantly in the $M_s=2.5-3$ TeV region giving essentially the SM results above about 5.5-6 TeV. The reverse situation is observed for the case where both $W$'s are transversely polarized, denoted by $TT$. The mixed case, denoted by the sum $TL+LT$, is also seen to grow from near zero to a modest value as $M_s$ increases and the SM limit is reached. Clearly a measurement of these final $W$ polarizations will allow us to probe respectable values of $M_s$. In order to determine the polarization fractions of the final state $W$'s one needs to examine the correlations in the angular distributions of the fermion decay products relative to the $W$'s original direction. If $\chi$ is the angle of one of the fermions with respect to the $W$ direction in the $W$ rest frame then transverse(longitudinal) $W$'s lead to an angular distribution $\sim 1 +(-) \cos^2 \chi$. Thus by measuring the correlation for both $W$'s the relevant $LL,TT$ and $LT+TL$ fractions can be extracted. In our numerical exercise we will assume that this can be done with an efficiency of $\sim 50\%$ for the $WW$ all hadronic final states ($W$ mass reconstruction removing combinatoric problems) as well as for the $q\bar q\ell \nu$ final state with no significant backgrounds present in either case due to the very large $\gamma \gamma \rightarrow W^+W^-$ cross section. \vspace*{-0.5cm} \noindent \begin{figure}[htbp] \centerline{ \psfig{figure=discr.res1.ps,height=9.0cm,width=11.5cm,angle=90}} \vspace*{5mm} \centerline{ \psfig{figure=discr.res2.ps,height=9.0cm,width=11.5cm,angle=90}} \vspace*{0.1cm} \caption{$M_s$ discovery reach from the process $\gamma \gamma \rightarrow W^+W^-$ at a 1 TeV $e^+e^-$ collider as a function of the integrated luminosity for the different initial state polarizations assuming (a)$\lambda=1$ or (b)$\lambda=-1$. From top to bottom on the right hand side of the figure the polarizations are $(-++-)$, $(+---)$, $(++--)$, $(+-+-)$, $(+---)$, and $(++++)$.} \label{fig3} \end{figure} \vspace*{0.4mm} By combining a fit to the total cross sections and angular distributions as well as the $LL$ and $LT+TL$ helicity fractions for various initial state polarization choices we are able to discern the discovery as well as the $95\%$ CL exclusion reaches for $M_s$. Given the rather steep behaviour of the both the SM-gravity and pure gravity terms in the cross section with $M_s$, {\it i.e.}, $M_s^{-4}$ and $M_s^{-8}$ respectively, we do not expect these two reaches to differ significantly. In performing this analysis, in addition to the above efficiencies and angular cuts, we have assumed an overall integrated luminosity uncertainty of $1\%$. The results of this analysis for $\lambda=\pm 1$ and the six possible initial state polarizations are displayed in Figs.\ref{fig3} and \ref{fig4} which show both reaches as functions of the total integrated luminosity. Note both the strong sensitivity of the two reaches to the initial electron and laser polarizations as well as the large values obtainable particularly for the $(-++-)$ choice. In this particular case with 100 $fb^{-1}$ of integrated luminosity the discovery reach is almost $11\sqrt s$ for either sign of $\lambda$, which is greater than any other K-K graviton exchange process so far examined{\cite {pheno}}. Clearly, a more detailed analysis of these reaches, including both radiative corrections and detector effects, is more than warranted. \vspace*{-0.5cm} \noindent \begin{figure}[htbp] \centerline{ \psfig{figure=helicity.res3.ps,height=9.0cm,width=11.5cm,angle=90}} \vspace*{5mm} \centerline{ \psfig{figure=helicity.res4.ps,height=9.0cm,width=11.5cm,angle=90}} \vspace*{0.1cm} \caption{$95\%$ CL exclusion reach for $M_s$ at a 1 TeV $e^+e^-$ collider for the same cases as shown in the previous figure.} \label{fig4} \end{figure} \vspace*{0.4mm} There are of course many more observables available that are sensitive to the presence of the K-K graviton tower exchange and which can be included in a more detailed global fit; here we only mention two of them. First, we can construct the invariant mass distribution of the two $W$'s in the final state approximately $90\%$ of the time and thus determine the differential cross section $d\sigma/ dM_{WW}$. After imposing the $|z|<0.8$ cut we would expect that the SM result will gradually fall off with increasing $M_{WW}$. However, since the K-K graviton tower exchange contribution grows quite rapidly with increasing $\hat s$, a modest value of $M_s$ will leads to an observable event excess at large values of $M_{WW}$. This is in fact exactly what we find as shown in Fig.\ref{fig5} for the sample choice of the initial state polarization $(-++-)$. Clearly a fit to this just distribution alone with the large statistics available will provide an additional probe of $M_s$ in the range beyond 4 TeV. \vspace*{-0.5cm} \noindent \begin{figure}[htbp] \centerline{ \psfig{figure=massdis.res.ps,height=13cm,width=16cm,angle=90}} \vspace*{0.1cm} \caption[*]{$WW$ mass distribution for $\gamma \gamma \rightarrow W^+W^-$ at a 1 TeV $e^+e^-$ collider for the initial polarization of $(-++-)$ in the SM(red), the case of graviton exchange with $M_s$=2.5 TeV for both values of $\lambda$(green and blue, respectively) and for the corresponding scenario with $M_s$=3.5 TeV(magenta and cyan, respectively). In all cases a cut of $|z|<0.8$ has been imposed.} \label{fig5} \end{figure} \vspace*{0.4mm} Second, for the six possible initial state polarizations one can construct three independent polarization asymmetries of the form \begin{equation} A_{pol}={{\sigma(P_{e1},P_{l1},P_{e2},P_{l2})-\sigma(P_{e1},P_{l1},-P_{e2}, -P_{l2})}\over {\sigma(P_{e1},P_{l1},P_{e2},P_{l2})+\sigma(P_{e1},P_{l1}, -P_{e2},-P_{l2})}}\,. \end{equation} These asymmetries can also be made angular dependent, $A_{pol}(z)$, by interpreting the cross sections in both the numerator and denominator as differential in $z$ and taking advantage of the large statistics available. Further one may separately examine asymmetries constructed from distinct final state polarization states. Such a general class of asymmetries involving differences in initial state photon helicities form the basic ingredients of the Drell-Hearn-Gerasimov Sum Rule(DHG){\cite {dhg}} which has be applied, {\it e.g.} , in the search for anomalous couplings between photons and $W,Z$'s{\cite {old}}. In order to prove the usefulness of these asymmetries let us first examine their angular dependencies in both the SM and when K-K towers of gravitons{\cite {gold}} are exchanged; this is shown in Fig.\ref{fig6} for a 1 TeV collider. Note that as $z\rightarrow \pm 1$ the SM dominates due to the large magnitude of the $u$- and $t$-channel poles. Away from the poles the three asymmetries all show a significant sensitivity to the K-K tower of graviton exchange. Although these asymmetries are not very big the large statistics of the data samples obtainable for this channel indicate that they will be very well determined since many systematic errors will also cancel in forming the cross section ratios. \vspace*{-0.5cm} \noindent \begin{figure}[htbp] \centerline{ \psfig{figure=dhg.res.ps,height=13cm,width=16cm,angle=90}} \vspace*{0.1cm} \caption[*]{Differential polarization asymmetries for $\gamma \gamma \rightarrow W^+W^-$ at a 1 TeV $e^+e^-$ collider for the SM(solid) as well with graviton tower exchange with $M_s$=2.5 TeV with $\lambda=\pm 1$(the dotted and dashed curves). We label the three cases shown by the first entry in the numerator in the definition of $A_{pol}$. Red represents an initial polarization of $(++++)$, green is for the choice $(+++-)$ and blue is for the case $(-++-)$.} \label{fig6} \end{figure} \vspace*{0.4mm} As is well known{\cite {old}} in order to satisfy the DHG sum rule the {\it integrated} asymmetry must pass through a zero (by the mean value theorem) for some value of $\hat s$. The exact position of the zero has been shown to be sensitive to new physics and to any applied kinematic cuts. What happens in the case where the SM is augmented by an exchange of a K-K tower of gravitons and how does $A_{pol}$ vary with $M_{WW}$? Fig.\ref{fig7} address these question directly for all three independent asymmetries after employing the $|z|<0.8$ cut. The position of the zero is seen to be the same in all three cases and quite insensitive to even low values of $M_s \sim 2.5$ TeV. This is due to the rather unfortunate fact that the zero occurs in the low $M_{WW}$ region where contributions from finite $M_s$ are very difficult to observe. The behaviour of $A_{pol}$ at larger values of $M_{WW}$ {\it is}, however, very sensitive to graviton exchange indicating a strong modification of the Sum Rule integral itself. Since the K-K graviton tower exchange is represented by an effective dimension-eight operator it can be shown for all initial state helicities that the integral in the Sum Rule diverges. This should be expected since the new operator is non-renormalizable. \vspace*{-0.5cm} \noindent \begin{figure}[htbp] \centerline{ \psfig{figure=dhgzero.res.ps,height=13cm,width=16cm,angle=90}} \vspace*{0.1cm} \caption[*]{Integrated polarization asymmetries for $\gamma \gamma \rightarrow W^+W^-$ at a 1 TeV $e^+e^-$ collider as functions of the $WW$ invariant mass. The labels for the various curves are as in the previous figure and a cut of $|z|<0.8$ has been applied.} \label{fig7} \end{figure} \vspace*{0.4mm} It is clear from the discussion above that there are a large number of observables that can be combined into a global fit to probe very high values of $M_s$ in comparison to the collider energy. It should be noted however that due to the large statistics available the eventually determined discovery reach for $M_s$ using the $\gamma \gamma \rightarrow W^+W^-$ process will strongly depend on the size and variety of the experimental systematic errors. \subsection{$\gamma \gamma \rightarrow ZZ$} The process $\gamma \gamma \rightarrow ZZ$ does not occur at the tree level in the SM or MSSM. At the one loop level in the SM the dominant contribution{\cite {loop2}} arises from $W$ and fermion box diagrams and triangle graphs with $s$-channel Higgs boson exchange. (In SUSY models, additional contributions arise due to sfermion and gaugino loops as well as the other Higgs exchanges.) The $W$ bosons loops have been shown to be the dominant contribution with the fermions interfering destructively. This would seem to naively imply that this channel is particularly suitable for looking for new physics effects since the SM and MSSM rates will be so small due to the loop suppression. The SM cross section (which peaks in the forward and backward directions), after a cut of $|z|<0.8$, is found to be $\sim 80$ fb and almost purely transverse away from Higgs boson resonance peaks. The size of this cross section being only $\sim 10^{-3}$ of that for $W^+W^-$ will make the $ZZ$ final state difficult to find. Since it is not anticipated that the $W$ and $Z$ masses will be very well separated in the jet-jet channel at this level of rejection, we must demand that at least one of the $Z$'s decay leptonically reducing the effective number of useful $Z$ by a factor $\simeq 10$. (We can regain some reasonable fraction of this suppression depending on the practicality of the $q\bar q \nu \bar \nu$ final state. While this state is useful for constructing angular distributions assuming a very hermetic detector it cannot be used, {\it e.g.} , to probe the polarization fractions of both of the $Z$'s. The addition of this final state for other analyses would yield an efficiency of $\sim 50\%$ instead of $\sim 10\%$.) A luminosity of 100 $fb^{-1}$ thus yields a sample of only about 800(4000) $Z$ pairs after these simple cuts and efficiencies are employed(assuming the $q\bar q \nu \bar \nu$ final states are included). \vspace*{-0.5cm} \noindent \begin{figure}[htbp] \centerline{ \psfig{figure=gamgravv.res1.ps,height=13cm,width=16cm,angle=90}} \vspace*{0.1cm} \caption[*]{Differential cross section for $\gamma \gamma \rightarrow ZZ$ at a 1 TeV $e^+e^-$ collider due to the exchange of a K-K tower of gravitons assuming $M_s=3$ TeV. From top to bottom in the center of the figure the initial state helicities are $(-++-)$, $(+-+-)$, $(+---)$, $(+++-)$, $(++--)$, $(++++)$.} \label{fig8} \end{figure} \vspace*{0.4mm} In the case of the ADD scenario the tree level K-K graviton tower contribution is now also present and is given by Eq.(2). Neglecting the loop-order SM contributions for the moment we obtain the resulting polarization-dependent differential cross sections shown in Fig.~\ref{fig8} where we have assumed $\sqrt s=1$ TeV and $M_s=3$ TeV for purposes of demonstration. Note that since this is the pure K-K graviton tower term there is no dependence here on the sign of $\lambda$. This cross section is found to scale with $s$ and $M_s$ as $\sim s^3/M_s^8$ and in contrast to the SM case is observed to peak at $90^o$. A short analysis along the lines of that performed above shows that essentially all of the $Z$'s in the final state are completely longitudinal with the $LL$ fraction being $\sim 99\%$ for the six possible initial state polarizations. Applying the same cuts, efficiencies and luminosities as above we find that for $M_s=4$ TeV this pure graviton contribution will only lead to 11(55) additional events beyond SM expectations which is not a huge excess. However, since the graviton tower cross section rises as $\hat s^3$ while the luminosities are falling off slowly, making a cut on the $ZZ$ invariant mass, $M_{ZZ}>550$ GeV, one finds that essentially all of the K-K-induced events lie above this cut (for the assumed value of the integrated luminosity) and would appear similar to a resonance-almost a broad Higgs-like{\cite {loop2}} bump. However, it would be doubtful that such an excess would be observed if $M_s>5$ TeV due to the $\sim M_s^{-8}$ scaling behaviour of the cross section unless significantly higher luminosities were available and systematic errors were very much under control. In addition, the use of the longitudinal polarization fraction would not seem to gain us any further reach. Although a detailed study of the loop-induced SM-graviton tower exchange interference terms have not yet been performed it is difficult to see how the search reach in this channel can exceed $\simeq 5$ TeV given the small magnitudes of the cross sections involved. Thus the anticipated overall reach for the $\gamma \gamma \rightarrow ZZ$ process is reasonably similar to that found for $\gamma \gamma \rightarrow \gamma \gamma$ {\cite {pheno}}. \section{Summary and Conclusions} $\gamma \gamma \rightarrow W^+W^-,ZZ$ offer new channels in which to search for the influence of graviton tower exchange, naively, each with their own individual strengths and weaknesses. As discussed above such channels are particularly clean for searches of the effects of graviton exchange since gauge boson towers cannot contribute to the cross sections at tree-level. We have found that \begin {itemize} \item The SM cross section for $\gamma \gamma \rightarrow W^+W^-$ is very large even after reasonable angular cuts are applied providing enormous statistics to look for new physics. This large cross section leads to an amplification of the size of the SM-graviton interference terms. The final state is quite clean there being little backgrounds due to rates alone and $90\%$ of the decay products are useable for analyses. The differential cross section as well as the polarization of the $W$'s in the final state were found to be quite sensitive to graviton exchanges especially for certain initial state electron and laser polarizations. The $W$-pair invariant mass distribution and the analogs of the Drell-Hearn-Gerasimov polarization asymmetries were also shown to be able to probe large values of $M_s$. Fitting the cross section and final state polarizations, after cuts, efficiencies and systematic errors were included was shown to lead to search reaches $\sim 11\sqrt s$, the largest of any of the graviton exchange processes so far examined. Detailed simulations of this channel should be performed which include the other observables, radiative corrections and detector effects to verify or improve upon this reach estimate. \item $\gamma \gamma \rightarrow ZZ$ would also appear to be an excellent process to probe for large values of $M_s$ since it only occurs at the loop level in the SM and MSSM. However both the SM and K-K tower cross sections are quite small for $M_s>4$ TeV even if the most advantageous initial state polarization choice, $(-++-)$, is assumed. Some assistance is gained by the fact that almost all of the $ZZ$ events will lie above a cut of $M_{ZZ}>550-600$ GeV and that they are almost purely longitudinally polarized. It seems unlikely, however, that the search reach in this channel can much exceed 5 TeV unless very large data samples become available. \end{itemize} The physics on large extra dimensions is quite exciting and may reveal itself at collider experiments in the not too far distant future. \vskip1.0in \noindent{\Large\bf Acknowledgements} The author would like to thank J.L. Hewett, N. Arkani-Hamed, J. Wells, T. Han, J. Lykken, M. Schmaltz and H. Davoudiasl for multi-dimensional discussion related to this work. He would also like to thank M. Berger, D. Dicus and G.J. Gounaris for discussions related to the SM helicity amplitudes for $\gamma \gamma \rightarrow ZZ$. \newpage \def\MPL #1 #2 #3 {Mod. Phys. Lett. {\bf#1},\ #2 (#3)} \def\NPB #1 #2 #3 {Nucl. Phys. {\bf#1},\ #2 (#3)} \def\PLB #1 #2 #3 {Phys. Lett. {\bf#1},\ #2 (#3)} \def\PR #1 #2 #3 {Phys. Rep. {\bf#1},\ #2 (#3)} \def\PRD #1 #2 #3 {Phys. Rev. {\bf#1},\ #2 (#3)} \def\PRL #1 #2 #3 {Phys. Rev. Lett. {\bf#1},\ #2 (#3)} \def\RMP #1 #2 #3 {Rev. Mod. Phys. {\bf#1},\ #2 (#3)} \def\NIM #1 #2 #3 {Nuc. Inst. Meth. {\bf#1},\ #2 (#3)} \def\ZPC #1 #2 #3 {Z. Phys. {\bf#1},\ #2 (#3)} \def\EJPC #1 #2 #3 {E. Phys. J. {\bf#1},\ #2 (#3)} \def\IJMP #1 #2 #3 {Int. J. Mod. Phys. {\bf#1},\ #2 (#3)}
\section{INTRODUCTION} There have been extensively studied for years about kaon condensation and its implications on neutron stars at low temperature \cite{lee}. In 1994 Brown and Bethe proposed the low-mass black hole (BH) scenario, based on the large softening of EOS due to kaon condensation \cite{bro}. It is produced as a consequence of the {\it delayed collapse} from a protoneutron star, different from the usual BH formation. This scenario should be very attractive in the light of recent observations on the mass of neutron stars, SN1987A or future observation of neutrinos associated with supernova explosions. Some numerical simulations based on the general relativity have been already performed for the delayed collapse \cite{bau,pon}. In ref.\cite{bau} they studied the delayed collapse by using the EOS of kaon condensed phase at $T=0$, though temperature is very much increased there. Also neutrino trapping is another important factor to be considered for protoneutron stars. There have been some attempts to treat thermal fluctuations in relation to kaon condensation \cite{pra}, but there seems to be no successful theory on the basis of chiral symmetry. Recently we have proposed a formalism to treat this problem \cite{tat}, in relation to protoneutron stars. Here we briefly explain how our formalism gives the thermodynamic potential, and show some results about EOS and structure of protoneutron stars. \section{THERMODYNAMIC POTENTIAL} The kaon condensed state can be decribed as a chiral-rotated one from the meson vacuum \cite{tatb}; actually we can discuss kaon condensation in almost model-independent way within the mean-field approximation. However, if we intend to study the phenomenon further by taking into account the effect of fluctuations, it is useful to invoke the effective Lagrangian like the nonlinear sigma model. \subsection{Path integral} We start with the partition function $Z_{chiral}$ for the nonlinear sigma model ${\cal L}_{chiral}$, \begin{equation} Z_{chiral}=N\int [dU][dB][d\bar B] \exp[S_{chiral}^{eff}], \label{aa} \end{equation} with the effective action, \begin{equation} S_{chiral}^{eff}=\int_0^\beta d\tau\int d^3x \left[{\cal L}_{chiral}(U, B)+\delta{\cal L}(U, B)\right], \end{equation} where $\delta{\cal L}(U, B)$ is the newly-appeared symmetry-breaking (SB) term due to the introduction of chemical potentials \cite{tat}. In evaluating the integral (\ref{aa}), we introduce the local coordinate around the condensate on the chiral manifold, which is equivalent with the following parametrization \cite{tat}, \begin{equation} U\equiv \xi^2=\zeta U_f\zeta(\xi=\zeta U_f^{1/2} u^\dagger=u U_f^{1/2}\zeta), \quad \zeta=\exp(\sqrt{2}i\langle M\rangle/f), \end{equation} where $\langle M\rangle$ is the condensate, $\langle M\rangle=V_+\langle K^+\rangle+V_-\langle K^-\rangle$, with $K^{\pm}=(\phi_4\pm i\phi_5)/\sqrt{2}$, $\theta^2\equiv 2K^+K^-/f^2$ and $V_\pm=F_4\pm iF_5$, while $U_f =\exp[2iF_a\phi_a/f]$ means the fluctuation field. Accordingly defining a new baryon field $B'$ by way of $ B'=u^\dagger B u, $ we can see that \begin{eqnarray} {\cal L}_{chiral}(U, B)={\cal L}_0(U, B)+{\cal L}_{SB}(U, B)&\longrightarrow& {\cal L}_0(U_f, B')+{\cal L}_{SB}(\zeta U_f\zeta,u B' u^\dagger)\nonumber\\ \delta{\cal L}(U, B)&\longrightarrow&\delta{\cal L}(\zeta U_f\zeta,u B' u^\dagger) \end{eqnarray} Thus all the dynamics of kaons and baryons in the condensed phase are completely prescribed by the non-invariant terms ${\cal L}_{SB}, \delta{\cal L}$ under chiral transformation; it is to be noted that the meson mass is included in ${\cal L}_{SB}$ and the Tomozawa-Weinberg term is in $\delta{\cal L}$. \subsection{Dispersion relation} The effective action for the kaon-nucleon sector can be represented as \begin{equation} S^{eff}_{chiral}=S_c+S_K+S_N+S_{int}, \end{equation} where $S_c$ is the previous classical-kaon action \cite{lee} and $S_N$ the nucleon action discarded in HBL. The sum $S_K+S_{int}$ gives the effective kaon action, \begin{equation} S_K^{eff}=-\frac{1}{2}\sum_{n, {\bf p}}(\phi_4(-n,-{\bf p}), \phi_5(-n,-{\bf p})) D^{eff}\left( \begin{array}{c} \phi_4(n,{\bf p})\\ \phi_5(n,{\bf p}) \end{array} \right) +..., \end{equation} with the inverse thermal Green function $D^{eff}$. Looking for the zeros in $D^{eff}$, we find two solutions $E_\pm$; $E_-$ corresponds to the Goldstone mode and exhibits the Bogoliubov spectrum, \begin{equation} E_-^2\sim \frac{c_3}{2C^2}p^2+\frac{p^4}{4C^2}+... \end{equation} where $C$ means an effective mass for kaons and $c_3$ the product of the charge density and the $KK$ scattering length \cite{tat}. We shall see the importance of the thermal kaon loops due to the Goldstone mode. The origin of this Goldstone mode is easily understood by observing that the kaon-condensed state is no longer invariant with respect to the $V$-spin rotation, while the effective Lagrangian is still inavariant. In other words, we can schematically say that the newly-appeared SB term $\delta{\cal L}$ completely cancels the original SB term ${\cal L}_{SB}$, and gives rise to a {\it new} spontaneous symmetry breaking (SSB) instead. \section{RESULTS} First we show EOS thus obtained for the isothermal and isentropic cases (Fig.~1). \begin{figure}[htb] \epsfxsize=1.0\textwidth \epsffile{3figu.eps} \caption{Sketch of EOS and graphical construction for the equilibrium curve in the isothermal case (a). EOS for the isothermal case (b) and the isentropic case (c).} \label{fig:largenenough} \end{figure} Since it exhibits the first-order phase transition (FOPT), we prescribe, for simplicity, the Maxwell construction by connecting the equal-pressure densities for the normal $(N)$ and condensed $(K)$ states. This means that there is no coexistent phase for the isothermal matter. On the other hand the coexistent phase appears in the isentropic matter due to the variation of temperature density by density. The magnitude of FOPT is so large that we shall see the existence of the gravitationally unstable region in the branch of neutron stars (see Fig. 2(b)). The effect of thermal fluctuations should be self-evident there. For protoneutron stars, the isentropic situation should be more relevant. In Fig.2(a) we present the temperature profile inside for protoneutron-star matter the entropy $s=1, 2$. It is a monotonically increasing function with respect to density, and takes the maximum value of several tens MeV at the center. The difference from no kaon case is rather small. In Fig.2(b) we depict the mass-radius relation for protoneutron stars with isentropic structure. The branch of protoneutron stars is clearly separated into two parts by the gravitationally unstable region. We can see that thermal effects are insufficient to support higher mass; on the contrary, the maximum mass is a little bit reduced at $T\neq 0$. \begin{figure}[htb] \begin{minipage}{0.49\textwidth} \epsfsize=0.95\textwidth \epsffile{usr.eps} \end{minipage}% \hfill~% \begin{minipage}{0.49\textwidth} \epsfsize=0.95\textwidth \epsffile{MRr.eps} \end{minipage}% \caption{(a)Temperature profile of the kaon-condensed matter. (b)Mass-radius relation for protoneutron stars (solid lines). Positive slope region means the gravitationally unstable branch. The symbols $N, C, K$ correspond to those in Fig. 1(a). } \label{fig:toosmall} \end{figure} \section{SUMMARY AND CONCLUDING REMARKS} A systematic formulation to include fluctuations around the condensate is presented by introducing the local coordinate on the chiral manifold. This procedure makes the aspect of chiral rotation prominant for the kaon condensation. Using this we obtained the dispersion relation for the kaonic mode; one is the Goldstone mode as a consequence of the SSB of the $V$-spin symmetry, while the other is a very massive mode. The EOS is obtained, in the HBL, for the isothermal and isentropic cases, where the role of thermal kaons should be noted. The EOS exhibits FOPT and it might be interesting to see the appearance of the coexistent phase in the isentropic case. On the other hand, the temperature profile is little changed from that in no kaon matter. These results may be relevant for the delayed collapse during the initial cooling era, where neutrinos are no longer trapped. Hence it is interesting to observe how these results affect the collapsing process and the profile of the neutrino liminocity by dynamical simulations. The maximum mass of protoneutron stars is around $1.6M_\odot$ and is not larger than that of cold neutron stars in this calculation, which suggests more elaborate study is needed to include nucleon dynamics and the neutrino trapping for observing the delayed collapse.
\section*{Figure Captions\markboth {FIGURECAPTIONS}{FIGURECAPTIONS}}\list {Figure \arabic{enumi}:\hfill}{\settowidth\labelwidth{Figure 999:} \leftmargin\labelwidth \advance\leftmargin\labelsep\usecounter{enumi}}} \let\endfigcap\endlist \relax \def\tablecap{\section*{Table Captions\markboth {TABLECAPTIONS}{TABLECAPTIONS}}\list {Table \arabic{enumi}:\hfill}{\settowidth\labelwidth{Table 999:} \leftmargin\labelwidth \advance\leftmargin\labelsep\usecounter{enumi}}} \let\endtablecap\endlist \relax \def\reflist{\section*{References\markboth {REFLIST}{REFLIST}}\list {[\arabic{enumi}]\hfill}{\settowidth\labelwidth{[999]} \leftmargin\labelwidth \advance\leftmargin\labelsep\usecounter{enumi}}} \let\endreflist\endlist \relax \def\list{}{\rightmargin\leftmargin}\item[]{\list{}{\rightmargin\leftmargin}\item[]} \let\endquote=\endlist \makeatletter \newcounter{pubctr} \def\@ifnextchar[{\@publist}{\@@publist}{\@ifnextchar[{\@publist}{\@@publist}} \def\@publist[#1]{\list {[\arabic{pubctr}]\hfill}{\settowidth\labelwidth{[999]} \leftmargin\labelwidth \advance\leftmargin\labelsep \@nmbrlisttrue\def\@listctr{pubctr} \setcounter{pubctr}{#1}\addtocounter{pubctr}{-1}}} \def\@@publist{\list {[\arabic{pubctr}]\hfill}{\settowidth\labelwidth{[999]} \leftmargin\labelwidth \advance\leftmargin\labelsep \@nmbrlisttrue\def\@listctr{pubctr}}} \let\endpublist\endlist \relax \makeatother \newskip\humongous \humongous=0pt plus 1000pt minus 1000pt \def\mathsurround=0pt{\mathsurround=0pt} \def\eqalign#1{\,\vcenter{\openup1\jot \mathsurround=0pt \ialign{\strut \hfil$\displaystyle{##}$&$ \displaystyle{{}##}$\hfil\crcr#1\crcr}}\,} \newif\ifdtup \def\panorama{\global\dtuptrue \openup1\jot \mathsurround=0pt \everycr{\noalign{\ifdtup \global\dtupfalse \vskip-\lineskiplimit \vskip\normallineskiplimit \else \penalty\interdisplaylinepenalty \fi}}} \def\eqalignno#1{\panorama \tabskip=\humongous \halign to\displaywidth{\hfil$\displaystyle{##}$ \tabskip=0pt&$\displaystyle{{}##}$\hfil \tabskip=\humongous&\llap{$##$}\tabskip=0pt \crcr#1\crcr}} \relax \def\begin{equation}{\begin{equation}} \def\end{equation}{\end{equation}} \def\begin{eqnarray}{\begin{eqnarray}} \def\end{eqnarray}{\end{eqnarray}} \def\bar{\partial}{\bar{\partial}} \def\bar{J}{\bar{J}} \def\partial{\partial} \def f_{,i} { f_{,i} } \def F_{,i} { F_{,i} } \def f_{,u} { f_{,u} } \def f_{,v} { f_{,v} } \def F_{,u} { F_{,u} } \def F_{,v} { F_{,v} } \def A_{,u} { A_{,u} } \def A_{,v} { A_{,v} } \def g_{,u} { g_{,u} } \def g_{,v} { g_{,v} } \def\kappa{\kappa} \def\rho{\rho} \def\alpha{\alpha} \def {\bar A} {\Alpha} \def\beta{\beta} \def\Beta{\Beta} \def\gamma{\gamma} \def\Gamma{\Gamma} \def\delta{\delta} \def\Delta{\Delta} \def\epsilon{\epsilon} \def\Epsilon{\Epsilon} \def\p{\pi} \def\Pi{\Pi} \def\chi{\chi} \def\Chi{\Chi} \def\theta{\theta} \def\Theta{\Theta} \def\mu{\mu} \def\nu{\nu} \def\omega{\omega} \def\Omega{\Omega} \def\lambda{\lambda} \def\Lambda{\Lambda} \def\s{\sigma} \def\Sigma{\Sigma} \def\varphi{\varphi} \def{\cal M}{{\cal M}} \def\tilde V{\tilde V} \def{\cal V}{{\cal V}} \def\tilde{\cal V}{\tilde{\cal V}} \def{\cal L}{{\cal L}} \def{\cal R}{{\cal R}} \def{\cal A}{{\cal A}} \def{\cal{G} }{{\cal{G} }} \def{\cal{D} } {{\cal{D} } } \defSchwarzschild {Schwarzschild} \defReissner-Nordstr\"om {Reissner-Nordstr\"om} \defChristoffel {Christoffel} \defMinkowski {Minkowski} \def\bigskip{\bigskip} \def\noindent{\noindent} \def\hfill\break{\hfill\break} \def\qquad{\qquad} \def\bigl{\bigl} \def\bigr{\bigr} \def\overline\del{\overline\partial} \def\relax{\rm I\kern-.18em R}{\relax{\rm I\kern-.18em R}} \def$SL(2,\IR)_{-k'}\otimes SU(2)_k/(\IR \otimes \tilde \IR)${$SL(2,\relax{\rm I\kern-.18em R})_{-k'}\otimes SU(2)_k/(\relax{\rm I\kern-.18em R} \otimes \tilde \relax{\rm I\kern-.18em R})$} \def Nucl. Phys. { Nucl. Phys. } \def Phys. Lett. { Phys. Lett. } \def Mod. Phys. Lett. { Mod. Phys. Lett. } \def Phys. Rev. Lett. { Phys. Rev. Lett. } \def Phys. Rev. { Phys. Rev. } \def Ann. Phys. { Ann. Phys. } \def Commun. Math. Phys. { Commun. Math. Phys. } \def Int. J. Mod. Phys. { Int. J. Mod. Phys. } \def\partial_+{\partial_+} \def\partial_-{\partial_-} \def\partial_{\pm}{\partial_{\pm}} \def\partial_{\mp}{\partial_{\mp}} \def\partial_{\tau}{\partial_{\tau}} \def \bar \del {\bar \partial} \def {\bar h} { {\bar h} } \def \bphi { {\bar \phi} } \def {\bar z} { {\bar z} } \def {\bar A} { {\bar A} } \def {\tilde {A }} { {\tilde {A }}} \def {\tilde {\A }} { {\tilde { {\bar A} }}} \def {\bar J} {{\bar J} } \def {\tilde J} { {\tilde {J }}} \def {1\over 2} {{1\over 2}} \def {1\over 3} {{1\over 3}} \def \over {\over} \def\int_{\Sigma} d^2 z{\int_{\Sigma} d^2 z} \def{\rm diag}{{\rm diag}} \def{\rm const.}{{\rm const.}} \def\relax{\rm I\kern-.18em R}{\relax{\rm I\kern-.18em R}} \def^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1}{^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1}} \def$SL(2,\IR)\otimes SO(1,1)^{d-2}/SO(1,1)${$SL(2,\relax{\rm I\kern-.18em R})\otimes SO(1,1)^{d-2}/SO(1,1)$} \def$SL(2,\IR)_{-k'}\otimes SU(2)_k/(\IR \otimes \tilde \IR)${$SL(2,\relax{\rm I\kern-.18em R})_{-k'}\otimes SU(2)_k/(\relax{\rm I\kern-.18em R} \otimes \tilde \relax{\rm I\kern-.18em R})$} \def$SO(d-1,2)_{-k}/ SO(d-1,1)_{-k}${$SO(d-1,2)_{-k}/ SO(d-1,1)_{-k}$} \def$SO(d-1,2)/ SO(d-1,1)${$SO(d-1,2)/ SO(d-1,1)$} \defPoisson--Lie T-duality{Poisson--Lie T-duality} \def{\cal M}{{\cal M}} \def\tilde V{\tilde V} \def{\cal V}{{\cal V}} \def\tilde{\cal V}{\tilde{\cal V}} \def{\cal L}{{\cal L}} \def{\cal R}{{\cal R}} \def{\cal A}{{\cal A}} \def{\tilde X}{{\tilde X}} \def{\tilde J}{{\tilde J}} \def{\tilde P}{{\tilde P}} \def{\tilde L}{{\tilde L}} \begin{document} \renewcommand{\thesection.\arabic{equation}}}{\thesection.\arabic{equation}} \newcommand{\begin{equation}}{\begin{equation}} \newcommand{\eeq}[1]{\label{#1}\end{equation}} \newcommand{\begin{eqnarray}}{\begin{eqnarray}} \newcommand{\eer}[1]{\label{#1}\end{eqnarray}} \newcommand{\eqn}[1]{(\ref{#1})} \begin{titlepage} \begin{center} \hfill CERN-TH/99-112\\ \hfill hep--th/9904188\\ \vskip .8in {\large \bf Duality-invariant class of two-dimensional field theories} \vskip 0.6in {\bf Konstadinos Sfetsos} \vskip 0.1in {\em Theory Division, CERN\\ CH-1211 Geneva 23, Switzerland\\ {\tt <EMAIL>}}\\ \vskip .2in \end{center} \vskip .6in \centerline{\bf Abstract } \vskip 0,2cm \noindent We construct a new class of two-dimensional field theories with target spaces that are finite multiparameter deformations of the usual coset $G/H$-spaces. They arise naturally, when certain models, related by Poisson--Lie T-duality, develop a local gauge invariance at specific points of their classical moduli space. We show that canonical equivalences in this context can be formulated in loop space in terms of parafermionic-type algebras with a central extension. We find that the corresponding generating functionals are non-polynomial in the derivatives of the fields with respect to the space-like variable. After constructing models with three- and two-dimensional targets, we study renormalization group flows in this context. In the ultraviolet, in some cases, the target space of the theory reduces to a coset space or there is a fixed point where the theory becomes free. \vskip 4cm \noindent CERN-TH/99-112\\ April 1999\\ \end{titlepage} \vfill \eject \def1.2{1.2} \baselineskip 16 pt \noindent \def{\tilde T}{{\tilde T}} \def{\tilde g}{{\tilde g}} \def{\tilde L}{{\tilde L}} \section{Introduction} Cosets $G/H$ as target spaces in 2-dim field theories have been extensively studied in the literature, as they provide examples of spaces other than group manifolds, which give rise to integrable models.\footnote{Examples include the $O(N)$ \cite{ON1}, the principal chiral \cite{PC} and the Gross--Neveu models \cite{GN1}, for which the complete $S$-matrix was found through the existence of higher-spin-conserved currents that lead to its factorization property. Building on work in \cite{PWW}, comparison between the $S$-matrix results and those obtained by perturbative techniques in the ultraviolet (UV) regime was made for the $O(N)$ $\s$-model \cite{ON2}, the Principal Chiral models for $SU(N)$ \cite{SUN}, $SO(N)$ and $Sp(N)$ \cite{sosp} and the $O(N)$ Gross--Neveu model \cite{GN2}, finding perfect agreement.} It is always of interest to find integrable deformations \cite{rajeev}--\cite{sausage} of such models and if possible classify them. In the ordinary (undeformed) coset models one starts with the usual Wess--Zumino action for a group, with Lagrangian density proportional to ${\rm Tr}(\partial_i g^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} \partial^i g)$, and then restricts the trace to the coset space only. Hence, this construction, but not the corresponding models, is quite trivial. Having in mind 2-dim field theories, with targets spaces representing continuous deformations of the latter coset spaces, we need models with non-trivial moduli as a starting point. Such an example was considered in \cite{PLsfe3}; we present in this paper the generalization of this to a class of theories. We found natural to start, in section 2, with 2-dim models related by Poisson--Lie T-duality\ \cite{KliSevI}, since these have indeed a non-trivial moduli space and, moreover, their classical equivalence has been established \cite{PLsfe1, PLsfe2}. Also, in some examples, there are hints that point towards the classical equivalence promoted into a quantum one at 1-loop in perturbation theory \cite{PLsfe3,toappp}. We will show that in some points in this moduli space a local (gauge-like) invariance is developing. Hence, at these points the configuration space is lower-dimensional and we discover in a unifying manner spaces that are deformations of the usual coset spaces. In addition, as a byproduct, we will obtain duals of these models that are classically canonically equivalent to them as 2-dim field theories. This equivalence is encoded in infinite-dimensional current algebras of the parafermionic type that we construct. We derive these from the infinite-dimensional algebras with a central extension, which were found in the proof of canonical equivalence of the Poisson--Lie T-duality-related models in \cite{PLsfe2}. The corresponding generating functionals have the new feature that they are not linear in the derivatives of the fields with respect to the space-like variable. This is in contrast with the cases of Abelian duality \cite{AALcan}, non-Abelian duality in Principal Criral \cite{zacloz,loz} and more general \cite{sfepara} models, as well as for Poisson--Lie T-duality\ (and its possible generalizations) \cite{PLsfe2,PLsfe3}. They are, instead, non-polynomial functions of these derivatives. Many of these aspects are explicitly demonstrated in section 3, with a particular example. In section 4 we discuss the renormalization group (RG) flow in this context. As in \cite{PLsfe3}, we emphasize that taking the classical limit that leads to the lower-dimensional models and then studying the RG flow does not necessarily imply that this limit would correspond to a fixed point of the RG flow, i.e. the two procedures do not commute. There is, however, a particular domain in parameter space, where for a wide range of energies in the UV, the description is effectively perturbative with a UV-stable fixed point corresponding to the point where the gauge invariance develops. Then the model becomes effectively a two-dimensional one. We end the paper with section 5, containing concluding remarks and a discussion on future directions of this research. We have also written an appendix, where some mathematical aspects of our proofs are worked out explicitly. \section{General formulation} \setcounter{equation}{0} In this section we first show how new 2-dim field theories, with target spaces representing deformed coset spaces, arise in the context of Poisson--Lie T-duality-related $\s$-models. We then present a duality-invariant formulation and show that canonical equivalences are encoded into algebras of the parafermionic-type in loop space. \subsection{Formulation using Poisson--Lie T-duality-related $\s$-models} The form of 2-dim $\s$-model actions related by Poisson--Lie T-duality\ (in the absence of spectator fields) is \cite{KliSevI} \begin{equation} S= {1\over 2 \lambda} \int E_{AB} L^A_M L^B_N \partial_+ X^M \partial_- X^N~ , ~~~~ E= (E_0^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} + \Pi)^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1}~ , \label{action1} \end{equation} and \begin{equation} \tilde S= {1\over 2 \lambda} \int \tilde E^{AB} {\tilde L}_{AM} {\tilde L}_{BN} \partial_+ \tilde X^M \partial_- \tilde X^N~ , ~~~~ \tilde E=(E_0 + \tilde \Pi)^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1}~ . \label{action2} \end{equation} The field variables in \eqn{action1} are $X^M$, $\mu=1,2,\dots ,d_G$ and parametrize an element $g$ of a group $G$. We also introduce representation matrices $\{T_A\}$, with $A=1,2,\dots, d_G$ and the components of the left-invariant Maurer--Cartan forms $L^A_M$. The light-cone coordinates on the 2-dim space-time are $x^\pm ={1\over 2} (t \pm x)$, whereas $\lambda$ denotes the overall coupling constant, which is assumed to be positive. Similarly, for \eqn{action2} the field variables are $\tilde X^M$, where $\tilde X^\mu$, $\mu=1,2,\dots ,d_G$, parametrize a different group $\tilde G$, whose dimension is, however, equal to that of $G$. Accordingly, we introduce a different set of representation matrices $\{\tilde T^A\}$, with $A=1,2,\dots, d_G$, and the corresponding components of the left-invariant Maurer--Cartan forms $\tilde L_{AM}$. In \eqn{action1} and \eqn{action2}, $E_0$ is a constant $d_G \times d_G$ matrix, whereas $\Pi$ and $\tilde \Pi$ are antisymmetric matrices with the same dimension as $E_0$, but they depend on the variables $X^M$ and $\tilde X^M$ via the corresponding group elements $g$ and $\tilde g$. They are defined as \cite{KliSevI} \begin{equation} \Pi^{AB} = b^{CA} a_C{}^B ~ , ~~~~~~ \tilde \Pi_{AB} = \tilde b_{CA} \tilde a^C{}_B ~ , \label{pipi} \end{equation} where the matrices $a(g)$, $b(g)$ are constructed using \begin{equation} g^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} T_A g = a_A{}^B T_B~ ,~~~~ g^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} \tilde T^A g = b^{AB} T_B + (a^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1})_B{}^A \tilde T^B~ , \label{abpi} \end{equation} and similarly for $\tilde a(\tilde g)$ and $\tilde b(\tilde g)$. Consistency restricts these to obey \begin{equation} a(g^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1}) = a^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1}(g)~ ,~~~~ b^T(g)= b(g^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1})~ ,~~~~ \Pi^T(g) = - \Pi(g) ~ , \label{conss} \end{equation} and similarly for the tilded ones. There is also a bilinear invariant $\langle{\cdot|\cdot \rangle}$ with the various generators obeying \begin{equation} \langle{T_A|T_B\rangle}= \langle{\tilde T^A|\tilde T^B \rangle}= 0~ , ~~~~ \langle{T_A|\tilde T^B \rangle} = \delta_A{}^B ~ . \label{bili} \end{equation} Finally, we note that the choice of possible groups $G$ and $\tilde G$ is restricted by the fact that \cite{KliSevI} their corresponding Lie algebras must form a pair of maximally isotropic subalgebras into which the Lie algebra of a larger group $D$, known as the Drinfeld double, can be decomposed \cite{alemal}. Let us consider two subgroups $H\in G$ and $\tilde H\in \tilde G$ with $d_H = d_{\tilde H}$. Accordingly we split the Lie-algebra indices as $A=(a,\alpha)$, where Latin and Greek indices refer to subgroup and coset spaces, respectively. Then we may separate the various matrices appearing in (\ref{action1}) and \eqn{action2} into blocks as \begin{equation} (E_0^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1})^{AB}= \pmatrix{E_0^{ab} & E_2^{a\beta}\cr E_3^{\alpha b} & E_1^{\alpha\beta}\cr }\ , \label{bllo} \end{equation} and \begin{equation} \Pi^{AB}= \pmatrix{\Pi_0^{ab} & \Pi_2^{a\beta}\cr -\Pi_2^{b \alpha } & \Pi_1^{\alpha\beta}}\ ,\qquad \tilde \Pi_{AB}= \pmatrix{(\tilde \Pi_0)_{ab} & (\tilde \Pi_2)_{a\beta}\cr -(\tilde \Pi_2)_{b \alpha } & (\tilde \Pi_1)_{\alpha\beta}}\ . \label{bllo1} \end{equation} \noindent We would like to take a limit in the model (\ref{action1}) and its dual (\ref{action2}) such that the number of fields $X^M$ (and ${\tilde X}^M$) is reduced by $d_H$. We would call the remaining variables by $X^\mu$ (and $\tilde X^\mu$) with $\mu=1,2,\dots , d_{G/H}$. Consider the limit \begin{equation} E_0^{ab}\to \infty\quad \Longleftrightarrow \quad (E_0^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1})_{ab}\to 0\ , \label{eoin} \end{equation} in a uniform way for all matrix elements. This means that ratios of matrix elements remain constant in this limit. Using (\ref{bllo}) we find that in the limit \eqn{eoin} \begin{equation} E_0 \approx \pmatrix{0& 0 \cr 0 & E_1^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1}} \ . \label{eoap} \end{equation} Then, the actions (\ref{action1}) and (\ref{action2}) take the form \begin{equation} S= {1\over 2 \lambda} \int \Sigma_{\alpha\beta} L^\alpha_\mu L^\beta_\nu \partial_+ X^\mu \partial_- X^\nu~ , \qquad \Sigma = (E_1 + \Pi_1)^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1}~ , \label{actn1} \end{equation} and \begin{equation} \tilde S= {1\over 2 \lambda} \int \tilde \Sigma^{AB} {\tilde L}_{A\mu} {\tilde L}_{B\nu} \partial_+ \tilde X^\mu \partial_- \tilde X^\nu~ , \qquad \tilde \Sigma=\pmatrix{\tilde \Pi_0 & \tilde \Pi_2 \cr -\tilde \Pi_2 & E_1^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} +\tilde \Pi_1 }^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1}~ . \label{actn2} \end{equation} Notice that in (\ref{actn1}) $\Sigma_{\alpha\beta}$ are elements of a $d_{G/H}\times d_{G/H}$ matrix, whereas in (\ref{actn2}) $\tilde \Sigma^{AB}$ are elements of a $d_G\times d_G$ one. We have anticipated that the number of variables in (\ref{actn1}) and (\ref{actn2}) has been reduced to $d_{G/H}$ upon taking the limit (\ref{eoin}). However, this does not happen automatically, but depends on whether or not certain conditions, as we will next prove, are fulfilled. In order to reduce the dimensionality of (\ref{action1}) we should prove that, after taking the limit (\ref{eoin}), a local gauge invariance develops, which suffices to gauge-fix $d_H$ degrees of freedom in the actions \eqn{action1} and similarly for \eqn{action2}. For \eqn{action1} consider the transformation \begin{equation} g\to g h \ ,\qquad h(x^+,x^-) \in H\ . \label{trfin} \end{equation} In its infinitesimal form it reads $\delta g = i g \epsilon_a T^a$. We may show that this induces the following transformations: \begin{equation} \delta L_\pm^\alpha = f_{b\gamma}{}^\alpha \epsilon^b L_\pm^\gamma\ ,\qquad \delta X^M = L^M_a \epsilon^a\ . \label{indd} \end{equation} Using these and the relation (A.6) of \cite{PLsfe2}, specialized for coset space indices\footnote{Our symmetrization and antisymmetrization conventions are $(ab)=ab+ba$ and $[ab]=ab-ba$.} \begin{equation} L^M_a \partial_M \Pi_1^{\gamma\delta} = - \tilde f^{\gamma\delta}{}_a - f_{a \beta}{}^{[\gamma} \Pi_1^{\delta]\beta}\ , \label{hgj} \end{equation} we may prove that (\ref{actn1}) is invariant under the gauge transformation (\ref{trfin}), provided that the following condition holds: \begin{equation} \tilde f^{\alpha\beta}{}_c + f_{c\gamma}{}^\alpha E_1^{\gamma\beta} + f_{c\gamma}{}^\beta E_1^{\alpha\gamma}= 0\ , \label{rel1} \end{equation} or equivalently \begin{equation} f_{c\gamma}{}^{(\alpha} S^{\beta)\gamma} = 0\ , \qquad \tilde f^{\alpha\beta}{}_c + f_{c\gamma}{}^{[\beta} A^{\alpha]\gamma} = 0\ , \label{rel2} \end{equation} where we have denoted by $S^{\alpha\beta}$ and $A^{\alpha\beta}$ the symmetric and antisymmetric parts of the matrix $E_1^{\alpha\beta}$. When the conditions (\ref{rel1}) are satisfied then we may gauge-fix $d_H$ parameters in the group element $g\in G$. The most efficient way is to parametrize the group element $g\in G$ as $g=\kappa h$, where $h\in H$ and $\kappa\in G/H$, and then set $h=I$. It can be easily seen that this completely fixes the gauge freedom. There are $d_{G/H}^2 d_H$ algebraic conditions in \eqn{rel1} for the $d_{G/H}^2$ elements of the matrix $E_1$. Hence, it is not at all obvious that they can be fulfilled for a general Drinfeld double and then for any arbitrary choice of the subgroup $H\subset G$. An obvious simplification occurs when $\tilde G$ is an Abelian group. Then $\Pi^{AB}=0$, $\tilde f^{AB}{}_C=0$ and eq. (\ref{rel1}) is solved by $E_1^{\alpha\beta} \sim \delta^{\alpha\beta}$. Then (\ref{actn1}) with $\Sigma_{\alpha\beta}\sim \delta_{\alpha\beta}$ takes the form of the usual $\s$-model action on the coset $G/H$ space. Accordingly \eqn{actn2} represents its usual non-Abelian dual. Hence, when both groups $G$ and $\tilde G$ are non-Abelian, the models (\ref{actn1}) and (\ref{actn2}) are deformations of the usual models on coset spaces $G/H$ and of their non-Abelian duals. \subsection{ Duality-invariant formulation} We would like to find a duality-invariant action, from which the $\s$-models \eqn{actn1} and \eqn{actn2} originate. It is natural to start with the manifestly Poisson--Lie T-duality-invariant action of \cite{DriDou} from which the $\s$-models \eqn{action1} and \eqn{action2} originate. This action is defined in the Drinfeld double as \cite{DriDou}, \begin{equation} S(l) = I_{0}(l) + {1\over 2\pi} \int dx dt \langle{l^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} \partial_x l |R| l^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} \partial_x l \rangle} \ , \label{actiL} \end{equation} where $I_{0}(l)$ is the WZW action for a group element $l\in D$. The operator $R$ is defined as \cite{DriDou} \begin{equation} R = |R^+_A {\rangle} \eta^{AB} \langle{ }R^+_B| + |R^-_A {\rangle} \eta^{AB} \langle{ }R^-_B| ~ , \label{RRpRm} \end{equation} with \begin{equation} R^\pm_A = T_A \pm (E_0^\pm)_{AB} \tilde T^B ~ , ~~~~ \eta_{AB}= (E_0^+)_{AB} + (E_0^-)_{AB} ~ , \label{rree} \end{equation} \noindent where we have used the notation $E_0^+= E_0$ and $E_0^- = E_0^T$. In the limit \eqn{eoin} we have \begin{equation} \pmatrix{R^\pm_a \cr R_\alpha^\pm } \approx \pmatrix{T_a \cr T_\alpha \pm (E_1^\pm)^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1}_{\alpha\beta} \tilde T^\beta}\ . \label{hjf} \end{equation} \noindent Using this and the conditions \eqn{rel1}, one can show that \eqn{actiL}, in the limit \eqn{eoin}, develops the gauge invariance \begin{equation} l\to l h\ ,\qquad h(t,x)\in H\ , \label{sd} \end{equation} provided that the following constraint is obeyed \begin{equation} \langle l^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} \partial_x l| T_a\rangle = 0\ , \qquad \forall\ T_a\ . \label{coonn} \end{equation} In order to avoid introducing this constraint we may use gauge fields instead. Indeed, consider the action \begin{equation} S(l,A_t) = I_{0}(l) + {1\over 2\pi} \int \langle{l^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} \partial_x l |R_{g/h}|l^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} \partial_x l \rangle} -{1\over \pi} \int \langle l^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} \partial_x l| A_t\rangle \ , \label{gauac} \end{equation} \noindent where $A_t$ takes values in the Lie algebra of $H$, i.e. $A_t = A_t^a T_a$. The operator $R_{g/h}$ is defined as the restriction in $G/H$ of the corresponding operator in \eqn{rree} \begin{equation} R_{g/h} = |R^+_\alpha {\rangle} \eta_1^{\alpha\beta} \langle{ }R^+_\beta| + |R^-_\alpha {\rangle} \eta_1^{\alpha\beta} \langle{ }R^-_\beta| ~ , \label{RRgh} \end{equation} where \begin{equation} R^\pm_\alpha = T_\alpha \pm (E_1^\pm)^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1}_{\alpha\beta} \tilde T^\beta\ ,\qquad (\eta_1)_{\alpha\beta}= (E_1^+)^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1}_{\alpha\beta} + (E_1^-)^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1}_{\alpha\beta} ~ , \label{rrgh} \end{equation} \noindent and $\eta_1^{\alpha\beta}$ is the inverse matrix of $(\eta_1)_{\alpha\beta}$. It can be shown that \eqn{gauac} is gauge-invariant under \eqn{sd} and the corresponding transformation for the gauge field \begin{equation} A_t \to h^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} (A_t - \partial_t) h \ , \label{traa} \end{equation} provided that $R_{g/h}$ is invariant under the similarity transformation \begin{equation} h R_{g/h} h^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} = R_{g/h} \ . \label{ghds} \end{equation} In order to prove \eqn{ghds} we first show that \begin{equation} h R^\pm_\alpha h^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} = \Delta^\pm_\alpha{}^\beta(h) R^\pm_\beta \ ,\qquad \Delta^\pm_\gamma{}^\alpha(h) \Delta^\pm_\delta{}^\beta(h) \eta_1^{\gamma\delta} = \eta_1^{\alpha\beta}\ , \label{ashjg} \end{equation} \noindent for some $h$-dependent matrix $\Delta^\pm_\alpha{}^\beta$. After repeatedly using \eqn{abpi} and a lengthy computation we find that such a matrix exists and is given by \begin{equation} \Delta^\pm_\alpha{}^\beta(h) = (E_1^\pm)^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1}_{\alpha\gamma}(E_1^\pm)^{\delta\beta} a_\delta{}^\gamma(h) \ , \label{dmatr} \end{equation} provided that the following condition holds:\footnote{The various algebraic manipulations are facilitated by the fact that matrix elements $a(h)_A{}^B$ and $b(h)^{AB}$ vanish if their indices are not both Greek or Latin.} \begin{equation} a_{\gamma}{}^{\alpha}(h) \alpha_\delta{}^\beta(h) (E_1^\pm)^{\gamma\delta} = (E_1^\pm)^{\alpha\beta} \pm \Pi^{\alpha\beta}(h) \ , \label{kjd} \end{equation} or equivalently, splitting into the symmetric and antisymmetric parts, \begin{equation} \alpha_\gamma{}^\alpha(h) \alpha_\delta{}^\beta(h) S^{\gamma\delta} = S^{\alpha\beta} \ ,\qquad \alpha_\gamma{}^\alpha(h) \alpha_\delta{}^\beta(h) A^{\gamma\delta} = A^{\alpha\beta} + \Pi^{\alpha\beta}(h)\ . \label{kjd1} \end{equation} At first sight it seems that \eqn{kjd} is more restrictive than the corresponding conditions in \eqn{rel1}, since, unlike \eqn{rel1}, they are valid for finite-gauge transformations. However, we show in the appendix that \eqn{rel1} actually implies \eqn{kjd}. In the remainder of this subsection, we consider the classical equations of motion for the (manifestly) duality and gauge-invariant action \eqn{gauac}. Its variation with respect to all fields is \begin{equation} \delta S(l,A_t) = -{1\over \pi} \int \delta(l^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} \partial_x l) \left(l^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} \partial_t l - R_{g/h} l^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} \partial_x l + A_t\right) + \delta A_t l^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} \partial_x l \ . \label{eqmo} \end{equation} \noindent Specializing to subgroup and coset space indices, we find the equations of motion \begin{eqnarray} \delta(l^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} \partial_x l) &:& \qquad \langle l^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} \partial_\pm l | R^\mp_\alpha \rangle = 0 \ ,\quad \langle l^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} \partial_t l | T_a \rangle = 0\ , \nonumber \\ \delta A_t &:& \qquad \langle l^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} \partial_x l | T_a \rangle = 0\ , \label{jgk} \end{eqnarray} where we have used also the fact that, because of \eqn{bili}, $\langle A_t |T_a\rangle =0$. Hence, the constraint \eqn{coonn} follows as the equation of motion for $A_t$. Using \eqn{hjf}, the equations of motion in \eqn{jgk} can be cast into the form $\langle l^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} \partial_\pm l|R^\mp_A\rangle =0$. These have the same form as the equations of motion for the action \eqn{actiL} \cite{DriDou}. We finally note that the action \eqn{actiL} is manifestly invariant under the transformation $l \to l_0(t) l$ for some $t$-dependent group element $l_0\in D$ \cite{DriDou}. By introducing gauge fields this symmetry can be promoted into a gauge symmetry with $l_0$ a function of $t$ and $x$. This type of gauge invariance, though interesting enough in its own right to be further investigated, has no apparent relation to the one we have just discussed. \subsection{The canonical transformation} Poisson--Lie T-duality-related models are canonically equivalent under the transformation \cite{PLsfe1,PLsfe2} \begin{eqnarray} && \tilde P^A = J^A \ ,\qquad \tilde J_A = P_A\ , \nonumber \\ && J^A= L^A_x + \Pi^{AB} P_B\ ,\qquad \tilde J_A= {\tilde L}_{x A} + \tilde \Pi_{AB} \tilde P^B \ . \label{cabn} \end{eqnarray} This transformation preserves the equal-time Poisson brackets of the conjugate pairs of variables $(J^A,P_A)$ and $(\tilde J_A,\tilde P^A)$ given by \cite{PLsfe2}\footnote{We will not display explicitly the 2-dim space-time dependence of the phase-space variables involved in the various Poisson brackets. It is understood that the first one in the bracket is always evaluated at $x$ and the second one at $y$, whereas the $t$-dependence is common. Also, compared with \cite{PLsfe2}, we have restored in the various Poisson brackets the dependence on the scale $\lambda$.} \begin{eqnarray} \{J^A,J^B\} & = & \tilde f^{AB}{}_C J^C \delta(x\!-\!y) ~ , \nonumber \\ \{P_A,P_B\} & = & f_{AB}{}^C P_C \delta(x\!-\!y) ~ , \label{JPJP} \\ \{J^A,P_B\} & = & \left( f_{BC}{}^A J^C - \tilde f^{AC}{}_B P_C \right) \delta(x\!-\!y) + {1\over \lambda} \delta_B{}^A \delta^\prime(x\!-\!y) ~ , \nonumber \end{eqnarray} and \begin{eqnarray} \{\tilde J_A,\tilde J_B\}& =& f_{AB}{}^C \tilde J_C \delta(x\!-\!y) ~ , \nonumber \\ \{\tilde P^A,\tilde P^B\}& = & \tilde f^{AB}{}_C \tilde P^C \delta(x\!-\!y) ~ , \label{tJPJP} \\ \{\tilde J_A,\tilde P^B\}& =&\left( \tilde f^{BC}{}_A \tilde J_C - f_{AC}{}^B \tilde P^C \right) \delta(x\!-\!y) + {1\over \lambda} \delta_A{}^B \delta^\prime(x\!-\!y) ~ , \nonumber \end{eqnarray} where $\epsilon(x\!-\!y)$ is the antisymmetric step function that equals $+1$($-1$) for $x>y$ ($x<0$). Notice that the above Poisson brackets are independent of the details of the $\s$-models related by Poisson--Lie T-duality. They are simply the central extensions, in loop space, of the usual Lie-(bi-)algebras defined in the Drinfeld double. One may also show that the Hamiltonians of the two dual actions \eqn{action1} and \eqn{action2} are equal \cite{PLsfe1} as required for canonical transformation with no explicit $t$-dependence. After some algebraic manipulations, these Hamiltonians can be written as \begin{equation} H = {\lambda\over 2} J^A (G_0-B_0 G_0^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} B_0)_{AB} J^B + {\lambda\over 2} P_A (G_0^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1})^{AB} P_B - \lambda J^A(B_0 G_0^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1})_A{}^B P_B \ , \label{haml} \end{equation} and \begin{equation} \tilde H ={\lambda\over 2} \tilde J_A (\tilde G_0-\tilde B_0 \tilde G_0^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} \tilde B_0)^{AB} \tilde J_B + {\lambda\over 2} \tilde P^A (\tilde G_0^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1})_{AB} \tilde P^B - \lambda \tilde J_A (\tilde B_0 \tilde G_0^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1})^A{}_B \tilde P^B \ , \label{hamldu} \end{equation} where $G_0$ and $B_0$ are the symmetric and antisymmetric parts of $E_0^+$ and similarly $\tilde G_0$ and $\tilde B_0$ are the symmetric and antisymmetric parts of $(E_0^+)^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1}$.\footnote{The proof that $\tilde H=H$ uses the fact that \begin{equation} \tilde G_0^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} = G_0-B_0 G_0^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} B_0\ ,\qquad \tilde B_0 \tilde G_0^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} = - G_0^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} B_0\ , \label{gobo} \end{equation} as well as the similar expressions obtained by interchanging tilded and untilded symbols. } Notice that in the limit \eqn{eoin} the conjugate momenta $P_a$ vanish. This is consistent with the development of a local gauge invariance \eqn{trfin}. At the level of the Poisson brackets the vanishing of $P_a$, together with its conjugate $J^a$, has to be imposed as a constraint. In fact they form a set $\varphi_a=(P_a,J^a)$ of second-class constraints. We may see that in the limit \eqn{eoin} and upon using \eqn{rel2}, the Hamiltonians \eqn{haml} and \eqn{hamldu} reduce to \begin{equation} H_{G/H} = {\lambda\over 2} (S^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1})_{\alpha\beta} J^\alpha J^\beta + {\lambda\over 2} S^{\alpha\beta} P_\alpha P_\beta + \lambda (S^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} A)_\alpha{}^\beta J^\alpha P_\beta\ , \label{amco1} \end{equation} and \begin{equation} \tilde H_{G/H} = {\lambda\over 2} S^{\alpha\beta} \tilde J_\alpha \tilde J_\beta + {\lambda\over 2} (S^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1})_{\alpha\beta} \tilde P^\alpha \tilde P^\beta + \lambda (S^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} A)_\alpha{}^\beta \tilde P^\alpha \tilde J_\beta\ . \label{amco2} \end{equation} We may show, with the help of \eqn{JPJP} and \eqn{tJPJP}, that $\{H_{G/H},P_a\}=\{H_{G/H},J^a\}\simeq 0$ (weakly). Hence, no new constraints are generated by the time $t$-evolution. In general (see, for instance, \cite{dirac}), in the presence of a set of second-class constraints $\{\varphi_a\}$, one computes the antisymmetric matrix associated with their Poisson brackets $D_{ab}=\{\varphi_a,\varphi_b\}$. When $D_{ab}$ is invertible one simply postulates that the usual Poisson brackets are replaced by Dirac brackets, defined as \begin{equation} \{A,B\}_D = \{A,B\} - \{A,\varphi_a\} (D^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1})^{ab} \{\varphi_b,B\}\ , \label{diir} \end{equation} for any two phase-space variables $A$ and $B$. In our case we compute the (infinite-dimensional) matrix \begin{equation} D(x,y) = {1\over \lambda} \pmatrix{0 & \delta_a{}^b\cr \delta_a{}^b & 0 } \delta^\prime(x\!-\!y) \ , \label{dxy} \end{equation} with inverse \begin{equation} D^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1}(x,y) ={\lambda\over 2} \pmatrix{0 & \delta_a{}^b\cr \delta_a{}^b & 0 } \epsilon(x\!-\!y) \ . \label{dxyinv} \end{equation} Then the Dirac brackets can be computed using \eqn{diir}. We find (for notational convenience in the rest of the paper, we omit the subscript $D$ from the Dirac brackets): \begin{eqnarray} \{J^\alpha,J^\beta\}&=&\tilde f^{\alpha\beta}{}_\gamma J^\gamma \delta(x\!-\!y) - {\lambda\over 2} \epsilon(x\!-\!y) F_1^{\alpha\beta}(x,y) \ ,\nonumber \\ F_1^{\alpha\beta}(x,y) &\equiv & (f_{c\gamma}{}^\alpha \tilde f^{c\beta}{}_\delta + f_{c\delta}{}^\beta \tilde f^{c\alpha}{}_\gamma) J^\gamma(x) J^\delta(y) \label{jjpa1}\\ && - \tilde f^{\alpha\gamma}{}_c \tilde f^{c\beta}{}_\delta P_\gamma(x) J^\delta(y) - \tilde f^{\beta\gamma}{}_c \tilde f^{c\alpha}{}_\delta P_\gamma (x) J^\delta(x) \ , \nonumber \end{eqnarray} \begin{eqnarray} \{P_\alpha,P_\beta\} & = & f_{\alpha\beta}{}^\gamma P_\gamma \delta(x\!-\!y) -{\lambda\over 2} \epsilon(x\!-\!y) (F_2)_{\alpha\beta}(x,y)\ , \nonumber \\ (F_2)_{\alpha\beta}(x,y) &\equiv &(\tilde f^{c\gamma}{}_\alpha f_{c\beta}{}^\delta + \tilde f^{c\delta}{}_\beta f_{c\alpha}{}^\gamma) P_\gamma(x) P_\delta(y) \label{jjpa2}\\ && - f_{\alpha\gamma}{}^c f_{c\beta}{}^\delta J^\gamma(x) P_\delta(y) - f_{\beta\gamma}{}^c f_{c\alpha}{}^\delta J^\gamma (y) P_\delta(x) \ , \nonumber \end{eqnarray} \begin{eqnarray} \{ J^\alpha,P_\beta\} & = & \left(f_{\beta\gamma}{}^\alpha J^\gamma -\tilde f^{\alpha\gamma}{}_\beta P_\gamma\right)\delta(x\!-\!y) + {1\over \lambda} \delta^\alpha{}_\beta \delta^\prime(x\!-\!y) -{\lambda\over 2} \epsilon(x\!-\!y) F^\alpha_{3\beta}(x,y)\ , \nonumber \\ F^\alpha_{3\beta}(x,y) &\equiv & \left( f_{c\gamma}{}^\alpha J^\gamma(x) - \tilde f^{\alpha\gamma}{}_c P_\gamma(x)\right) \left( f_{\beta\delta}{}^c J^\delta(y) - \tilde f^{c\delta}{}_\beta P_\delta(y) \right) \label{jjpa3}\\ && + \tilde f^{\alpha c}{}_\gamma f_{c\beta}{}^\delta J^\gamma(x) P_\delta(y)\ . \nonumber \end{eqnarray} Notice the parafermionic character of this algebra,\footnote{This is reminiscent of the parafermionic algebras that appeared \cite{paraa} in the study of classical aspects of exact conformal field theories corresponding to gauged WZW models.} which is encoded in the terms containing $\epsilon(x\!-\! y)$. The Dirac brackets for the pair $(\tilde J_\alpha,\tilde P^\alpha)$ are obtained from \eqn{jjpa1}--\eqn{jjpa3} by replacing untilded symbols by tilded ones and vice versa. It is instructive to write down the Dirac brackets for the case that the group $\tilde G$ is Abelian, i.e. $\tilde f^{AB}{}_C=0$. We find \begin{eqnarray} \{J^\alpha,J^\beta\} &=& 0\ , \nonumber \\ \{P_\alpha,P_\beta\} &= & f_{\alpha\beta}{}^\gamma P_\gamma \delta(x\!-\!y) \nonumber \\ && +\ {\lambda\over 2} \epsilon(x\!-\!y) \left( f_{\alpha\gamma}{}^c f_{c\beta}{}^\delta J^\gamma(x) P_\delta(y) + f_{\beta\gamma}{}^c f_{c\alpha}{}^\delta J^\gamma (y) P_\delta(x) \right) \ , \label{jghgh}\\ \{ J^\alpha,P_\beta\} & =& f_{\beta\gamma}{}^\alpha J^\gamma \delta(x\!-\!y) + {1\over \lambda} \delta^\alpha{}_\beta \delta^\prime(x\!-\!y) -{\lambda\over 2} \epsilon(x\!-\!y) f_{c\gamma}{}^\alpha f_{\beta\delta}{}^c J^\gamma(x) J^\delta(y) \ . \nonumber \end{eqnarray} The above Dirac brackets can also be obtained from the ones in \eqn{jjpa1}--\eqn{jjpa3} via a contraction that Abelianizes the group $\tilde G$, i.e. $J^\alpha\to {1\over \epsilon} J^\alpha$, $\lambda\to \epsilon \lambda$, $\epsilon\to 0$. \section{An explicit example} \setcounter{equation}{0} In this section we explicitly demonstrate many of the general aspects developed in section 2, using 3- and 2-dim models related by Poisson--Lie T-duality. That includes the explicit construction of the metric and antisymmetric tensor fields, of the Dirac-bracket algebra for canonical equivalence, and also of the corresponding generating functional. \subsection{The Drinfeld double} Our example will be based on the 6-dim Drinfeld double considered in \cite{PLsfe3,toappp,lledo}, which we first review by following \cite{PLsfe3}.\footnote{Recently, a classification was made of all possible Drinfeld doubles based on the 3-dim real Lie algebras (Bianchi algebras) \cite{JR}. It will be interesting to use them for the construction of more examples that could be useful for the investigation of various issues presented in this and the following section.} It is just the non-compact group $SO(3,1)$ with $G=SU(2)$ and dual $\tilde G =E_3 ={\rm solv }(SO(3,1))$ given by the Iwasawa decomposition of $SO(3,1)$ \cite{helgason}. The associated 3-dim algebras $su(2)$ and $e_3$ have generators denoted by $\{T_A\}$ and $\{\tilde T^A\}$, where $A=1,2,3$. Leaving aside the details we only present the elements that are necessary in this paper. It is convenient to split the index $A=(3,\alpha)$, $\alpha=1,2$. The non-vanishing structure constants for the algebras $su(2)$ and $e_3$ are \begin{equation} f_{\alpha\beta}{}^3 = f_{3\alpha}{}^\beta = \epsilon_{\alpha\beta}\ ,\qquad \tilde f^{3\alpha}{}_\beta =\delta_{\alpha\beta}\ , \label{struu} \end{equation} where our normalization is such that $\epsilon_{12}=\delta_{11} =1$. We parametrize the $SU(2)$ group element in terms of the three Euler angles $\phi$, $\psi$ and $\theta$. It is represented by the $4 \times 4$ block-diagonal matrix \begin{equation} g_{SU(2)}= {\rm diag}( g, g ) ~ , \label{gsu2} \end{equation} where \begin{equation} g = e^{{i\over 2}\phi \s_3} e^{{i\over 2}\theta \s_2} e^{{i\over 2}\psi \s_3} = \pmatrix{ \cos {\theta\over 2} e^{{i\over 2} (\phi +\psi)} & \sin {\theta\over 2} e^{{i\over 2} (\phi -\psi)} \cr - \sin {\theta\over 2} e^{-{i\over 2} (\phi -\psi)} & \cos {\theta\over 2} e^{-{i\over 2} (\phi +\psi)} \cr } ~ . \label{su211} \end{equation} Also the group element of $E_3$ is parametrized in terms of three variables $y_1$, $y_2$ and $\chi$ and represented by the following $4\times 4$ block-diagonal matrix \begin{equation} {\tilde g}_{E_3}= {\rm diag}({\tilde g}_+,{\tilde g}_-) ~ , \label{ge3} \end{equation} where \begin{eqnarray} && {\tilde g}_+ = \pmatrix { e^{+ {\chi \over 2}} & \chi_+ \cr 0 & e^{-{\chi\over 2}} \cr} ~ ,~~~~ {\tilde g}_- = \pmatrix { e^{- {\chi \over 2}} & 0 \cr \chi_- & e^{+{\chi\over 2}} \cr} ~ , \nonumber \\ && \chi_\pm = \pm e^{-{\chi\over 2}} (y_1 \mp i y_2) ~ . \label{ge22} \end{eqnarray} The Maurer--Cartan forms in the parametrization of the $SU(2)$ group element \eqn{su211} are \begin{eqnarray} L^1 & =& \cos \psi \sin \theta d\phi - \sin \psi d\theta ~ , \nonumber \\ L^2 & = & \sin \psi \sin \theta d\phi + \cos \psi d\theta ~ , \label{mausu2} \\ L^3 & = & d\psi + \cos \theta d\phi ~ . \nonumber \end{eqnarray} Similarly, using the parametrization \eqn{ge3} for the $E_3$ group element we find \begin{equation} {\tilde L}_1 = e^{-\chi } dy_1 ~ ,~~~~ {\tilde L}_2 = e^{-\chi } dy_2 ~ ,~~~~ {\tilde L}_3 = d\chi ~ . \label{maue3} \end{equation} The antisymmetric matrices $\Pi$ and $\tilde \Pi$ are \begin{equation} \Pi = \pmatrix { 0 & -\sin\psi \sin\theta & \cos\psi \sin\theta \cr \sin\psi \sin\theta & 0 & 1- \cos \theta \cr - \cos \psi \sin \theta & \cos \theta -1 & 0 \cr } ~ , \label{piii} \end{equation} and \begin{equation} \tilde \Pi= \pmatrix{0 & -y_2 e^{-\chi} & y_1 e^{-\chi} \cr y_2 e^{-\chi}& 0 & -{1\over 2} \left(1-(1+ y_1^2 + y_2^2) e^{-2 \chi}\right) \cr - y_1 e^{-\chi} & {1\over 2} \left(1-(1+ y_1^2 + y_2^2) e^{-2 \chi}\right) & 0 \cr } ~ . \label{ttpiij} \end{equation} \subsection{Explicit three- and two-dimensional models} Consider the $\s$-model action \eqn{action1} for the case of our double based on $SO(3,1)$. Let us single-out the 1-dim subgroup $H\simeq U(1)$ that is generated by $T_3$. For our purposes it will be sufficient to use the following form for the $3\times 3$ matrix $E_0^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1}$ \begin{equation} E_0^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} = \pmatrix{(1+g)^{-1} a & 0 & 0\cr 0 & a & b-1 \cr 0 & 1-b & a}\ , \label{eoii} \end{equation} where we have kept the conventions of \eqn{bllo} for the enumeration of the matrix elements. Using \eqn{mausu2}, \eqn{piii} and \eqn{eoii}, it is the easy to compute the metric and antisymmetric tensor fields corresponding to \eqn{action1}. We find a metric given by \begin{equation} ds^2 = {a\over V} \left( (L^1)^2 +(L^2)^2 +(g+1)(L^3)^2 + {1+ g\over a^2} \big( (b \cos\theta -1)\ d\phi +(b-\cos\theta)\ d\psi\big)^2 \right) \ , \label{meett} \end{equation} and an antisymmetric tensor given by \begin{equation} B= 2{ \sin\theta \over V}\ d\theta\wedge \Big((g+1) d\psi + (b + g \cos\theta) d\phi\Big)\ , \label{btht} \end{equation} where \begin{equation} V \equiv a^2 + (b-\cos\theta)^2 + (1+g)\sin^2\theta\ . \label{veo} \end{equation} Notice that the antisymmetric tensor can be (locally) gauged away since the corresponding 3-form field strength is zero. Also, for our purposes, we will not need the explicit expressions for the metric and antisymmetric tensor corresponding to the dual $\s$-model \eqn{action2}. For $b=1$, but general $a$ and $g$, the above example (with its dual) was considered in \cite{PLsfe3} (also in \cite{toappp} for $a=b=1$ and $g=0$). We would like to take the analogue of the limit \eqn{eoin}. It is clear that in our case this corresponds to letting $g\to -1$. Comparing \eqn{eoii} to \eqn{bllo} we see that the $2\times 2$ matrix $E_1$ is \begin{equation} E_1 = \pmatrix{a & b-1 \cr 1-b & a }\ . \label{e11} \end{equation} It is easily seen that this is the most general $2\times 2$ matrix that solves \eqn{rel1}, with structure constants given by \eqn{struu}. In agreement with our general discussion, the $\s$-model action with metric \eqn{meett} and antisymmetric tensor \eqn{btht} develops a local invariance under the transformation \begin{equation} \delta \psi=\epsilon(t,x)\ . \label{trg1} \end{equation} This allows to gauge-fix the variable $\psi =0$. Explicitly computing \eqn{actn1} we find that the metric and antisymmetric tensors are given by \begin{eqnarray} ds^2 & = &{a \over a^2 +(b - \cos\theta)^2 } \ \left(d\theta^2 + \sin^2\theta d\phi^2\right) \ , \nonumber \\ B & = & 2 {\sin\theta (b- \cos\theta) \over a^2 +(b-\cos\theta)^2 }\ d\theta \wedge d\phi \ . \label{dsbth} \end{eqnarray} Equivalently, the same result follows if we set $g=-1$ directly into the expressions for the metric \eqn{meett} and antisymmetric \eqn{btht} tensors. Similarly, the dual model action \eqn{actn2} is invariant under the local transformation \begin{equation} \delta y_\alpha = \epsilon(t,x) \epsilon_{\alpha\beta} y_\beta\ . \label{trg2} \end{equation} Hence, we may evaluate \eqn{actn2} in the gauge $y_1=0$. The corresponding metric (the antisymmetric tensor turns out to be zero) is found to be \begin{eqnarray} && ds^2 = {a_1/2 \over 1+ a_1 z} \left( {dz^2\over \rho^2} + \left( d\rho + \Bigg[ {b-1\over a} + {z-a_1 \rho^2/4\over 1+ a_1 z}\Bigg] {dz\over \rho} \right)^2 \right)\ , \nonumber \\ && a_1 \equiv {2 a\over a^2 + (b-1)^2} \ , \label{dsbth1} \end{eqnarray} where we have changed variables as $y_2^2 = {1\over 4} \rho^2 a_1^2 $ and $e^{2 \chi}= 1+ a_1 z$. The metric\footnote{If we set $b=1$ and redefine $a\to 2/a $ and $\lambda\to \lambda a/2$ the metric \eqn{dsbth} and its dual \eqn{dsbth1} become those derived in \cite{PLsfe3} using a limiting procedure, equivalent to \eqn{eoin}. The deeper reason that validates such a procedure is, as we have shown in the present paper, the development of a local gauge invariance in this limit.} in \eqn{dsbth} is free of singularities (since \eqn{trg1} has no fixed point) and represents a deformed 2-sphere. In contrast, \eqn{dsbth1} is singular for $r=0$. This is related to the fact that $y_1=y_2=0$ is a fixed point of the gauge transformation \eqn{trg2}. The singularity at $1+a_1 z=0$ is only a coordinate singularity and can be removed by an appropriate change of variables. It is worth while to consider some analytic continuations of the models \eqn{dsbth} and its dual \eqn{dsbth1}. If we let $\theta\to i r$, where $r\in [0,\infty)$, and also we change the sign of the overall coupling constant $\lambda$, then \eqn{dsbth} becomes \begin{eqnarray} ds^2 & = &{a \over a^2 +(b - \cosh r)^2 } \ \left(dr^2 + \sinh^2 r d\phi^2\right) \ , \nonumber \\ B & = & 2 {\sinh r (b - \cosh r) \over a^2 +(b-\cosh r)^2 }\ d r \wedge d\phi \ . \label{dsban} \end{eqnarray} The corresponding analytic continuation in the dual metric \eqn{dsbth1} should be $\rho\to i \rho$, with a parallel change of sign in the overall coupling constant. The metric in \eqn{dsban} is reduced to the Euclidean $AdS_2$ metric if we rescale the coupling constant $\lambda\to \lambda/a$ and then take the limit $a\to \infty$ (keeping the new coupling finite). However, for generic values of the constant $a$, it represents a space that is topologically a cigar. Indeed, for $r\to 0$ we get the 2-dim Euclidean space $E^2$ in polar coordinates, whereas for $r\to \infty$ we get, after an appropriate change of variables, $R^1 \times S^1$. For $b>0$, the cigar-shaped space develops a ``pump'' corresponding to the maximum of the metric components $G_{\phi\phi}$ at \begin{equation} \cosh r= {\sqrt{(1+a^2+b^2)^2-4 b^2} + 1+a^2+b^2\over 2 b}\ . \label{hgsk} \end{equation} We note that the cigar-shape topology is also a characteristic of the Euclidean black hole corresponding to the coset $SL(2,\relax{\rm I\kern-.18em R})/U(1)$ exact conformal field theory \cite{witbla}. However, in our case the model \eqn{dsban} is not conformal. The Drinfeld double for \eqn{dsban} and its dual model is $SO(2,2)$, with $G=SL(2,\relax{\rm I\kern-.18em R})$, instead of $SU(2)$. \subsection{The Dirac brackets and the generating functional} The Dirac brackets for the conjugate variables in our example are most easily written down in the basis $J^\pm = J^1\pm i J^2$ and $P_\pm = P_1\pm i P_2$, where the non-zero structure constants are $f_{3\pm}{}^\pm=\pm$, $f_{+-}{}^3 =2$ and $\tilde f^{3\pm}{}_\pm =1$. Using \eqn{jjpa1}--\eqn{jjpa3} we obtain \begin{eqnarray} \{ J^\pm, J^\pm\} & = & \mp \lambda \epsilon(x\!-\!y) \underline{{ J^\pm(x) J^\pm(y)} } \ , \nonumber \\ \{ J^+,J^-\} & = & 0\ , \label{jjss1} \end{eqnarray} \begin{eqnarray} \{ P_\pm,P_\pm \} & = & + \lambda \epsilon(x\!-\!y) \left( \mp \underline{ {P_\pm(x) P_\pm(y)}} + J^\mp(x) P_\pm(y) + P_\pm(x) J^\mp (y) \right)\ , \nonumber \\ \{P_+,P_-\} & = & -\lambda \epsilon(x\!-\!y) \left( J^-(x) P_-(y) + P_+(x) J^+(y) \right)\ , \label{jjss2} \end{eqnarray} \begin{eqnarray} \{ J^\pm,P_\pm\} & = & {1\over \lambda} \delta^\prime(x\!-\! y) -\lambda \epsilon(x\!-\! y) \left( J^\pm(x) J^\mp(y) \mp \underline{{J^\pm(x) P_\pm(y)}} \right) \ , \nonumber \\ \{ J^\pm,P_\mp\} & = & \lambda \epsilon(x\!-\! y) \left( J^\pm(x) J^\pm(y) \pm \underline{{J^\pm(x) P_\mp(y)}} \right ) \ , \label{jjss3} \end{eqnarray} where the underlined terms should be omitted in the Abelian limit of the dual group $\tilde G=E_3$. In this case the above algebra provides a canonical equivalence between the $\s$-model for $S^2$ and its non-Abelian dual with respect to the left (or right) action of $SU(2)$. Note also that the generators $J^\pm$ form a subalgebra \eqn{jjss1}. The generating functional that demonstrates the classical equivalence between $\s$-models related by Poisson--Lie T-duality\ based on our Drinfeld double was explicitly constructed in \cite{PLsfe3}. In a slightly different form than that in \cite{PLsfe3}, it reads\footnote{We also correct a misprint in eq. (26) of \cite{PLsfe3}. In the expression for $B_\psi$ and in the argument for $\cot^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1}$, $(y_1 \cos\psi +y_2 \sin\psi)$ should be replaced by $(y_1 \cos\psi +y_2 \sin\psi)\tan{\theta\over 2}$.} \begin{eqnarray} F & =& \int dx \Big( A \partial_x \phi + (\psi +\alpha -\phi)\partial_x \chi -{2 \rho \tan^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} B \over \sqrt{1+\rho^2 \cos^2(\psi+\alpha)}} \partial_x (\rho \cos(\psi+\alpha)) \Big) \ , \nonumber\\ A & \equiv & - \ln \left( e^{2\chi} \cos^2{\theta\over 2} + e^\chi \rho \sin\theta \sin(\psi+\alpha) + (1+\rho^2) \sin^2{\theta\over 2} \right)\ , \label{fab}\\ B &\equiv & {e^\chi \cot {\theta\over 2} + \rho \sin(\psi+\alpha) \over \sqrt{1+\rho^2 \cos^2(\psi+\alpha)} }\ ,\qquad (y_1,y_2)=\rho (\sin\alpha,\cos\alpha)\ . \nonumber \end{eqnarray} Notice that the above generating functional depends only on the combination $\psi +\alpha$; it is therefore invariant under the $U(1)$ gauge transformation $\delta \psi = \epsilon$ and $\delta \alpha = -\epsilon$. The generating functional for the deformed coset models \eqn{dsbth} is obtained by solving the equation ${\delta F \over \delta \psi} =0$ (equivalently ${\delta F \over \delta \alpha} =0$) for $\psi +\alpha$ and inserting the result back into \eqn{fab}.\footnote{Such a procedure is motivated by the fact that the variations ${\delta F \over \delta \psi}$ and $-{\delta F \over \delta \alpha}$, corresponding to the conjugate momenta $P_\psi$ and $P_\alpha$, are zero since the variables $\psi$ and $\alpha$ have dropped out of the corresponding dual $\s$-models because of the gauge invariance. Also, thanks to the latter, only one of these variations is independent. This procedure has an obvious generalization for the more general coset models constructed in section 2. For some similar considerations, see \cite{ouggroi2} and more recently \cite{stern}.} The result is a generating functional, which is non-polynomial in derivatives with respect to $x$. The obtained expressions are quite complicated and not very illustrative, so that we decided to present the corresponding result for the $\s$-model for $S^2$ and its non-Abelian dual. We start with the generating functional corresponding to the 2-dim $\s$-models for $S^3$ and its non-Abelian dual with respect to the left (or right) action of $SU(2)$ that was obtained in \cite{zacloz}. In our notation it is given by $F=-\int dx (y_1 L^1_x+ y_2 L^2_x + z L^3_x)$. This is easily modified to depend on the angles $\psi$ and $\alpha$ only through the combination $\psi+\alpha$, by adding the term $-\int dx \alpha\partial_x z$. Such a term, being dependent on the variables of only one of the dual models, can be absorbed as total derivative into the corresponding action and hence it does not affect the classical dynamics. Explicitly, the resulting generating functional is \begin{equation} F = - \int dx \Big((z \cos \theta + \rho \sin\theta \sin (\psi +\alpha) )\partial_x \phi + \rho \cos (\psi+\alpha) \partial_x \theta - (\psi+\alpha) \partial_x z \Big) \ . \label{fgh} \end{equation} The variation of $F$ with respect to $\psi+\alpha$ gives \begin{equation} \tan (\psi+\alpha) = { \partial_x z \sqrt{ \rho^2\left((\partial_x\theta)^2 + \sin^2\theta (\partial_x \phi)^2\right) - (\partial_x z)^2} - \rho^2\sin\theta \partial_x\theta \partial_x \phi \over (\partial_x z)^2 - \rho^2 (\partial_x \theta)^2}\ . \label{phisa} \end{equation} Substituting back into \eqn{fgh} we obtain \begin{equation} F = - \int dx \left(\sqrt{\rho^2\left( (\partial_x \theta)^2 + \sin^2 \theta (\partial_x \phi)^2 \right)- (\partial_x z)^2} + z \cos\theta \partial_x \phi - (\psi + \alpha) \partial_x z\right) \ , \label{fgh1} \end{equation} where $\psi+\alpha$ is given by \eqn{phisa}. The generating functional \eqn{fgh1} is non-polynomial in the derivatives of the fields with respect to $x$. In that sense it belongs to a new class of generating functionals, which depend not only on the fields of the two dual $\s$-models, but also on their first derivatives with respect to the space-like variable in a non-trivial way. For comparison, up to now, either in the case of non-Abelian duality \cite{zacloz,loz,sfepara} or for Poisson--Lie T-duality\ (and its possible generalizations) \cite{PLsfe1,PLsfe2}, there was no dependence of the generating functional on more than the first power of these derivatives (see, for example \eqn{fab}).\footnote{A generating functional of the type \eqn{fgh1}, containing first derivatives of the fields in a non-polynomial way, has appeared in a study on the canonical equivalence between Liouville and free field theories \cite{FS} (also \cite{FS1} as quoted in \cite{FS}).} Finally, we note that, according to the work in \cite{qufun}, generating functionals of the form \eqn{fgh1}, being non-linear, are expected to receive quantum corrections. Consequently, the corresponding duality rules relating the 2-dim field theories, as well as the algebra \eqn{jjpa1}--\eqn{jjpa3}, are expected to be quantum-corrected. \section{Renormalization group flow} \setcounter{equation}{0} In this section we study the 1-loop RG equations corresponding to the three-dimensional model \eqn{meett}, \eqn{btht}. We will show that there are no fixed points in the flow and also that the correct description of the models is a non-perturbative one. However, for large domains in parameter space and for a wide range of energies in the UV, the description is effectively perturbative and the model becomes a two-dimensional one. Finally, by performing some analytic continuations we will find three- and two-dimensional models with fixed points under the RG flow, where the theory becomes free. We begin this section with a short review of RG flow in 2-dim field theories with curved target spaces. and \eqn{dsbth}. The 2-dim $\s$-model corresponding to the metric \eqn{meett} and antisymmetric tensor \eqn{btht} is of the form \begin{equation} S={1\over 2\lambda} \int Q^+_{\mu\nu} \partial_+ X^\mu \partial_- X^\nu\ ,\qquad Q^+_{\mu\nu}\equiv G_{\mu\nu} + B_{\mu\nu}\ . \label{fjll} \end{equation} It will be renormalizable if the corresponding counter-terms, at a given order in a loop expansion, can be absorbed into a renormalization of the coupling constant $\lambda$ and (or) of some parameters labelled collectively by $a^i$, $i=1,2,\dots$ In addition, we allow for general field redefinitions of the $X^\mu$'s, which are coordinate reparametrizations in the target space. This definition of renormalizability of $\s$-models is quite strict and similar to that for ordinary field theories. A natural extension of this is to allow for the manifold to vary with the mass scale and the RG to act in the infinite-dimensional space of all metrics and torsions \cite{Frialv}. Further discussion of this generalized renormalizability will not be needed for our purposes. Perturbatively, in powers of $\lambda$, we express the bare quantities, denoted by a zero as a subscript, as \begin{eqnarray} &&\lambda_0 = \mu^\epsilon \lambda \left(1+ {J_1(a)\over \pi \epsilon} \lambda + \cdots \right) \equiv \mu^\epsilon \lambda \left (1+ {y_\lambda \over \epsilon} + \cdots \right ) ~ , \nonumber \\ &&a^i_0 = a^i + {a^i_1(a) \over \pi \epsilon} \lambda + \cdots \equiv a^i \left (1+ {y_{a^i} \over \epsilon} + \cdots \right ) ~ , \label{expp}\\ &&X^\mu_0 = X^\mu + {X^\mu_1(X,a) \over \pi \epsilon} \lambda + \cdots ~ . \nonumber \end{eqnarray} The ellipses stand for higher-order loop- and pole-terms in $\lambda$ and $\epsilon$ respectively. Then, the beta-functions up to one loop are given by $\beta_\lambda = \lambda^2 {\partial y_\lambda\over \partial \lambda} = {\lambda^2\over \pi} J_1$ and $ \beta_{a^i} = \lambda a^i {\partial y_{a^i}\over \partial \lambda}= {\lambda \over \pi} a^i_1$, where, as usual, $\beta_\lambda = {d\lambda\over dt}$, $\beta_{a^i} = {da^i\over dt}$ and $t=\ln \mu$. The equations to be satisfied by appropriately choosing $J_1,a^i_1$ and $X_1^\mu$ are given by \begin{equation} {1\over 2} R^-_{\mu\nu} = -J_1 Q^+_{\mu\nu} + \partial_{a^i} Q^+_{\mu\nu} a^i_1 + \partial_\lambda Q^+_{\mu\nu} X_1^\lambda + Q^+_{\lambda\nu} \partial_\mu X_1^\lambda + Q^+_{\mu\lambda} \partial_\nu X_1^\lambda\ , \label{1loop} \end{equation} where $R^-_{\mu\nu}$ are the components of the ``generalized'' Ricci tensor defined with a connection that includes the torsion, i.e. with $\Gamma^{\mu}_{\nu\rho} - {1\over 2} H^\mu{}_{\nu\rho}$. The corresponding counter-terms were computed in the dimensional regularization scheme (see, for instance, \cite{Osborn}). \subsection{Models with no fixed points} \subsubsection{Three-dimensional models} In the metric \eqn{meett} there are three parameters $a$, $b$ and $g$ and the three Euler angles $\theta,\psi$ and $\phi$ will be denoted by $X^\mu$. Also for the antisymmetric tensor in \eqn{btht} we have $H^\mu{}_{\nu\rho}=0$. Examining \eqn{1loop} we find that the coupling $\lambda$ and the coordinates $(\theta,\psi,\phi)$ do not renormalize and therefore the corresponding beta-functions are zero.\footnote{We believe that the non-renormalization of the overall coupling constant $\lambda$ will persist in general for all Poisson--Lie T-duality-related models with actions \eqn{action1} and \eqn{action2}. On the other hand, models corresponding to a limit of \eqn{meett} and with target space $S^3$ or its deformation along a direction in the Cartan subalgebra of $SU(2)$ \cite{ouggroi} (see also the comments after \eqn{abgli} below), have an overall coupling constant that gets renormalized \cite{ouggroi}. The reason for this apparent paradox is that, in these models, the overall coupling constant is related to our $\lambda$ by rescalings, such as those described in footnote 8, with parameters that get renormalized.} In contrast, for the parameters $a,b$ and $g$ we find \begin{eqnarray} \beta_a& = & {\lambda\over 4\pi} {1+a^2-b^2\over a^2} \Big((g-1) a^2 + (g+1) (b^2-1)\Big)\ , \nonumber\\ \beta_b & = & {\lambda\over 2\pi} {b\over a} \Big((g-1) a^2 + (g+1) (b^2-1)\Big)\ , \label{abg1}\\ \beta_g &=& {\lambda\over 2\pi} {1+g\over a} \Big(g (1+a^2) + (g+2) b^2\Big)\ . \nonumber \end{eqnarray} This system of coupled non-linear equations\footnote{Presumably, the dual to the \eqn{meett}, \eqn{btht} model will also have the same beta-functions \eqn{abg1}. We also note that it is highly non-trivial that the change of the matrix \eqn{eoii} under the RG eqs. \eqn{abg1} preserves its form. For example, had we started, as in \cite{toappp}, with a matrix $E_0^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} $ proportional to the identity, then eqs. \eqn{abg1} would have generated off-diagonal elements. This is clearly seen by computing the right-hand sides of eqs. \eqn{abg1} for $a=b=1$ and $g=0$. By doing so we obtain the infinitesimal change of $E_0^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1} =I$ (to lowest order) and eq. (102) of \cite{toappp} (with $c=1$).} can be considerably simplified. First, using \eqn{abg1}, we may easily show that there is a RG-flow-invariant defined as \begin{equation} {a^2+b^2+1\over b} \equiv 2 \nu = {\rm const.}\ , \label{invv} \end{equation} which implies that \begin{equation} a=\sqrt{(b_+ - b)(b-b_-)}\geq 0\ ,\qquad b_\pm\equiv \nu\pm \sqrt{\nu^2-1}\ , \qquad |\nu|\geq 1 \ . \label{hja} \end{equation} Without loss of generality we may assume that $\nu>0$ since \eqn{invv} remains invariant under $\nu\to -\nu$ and $b\to -b$. Then, using the last two equations in \eqn{abg1} we may derive an equation for $b$ as a function of $g$ whose solution is \begin{equation} b=- g\left(\nu \pm \sqrt{\nu^2 -1+e^{-2 C} (1+1/g)^2}\ \right)\ , \label{bgy} \end{equation} where $C$ is a real constant, which is determined by the initial conditions for $b$ and $g$. The sign in front of the square root in \eqn{bgy} is changed when $g=0$, in order to ensure the continuity of $b$ as a function of the energy scale $t=\ln \mu$. Hence, the only differential equation we still have to solve is the one for $g$, which, after using \eqn{invv}, takes the form \begin{equation} \beta_g = {\lambda\over \pi} {b \over a}\ (g+1) (b+\nu g)\ , \label{abg11} \end{equation} where $a$ and $b$ are determined by \eqn{hja} and \eqn{bgy}. Since the RG equations are real, $a^2$ will stay strictly non-negative and therefore $b$ will oscillate with $t=\ln \mu$ between its minimum and maximum values $b_-$ and $b_+$, where $a=0$. When $a\simeq 0$, for finite values of the overall coupling constant $\lambda$, the curvature for the metric \eqn{meett} approaches infinity and the perturbative expansion of the RG equations becomes meaningless. We have seen that the correct description of the theory is a genuine non-perturbative one. Neverthelss, for $\nu\gg 1$ we will show that there exists a wide range of energies in the UV, where the description is effectively perturbative. Moreover, there exists a fixed point at $g=-1$ where the theory has effectively a 2-dim target space. Indeed, using \eqn{invv}, we have that $a^2\simeq 2 \nu b\gg 1$ when $\nu \gg 1$. Hence, in that limit and after redefining $\lambda\to \lambda/a$ we may simplify the RG eqs. \eqn{abg1} as \begin{eqnarray} \beta_\lambda& \simeq & -{\lambda^2\over 4\pi} (1-g)\ , \nonumber\\ \beta_g &\simeq & {\lambda\over 2\pi}\ g (1+g) \ , \label{abgli}\\ \beta_b & \simeq &- {\lambda\over 2\pi}\ (1-g) b \ . \nonumber \end{eqnarray} Then the metric \eqn{meett} becomes \begin{eqnarray} ds^2 & =& (L^1)^2 + (L^2)^2 + (1+g) (L^3)^2 \nonumber \\ & = & d\theta^2 + \sin^2\theta d\phi^2 + (1+g) (d\psi + \cos\theta d\phi)^2\ , \label{jef} \end{eqnarray} which is the deformed $SU(2)$ Principal Chiral model considered in \cite{ouggroi}. Also the first two of the above equations are those derived in \cite{ouggroi} for the corresponding coupling $\lambda$ and deformation parameter $g$. In the UV the solution of \eqn{abgli} is \begin{equation} \lambda \simeq {2\pi \over t} ,\qquad g \simeq -1 +{{\rm const.}\over t}\ , \qquad b\simeq {{\rm const.}\over t^2} \ ,\qquad {\rm as} \ t\to \infty \ . \label{asdj} \end{equation} Hence, in the UV $a^2 \simeq 2 \nu b \sim 2\nu/t^2$. Therefore if the condition \begin{equation} 1 \ll t \ll \nu^{1/2}\ , \label{sdsf} \end{equation} is fulfilled, then $a\gg 1 $ and the model is indeed described perturbatively by \eqn{jef}. The point $g=-1$ is a UV-fixed point, where the metric \eqn{jef} becomes $S^2$. However, outside the validity of \eqn{sdsf} the correct description is non-perturbative. \subsubsection{Two-dimensional models} Let us now return to the 2-dim models \eqn{dsbth} and \eqn{dsbth1}. As before, there is no wave-function renormalization for $\theta$ and $\phi$, and the beta-function for the coupling $\lambda$ is zero. For the couplings $a$ and $b$ the corresponding beta-functions can be obtained by simply setting $g=-1$ into \eqn{abg1}. The reason why such a procedure is consistent seems to be intimately related to the local invariance that reduces the 3-dim models into 2-dim ones. Hence, we have \begin{eqnarray} \beta_a& = & - {\lambda\over 2\pi}\ (1+a^2-b^2)\ , \nonumber\\ \beta_b & = & - {\lambda\over \pi}\ ab\ , \label{abg2} \end{eqnarray} which are nothing but the beta-functions for the 2-dim model \eqn{dsbth} as well as for its dual \eqn{dsbth1}.\footnote{In order to compare with $\beta_a$ and $\beta_\lambda$ as given by eq. (47) of \cite{PLsfe3}, one should remember that these correspond to the model \eqn{dsban} with $b=1$. Imposing that $b=1$ and further requiring that $\beta_b=0$ enforces a wave-function renormalization of the $\theta$, i.e. in \eqn{expp} we have $\theta_1 =-{1\over a} \sin\theta$, in order for the model to be 1-loop-renormalizable. Then it turns out that $\beta_a= -{\lambda\over 2\pi}(4+a^2)$. After taking into account the redefinitions of the various parameters, as described in footnote 8 of the present paper, this implies eq. (47) of \cite{PLsfe3}.} This is a strong hint that their classical equivalence can be promoted into a quantum one as well. Having said that we note, once again, that $g=-1$ is not a fixed point of the \eqn{abg11} in the UV. Since \eqn{invv} is still a RG invariant of \eqn{abg2}, it is clear that one variable between $a$ and $b$ is an independent one. Eliminating $a$ from \eqn{abg2} using \eqn{invv}, we obtain \begin{equation} \beta_b = -{\lambda\over \pi} b \sqrt{(b_+-b)(b-b_-)}\ . \label{ab11} \end{equation} Hence, the solution for $b$ as a function of the energy scale $t=\ln \mu$ oscillates between $b_+$ and $b_-$ as \begin{equation} {1\over b(t)} = \nu + \sqrt{\nu^2 -1} \sin {\lambda \over \pi} (t-t_0)\ , \label{osscc} \end{equation} where $t_0$ is an arbitrary reference scale. This means that the corresponding $\s$-model actions do not define local field theories and can be considered at most as effective actions for scales such that $b$ stays away from $b_\pm$. The usual $S^2$ metric and its non-Abelian dual with respect to the right (or left) action of $SU(2)$ are obtained from \eqn{dsbth} and \eqn{dsbth1} if we rescale the coupling constant $\lambda\to \lambda/a$ and then take the limit $a\to \infty$ (keeping the new coupling finite). However, this limit is problematic at the quantum level since the corresponding $\beta$-functions do not tend to the beta-function obtained by studying the 2-dim field theories based on $S^2$ (and its non-Abelian dual) by themselves \cite{PLsfe3}. The latter is, at one-loop, just $\beta_\lambda = -{\lambda^2\over 2\pi}$ and is consistent with the fact that these models are asymptotically free. It is formally obtained by the first of \eqn{abg2} in the limit $a\to \infty$ after we rescale $\lambda \to \lambda/a$ as described above. This limit does not correspond to any fixed point of \eqn{abg2}. It is easily seen that, from a RG theory view point, these models offer an effective description of the more general models \eqn{dsbth} and \eqn{dsbth1} in the case of $\alpha\simeq b\simeq \nu\gg 1$, which, according to \eqn{osscc}, occurs at scales ${\lambda\over \pi} (t-t_0)\simeq -{\pi \over 2} + {1\over \nu}\ {\rm mod}(2\pi)$. \subsection{Models with fixed points} \subsubsection{Three-dimensional models} We have seen that our model \eqn{meett}, \eqn{btht} does not have a true fixed point under the 1-loop RG eqs. \eqn{abg11}. Consider, however, the analytic continuation $\lambda\to -i \lambda$ and $a\to i a$. Then the metric and antisymmetric tensors become \begin{equation} ds^2 = {a\over V} \left( (L^1)^2 +(L^2)^2 +(g+1)(L^3)^2 - {1+ g\over a^2} \big( (b \cos\theta -1)\ d\phi +(b-\cos\theta)\ d\psi\big)^2 \right) \ , \label{meett1} \end{equation} and \begin{equation} B= 2 i {\sin\theta \over V}\ d\theta\wedge \Big((g+1) \wedge d\psi + (b + g \cos\theta) \wedge d\phi\Big)\ , \label{btht1} \end{equation} where instead of \eqn{veo} the function $V$ is given by \begin{equation} V \equiv a^2 - (b-\cos\theta)^2 - (1+g)\sin^2\theta\ . \label{veo1} \end{equation} The fact that the antisymmetric tensor is imaginary is bothersome if we want to describe models in 2-dim Minkowskian space-times. However, for Euclidean ones, the (locally) exact 2-form measures the charge of non-trivial instanton-like configurations. The perturbative expansion is completely independent of the antisymmetric tensor, but this will definitely play a r\^ole in a, yet lacking, non-perturbative formulation of the model. The 1-loop RG equations for the metric \eqn{meett1} are obtained from \eqn{abg1} by the analytic continuation we have described above. Then the analogue of \eqn{abg11} is given by \begin{equation} \beta_g = - {\lambda\over \pi} {b \over a}\ (g+1) (b+\nu g)\ , \label{abg111} \end{equation} where now \begin{equation} a=\sqrt{(b - b_+)(b-b_-)}\geq 0\ ,\qquad b_\pm\equiv \nu\pm \sqrt{\nu^2-1}\ , \label{hja1} \end{equation} and $b$ is still given by \eqn{bgy}. As before, we will assume that $\nu>0$ with no loss of generality. However, now $\nu$ does not have to be larger than or equal to 1, as in \eqn{hja}, in order to ensure reality for $a$. If $\nu<1$ then $b_\pm$ are complex conjugate of each other and, unlike the case when they are real, $b$ can take any real value without spoiling the reality of the parameter $a$. However, now the condition $|1+1/g|\geq e^C \sqrt{1-\nu^2}$ has to be fulfilled in order for $b$ to remain real. If on the other hand $\nu>1$, then $b_\pm$ are both real and the reality condition for $a$ requires that $b\geq b_+>b_-$ or $b\leq b_-<b_+$. Since $0<b_-<b_+$, it turns out that there are fixed points for initial conditions where $b$ is less than $b_-$. Consider first the RG eq. \eqn{abg111} near the point with $g=1/(e^C-1)$, $b=0$ and $a=1$. It can be written as (we take the lower sign in \eqn{bgy}): \begin{equation} \beta_g \simeq {\lambda\over \pi} g^* (g-g^*)\ ,\qquad g^* \equiv 1/(e^C-1)\ . \label{bg1} \end{equation} The same equation near the different point with $g=-1/(e^C+1)$, $b=0$ and $a=1$ takes the form \begin{equation} \beta_g \simeq {\lambda\over \pi} \tilde g^* (g-\tilde g^*)\ ,\qquad \tilde g^*\equiv -1/(e^C+1)\ . \label{bg2} \end{equation} For $e^C>1$ we have $-{1\over 2} < \tilde g^* <0< g^* $. Hence, for $e^C>1$ we have an IR-stable point at $g=g^*$ as well as a UV-stable point at $g=\tilde g^*$. For $0<e^C<1$, we have that $g^*<-1<\tilde g^*<-{1\over 2}$. Therefore, for $0<e^C<1$ there are two UV-stable points at $g=g^* $ and at $g=\tilde g^*$. In all cases the background \eqn{meett1}, \eqn{btht1} flows, either in the IR or in the UV, towards the background with \begin{eqnarray} ds^2& = & -{1\over g_0} \left( {d\theta^2 \over \sin^2\theta } - g_0 d\phi^2 +(g_0+1) d\psi^2 \right)\ , \nonumber\\ B& = & -{2 i \over g_0} {1\over \sin\theta} d\theta \wedge \Big((g_0+1) d\psi + g_0 \cos\theta d\phi \Big)\ , \label{ddsf} \end{eqnarray} where $g_0$ represents any of the two fixed points $g^*$ or $\tilde g^*$. This represents a free theory, as can be seen by changing variables as $\sin\theta = {1\over \cosh y}$. It is interesting to note that in the case $e^C>1$ the signature of the metric in \eqn{ddsf} is $(-+-)$ in the IR fixed point $g_0=g^*$ and $(+++)$ in the UV fixed point $g_0=\tilde g^*$. Also in the case of $0<e^C<1$ the signature at the $g_0=g^*$ UV-stable point is $(++-)$, but in the other UV-stable point at $g_0=\tilde g^*$ it is $(+++)$. Hence, only at $g=\tilde g^*$ the metric has Euclidean signature and we expect a well-defined field-theoretical description. Let us also note that for $\nu\gg 1$ the RG flow is described, as before, by \eqn{abgli}, \eqn{asdj} and the corresponding $\s$-model is again \eqn{jef}, provided \eqn{sdsf} is satisfied. \subsubsection{Two-dimensional models} Now we turn to the 2-dim model \eqn{dsbth} after the same analytic continuation as before, $a\to i a$ and $\lambda\to -i \lambda$: \begin{eqnarray} ds^2 & = &{a \over a^2 -(b - \cos\theta)^2 } \ \left(d\theta^2 + \sin^2\theta d\phi^2\right) \ , \nonumber \\ B & = & 2 i {\sin\theta (b- \cos\theta) \over a^2 -(b-\cos\theta)^2 }\ d\theta \wedge d\phi \ . \label{dsbbh} \end{eqnarray} The 1-loop RG equation corresponding to \eqn{ab11} is \begin{equation} \beta_b = -{\lambda\over \pi} b \sqrt{(b-b_+)(b-b_-)}\ . \label{ac11} \end{equation} The form of the solution for $b$ as a function of the energy scale $t=\ln \mu$ depends on whether or not $\nu$ is smaller or larger than 1. We find \begin{equation} {1\over b(t)} = \nu + \sqrt{1-\nu^2} \sinh{\lambda \over \pi} (t-t_0)\ ,\quad {\rm if }\quad \nu<1\ ,\quad -{\pi\over 2\lambda} \ln\left({1+\nu\over 1-\nu}\right) \leq t-t_0 < \infty\ , \label{odcc} \end{equation} where $t_0$ denotes again an arbitrary reference scale. We see that in the UV there is a fixed point at $b=0$ (and $a=1$). The lower bound for $t$ above is needed for $b$ to stay positive, since only then is \eqn{odcc} a solution of \eqn{ac11}. For the case of $\nu>1$, we have to distinguish the solutions between those with $b\leq b_-$ and those with $b\geq b_+$. In the former case we obtain \begin{equation} {1\over b(t)} = \nu + \sqrt{\nu^2-1} \cosh{\lambda \over \pi} (t-t_0)\ ,\quad {\rm if }\quad \nu>1\ ,\quad t\geq t_0\ , \label{jqws} \end{equation} and \begin{equation} {1\over b(t)} = \nu - \sqrt{\nu^2-1} \cosh{\lambda \over \pi} (t-\tilde t_0)\ ,\quad {\rm if }\quad \nu>1\ ,\quad {\pi\over 2\lambda} \ln\left({\nu+1\over \nu-1}\right) \leq t-\tilde t_0 < \infty\ , \label{jqws1} \end{equation} where, as before, $t_0$ and $\tilde t_0$ are arbitrary reference scales. For the trajectory given by \eqn{jqws}, $b$ stays positive. It starts at $b=b_-$ for $t=t_0$, and ends at $b=0$ for $t\to \infty$. For the trajectory given by \eqn{jqws1}, $b$ is always negative and starts at $b=-\infty$ for $t-\tilde t_0={\pi\over 2\lambda} \ln\left({\nu+1\over \nu-1}\right)$ and ends at $b=0$ for $t\to \infty$. Hence, we see that $b=0$ is a UV fixed point. Also as we lower the scale $t$ towards the IR, the solution becomes singular in both cases. In any case, we then run into non-perturbative regimes. For trajectories in the region $b\geq b_+$, the solution is still given by \eqn{jqws1}, but with $-{\pi\over 2\lambda} \ln\left({\nu+1\over \nu-1}\right) \leq t-\tilde t_0 \leq 0$. In the lower limit $b\to \infty$ and in the upper limit $b=b_+$. Hence in that case we have a singular behaviour of the 1-loop RG equations towards the IR as well as the UV. As we have mentioned, in those cases the corresponding 2-dim field theory is not well defined at the quantum level and can be considered only as an effective field theory at scales away from the singularities. The 2-dim model corresponding to the fixed point at $b=0$ is obtained by setting $g_0=-1$ in \eqn{ddsf} \begin{eqnarray} ds^2 & = &{ 1 \over \sin^2\theta } \ \left(d\theta^2 + \sin^2\theta d\phi^2\right) \ , \nonumber \\ B & = & - 2 i \cot\theta\ d\theta \wedge d\phi \ . \label{dsbth11} \end{eqnarray} The fact that \eqn{dsbth} approaches a free-field conformal field theory at the fixed point is similar to the case of an integrable model (different from \eqn{dsbth}), representing also a 1-parameter deformation of $S^2$, that was considered in \cite{sausage}. It is interesting to investigate whether or not \eqn{dsbth} represents also an integrable perturbation of $S^2$. \section{Concluding remarks} We have constructed a new class of 2-dim field theories with target spaces corresponding to deformations of coset spaces $G/H$. Our models correspond to special points of the classical moduli space of models related by Poisson--Lie T-duality, where a local invariance develops. A classification of all possible models that arise with such a procedure is an interesting open problem and can be done by analyzing the general conditions \eqn{rel1}, or equivalently \eqn{kjd}. By construction these models come in dual pairs. The corresponding generating functionals depend non-polynomially on the derivatives of the fields with respect to the space-like variable. The latter feature is also manifested in an underlying infinite-dimensional algebra with a central extension of the parafermionic type. It would also be interesting to uncover the relation of our models to those in \cite{KliSevdressed}. We have also performed a quite general RG flow analysis using specific models with 3- and 2-dim target spaces. As in \cite{PLsfe3}, we conclude that quantum aspects of the lower dimensional models do not necessarily follow by taking the same classical limit as that used to relate the corresponding 2-dim field-theoretical classical actions. Concretely, the beta-function equations for the lower-dimensional models follow from those of the original models by just setting some parameters to their prescribed values (see \eqn{abg1} and \eqn{abg2}). However, these values do not necessarily correspond to any fixed points of the solutions of these equations. Using our 3-dim example we saw that in a large domain in parameter space, and for a wide range of energies in the UV, the description is effectively perturbative with a UV-fixed point exactly where the local gauge invariance develops. We believe that this feature will persist for more general models related by Poisson--Lie T-duality. In that respect it would be very interesting to study the RG flow in general using \eqn{action1} and \eqn{action2} and possibly to formulate this flow in a duality-invariant way. \bigskip\bigskip \centerline{\bf Acknowledgements} I would like to thank P. Forgacs for interesting discussions related to renormalization group flow in 2-dim field theories and for bringing \cite{sausage} into my attention. I also thank C. Zachos for an e-mail correspondence. \bigskip\bigskip \centerline{\bf Note added} I would like to thank C. Klimcik and the referee for informing me that the models constructed here have been studied before, at a purely classical level, in \cite{KliSevdressed}. The present paper provides an alternative, more natural to physicists, complementary view point on the classical origin of these models. Moreover, we have further elucidated their structure by providing explicit examples and exploring quantum aspects of the renormalization group flow.
\section{Introduction} \label{intro} In many physical phenomena percolation effects play an important role \cite{Perc}. Particularly, some dilute magnets are well described, in what concerns magnetic phase transitions, by uncorrelated diluted models \cite{Stinch}. In these models, magnetic sites on a lattice are randomly replaced by non-magnetic ones, and a bond links each pair of occupied (magnetic) first neighbours. At zero temperature, the problem is purely geometrical and is described in the following way. Sites on a lattice are randomly occupied with probability $p$, while all bonds are considered to be present. A cluster is then defined as a collection of present sites which are connected to each other through steps between occupied first-neighbours. As $p$ increases, an infinite cluster appears for the first time at a critical probability $p_{c}$, which is lattice dependent. In analogy with thermal critical phenomena, some quantities are singular at the critical point, following a power-law behavior near $p_{c}$ \cite{Perc}. Typical examples are the probability $P(p)$ that a site belongs to the infinite cluster, which behaves as $P(p) \sim (p-p_c)^{\beta}$ near $p_c$, and the correlation length $\xi(p)$, which diverges at $p_c$, according to $\xi(p) \sim |p-p_c|^{-\nu}$. Nevertheless, some systems were found to be better described by {\it correlated} percolation models, where the presence of sites (or bonds) depends also on their neighbourhood. Typical examples of correlated percolation models are bootstrap percolation \cite{Corr} and site-bond correlated percolation \cite{Rec}. The physical motivation for the introduction of the former model comes from diluted magnetic systems where competition between exchange (which favour a magnetic ground state) and crystal-field (which leads to a non-magnetic ground state) interactions takes place. To mimic this competition at zero temperature, the bootstrap percolation model was introduced \cite{cha,NSB}: in this model, sites on a lattice are randomly occupied with probability $p$ but only those with at least $m$ occupied first-neighbours remain occupied. In the stable final configuration, all occupied sites have at least $m$ occupied first-neighbours or the whole lattice is empty. As our purpose in this paper is to study one implementation of the bootstrap percolation model, we briefly review some of its properties in what follows (for a thourough discussion of the results up to 1990, see Reference \cite{Corr}). The $m=0$ case regains the usual (uncorrelated) site percolation model, where the transition is continuous and $p_c < 1$ in two and three dimensions. More specifically, the most precise evaluations of the critical exponents for three-dimensional lattices are $\nu=1/y_p=0.875 \pm 0.008$ and $\beta=0.412 \pm 0.010$ from simulation \cite{Ziff1} and $\nu=1/y_p=0.872 \pm 0.070$ and $\beta=0.405 \pm 0.025$ from series \cite{series}. In the $m=1$ case, only isolated sites are removed by the culling process: these sites do not contribute to the critical behaviour and $p_c$ and all critical exponents remain the same as for usual percolation. For $m=2$, on the other hand, isolated clusters with two sites are eliminated, as well as some dangling structures of more compact clusters. The elimination of these dangling structures, however, does not break the infinite cluster (whenever present) and, therefore, the critical probability is the same as for usual percolation. Moreover, as the exponent $\nu$ is also connected to the formation of this infinite cluster, its value is the same for $m=2$ and $m=0$. In what concerns ``field'' exponents ($\beta$, for instance), previous results for two-dimensional systems indicate a higher value of this exponent for $m=2$ than for usual percolation \cite{NSB2}. Nevertheless, it has been shown later, through simulations on bigger lattices in two dimensions \cite{belita} and general arguments applied to both two and three dimensions \cite{Branco}, that the exponent $\beta$ is the same for uncorrelated and $m=2$ bootstrap percolation models. Let us now turn our atention to higher values of $m$. It is generally believed that, for any value of $m$ where only infinite clusters can survive the culling process, $p_c=1$. This is indeed the case for $m=2d-1$ on hypercubic lattices ($d$ is the dimension of the lattice) \cite{robert}, for $m=4$ on cubic \cite{robert} and triangular \cite{Corr} lattices and for $m=5$ on the triangular lattice \cite{kogut}. Moreover, it has been shown that, for these cases, the usual finite-size scaling relation does not hold. This finite-size scaling predicts that, if a suitable definition of a finite-lattice ``critical'' point, $p_{av}$, is made, this point will approach the critical point in the thermodynamic limit, $p_c$, as: \begin{equation} p_c - p_{av} \sim L^{-1/\nu}, \label{fss1} \end{equation} where $L$ is the linear size of the finite system, such that $L \gg 1$, and $\nu$ is the usual critical exponent. This finite-size behaviour is indeed observed for $m \leq 2$, but fails for high values of $m$. For $m=2d-1$ on hypercubic lattices, it has been proven that $p_c - p_{av} \sim 1/(\log^{d-1}L)$ \cite{aizenman}, e.g., proportional to $1/\log(\log L)$ for $d=3$ dimensions. Also for $m=4$ on the cubic lattice, the correct finite-size behaviour is $p_c - p_{av} \sim 1/\log (\log L)$, with $p_c=1$ \cite{cerf}. These results for high $m$ have been conjectured or tested in numerical simulations \cite{vanenter}. The results stated in the previous paragraphs do not apply to the $m=3$ case on the cubic lattice. For this model, one expects $p_c$ to be above the value for uncorrelated percolation (where $p_c=0.311605 \pm 0.000005$ \cite{Acha}), since the infinite cluster for usual percolation at $p_c$ is not stable with respect to the culling process for $m=3$. On the other hand, since finite clusters are still stable for this value of $m$ on the cubic lattice, it is expected that $p_c<1$. Numerical simulations confirmed this scenario \cite{kogut,as}, although the relative small sizes used in those works indicate that the values might not be precise (we will return to this point later). In what concerns the critical exponents, previous results indicate that the exponent $\nu$ is the same as for usual percolation \cite{as} but $\beta$ is higher than its uncorrelated counterpart \cite{kogut,as}. However, in neither of these works a extrapolation to the thermodynamic limit is attempted, leaving the possibility that finite-size effects might be the reason for the discrepancy in the values of $\beta$. This possibility was first proven to be right, in the context of two-dimensional bootstrap percolation models, in References \cite{belita} and later confirmed for $m=3$ on the triangular lattice \cite{arcangelis}. The possibility of a new universality class for bootsrap percolation with $m=3$ on the cubic lattice is the problem we address in this work. We resort to numerical simulation methods, which, together with finite-size scaling analysis, allowed us to obtain more precise values for the critical parameters. We also study the so-called critical spanning probability, i.e., the probability that a given lattice has a cluster connecting its boundaries at criticality \cite{langlands,Ziff2}. This quantity shows some degree of universality: it depends on the dimension and shape of the system and on the specific boundary condition but not on the lattice type (simple cubic of f.c.c., for instance) and on the particular kind of percolation (site or bond) \cite{Acha,Ziff2,Lin}. It is then interesting to see whether it remains invariant for percolation models such that long-range correlation is involved, like bootstrap percolation. The remainder of the paper is organized as follows. In the next section we present the method and discuss some technical details, as well as the results for the critical parameters. In Section \ref{csp} we discuss some previous results concerning the critical spanning probability for percolation and our results for bootstrap percolation. Finally, we summarize our results in the last section. \section{Method and Results} \label{mr} The method we use is connected to real-space renormalization-group and finite-size scaling procedures \cite{MCRG}. The approach needs that precise values of the physical quantities are available; these are only obtained for high values of the linear system size $L$. We then studied finite systems of size $L^3$, with $32 \leq L \leq 480$; from the results for $L \gg 1$, it is possible to use finite-size scaling techniques to extrapolate to the thermodynamic limit ($L=\infty$). The critical probability $p_c$ and the critical exponent $\nu$ are calculated as follows. For a lattice of size $L$, we occupy each site with probability $p$, apply the bootstrap condition and test the lattice for percolation (here we define percolation as the presence of a cluster which connects the bottom and top planes of the finite cubic lattice; we discuss this and other technical points below). Our finite-size estimate of the critical probability, $p^*$, is taken as the value of $p$ at which the cell percolates for the first time, when $p$ is increased from zero; this procedure is made for ${\cal N}$ different runs (which correspond to ${\cal N}$ different seeds to the random number generator). Each run leads to a different value of $p^*$, since the lattice is finite. We take the average of the ${\cal N}$ values of $p^*$ as our estimate of $p_{av}$ (see Section \ref{intro}). It is then assumed that $p_{av}$ will approach $p_c$ as given by Equation \ref{fss1}. Moreover, it is possible to calculate the width $\sigma = \sqrt{<p^{*^2}> - p_{av}^2}$, which behaves as ($<p^{*^2}>$ stands for the average of $p^{*^2}$ over the ${\cal N}$ realizations): \begin{equation} \sigma \sim L^{-1/\nu}. \label{fss2} \end{equation} From the previous equation and a $\log-\log$ plot of $\sigma \times L$, it is then possible to obtain the value of the critical exponent $\nu$. Tha data is depicted in Figure \ref{sig}; from the slope of the straight line we obtain $\nu=0.89 \pm 0.04$. We compare this value with other evaluations in Table \ref{crit}; within the numerical precision, this exponent is the same for bootstrap percolation with $m=3$ and ordinary percolation on the cubic lattice and agrees with previous evaluations of $\nu$ for bootstrap percolation with $m=3$ on the cubic lattice. Note from the graph that the straight line regime is achieved for $L \geq 128$, while for uncorrelated percolation this regime sets in for smaller values of $L$. This is expected, since finite-size effects are stronger for correlated percolation problems than for their uncorrelated counterparts \cite{NSB,NSB3}. Therefore, we neglect the data for $L \leq 96$ and used only lattices with $128 \leq L \leq 480$ in our linear regression analysis. \begin{figure} \epsfxsize=9.5cm \begin{center} \epsffile{branco1.eps} \caption{Log-log plot $\sigma \times L$, for $32 \leq L \leq480$. The linear size $L$ is always a multiple of 32. The dotted line is a linear fit of the data {\it for $128 \leq L \leq 480$ only} (see text).} \label{sig} \end{center} \end{figure} In order to calculate the critical threshold, we resort to Equations \ref{fss1} and \ref{fss2}, which imply that, for $L \gg 1$: \begin{equation} p_c - p_{av} \sim \sigma . \label{fss3} \end{equation} This is a convenient way to calculate $p_c$, since it does not depend on the value of $\nu$; such dependence would appear if Equation \ref{fss1} was used. It is shown in Figure \ref{sigma} a plot of $p_{av} \times \sigma$: $p_c$ is given by the linear coeficient and the value obtained is $p_c=0.57256 \pm 0.00006$. This value is slightly above the one calculated in Reference \cite{as}, in which the value of $L$ varied from 10 to 110. If we use the same range in our calculation the extrapolated value of $p_c$ is consistent with the result of \cite{as} (see Table \ref{crit}). \begin{figure} \epsfxsize=9.5cm \begin{center} \leavevmode \epsffile{branco2.eps} \caption{Graph of $p_{av} \times \sigma$, for $32 \leq L \leq480$. The linear size $L$ is always a multiple of 32. The dotted line is a linear fit of the data {\it for $128 \leq L \leq 480$ only} (see text). Although not depicted in the figure, there are error bars in the values of $\sigma$, which were taken into account in the evaluation of $p_c$.} \label{sigma} \end{center} \end{figure} Let us know discuss some technical points. An important quantity is the value of $p$ at which the lattice percolates for the first time, when $p$ is increased from zero. One needs to define what ``percolate'' means for a finite lattice. In this work, we used the rule called ${\cal R}_1$ in Reference \cite{MCRG}, i.e., a lattice percolates if, after the bootstrap culling process, there is a path of present sites which links the boundaries of the lattice in a fixed direction (vertical, say). There are other possible definitions (see \cite{MCRG}) and it is expected that all of them lead to the same value of $p_c$ in the thermodynamic limit. Note, however, that the value of the critical spanning probability does depend on this definition, as we will discuss im Section \ref{csp}. The numerical procedure we used to test for percolation is the Hoshen-Kopelman algorithm \cite{HK}: for usual percolation, it requires the storage of only one plane. For bootstrap percolation, on the other hand, the bootstrap iteration needs the storage of the whole lattice, due to correlation effects. To cope with this drawback, we store the lattice in bits, instead of words: this saves memory and time, since the updates connected to the bootstrap rule can be made in parallel for a set of 32 sites \cite{stauffer}. To define all six first-neighbours of sites at the boundaries of the lattice, periodic boundary conditions are used. It is expected that the boundary condition does not affect the values of the critical parameters in the thermodynamic limit, since it is a ``surface'' effect. Finally, let us mention that the number of realizations ${\cal N}$ (see Section \ref{intro}) varied from 12000 for $L=32$ to 640 for $L=480$; the errors were calculated as three times the standard deviation for subsets of the total number of realizations. \begin{table} \caption{Critical parameters for $m=3$ bootstrap percolation and comparison with previous results for this model and for usual percolation.} \begin{tabular}{|l|l|l|} \hline & m=3 bootstrap percolation & usual percolation \\ \hline\hline $p_{c}$ & 0.57256$\pm$0.00006$^a$ & \\ & 0.5717$\pm$0.0005$^b$ & 0.311605$\pm$0.000005$^d$ \\ & 0.568$\pm$0.002$^c$ & \\ \hline $\nu$ & 0.89$\pm$0.04$^a$ & 0.875$\pm$0.008$^e$ \\ & 0.876$\pm$0.010$^b$ & 0.872$\pm$0.070$^f$ \\ \hline $\beta$ & 0.37$\pm$0.03$^a$ & 0.41$\pm$0.02$^e$ \\ & 0.6$\pm$0.1$^b$ & 0.40$\pm$0.06$^f$ \\ & 0.82$\pm$0.04$^c$ & \\ \hline \end{tabular} \\ $^a$ Present work. $^b$ Reference \cite{as}. $^c$ Reference \cite{kogut}.\\ $^d$ Reference \cite{Acha}. $^e$ Reference \cite{Ziff1}. $f$ Reference \cite{series} \label{crit} \end{table} To have access to the ``magnetic'' scaling power $y_{h}$, one possible way is to define a ``ghost''- site, which is linked to all sites of the lattice with probability $h$ \cite{MCRG}. Within a real-space renormalization group framework, it is possible to calculate the eigenvalue $\lambda_{h}$ through \mbox{$\lambda_{h}=<n>/p_{c}$}, where $<n>$ is the average number of occupied sites linked to one of the boundary planes of the finite cubic lattice at the critical threshold, $p=p_c$. The value of $p_c$ is obtained as explained above and is then used to calculate $<n>$ and hence $\lambda_{h}(L)$, averaging over ${\cal M}$ configurations. The value of ${\cal M}$ varied from $\sim 320000$, for the smaller lattices, to $\sim 20000$, for the bigger ones. We use two procedures to obtain $y_h$ in the thermodynamic limit. The first one is based on the fact that, for $L \rightarrow \infty$: \begin{equation} \lambda_h(L) = L^{y_h}, \label{fss4} \end{equation} with $y_h$ independent of $L$; this equation leads to a straight line in a $\log-\log$ plot of $\lambda_h \times L$, for $L \gg 1$. As depicted in Figure \ref{yh}, this is indeed the case for $128 \leq L \leq 480$. From a linear fitting of the data for this range of $L$, and resorting to the scaling relation $\beta = (d-y_h) \nu$, we obtained the value $\beta= 0.37 \pm 0.03$ (see Table \ref{crit}). Alternatively, we could take into account correction-to-scaling terms, through: \begin{equation} \lambda_h = L^{y_h} \left( 1 + B/L \right). \label{fss5} \end{equation} From this equation, we see that local slopes of $\log \lambda_h \times \log L$ provide estimates of $y_h(L)$, which, when extrapolated to $L\rightarrow \infty$, leads to an evaluation of $y_h$ in the thermodynamic limit. We have applied this procedure for $32 \leq L \leq 480$, using three consecutive points to calulate the local slopes and extrapolating to $L \rightarrow \infty$ through a $y_h(L) \times 1/L$ graph. The value thus obtained for $\beta$ is the same as for the first procedure. In Table \ref{crit} we see that our estimate of $\beta$ disagrees with the values calculated in References \cite{kogut} and \cite{as}. In evaluating this exponent, the former uses lattices of linear size $L=35$, while the latter uses $L=80$; neither attempted to make an extrapolation to the thermodynamic limit. From Table \ref{crit}, we can infer that our estimate of $\beta$ for the bootstrap model we study is, within the error bars, equal to the corresponding value of this exponent for usual (uncorrelated) percolation. Since we have already seen that $\nu$ is also the same for both models, we can draw the conclusion that usual percolation and bootstrap percolation with $m=3$ on the cubic lattice belong to the same universality class. This result contradicts References \cite{kogut} and \cite{as}; we believe this is caused by the small lattices used in those works. \begin{figure} \epsfxsize=9.5cm \begin{center} \leavevmode \epsffile{branco3.eps} \caption{Graph of $\log(\lambda_h(L)) \times \log(L)$, for $32 \leq L \leq480$. The linear size $L$ is always a multiple of 32. The dotted line is a linear fit of the data {\it for $128 \leq L \leq 480$ only} (see text) and $y_h$ is the angular coefficient of this line. The error bars are smaller than the data points.} \label{yh} \end{center} \end{figure} \section{Critical spanning probability} \label{csp} It has been established some time ago that the critical spanning probability, $R(p_c)$, defined as the probability of spanning a lattice at the critical point, shows some degree of universality \cite{langlands,cardy}. More precisely, this quantity does not depend on the lattice type and on the kind of percolation. This result contradicts the assumption made on early applications of real-space renormalization group procedures to percolation. In those, it was assumed that the critical spanning probability is equal to the critical percolation threshold, $p_c$ and, therefore, lattice dependent \cite{MCRG}. However, later numerical tests confirmed and expanded the universality proposal \cite{Acha,Ziff2,Lin}. Nevertheles, no study has been made on correlated models, to the best of our knowledge. While it is expected that {\it short-range} correlations do not change the universality scenario \cite{NSB}, it is not clear whether $R(p_c)$ changes when {\it long-range} correlations are introduced. A convenient model to test these possibilities is the bootstrap percolation one. While the behaviour of $R(p_c)$ is trivial for the cases where $p_c=1$, for $m=3$ on the simple cubic lattice the presence of correlation may lead to a non trivial behavior. We study this possibility using numerical simulation on simple cubic lattices of size $L^3$, with $32 \leq L \leq 480$. The programs and algorithms used are essentially the same as the ones described in the previous section. The basic procedure is to generate ${\cal M}$ independent runs for each lattice size and compute the fraction of those which percolate after the bootstrap condition is used and a stable configuration is reached. The values of ${\cal M}$ are the same as those used in the calculation of $y_h$ (see previous section). The results are depicted in Figure \ref{rpc}: we can infer that $R(p_c)=0.270 \pm 0.005$, where the error bar is a rough estimate. There are two previous calculation of $R(p_c)$ for the uncorrelated site percolation on the cubic lattice with free boundary conditions: they lead to $R(p_c)=0.265 \pm 0.005$ \cite{Lin} and $R(p_c)=0.28$ \cite{Acha}. Our value agrees, within the numerical error, with the first one but the result of Reference \cite{Acha} cannot be ruled out. It is then reasonable to infer that usual percolation and $m=3$ bootstrap percolation on the cubic lattice belong to the same universality class, also in what regards the critical spanning probability. \begin{figure} \epsfxsize=9.5cm \begin{center} \leavevmode \epsffile{branco4.eps} \caption{Graph of $R(p_c) \times \log(L)$, for $32 \leq L \leq 480$. The linear size $L$ is always a multiple of 32. The dotted line is a guide to the eye and indicates the value 0.270, taken as $R(p_c)$ in the limit $L \rightarrow \infty$.} \label{rpc} \end{center} \end{figure} \section{Summary} \label{summ} We calculate, using numerical simulations and finite-size scaling techniques, the critical parameters of the bootstrap percolation model with $m=3$ on the cubic lattice, using finite lattices of size $L^3$, with $32 \leq L \leq 480$. Our evaluations of $\nu$ and $\beta$ strongly support the conclusion that usual percolation and the model we study belong to the same universality class. This result disagrees with previous calculations \cite{kogut,as}; we believe this is due to the small sizes used in those works. To support this assumption, we note that the finite-size scaling assumptions hold for lattices of linear size $L \geq 128$, which are higher then the sizes studied in previous works. The critical spanning probability, $R(p_c)$, is also calculated. It has been shown that this quantity shows some degree of universality, but, to the best of our knowledge, no study concerning correlated percolation models have been done so far. Our result for bootstrap percolation with $m=3$ on the cubic lattice, $R(p_c)=0.270 \pm 0.005$, is, within the numerical accuracy, the same value as for usual percolation with free boundary conditions. Therefore, we can infer that $R(p_c)$ is not sensitive to short range correlations and even to some long range correlations, like the one studied in this paper. \vskip 0.7cm We would like to thank Dr. D. Stauffer for fruitful discussions at early stages of this work and for a critical reading of the manuscript.
\section{Introduction} \label{sec:intro} Although the Wisconsin H-Alpha Mapper (WHAM) Sky Survey will present the first view of the distribution and kinematics of the Warm Ionized Medium (WIM), we still know little about the physical conditions in this gas or how they are produced. Several studies have measured constraints on important parameters toward specific regions of the WIM, but none has yet provided a large sample of measurements to help test how environmental conditions (\emph{e.g.}\ distance from \ion{H}{2}\ regions, distance from the Galactic plane, etc.) might affect these physical conditions. Reynolds (1985a\markcite{rjr85a}) derived the first estimate of the temperature of the WIM by combining measured line widths of [\ion{S}{2}] and H$\alpha$\ emission toward the same regions. He was able to place limits between 5,000 K and 20,000 K, with a mean of about 8,000 K, on the temperature of a sample of 21 WIM observations. Ionization conditions have been probed even less, with most of our current information coming from observations in the plane of the Galaxy. Reynolds (1985b\markcite{rjr85b}) showed that [\ion{O}{3}] $\lambda 5007$ emission is very faint in the WIM, placing limits on the amount of O$^{++}$ (needing $h\nu > 35.1$ eV) in the WIM. More conclusively, Reynolds \& Tufte (1995\markcite{rt95}) and Tufte (1997\markcite{t97}) have shown via observations of the He I $\lambda 5876$ recombination line that the fraction of He$^{+}$ (needing $h\nu > 24.6$ eV) in this gas is also quite low in comparison with traditional O star \ion{H}{2}\ regions. The recent detection of [\ion{O}{1}] $\lambda 6300$ from three WIM directions (Reynolds, et al.\ 1998\markcite{rhth98}) places further limits on the ionizing spectrum. The low [\ion{O}{1}]/H$\alpha$\ ratios seen in all three directions suggests that the H is mostly H$^{+}$ in the H$\alpha$\ emitting gas. The advent of WHAM provides a unique opportunity to trace some of these diagnostics over a larger region of the sky. This paper presents the first velocity-resolved study of [\ion{S}{2}] and [\ion{N}{2}] from the diffuse gas of our Galaxy. A region of the sky that includes the Perseus arm was chosen for examination since the H$\alpha$\ emission from the WIM in this arm is well separated in velocity from local WIM emission and can be traced to high Galactic latitude, allowing an exploration of the emission with distance, $z$, from the Galactic plane. Such observations can be directly compared to studies of edge-on galaxies, most notably those by Rand (1997\markcite{rand97}, 1998\markcite{rand98}) of NGC 891. The primary trends found in his work are borne out here and extended to even fainter emission measures. Specifically, [\ion{S}{2}]/H$\alpha$\ and [\ion{N}{2}]/H$\alpha$\ ratios increase with $|z|$, while [\ion{S}{2}]/[\ion{N}{2}] ratios remain nearly constant. A brief description of the instrument and observations are presented in \S\ref{sec:obs}. The results from the data, including a measurement of the electron scale height in the Perseus arm and an investigation of the trends in line ratios, are presented in \S\ref{sec:results}. In \S\ref{sec:discussion}, we present our interpretation that these line ratios provide direct measurements of the temperature and ionization state of the WIM. Finally, our conclusions are summarized in \S\ref{sec:summary}. \section{Observations} \label{sec:obs} \subsection{The WHAM Instrument} \label{sec:wham} Due to the extended nature of the WIM, large-aperture Fabry-Perot detection techniques have proven to be quite successful in studying the WIM (Reynolds et al.\ 1990\markcite{rrsh90}). For WHAM, a 6-inch, dual-etalon Fabry-Perot spectrometer delivers the high-efficiency, high-resolution observations (Tufte 1997\markcite{t97}; Reynolds et al.\ 1998\markcite{rthjp98}). The dual-etalon design provides superior order rejection over single-etalon systems to achieve a much larger free spectral range while maintaining high spectral resolution (12 km~s$^{-1}$). An all-sky siderostat housing a 0.6-meter lens guides light to the etalon chambers in an environmentally controlled trailer. The center wavelength of the 200 km~s$^{-1}$\ spectral window is selected by changing the pressure of a high index of refraction gas (SF$_6$) in sealed etalon chambers. A narrow-band ($\sim 20$ \AA\ FWHM) interference filter provides additional spectral isolation near the desired wavelength. The resulting spectrum, in the form of a Fabry-Perot ring pattern, is imaged onto a cryogenically cooled Tek 1024$\times$1024 CCD. The spectrum covers a 200 km~s$^{-1}$\ interval at 12 km~s$^{-1}$\ spectral resolution. Coatings on the etalons and optics allow the placement of the 200 km~s$^{-1}$\ wide spectral window anywhere between 4800 and 7300 \AA. In this ``spectral mode,'' the image contains only the average spectrum within the one-degree diameter circular beam, avoiding confusion between spectral features and angular structures of the source (or stars) within the beam. WHAM also has a set of optics that can be placed in the post-etalon beam to provide an extremely narrow-band (adjustable between 12-200 km~s$^{-1}$), 1\arcmin\ angular resolution \emph{image} of the sky within a one-degree field of view. Only the spectral mode (with one-degree angular resolution) is used for the observations presented here. Further details on the WHAM spectrometer and its calibration can be found in Tufte (1997\markcite{t97}). WHAM is currently located at Kitt Peak National Observatory. Using the WIYN messaging system designed by Percival (1994\markcite{p94}, 1995\markcite{p95}) and custom hardware designed by the {\em Space Astronomy Lab} at Wisconsin, WHAM is a completely remotely operated telescope system. All functions of the siderostat and spectrometer are routinely operated from our Wisconsin offices. Details on the hardware and software used to operate WHAM and to automate the survey can be found in Haffner (1999\markcite{h99}). \subsection{The Data} \label{sec:data} The H$\alpha$\ observations of the 1100 deg$^2$ region $\ell = 123\arcdeg$ to $164\arcdeg$, $b = -6\arcdeg$ to $-35\arcdeg$ that is the focus of this study were obtained between 1997 July and 1997 October as a part of the WHAM H$\alpha$\ Sky Survey (Haffner 1999\markcite{h99}). Additional observations at the same survey grid points but with the 200 km~s$^{-1}$\ spectral window tuned near the [\ion{S}{2}] $\lambda 6716$ and [\ion{N}{2}] $\lambda 6583$ lines were obtained in 1997 October, 1997 November, and 1998 June. The region presented here is sampled by 1360 H$\alpha$, 1360 [\ion{S}{2}], and 1360 [\ion{N}{2}] spectra, each representing the integrated emission within WHAM's one-degree beam. The integration time for each spectrum was 30 s for H$\alpha$\ and 60 s for the other two emission lines. The [\ion{S}{2}] and [\ion{N}{2}] observations were converted into spectra using the same steps as those for H$\alpha$\ presented in Haffner (1999\markcite{h99}). Separate white-light flat fields were created for each of these additional filters and used in place of the H$\alpha$\ flat field. There are no strong terrestrial components in [\ion{S}{2}] or [\ion{N}{2}] comparable to the geocoronal emission in the H$\alpha$\ spectra. For the [\ion{S}{2}] observations, the weak terrestrial (Tucson lights) Ne I line at 6717.04 \AA\ was used for velocity calibration. For the [\ion{N}{2}] observations, spectra of H$\alpha$\ and [\ion{N}{2}] from bright emission nebulae were compared and the velocity frame of [\ion{N}{2}] was adjusted so that the mean velocity of the nebular [\ion{N}{2}] emission matched that of the H$\alpha$. By using the terrestrial lines in the H$\alpha$\ spectra (geocoronal H$\alpha$) and [\ion{S}{2}] spectra (Ne I), the velocity scale in each spectrum is calibrated to better than 1 km~s$^{-1}$. Intensities were calibrated by applying a correction factor to the [\ion{S}{2}] and [\ion{N}{2}] observations (0.94 and 1.15, respectively) based on the transmission differences between these filters and the H$\alpha$\ filter. As in the H$\alpha$\ observations, the [\ion{S}{2}] and [\ion{N}{2}] spectra are contaminated by weak atmospheric lines (Haffner, Reynolds, \& Tufte 1998\markcite{hrt98}; Haffner 1999\markcite{h99}). The brightest of these in our [\ion{S}{2}] spectra is the Ne I line located $+26.8$ km~s$^{-1}$\ from the geocentric zero of [\ion{S}{2}] $\lambda6716$. As noted above, it proves to be a useful wavelength calibrator. To accurately characterize the parameters of these faint lines in the [\ion{S}{2}] and [\ion{N}{2}] spectra, we followed the same strategy outlined in Haffner (1999\markcite{h99}) for H$\alpha$. We first created an average spectrum for each survey ``block'' ($\approx49$ pointings taken sequentially). Using this high signal-to-noise spectra, a single Gaussian was fitted to each of the atmospheric lines and then subtracted from the individual pointings. Such corrections account for hourly variations in these faint lines (typically $\sim$0.02 R). A list of the atmospheric lines in our [\ion{S}{2}] and [\ion{N}{2}] observations is given in Table~\ref{tab:atlines}. The potential atmospheric line at $-40$ km~s$^{-1}$\ in [\ion{N}{2}] is coincident with interstellar emission from the Perseus arm in all spectra presented here. The implications of such a line are discussed in \S\ref{sec:discussion}. \section{Results} \label{sec:results} \subsection{Overview} \label{sec:overview} Figure~\ref{fig:spectra} shows samples of H$\alpha$, \ion{N}{2}, and \ion{S}{2}\ spectra from the region under study. The Local (centered near 0 km~s$^{-1}$) and Perseus arm (centered near $-40$ to $-60$ km~s$^{-1}$) emission are shown as well-separated components in the right-hand panels, where $\ell < 150\arcdeg$, while Local gas dominates in the left panels ($\ell > 150\arcdeg$). Figures~\ref{fig:loc-lr}a and \ref{fig:per-lr}a display H$\alpha$\ intensity maps integrated over $\ensuremath{v_{\rm LSR}} = -10$ to $+10$ km~s$^{-1}$\ (Figure~\ref{fig:loc-lr}) and $\ensuremath{v_{\rm LSR}} = -30$ to $-50$ km~s$^{-1}$\ (Figure~\ref{fig:per-lr}), sampling the Local and Perseus arms, respectively. The boxes overplotted in them denote the $10\arcdeg \times 12\arcdeg$ region previously mapped in H$\alpha$\ by Reynolds (1980\markcite{rjr80}) and compared to \ion{H}{1}\ observations in detail by Reynolds, et al.\ (1995\markcite{rtkmh95}). The [\ion{S}{2}] and [\ion{N}{2}] maps look qualitatively similar to H$\alpha$\ and are not reproduced here. The line ratios images [\ion{N}{2}]/H$\alpha$\ (Figures~\ref{fig:loc-lr}b and \ref{fig:per-lr}b), [\ion{S}{2}]/H$\alpha$\ (Figures~\ref{fig:loc-lr}c and \ref{fig:per-lr}c), and [\ion{S}{2}]/[\ion{N}{2}] (Figures~\ref{fig:loc-lr}d and \ref{fig:per-lr}d) are discussed below in \S\ref{sec:lr}. Two large H$\alpha$\ emitting regions dominate the emission at lower latitudes, above $b = -20\arcdeg$. The brightest emission arises from NGC 1499 (California Nebula) excited by the O7.5I star $\xi$ Per located $540\pma{330}{150}$ pc away (\emph{Hipparcos}; Perryman 1997\markcite{hipparcos}) at $\ell = 160\fdg37$, $b = -13\fdg11$. Emission in the nebula itself, a $5\arcdeg \times 2\arcdeg$ crescent centered near $\ell = 160\arcdeg$, $b = -13\arcdeg$, peaks at over 250 R (1 R $= 10^6/4\pi$ ph cm$^{-2}$ s$^{-1}$ sr$^{-1} = 2.4\times10^{-7}$ erg cm$^{-2}$ s$^{-1}$ sr$^{-1}$ at H$\alpha$) and is centered around $\ensuremath{v_{\rm LSR}} = -5$ km~s$^{-1}$, while faint, extended emission above 5 R centered near $\ensuremath{v_{\rm LSR}} = +5$ km~s$^{-1}$\ continues 5--7$\arcdeg$ away from the nebula. This extended \ion{H}{2}\ region has been studied in some detail by Reynolds (1988\markcite{rjr88}), who first hypothesized that the bar of emission ($I_{\mathrm{H}\alpha} \approx 8$ R) running from $\ell = 150\arcdeg$, $b = -20\arcdeg$ to $\ell = 142\arcdeg$, $b = -10\arcdeg$ centered between $\ensuremath{v_{\rm LSR}} = -5$ to $+5$ km~s$^{-1}$\ may also be ionized by the star. The second large emission region, centered near $\ell = 131\arcdeg$, $b = -11\arcdeg$, is a superposition of separate but equally bright emission regions in the Local and Perseus arms (see Figure~\ref{fig:spectra}). Emission from the Local arm is near $\ensuremath{v_{\rm LSR}} = 0$ km~s$^{-1}$\ and has a more circular appearance reminiscent of a classical Str\"{o}mgren sphere (see Figure~\ref{fig:loc-lr}a). This faint \ion{H}{2}\ region has a relatively constant brightness of 5--6 R over most of its $13\arcdeg$ diameter area. The most promising candidate for an ionizing source for this region is $\phi$ Per, a B0.5e + sdO binary system located at $\ell = 131\fdg3$, $b = -11\fdg3$ and $220\pma{43}{32}$ pc away (\emph{Hipparcos}; Perryman 1997\markcite{hipparcos}; Gies et al.\ 1998\markcite{gea98}). At that distance, the region is 50 pc in diameter. Some details of this region can be found in Haffner (1999\markcite{h99}). A future study (Haffner, Reynolds, \& Tufte 1999\markcite{hrt99}) examining the properties of B-star \ion{H}{2}\ regions in the WHAM H$\alpha$\ Sky Survey will expand on this work. Emission from the Perseus arm in this direction at $-60 < \ensuremath{v_{\rm LSR}} < -40$ is much more amorphous in shape (Figure~\ref{fig:per-lr}a). At higher latitudes and at velocities farther from the LSR, the H$\alpha$\ emission becomes quite filamentary. Some of the more notable structures include a faint bar in Figure~\ref{fig:per-lr}a running from $\ell = 150\arcdeg$, $b = -33\arcdeg$ to $\ell = 160\arcdeg$, $b = -25\arcdeg$; several, nearly vertical filaments near $\ell = 130\arcdeg$, $b = -20\arcdeg$; and the shell-like region and associated dissecting filament (more visible at higher velocities; see Haffner 1999\markcite{h99}) centered on the box in Figure~\ref{fig:per-lr}a. The last of these was previously studied in detail by Ogden \& Reynolds (1985\markcite{or85}) and Reynolds et al.\ (1995\markcite{rtkmh95}). \subsection{Scale Height of the WIM} \label{sec:hvsb} Since the Perseus arm gas is well separated in velocity from the Local emission, we can explore the decrease of this emission as a function of the height above the Galactic plane. To extract information about the scale height from the intensity distribution, we first assume that the gas density is primarily a function of $z$ and has the form \begin{equation} \label{eq:gasexp} n_e(z) = n_e^0\;e^{-|z|/H}\;\mathrm{cm}^{-3}, \end{equation} where $n_e^0$ is the mid-plane density and $H$ is the scale-hight of the WIM. The H$\alpha$\ intensity is related to the gas density by the equation \begin{equation} \label{eq:itoem} 2.75\;T_4^{0.9}\,\ensuremath{I_{\mathrm{H}\alpha}} = \int \phi\;n_e^2\;dl, \end{equation} where $\phi$ is the filling fraction of the emitting gas, $T_4$ is the temperature of the gas in $10^4$ K, and $\ensuremath{I_{\mathrm{H}\alpha}}$ is in R. As long as the temperature, filling factor, and pathlength through the emitting gas, $\int dl = L$, are not functions of $z$ we then have \begin{equation} \label{eq:ivsz} \ensuremath{I_{\mathrm{H}\alpha}} = \frac{\phi\,(n_e^0)^2\,L}{2.75\,T_4^{0.9}}\;e^{-2|z|/H} = \ensuremath{I_{\mathrm{H}\alpha}}^0\;e^{-2|z|/H}, \end{equation} where $\ensuremath{I_{\mathrm{H}\alpha}}^0$ is the intensity at the mid-plane. Finally, substituting $z = D\,\tan |b|$, where $D$ is the distance to the arm and $b$ is the Galactic latitude of an observation, and taking the natural log, we arrive at \begin{equation} \label{eq:ivstanb} \ln \ensuremath{I_{\mathrm{H}\alpha}} = \ln \ensuremath{I_{\mathrm{H}\alpha}}^0 - \frac{2D}{H} \tan |b|. \end{equation} Figure~\ref{fig:havstanb} plots this relationship for a selected region of the map. Each point represents the median emission between $\ell=125\arcdeg$ and $152\arcdeg$ for each Galactic latitude. The median of the emission in each latitude slice is used instead of an average to reduce the effect of local emission enhancements (\emph{e.g.}\ faint \ion{H}{2}\ regions). The limited longitude range is chosen to avoid the gradual blending of the Perseus arm component with the Local gas as the longitude approaches $180\arcdeg$. Three radial velocity intervals are presented in the plot, reflecting total ($\ensuremath{v_{\rm LSR}} = -100$ to $+100$ km~s$^{-1}$), Local ($\ensuremath{v_{\rm LSR}} = -25$ to $+100$ km~s$^{-1}$), and Perseus arm ($\ensuremath{v_{\rm LSR}} = -100$ to $-25$ km~s$^{-1}$) emission. The integration ranges are expanded from the limited ranges in Figures~\ref{fig:loc-lr} and \ref{fig:per-lr} to include the full emission profile in our calculations. The error bars for the Perseus arm are an estimate of the physical scatter along a latitude slice calculated from the average deviation about the median of emission within that slice. Also displayed in Figure~\ref{fig:havstanb} is a linear fit to Perseus arm points that are above $b = -15\arcdeg$. Closer to the plane, \ion{H}{2}\ regions and extinction play a significant role and these points are excluded from the fit. The resulting fit parameters are $\ln \ensuremath{I_{\mathrm{H}\alpha}}^0 = 1.73\pm0.19$ and $2D/H = 4.91\pm0.39$, leading to a mid-plane intensity for the WIM of $5.7\pm0.2$ R and a scale height of $1.0\pm0.1$ kpc if the Perseus arm is at $D = 2.5$ kpc (see Reynolds et al.\ 1995\markcite{rtkmh95} and references therein). The dotted line in Figure~\ref{fig:havstanb} presents a manual fit to the lower envelope of the Perseus arm emission. The resulting parameters are essentially the same: $H = 1.0$ kpc and $\ensuremath{I_{\mathrm{H}\alpha}}^0 = 4.2$ R. The 1.0 kpc scale height appears to be valid for $|z|$ from 750 pc to at least 1700 pc, the boundary of this studied region. As mentioned above, this estimate assumes that the filling fraction and temperature do not depend on $|z|$. If in fact $\phi$ changes as a function of height above the plane, it affects our estimate of the electron scale height. As shown below in \S\ref{sec:discussion}, these new observations suggest that the temperature in the WIM rises with distance from the Galactic plane. Such an effect would increase our estimate of the scale height. Finally, we also note here that the full role of interstellar extinction has not been taken into account within our restricted fitting region. Future observations of H$\beta$ will allow a straightforward correction of the H$\alpha$ data for this attenuation. A simplistic estimate of the magnitude of this correction within our fitted region suggests a decrease in the inferred scale height of $\sim10\%$, somewhat offsetting the expected effect of $T(|z|)$. \subsection{Line Ratios} \label{sec:lr} Figures~\ref{fig:loc-lr} and \ref{fig:per-lr} also present pseudo-color images of (b) [\ion{N}{2}]/H$\alpha$, (c) [\ion{S}{2}]/H$\alpha$, and (d) [\ion{S}{2}]/[\ion{N}{2}] line ratios integrated over $\ensuremath{v_{\rm LSR}} = -10$ to $+10$ km~s$^{-1}$\ (Figure~\ref{fig:loc-lr}) and $\ensuremath{v_{\rm LSR}} = -30$ to $-50$ km~s$^{-1}$\ (Figure~\ref{fig:per-lr}), sampling the Local and Perseus arms, respectively. The [\ion{N}{2}]/H$\alpha$\ and [\ion{S}{2}]/H$\alpha$\ maps in both arms generally anti-correlate with H$\alpha$\ emission except for the $\phi$ Per \ion{H}{2}\ region centered near $\ell = 131\arcdeg$, $b = -11\arcdeg$ mentioned in \S\ref{sec:overview} above. [\ion{S}{2}]/[\ion{N}{2}] ratios are generally lower in the the two brighter \ion{H}{2}\ regions in the Local arm, but otherwise show little additional correspondence to H$\alpha$\ maps. In Figures~\ref{fig:siivsha} and \ref{fig:niivsha}, [\ion{S}{2}]/H$\alpha$\ and [\ion{N}{2}]/H$\alpha$\ are plotted against H$\alpha$\ intensity. The radial velocity ranges for each panel of the figures are restricted to separate the Local and Perseus arm emission (see Figure~\ref{fig:spectra}). As above, the longitudinal extent of the Perseus observations is limited to $\ell < 152\arcdeg$. We denote the pointings within the newly discovered $\phi$ Per \ion{H}{2}\ region (see \S\ref{sec:overview}) with a square symbol. Note that these points appear to form the upper envelope of the [\ion{S}{2}]/H$\alpha$\ ratios for their H$\alpha$\ intensities (Figure~\ref{fig:siivsha}) and actually form a distinctly different horizontal branch in the [\ion{N}{2}]/H$\alpha$\ ratios (Figure~\ref{fig:niivsha}). In general, both figures show a marked increase in these ratios with decreasing H$\alpha$\ intensity. For comparison to a more traditional \ion{H}{2}\ region in this mapped area, the horizontal, solid line at the right edge of both figures shows the average ratio from the six pointings with $\ensuremath{I_{\mathrm{H}\alpha}} > 50$ R near NGC 1499: [\ion{S}{2}]/H$\alpha$\ = 0.14 and [\ion{N}{2}]/H$\alpha$\ = 0.39. Ratios for a brighter \ion{H}{2}\ region, our calibration source NGC 7000 ($\ensuremath{I_{\mathrm{H}\alpha}} = 800$ R), are [\ion{S}{2}]/H$\alpha$\ = 0.06 and [\ion{N}{2}]/H$\alpha$\ = 0.20 (see also Reynolds 1985a\markcite{rjr85a}). Figure~\ref{fig:siivsnii} plots these two line ratios against one another. In both the Local and Perseus arms, the ratios appear to be strongly correlated. In this figure, the $\phi$ Per (B-star) and $\xi$ Per (O-star) \ion{H}{2}\ regions are highlighted with square and triangle symbols, respectively. Both appear to occupy specific loci that are different not only from one another, but from the general WIM background. The Perseus arm panel is again restricted to $\ell < 152\arcdeg$. As discussed in \S\ref{sec:hvsb}, the general trend in H$\alpha$\ intensity is that it decreases as the observed gas lies farther from the Galactic plane. Figure~\ref{fig:lrvb} shows that the relationship of the [\ion{S}{2}]/H$\alpha$\ and [\ion{N}{2}]H$\alpha$\ ratios versus H$\alpha$\ intensity described above translate directly into an increase of these ratios with increasing $|b|$ (or $|z|$) in the Perseus arm. The figure shows the three line ratios [\ion{S}{2}]/H$\alpha$, [\ion{N}{2}]/H$\alpha$, and [\ion{S}{2}]/[\ion{N}{2}] as a function of $\tan |b|$. Both panels are restricted to the longitude range $\ell = 125\arcdeg$--$152\arcdeg$ for consistency. Note also that the [\ion{S}{2}]/[\ion{N}{2}] ratio is relatively constant, with perhaps a slight increase in the Local arm with increasing $|b|$. \section{Discussion} \label{sec:discussion} The tight correlation of [\ion{S}{2}] and [\ion{N}{2}] emission in Figure~\ref{fig:siivsnii} results in a fairly constant value of [\ion{S}{2}]/[\ion{N}{2}] over a wide variety of H$\alpha$\ intensities and latitudes in both arms. This result has been seen in numerous studies of the extragalactic WIM analogue, commonly referred to as the diffuse ionized gas (DIG) (Otte \& Dettmar 1999\markcite{od99}; Rand 1998\markcite{rand98}; Wang, Heckman, \& Lehnert 1997\markcite{whl97}; Greenawalt, Walterbos, \& Braun 1997\markcite{gwb97}). As Rand covers in detail, current photoionization models in which variations in [\ion{S}{2}]/H$\alpha$\ and [\ion{N}{2}]/H$\alpha$\ arise primarily from variations in the ionization parameter (\emph{e.g.}\ Domg\"orgen \& Mathis 1994\markcite{dm94}; Sokolowski 1994\markcite{s94}; see also Bland-Hawthorn, Freeman, \& Quinn 1997\markcite{bfq97}) have difficulty keeping this ratio constant as [\ion{S}{2}]/H$\alpha$\ and [\ion{N}{2}]/H$\alpha$\ rise. Abundance gradients and \ion{H}{2}\ region contamination can be a larger problem in interpreting such ratios in extragalactic studies with the large physical sizes represented by a typical data point. Our new observations within the Galaxy provide smaller scales and adequate velocity resolution to separate radial components of gas, avoiding some of these problems. We suggest that the tight correlation of [\ion{S}{2}]/[\ion{N}{2}] seen in our data arises simply from the dominance of S$^{+}$ and N$^{+}$ ionization states in this gas together with the similar excitation potential of these two emission lines. With increasing electron temperature, [\ion{S}{2}]/H$\alpha$\ and [\ion{N}{2}]/H$\alpha$\ ratios rise together with no change in the [\ion{S}{2}]/[\ion{N}{2}] ratio. Using this premise, namely, that variations in these ratios are due primarily to variations in electron temperature, we attempt to derive estimates of the electron temperature and ionization state of S within our sample. We start with the basic equation for the intensity of emission lines from collisionally excited ions (Osterbrock 1989\markcite{o89}), \begin{equation} \label{eq:collisional} I_\nu\ \mathrm{(ph\ s^{-1}\ cm^{-2}\ ster^{-1})} = \frac{f_\nu}{4\pi}\, \int n_i\,n_e\;\frac{8.63 \times 10^{-6}}{T^{0.5}}\,\frac{\Omega(i,j)}{\omega_i}\,e^{-\frac{E_{ij}} {k T}}\,dl, \end{equation} where $\Omega(i,j)$ is the collision strength of the transition, $\omega_i$ is the statistical weight of the ground level, $E_{ij}$ is is the energy of the upper level of the transition above the lower, and $f_\nu$ is the fraction of downward transitions that produce the emission line in question. We adopt the following parameterized representations of the collision strengths supplied by Aller (1984\markcite{a84}), valid from 5000 K to 20,000 K: \begin{eqnarray} \label{eq:css} \mathrm{S^+\!:}\ \Omega(^4S_{3/2},^2D_{5/2}) & = & 4.19 \times T_4^{-0.093}\\ \label{eq:csn} \mathrm{N^+\!:}\ \Omega(^3P,^1D) & = & 2.68 \times T_4^{0.026} \end{eqnarray} where $T_4$ is in units of $10^4$ K. A term to account for extinction has not been included in Equation~\ref{eq:collisional}, since the three lines we are comparing are close in wavelength. To arrive at any estimates, we must be able to compute the path averaged integrals in these equations. Since there is little knowledge yet of the details of the spatial distribution of the physical parameters in the WIM, we take the simplest approach and assume that physical values do not vary along a path. For the ratios we calculate below, we need to make this assumption for the temperature, ionization, and abundance of the gas. Our estimates should then be viewed as representative values, rather than accurately reflecting local conditions from where the emission arises. Since we are able to cleanly separate Local and Perseus arm gas, large-scale Galactic variations can be probed along lines of sight, even with this assumption. This simplification also implies that the line emission from the three ions originates from the same location. This is probably a valid assumption in the WIM, unless local processes on small scales (\emph{e.g.}\ shocks) dominate the ionization structure. In the brighter, discrete \ion{H}{2}\ regions, however, this assumption most likely breaks down since the ionization structure of the nebula could be changing significantly within our beam. With this simplification, we arrive at these two equations for the intensity of [\ion{S}{2}] and [\ion{N}{2}] emission, \begin{equation} \label{eq:sii} I_{6716}\ \mathrm{(R)} = 2.79 \times 10^{5}\;\left(\frac{H^+}{H}\right)^{-1} \left(\frac{S^{\rule{0ex}{1ex}}}{H}\right) \left(\frac{S^+}{S}\right)\;T_4^{-0.593}\;e^{-2.14/T_4}\;EM, \end{equation} and \begin{equation} \label{eq:nii} I_{6583}\ \mathrm{(R)} = 5.95 \times 10^{4}\;\left(\frac{H^+}{H}\right)^{-1} \left(\frac{N^{\rule{0ex}{1ex}}}{H}\right) \left(\frac{N^+}{N}\right)\;T_4^{-0.474}\;e^{-2.18/T_4}\;EM \end{equation} where $T_4$ is in units of $10^4$ K and $EM = \int n_e^2\,dl$ is the emission measure in units of cm$^{-6}$ pc. Note also that $I$ is now in R. The [\ion{S}{2}]/[\ion{N}{2}] intensity ratio is then \begin{equation} \label{eq:siinii} \frac{I_{6716}}{I_{6583}} = 4.69\; \frac{\left(\frac{S^{\rule{0ex}{1ex}}}{H}\right)} {\left(\frac{N^{\rule{0ex}{1ex}}}{H}\right)}\, \frac{\left(\frac{S^+}{S}\right)}{\left(\frac{N^+}{N}\right)}\; e^{0.04/T_4}\;T_4^{-0.119}. \end{equation} We use the solar (S/H) $= 1.86\times10^{-5}$ from Anders \& Grevesse (1989\markcite{ag89}) and (N/H) $= 7.5\times10^{-5}$ from Meyer, Cardelli, \& Sofia (1997\markcite{mcs97}) for the gas-phase abundances of S and N. Because of the similar first ionization potentials of N and H (14.5 and 13.6 eV) and a weak charge-exchange reaction, N$^+$/N$^0$ tracks H$^+$/H$^0$ in photoionization models (Sokolowski 1994\markcite{s94}), so that in H$\alpha$\ emitting regions, where the fractional ionization of H has been measured to be near unity (Reynolds et al.\ 1998\markcite{rhth98}), N$^+$/N $\rightarrow$ 1 also. However, with a second ionization potential of 29.6 eV, N most likely ionizes no higher than N$^+$. Recent photoionization models run by Howk \& Savage (1999\markcite{hs99}) over a wide range of temperatures and ionization parameters show that even if the ionizing spectrum in the WIM is as hard as $T_e\approx40,000$ K, N$^{++}$/N $< 0.3$. Furthermore, Reynolds \& Tufte (1995\markcite{rt95}) and Tufte (1997\markcite{t97}) found that He$^+$/He ranges from 0.3 to about 0.6 in the WIM, suggesting a spectrum that is significantly softer than 40,000 K, and therefore implying a very small N$^{++}$/N ratio. Sulfur, on the other hand, has a low enough first ionization potential, 10.4 eV, that it is most likely fully ionized in the gas we are studying. In addition, its second potential, 23.4 eV, lies slightly below the neutral He edge at 24.6 eV. The existence of He$^+$ in the WIM (Reynolds \& Tufte 1995\markcite{rt95}; Tufte 1997\markcite{t97}) and the model calculations of Howk \& Savage (1999\markcite{hs99}) both imply that indeed some S will be ionized to S$^{++}$. In the limit where S$^+$/S and N$^+$/N are both unity and at a typical WIM temperature of $T_4=0.8$, Equation~\ref{eq:siinii} gives $I_{6716}/I_{6583} = 1.26$. This relationship is plotted as the solid line in Figure~\ref{fig:loc-grid}. Dotted lines of lower slope show the effect of decreasing S$^+$/S (presumably due to an increase in S$^{++}$) while keeping N$^+$/N $= 1$. In this interpretation, the data imply that S$^+$/S ranges between about 0.25 and 0.8, with an average value for the Local diffuse background of 0.6--0.65. Note that the location of the \ion{H}{2}\ region symbols (filled triangles and squares) in Figure~\ref{fig:loc-grid} indicate that these regions have systematically lower S$^+$/S than the diffuse background. In these cases the lower ratios are more likely due to the expected behavior that closer to the ionizing source, where the ionization parameter is large, more S$^+$ is ionized to S$^{++}$, producing a corresponding decrease in the observed [\ion{S}{2}] intensity relative to H$\alpha$\ (Domg\"orgen \& Mathis 1994\markcite{dm94}; Sokolowski 1994\markcite{s94}). Just as the [\ion{N}{2}]/H$\alpha$\ and [\ion{S}{2}]/H$\alpha$\ ratios can together give us information on the ionization state of S, examining the ratios alone can provide some estimate of the temperature. From Equations~\ref{eq:itoem} and \ref{eq:nii} the [\ion{N}{2}]/H$\alpha$\ ratio is given by \begin{equation} \label{eq:niiha} \frac{I_{6583}}{\ensuremath{I_{\mathrm{H}\alpha}}} = 1.63\times10^5\;\left(\frac{H^+}{H}\right)^{-1} \left(\frac{N^{\rule{0ex}{1ex}}}{H}\right) \left(\frac{N^+}{N}\right)\;T_4^{0.426}\;e^{-2.18/T_4}. \end{equation} Since N$^+$/N $\approx$ H$^+$/H (see above), the H and N ionization factors can be removed from Equation~\ref{eq:niiha}, with the result that for a a given (N/H) abundance, [\ion{N}{2}]/H$\alpha$\ is only a function of the electron temperature. As an example, the typical WIM temperature of $T_4 = 0.8$ then leads to [\ion{N}{2}]/H$\alpha$\ $= 0.72$. Lines of constant temperature are plotted on Figure~\ref{fig:loc-grid} as dashed vertical lines. With a better estimate for temperature from this line ratio, we also recompute Equation~\ref{eq:siinii}, including the temperature dependencies this time, to form the dash-dotted line in Figure~\ref{fig:loc-grid}. This correction is small. Figure~\ref{fig:per-grid} shows the same grid of S$^+$/S and $T$ overplotted on the Perseus arm ratios. The limits placed on S$^{++}$/S in the WIM by Howk \& Savage (1999\markcite{hs99}) are at odds with the significant number of points below the S$^+$/S $= 0.5$ line in the Perseus arm. As noted in \S\ref{sec:obs}, there may be an atmospheric line within the integration range of Perseus arm [\ion{N}{2}] emission. By comparing [\ion{S}{2}] and [\ion{N}{2}] spectra toward the same locations, we derive an upper limit of $0.1$ R for the intensity of this line. The maximum effect of removing this line is displayed for a few sample points in Figure~\ref{fig:per-grid} by translating the solid sample error bars points to the location of the dotted error bars. Such a correction would imply a slightly higher S$^+$/S ratio on average for the Perseus arm, but not enough to explain the entire difference between the arms. We may be probing a harder radiation field in the Perseus arm gas than their model input spectra or our assumed value of S/H may be too high. A reduction in the gas phase sulfur abundance by 0.5--0.75 (0.3--0.1 dex) would bring our inferred S$^{++}$/S fractions more in line with their models. The Perseus arm gas appears to have systematically lower S$^+$/S than the Local arm, with an average value closer to 0.5. Equation~\ref{eq:siinii} shows that differences in abundances between the Local and Perseus arm could change the slope of the [\ion{S}{2}]/H$\alpha$\ and [\ion{N}{2}]/H$\alpha$\ relationship. However, Afflerbach, Churchwell, \& Werner (1997\markcite{acw97}) measured the Galactic abundance gradients in ultra-compact \ion{H}{2}\ regions for O, S, and N. Using their best fits, N/H falls off faster than S/H with increasing Galactocentric distance so that S/N is $\sim 5\%$ higher in the Perseus arm. This is the opposite effect needed to explain the difference in the Perseus and Local arm emission ratios. More likely, Figure~\ref{fig:per-grid} suggests that this region of the Perseus arm has a higher fraction of S$^{++}$ than the Local arm region. Rand (1998\markcite{rand98}) observed an increase in [\ion{O}{3}]/H$\beta$\ with height above the plane of NGC 891 and concluded that an additional source of ionization is needed beyond photoionization from hot stars in the disk. In a region where O$^{++}$ ions are being maintained, S$^+$ becomes a less dominant species. This is consistent with the inferred differences in S$^+$/S between the Perseus arm (Figure~\ref{fig:per-grid}) and Local arm (Figure~\ref{fig:loc-grid}) data, since lines of sight pass through the Perseus arm at much higher $|z|$ than for Local gas. One problem with this interpretation is that we do not see a substantial fall in the [\ion{S}{2}]/[\ion{N}{2}] ratio with increasing $|z|$ in the Perseus arm (Figure~\ref{fig:lrvb}). Instead of an additional ionizing source, it is also possible that there are marked differences in the strength of the halo ionizing field between the Local and Perseus arms. For example, NGC 891 shows considerable variation in the extent of extended H$\alpha$\ emission along the galactic plane (Rand, Kulkarni, \& Hester 1990\markcite{rkh90}; Rand 1998\markcite{rand98}). The scatter of data points suggests significant variations in the electron temperature within the mapped region, from about 6,000 K to 10,000 K. The most striking result from these temperature estimates comes when re-examining the trend of the [\ion{N}{2}]/H$\alpha$\ ratio in Figures~\ref{fig:per-lr}b and \ref{fig:lrvb}. From the rise in this ratio with increasing $|z|$ and Equation~\ref{eq:niiha} (or the grid of Figure~\ref{fig:per-grid}), we infer that the temperature in the halo rises from about 7,000 K at $|z| = 0.75$ kpc ($b \sim -17\arcdeg$) to over 10,000 K at $|z| = 1.75$ kpc ($b \sim -35\arcdeg$). If the electron temperature is truly rising with $|z|$ as our results suggest, the calculation of the scale height of H$\alpha$\ emitting gas (\S\ref{sec:hvsb}) becomes more complicated. If we confirm this effect with future observations (see \S\ref{sec:summary}), the net effect (Equation~\ref{eq:itoem}) is that the true scale height of the gas is larger than the value calculated in \S\ref{sec:hvsb} by about 20\%. Note also that since fainter regions tend to have higher [\ion{N}{2}]/H$\alpha$\ (Figure~\ref{fig:niivsha}), our analysis implies that fainter regions are hotter. Since $EM \propto \ensuremath{I_{\mathrm{H}\alpha}} T^{0.9}$, higher temperature regions naturally produce lower H$\alpha$\ intensities for a given $EM$; however, even after converting \ensuremath{I_{\mathrm{H}\alpha}}\ to $EM$ using a varying temperature estimate from the [\ion{N}{2}]/H$\alpha$\ observations, regions with lower $EM$ are still hotter. Since $EM$ depends on $\phi n_e^2$ and the pathlength through the emitting region, one of the two must be smaller in these fainter, hotter regions. Pathlength is an inherently unphysical parameter to link to temperature and we can think of no reason why our data set would bias shorter pathlength sightlines toward higher temperature regions. Thus, we conclude that the hotter regions most likely have smaller $\langle n_e^2 \rangle$. Since $n_e$ in our mapped region of the Perseus arm decreases with increasing $|z|$ (see \S\ref{sec:hvsb}), we are not able to ascertain from these data whether $n_e$ or $|z|$ is the fundamental parameter influencing the electron temperature. One possible explanation for the discrepancy between these results and photoionization models (\emph{e.g.}\ Domg\"orgen \& Mathis 1994\markcite{dm94}; Sokolowski 1994\markcite{s94}) is that at high $|z|$ or low $n_e$ an additional heating source may dominate over the heating from photoionization (Bland-Hawthorn et al.\ 1997\markcite{bfq97}). Since the heating per volume from photoionization is proportional to $n^2_e$ (it is limited by recombination), a heating term that is proportional to $n_e$ or did not depend on $n_e$ at all would become dominant at sufficiently low densities. Such an effect would decouple the heating from the ionization of atoms, keeping [\ion{S}{2}]/[\ion{N}{2}] constant, while driving up the [\ion{S}{2}] and [\ion{N}{2}] intensities relative to H$\alpha$. We also note that if such a process exists in the halo, the increase in [\ion{O}{3}]/H$\beta$\ seen by Rand (1998\markcite{rand98}) in the halo of NGC 891 could be a direct consequence of this additional heating source, and would not require a secondary ionization source. Possible heating sources include, for example, the dissipation of interstellar turbulence (Minter \& Spangler 1992\markcite{ms97}) and the photoionization of interstellar grains (Reynolds \& Cox 1992\markcite{rc92}). The idea of an additional heating source will be explored further in a related work (Reynolds, Haffner, \& Tufte 1999\markcite{rht99}). \section{Summary} \label{sec:summary} We have presented the first comprehensive, velocity-resolved maps of [\ion{N}{2}] and [\ion{S}{2}] in the Galaxy. Combining the H$\alpha$\ maps from the WHAM H$\alpha$\ Sky Survey in this region, we draw the following conclusions: \begin{enumerate} \item We confirm earlier estimates of the distribution of H$\alpha$\ emission in the Perseus arm (Reynolds 1997\markcite{rjr97}). With continuous coverage from $b = -6\arcdeg$ to $-35\arcdeg$ ($|z| = 300$ to $1750$ pc at $d = 2.5$ pc), we find that the electron distribution derived from the emission is described well by an exponential with a scale height of $H = 1.0\pm0.1$ kpc. \item \mbox{[\ion{S}{2}]}/H$\alpha$\ and [\ion{N}{2}]/H$\alpha$\ emission in our Galaxy shows the same rise with increasing $|z|$ as found in observations of edge-on galaxies (Rand 1998\markcite{rand98}; Golla et al.\ 1996\markcite{gdd96}). These ratios appear to be correlated with each other over a wide range of emission measures and $|z|$. \item We suggest that the correlation between [\ion{N}{2}]/H$\alpha$\ and [\ion{S}{2}]/H$\alpha$\ can be attributed to the fact that changes in these ratios are primarily due to changes in electron temperature rather than ionization parameter, and that scatter in the relationship is caused by variations in S$^+$/S. The maps of [\ion{N}{2}]/H$\alpha$\ and [\ion{S}{2}]/[\ion{N}{2}] (Figures~\ref{fig:loc-lr} and \ref{fig:per-lr}) can then be viewed as distributions of electron temperature and ionization state, respectively. With this interpretation, we derive temperatures in the WIM of 6,000 K to 10,000 K and S$^+$/S ratios of 0.3 to 0.8. The average value of S$^+$/S appears to be lower in the Perseus arm. Higher ionization and lower temperatures are mostly found in the two \ion{H}{2}\ regions surrounding the O star $\xi$ Per and the B star system $\phi$ Per. Differences in temperature between regions may be able to be directly confirmed with observations of the faint, higher-level [\ion{N}{2}] $\lambda$5755 transition, which when combined with our $\lambda$6483 measurements can directly measure the electron temperatures in the emitting gas. \end{enumerate} We thank Kurt Jaehnig and Jeff Percival of the University of Wisconsin's \emph{Space Astronomy Lab} for their dedicated engineering support of WHAM; Nikki Hausen, Mark Quigley, and Brian Babler for their contributions to the data analysis; and Trudy Tilleman for essential night-sky condition reports from Kitt Peak, which have made remote observing possible. We acknowledge the use of the SIMBAD database, operated at CDS, Strasbourg, France. This work is supported by the National Science Foundation through grant AST9619424. \newpage
\section{Introduction} The \einstein\ observatory performed a survey of the Small Magellanic Cloud (SMC) in which two sources with unusually soft spectra, 1E~0056.8-7154 and 1E~0035.4-7230, were detected (Seward \& Mitchell 1981). 1E~0035.4-7230 has been identified in the optical with a variable blue star with a strong UV excess and a 4.1~hour orbital period (Orio et al. 1994). The supersoft nature of the source was confirmed by subsequent ROSAT observations (Kahabka et al. 1994; Kahabka 1996). Supersoft sources are most probably accreting white dwarfs (WDs) which are burning hydrogen steadily (van den Heuvel et al. 1992). An orbital period of 0.1719~days was found in optical and X-ray data (Crampton et al. 1997; see van Teeseling et al. 1998 for a recent discussion). The optical spectrum shows pronounced emission at He{\sc ii}~$\lambda$4686 and near H$\alpha$. The H$\beta$ line is seen in absorption (Cowley et al. 1998). The mass function derived from the He{\sc ii}~$\lambda$4686 line radial velocities indicates a donor star mass of $\sim$${\rm 0.4\ M_{\odot}}$ and a WD mass of $\sim$${\rm0.5-1.4\ M_{\odot}}$. A scenario to account for the evolutionary state of the source has been proposed by Kahabka \& Ergma (1997) in which the system is a cataclysmic variable (CV) with a low-mass WD accreting below the steady-state burning rate. The recurrent flashes in this system are assumed to be mild and leave a fraction of the envelope as a nuclear reservoir. The WD is heated to a high temperature during its evolution (Sion \& Starrfield 1994). The Sion \& Starrfield models of a ${\rm 0.6\ M_{\odot}}$ and a ${\rm 0.7\ M_{\odot}}$ WD predict an asymptotic temperature of $\sim$$\rm 3.25\ 10^5~$K and a luminosity of $\sim$$\rm 10^{37}$~erg~s$^{-1}$. X-ray spectroscopy allows the WD mass and temperature to be estimated, so allowing the evolutionary status of the system to be probed. If 1E~0035.4-7230 harbors a massive WD (${\rm M_{\rm WD} > 0.7-0.8\ M_{\odot}}$) then the Sion \& Starrfield model may no longer be applicable as the envelope is probably expelled during recurrent nova outbursts. This work is part of a program to systematically observe the brighter supersoft sources with the BeppoSAX satellite (Parmar et al. 1997a, 1998; Hartmann et al. 1999). \section{Observation} The Low-Energy Concentrator Spectrometer (LECS) onboard \sax\ is an imaging gas scintillation proportional counter sensitive in the energy range 0.1--10.0~keV with a circular field of view of 37$'$ diameter (Parmar et al. 1997b). Its energy resolution is a factor $\sim$2 better than that of the ROSAT Position Sensitive Proportional Counter (PSPC). 1E~0035.4-7230\ was observed between 1998 January 4 10:27 and January 6 01:21~UTC. Good data were selected from intervals when the minimum elevation angle above the Earth's limb was $>$5$^\circ$ and when the high voltage levels were nominal using the SAXDAS data analysis package. Since the LECS was only operated during satellite night-time, this gave a total on-source exposure of 37~ks. Examination of the LECS image shows a source at a position consistent with that of 1E~0035.4-7230. A spectrum was extracted centered on the source centroid using a radius of 8$'$. This radius was chosen to include 95\% of the 0.28~keV photons. The spectrum was rebinned to have $>$20 counts in each bin to allow the use of the $\rm \chi^2$ statistic. The LECS response matrix from the 1997 September release was used in the spectral analysis. The background spectrum was extracted from the image itself using a semi-annulus centered on 1E~0035.4-7230 with inner and outer radii of 14$'$ and 19$'$, respectively. A correction for telescope vignetting was applied (see Parmar et al. 1999). The 1E~0035.4-7230\ count rate above background is 0.008~s$\rm ^{-1}$. Examination of the extracted spectrum shows that the source is only detected in a narrow energy range and only the 21 rebinned channels corresponding to energies between 0.13 and 1.7~keV were used for spectral fitting. \section{BeppoSAX spectral fit} Initially only BeppoSAX data was used to constrain the parameters derived from different spectral models. The fits indicate the presence of a C edge at $\sim$0.5~keV, most likely due to C~{\sc vi} and this provides strong support for a hot and nuclear burning WD atmosphere model. \begin{figure}[htbp] \centering{ \vbox{\psfig{figure=8569_f1.ps,width=8.0cm,% bbllx=1.0cm,bblly=0.5cm,bburx=22.0cm,bbury=10.1cm,clip=}}\par } \caption[]{BeppoSAX count spectrum of the 37~ks observation of 1E~0035.4-7230 together with the best-fit non-LTE spectral fit} \end{figure} \begin{figure}[htbp] \centering{ \vbox{\psfig{figure=8569_f2a.ps,width=7.0cm,% bbllx=1.0cm,bblly=0.5cm,bburx=22.0cm,bbury=10.0cm,clip=}}\par \vbox{\psfig{figure=8569_f2b.ps,width=7.0cm,% bbllx=1.0cm,bblly=0.5cm,bburx=22.0cm,bbury=10.0cm,clip=}}\par \vbox{\psfig{figure=8569_f2c.ps,width=7.0cm,% bbllx=1.0cm,bblly=0.5cm,bburx=22.0cm,bbury=10.0cm,clip=}}\par \vbox{\psfig{figure=8569_f2d.ps,width=7.0cm,% bbllx=1.0cm,bblly=0.5cm,bburx=22.0cm,bbury=10.0cm,clip=}}\par \vbox{\psfig{figure=8569_f2e.ps,width=7.0cm,% bbllx=1.0cm,bblly=0.5cm,bburx=22.0cm,bbury=10.0cm,clip=}}\par } \caption[]{(a) Best-fit blackbody model. The blackbody kT is 39~eV and the column ${\rm 4.7\ 10^{20}}$ ${\rm H-atoms~cm^{-2}}$.(b) Best-fit blackbody model with an edge at 0.44~keV. The optical depth is larger than 1.8 (68\% confidence). The blackbody kT is 79~eV and the column ${\rm 2.2\ 10^{20}}$ ${\rm H-atoms~cm^{-2}}$. (c) Best-fit blackbody model with a C~{\sc v} edge at 0.391~keV and a C~{\sc vi} edge at 0.487~keV. The optical depths are 0.4 and $>$1.3, respectively. The blackbody kT is 63~eV and N${\rm _H}$\ ${\rm 3.1\ 10^{20}}$ ${\rm H-atoms~cm^{-2}}$. (d) Best-fit LTE WD atmosphere model. The kT is $\rm 5.5\ 10^5$~K and N${\rm _H}$\ ${\rm 3.3\ 10^{20}}$ ${\rm H-atoms~cm^{-2}}$. (e) Best-fit non-LTE WD atmosphere model. The kT is $3.9\ 10^5$~K and N${\rm _H}$\ ${\rm 3.5\ 10^{20}}$ ${\rm H-atoms~cm^{-2}}$} \end{figure} \subsection{Blackbody spectral fits} The BeppoSAX data were first fit with a blackbody spectral model. The fit is acceptable with a $\chi^2$ of 8.5 for 18 degrees of freedom (dof). The blackbody temperature, kT, is (20--52)~eV and the column, N${\rm _H}$, ${\rm (2.2-11)\ 10^{20}}$ ${\rm H-atoms~cm^{-2}}$. The best-fit luminosity (60~kpc) is ${\rm 10^{37}\ erg\ s^{-1}}$, but it is not well constrained (cf. Tab.1). Next edges were included in the spectral model. \begin{equation}{\rm Bbody \times f(E)} \end{equation} where \begin{equation}{\rm f(E) = \left\{ \begin{array}{c@{\quad:\quad}l} 1. & E<E_{\rm edge} \\ exp(-\tau \times ({\frac{E}{E_{\rm edge}}})^3) & E\ge E_{\rm edge} \end{array} \right. } \end{equation} Edges are expected to be present in the atmospheres of hot WDs. Calculations of synthetic spectra have shown that such edges are the dominant spectral features in this energy range. These are likely to be Lyman edges, i.e. ionizations from the ground state of C~{\sc v} and C~{\sc vi} (at energies of 0.391 and 0.487~keV). A single edge with variable energy was first included. As the fit did not allow to constrain the energy of the edge we fixed the edge energy to 0.44~keV, the mean expected energy of the C~{\sc v} and the C~{\sc vi} edge of 0.439~keV. The optical depth, $\tau>1.8$, at the 68\% confidence level. At the 90\% confidence level $\tau$ cannot be constrained. The $\chi^2$ is 7.1 for 17 dof. The best-fit blackbody kT is 79~eV. The N${\rm _H}$\ of ${\rm (0.8-8.8)\ 10^{20}}$ ${\rm H-atoms~cm^{-2}}$ is in agreement with the foreground column of ${\rm 3.2\ 10^{20}}$ ${\rm H-atoms~cm^{-2}}$ (derived from the 21-cm map of Stanimirovic et al. 1999). The luminosity is ${\rm (0.7-8.4)\ 10^{36}\ erg\ s^{-1}}$. We applied the F-test statistics and find a probability of $\sim$70\% that a fit including a C-edge is more likely than a fit without an edge (leaving the blackbody temperature in the fit as a free parameter). Next edges fixed at the energies of the C~{\sc v} and C~{\sc vi} edges were included in the spectral model. The best-fit optical depths are ${\rm \tau_{\rm C~V}\le 1.8}$ and ${\rm \tau_{\rm C~VI}\ge 1.3}$ at the 68\% confidence level (${\rm \tau_{\rm C~V}\le 3.0}$ at the 90\% confidence level). The best-fit blackbody kT is 63~eV. The N${\rm _H}$\ is ${\rm (2.0-6.1)\ 10^{20}}$ ${\rm H-atoms~cm^{-2}}$, and is in agreement with the galactic foreground column. The luminosity is ${\rm (1.0-9.6)\ 10^{36}\ erg\ s^{-1}}$. \subsection{WD atmosphere spectral fits} A Blackbody spectrum including absorption edges of C~{\sc v} and C~{\sc vi} can only be an approximation to the true spectral distribution. LTE and non-LTE WD atmosphere spectra may be the better approximation (Heise et al. 1994; Hartmann \& Heise 1997). We therefore applied LTE and non-LTE WD atmosphere spectra to the BeppoSAX spectrum of 1E~0035.4-7230 assuming reduced abundances for C, N, O and Ne, also in Section 4.2. The LTE fit is acceptable with a $\chi^2$ of 8.0 for 17 dof. The effective kT is $(4.1-5.9)\ 10^5$~K, the N${\rm _H}$\ is ${\rm (1.6-6.5)\ 10^{20}}$ ${\rm H-atoms~cm^{-2}}$ and the best-fit luminosity ${\rm 3.4\ 10^{36}\ erg\ s^{-1}}$, the luminosity is not well constrained. The non-LTE fit is also acceptable with a $\chi^2$ of 8.0 for 17 dof. The best-fit effective kT of $(3.3 \pm _{0.3} ^{0.5})\ 10^5$~K is significantly lower than that derived from the LTE fit. The N${\rm _H}$\ is ${\rm (2.3-6.6)\ 10^{20}}$ ${\rm H-atoms\ cm^{-2}}$. The luminosity is in the range ${\rm (0.5-3.8)\ 10^{37}\ erg\ s^{-1}}$. \begin{table*} \caption[]{Results of spectral fitting of the BeppoSAX~LECS spectrum (0.1--1.7~keV) and combined BeppoSAX~LECS and ROSAT~PSPC spectrum of 1E~0035.4-7230. The (0.1--2.4)~keV luminosity L$_{36}$ ($\rm 10^{36}\ erg\ s^{-1}$) and for the combined ROSAT~PSPC and BeppoSAX~LECS spectral fit also the bolometric luminosity L$_{36}$ ($\rm 10^{36}\ erg\ s^{-1}$) is given. 90\% confidence parameter ranges are given ($^{a}$ 68\% confidence parameter range is given).} \begin{flushleft} \begin{tabular}{lccccccccc} \hline \noalign{\smallskip} \multicolumn{9}{c}{BeppoSAX LECS} \\ \noalign{\smallskip} Model &N${\rm _H}$&kT&${\rm E_{\rm C~{V}}}$&${\rm E_{\rm C~{VI}}}$ &${\rm \tau_{\rm C~V}}$&${\rm \tau_{\rm C~VI}}$ & L$_{36}$ && $\chi$/dof \\ & & & \multicolumn{4}{c}{Edge} & (0.1-2.4~keV)& \\ & ($10^{20}\ cm^{-2}$) &(eV; &\multicolumn{2}{c} {(keV)} & & & & \\ & &$10^5$~K) & & & & & & \\ \noalign{\smallskip} \hline \noalign{\smallskip} bbody & $4.7\pm_{2.5}^{6.8}$&$39\pm_{19}^{13}$& & & & &$10\pm_{9.3}^{>10}$ &&0.47\\ & &$4.5\pm_{2.2}^{1.5}$& & & && \\ \noalign{\smallskip} bbody+edge& $2.2\pm_{1.4}^{6.6}$&$79\pm_{55}^{19}$&\multicolumn{2}{c}{0.44}&\multicolumn{2}{c}{$>1.8^{a}$}&$2.2\pm_{1.5}^{6.2}$&&0.42\\ & &$9.2\pm_{6.3}^{2.2}$& & & && \\ \noalign{\smallskip} bbody+C{\sc v}+C{\sc vi} edge& $3.1\pm_{2.1}^{3.0}$&$63\pm_{23}^{33}$&0.391&0.487&$0.4\pm_{0.4}^{2.6}$ &$>1.3^{a}$ &$2.8\pm_{1.8}^{6.8}$ &&0.57\\ & &$7.3\pm_{2.2}^{2.9}$& & & && \\ \noalign{\smallskip} LTE WD atm. & $3.3\pm_{1.7}^{3.2}$&$47\pm_{12}^{3}$& & & & &$3.4\pm_{3.1}^{82}$&&0.47\\ & &$5.5\pm_{1.4}^{0.4}$& & & && \\ \noalign{\smallskip} Non-LTE WD atm.& $4.0\pm_{1.7}^{2.6}$&$28\pm_{3}^{4}$& & & & &$8.6\pm_{3}^{29}$ &&0.47\\ & &$3.3\pm_{0.3}^{0.5}$& & & & & && \\ \noalign{\smallskip} \noalign{\smallskip} \hline \noalign{\smallskip} \multicolumn{9}{c}{BeppoSAX LECS + ROSAT PSPC} \\ \noalign{\smallskip} Model & N${\rm _H}$ & kT & ${\rm E_{C~{V}}}$ & ${\rm E_{C~{VI}}}$ & ${\rm \tau_{\rm C~V}}$ & ${\rm \tau_{\rm C~VI}}$ & L$^{c}_{36}$ &scaling& $\chi$/dof\\ & & & \multicolumn{4}{c}{Edge} &(0.1-2.4~keV)&factor$^{b}$&\\ & ($10^{20}\ cm^{-2}$) &(eV; &\multicolumn{2}{c}{(keV)}& & & (bolometric) && \\ & &$10^5$~K) & & & & & && \\ \noalign{\smallskip} \hline \noalign{\smallskip} bbody & $4.8\pm_{0.8}^{1.1}$ &$43\pm_{4}^{4}$& & & & &$9.3\pm_{4.2}^{11}$&$1.68\pm_{0.19}^{0.23}$&0.68 \\ & &$5.0\pm_{0.5}^{0.5}$ && & & &$12\pm_{9}^{42}$&\\ \noalign{\smallskip} bbody+edge & $3.8\pm_{0.8}^{1.1}$&$63\pm_{15}^{13}$&\multicolumn{2}{c}{$0.44\pm_{0.08}^{0.03}$}&\multicolumn{2}{c}{$5.4{\pm_{2.7}^{3.6}}^{a}$}&$4.2\pm_{1.5}^{3.8}$&$1.67\pm_{0.17}^{0.23}$&0.62\\ & &$7.3\pm_{1.7}^{1.5}$ && & & &$5.2\pm_{2.0}^{0.9}$ & & \\ \noalign{\smallskip} bbody+C{\sc v}+C{\sc vi} edge & $3.9\pm_{0.8}^{1.2}$ &$65\pm_{15}^{16}$ &0.391&0.487 &$1.6{\pm_{0.7}^{1.0}}^{a}$ &$3.9{\pm_{2.8}^{3.6}}^{a}$&$4.2\pm_{1.4}^{3.3}$&$1.68\pm_{0.12}^{0.14}$&0.63\\ & &$7.5\pm_{1.7}^{1.5}$ && & & &$5.4{\pm_{3.5}^{3.6}}$ & & \\ \noalign{\smallskip} LTE WD atm.$^{d}$ & $3.6\pm_{0.4}^{0.4}$ &$47\pm_{1}^{1}$ & & & & &$2.0\pm_{0.8}^{1.4}$&$1.68\pm_{0.1}^{0.2}$&0.80\\ & &$5.5\pm_{0.1}^{0.1}$ & & & & & $4.8\pm_{0.8}^{1.1}$& \\ \noalign{\smallskip} Non-LTE WD atm.$^{d}$& $4.1\pm_{0.3}^{0.4}$ &$28\pm_{0.8}^{0.6}$& & & & &$9.8\pm_{1.4}^{4.0}$&$1.30\pm_{0.14}^{0.15}$&0.77\\ & &$3.25\pm_{0.09}^{0.07}$& & & & &$10.9\pm_{0.3}^{0.3}$&\\ \noalign{\smallskip} \noalign{\smallskip} \hline \end{tabular} \end{flushleft} $^{b}$ the flux scaling factor is ($\rm flux_{ROSAT PSPC}$ / ($\rm flux_{BeppoSAX}$), $^{c}$ the luminosity is given for the BeppoSAX spectrum. $^{d}$ 68\% confidence errors are given. In the text estimates for the 90\% errors are given as twice the 68\% errors. \end{table*} \begin{figure}[htbp] \centering{ \vbox{\psfig{figure=8569_f3a.ps,width=7.0cm,% bbllx=1.0cm,bblly=0.5cm,bburx=22.0cm,bbury=10.0cm,clip=}}\par \vbox{\psfig{figure=8569_f3b.ps,width=7.0cm,% bbllx=1.0cm,bblly=0.5cm,bburx=22.0cm,bbury=10.0cm,clip=}}\par \vbox{\psfig{figure=8569_f3c.ps,width=7.0cm,% bbllx=1.0cm,bblly=0.5cm,bburx=22.0cm,bbury=10.0cm,clip=}}\par \vbox{\psfig{figure=8569_f3d.ps,width=7.0cm,% bbllx=1.0cm,bblly=0.5cm,bburx=22.0cm,bbury=10.0cm,clip=}}\par \vbox{\psfig{figure=8569_f3e.ps,width=7.0cm,% bbllx=1.0cm,bblly=0.5cm,bburx=22.0cm,bbury=10.0cm,clip=}}\par } \caption[]{(a) Best-fit blackbody model. The blackbody kT is 43~eV and N${\rm _H}$\ is ${\rm 4.9\ 10^{20}}$ ${\rm H-atoms~cm^{-2}}$. (b) Best-fit blackbody model with an edge at 0.44~keV. The optical depth is 5.4. The blackbody kT is 63~eV and N${\rm _H}$\ is ${\rm 3.8\ 10^{20}}$ ${\rm H-atoms~cm^{-2}}$. (c) Best-fit blackbody model with a C~{\sc v} edge at 0.391~keV and a C~{\sc vi} edge at 0.487~keV. The optical depths are 1.6 and 3.9, respectively. The blackbody kT is 65~eV and N${\rm _H}$\ is ${\rm 3.9\ 10^{20}}$ ${\rm H-atoms~cm^{-2}}$. (d) Best-fit LTE WD atmosphere model. The kT is $5.5\ 10^5$~K, and the N${\rm _H}$\ ${\rm 3.6\ 10^{20}}$ ${\rm H-atoms~cm^{-2}}$. (e) Best-fit non-LTE WD atmosphere model. The kT is $3.25\ 10^5$~K, and the N${\rm _H}$\ ${\rm 4.1\ 10^{20}}$ ${\rm H-atoms~cm^{-2}}$.} \end{figure} \section{Combined BeppoSAX and ROSAT PSPC spectral fits} In a second step we used the BeppoSAX spectrum together with a ROSAT spectrum obtained from a 13~ks exposure in 1993 April in a combined (multi-instrument) spectral fit to better constrain the spectral parameters. Fits to the ROSAT data alone are presented in Kahabka et al. (1994). Again the fits indicate the presence of a C edge at $\sim$0.5~keV, most likely due to C~{\sc vi} and are in strong support for a hot and nuclear burning WD nature of the source. \subsection{Blackbody spectral fits} The data were first fit with a blackbody spectral model. The fit is acceptable with a $\chi^2$ of 40 for 59 dof. The blackbody kT is (39--47)~eV, the N${\rm _H}$\ of ${\rm (4.1-6.0)\ 10^{20}}$ ${\rm H-atoms~cm^{-2}}$ is greater than the foreground column of ${\rm 3.2\ 10^{20}}$ ${\rm H-atoms~cm^{-2}}$. The bolometric luminosity is ${\rm (1-5)\ 10^{37}\ erg\ s^{-1}}$. There is no indication of the existence of neutral circumstellar matter (e.g. a nebula) in this system. Any low density matter in the binary system (mass-loss due to winds) found from the modeling of the line profiles of the Balmer lines has been shown to be highly ionized (cf. Asai et al. 1998 for the supersoft source CAL~87 located in the LMC). Next a single edge with variable energy and depth was included in the spectral model. The fit is acceptable with a $\chi^2$ of 35 for 56 dof. The best-fit indicates an edge energy in the range (0.36--0.47)~keV with a best-fit value of 0.44~keV. This value is consistent with the mean expected energy of the C~{\sc v} and the C~{\sc vi} edge (at 0.391~keV and 0.487~keV, respectively) which is 0.439~keV. The derived energy range indicates an ionization state to be somewhere between C~{\sc v} and C~{\sc vi}. The optical depth is ${\rm \tau_{\rm C}}>1.1$ at the 90\% confidence level. In this model the blackbody temperature is found to be 48--76~eV. The N${\rm _H}$\ of ${\rm (3.0-5.0)\ 10^{20}}$ ${\rm H-atoms~cm^{-2}}$ is in agreement with the foreground column. We applied the F-test statistics and find a probability of $\sim$70\% that a fit including a C-edge is more likely than a fit without an edge (leaving the blackbody temperature in the fit as a free parameter). As before, the edge energies were next fixed at the values appropriate to C~{\sc v} and C~{\sc vi} edges in the spectral fit. The fit is acceptable with a $\chi^2$ of 35 for 56 dof. The best-fit optical depths are ${\rm \tau_{\rm C~V}}$= 0.9--2.6 and ${\rm \tau_{\rm C~VI}}$=1.1--7.5 at 68\% confidence, ${\rm \tau_{\rm C~V}}<2.9$ at 90\% confidence. The blackbody kT is in the range 50--81~eV. Using the Saha equation we may constrain the electron density (cf. Sect.~5.1). The N${\rm _H}$\ of ${\rm (3.1-5.1)\ 10^{20}}$ ${\rm H-atoms\ cm^{-2}}$ is in agreement with the galactic foreground column. \subsection{WD atmosphere spectral fits} The LTE fit assuming reduced abundances is acceptable with a $\chi^2$ of 47 for 59 dof. The effective kT is $(5.5\pm 0.2)\ 10^5$~K, the N${\rm _H}$\ is ${\rm (3.6 \pm 0.8) 10^{20}}$ ${\rm H-atoms\ cm^{-2}}$, the bolometric luminosity is ${\rm (3-7)\ 10^{36}\ erg\ s^{-1}}$. With non-LTE WD atmosphere spectrum and assuming reduced abundances (cf. section~3.2) significantly lower kT of $(3.25 \pm 0.16)\ 10^5$~K is found. The derived N${\rm _H}$\ is ${\rm (4.1 \pm0.8)\ 10^{20}}$ ${\rm H-atoms~cm^{-2}}$. The fit is acceptable with a $\chi^2$ of 46 for 58 dof. The bolometric luminosity is $\rm (1.0-1.2)\ 10^{37}\ erg\ s^{-1}$. \section{Discussion} We find strong evidence in the BeppoSAX and ROSAT spectra for the presence of an edge at $\sim$0.44~keV in the spectrum of 1E~0035.4-7230. This edge originates from a combination of C~{\sc v} and C~{\sc vi} with a large optical depth $\tau > 1.1$ (90\% confidence). The temperatures and absorbing columns derived from the blackbody fits should be treated with caution and more reliable estimates are probably obtained with the LTE or non-LTE WD atmosphere model fits. The resulting chi-squared values do not allow to decide between both models. Therefore, the temperature is constrained to the range $(3-6)\ 10^5$~K and the ionization state of C is likely between C~{\sc v} and C~{\sc vi}. The discovery of deep C edges in the spectrum of 1E~0035.4-7230 has implications as it is assumed that the metallicity of the SMC material is heavily reduced (by a factor 3--20, Pagel 1993). Reduced CNO abundances result in less violent outbursts. This can sustain nuclear burning for long periods of time without the requirement of additional supply due to accretion e.g. by X-ray induced heating of the donor star (van Teeseling \& King 1998; King \& van Teeseling 1998). It is unclear what enhances the C abundance in the envelope of the WD. If the WD goes through recurrent weak outbursts then the CO core may be affected (i.e. eroded) during the outburst. Eroded core material (i.e. C and O) may be mixed into the envelope. But as no N is supplied, it cannot affect the CNO cycle, i.e. it will not enhance the strength of the outburst. The soft X-ray spectrum of 1E~0035.4-7230 ressembles that of the symbiotic nova RX~J0048.4-7332 in the SMC (Jordan et al. 1996). For this source a strong C~{\sc v} edge is also required in the spectral modelling. We applied to the {\sl ROSAT} {\sl PSPC} spectrum of this source a blackbody spectrum including a C~{\sc v} and a C~{\sc vi} edge and find ${\rm \tau_{\rm C~V}\ge 5.8}$ and ${\rm \tau_{\rm C~VI}\ge 0.57}$ (68\% confidence). The best-fit blackbody temperature is $\sim$$4\ 10^5$~K, somewhat above the temperature derived from the non-LTE fit by Jordan et al. (1996). These authors find a strong stellar wind with a mass-loss rate of $\sim$${\rm 10^{-6}\ M_{\odot}\ yr^{-1}}$ for the symbiotic nova. \subsection{Constraints inferred from the Saha equation} Using the Saha equation we can predict the ratio of the depth in the C~{\sc v} and C~{\sc vi} edges assuming realistic values for the electron density in hot atmospheres of WDs and compare this ratio with that derived from observation. Considering only the ground state of ions, the Saha equation is \begin{equation}{\rm \frac{N^+}{N} N_e = \big(\frac{2\pi m_e kT}{h^2}\big)^{3/2}\ \frac{2 g^+}{g}\ e^{-\chi_i/kT}.} \end{equation} Here N is the particle density and $+$ designates the ionized state, ${\rm N_e}$ is the electron density, g and ${\rm g^+}$ are the statistical weights of the ions, and ${\rm \chi_i}$ is the ionization energy. For C~{\sc v} and C~{\sc vi} these weights are ${\rm g_{C~{\sc V}}=1}$ and ${\rm g_{C~{\sc VI}}=2}$ with ${\rm \chi_{C~{\sc V}-C~{\sc VI}} = 6.3\ 10^{-17}\ J = 392~eV}$. We assume for the optical depth $d\tau = N \sigma ds$ and $d\tau^+ = N^+\sigma^+ds^+$. $\sigma$ is the threshold photoionization cross-section and ds is the radial distance travelled by a photon. The definition of an absorption edge in Section 3.1 allows us to put $ds = ds^+$ since all blackbody radiation passes through the same absorbing slab of material. Then we get ${\rm d\tau = N\sigma ds}$ and ${\rm d\tau^+ = N^+\sigma^+ ds}$ gives \begin{equation}{\rm \frac{N^+}{N} \approx \frac{\tau^+\sigma}{\tau \sigma^+}} \end{equation} and \begin{equation}{\rm N_e = \frac{\tau\sigma^+}{\tau^+\sigma}\big(\frac{2\pi m_e kT}{h^2} \big)^{3/2} \frac{2g^+}{g} e^{-\chi_i/kT}} \end{equation} with $\rm \sigma=4.7\ 10^{-19}\ cm^{-2}$ and $\rm \sigma^+=1.810^{-19}\ cm^{-2}$ (Verner et al. 1996). For C~{\sc v} and C~{\sc vi} the electron density can be written as \begin{equation}{\rm N_e = 1.1710^{23}\ T_5^{3/2} \frac{\tau}{\tau^+} e^{-\chi_i/kT}\ cm^{-3}} \end{equation} with the temperature ${\rm T_5}$ in units of $10^5$~K. Constraining the temperature to be in the range $(4-6)\,10^5$~K gives \begin{equation}{\rm N_e = (0.11-8.8)\,10^{20}\ \frac{\tau}{\tau^+}\ cm^{-3}.} \end{equation} For high gravity (log~g=9.0) WD atmospheres electron densities in the range (${\rm 10^{19}-10^{20}\ cm^{-3}}$) are expected. Such densities are achieved for ratios in the range ${\rm \frac{\tau}{\tau^+}}$ ${\rm \sim0.1-1}$. The range derived from the blackbody and edge fit ${\rm \frac{\tau}{\tau^+}}$ ${\rm \sim0.1-2}$ is therefore consistent with that predicted. \subsection{The evolutionary state of 1E~0035.4-7230} The temperature and luminosity derived from the different spectral models can be compared with the asymptotic temperature of $3.25\ 10^5$~K and the bolometric luminosity of $\mathrel{\hbox{\rlap{\lower.55ex \hbox {$\sim$}$${\rm 10^{37}\ erg\ s^{-1}}$ predicted by Sion \& Starrfield (1994) for a low-mass (${\rm M_{\rm WD}\sim 0.6-0.7\ M_{\odot}}$) WD accreting at a rate of $\sim$${\rm 10^{-8}\ M_{\odot}\ yr^{-1}}$ and burning hydrogen steadily for thousands of years. With the non-LTE model a kT of $(3.25 \pm 0.16)\ 10^5$~K and a bolometric luminosity of ${\rm (1.0-1.2)\ 10^{37}\ erg\ s^{-1}}$ are derived, in agreement with the predictions from the Sion \& Starrfield model. The LTE model gives a higher effective kT of $5.5\pm0.2\ 10^5$~K and a somewhat lower bolometric luminosity of ${\rm (3-7)\ 10^{36}\ erg\ s^{-1}}$. Such a temperature is consistent with a more massive WD ${\rm M_{\rm WD}\sim 1.1\ M_{\odot}}$, assuming the stability relation in Iben (1982) (see Kahabka 1998). However the predicted luminosity is at least a factor of 10 too low in order to be explained by steady nuclear burning. This could be either due to obscuration of the source which is seen at a large inclination of $\sim$40\degree\ (cf. van Teeseling et al. 1998), or due to reduced nuclear burning. The X-ray luminosity of the source appeared not to have varied by more than $\sim1.30 \pm 0.3$ between the ROSAT and BeppoSAX observations. We cannot exclude that this change in luminosity could be of instrumental nature due to the uncertain cross-calibration of the BeppoSAX LECS and the ROSAT PSPC instruments. It also may be due to changing X-ray scattering in the binary system. If it is due to a long-term change in the luminosity caused by a cooling of the WD envelope then a temperature change by a factor of $1.07\pm 0.05$ would be required to account for the derived change in luminosity. The extreme softness of the spectrum of 1E~0035.4-7230 allows constraints to be set on the existence of a spectrally hard component in this system. Such a component may be expected if there is a strong colliding wind present. A wind with a mass-loss rate of $\sim$${\rm 10^{-6}\ M_{\odot}\ yr^{-1}}$ with a terminal velocity of $\sim$${\rm 500\ km\ s^{-1}}$ from a low-mass WD and a mass-loss rate of $\sim$${\rm 10^{-7}\ M_{\odot}\ yr^{-1}}$ with a lower terminal velocity from the donor star (as e.g. proposed by van Teeseling \& King 1998) should give rise to a colliding wind. We do not detect such a component from the BeppoSAX observation. We derive a 90\% confidence upper limit luminosity for a bremsstrahlung component with a temperature of 0.5~keV (assuming a distance of 60~kpc) of ${\rm 2\ 10^{35}\ erg\ s^{-1}}$. This upper limit is too large to compare with the luminosity of colliding winds observed e.g. by Jordan et al. (1994) in the galactic symbiotic nova RR~Tel of $\sim$${\rm 2\ 10^{32}\ erg\ s^{-1}}$. \begin{acknowledgements} This research was supported in part by the Netherlands Organization for Scientific Research (NWO) through Spinoza Grant 08-0 to E.P.J. van den Heuvel. The \sax\ satellite is a joint Italian--Dutch programme. We thank the staff of the \sax\ Science Data Center for their support. \end{acknowledgements}
\section{Introduction} The dynamics of first order phase transitions is a fundamental ingredient in particle physics and in condensed matter. First order phase transitions occur via the nucleation of bubbles of the true vacuum phase in the metastable or false vacuum phase. At large temperature it is mediated by thermal or overbarrier activation and at low temperatures via quantum nucleation. First order phase transitions are conjectured to occur in QCD and in the Electroweak theory. In QCD a first order phase transition {\em could} describe the hadronization of the quark-gluon plasma, possibly produced in the early Universe at about $10^{-5}$ seconds after the Big Bang or in relativistic heavy ion collisions\cite{shuryak,meyer,kapu}. In the Electroweak theory a first order phase transition is argued to provide the non-equilibrium setting for baryogenesis\cite{cohen,ruba,dolgov}. In early universe cosmology, first order phase transitions had been proposed as mechanisms to generate the inflationary stage\cite{linde,kolb,brand}. In condensed matter physics thermal activation results in the nucleation of bubbles of the lowest free energy phase in binary fluids\cite{miguel,langercond} and also of the decay of metastable dimerized states in quasi-one dimensional charge density wave systems and non-degenerate organic conductors\cite{yulu,boyancondbub}. The most comprehensive microscopic theory of nucleation via thermal activation was presented by Langer\cite{langer} and was later extended to quantum field theory to account both for thermal activation as well as for quantum nucleation\cite{coleman,stone,affleck,linde2}. An approach to describe nucleation in a non-steady state situation in real time has been advocated in\cite{sphal}. In the limit in which nucleation is dominated by thermal activation and overbarrier transitions, these bubbles are produced via large thermal fluctuations. These bubbles will grow whenever their radius is larger than a critical value and collapse if it is smaller. Supercritical bubbles will grow to convert the metastable phase into the stable phase or until they percolate achieving the total conversion of the metastable phase. For slightly supercritical bubbles an important {\em dynamical} quantity is the growth rate of a bubble \begin{equation} \Omega = \frac{d}{dt}\left[\ln\left|\frac{R(t)}{R_c}-1\right|\right] ; ~~ R(t) \geq R_c \end{equation} \noindent with $R(t)$ the (time dependent) radius of a slightly supercritical nucleating bubble and $R_c$ is the critical radius\cite{langer,kapu,miguel}. Langer's theory provides the nucleation rate per unit volume per unit time given by \begin{equation} I= \Omega ~ {\cal D}~ e^{-\frac{F_{b}}{T}} \end{equation} \noindent where $F_b$ is the free energy of a critical bubble and ${\cal D}$ is proportional to the ratio of the determinants of the fluctuation operators around the bubble configuration to that around the homogeneous metastable state\cite{langer,kapu,miguel}. The regime of applicability of homogeneous nucleation theory is for $F_b >> T$, for $F_b \approx T$ small amplitude thermal fluctuations can trigger the phase transition and nucleation and spinodal decomposition can no longer be distinguished. The ratio of determinants in ${\cal D}$ has been computed analytically and numerically in several important cases\cite{maclerran,baacke,ramos}. Csernai and Kapusta\cite{kapu} have studied the growth rate of hadronic bubbles in a quark gas by extending Langer's theory to the relativistic case. These authors studied a coarse-grained field theory in terms of a relativistic hydrodynamic description with viscous terms in the energy momentum tensor and the baryon current. Their conclusion was that the growth rate is determined by the coefficients of the shear and bulk viscosities and that hadronic bubbles do not grow when these coefficients vanish. Another approach presented by Ruggeri and Friedman\cite{ruggeri} also based on baryon-free relativistic hydrodynamics but with a different treatment of the heat conduction and energy flow reached a different conclusion, th at the growth rate is non-zero even for vanishing bulk and shear viscosities and that viscosity effects are subleading for small viscosities. If the hadronization phase transition is of first order there are potentially important experimental signatures associated with the homogeneous nucleation of the hadronic phase some of which had been studied in reference\cite{kapusta2}. Homogeneous nucleation of the quark-gluon plasma has been recently studied with a bag model of the equation of state for the quark and hadron phases and the different predictions above have been compared\cite{zabrodin}. The formulation and results of Csernai and Kapusta had recently been used to study first order quark-hadron phase transition in the Early Universe for a first order hadronization transition\cite{chandra} and more recently homogeneous nucleation has been tested numerically in $2+1$ dimensional systems with qualitative agreement to the standard result\cite{tetradis}. An alternative phenomenological description of the growth rate based on dissipative hydrodynamics combined with the finite temperature effective potential has been presented in \cite{ignatius}. In this work analytical and numerical studies reveal a dependence of the growth rate on the phenomenological dissipative coefficient. Khlebnikov\cite{kleb} and Arnold\cite{arnold} have studied dissipative effects that slow down the growth rate of a supercritical bubble by coupling the order parameter that describes nucleation to {\em other fields} (with a trilinear coupling) and applie d the fluctuation dissipation relation to one loop order in the coupling to the other field. Finally we must mention a numerical approach to the description of nucleation, based on a phenomenological {\em Markovian} Langevin dynamics in terms of the finite temperature effective potential (typically computed to one loop order) with a (local) friction coefficient and a Gaussian white noise term related by the fluctuation-dissipation theorem\cite{gleiser,haas}. The use of the finite temperature effective potential in the description of the inhomogeneous bubble configuration that seeds the nucleation of the phase of lowest free energy is an important ingredient, in particular in coarse-grained phenomenological descriptions\cite{kapu}. Finite temperature effects are included in the coarse grained free energy which describes long wavelength physics by integrating o ut short wavelength fluctuations which are in thermodynamic equilibrium\cite{liu,kunz}. These finite temperature corrections modify quantitatively and sometimes substantially the equilibrium free energy functional, for example the position of the minima, masses and couplings. These are important parameters that determine the profile of the bubble configuration, since the asymptotic behavior of the bubble is determined by the position of the minima of the equilibrium free energy. The importance of the growth rate for describing the dynamics of nucleation in the hadronization and the Electroweak phase transitions as well as the practical importance of nucleation in quasi-one dimensional organic conductors justifies a continued effort to understand from a {\em microscopic} perspective the influence of dissipation, viscosity and friction upon the growth rate of supercritical nucleating bubbles. {\bf Focus and Strategy:} Our goal in this article is precisely to study dissipative effects upon the growth rate $\Omega$ from a more {\em microscopic} point of view in model quantum field theories that describe the main features of nucleation. The simplest model to study nucleation and a first order phase transition is a $\phi^4$ field theory with a small explicit symmetry breaking term that breaks the degeneracy between ground states and thus leads to the existence of a metastable state. Although arguably this model could hardly describe the features of the quark-hadron or electroweak phase transitions, our hope is to extract robust phenomena that will be generic to the physics of nucleation and that could enlighten the effect of viscosity on the growth rate. The study begins by identifying the inhomogenous bubble configuration which is a solution of the static equations of motion. We include the finite temperature effects by considering the equation of motion in terms of the {\em finite temperature effective potential}, this is achieved consistently by the addition of finite temperature counterterms to the Lagrangian. The quadratic fluctuations around this bubble solution feature an unstable mode that describes small perturbations from the critical bubble. The instability is a manifestation of the growth (or collapse) of supercritical (or subcritical) bubbles and the unstable eigenvalue is directly related to the growth rate of slightly supercritical bubbles. Besides this unstable mode there are modes of zero frequency corresponding to translations of this configuration and modes of positive frequency that correspond to stable fluctuations. The main strategy is to consider the dynamics of the coordinate that determines the unstable direction in functional space, and treating it as a {\em collective coordinate}. We then focus on the {\em interaction} of this coordinate with those representing the stable fluctuations by expanding the original Lagrangian in terms of these coordinates in function space and recognize the interaction terms between these coordinates. The translational mode is anchored and ignored since we are not interested in studying the dynamics of the translational degree of freedom but that of the growth of a supercritical bubble. We assume that the stable degrees of freedom are in thermal equilibrium and that they serve as a ``bath'' with which the unstable coordinate interacts. Integrating out the degrees of freedom associated with the stable fluctuations we obtain an effective description of the dynamics for the unstable coordinate which includes the ``viscosity and friction'' effects associated with the transfer of energy to the stable modes. This description allows us to obtain the equations of motion for the expectation value of the unstable coordinate wherefrom we can obtain the {\em corrections} to the growth rate from the interaction with the stable degrees of freedom. Furthermore, integrating out the stable degrees of freedom in the non-equilibrium functional allows us to extract the effective Langevin equation for the unstable coordinate and to obtain consistently the noise terms. We emphasize that this approach is {\em fundamentally different} from the previous studies described above. First of all, we are {\em not} computing the ratio of determinants that enter in the rate expression. This ratio of determinants only involves the quadratic fluctuations but {\em not} the interaction between the coordinates associated with the fluctuations. Our approach is different from that of references\cite{kleb,arnold}, in these references the coupling of the order parameter associated with the nucleating field was to {\em other independent fields}. Contrary to this we consider the coupling of the unstable coordinates to the stable fluctuations involving the {\em same} field. Our approach starts from the microscopic theory and does not rely on a hydrodynamic description, however as it will be seen later, long-wavelength hydrodynamic fluctuations associated with surface waves provide the largest contribution to viscosity effects in three spatial dimensions. In a very specific sense our program obtains the effective action for the degree of freedom that represents departures from a critical bubble by integrating out the degrees of freedom describing stable fluctuations, but of the {\em same field}. Our approach must necessarily rely on several different approximations: i) weak coupling to allow a perturbative expansion of the effective action and the real time equations of motion, ii) the thin wall approximation in which the critical radius is much larger than the width of the bubble wall. This approximation is necessary to be able to provide quantitative answers. Finally as it will be justified later the important fluctuations can be described by the {\em classical} limit of the finite temperature distribution functions. We carry out this program to lowest order in the quartic coupling and to leading order in the thin wall approximation both in $1+1$ and $3+1$ dimensions. The motivation to study the lower dimensional case is provided by its potential application in experimentally relevant condensed matter systems\cite{yulu,boyancondbub}. {\bf Summary of results:} Our main results can be summarized as a consistent perturbative correction to the growth rate from ``dynamical viscosity'' effects associated with the interaction of the radius of the bubble with the stable quantum and thermal fluctuations. In $1+1$ dimensions the growth rate for bubbles is given by eqn. (\ref{estimate}) with $R_c,\xi$ the radius and width of the critical bubble respectively, $\lambda$ the quartic self-coupling and $F[R_c/\xi]$ a slowly varying dimensionless function that depends on the details of the potential and spectrum of stable fluctuations, in particular scattering states. In $3+1$ dimensions we find that in the thin wall limit the important low energy excitations are {\em hydrodynamic fluctuations} associated with surface waves. These are identified as {\em quasi-Goldstone} bosons and give the dominant contribution to viscosity effects on the growth rate. These low energy excitations are a robust feature of {\em any} model and not particular to the one considered in this article, however, the {\em coupling} of the unstable radial coordinate to these fluctuations depends on the model. To lowest order in the coupling the growth rate is given by eqn. (\ref{finalgrowthrate}). Furthermore, we obtain the effective Langevin equation for the unstable coordinate that reveals the correct microscopic correlations of the noise term showing that the coupling to the hydrodynamic modes determine that the noise is correlated on time scales $\Omega^{-1}$. The article is organized as follows: section II introduces the model, describes the strategy in detail and sets up the general form of the correction to the growth rate to lowest order (one loop) and in the classical limit. In section III the case of $1+1$ dimensions is analyzed in detail within the model considered. In section IV we study the $3+1$ dimensional case, and discuss the low energy fluctuations associated with surface waves, providing the argument that these are quasi-Goldstone hydrodynamic fluctuations. We argue that these low energy fluctuations are present independent of the model and will dominate the viscosity corrections to the growth rate. The conclusions are presented in section V, along with a critical discussion on the validity of our results to describe the growth rate of supercritical bubbles in coarse-grained descriptions of a first order quark-hadron phase transition. An appendix presents the effective Langevin equation for small departures of the critical radius and analyzes the generalized fluctuation-dissipation relation. \section{The model: generalities and strategy} We consider a real scalar field, $\phi(x)$, whose dynamics is determined by the following Lagrangian density \begin{equation} \label{lagrangeana} {\cal L}={1\over 2}(\partial_\mu\phi)(\partial^\mu\phi)- V(\phi)\label{model}, \end{equation} where $V(\phi)$ is a double well potential with a metastable minimum at $\phi_-$ and a stable minimum at $\phi_+$. It proves convenient to parametrize it in the following form: \begin{equation} V(\phi)=\lambda (\phi-\phi_-)^2 \phi(\phi-\phi^*). \label{pot} \end{equation} This form of the potential is depicted in fig.(\ref{potfig}) and is very similar to the coarse-grained free energy proposed in\cite{kapu} identifying the local energy variable {\em e} in that reference with the scalar field $\phi$, and also similar to the effective potential description used in numerical simulations for the Electroweak theory\cite{gleiser,haas,kunz}. The stable minimum $\phi_+$ is given by \begin{equation} \phi_+ \,=\, \frac{1}{8} \left( 2 \phi_- + 3 \phi_* - \sqrt{4 \phi_-^2 + 9 \phi_*^2 -4 \phi_- \phi_*} \right). \end{equation} where the point $\phi_*$ is parametrized by the mass $m$ of small oscillations around the {\em metastable} minimum $\phi_-$ as \begin{equation} \phi^*\equiv\phi_-\left(1-\epsilon\right) \end{equation} where \begin{equation} \epsilon \equiv {m^2\over 2\lambda\phi_-^2} \;\;\;\; \mbox{and} \;\;\;\; m^2 \equiv V''(\phi_-) \label{epsln} \end{equation} and $\lambda$ is the quartic coupling. Although this parametrization does not look familiar it will prove advantageous later. In particular, $(1-\epsilon)$ gives a measure of the free energy difference between the true and the false vacua and is therefore related to the supercooling temperature. As $\epsilon \rightarrow 1$, the two minima become degenerate and the condition $\epsilon\approx 1$ defines the thin-wall limit (this will become clear later when we analyze the bubble profile), when the radius of the nucleating bubble is much larger than the width of its surface $\xi$ (or the correlation length). This corresponds to the case of small supercooling $T<T_c$ but $T/T_c \approx 1$ in the description of reference\cite{kapu}. In the thin wall limit, we find that $\phi_+$ has the following simple form \begin{equation} \phi_+ \,=\, -\,\frac{\phi_-}{2}(\epsilon-1) + {\cal O}((\epsilon-1)^2). \label{phipls} \end{equation} \subsection{The counterterms:} We anticipate that there will be renormalization of the parameters in the potential, not only arising from ultraviolet divergences but more importantly from medium effects, i.e. finite temperature contributions to the mass, couplings and explicit symmetry breaking terms. In particular in three spatial dimensions we expect a correction to the mass and the linear symmetry breaking term both proportional to $T^2,T,\cdots$. The origin of these finite temperature corrections are the usual tadpole diagrams, the correction to the linear symmetry breaking term is a consequence of the trilinear coupling. These finite temperature corrections are the usual ones obtained in an effective potential description and provide a finite renormalization to the potential. Our aim is to describe the dynamical aspects of the thermal fluctuations that are responsible for thermal activation and the nucleation of the true phase in the false vacuum phase. These are the solutions of the static field equations but {\em in presence of the thermal fluctuations}, i.e. with the potential renormalized by the finite temperature effects. That is to say, the potential that must enter in the field equations must be the finite temperature effective potential\cite{liu,gleiser,kunz,tetradis,haas}. The finite temperature effective potential includes the finite temperature corrections for static and homogeneous field configurations. Including these finite temperature effects in the equations o f motion guarantees that the {\em inhomogeneous} bubble configurations will asymptotically tend to the homogeneous values of the field representing the {\em correct} extrema of the equilibrium free energy. Furthermore, these finite temperature renormalizations will determine the {\em correct} scales for the bubble size, its width and the unstable frequency associated with the growth of a supercritical bubble. We account for these finite temperature renormalizations by replacing the parameters in the potential by the finite temperature renormalized parameters and adding counterterms to the Lagrangian. The counterterms are then found consistently in a coupling or loop expansion by requesting that they cancel the contributions from the tadpole (or similar) terms {\em completely}, i.e. including the finite temperature parts. This prescription is at the heart of using the finite temperature effective potential to study the solutions that lead to the decay of the metastable state\cite{kapu,liu,gleiser,kunz,tetradis,haas}. Thus the Lagrangian density becomes \begin{equation} {\cal L}[\phi] = {1\over 2}(\partial_\mu\phi)(\partial^\mu\phi)- V(\phi,T)+\delta {\cal L}_{ct}[\phi] \label{counter} \end{equation} where the potential $V(\phi,T)$ is of the same form as in (\ref{pot}) but in terms of the parameters that include the finite temperature (and ultraviolet) renormalizations, i.e. $\lambda(T), m(T),\epsilon(T)$ etc. and the counterterm Lagrangian density is of the form \begin{equation} \delta {\cal L}_{ct} = \delta \lambda \phi^4 + \delta g \phi^3+ \delta M^2 \phi^2 + H\phi \label{countertermlagra} \end{equation} where space-time translational invariance dictates that the counterterms are constant. These counterterms will be consistently chosen in the loop expansion to cancel the contribution from local tadpoles. We will argue below that this procedure renormalizes consistently in perturbation theory the {\em static} properties of the bubble, in particular the width and radius of a critical bubble. In particular the free energy density studied in\cite{kapu,gleiser,haas,kunz,liu} is precisely of this form with parameters that depend on temperature. At this point we may consider adding to the counterterms a wave-function renormalization which would involve corrections for inhomogeneous configurations. A wave-function renormalization will arise at one loop order and beyond because of the trilinear coupling. However, computing this correction in the usual loop expansion around a homogeneous background is not very relevant for the bubble configuration. Since our goal is to obtain the {\em real time} equations of motion for particular fluctuations around the bubble configuration, these renormalizations will arise automatically in the equations of motion obtained to one loop order. Thus the counterterms that we include in the Lagrangian density are only those for {\em static} renormalization thus accounting for the correct scales and asymptotic values of the classical solution and obtain the equivalent of the wave-function renormalization directly from the real time evolution of fluctuations around the bubble configuration. This procedure will become clear when we obtain explicitly the evolution equations in the next sections. We will focus our attention on the cases of one and three spatial dimensions. Although the case of one spatial dimension is of limited interest in particle physics, it is important in condensed matter and statistical physics of quasi-one dimensional systems\cite{yulu,boyancondbub}. Furthermore most aspects of the treatment are general and the one-dimensional case provides a somewhat simpler setting to introduce the main strategy as well as to explore the general features. The three dimensional case will be studied in detail subsequently. \subsection{The bubbles:} Before proceeding to the specific situations, we now consider general features of the bubble solutions and the quadratic fluctuations around them to focus on the precise strategy to follow. A critical bubble is a solution of the {\em static} field equations, which in terms of the (effective) potential $V(\phi,T)$ is given in one spatial dimension by \begin{equation} \left. -\frac{d^2\phi_b}{dx^2}+\frac{\partial V(\phi,T)}{\partial\phi}\right|_{\phi_b}\,=\,0 \label{1Dbubble} \end{equation} \noindent with boundary conditions such that $ \phi_{b} \rightarrow \phi_- $ for $\vert x \vert \rightarrow \infty $. Such a solution describes a ``bubble'' like configuration which approaches the false vacuum at asymptotically large distances and probes the true vacuum in a localized region in space of size $2R_c$ with $R_c$ the critical ``radius''. Such a solution will be found in detail in section III. In three spatial dimensions, the critical bubble is a radially symmetric static field configuration solution of the following equation \begin{equation} \left. - \frac{d^2 \phi_{b}}{dr^2} - \frac{2}{r} \frac{d \phi_{b}}{dr} + \frac{ \partial V(\phi,T)}{\partial\phi}\right|_{\phi_b}\,=\,0 \label{3Dbubble} \end{equation} \noindent with the boundary condition $\phi_{b}(r) \rightarrow \phi_-$ as $r \rightarrow \infty$. It corresponds to a field configuration that starts close to the true vacuum, $\phi_+$, and tends to the false vacuum, $\phi_-$, at asymptotically large radial distance. The solution of (\ref{3Dbubble}) will be studied in detail in the thin wall approximation in section IV. In both cases the change from the true vacuum to the false one occurs around the radius of the bubble, $R_c$, over a distance $\xi(T) \sim 1/m$ that defines the wall thickness of the bubble and is related to the correlation length in the metastable phase. We will study the non-equilibrium dynamical viscosity of nucleating bubbles in the thin wall limit in which the radius of the bubble is much larger than the wall thickness; i.e. $R \gg \xi$. These field configurations are parametrized by two important coordinates: $\vec{x}_0$ which describes the position of the center of the bubble and the radius $R$, i.e. a bubble configuration is of the form $\phi(\vec x - \vec{x}_0;R)$ and a critical bubble corresponds to $R=R_c$ determined by the solution to the equations of motion (\ref{1Dbubble},\ref{3Dbubble}). The coordinate ${\vec x}_0$ is associated with translational invariance and typically treated as a collective coordinate\cite{rajaraman}, while the radius $R$ determines the size of the bubble and will be treated also as a collective coordinate (see below). Since we are not concerned here with the dynamics of the translational degrees of freedom, but rather with that of the radius, we ``clamp'' the collective coordinate $\vec{x}_0$ by fixing the center of the bubble at $\vec{x}_0=0$. Integration over this collective coordinate results in the typical volume factor\cite{langer,coleman} and is not relevant for our discussion. In $d$ space dimensions, we can use the radius of the bubble $R$ as a variational parameter and introduce the total energy of a bubble configuration with radius $R$ by \begin{equation} E_{var}(R) = \int d^d x\;\left[\frac{1}{2}\left(\nabla \phi_{b}(\vec x,R) \right)^2 + V(\phi_{b}(\vec x, R))\right]. \label{bubenergy} \end{equation} and in the limit $R/\xi(T)>>1$ (thin wall) has two main contributions: a volume contribution proportional to $R^d$ which is negative and arises from the region of the bubble that probes the true vacuum which has lower (free) energy, and a surface term which is positive and arises from the region of the bubble corresponding to the wall, includes gradient terms and is therefore proportional to the area of the surface of the wall, $R^{d-1}$. In one spatial dimension the ``surface'' of the bubble corresponds to two points and the gradient energy saturates to a constant independent of the radius for large bubble radius. The general behavior of the total energy of a bubble configuration as a function of $s=R/\xi$ is depicted in fig. (\ref{e1dfig}) for one spatial dimension and in fig. (\ref{e3dfig}) for three spatial dimensions (see sections III and IV). The maximum of the energy function determines the value of the critical radius $R_c$ as discussed in detail in the next sections. Clearly a critical bubble is an {\em unstable} static solution of the equations of motion. For $R<R_c$ the surface energy term dominates and the bubble shrinks into the false vacuum phase, for $R>R_c$ the volume energy dominates and the bubble will grow as the gain in the volume energy is greater than the cost in elastic surface energy. A more clear description is obtained in the case of three spatial dimensions for a spherically symmetric bubble. Let $\sigma$ be the surface tension (energy per unit area) of the spherical bubble configuration separating the metastable phase from the stable one. Then the total energy of the bubble is \begin{equation} E_{var}(R) = -\frac{4 \pi}{3} \Delta {\cal F} R^3 + 4 \pi \sigma R^2 + \cdots \label{classengy} \end{equation} where $\Delta {\cal F}= \left| V(\phi_+,T) - V(\phi_-,T) \right|$, and the dots stand for corrections that are subleading in the thin wall approximation, see section IV (eqn. (\ref{energy3dbubb}) for details. These (small) corrections will be neglected for the arguments presented below. The energy attains its maximum $E_c$ at the critical radius \begin{equation} R_c = \frac{2 \sigma}{\Delta {\cal F}} \label{rc} \end{equation} and it is given by \begin{equation} E_c \equiv E_{var}(R_c)= \frac{4 \pi \sigma}{3} R_c^2 \,= \, \frac{16 \pi \sigma^3}{3 \left(\Delta {\cal F} \right)^2}. \label{ec} \end{equation} Near the maximum of the energy function, it can be written as an expansion in terms of the (small) departures from the critical radius as \begin{equation} E_{var}(R) = E_c - \frac{1}{2}\omega^2 (R-R_c)^2 + \cdots \label{energyofR} \end{equation} where the frequency \begin{equation} -\omega^2 \equiv \left. \frac{d^2 E_{var}(R)}{dR^2} \right|_{R_c} = - 8 \pi \sigma \label{wsecdv} \end{equation} is independent of the critical radius and $\Delta {\cal F}$ and as it will be seen in detail below, it is related to the growth rate of slightly supercritical bubbles. \subsection{Fluctuations} The study of the fluctuations around the classical bubble configuration begins by studying the spectrum of the fluctuation operator \begin{eqnarray} && \left[-\frac{d^2}{dx^2}+V''(\phi_b(x,R_c))\right] {\cal U}_n(x) = \omega^2_n {\cal U}_n(x) ~~\mbox{in 1D} \label{1dfluc}\\ && \left[-\vec{\nabla}^2+V''(\phi_b(\vec x,R_c))\right] {\cal U}_n(\vec x) = \omega^2_n {\cal U}_n(\vec x) ~~\mbox{in 3D} \label{3dfluc} \end{eqnarray} where the prime represents differentiation with respect to the field $\phi$. Taking a spatial derivative of the equation of motion satisfied by the bubble configuration it is straightforward to find that \begin{equation} {\cal U}_0(x,R_c) \propto \frac{d\phi_b(x,R_c)}{dx} \label{1dzeromode} \end{equation} is a solution of the one dimensional eigenvalue problem (\ref{1dfluc}) with $\omega^2_0=0$ and that \begin{equation} {\cal U}_0(\vec x,R_c) \propto \frac{\vec x}{|\vec x|}\frac{d\phi_b(r,R_c)}{dr} \label{3dzeromode} \end{equation} are eigenfunctions of the three dimensional fluctuation operator with zero eigenvalue. These are the zero modes arising from translational invariance\cite{langer,coleman,rajaraman}. In one dimension the zero mode is an odd function of $x$ since the bubble solution is even, therefore it has a node. Hence there must be another solution of the Schroedinger like operator (\ref{1dfluc}) that has no nodes and has a smaller and therefore {\em negative} eigenvalue. The form of the bubble solution in the thin wall approximation is that of a widely separated kink-antikink pair, as it will be shown in detail in the next section we find that the eigenfunction corresponding to the negative eigenvalue must be ${\cal U}_{-1} \propto d\phi_b(x,R)/dR |_{R_c}$. This can be understood simply from the fact that the zero mode associated with the bubble solution is the {\em antisymmetric} combination of the zero modes associated with the individual kinks\cite{rajaraman}, hence the eigenfunction with lower eigenvalue must be the {\em symmetric} combination and the form of the bubble profile immediately leads to the conclusion that is the derivative with respect to the radius (see section III, eqn. (\ref{1dunstmode})). In the three dimensional case, the zero modes (\ref{3dzeromode}) correspond to the angular momentum $l=1$ spherical harmonics, therefore there must be an $l=0$ (spherically symmetric) solution of the Schroedinger-like eigenvalue problem (\ref{3dfluc}) with a {\em negative eigenvalue}. In a later section we study in detail the bubble solution and the spectrum of fluctuations around it and conclude that the eigenfunction corresponding to the negative eigenvalue is ${\cal U}_{-1} \propto d\phi_b(r,R)/dR|_{R_c}$ in the thin wall approximation (see section IV, eqn. (\ref{spcquasi}) for ${\cal U}_{000}$). Therefore the respective spectra of the fluctuation operators are \begin{eqnarray} && {\cal U}_{-1}(\vec x,R_c) = \sqrt{N_{-1}}~ \left.\frac{d\phi(\vec x,R)}{dR}\right|_{R_c}, ~{\omega^2_{-1} = -\Omega^2 \; ; \; \Omega^2 >0 } \; \; ; \; \; {\cal U}_{0}(\vec x,R_c) = \sqrt{N_0}~ \vec{\nabla}\phi(\vec x,R_c), ~{\omega^2_0 = 0} \nonumber \\ && {\cal U}_{n>0}(\vec x,R_c), ~{\omega^2_n > 0} \label{spectrum} \end{eqnarray} with $N_{-1}; N_0$ normalization factors (chosen real). The classical bubble solution corresponds to a saddle point in functional space, the mode ${\cal U}_{-1}$ determines an unstable direction. The field operator can now be expanded in the complete basis of eigenfunctions of (\ref{1dfluc},\ref{3dfluc}) in either case and write in general \begin{equation} \phi(\vec x,t)=\phi_b(\vec x-\vec{x}_0,R_c)+ q_{-1}(t){\cal U}_{-1}(\vec x-\vec{x}_0)+ q_0(t){\cal U}_{0}(\vec x-\vec{x}_0)+\sum_{n>0} q_n(t){\cal U}_n(\vec x-\vec{x}_0), \label{fieldexpansion} \end{equation} Whereas the bound state eigenfunctions are chosen real, the scattering states are chosen to satisfy the hermiticity condition ${\cal U}_{n}(\vec x) = {\cal U}^*_{-n}(\vec x); ~~q_n(t)= q^*_{-n}(t)$. The quanta associated with the coordinates $q_{n>0}$ for the stable modes will be referred generically as ``mesons'', and the operators $q_{n>0}$ create and annihilate meson states. From (\ref{spectrum}) it is clear that the modes ${\cal U}_0 \; ; \; {\cal U}_{-1}$ correspond to shifts in the position of the bubble, $\vec{x}_0$ and the radius $R$ respectively. The collective coordinate treatment\cite{rajaraman} absorbs the zero mode ${\cal U}_0$ into a definition of a new quantum mechanical degree of freedom $\hat{\vec{x}}_0(t)$ and expands the field operator in terms of the directions perpendicular to this zero mode, \begin{equation} \phi(\vec x,t)=\phi_b(\vec x-\hat{\vec{x}}_0(t),R_c)+ q_{-1}(t){\cal U}_{-1}(\vec x-\hat{\vec{x}}_0(t))+ \sum_{n>0} q_n(t){\cal U}_n(\vec x-\hat{\vec{x}}_0(t))\label{expre} \end{equation} We are not interested in the dynamics of the translational collective coordinate, therefore we will ``clamp'' the position of the bubble and focus solely on the dynamics of the unstable coordinate $q_{-1}(t)$. Technically this is achieved by inserting a functional delta function for $\hat{\vec{x}}_0(t)$ in the path integral, the corresponding Jacobian from the change of field variables to the collective coordinate\cite{rajaraman} leads to the usual volume factor\cite{langer,coleman}, hence in what follows we will set $\hat{\vec{x}}_0(t)=0$. We note that for small amplitudes of the unstable mode we can write the expansion (\ref{expre}) as \begin{equation} \phi(\vec x,t)\approx \phi_b(\vec x,R(t))+ \sum_{n>0} q_n(t){\cal U}_n(\vec x) \; ; \; \; R(t)=R_c+q_{-1}(t)\sqrt{N_{-1}}\label{expre2} \end{equation} that allows us to identify $R(t)-R_c=q_{-1}(t)\sqrt{N_{-1}}$ and the coordinate $q_{-1}$ is associated with small departures from the critical radius. \subsection{The relation between $\Omega^2$ and $\omega^2$} At this stage we recognize that the fluctuation mode ${\cal U}_{-1}(\vec x)\propto d\phi_b(\vec x,R_c)/ dR_c$ is the (unstable) functional direction along which the bubble either grows or collapses, therefore the coordinate $q_{-1}(t) \propto (R(t)-R_c)$ describes the small fluctuations around the critical bubble. An immediate question is: what is the relation between the frequency $\omega^2$ in (\ref{energyofR},\ref{wsecdv}) and $\Omega^2$, the eigenvalue associated with ${\cal U}_{-1}$ (\ref{spectrum})? The answer to this important question is obtained by taking the second derivative of the total energy in eq. (\ref{bubenergy}) with respect to the radius of the bubble $R$, which results in the following equation \begin{equation} \frac{d^2 E_{var}(R)}{d R^2} \,=\, \int d^dr \, \left\{ \left[\vec{\nabla}\left(\frac{d\phi_b}{dR}\right)\right]^2+ \vec{\nabla}\phi_b.\vec{\nabla}\left(\frac{d^2\phi_b}{dR^2}\right) + \left(\frac{d^2\phi_b}{dR^2}\right) \frac{\partial V(\phi_b)}{\partial \phi} + \left(\frac{d\phi_b}{dR}\right)^2 \frac{\partial^2 V(\phi_b)}{\partial \phi^2}\right\} \label{d2ER} \end{equation} Integrating the first two terms in eq. (\ref{d2ER}) by parts using the boundary conditions for the bubble solution and eqs. (\ref{1Dbubble},\ref{3Dbubble},\ref{1dfluc},\ref{3dfluc}), which are evaluated at the critical radius $R_c$, we obtain the following relation \begin{equation} \left.\frac{d^2 E_{var}(R)}{d R^2}\right|_{R=R_c} \,=\, - \omega^2 \,=\, -\frac{\Omega^2}{N_{-1}} \label{omegarelation} \end{equation} in terms of the normalization factor of the unstable mode (see (\ref{spectrum})) which will be found for the one and three dimensional cases separately below. We note that the relation (\ref{expre2}) determines that in the absence of interactions with other degrees of freedom the growth rate of a supercritical bubble is given by \begin{equation} \frac{d}{dt}\left[\ln\left|\frac{R(t)}{R_c}-1\right|\right]= \Omega \label{growthrate} \end{equation} for small departures from the critical radius. Equation (\ref{omegarelation}) is very useful since it relates the second derivative of the total {\em classical} variational energy $E_{var}(R)$ (the energy of the classical bubble as a function of its radius) to the eigenfrequency associated with the {\em quantum} unstable mode. Usually it is a difficult task to find the spectrum of the quantum fluctuation operator, and consequently the frequency of the unstable mode. The above relation can be used to find the value of the unstable frequency, or the normalization factor $N_{-1}$ if the frequencies are known. \subsection{The strategy} Our goal is to study the effects of quantum and thermal fluctuations upon the {\em real time} evolution of a bubble whose radius is slightly larger than critical. This is achieved by treating the coordinate $q_{-1}(t)\propto (R(t)-R_c)$ as a {\em collective coordinate}, i.e. the radius of the bubble is treated as a fully quantum mechanical variable interacting with the other degrees of freedom corresponding to the stable fluctuations and described by the coordinates $q_n(t)$. The ``zero mode'' (translational degree of freedom) is clamped and frozen since we are only interested in describing the dynamics of the unstable mode. The interaction of the stable degrees of freedom with the collective coordinate describing the departure from the critical radius will introduce viscosity effects and slow down the growth of a supercritical bubble. This will result in a smaller growth rate and our aim is to precisely compute this viscosity effect on the growth rate for small departures from the critical radius. By including the finite temperature {\em static} counterterms in the Lagrangian and requesting that these be cancelled consistently in the perturbative expansion, we can isolate the static renormalization to the unstable frequency from the {\em dynamical} viscosity effects associated with the energy transfer to the stable modes. This program begins by expanding the field in terms of the unstable and stable coordinates (after clamping the translational mode) as \begin{equation} \phi(x,t)=\phi_b(x,R_c) + q_{-1}(t) {\cal U}_{-1}(x)+ \sum_{p} q_p(t) {\cal U}_p(x), \label{expansion} \end{equation} where the summation index $p$ runs over all stable, bound and scattering, states other than the translational and unstable modes. Throughout the index $p$ refers to both scattering and bound states. Using the above expansion, eq. (\ref{expansion}), of the field $\phi(x,t)$, the Lagrangian can be shown to have the following form: \begin{equation} L[q_{-1},q_p]\,=\,L_0[q_{-1}] + L_0[q_p] + L_I[q_{-1},q_p] \end{equation} where \begin{eqnarray} L_0[q_{-1}]&=&\frac{1}{2}~\left\{ \dot q_{-1}^2(t)+ \Omega^2 q_{-1}^2(t)+ \delta \Omega^2 q_{-1}^2(t) + hq_{-1}(t)\right\} \label{unlag} \\ L_0[q_{p}] &=& \frac{1}{2}~\sum_{p}~ \left\{ \dot{q}_p(t)\dot{q}_{-p}(t)-\omega_p^2 q_p(t)q_{-p}(t)\right\} \label{scatlag} \\ L_I[q_{-1},q_p] &=& - \sqrt{\lambda}~q_{-1}(t) \sum_{p,p'}{\cal B}_{pp'}q_p(t)q_{p'}(t) \, - \, q_{-1}^2(t) \left\{ \sqrt{\lambda} \sum_{p}{\cal B}_pq_p(t) + \lambda\sum_{p,p'}{\cal A}_{pp'} q_p(t)q_{p'}(t)\right\} \nonumber \\ && -\; q_{-1}^3(t) \,\left\{\sqrt{\lambda}\,{\cal B}_{-1} \,+\,\lambda\, \sum_{p} {\cal A}_pq_p(t) \right\} -\,\lambda \, {\cal A}_{-1}\,q_{-1}^4(t)+\mbox{h.o.t}+ L_{ct}[q_{-1},q_p], \label{lagrainte} \end{eqnarray} where we have used the equations of motion satisfied by the bubble configurations, and equations (\ref{1dfluc}) and (\ref{3dfluc}). The terms $\delta \Omega^2 q^2_{-1} \; ; \; h q_{-1}$ are the quadratic and linear terms in $q_{-1}$ from the counterterm Lagrangian. The terms denoted by h.o.t. in (\ref{lagrainte}) correspond to higher order interactions that are not important to the order that we are studying. The term $L_{ct}[q_{-1},q_p]$ arise from the non-linear (cubic and higher) terms in the coordinates from the counterterm Lagrangian, they do not need to be specified since they will be requested to cancel tadpole and static terms in the equations of motion for the collective coordinate $q_{-1}$ and do not contribute to lowest order (${\cal O}(\lambda)$). The matrix elements are given by \begin{eqnarray} {\cal B}_{pp'} & \equiv & \frac{1}{2\sqrt{\lambda}}\int_{-\infty}^{+\infty}~dx'~ {\cal U}_{-1}(x'){\cal U}_p(x'){\cal U}_{p'}(x')V'''(\phi_b(x',R_c)), \label{Bpp}\\ & & \nonumber \\ {\cal B}_p & \equiv & \frac{1}{2\sqrt{\lambda}}\int_{-\infty}^{+\infty}~dx'~ {\cal U}_{-1}^2(x'){\cal U}_p(x')V'''(\phi_b(x',R_c)), \label{Bp}\\ & & \nonumber \\ {\cal B}_{-1} & \equiv & \frac{1}{6\sqrt{\lambda}}\int_{-\infty}^{+\infty}~dx'~ {\cal U}_{-1}^3(x')V'''(\phi_b(x',R_c)), \label{B}\\ & & \nonumber \\ {\cal A}_{pp'} & \equiv & 6\int_{-\infty}^{+\infty}~dx'~ {\cal U}_{-1}^2(x'){\cal U}_p(x'){\cal U}_{p'}(x'), \label{Akk} \\ {\cal A}_{p} & \equiv & 4\int_{-\infty}^{+\infty}~dx'~ {\cal U}_{-1}^3(x'){\cal U}_p(x'), \label{Ak} \\ & & \nonumber \\ {\cal A}_{-1} & \equiv & \int_{-\infty}^{+\infty}~dx'~ {\cal U}_{-1}^4(x'), \label{A} \end{eqnarray} so that the $\lambda$-dependence is explicit in each term of the effective action. We thus see that the above Lagrangian describes a quantum mechanical degree of freedom $q_{-1}$ interacting with a bath of infinitely many degrees of freedom $q_p$, as well as with self-interaction. The self interaction of the coordinate $q_{-1}$ will be neglected because we are only interested in the viscosity effects arising from interaction with the stable degrees of freedom. The non-linear terms in $q_{-1}$ are neglegted because we will extract the growth rate for small departures from the critical bubble. The bath of mesons will introduce viscosity effects on evolution of the coordinate associated with departures from the critical radius and will result in a correction to the growth rate. These effects can be studied in a consistent manner by integrating out the meson degrees of freedom in a consistent perturbative expansion, thus obtaining the non-equilibrium effective action for $q_{-1}$. The variational derivative of this effective action leads to the equation of motion that includes the effects of the meson bath. This non-equilibrium effective action is obtained up to one loop order in appendix B. The effective equation of motion is a Langevin equation with a non-Markovian viscosity kernel and a Gaussian noise term (to one loop order) whose correlations are related to the viscosity kernel via the Fluctuation-Dissipation relation. The details are provided in appendix B. Alternatively we obtain the equation of motion for the expectation value of the coordinate $q_{-1}$ directly by means of linear response. We have already accounted for the {\em static} renormalization of the unstable frequency $\Omega$ by introducing the finite temperature counterterms and using the finite temperature effective potential in the equations of motion. However, we anticipate that there will arise further corrections to the growth rate $\Omega$ from {\em velocity dependent} terms in the effective equations of motion. These are truly {\em dynamical} viscosity effects which cannot be captured by a static calculation. The viscosity terms that are a function of the time derivative of the unstable coordinate will be revealed by obtaining the effective non-equilibrium equation of motion for this coordinate by integrating out the stable modes, which act as a bath for the quantum mechanical degree of freedom $q_{-1}$. We will restrict our study to small departures from the critical radius and obtain the linearized equation of motion for the expectation value of $q_{-1}$. This is consistent with the definition of the growth rate as the logarithmic derivative of the radius for small departures from the critical value. As is clear from (\ref{unlag}), in the absence of interactions the Hamiltonian for the unstable coordinate $q_{-1}$ is that of an {\em inverted harmonic oscillator} (of negative frequency squared) and small departures from the unstable equilibrium value $q_{-1} =0$ will grow exponentially with the growth rate $\Omega$. \subsection{The initial value problem:} We study the time evolution of the expectation value of the unstable coordinate $Q =\langle q_{-1} \rangle$ by proposing an initial state described by a density matrix which is a tensor product of the density matrix for the unstable coordinate and a density matrix that describes all the stable modes in thermal equilibrium at a given temperature. Therefore \begin{equation} \rho(0)= \rho_{-1}(0)\otimes \rho_s(0) \label{rhoinitial} \end{equation} Where $\rho_{-1}$ describes a pure state in which the expectation value of the unstable coordinate is $Q_0$, and $\rho_s(0)$ is a thermal density matrix for the stable modes. As mentioned above the translational mode is clamped. The time evolution of the initially prepared density matrix is given by Liouville's equation, whose solution is \begin{equation} \rho(t)= U(t)\rho(0)U^{-1}(t) \label{timeevoldens} \end{equation} with $U(t)$ being the unitary time evolution operator. As described in detail in reference\cite{boysinghlee} the time evolution can be cast in terms of a time dependent Hamiltonian in which the interaction is switched on at the initial time $t=0$. Alternatively the initial value problem can be cast in terms of linear response to an external source adiabatically turned-on from $t=-\infty$ that determines the initial preparation\cite{rgir}. Both approaches are equivalent and the reader is referred to\cite{boysinghlee,rgir} for more details on the initial value problem. The equation of motion in real time is obtained by using the generating functional of non-equilibrium Green's functions which requires a path integral along a contour in complex time and the following effective Lagrangian\cite{keldysh,nos}: \begin{equation} L_{eff}= L[q^+_{-1},q^+_p]-L[q^-_{-1},q^-_p] \end{equation} where the labels $\pm$ refer to the forward $(+)$ and backward $(-)$ branches along the complex time contour\cite{keldysh,nos}. The equation of motion for the expectation value $Q(t) = \left<q_{-1}(t)\right>$ is obtained by performing the shift \begin{equation} q^{\pm}_{-1}(t) = Q(t) + \tilde{q}^{\pm}(t) ~~;~~ \left<\tilde{q}^{\pm}(t)\right>=0 \label{shift} \end{equation} Imposing the condition $\left<\tilde{q}^{\pm}(t)\right>=0$ to all orders in perturbation theory leads to the retarded equation of motion for $Q(t)$. For a detailed presentation of this method in many other situations see\cite{nos}. The important ingredients in this program are the real time Green's functions for the stable coordinates which are assumed to be in thermal equilibrium. These are the following \begin{equation} \begin{array}{lclcl} \left< q^+_p(t)q^+_{p'}(t^\prime) \right> & =& -i\,\delta_{p,-{p'}} G^{++}_{p}(t,t^\prime) & = & -i \delta_{p,-{p'}} \left[{\cal G}^>_p(t,t')\Theta(t-t')+{\cal G}^<_p(t,t')\Theta(t'-t) \right] \\ & & & & \vspace{-2ex}\\ \left< q^-_p(t)q^-_{p'}(t^\prime) \right> & =& -i\,\delta_{p,-{p'}}G^{--}_{p}(t,t^\prime) & = & -i \delta_{p,-{p'}} \left[{\cal G}^>_p(t,t')\Theta(t'-t) + {\cal G}^<_p(t,t')\Theta(t-t') \right] \\ & & & & \vspace{-2ex}\\ \left< q^+_p(t)q^-_{p'}(t^\prime) \right> & =& i\,\delta_{p,-{p'}}G^{+-}_{p}(t,t^\prime) & = & -i\,\delta_{p,-{p'}}\,{\cal G}^<_p(t,t') \\ & & & & \vspace{-2ex}\\ \left< q^-_p(t)q^+_{p'}(t^\prime) \right> & =& i\,\delta_{p,-{p'}}G^{-+}_{p}(t,t^\prime) & = & -i\, \delta_{p,-{p'}} \, {\cal G}^>_{p}(t,t')= -i\, \delta_{p,-{p'}} \,{\cal G}^<_p(t',t), \end{array} \label{greens} \end{equation} where \begin{eqnarray} {\cal G}^>_p(t,t') & = & \frac{i}{2\omega_p} \bigg[ \left(1\,+\,n_p\right)\mbox{exp} \left\{-i\omega_p(t-t')\right\} \,+\,n_p\,\mbox{exp}\left\{i\omega_p(t-t')\right\} \bigg] \nonumber \\ {\cal G}^<_p(t,t') & = & \frac{i}{2\omega_p} \bigg[ \left(1\,+\,n_p\right)\mbox{exp} \left\{i\omega_p(t-t')\right\} \,+\,n_p\,\mbox{exp}\left\{-i\omega_p(t-t')\right\} \bigg]\nonumber \\ n_p & = & \frac{1}{ \mbox{exp}\left\{ \beta\omega_p \right\}\,-\,1 }. \label{ggreater} \end{eqnarray} We will carry our derivation of the equation of motion in the linear theory where we will neglect nonlinear terms, ${\cal O}(Q^2)$ and higher, and furthermore we will neglect the self-interaction of the unstable coordinate (cubic and quartic terms). As mentioned before, the linearization of the equation of motion is consistent with the focus on the dissipative corrections to the growth rate, which is defined for small departures from the critical radius. In this case, the non-equilibrium effective action $S_{eff}$ becomes \begin{eqnarray} \label{acao1} S_{eff} &=&\int_{-\infty}^{+\infty} dt \,\left\{L_0[\tilde{q}^+(t)] + L_0[q^+_p(t)]\,+\, \tilde{q}^+(t) \left[ -\ddot{Q}(t) +(\Omega^2 + \delta\Omega^2)Q(t) - \sqrt{\lambda}~\theta(t) \sum_{p,p'}{\cal B}_{p,p'}q_p^+(t)q_{p'}^+(t) \right. \right.\nonumber \\ && \left. \left. \,-\, 2\lambda ~ \theta(t) \sum_{p,p'}{\cal A}_{p_1 p_2} q_{p_1}^+(t)q_{p_2}^+(t) Q(t) \,+\, h \right] \;-\;\sqrt{\lambda}~\theta(t) \, Q(t) \sum_{p,p'}{\cal B}_{pp'} q_p^+(t) q_{p'}^+(t) \;-\; \left[ (+)\leftrightarrow (-) \right]\; \right\}, \nonumber \\ & & \end{eqnarray} where we have written only the terms that are relevant to the lowest order calculation ${\cal O}(\lambda)$ which is the focus of the present discussion and included the time dependence of the Hamiltonian to set up the initial value problem\cite{boysinghlee}. Imposing the condition $\langle \tilde q^+_{-1} (t)\rangle =0$, up to ${\cal O}(\lambda)$, we obtain the following equation to one-loop order \begin{eqnarray} &&i \int_{-\infty}^\infty dt' \left< \tilde q^+ (t) \tilde q^+ (t') \right> \left[\ddot{Q} (t') \, - \, \Omega^2 \, Q(t')- \Delta Q(t')-{\cal H} -2i\lambda \sum_{p_1,p_2,p_3,p_4}{\cal B}_{p_1 p_2}{\cal B}_{p_3 p_4} \times \right. \nonumber \\ && \left. \int_{-\infty}^{+\infty}~dt''~\bigg\{ \langle q_{p_1}^+(t')q_{p_3}^+(t'')\rangle \langle q_{p_2}^+(t')q_{p_4}^+(t'')\rangle - \langle q_{p_1}^+(t')q_{p_3}^-(t'')\rangle \langle q_{p_2}^+(t')q_{p_4}^-(t'')\rangle\bigg\} Q(t'')\right] \; = 0 \nonumber \end{eqnarray} with \begin{eqnarray} \Delta & = &\delta \Omega^2 -2\lambda \sum_{p,p'}{\cal A}_{p,p'}<q^+_p(t')q^+_{p'}(t')> \nonumber \\ {\cal H} & = & h - \sqrt{\lambda}\sum_{p,p'}{\cal B}_{p,p'}<q^+_p(t')q^+_{p'}(t')> \label{eqnmot} \end{eqnarray} Since the correlation $ \left< \tilde q^+ (t) \tilde q^+ (t') \right>$ does not vanish for all values of $t$ and $t'$, this implies that the quantity between the square brackets must vanish\cite{nos}. Furthermore, we now choose the counterterm in such a way to ensure that ${\cal H}=0$. We thus obtain the equation of motion for the expectation value of the unstable coordinate for $t>0$ \begin{equation} \ddot{Q} (t) \,-\, \Omega^2 \, Q(t)-\Delta Q(t) - \lambda \, \int_{0}^{t}~dt'~ \Sigma (t-t') Q(t')=0 ~~ ; ~~ \dot{Q}(t=0) =0 ; Q(t=0)=Q_0 \label{eqmv2} \end{equation} where the self energy $\Sigma (t-t')$ is given by \begin{eqnarray} \Sigma (t-t') & = & \sum_{p,p'} \frac{{\vert{\cal B}_{pp'}\vert}^2}{\omega_{p}\omega_{p'}} \left\{(1+n_p+n_{p'}) \sin[(\omega_p+\omega_{p'})(t-t')] \,-\, (n_p-n_{p'}) \sin[(\omega_p-\omega_{p'})(t-t')] \right\},\label{sigmat} \end{eqnarray} and we have used the fact that $ {\cal B}^*_{p,p'}\,=\,{\cal B}_{-p,-p'}$. The two different terms in the self energy, proportional to the sum and difference of frequencies respectively have a simple but important interpretation. The first term proportional to the sum of frequencies corresponds to the process in which the coordinate $q_{-1}$ ``decays'' into two mesons with probability $(1+n_p)(1+n_{p'})$ minus the ``recombination'' process with probability $n_p n_{p'}$. The second term proportional to the difference of frequencies originates in Landau damping and corresponds to the scattering of the unstable coordinate with mesons in the medium, with probability $(1+n_{p'})n_p$ minus the reverse process with probability $(1+n_{p})n_{p'}$\cite{weldon}. The counterterm $\delta \Omega^2$ will be chosen to cancel the tadpole contribution to $\Delta$ (see eqn. (\ref{eqnmot})) as well as the {\em static} contribution from the self-energy, since this contribution is a {\em static} renormalization of the unstable frequency associated with a stationary bubble. \subsection{Viscosity corrections to the growth rate} In order to obtain the influence of thermal and quantum fluctuations on the growth rate of the bubble, we must solve the equation of motion (\ref{eqmv2}). This is achieved via the Laplace transform. Introducing the Laplace transforms of $Q(t), \Sigma(t)$ as $\tilde{Q}(s), \tilde{\Sigma}(s)$ respectively with $s$ the Laplace variable. The Laplace transform of (\ref{eqmv2}) with the specified initial conditions is given by \begin{equation} \tilde{Q}(s)=\frac{s \, Q_0} {s^2-(\Omega^2+\Delta) -\lambda \tilde{\Sigma}(s)}, \label{eqsI} \end{equation} where $\tilde{\Sigma}(s)$ is given by \begin{equation} \tilde\Sigma(s) = \sum_{p,p'}\frac{|{\cal B}_{p,p'}|^2}{\omega_{p}\omega_{p'}} \left\{(1+n_p+n_{p'})\frac{\omega_p+\omega_{p'}}{s^2+(\omega_{p}+\omega_{p'})^2} \,-\, (n_p-n_{p'})\frac{\omega_p-\omega_{p'}}{s^2+(\omega_{p}-\omega_{p'})^2} \right\} \label{sigmas} \end{equation} We now isolate the static contribution by substracting \begin{equation} \tilde{\Sigma}(0) = lim_{s\rightarrow 0} \tilde{\Sigma}(s) \label{subs} \end{equation} from the Laplace transform of the self-energy. Using the identity \begin{equation} lim_{s\rightarrow 0} \frac{W}{s^2+W^2} = {\cal P}\left(\frac{1}{W}\right) \label{PP} \end{equation} we now fix the counterterm $\delta \Omega^2$ such that $\Delta + \lambda \tilde{\Sigma}(0)=0$. Introducing \begin{equation} \tilde\Gamma(s) = \sum_{p,p'}\frac{|{\cal B}_{p,p'}|^2}{\omega_{p}\omega_{p'}} \left\{ \frac{1+n_p+n_{p'}}{\omega_p+\omega_{p'}} \frac{s^2}{s^2+(\omega_{p}+\omega_{p'})^2} \,-\, (n_p-n_{p'}){\cal P} \left(\frac{1}{\omega_p-\omega_{p'}}\right) \frac{s^2}{s^2+(\omega_{p}-\omega_{p'})^2} \right\} \label{gamas} \end{equation} the Laplace transform of the equation of motion becomes \begin{equation} \tilde{Q}(s)=\frac{s\,Q_0} {s^2-\Omega^2 +\lambda \tilde{\Gamma}(s)}, \label{eqs} \end{equation} In order to avoid cluttering of notation we will not write explicitly the ${\cal P}$ in eqn. (\ref{gamas}) in what follows, but it must always be understood that the term $1/(\omega_p-\omega_{p'})$ actually refers to its principal part. This principal part prescription arises from the subtraction in the limit of vanishing $s$ which is the equivalent of the frequency in the real time domain, since the Laplace transform requires that $s\rightarrow i\omega +\epsilon$ with $\omega$ the frequency\cite{boysinghlee}. At this stage it becomes clear that the procedure of absorbing the local (zero frequency limit) contributions to the self energy in the {\em static} renormalization of the growth rate $\Omega$ provides the correct description of the dynamics. Viscosity and dissipative effects only arise from the time dependence of the expectation value of the coordinate and the {\em static} medium renormalization had already been absorbed into the definition of the growth rate as the limit of zero frequency. Viscosity and dissipation arise from the transfer of energy of the bubble wall to other excitations (mesons) and the growth slows down because of these processes. An important distinction arises in this case as compared to the familiar situation in field theory in which complex poles determine the renormalization to the mass from the {\em real} part of the self-energy on shell and the decay width from the {\em imaginary} part of the self-energy on shell. In this case, however, because the frequency associated with the unstable mode is {\em imaginary}, the real part of the self-energy will renormalize the {\em growth rate} whereas the imaginary part (if any) at the position of the (purely imaginary pole) will provide {\em oscillatory} contributions to the bubble dynamics. Therefore, we emphasize that {\em viscosity} effects that will diminish the growth rate are determined by the {\em real part of the self-energy} at the position of the pole. This is a striking difference from the usual case in which damping and viscosity are associated with the imaginary part of the self-energy on shell. The real time dynamics of the bubble growth, $Q(t)$, is given by the inverse Laplace transform \begin{equation} Q(t) = \frac{1}{2\pi i} \int_C e^{st} \; \tilde{Q}(s) \;ds \;, \label{contint} \end{equation} where $\tilde{Q}(s)$ is given by eq.(\ref{eqs}) and $C$ refers to the Bromwich contour running along the imaginary axis to the right of all the singularities of $\tilde{Q}(s)$ in the complex $s$-plane. The analytic structure of $\tilde{Q}(s)$ consists of cuts along the imaginary axis in the $s$-plane and poles. The two different processes of decay into meson pairs (and recombination) and Landau damping discussed above yield two different cut structures. The first term in eq. (\ref{gamas}) gives a two-meson cut and the contribution from Landau damping determines a cut structure that includes the origin in the s-plane. The structure of these cuts depend on the matrix elements of the interaction as well as the full spectrum of excitations and will be investigated in detail for particular cases in the next sections. The growth rate of the bubble with the quantum fluctuation effects included is given by the pole $s_p$ of $\tilde{Q}(s)$, eq. (\ref{eqs}), which satisfies the following relation \begin{equation} s_p^2\,-\,\Omega^2\,+\,\lambda \tilde{\Gamma}(s_p) \,=\, 0. \label{polecondition} \end{equation} To first order in $\lambda$, the pole of $\tilde{Q}(s)$ which corresponds to a growing bubble is given to lowest order (${\cal O}(\lambda)$) by \begin{equation} s_p = \pm(\Omega +\lambda\Omega^{(1)}) \end{equation} where the first order quantum and thermal fluctuation correction $\lambda\Omega^{(1)}$ to the bubble growth is given by \begin{eqnarray} \lambda \Omega^{(1)} &= & -\lambda \frac{\tilde{\Gamma}(\Omega)}{2\Omega} \nonumber \\ &=& - \frac{\lambda}{2} \sum_{p,p'}\frac{|{\cal B}_{p,p'}|^2}{\omega_{p}\omega_{p'}} \left\{\frac{(1+n_p+n_{p'})}{(\omega_p+\omega_{p'})}\left( \frac{\Omega}{\Omega^2+(\omega_{p}+ \omega_{p'})^2} \right) \right. \nonumber \\ & & \hspace{1.35in} - \left.\;\frac{n_p-n_{p'}}{\omega_{p} - \omega_{p'}}\left(\frac{\Omega}{\Omega^2+(\omega_{p}- \omega_{p'})^2} \right) \right\} \label{delomega} \end{eqnarray} We note that both terms inside bracket in (\ref{delomega}) are {\em positive} since $n_p < n_{p'}$ for $\omega_p > \omega_{p'}$, we recall again that the term $1/({\omega_{p} - \omega_{p'}})\equiv {\cal P}(1/({\omega_{p} - \omega_{p'}}))$ from the discussion following eqn. (\ref{gamas}). Therefore we conclude that the correction to the growth rate is {\em negative} i.e. the dissipative effects of the coupling to mesons results in a smaller growth rate of supercritical bubbles. Neither the frequencies of oscillations around the stationary bubble nor the matrix elements are quantum mechanical in origin. However the one loop contribution ${\cal O}(\lambda)$ to the self-energy is of quantum origin. If we restore $\hbar$ in the expression above this results in $\lambda \rightarrow \lambda \hbar \; ; \; T \rightarrow T/\hbar$. Obviously neither the eigenvalues of the fluctuation operator nor the matrix elements depend on $\hbar$ but the one loop contribution to the self-energy includes the $\hbar$ from the coupling (one loop) as well as from the temperature factors. \subsection{Classical Limit} In the next sections we will see that the correction to the growth rate is dominated by low lying excitations and for high temperatures these will be such that $\omega_p << T$. In this case we can invoke the classical limit which is best understood by restoring $\hbar$ as mentioned above \begin{equation} n_k=\frac{1}{e^{\hbar \beta\omega_k}-1} \approx \frac{T}{ \hbar \omega_k } >> 1 \end{equation} In this form, \begin{equation} n_k-n_{k'}\approx \frac{T}{\hbar} \left( \frac{1}{\omega_{k}}-\frac{1}{\omega_{k'}}\right) =-\frac{T}{\hbar} \frac{\omega_{k}-\omega_{k'}}{\omega_{k}\omega_{k'}} \end{equation} and in the high temperature limit we further approximate \begin{equation} 1+n_k+n_{k'}\approx 1+ \frac{T}{\hbar} \frac{\omega_{k}+\omega_{k'}}{\omega_{k}\omega_{k'}} \approx \frac{T}{\hbar} \frac{\omega_k+\omega_{k'}}{\omega_{k}\omega_{k'}} \end{equation} These approximations lead to a simplified expression for the lowest order correction to the growth rate \begin{eqnarray} \delta\Omega &=& - \frac{\lambda T}{2 } \sum_{p,p'}\frac{|{\cal B}_{p,p'}|^2}{\omega_{p}^2 \omega_{p'}^2} \left\{ \frac{\Omega}{\Omega^2+(\omega_{p}+ \omega_{p'})^2} + \frac{\Omega}{\Omega^2+(\omega_{p}- \omega_{p'})^2} \right\} \label{delomegac} \end{eqnarray} where the product $\lambda T$ is independent of $\hbar$ and displays clearly the classical limit. The classical limit will be justified in each particular case in the next sections. The expression (\ref{delomegac}) is the final form of the one loop ${\cal O}(\lambda)$ correction to the growth rate arising from {\em dynamical viscosity} since all of the static contributions had been absorbed by the counterterms. This is as far as we can pursue in a general manner without addressing the details of the spectrum of fluctuations around the bubble solution. In the next two sections we study the details of the model determined by the Lagrangian (\ref{model}) with the potential (\ref{pot}) for the cases of $1+1$ and $3+1$ dimensions. \section{The 1+1 Dimensional case} As mentioned in the introduction, the $1+1$ dimensional case is relevant in statistical and condensed matter physics. Quantum field theory models based on the Lagrangian (\ref{lagrangeana}) with potentials with a stable and metastable state are proposed to describe the low energy phenomenology of quasi-one dimensional charge density wave systems\cite{yulu}, and therefore their relevance in these physical situations warrants the study of this case. For $V(\phi)$ given by eq. (\ref{pot}), the solution to the static classical equation of motion (\ref{1Dbubble}) for one spatial dimension can be found exactly\cite{boyancondbub,sphal}. The critical bubble is found to be given by\cite{boyancondbub,sphal} \begin{equation} \label{bubble1d} \phi_b(x,s_0)=\phi_-+{m\over 2\sqrt{2\lambda}} \left\{\tanh\left[\frac{x}{\xi} +s_0\right]-\tanh \left[\frac{x}{\xi}-s_0\right]\right\} ~~;~~ \xi =\frac{2}{m}, \end{equation} where $\xi$ is the width of the bubble wall and $s_0$ is given in terms of the critical radius $R_c$ by \begin{eqnarray} s_0 \,\equiv\, \frac{R_c}{\xi} \,\equiv\, {1\over 2}{\rm cosh^{-1}} \left({\epsilon+1\over\epsilon-1}\right) \label{s0} \end{eqnarray} with $\epsilon$ given by eq. (\ref{epsln}). The above solution corresponds to a kink-antikink pair centered at $x = 0$, and separated by a distance $R_c$, and is displayed in fig. (\ref{critbub}). This is the one dimensional bubble that starts at the false vacuum $\phi_-$ almost reaches the true vacuum $\phi_+$ and returns to $\phi_-$ and it is similar to the polaron solution found in quasi-one-dimensional polymers \cite{boyancondbub}. The total energy of the bubble, given by eq. (\ref{bubenergy}), as a function of its dimensionless radius $s$ can be calculated from the field $\phi_b(x,s_0)$ by replacing $s_0=R_c/\xi \rightarrow s=R/\xi$ in eq. (\ref{bubble1d}). The gradient term $(d\phi_b/dx)^2$ contributes to the surface energy while $V(\phi_b)$ has two contributions; surface contribution and volume contribution. Substituting $V(\phi_b)$ in eq. (\ref{bubenergy}) and evaluating the integral, one finds that the total energy of the bubble is given by \begin{equation} E_{var}(s) \,=\, E_{sur}(s)\,+\,E_{vol}(s) \end{equation} with \begin{eqnarray} E_{sur}(s) & = & {\frac{-4\,{m^3}}{3\,\lambda }} + {\frac{4\,{m^3}\,s }{\left( -1 + {e^{4\,s}} \right) \,\lambda }} + {\frac{3\,{m^3}\,\left( -1 + {e^{8\,s}} - 8\,{e^{4\,s}}\,s \right) \, \left( 2 + \eta \right) }{4\,{{\left( -1 + {e^{4\,s}} \right) }^2}\, \lambda }} \nonumber \\ & & \; + \; {\frac{4\,{e^{4\,s}}\,{m^3}\, \left( 1 - {e^{4\,s}} + 2\,\left( 1 + {e^{4\,s}} \right) \,s \right) } {{{\left( -1 + {e^{4\,s}} \right) }^3}\,\lambda }} \\ E_{vol}(s) & = & -{ \frac{\eta \,{m^3} }{\lambda }}\,s \end{eqnarray} where $\eta$ is defined as \begin{equation} \eta \, \equiv \, \sqrt{\epsilon} \, +\, \frac{1}{\sqrt{\epsilon}} \,-\, 2. \label{eta} \end{equation} The volume contribution grows linearly with the radius of the bubble while the surface contribution saturates at about the critical radius of the bubble, $s_0$, and attains the following asymptotic value \[ \lim_{s \rightarrow \infty} E_{sur}(s) \rightarrow \frac{{m^3}\,\left( 2 + 9\, \eta \right) }{12\,\lambda }. \] In the thin wall limit ($\eta \rightarrow 0$), the surface energy is simply twice the kink mass\cite{rajaraman}, as the kink-antikink pair becames widely separated and the exponential interaction between kink and antikink becomes negligible. The total energy of the bubble is depicted in fig.(\ref{e1dfig}). It attains its maximum, $E_c$, at the critical radius $s_0$ where \begin{equation} E_c \,= \frac{m^3}{24\lambda}\left[ 4 + 12\,\eta + 3\,\eta^2 - 3\,s_0\,\eta \, \left( 8 + 6\,\eta + \eta^2 \right) \right] \end{equation} with \begin{eqnarray} s_0 \,=\, {1\over 2}{\rm cosh^{-1}} \left({\epsilon+1\over\epsilon-1}\right) \,=\, \frac{1}{4}\,\ln\left[\frac{4 + \eta}{\eta}\right] \end{eqnarray} In the thin wall limit, the above expressions simplify and we find \[ \eta \,\simeq \, \frac{(\epsilon-1)^2}{4} \,+\, {\cal O}((\epsilon-1)^3) \] and \begin{equation} E_c \,= \frac{m^3}{6 \lambda} \,+\, {\cal O}((\epsilon-1)^2) \end{equation} To study the fluctuations around the bubble solution we need the complete set of eigenfunctions $\{{\cal U}_n\}$ satisfying \begin{equation} \left[ -\frac{d^2}{dx^2}+V''(\phi_b)\right] {\cal U}_n(x)=\omega_n^2{\cal U}_n(x), \label{eigen} \end{equation} with \[ V''(\phi_b)\,=\,m^2-\frac{3\,m^2}{2}\mbox{sech}^2\left[\frac{x}{\xi}+s_0\right] - \frac{3\,m^2}{2}\mbox{sech}^2\left[\frac{x}{\xi}-s_0\right]. \] Although the spectrum of eigenfunctions and eigenvalues is known exactly in the case of one kink or antikink\cite{rajaraman}, for the case of the kink-antikink pair there are not known results that we are aware of. Solving for the eigenfunctions $\{{\cal U}_n\}$ in this case is a difficult task but in the thin wall limit, the above potential consists of two identical and widely separated wells centered at $x=\pm R_c$. The spectrum of each potential well is known in the literature\cite{rajaraman}. It consists of a zero frequency mode localized in the well, an excited bound state with frequency $3m^2/4$ that is also localized in the well and a continuum of scattering states. We can use approximate methods such as the linear combination of atomic orbitals approximation (LCAO) (available in elementary textbooks) to provide a reliable estimate for low lying bound states of the above potential from the spectrum of the single potential well. In this method, the ground state of the potential $V''(\phi_b)$ is the {\em symmetric} linear combination of the ground states of the single well located at $x=\pm R_c$ while the first excited state of $V''(\phi_b)$ is the {\em anti-symmetric} linear combination. In the thin wall limit, these two states are given by \begin{equation} {\cal U}_{-1}(x,s_0)=\frac{m}{2\sqrt{E_{c}}}\frac{d\phi_b} {ds_0}(x,s_0) \;\;\;\; \mbox{with} \;\;\;\; \omega^2_{-1}\,\equiv\, -\Omega^2 \simeq \,-24 m^2 e^{-4 s_0}\,=\,-6 \eta m^2 \label{1dunstmode} \end{equation} and \begin{equation} {\cal U}_0(x,s_0)=\frac{1}{\sqrt{E_{c}}}\frac{d\phi_b}{dx}(x,s_0) \;\;\;\; \mbox{with} \;\;\;\; \omega^2_{0}\,= \,0. \end{equation} These are the unstable, ${\cal U}_{-1}$, and the zero, ${\cal U}_{0}$, modes which we discussed earlier, see eqs. (\ref{1dzeromode}) and (\ref{spectrum}). Obviously ${\cal U}_0(x)$ is associated with translations since it is the spatial derivative of the bubble solution and must correspond to a vanishing eigenvalue by translational invariance. Since it is antisymmetric there has to be a nodeless eigenfunction of smaller frequency. The symmetric combination is ${\cal U}_{-1}(x)$ and since the two combinations will be split off in energy by a tunneling amplitude that is exponentially small in the distance between the kink and the antikink, the unstable frequency must be negative and exponentially small in this separation as is clearly displayed in eqn. (\ref{1dunstmode}). Besides the unstable and the zero modes, there are two bound states that have energies $\sim (3 m^2/4) \pm \Delta E(s_0)$ with energy difference which is again exponentially small in $s_0$ and correspond to the symmetric and antisymmetric linear combinations of the bound states of the kink in $\phi^4$ theory\cite{rajaraman} localized at $x=\pm R_c$. Since $V''(\phi_b)$ does not exactly reach $m^2$ near the center of the bubble, the potential in the Schroedinger equation could allow for shallow bound states near the scattering continuum with binding energies that are exponentially small in the variable $s_0$. These bound states if present are extremely difficult to obtain. Finally there is the scattering continuum region of the spectrum characterized by functions ${\cal U}_k$, whose eigenfunctions are asymptotically phase shifted plane waves with eigenvalues $\omega_k^2= k^2+m^2$ In order to compute the matrix elements that enter in the expression for the correction to the growth rate (\ref{delomegac}) we need either the exact form of the eigenfunctions or an excellent approximation to them. Whereas we are confident of our analysis regarding the low lying bound states, we lack a full understanding of shallow bound and continuum states. Such an understanding requires a detailed study of the spectrum which certainly lies beyond the scope of this article. Although we do not have a complete understanding of the spectrum of eigenfunctions and therefore we do not have even a good approximation to the matrix elements, we can, however, provide some physically reasonable assumptions complemented with dimensional arguments to provide an estimate for the corrections in this case. We begin by noting that in the one kink case, the potential that enters in the Schroedinger equation for the fluctuation is reflectionless\cite{rajaraman} and the scattering states for the one kink (or antikink) case have a transmission amplitude which is a pure phase. Thus we expect that in the thin wall limit when the kink-antikink pair separation is much larger than the width of an isolated kink the wave functions of the scattering states will acquire a phase shift that is at least twice as large as that in the one kink case and will have a rather smooth dependence on the kink-antikink separation. Because the potentials for each kink are reflectionless\cite{rajaraman} we {\em assume} that in the thin wall limit the reflection coefficients of each potential well are very small and therefore there are no substantial interference effects in the region between the kink and the antikink. Under these assumptions the matrix elements ${\cal B}_{p,p'}$ will be similar to those calculated in reference\cite{alamoudi2} for a single kink case and fall off very fast at large momenta justifying the classical limit\cite{alamoudi2}. Hence, under these suitable assumptions the matrix elements are smooth functions of $s_0=R_c/\xi$. The contribution from the two-meson cut, i.e. the first term in the bracket in (\ref{delomegac}) is proportional to $\Omega$ because the $\Omega$ in the denominator can be neglected in comparison with the frequencies for the meson states $\omega_p \approx {\cal O}(m)$. Since in $1+1$ dimensions the coupling $\lambda$ has dimensions of $(\mbox{mass})^2$ the correction to the growth rate arising from the two meson cut is of the form \begin{equation} \delta \Omega_{2mes} \approx -\frac{\lambda T}{m^3}\Omega F[s_0]\label{2meson} \end{equation} \noindent with $F[s_0]$ a dimensionless slowly varying function of $s_0=R_c/\xi$ which is rather difficult to calculate and can only be obtained from a detailed knowledge of the eigenfunctions. The contribution from the Landau damping cut is more complicated to extract. The second term in the bracket in (\ref{delomegac}) has the form of a Lorentzian and since $\Omega$ is exponentially small it is a function that is strongly peaked at $\omega_p = \omega_{p'}$ and the sum (integral) over $p,p'$ is dominated by a region of width $\Omega$ near $\omega_p=\omega_{p'}$. Assuming that the matrix elements are smooth functions of momentum, in this one dimensional case the integral over a small region $\omega_p-\omega_{p'} \approx \Omega$ can be done by taking a narrow Lorentzian and integrating over the relative momentum within this region\cite{note}, leading to a contribution of the form $-(\lambda T/m^2) C[s_0]$ where $C[s_0]$ is a smooth function of its argument that can only be calculated from a detailed knowledge of the scattering wavefunctions. Hence \begin{equation} \delta \Omega_{LD} \approx -(\frac{\lambda T}{m^2})C[s_0]\label{LD} \end{equation} and the total shift in the frequency is given by $\delta \Omega = \delta \Omega_{2mes}+\delta \Omega_{LD}$. Thus we conclude that although we do not have a complete knowledge of the eigenfunctions and therefore cannot provide a complete calculation of the correction to the growth rate, suitable assumptions based on the properties of the spectrum of the one kink case combined with dimensional arguments suggest that to lowest order in the coupling and in the thin wall approximation the growth rate of a slightly supercritical bubble is given by \begin{equation} \Omega \approx \frac{4\sqrt{6}}{\xi} e^{-\frac{2 R_c}{\xi}}\left[1-\frac{\lambda T\xi^3}{8} F[\frac{R_c}{\xi}]\right]- \frac{\lambda T^2 \xi^2}{4}C[\frac{R_c}{\xi}]+\cdots \label{estimate} \end{equation} \noindent with $F[\frac{R_c}{\xi}], C[\frac{R_c}{\xi}]$ slowly varying dimensionless functions of their argument. Obviously a full calculation of $F[\frac{R_c}{\xi}], C[\frac{R_c}{\xi}]$ requires a detailed understanding of the spectrum, in particular the scattering states. However it is clear from (\ref{estimate}) that the validity of a perturbative expansion places a severe constrain on the coupling constant and the value of the temperature, in particular the second contribution in (\ref{estimate}) arising from Landau damping gives the leading correction in the thin wall approximation and signals a potential breakdown of perturbation theory in this limit. A more detailed understanding of this possiblity requires a better knowledge of the scattering matrix elements, this is an extremely difficult problem that depends on the details of the potential and lies outside the scope of this article. \section{The 3+1 Dimensional case} \subsection{General aspects:} We now study the $3+1$ dimensional case which is more relevant from the point of view of particle physics, however before focusing on a particular form of the potential and bubble profile, we can study fundamental model independent properties of the $3+1$ dimensional case that will determine very robust predictions for the corrections to the bubble growth. The static bubble configuration $\phi_b(r,R_c)$ is radially symmetric and satisfies the static equation (\ref{3Dbubble}). To study quantum fluctuations around the critical bubble configuration, we need to find the spectrum of the fluctuation operator ${\cal M}$ which in 3 spatial dimensions is given by \begin{equation} {\cal M} \,=\left. \,-\nabla^2 \,+\, \frac{\partial^2 V(\phi)}{\partial\phi^2}\right|_{\phi_b(r,R_c)} \label{fluctopr} \end{equation} Since the critical bubble solution is radially symmetric, we write the eigenfunctions ${\cal U}_{nlm}(r,\theta,\varphi)$ of the differential operator ${\cal M}$ as a product of spherical harmonics $Y_{lm}(\theta,\varphi)$ and radial functions $\psi_{nl}(r)$ that satisfy \begin{equation} \left\{-\frac{d^2 }{d r^2} \,-\, \frac{2}{r} \frac{d}{dr} \,+\, \frac{l(l+1)}{r^2}\,+\, \frac{\partial^2 V(\phi_{b})}{\partial\phi^2}\right\}\psi_{nl}(r)\,=\, \omega^2_{n} \psi_{nl}(r) \label{eigen3d} \end{equation} Because of the translational invariance of the Lagrangian (\ref{lagrangeana}), there is a 3-fold degenerate zero mode given by the eigenfunction $\propto \nabla \phi_b(\vec x,R_c) \propto Y_{1;\pm 1,0} d\phi_b / dr$ which correspond to translations of the bubble in 3-dimensions with no energy cost. These are the Goldstone modes of the spontaneously broken translational invariance. This can be easily seen by taking the derivative of eq. (\ref{3Dbubble}) with respect to $r$ which results in the following equation \begin{equation} \left\{-\frac{d^2 }{d r^2} \,-\, \frac{2}{r} \frac{d}{dr} \,+\, \frac{2}{r^2}\,+\, \frac{\partial^2 V(\phi_{b})}{\partial\phi^2}\right\}\frac{d\phi_b}{ dr}\,=\,0 \label{zeromod} \end{equation} In addition to the Goldstone mode, there are other low-lying excitations corresponding to different values of $l\neq 1$. The eigenfunctions of these excitations and their eigenvalues can be obtained by writing the $l=1$ term in the above equation as \begin{equation} \frac{2}{r^2}\,=\, \frac{l(l+1)}{r^2}\,-\,\frac{(l-1)(l+2)}{r^2} \label{2overr}. \end{equation} and we re-write equation (\ref{zeromod}) in the following form \begin{eqnarray} &&\left\{-\frac{d^2 }{d r^2} \,-\, \frac{2}{r} \frac{d}{dr} \,+\, \frac{l(l+1)}{r^2}\,+\, \frac{\partial^2 V(\phi_{b})}{\partial\phi^2}\right\} \frac{d\phi_b}{dr}\,= \frac{(l-1)(l+2)}{R^2_c}\frac{d\phi_b}{dr}+ \delta V(r)\frac{d\phi_b}{dr} \label{lowener} \end{eqnarray} where \begin{eqnarray} && \delta V(r) = (l-1)(l+2)\left[\frac{1}{r^2}-\frac{1}{R^2_c} \right] \label{deltaV} \end{eqnarray} Since the function $d\phi_b/ dr$ is strongly localized at $r=R_c$ in the thin wall limit, the second term on the right hand side is a {\em small localized perturbation}. Therefore the unperturbed lowest lying eigenvalues are given by \begin{equation} \omega^2_{0l}\,=\, \frac{(l-1)(l+2)}{R_c^2}. \label{goldex} \end{equation} In the thin wall limit, $\xi/R_c <<1 $ we find that the lowest order correction (in $\delta V$) to these eigenvalues is of order ${\cal O} (\xi^2/ R_c^{2})$. For details see\cite{langer,wallace} and appendix \ref{correction}. This analysis reveals that there is a band of low lying modes with eigenfunctions \begin{equation} {\cal U}_{0lm}(r,\theta,\varphi) = \sqrt{N_{0l}} Y_{lm}(\theta,\varphi) \frac{d\phi_b(r,R_c)}{dr} \label{lolyin} \end{equation} with $N_{0l}$ the normalization constants, corresponding to the $2l+1$-fold degenerate eigenvalues given by (\ref{goldex}). Using equations (\ref{wsecdv},\ref{omegarelation},\ref{spectrum}) and (\ref{goldex}) with $l=0$ (corresponding to the unstable mode) we find the normalization to be given by \begin{equation} N_{0l} = \frac{1}{R^2_c\sigma} \label{norma} \end{equation} in terms of the critical radius and the surface tension. We summarize several noteworthy features of these low-lying solutions: \begin{itemize} \item {These excitations become Goldstone modes in the limit as $R_c \rightarrow \infty$, i.e. in the limit in which the radius of curvature of the bubble goes to infinity. This statement will be understood in detail below in connection with the case of flat interfaces in $3+1$ dimensions.} \item{The eigenfunction with the lowest eigenvalue, corresponds to $l=0$, i.e. a spherically symmetric solution with a negative eigenvalue given by\cite{langer,kapu,wallace} \begin{equation} \omega^2_{00}\,\equiv\, -\Omega^2\,=\, -\, \frac{2}{R_c^2}. \end{equation} This is the unstable mode which corresponds to a spherically symmetric expansion or contraction of the bubble and therefore corresponds to the unstable functional direction. The coordinate associated with this mode is the displacement from the critical radius.} \item{The three Goldstone modes ${\cal U}_{0,1,0}~;~{\cal U}_{0,1,\pm 1}$ are the translation modes.} \item{The higher energy modes with $l\geq 2$ are excitations on the surface of the bubble, or surface waves with energies given by \begin{equation} \omega^2_{0l} = \frac{(l-1)(l+2)}{R^2_c} = \frac{\Omega^2}{2}(l-1)(l+2) ; ~~ l \geq 2 \label{quasigold} \end{equation} .} \end{itemize} These low lying modes will play a dominant role and we will refer to them collectively as ${\cal U}_{0lm}$ with eigenvalues given by (\ref{goldex}) and whose normalized eigenfunctions (\ref{lolyin}) are simple functions of the bubble configuration which for the potential (\ref{pot}) in the thin wall approximation are given by eqn. (\ref{spcquasi}) below. That the modes with $l\geq2$ can be identified as wiggles of the bubble surface, or surface waves, can be seen from the following expansion\cite{wallace,safran} where as discussed before the translational modes is not included because it is ``clamped'' \begin{eqnarray} \phi(r,t) & = & \phi_b(r-R_c) + q_{-1}(t) {\cal U}_{000}(r,\theta,\varphi) + \sum_{l\geq 2;m} a_{lm}(t) Y_{lm}(\theta,\varphi) \; \left.\frac{d\phi_b}{dr}\right|_{R_c} + \cdots \nonumber \\ & \simeq & \phi_b(r-R(\theta,\varphi,t)) + q_{-1}(t) {\cal U}_{000}(r,\theta,\varphi) + \cdots\label{surfaceflucs} \end{eqnarray} where \[ R(\theta,\varphi,t) = R_c - \sum_{l\geq 2;m} a_{lm}(t) Y_{lm}(\theta,\varphi) \] This is an important identification that we emphasize: {\em this band of low-lying modes describes fluctuations of the surface of the bubble}. We will argue below that these excitations dominate the infrared behavior of the viscosity correction and will provide the largest contribution to the viscosity coefficient. The low lying spectrum described above is fairly general and only depends on the existence of a thin wall bubble. Depending on the form of potential $V(\phi)$, there might be other bound states. An analysis similar to that leading to the band of low-lying excitations reveals that for the potential (\ref{pot}) there is another band of rotational bound state excitations that starts near $\omega^2_{10} \simeq 3m^2/4$. In the thin wall limit the radial wave function for the lowest state in this band is given by the bound state of energy $3m^2/4$ of the $\phi^4$ theory in $1+1$ dimensions\cite{rajaraman} which is also localized at the wall. The eigenfunctions and eigenvalues for this rotational band of excitations are given by (see appendix A) \begin{equation} {\cal U}_{1lm}(r,\theta,\varphi) = \sqrt{N_{1l}} Y_{lm}(\theta,\varphi) \psi_1(r)~ ; ~~~ \omega^2_{1l}\approx \frac{3m^2}{4}+\frac{l(l+1)}{R^2_c}+{\cal O}(\frac{\xi}{R_c}) \label{nextband} \end{equation} \noindent where $\psi_1(r)$ is given by the bound state of the theory with potential (\ref{pot}) in one spatial dimension\cite{rajaraman}, the eigenfunctions in this band are given in eq. (\ref{nextbs}) below. Finally there is a continuum of scattering eigenstates with eigenvalues $\omega_k^2 = k^2 + m^2$. As will be discussed later, the contribution to the growth rate from the rotational band (\ref{nextband}) and of the scattering states is subleading in the thin wall limit. The maximum value of angular momentum available for the low-lying part of the spectrum (\ref{goldex}) is limited by the edge of the continuum spectrum or the presence of higher bound states, hence $l^2_{max}/R^2_c \leq m^2$ or $l^2_{max} \leq (mR_c)^2 = (R_c/\xi)^2$. Therefore in the thin wall limit the maximum value of the angular momentum $l_{max}>>1$. \subsection{Planar interfaces, surface waves and (quasi) Goldstone bosons} The low lying spectrum of eigenfunctions given by (\ref{lolyin}) with eigenvalues (\ref{goldex}) has a simple physical origin, which can be understood by noticing that in the limit $R_c \rightarrow \infty$ the discrete spectrum becomes a continuum. In the limit when the radius of the critical bubble is very large $R_c \rightarrow \infty$, the interface between the two phases $\phi_+$ and $\phi_-$ becomes planar and the two phases become degenerate. Let us consider a static planar interface configuration corresponding to a domain wall along the z-axis in three spatial dimensions. Such a configuration satisfies the following equation \cite{rajaraman,wallace} \begin{equation} - \frac{d^2 \phi_{w}(z)}{dz^2} + \frac{ \partial V(\phi_{w})}{\partial\phi}\,=\,0 \label{walleqn} \end{equation} where $\phi_{w}(z)$ satisfies the boundary condition $\phi_{w}(z) \rightarrow \phi_\pm$ as $z \rightarrow \mp\infty$ and $z$ is the coordinate perpendicular to the planar interface. The quantum fluctuations around the classical static wall solution $\phi_{w}(z)$ are given by the spectrum of the differential operator \begin{equation} {\cal M} \,=\,-\nabla^2 \,+\, \frac{\partial^2 V(\phi_{w}(z))}{\partial\phi^2} \label{wfluctopr} \end{equation} Since the domain wall only depends on the coordinate z, the differential operator for the fluctuations is separable in terms of eigenfunctions $\psi_{\vec{q}_{\bot}}(z,\vec{x}_{\bot})=e^{i\vec {q}_{\bot}\cdot \vec{x}_{\bot}}\psi_n(z)$, where $\vec{x}_\bot$ denotes the transverse coordinates to $z$, namely $x$ and $y$, and $\vec{q}_{\bot}$ the transverse momentum. The functions $\psi_n(z)$ are so lution of the following eigenvalue problem \begin{equation} \left[-\frac{d^2}{dz^2}+\vec{q}^2_{\bot} \,+\, \frac{\partial^2 V(\phi_{w}(z))}{\partial\phi^2}\right]\psi_n(z) = \omega^2_n(\vec{q}_{\bot})\psi_n(z) \label{flucwall} \end{equation} Taking the derivative of eq. (\ref{walleqn}) with respect to $z$, i.e. \begin{equation} \left[ - \frac{d^2}{dz^2} + \frac{ \partial^2 V(\phi_{w})}{\partial\phi^2} \right] \frac{d \phi_{w}}{dz}\,=\,0 \end{equation} and comparing to eq. (\ref{wfluctopr}) we see that $d\phi_w/dz$ is the zero mode which corresponds to the translational invariance\cite{rajaraman}. Therefore the eigenfunctions \begin{equation} \psi_{\vec{q}_{\bot}} = \mbox{exp}\left(i \mbox{{$\vec{q}_\bot\cdot \vec{x}_\bot$}} \right) \frac{d \phi_{w}}{dz} \label{capqs} \end{equation} have eigenvalues $\vec{q}^2_\bot$. These are Goldstone modes associated with translational invariance and represent excitations of the surface of the planar interface $\phi_w(z)$ since \begin{eqnarray} \phi(\vec{r}) & = & \phi_w(z-z_0) + \sum_{q_\bot} a_{q_\bot} \mbox{exp}\left(i \mbox{\boldmath{$q_\bot . x_\bot$}} \right) \frac{d \phi_{w}}{dz} \nonumber \\ & \simeq & \phi_w(z-z_\bot(\mbox{\boldmath{$x_\bot$}})) \nonumber \end{eqnarray} where $z_0$ is the position of the planar interface and \[ z_\bot(\mbox{\boldmath{$x_\bot$}}) = z_0 - \sum_{q_\bot} a_{q_\bot} \mbox{exp}\left(i \mbox{\boldmath{$q_\bot. x_\bot$}} \right) \] This is clearly similar to the case of the spherical bubble eq. (\ref{surfaceflucs}) and describes the same physics, i.e. fluctuations of the surface. In the case of a planar interface these surface waves are also called capillary waves, and describe the {\em hydrodynamic} modes of long-wavelength fluctuations of interfaces in systems with two degenerate phases separated by an interface\cite{safran}. In the case of degenerate phases, such as for example a liquid-gas or an Ising system, these surface waves are Goldstone modes associated with the breakdown of translational invariance by the presence of the interface. For a spherical bubble in the thin wall limit these surface waves acquire a gap given by the inverse radius (proportional to the Gaussian curvature of the surface). Therefore in analogy with the case of interfaces for degenerate separated phases, we identify these surface fluctuations as {\em quasi Goldstone modes}. Since, as argued before the maximum frequency of the surface waves is $< m$ they are {\em classical} in the high temperature limit $T>>m$. Hence the surface waves are identified as {\em classical hydrodynamic fluctuations} of the bubble shape and quasi-Goldstone modes in the thin wall limit. We want to emphasize that these low energy excitations are a robust feature of the thin wall approximation and are {\em model independent}. Having studied in detail the {\em general} aspects of fluctuations around a thin wall bubble, we now focus on the specific details of the theory with a potential given by (\ref{pot}) so as to be able to compute the matrix elements and provide a quantitative analysis of the viscosity effects. \subsection{$\phi^4$ theory: specifics} The critical bubble solution that satisfies eq. (\ref{3Dbubble}) with the potential (\ref{pot}) is not in general an elementary function, but in the thin wall limit the critical radius of the bubble $mR_c \sim R_c/\xi >>1$ and the function $d\phi_b / dr$ is localized near $R_c$ which makes the ``friction" term, $r^{-1} d\phi_b/dr \propto 1/(mR_c) \sim \xi / R_c <<1$. In this limit, the critical bubble solution is found to be \begin{equation} \phi_b(r,R_c) \,=\, \phi_-\,+\,{m\over 2\sqrt{2\lambda}} \left\{1 - \tanh \left[\frac{r-R_c}{\xi}\right]\right\}; ~~~\xi= {2\over m} \label{bubble} \end{equation} It corresponds to a field configuration that starts around $r=0$ at the true vacuum $\phi_+$, given by eq. (\ref{phipls}), and goes to the false vacuum, $\phi_-$, as $r \rightarrow \infty$ with a surface width $\xi =m/2$ and a critical radius $R_c$. Having specified the critical bubble solution $\phi_b$, we now go back and determine explicitly the general expressions which we discussed in the previous sections. The total energy of the bubble as a function of its radius $R$, given by eq.(\ref{bubenergy}), can be calculated from the field $\phi_b(r,R_c)$ by replacing $R_c$ in eq. (\ref{bubble}) by $R$. The gradient term $\left(\nabla \phi_{b} \right)^2$ contributes to the surface energy while $V(\phi_b)$ has two contributions: a surface contribution $V_s(\phi_b)$ and volume contribution $V_v(\phi_b)$ given by \begin{eqnarray} V_s(\phi_b) & = & \frac{1}{2} \left(\nabla \phi_{b}(r) \right)^2 \,+\, \frac{\eta \,m^4}{32\,\lambda }\,\left\{ \mbox{sech}^2\left[\frac{r-R}{\xi} \right] \left( 3 - \tanh\left[\frac{r-R}{\xi} \right] \right) \right\} \\ V_v(\phi_b) & = & -\frac{\eta \,m^4}{8\,\lambda } \left(1 - \tanh\left[\frac{r-R}{\xi} \right] \right) \end{eqnarray} where $\eta$ is defined by eq. (\ref{eta}). Substituting the above expressions in eq. (\ref{bubenergy}) and evaluating the integral, we find that the total energy of the bubble is given by \begin{equation} E_{var}(R) \,=\, \frac{ 4 \,\pi}{3} \, V(\phi_+)\,R^3 \,+\, {\frac{4\pi {R^2}{m^3}}{12\lambda}}\,\left[ 1 + \frac{9\,\eta }{2} - \frac{3\eta }{2}\frac{\xi}{R}+ \left(\frac{\pi^2 - 6}{12}+\frac{3\pi^2\eta}{8} \right) \frac{\xi^2}{R^2}\right] \label{energy3dbubb} \end{equation} where in the thin wall limit \[ V(\phi_+) \,\simeq \, -\frac{ \eta\,{m^4}}{4\,\lambda }\,+\, {\cal O}((\epsilon-1)^3). \] The total energy attains its maximum at \begin{equation} R_c \,=\, \frac{1}{12\xi \eta}\left[2 + 9\,\eta + \sqrt{4 + 36\,\eta + 45\,\eta^2}\right]. \end{equation} Using the fact that $\eta$ is small in the thin wall limit, we find that the critical radius $R_c$ is given by \begin{equation} R_c \,=\, \frac{1}{3\,\eta \xi}\,\left[ 1 + {\cal O}((\epsilon-1)^2)\right]\;\; \rightarrow\;\; \frac{\xi}{R_c} \,\simeq \,3 \, \eta \end{equation} and the total critical energy is given by \begin{equation} E_c \,=\,\frac{4 \,\pi m}{81 \,\lambda \, \eta^2}\left(1 \, + \,{\cal O}((\epsilon-1)^2) \right) \end{equation} which is equivalent to eqs. (\ref{rc}) and (\ref{ec}) with the surface tension given by \begin{equation} \sigma \,=\,\frac{m^3}{12 \, \lambda}\label{surfacetension} \end{equation} The low-lying fluctuation modes ${\cal U}_{0lm}(\theta,\varphi,r)$ eqn. (\ref{lolyin}) are given by \begin{eqnarray} {\cal U}_{0lm}(r,\theta,\varphi) & = & \frac{\sqrt{6\,m}}{4\,R_c} \, \mbox{sech}^2\left[\frac{r-R_c}{\xi}\right] Y_{lm}(\theta,\varphi) \;\;\;\mbox{with}\;\;\;\omega^2_{0l}\,=\, \frac{(l-1)(l+2)}{R_c^2} \left[1+{\cal O} \left( \frac{\xi^2}{R_c^2}\right) \right]\label{spcquasi} \end{eqnarray} \noindent and the next rotational band of bound states is given by (for details see appendix A) \begin{eqnarray} {\cal U}_{1lm}(r,\theta,\varphi) & \simeq & \sqrt{N_{1l}} \, \mbox{sech}\left[\frac{r-R_c}{\xi}\right]\mbox{tanh}\left[\frac{r-R_c}{\xi}\right] Y_{lm}(\theta,\varphi) \nonumber \\ & &\mbox{with}\;\;\;\omega^2_{1l}\,=\, \frac{3m^2}{4}+\frac{l(l+1)}{R^2_c}+{\cal O}\left(\frac{\xi}{R_c}\right) \label{nextbs} \end{eqnarray} \subsection{Corrections to the bubble growth rate} From eq. (\ref{delomegac}), the corrections from quantum and thermal fluctuations to the bubble growth in the present case has the following form in the classical limit \begin{equation} \delta\Omega = -\frac{\lambda T}{2 } \sum_{qlm,q'l'm'}\frac{|{\cal B}_{qlm,q'l'm'}|^2}{\omega_{ql}^2\,\omega_{q'l'}^2}\left\{ \frac{\Omega}{\Omega^2+(\omega_{ql}+\omega_{q'l'})^2} +\frac{\Omega}{\Omega^2+(\omega_{ql}-\omega_{q'l'})^2} \right\} \label{delomega3d} \end{equation} where the index $q$ runs over bound and scattering states and \begin{equation} {\cal B}_{qlm,q'l'm'} \, \equiv \, \frac{1}{2\sqrt{\lambda}}\int ~d^3r~ V'''(\phi_{b}){\cal U}_{-1}(r){\cal U}_{qlm}({\mathbf r}){\cal U}_{q'l'm'}({\mathbf r}) \label{bqlm} \end{equation} with \begin{equation} V'''(\phi_{b})\,=\, - 6 m \sqrt{2 \lambda} \tanh\left[\frac{r-R_c}{\xi} \right] \end{equation} Since $V'''$ is spherically symmetric, the angular integral leads to \begin{equation} {\cal B}_{qlm,q',l',m'} = {\cal B}_{q,q',l} \delta_{l,l'}\delta_{m,m'} \end{equation} with \begin{equation} {\cal B}_{q,q',l}\, \equiv \, \frac{(3m)^{3/2}}{4\sqrt{\pi}\, R_c} \sqrt{N_{ql}} \sqrt{N_{q'l}} \int ~r^2 dr~\tanh\left[\frac{r-R_c}{\xi} \right] \mbox{sech}^2 \left[\frac{r-R_c}{\xi} \right]\psi_{ql}(r)\psi_{q'l}(r) \label{bql} \end{equation} We now note that the bound state energies only depend on $l$, and that the energy of the scattering states is independent of $l$. Therefore there is {\em no Landau damping contribution} from the bound states as consequence of the principal part prescription which subtracts the contribution from $\omega_p = \omega_{p'}$ in the Landau damping term as discussed in detail below equations (\ref{gamas}, \ref{delomega}). Therefore the correction to the growth rate can be written as $\delta \Omega = \delta \Omega_b + \delta \Omega_s$ with the correction from the bound states given by \begin{equation} \delta\Omega_b = -\frac{\lambda T}{2} \sum_{n,n'=0,1} \sum_{l}\frac{(2l+1)| {\cal B}_{n,n',l}|^2}{\omega_{nl}^2\,\omega_{n'l}^2} \frac{\Omega}{\Omega^2+(\omega_{nl}+\omega_{n'l})^2} \label{corr3dbound} \end{equation} \noindent and that from the scattering states given by \begin{equation} \delta\Omega_s = -\frac{\lambda T}{2} \sum_{k,k'\neq 0,1} \sum_{l} \frac{(2l+1)| {\cal B}_{k,k',l}|^2}{ \omega_{k}^2 \omega_{k'}^2} \left\{ \frac{\Omega}{\Omega^2+(\omega_{k}+\omega_{k'})^2} \,+\, \frac{\Omega}{\Omega^2+(\omega_{k}-\omega_{k'})^2} \right\} \label{corr3dscat} \end{equation} The bound state correction has three contributions; the {\em quasi-Goldstone} modes contribution corresponding to surface waves with $\omega^2_{0l} = \Omega^2 (l-1)(l+1)/2; l\geq 2$, the contribution from the higher energy rotational band of bound states near $\omega^2 \geq 3m^2/4$ and the mixed contribution from the quasi-Goldstone and the higher energy bound states modes. The denominators in $\delta \Omega_b$ are of order $ \Omega^6$ for the quasi-Goldstone modes contribution as compared to $m^6$ for the mixed and the higher energy bound states contributions. Since in the thin wall approximation $m/\Omega \propto R_c/ \xi >>1$, the largest contribution arises from the quasi-Goldstone surface modes with the lowest energy denominators. It can be easily seen that the matrix elements cannot compensate for the difference in powers of $R_c$ and that in fact for the higher energy bound states these matrix elements are {\em smaller} than those for the surface modes because the wave functions ${\cal U}_{1lm}$ actually vanish at the position of the bubble wall (see eqn. \ref{nextbs})). The contribution from the scattering states is also seen to be much smaller than that from the surface waves. The frequencies for the continuum states $\omega_k \geq m >>\Omega$ and the two meson cut give a contribution of order $\Omega$ (since in the denominators the $\Omega$ can be neglected as compared to $m$). For the Landau damping cut, the argument is similar to that of the case of one space dimension. For $\Omega <<m$ this contribution has a Lorentzian shape of width $\approx \Omega$, and the integral over the momenta can be performed in the narrow width approximation. The subtraction of the static contribution guarantees that the integral is dominated by the Lorentzian\cite{note} and the region $\omega_k \approx \omega_{k'}$ can be integrated by changing to relative variables and now in three spatial dimensions the phase space in the region $\omega_k - \omega_{k'} \approx \Omega$ give extra powers of $\Omega$ as compared to the one dimensional case. Furthermore, the matrix elements are smooth functions of the radius of the bubble as can be understood simply by a scattering argument from a sharply peaked potential in three spatial dimensions. These matrix elements do not introduce any singularity in the limit $R_c/\xi >>1$, hence the contribution of the scattering states is at least proportional to $\Omega$ and is therefore subleading in the thin wall limit. This analysis leads to the conclusion that the largest contribution to the correction to the growth rate is given by the {\em quasi-Goldstone modes}, i.e. the surface waves, since these are the lowest lying excitations and hence provide the smallest energy denominators. The matrix element ${\cal B}_{0,0,l}$ for the surface waves can be calculated easily and we find \begin{equation} {\cal B}_{0,0,l} \,=\,\sqrt{\frac{3 m}{\pi}} \frac{2}{5 R_c^2} \,=\,\sqrt{\frac{3 m}{\pi}} \frac{\Omega^2}{5} \end{equation} \noindent leading to the following correction \begin{equation} \delta\Omega_{sw} \,=\, -\frac{6 \lambda \alpha}{25 \pi} \frac{m T}{\Omega} \end{equation} where \begin{equation} \alpha \,\equiv \, \sum_{l=0}^{R_c/\xi} \frac{2 l + 5}{(l+1)^2 (l+4)^2 \left(1+2(l+1)(l+4) \right)} \simeq 0.039 \label{alpha} \end{equation} The above series converges rapidly and only the first few terms contribute to the sum, we have evaluated the sum numerically with $R_c/\xi =10$. Hence we summarize one of the main results of this article: the lowest order correction to the bubble growth for three dimensional bubbles is dominated by viscosity effects arising from the excitation of long-wavelength surface waves and is given by \begin{equation} \delta \Omega = -0.003 \lambda \Omega T\xi \left(\frac{R_c}{\xi}\right)^2 \label{fincorr3d} \end{equation} Therefore to lowest order in $\lambda$ (one-loop) and to leading order in the thin wall approximation we find that the growth rate of slightly supercritical bubbles is given by \begin{equation} \Omega = \frac{\sqrt{2}}{R_c} \left[1- 0.003\lambda T \xi \left(\frac{R_c}{\xi}\right)^2 \right] \label{finalgrowthrate} \end{equation} \noindent with $\xi$ and $R_c$ are the width and radius of the critical bubble. This is one of the important results of our study. The validity of perturbation theory places a very stringent constraint on the quartic coupling constant in the thin wall limit $R_c/\xi >>1$ and in the classical limit $T/m \sim T\xi >>1$. The validity of the classical limit in this case is warranted: we are studying the dynamics of nucleation via thermal activation for temperatures below the critical temperature $T< T_c \approx m/\sqrt{\lambda}$ but for temperatures much larger than the energy of the low lying excitations, the relevant regime for thermal activation is $m/\sqrt{\lambda}> T >>m$ in the weak coupling limit. Furthermore the low-lying excitations with frequencies $<<m$ are obviously classical. \section{Conclusions and discussion} The focus of this article is to provide a microscopic calculation of the growth rate of slightly supercritical nucleation bubbles. The model under consideration is a $\phi^4$ scalar theory with an explicitly symmetry breaking term that produces a metastable and a stable ground state, we studied the case of nucleation in $1+1$ dimensions as well as $3+1$ dimensions. The former is relevant in the case of quasi-one dimensional charge density wave systems and organic conductors. We begin our analysis by obtaining the critical bubble solution by including finite temperature effects in the potential that enters in the classical equation of motion, counterterms are added to the Lagrangian to compensate for the finite temperature corrections consistently in a perturbative expansion. Our approach to obtaining the growth rate is very different from previous treatments in that we begin by expanding the quantum field around the critical bubble in terms of the quadratic fluctuations around the critical bubble configuration. These fluctuations describe an unstable direction associated with small departures from the critical radius, translational zero modes and stable fluctuations. The translational modes are anchored by fixing the center of the bubbles and we treat explicitly the interaction between the coordinate associated with the growth (or collapse) of the bubble with those associated with the stable fluctuations. We obtain the growth rate by obtaining the effective linearized equation of motion for the unstable coordinate by integrating out the coordinates associated with the stable fluctuations. Two different approximations are involved; i) a weak coupling expansion in terms of $\lambda$ the quartic self-coupling, and ii) the thin wall approximation in terms of $\xi/R_c$ with $\xi$ the width of the bubble wall and $R_c$ the critical radius. The first approximation allows a consistent perturbative expansion of the self-energy of the unstable coordinate, the second allows a quantitative calculation of the relevant matrix elements, furthermore our analysis reveals that the important fluctuations are {\em classical} for temperatures $T_c>T>>m$, with $T_c$ the critical temperature and $m$ the mass of quanta in the metastable phase. In the one dimensional case we are able to provide an estimate for the growth rate given by eq. (\ref{estimate}) where the functions $F[R_c/\xi]~,~ C[R_c/\xi]$ depend in a detailed manner upon the scattering states solutions of the eigenvalue problem for the quadratic fluctuations, and clearly will depend on the details of the potential. This estimate points out the potential breakdown of perturbation theory in the thin wall limit. In the case of three dimensions we are able to extract some robust features that trascend the form of the potential and are solely a consequence of the thin wall limit. In particular we identify a rotational band of low lying excitations which describe {\em surface waves} i.e. ripples on the surface of the bubble. We establish the connection of these surface waves to the capillary waves of flat interfaces in the case of degenerate but phase separated thermodynamic states (such as the Ising or liquid - gas at coexistence), the surface waves are then identified as {\em classical long-wavelength hydrodynamic fluctuations}. The unstable coordinate couples to these hydrodynamic fluctuations and as a result the friction term arising from the self-energy is dominated by the coupling to these hydrodynamic modes. Clearly the {\em coupling} of the unstable coordinate to these hydrodynamic modes depends on the model, and in the case under consideration we find in the thin wall limit and to lowest order in perturbation theory the following expression for the growth rate \begin{equation} \Omega = \frac{\sqrt{2}}{R_c} \left[1- 0.003\lambda T \xi \left(\frac{R_c}{\xi}\right)^2 \right] \end{equation} We also obtain the effective {\em non-Markovian} Langevin equation for the coordinate describing small departures from the critical radius and establish the generalized fluctuation dissipation relation between the viscosity and the noise kernel. The noise is correlated on time scales comparable to $\Omega^{-1}$ precisely as a consequence of the coupling to the hydrodynamic modes and cannot be treated simply with white (delta function) correlations. {\bf Discussion:} Although we have studied a specific microscopic model, in the case of $3+1$ dimensions we have been able to identify some robust features that trascend the particular model. These are the existence and dominance of hydrodynamic fluctuations associated with surface waves in the thin wall limit. The coupling of the coordinate associated with small departures from the critical radius to these low energy fluctuations induces friction or viscosity corrections to the growth rate of slightly supercritical bubbles and in the weak coupling and thin wall limit these fluctuations give the largest contribution to the friction corrections. One of our original motivations is to make contact with previous studies of nucleation as applied to the quark-hadron phase transition. In particular Csernai and Kapusta\cite{kapu} have parametrized the coarse grained free energy that describes a quark-hadron first order phase transition in terms of a local energy variable that can be identified with our scalar field $\phi$. The form of the potential taken by these authors coincides with our potential $V(\phi)$ (\ref{pot}), with coefficients that depend on temperature, just as we have argued in this article. Their form of the critical bubble solution and the variational energy as a function of the radius of a bubble are very similar to those studied in this article. These authors have established that for about $1\%$ supercooling the critical radius for such a model is about $R_c \approx 12 \mbox{fm}$ the width of the wall is about $\xi \approx 0.7 \mbox{fm}$ and the thin wall approximation must be reliable in this regime. We can obtain the quartic self-coupling $\lambda$ in (\ref{pot}) from the parameters used in\cite{kapu} by relating the width of the bubble $\xi =2/m$ and the surface tension $\sigma$ to $\lambda$ via eqn. (\ref{surfacetension}). In\cite{kapu} the value of the surface tension for the particular quark-hadron model is $\sigma = 50 \mbox{MeV}/\mbox{fm}^2$ for $T \approx 200 \mbox{MeV}$, yielding a value $\lambda \approx 6$ which leads to a very large correction to the growth rate. Certainly the large value of the coupling invalidates the perturbative scheme and we cannot draw a definite conclusion as to the relevance of our lowest order estimate for the quark-hadron transition beyond the statement that the viscosity induced by the hydrodynamic fluctuations {\em could} result in a very large (negative) correction to the growth rate and therefore a rather small nucleation rate. For larger supercooling the critical radius becomes smaller and the thin wall approximation breaks down, but in this case nucleation and spinodal decomposition will be indistinguishable and homogenous nucleation theory may not be the proper description. However despite the limitations of the perturbative expansion and the thin wall approximation, we have provided a consistent approach to obtain friction or viscosity corrections to the growth rate from a {\em microscopic} perspective without invoking a phenomenological description. The observation that in the thin wall limit the most important corrections arise from the coupling to {\em classical} hydrodynamic fluctuations could perhaps pave the way to a systematic hydrodynamic treatment that would circumvent the weak coupling expansion and allow to extract non-perturbative physics. \section{Acknowledgements} D. B. thanks the N.S.F for partial support through the grant awards: PHY-9605186 and INT-9512798. S. M. A. thanks King Fahad University of Petroleum and Minerals (Saudi Arabia) for financial support. S. E. J. thanks CNPq for finantial support. F. I. T. thanks FAPEMIG for support. E. S. F. thanks CNPq and FAPERJ for support and D. G. B. thanks FAPERJ and CNPq for support through a binational collaboration.
\section{Observations} UMi dSph was observed in $B$ and $R$ with the Wide Field Camera (WFC) at the prime focus of the INT (2.5m) at the Roque de los Muchachos Observatory (Canary Islands). The WFC holds four 4096 $\times$ 2048 pixels EEV CCDs with a pixel size of 0.33 $\arcsec$, covering a total area of 0.27 square degrees for field. Figure 1 shows the fields selected for the first observing run of this project. \section{The color-magnitude diagram of Ursa Minor dSph} Figure 2 shows the [$B-R$, $"V"$] CMDs for the central body of UMi dSph (field A of Fig. 1). These diagrams show a prominent narrow RGB suggesting a small metallicity range in the stellar population. In addition, a well populated, blue extended horizontal branch and several ($\sim$ 150) stars in the RR Lyrae gap region are observed. The main turn-off corresponds to an age of $\sim 15$ Gyr for $Z=0.0004$, based on the Padua's stellar evolutionary models (Bertelli et al. 1994, and references therein). The few stars above the subgiant track ($0.6 \leq (B-R) \leq 1.0$; $ 20.5 <"V"< 22.5$) are probably binary stars although they could be the trace of a not so old population. Perhaps the most streaking feature is the well populated {\it blue plume}. A simple explanation is that it would be made of blue stragglers. However, normalized to the RGB, it contains three times more stars than low central density Galactic globular clusters, like NGC 5053 (Nemec \& Cohen 1989). This perhaps indicates that the blue-plume is the trace of an intermediate-age population. Nevertheless, photometry alone is not enough to distinguish blue stragglers from normal main sequence stars. A definitive answer will requires spectroscopic data. \begin{figure} \plotfiddle{martinez2_fig1.ps}{5cm}{0}{40}{40}{-150}{-80} \caption{Location of the WFC fields of our study of UMi. North is to the top and East to the left on the isopleth map by Irwin \& Hatzidimitriou (1995). The scale is 1.65 $\arcmin$/pixel.} \label{fig-1} \end{figure} \section{ An intermediate-age population in UMi?} We have done a preliminary estimate of the star formation history for the central body of UMi using the CMD of chip 4 (Figure 2) and assuming that the blue stars present in the CMDs are not blue-stragglers but an intermediate-age population. The result is shown in Figure 3 (left panel). A high SFR is required in an early epoch to account for the amount of HB and RGB stars. After $\sim 12$ Gyr ago, a low star formation activity extended until a recent epoch, is enough to produce the observed blue-plume. Interestingly, UMi would have maintain this intermediate-age SFR if it somehow would have manage to convert into new stars only a 5$\%$ of the material returned to the interstellar medium by older stellar generations. For illustrative purposes, right panel of Figure 3 shows the corresponding synthetic CMD. For simplicity, a constant metallicity ($Z=0.0004$) has been used and no observational effects have been simulated. \begin{figure} \plotfiddle{martinez2_fig2.ps}{8cm}{0}{60}{60}{-200}{-100} \caption{CMDs for the central body of UMi (field A). Each diagram corresponds to one of the four chips of the INT WFC. Chip 4 includes the nucleus of the galaxy.} \label{fig-1} \end{figure} \begin{figure} \plottwo{martinez2_fig3.ps}{martinez2_fig4.ps} \caption{The SFR computed for the central body of UMi (left) and the corresponding synthetic CMD (right).} \label{fig-1} \end{figure} \section{Searching for tidal tails} As a part of our project, we are searching for tidal tails in the periphery of UMi. A first test is the detection of relevant structures in the CMDs of these regions, such as RGB or HB traces. Figure 4 shows the CMDs of field C (shown in Fig. 1), situated over the major semiaxis of UMi and beyond its tidal radius in the direction of the Magellanic Stream. No traces of the galaxy are found in this field. A similar result is found for the field B (see Fig. 1). The lack of traces beyond the tidal radius may indicate the absence of extra-tidal stellar debris and suggests that UMi would not be in an advanced state of tidal disruption. \begin{figure} \plotfiddle{martinez2_fig5.ps}{8cm}{0}{60}{60}{-200}{-100} \caption{CMDs for the field C (see Fig. 1) situated in the outer regions of UMi, beyond its tidal radii. No trace of the galaxy is found.} \label{fig-1} \end{figure}
\section{Introduction} Since the introduction of the Renormalization--Group by Wilson \cite{Wilson}, we have learned much about critical phenomena and the striking feature of universality. Unfortunately, the majority of systems with nontrivial behaviour cannot be treated analytically. Nevertheless, there exist approximation schemes concerning finite systems which allow to estimate the properties of the corresponding infinite system by proper extrapolation to the thermodynamic limit. A very powerful method for the investigation of the critical properties of an infinite system along these lines is the socalled finite size scaling theory introduced by Fisher et al.~ (see \cite{Fisher}, \cite{Barber} and references therein). Although hypothesised before the advent of the Renormalization--Group, finite size scaling may be understood within the framework of the latter.\\ The main result of finite size scaling theory may be stated like this: In the vicinity of the critical point of a given infinite system, the system size dependence of certain thermal properties of the corresponding {\em finite} system is governed by properties (namely: the critical indices) of the {\em infinite} system. Or, formulated slightly differently: In spite of the fact, that the free energy density of a finite system is a completely analytic function of its variables, the system size dependence of certain derivatives of the free energy density are dictated by quantities which describe the non--analytic behaviour of the free energy density of the corresponding infinite system. All thermal properties of a finite system (volume% \footnote{Throughout this paper, we will use the language of lattice systems with discrete energies, where the number of degrees of freedom $N$ is equivalent to the volume $V$.}% $\,\,V\equiv N:=L^{\rm d}$) are given by logarithmic derivatives of the canonical partition function \begin{equation} Z_N({\rm T})\,:=\,\sum \limits_{x\in\Gamma}\,e^{-\beta H(x)} \qquad , \quad {\rm T}=\frac{1}{\beta} \qquad , \quad {\rm k}_B \equiv 1 \qquad , \end{equation} where $H(x)$ represents the energy of a particular microstate $x$ of the system and the sum runs over all possible microstates which constitute the space $\Gamma$ of all the states available to the system. With the definition of the microcanonical density of states \begin{equation} \Omega_N (\varepsilon) \,:=\,\sum \limits_{x\in\Gamma}\,\delta_{H(x),N\varepsilon} \end{equation} and the microcanonical specific entropy \begin{equation} \hat s_N(\varepsilon)\,:=\, \frac{1}{N}\,\ln \Omega_N(\varepsilon) \qquad , \end{equation} the canonical partition function reads \begin{equation} \label{partition_function} Z_N({\rm T})\,:=\,\sum\limits_{\varepsilon}\,e^{N\left( \hat s_N(\varepsilon)-\beta \varepsilon \right)} \qquad . \end{equation} Here, the sum runs over all possible energy--values of the finite system. As it is clearly visible, the system size dependence of the partition function is due to two causes. Namely, the size dependence of the microcanonical specific entropy $\hat s_N (\varepsilon)$ and the overall factor $N$ in the exponential, which we will call the trivial system size dependence. It is the aim of this paper to demonstrate that finite size scaling emerges naturally and in some sense trivially from the assumption that the critical properties of the infinite system are already contained in the microcanonical specific entropy of the finite system. Unfortunately, up to now we are able to show this only as far as the finite size scaling properties of the specific heat are concerned. Although we are never introducing a specific system explicitly by giving its Hamiltonian, we restrict our discussion to systems with short range interactions. For this reason, standard finite size scaling is applicable and hyperscaling holds% \footnote{At least if the dimensionality of the system is smaller than the upper critical dimension of the corresponding universality--class.}. Remark: In the thermodynamic limit, the entropy $\hat s_N(\varepsilon)$ is replaced by the Massieu--function $\hat s(\varepsilon,h/{\rm T})$ at zero magnetic field $h\,\,$ $(=:\hat s(\varepsilon))$ , which is the Legendre transform of the entropy $s(\varepsilon,m)$: \begin{equation} \lim_{N \to \infty} \hat s_N(\varepsilon) \,\,\,=\,\,\, \hat s(\varepsilon, \frac{h}{{\rm T}}=0) \,\,\,:=\,\,\, \left. \sup \limits_m \left\{ s(\varepsilon,m) + \frac{h}{{\rm T}} \, m \right\} \right|_{\frac{h}{{\rm T}}=0} \end{equation} Here, $m$ denotes the magnetization per particle. In this paper, the thus defined Massieu--function will be called "microcanonical specific entropy". In section \ref{s_infty}, we give an explicit form for the microcanonical specific entropy of a system which undergoes a continuous phase transition with a power law singularity of the specific heat. In section \ref{s_scale}, we show that the system size independence of the microcanonical specific entropy implies canonical finite size scaling. Namely, the scaling of the specific heat maximum (section \ref{c_max_scale}) and the scaling of the softening of the specific heat singularity (section \ref{c_broad_scale}). Since we are not able to proof the reverse direction of this statement, in section \ref{hints}, we give some hints onto the validity of the conjecture that the system size dependence of the microcanonical specific entropy is such weak not to impact the canonical finite size scaling. \section{Microcanonical specific entropy $vs.$~continuous phase transitions \label{s_infty}} The microcanonical specific entropy of a system which undergoes a continuous phase transition may be written as a sum of a singular part $\hat s_{sing}(\varepsilon)$ and a correction term $\hat s_{corr}(\varepsilon)$ which is needed to correctly describe the behaviour of the specific entropy outside the critical region. \begin{equation} \label{s_sing_reg} \hat s(\varepsilon) \, = \, \hat s_{sing}(\varepsilon) \, + \, \hat s_{corr}(\varepsilon) \end{equation} If the corresponding specific heat shows a power law singularity, the choice \begin{equation} \label{s_sing} \hat s_{sing}(\varepsilon)\,=\,\hat s_c+\beta_c(\varepsilon-\varepsilon_c)-\frac{\beta_c |\varepsilon_c|}{g} \left| \frac{\varepsilon-\varepsilon_c}{\varepsilon_c} \right|^g \left(\vphantom{\int} \Theta(\varepsilon_c-\varepsilon)A+\Theta(\varepsilon-\varepsilon_c)A^\prime \right) \end{equation} (with $g:=\frac{2-\alpha}{1-\alpha}$) for the singular part of the specific entropy yields the correct behaviour of the singular part of the specific heat% \footnote{ In the case of a logarithmic singularity the last term in $\hat s_{sing}(\varepsilon)$ should be replaced by a function of the form $ \,\,\left(\varepsilon-\varepsilon_c\right)^2\,/\,\ln \left|\varepsilon-\varepsilon_c\right|\, $.}. $\hat s_c$, $\varepsilon_c$ and $\beta_c$ are the values of the specific entropy, the specific energy and the inverse temperature at the critical point ($\beta_c:=1/{\rm T}_c$), the step--function $\Theta(x)$ is defined by $\Theta(x):=1$ $\forall \,x>0$ and $\Theta(x):=0$ $\forall \, x\le0$. Indeed, since the microcanonical specific heat is given by \begin{equation} c(\varepsilon):=-\frac{\beta (\varepsilon)^2}{s^{\prime\prime}(\varepsilon)}:= -\frac{s^\prime(\varepsilon)^2}{s^{\prime\prime}(\varepsilon)}\quad , \end{equation} differentiating the singular part (\ref{s_sing}) of the specific entropy twice with respect to the specific energy $\varepsilon$ yields \begin{equation} c_{sing}(\varepsilon)\,=\,\frac{\beta_c |\varepsilon_c|}{g-1}\, \left( \frac{\Theta(\varepsilon_c-\varepsilon)}{A}+\frac{\Theta(\varepsilon-\varepsilon_c)}{A^\prime} \right)\,\, \left| \frac{\varepsilon-\varepsilon_c}{\varepsilon_c} \right|^{-\frac{\alpha}{1-\alpha}} \end{equation} in the vicinity of the critical point $\varepsilon_c$. Alternatively, by going over from $\left|\frac{\varepsilon-\varepsilon_c}{\varepsilon_c}\right|$ to $\left|\frac{{\rm T}-{\rm T}_c}{{\rm T}_c}\right|$ via \begin{equation} \left| \frac{\beta (\varepsilon)-\beta_c}{\beta_c} \right|\, \approx \, \left|\frac{\varepsilon-\varepsilon_c}{\varepsilon_c}\right|^\frac{1}{1-\alpha} \left( \vphantom{\int} \Theta (\varepsilon_c-\varepsilon) A + \Theta (\varepsilon-\varepsilon_c ) A^\prime \right) \quad , \end{equation} the singular part of the microcanonical specific heat as a function of the reduced temperature ${\rm t}:=({\rm T}-{\rm T}_c)/{\rm T}_c$ is proportional to $|{\rm t}|^{-\alpha}$: \begin{equation} \label{c_sing} c_{sing}({\rm t})=(1-\alpha)\frac{|\varepsilon_c|}{{\rm T}_c} \left( \frac{\Theta({\rm T}_c-{\rm T})}{A^{1-\alpha}}+\frac{\Theta({\rm T}-{\rm T}_c)}{{A^\prime}^{1-\alpha}} \right)\, \left|{\rm t}\right|^{-\alpha} \end{equation} The correction term $\hat s_{corr}(\varepsilon)$ of the specific entropy yields no contribution to the singular behaviour of the specific heat if it obeys the following condition: \begin{equation} \label{s_reg_cond} \lim_{\varepsilon\to\varepsilon_c} \frac{\hat s_{corr}(\varepsilon)}{|\varepsilon-\varepsilon_c|^g}\,=\,0 \end{equation} \section{Microcanonical specific entropy $vs.$~finite size scaling of the canonical specific heat \label{s_scale}} In the rest of this paper, we will study finite size scaling properties of the canonical specific heat of a hypothetical $N$--particle system with specific entropy \begin{equation} \label{postulat} \hat s_N(\varepsilon) \, \equiv \, \hat s_{sing}(\varepsilon)\,+\,\hat s_{corr,N}(\varepsilon) \qquad \forall \quad N\,>\,N_0 \qquad . \end{equation} This implies the assumption that, at least for sufficiently large $N$, the singular contribution to $\hat s_N(\varepsilon)$ is identical to the singular part of the entropy of the infinite system. The correction term $\hat s_{corr,N}(\varepsilon)$ may show some $N$--dependence which should of course be consistent with the condition (\ref{s_reg_cond}). Having postulated the form of the entropy in (\ref{postulat}) the canonical specific heat $c_N({\rm T})$ of the $N$--particle system follows directly from (\ref{partition_function}). We shall compare the scaling properties of the thus determined specific heat $c_N({\rm T})$ with the results of conventional finite size scaling theory \cite{Binder}. At the critical temperature ${\rm T}_c$ of the infinite system the finite size scaling theory predicts for the value of the specific heat of the finite system \begin{equation} \label{hyper_k} c_N({\rm T}_c) \, \propto \, L^\frac{\alpha}{\nu} \qquad . \end{equation} In the finite system the singularity is smeared. Scaling theory predicts for the width of the specific heat anomaly: \begin{equation} \label{softening} \Delta {\rm T}(L) \, \propto \, \left(\frac{1}{L}\right)^\frac{1}{\nu} \qquad . \end{equation} Here, $\nu$ is the critical exponent of the correlation length $\xi ({\rm t})$ of the infinite system, $L$ is the linear dimension of the finite system and ${\rm T}_c$ denotes the critical temperature of the infinite system. We are now going to show, that the finite size scaling relations \begin{equation} \label{hyper_m} c_N ({\rm T}_c) \, \propto \, L^{\frac{{\rm d}\alpha}{2-\alpha}} \end{equation} and \begin{equation} \label{soft_m} \Delta {\rm T}(L) \, \propto \, \left(\frac{1}{L}\right)^\frac{\rm d}{2-\alpha} \end{equation} are direct consequences of the postulate (\ref{postulat}). Together with the validity of hyperscaling, (\ref{hyper_m}) implies (\ref{hyper_k}) and (\ref{soft_m}) implies (\ref{softening}). While (\ref{soft_m}) is established by numerical integration, (\ref{hyper_m}) can be shown analyticaly. \subsection{Canonical specific heat scaling at ${\rm T}_c$ \label{c_max_scale}} The first step consists in the calculation of the $n$--th moment of the specific energy of the $N$--particle system with respect to the canonical distribution at the critical temperature of the infinite system. \begin{equation} \label{n_moment} \left\langle \varepsilon^n \right\rangle_N ({\rm T}_c)\,\,=\,\, \frac{1}{Z_N({\rm T}_c)} \sum_\varepsilon \, \varepsilon^n\,e^{N(\hat s_N(\varepsilon)-\beta_c \varepsilon)} \end{equation} $Z_N({\rm T}_c)$ denotes the canonical partition function of the $N$--particle system at ${\rm T}_c$ (cf.~(\ref{partition_function})). For sufficiently large $N$, it is justified to replace the sum over all possible energy--values by an integration along the energy--axis. Since the correction term $\hat s_{corr,N}(\varepsilon)$ of the specific entropy will yield no contribution to the canonical quantities at ${\rm T}_c$ (for sufficiently large $N$ again), we are concerned with integrals of the type \begin{equation} {\rm I}_n \, := \, \int\limits_{-\infty}^\infty d\varepsilon \, \varepsilon^n \, \exp \left\{ -N\,\frac{\beta_c |\varepsilon_c |}{g}\, \left| \frac{\varepsilon-\varepsilon_c}{\varepsilon_c} \right|^g \,\left(\vphantom{\int}\Theta (\varepsilon_c - \varepsilon)A + \Theta (\varepsilon-\varepsilon_c) A^\prime \right)\, \right\} \end{equation} Defining new amplitudes $B:=A\beta_c|\varepsilon_c |^{1-g}/g$, $\,\,B^\prime:=A^\prime \beta_c|\varepsilon_c |^{1-g}/g$ and renaming $(\varepsilon-\varepsilon_c) \rightarrow \varepsilon$, we get \begin{equation} {\rm I}_n\,=\, \sum_{l=0}^n \, {n \choose l} \, \varepsilon_c^{n-l}\, \int\limits_0^\infty d\varepsilon \, \varepsilon^l \, \left( e^{-NB^\prime \varepsilon^g} + (-1)^l e^{-NB\varepsilon^g} \right) \end{equation} Substituting $x:=NB^\prime \varepsilon^g$ in the first and $x:=NB\varepsilon^g$ in the second term of the integral, we end up with the following expression for ${\rm I}_n$: \begin{eqnarray}\nonumber {\rm I}_n & = & \sum_{l=0}^n {n \choose l} \, \frac{\varepsilon_c^{n-l}}{g} \left( {B^\prime}^{-\frac{l+1}{g}} + (-1)^l B^{-\frac{l+1}{g}} \right) N^{-\frac{l+1}{g}} \int\limits_0^\infty dx \, x^{\frac{l+1}{g}-1}\,e^{-x} \\ [1ex] & = & \sum_{l=0}^n {n \choose l} \, \frac{\varepsilon_c^{n-l} \Gamma \left( \frac{l+1}{g} \right)}{g} \left( {B^\prime}^{-\frac{l+1}{g}} + (-1)^l B^{-\frac{l+1}{g}} \right) N^{-\frac{l+1}{g}} \end{eqnarray} Since we want to study the finite size scaling properties of the canonical specific heat $c_N({\rm T})$ \begin{equation} \label{c_kan} c_N({\rm T}_c) \, := \, N \beta_c^2 \left( \left\langle \varepsilon^2 \right\rangle_N({\rm T}_c)\,-\, \left\langle \varepsilon^{\vphantom{2}} \right\rangle^2_N ({\rm T}_c) \right) \end{equation} at the critical temperature ${\rm T}_c$ of the infinite system, we have to look at the second central moment of the specific energy with respect to the canonical distribution: \begin{eqnarray} \label{cent_mom} \lefteqn{ \hspace{-3cm} \left\langle \varepsilon^2 \right\rangle_N({\rm T}_c)\,-\, \left\langle \varepsilon^{\vphantom{2}} \right\rangle^2_N({\rm T}_c) \, = \,\frac{\rm I_2}{\rm I_0}\,-\,\left( \frac{\rm I_1}{\rm I_0} \right)^2 \,\, = } \nonumber \\[2ex] \phantom{m} \hspace{2cm} & = & N^{-\frac{2}{g}}\,\, \left[ \frac{\Gamma (\frac{3}{g})}{\Gamma (\frac{1}{g})} \,\, \frac{{B^\prime}^{-\frac{3}{g}}+B^{-\frac{3}{g}}}{{B^\prime}^{-\frac{1}{g}}+B^{-\frac{1}{g}}} \,-\, \left( \frac{\Gamma (\frac{2}{g})}{\Gamma (\frac{1}{g})} \,\, \frac{{B^\prime}^{-\frac{2}{g}}-B^{-\frac{2}{g}}}{{B^\prime}^{-\frac{1}{g}}+B^{-\frac{1}{g}}} \right)^2 \right]\,\propto\,N^{-\frac{2}{g}} \phantom{space} \end{eqnarray} With $g:=(2-\alpha)/(1-\alpha)$ and $N=L^{\rm d}$, {\rm d} being the dimension of configuration space, eq.~(\ref{hyper_m}) follows immediately from (\ref{cent_mom}). In conjunction with the hyperscaling relation \begin{equation} \label{hyper} {\rm d}\nu \, = \, 2-\alpha \qquad , \end{equation} this implies (\ref{hyper_k}). \subsection{Scaling of the specific heat width $\Delta{\rm T}$ \label{c_broad_scale}} The specific heat singularity (\ref{c_sing}) of the infinite system is rounded in the corresponding finite systems. Since in the canonical ensemble there are no phase transitions in finite systems, this effect seems to be quite natural. The standard (finite size scaling) argument for this softening is that the specific heat of the finite system saturates for those temperatures, where the correlation length $\xi({\rm t})$ of the infinite system becomes larger than the linear system size $L$. The correlation length $\xi_N({\rm t})$ of the finite system is bounded from above by a length which is of the order of magnitude of the linear system size $L$. For this reason, there is a temperature region within which the specific heat of the finite system deviates essentially from the specific heat of the infinite system.\\ Any measure of the width of this region will do. For numerical convenience, the width $\Delta {\rm T} (L)$ is defined as the temperature range were $c_N(T)$ is larger than 80\% of its maximum value. Having defined $\Delta {\rm T}(L)$, it is an easy matter to compute it by numerical integration via eqs.~(\ref{partition_function}) and (\ref{s_sing}) for any lattice size $L$. The various parameters appearing in (\ref{s_sing}) have not been chosen arbitrarily but we have taken the parameter set obtained by a fit to the (simulated) entropy data of a three--dimensional Ising model with linear system size $L=18$. The parameters are% \footnote{The value of $\hat s_c$ is not listed, because an additive constant in the specific entropy is quite irrelevant with respect to the physics described by that entropy. Likewise the value of $\varepsilon_c$ is irrelevant for the smearing.}% : $\varepsilon_c=-1.059$; $\alpha=0.1155$; $\beta_c=0.222684$; $A=0.091$; $A^\prime=0.150$.\\ Fig.~1 shows a log--log plot of $\Delta {\rm T} (L)$ {\em vs.~}$1/L$ where we have chosen $N=L^3=10^8,...,10^{10}$ together with a straight--line fit to the data points. The critical exponent $(2-\alpha)/{\rm d}$, which is just the inverse slope of the fitted straight line (cf.~(\ref{soft_m})), emerges to be $(2-\alpha)/{\rm d}=0.6275$ which is consistent with the value of $\nu=0.6282$ obtained by combining the input--value of $\alpha=0.1155$ with the hyperscaling relation (\ref{hyper}). \section{\label{hints}On the system size dependence of microcanonical specific entropies} In the previous section, we have shown that a system size independent microcanonical specific entropy implies the canonical finite size scaling relations. Unfortunately, we are not able to proof this statement in the reverse direction. Nevertheless, we can report at least two observations which are necessary (not sufficient) for the validity of the statement that the system size dependence of the microcanonical specific entropy has no considerable impact on canonical finite size scaling (if the systems are not chosen to be too small). 1) If the microcanonical specific entropy shows no system size dependence and if the critical properties of the infinite system are already contained in the entropy of the corresponding finite systems, then it should be possible to extract information about the critical exponent $\alpha$ by performing some fits of a function of the form (\ref{s_sing}) to the entropy data of finite systems obtained by, e.g., Monte Carlo simulation.\\ And indeed, it has already been shown \cite{Promberger}, that an entropy of the form (\ref{s_sing_reg}) with the singular part given in (\ref{s_sing}) fits the data of a $10 \times 10 \times 10$--Ising system very well and it will be shown elsewhere, that the same entropy with the same exponent $g$ but slightly modified parameters $\varepsilon_c$, $\beta_c$ is well suited to fit the data of larger 3D--Ising systems (the values of the critical exponent $\alpha$ emerging from this fits is well consistent with the respective value of $\alpha$ in the thermodynamic limit, i.e.~$\alpha_{fit} \in [.08;.12]$). 2) If the microcanonical specific entropy shows no substantial system size dependence, then it should make no difference if the value of the specific heat of a finite system at the critical temperature of the infinit system is calculated by use of the entropy of the finite system or by use of the entropy of the infinit system.\\ Fortunately, this can be checked in the case of the twodimensional Ising model, where the entropy of the infinit system can be calculated from Onsagers solution \cite{Onsager}. At zero external field, the internal energy per particle as a function of the inverse temperature reads as: \begin{equation} \label{Onsager} \epsilon(\beta)=-\coth (2\beta)\, \left[ 1+\frac{2}{\pi} \left( 2\tanh^2 (2\beta)-1 \right) K_1(q) \right] \end{equation} where \begin{equation} q:=\frac{2\sinh (2\beta)}{\cosh^2(2\beta)} \qquad \mbox{and} \quad K_1(q):= \int \limits_0^{\pi /2} {\rm d}\varphi \left( 1-q^2 \sin^2\varphi \right)^{1/2} \qquad , \quad J\equiv k_B \equiv 1 \quad . \end{equation} Here, $J$ denotes the Ising coupling constant. Since the inverse temperature $\beta$ is defined to be the derivative of the entropy $\hat{s}(\varepsilon)$ with respect to the energy $\varepsilon$, the entropy can be calculated according to \begin{equation} \hat{s}(\varepsilon) = const. + \int \limits_{\varepsilon_0}^\varepsilon {\rm d}\tilde{\varepsilon}\,\beta (\tilde{\varepsilon}) \end{equation} with arbitrary $\varepsilon_0$. $\beta (\varepsilon)$ is obtained by inverting eq.~(\ref{Onsager}). In the case of a logarithmic specific heat singularity, the canonical finite size scaling theory predicts \cite{Ferdinand} \begin{equation} \label{2dscale} c_N(T_c) \propto \ln (1/L) \end{equation} Having obtained the entropy of the infinit 2$D$ Ising system, it is an easy thing to calculate the critical point specific heat using eq.~(\ref{c_kan}). The result is shown in figure 2 \section{Conclusion} We have shown that the finite size scaling relations (\ref{hyper_m}) and (\ref{soft_m}) are trivial consequences of the postulate (\ref{postulat}): for sufficiently large $N$, the entropy of the finite system was assumed to be identical to the entropy of the infinite system at least in the vicinity of the critical point. In this context, "trivial" means that the softening of the specific heat singularity is caused solely by the trivial factor $N$ in the exponential of the canonical partition function (\ref{partition_function}). In the framework of this scenario it is therefore not astonishing that some properties of the finite system are governed by the critical indices of the infinite system: they are already contained in the entropy of the finite system but they are covered up by the averaging ("smearing") procedure of the canonical partition function (for a detailed discussion of this "smearing--effect" see \cite{Huller}). For this reason it seems to be plausible that, as far as finite systems are concerned, we are in some sense blinded by the canonical formalism which obscures the information already available in the microcanonical specific entropy.\\ Indeed, the hypothetical system which we have discussed may seem to be a rather strange construction but it is not as arbitrary as it seems to be since we have already shown that an entropy of the type (\ref{s_sing}) is well suited to fit the data of a $10^3$--3d--Ising system. Note that this is by no means the only example of a system with a microcanonical specific entropy $\hat s_N(\varepsilon)$ which shows $no$ substantial $N$--dependence. We will report about other examples elsewhere. \section{Acknowledgements} The author wants to thank Alfred H\"uller and Hajo Leschke for many stimulating discussions. \input{lit}
\section{Introduction} Large-scale radio halos in clusters of galaxies are diffuse radio sources with no apparent parent galaxy, typical sizes of 1 Mpc, low surface brightness and steep radio spectrum. They demonstrate the existence of relativistic electrons and large scale magnetic fields in the intracluster medium. The sources classified as {\it radio halos} are located at the cluster centers. Sources with similar properties have also been found at the cluster peripheries: they are called {\it relic} sources. Moreover, in some clusters with a central dominant galaxy, the relativistic particles can be traced out quite far, forming what is called a {\it mini-halo} (see e.g. 3C~84 in the Perseus cluster, Burns {\it et al.~} 1992). The radio halos and relics are a rare phenomenon. They are present in a few rich clusters, characterized by high X-ray luminosity and high temperature. The prototypical example of cluster-wide radio halo is Coma-C, in Coma (see Giovannini {\it et al.~} 1993 and references therein). The Coma cluster also contains the peripheral relic 1253+275, which is connected to Coma-C through a very low-brightness bridge of radio emission (see e.g. Feretti \& Giovannini 1998). Other radio halos studied in detail so far are those in A~2255 (Burns {\it et al.~} 1995, Feretti {\it et al.~} 1997a), and in A~2319 (Feretti {\it et al.~} 1997b). Well studied relic sources are those in A~2256 (R\"ottgering {\it et al.~} 1994), A~3667 (R\"ottgering {\it et al.~} 1997), and A~85 (Bagchi {\it et al.~} 1998). The properties of halos and relics are not yet well understood, because of the low number of known sources of this type. Also, it is not yet clear if radio halos and relics have a common origin and evolution, or should be considered as different classes of sources. Information on a larger sample of halos and relics is crucial to investigate their formation and evolution, and their relation to other cluster properties. To this aim we undertook the search for new candidates of radio halos and relics in the NRAO\footnote {The National Radio Astronomy Observatory is operated by Associated Universities, Inc., under contract with the National Science Foundation.} VLA Sky Survey (NVSS, Condon {\it et al.~} 1998). In this paper we report the detection of new halo and relic candidates, which significantly increase the number of diffuse sources in clusters. The paper is organized as follows: in Sect. 2 we describe the sample of clusters which were inspected in the NVSS, in Sect. 3 we present the results, in Sect. 4 we give comments on the individual sources and their clusters, in Sect. 5 we discuss the results. A Hubble constant H$_0$=50 km s$^{-1}$ Mpc$^{-1}$ and a deceleration parameter q$_0$=1 are assumed throughout. \section {Sample {\bf and Source} selection} We searched for cluster-wide radio sources in clusters using the public images of the NVSS. This radio survey was performed at 1.4 GHz with the Very Large Array (VLA) in the tightest configuration (D). It has an angular resolution of 45$^{\prime\prime}$~ (HPBW), a noise level of 0.45 mJy (1$\sigma$) and covers all the sky north of Declination = --40$^{\circ}$. As a cluster sample we used the sample of X-ray-brightest Abell-type clusters (XBACs) presented by Ebeling {\it et al.~} (1996), consisting of 283 clusters/subclusters from the catalogue of Abell, Corwin \& Olowin (1989, ACO) detected in the ROSAT All Sky survey (RASS) with X-ray flux f$_X>$ 5 10$^{-12}$ erg cm$^{-2}$ s$^{-1}$ in the 0.1-2.4 keV energy range. This is an all-sky, X-ray flux limited sample, complete in the galactic latitude range $\mid b \mid \ge$20$^{\circ}$~ and in the redshift interval z$\le$0.2, but it contains also 12 clusters at lower galactic latitude and 24 clusters with redshift greater than 0.2 that meet the flux criterion (see Tables 3 and 4 in Ebeling {\it et al.~} 1996). Because of the lack of short baselines, the NVSS is insensitive to extended structure larger than 15$^{\prime}$. Since the known radio halos are about 1 Mpc in size, sources of this type are missed in the NVSS if belonging to nearby clusters. We note indeed that the radio halo in Coma (z = 0.0232) is resolved out in the NVSS, because of the lack of short spacings. We have taken as a limiting redshift for the search of diffuse cluster-wide halos and relics the value z=0.044, which corresponds to a largest detectable linear size of about 1 Mpc. Considering this redshift limit, and taking also into account the declination limit of the NVSS, we ended with a sample of 207 clusters. For each of them, we searched for the NVSS image, and extracted a field of 1$^{\circ}$$\times$1$^{\circ}$, centered on the cluster position given in the XBACs catalogue. We remark here that this is the X-ray position and may be different from that of the optical cluster center. Moreover, in the case that the X-ray structure is double or more complex, the XBACs position refers to the centroid of the X-ray emission. The two clusters A~1773 and A~3888 fall in the few remaining gaps of the NVSS, and therefore no image is currently available for them. For a few other clusters, the field retrieved from the NVSS is slightly smaller than 1 square degree, because of the existence of gaps at one of the field edges; however, the available image covers a significant region of the clusters. In conclusion, we have finally inspected 205 clusters, for a total of about 0.062 steradians. The redshift distribution of the searched clusters with respect to the distribution of the whole XBACs catalogue is shown in Fig. 1. \begin{figure} \special{psfile=fig1.ps hoffset=-5 voffset=-325 hscale=60 vscale =60} \vspace{8 cm} \caption{ Redshift distribution of the total sample of the XBACs, with the shaded area representing the clusters inspected in the NVSS.} \label{} \end{figure} To search for the cluster diffuse sources, we used the NVSS images overlaid upon the optical images from the Digitized Palomar Sky Survey (PSS). We considered as diffuse cluster sources the radio features with surface brightness greater than the 3 $\sigma$ level, which were found to be: i) resolved, ii) not associated with bright galaxies, iii) not clearly related to extended radio galaxies, iv) not simply attributable to the blend of pointlike sources. Thanks to the large images retrieved from the NVSS, we are confident that we could easily recognize and avoid the side lobes of near strong sources possibly present because of dynamical range problems. We looked into the literature and in the VLA Faint Images of the Radio Sky at Twenty-centimeters (FIRST) Survey (Becker {\it et al.~} 1995) to discriminate the diffuse sources from discrete unrelated radio sources. In some cases, where no high resolution data are available and the radio structure is ambiguous, we considered the halo and relic candidates as uncertain. We included in our sample also a few cases where the evidence of an extended radio emission is marginal but the presence of a diffuse emission was already known from the literature (see notes to the individual clusters). \section {Results} With the above criteria, we have found diffuse extended emission in 29 clusters, listed in Table 1. In 9 clusters, the presence of a diffuse extended radio source is uncertain, because of the possible contamination of discrete radio galaxies or because the diffuse radio emission is very faint. These clusters are indicated by ``u'' in column 7 of Table 1. \begin{figure} \special{psfile=fig2.ps hoffset=-5 voffset=-325 hscale=60 vscale =60} \vspace{8 cm} \caption{ Redshift distributions of the clusters with NVSS image and of the clusters with a diffuse source (shaded).} \label{} \end{figure} The redshift distribution of clusters with diffuse sources with respect to that of the inspected clusters is presented in Fig. 2. We remind that we could have missed extended sources because of their low surface brightness. We also are aware that we can miss very extended sources ($>$ 1 Mpc) in the clusters at lowest redshifts because of the poor sampling of short baselines. This implies that the sample of diffuse sources found in this search might be slightly biased toward sources at high redshift. On the other hand, we note that in distant clusters diffuse extended sources could be, in some cases, considered as discrete sources because of the large beam. According to the literature, we classify the detected extended source as halos, if they are centrally located in the cluster, or relics if they are peripheral. In the case that peripheral relics are projected onto the cluster center, they would be classified as halo sources. More detailed observations will be necessary for a correct classification. The list presented in Table 1 contains most of the well known radio halos and relics, but also includes 18 clusters where this is the first indication of the existence of a diffuse source (see last column in Table 1). Among the latter, 11 are considered good candidates, while 7 are uncertain. The images of all the diffuse radio sources, overlaid onto the red images of the PSS, are given in Fig. 3. We present in Table 2 the parameters of the halo and relic candidates, excluding the uncertain diffuse sources, for which the measurement of flux density and extension is very difficult because strong contaminating sources are often present. The radio flux density refers to the extended emission after subtraction of obvious discrete sources. The size is the largest dimension of the diffuse radio emission. We are aware that flux densities and sizes can be quite underestimated, because of the low sensitivity to extended structures. Nevertheless, more than half of the sources given in Table 2 have a size larger than 1 Mpc. For the relics, we also give the projected distance between the approximate centroid of the radio emitting region and the X-ray cluster center. \begin{table*} \caption{List of cluster with halos or relics} \begin{flushleft} \begin{tabular}{llllllll} \hline \noalign{\smallskip} Name & z & RA (J2000) & DEC & T & L$_X$(0.1-2.4) & Radio Type & Previous \\ & & h~~m~~s & $^{\circ}$~~$^{\prime}$~~$^{\prime\prime}$ & keV & 10$^{44}$ erg s$^{-1}$ \\ \noalign{\smallskip} \hline \noalign{\smallskip} A~13 & 0.0943$^M$ & 00 13 32.2 & --19 30 03.6 & 4.3$^e$ & 2.24 & R & n \\ A~2744 & 0.3080 & 00 14 16.1 & --30 22 58.8 & 11.04$^{AF}$ & 22.05 & H & n \\ A~22 & 0.1310 & 00 20 38.6 & --25 43 19.2 & 6.3$^e$ & 5.31 & u & n \\ A~85 & 0.0555$^P$ & 00 41 48.7 & --09 19 04.8 & 6.2 & 8.38 & R & y \\ A~115 & 0.1971 & 00 55 59.8 & +26 22 40.8 & 9.8$^e$ & 14.57 & R & n \\ A~133 & 0.0603 & 01 02 45.1 & --21 52 48.0 & 3.8 & 3.57 & u & n \\ A~209 & 0.2060 & 01 31 50.9 & --13 36 28.8 & 9.6$^e$ & 13.75 & u & n \\ A~401 & 0.0739 & 02 58 56.9 & +13 34 22.8 & 7.8 & 9.88 & H & y \\ A~520 & 0.2030 & 04 54 07.4 & +02 55 12.0 & 8.33$^{AF}$ & 14.20 & H & n \\ A~545 & 0.1540 & 05 32 23.3 & --11 32 09.6 & 5.5 & 9.29 & H & n \\ A~548b & 0.0424$^{DHK}$ & 05 45 27.8 & --25 54 21.6 & 2.4 & 0.30 & R & n \\ A~665 & 0.1818 & 08 30 57.4 & +65 51 14.4 & 9.03$^{AF}$ & 16.22 & H & y\\ A~754 & 0.0542 & 09 09 01.4 & --09 39 18.0 & 8.7 & 8.01 & u & y \\ A~773 & 0.2170 & 09 17 54.0 & +51 42 57.6 & 9.29$^{AF}$ & 12.52 & H & n \\ A~1300 & 0.3071$^L$ & 11 31 54.9 & --19 54 50.4 & 5$^L$ & 23.40 & H & y \\ A~1664 & 0.1276 & 13 03 44.2 & --24 15 21.6 & 6.5$^{A}$ & 5.36 & R & n \\ A~1758a & 0.2800 & 13 32 45.3 & +50 32 52.8 & 8.7$^e$ & 11.22 & u & n \\ A~1914 & 0.1712 & 14 26 02.2 & +37 50 06.0 & 10.7$^e$ & 17.93 & H & n \\ A~2069 & 0.1145 & 15 24 09.8 & +29 55 15.6 & 7.8$^e$ & 8.74 & u & n \\ A~2142 & 0.0894 & 15 58 22.1 & +27 13 58.8 & 11.0 & 20.74 & u & y \\ A~2163 & 0.2080 & 16 15 49.4 & --06 09 00 & 13.83$^{AF}$ & 37.50 & H & y \\ A~2218 & 0.1710 & 16 35 52.8 & +66 12 50.4 & 7.05$^{AF}$ & 8.99 & H & y \\ A~2219 & 0.2281 & 16 40 22.5 & +46 42 21.6 & 12.42$^{AF}$ & 19.80 & u & n \\ A~2256 & 0.0581 & 17 04 02.4 & +78 37 55.2 & 7.5 & 7.05 & R & y \\ A~2255 & 0.0809 & 17 12 45.1 & +64 03 43.2 & 7.3 & 4.79 & H & y \\ A~2254 & 0.1780 & 17 17 46.8 & +19 40 48.0 & 7.2$^e$ & 7.19 & H & n \\ A~2319 & 0.0555 & 19 21 05.8 & +43 57 50.4 & 9.3$^{AF}$ & 13.71 & H & y \\ A~2345 & 0.1760 & 21 26 58.6 & --12 08 27.6 & 8.2$^e$ & 9.93 & R+R & n \\ A~2390 & 0.2329 & 21 53 36.7 & +17 41 32.2 1 & 10.13$^{AF}$ & 21.25 & u & n \\ \noalign{\smallskip} \hline \label{olog} \end{tabular} \end{flushleft} \end{table*} \vfill\eject Caption. Col. 1: cluster name; Col. 2: redshift; Cols. 3 and 4: coordinates of the X-ray cluster center; Col. 5: temperature, where ``e'' indicates that the temperature has been estimated from the L$_X$-kT relation; Col. 6: X-ray luminosity in the ROSAT band (0.1-2.4 keV); Col. 7: type of the diffuse radio source (H = halo, R = relic, u = uncertain); Col. 8: previous knowledge in the literature of the existence of a diffuse source in this cluster (n = no; y = yes, reference given in Sect. 4). \par The data in Cols. 2, 3, 4, 5 and 6 are taken from Ebeling {\it et al.~} (1996), except where a more recent reference is given. References are as follows: A = Allen {\it et al.~} 1995; AF: Allen \& Fabian 1998; DHK = Den Hartog \& Katgert 1996; L = L\'emonon {\it et al.~} 1997; M = Mazure {\it et al.~} 1996; P = Pislar {\it et al.~} 1997. \vfill\eject \begin{figure} \vspace{6 cm} \caption{ Radio images of the diffuse sources (contours), overlaid onto the optical image from the PSS (grey-scale). Contour levels are 0.9, 1.35, 2, 4, 8, 16, 32, 64, 128, 256 mJy/beam.} \label{} \end{figure} \begin{table} \caption{Parameters of the halo and relic candidates} \begin{flushleft} \begin{tabular}{llllll} \hline \noalign{\smallskip} Name & Flux & $\theta$ & LLS & Dist. & Power \\ & mJy & $^{\prime}$ & kpc & kpc & 10$^{24}$ W Hz$^{-1}$ \\ \noalign{\smallskip} \hline \noalign{\smallskip} A~13 & 34 & 6.4 & 880 & 150 & 1.30 \\ A~2744 & 38 & 5.4 & 1700 & -- & 15.5 \\ A~85 & 46 & 5.5 & 480 & 530 & 0.61 \\ A~115 & 80 & 6.2 & 1500 & 1050 & 1.34 \\ A~401 & 25 & 5.3 & 590 & -- & 0.59 \\ A~520 & 38 & 4.4 & 1080 & -- & 6.74 \\ A~545 & 41 & 7.4 & 1500 & -- & 4.18 \\ A~548b & 50 & 5.3 & 360 & 620 & 0.39 \\ A~665 & 31 & 4.8 & 1100 & -- & 4.41 \\ A~773 & 14 & 3.1 & 800 & -- & 2.84 \\ A~1300 & 14 & 2.5 & 780 & -- & 5.7 \\ A~1664 & 107 & 8.0 & 1400 & 1350 & 7.50 \\ A~1914 & 50 & 4.4 & 960 & -- & 6.31 \\ A~2163 & 55 & 6.0 & 1500 & -- & 10.2 \\ A~2218 & 9 & 2.3 & 510 & -- & 1.13 \\ A~2256 & 397 & 12.1 & 1100 & 590 & 5.77 \\ A~2255 & 45 & 5.2 & 630 & -- & 1.27 \\ A~2254 & 32 & 5.1 & 1140 & -- & 4.36 \\ A~2319 & 23 & 4.8 & 420 & -- & 0.30 \\ A~2345 & 92 & 5.4 & 1200 & 910 & 12.3 \\ & 69 & 7.0 & 1560 & 2050 & 9.20 \\ \noalign{\smallskip} \hline \label{olog} \end{tabular} \end{flushleft} \par\noindent Caption. Col. 1: cluster name; Col. 2: Flux density at 1.4 GHz, after subtraction of obvious discrete sources; Col. 3: maximum angular size; Col. 4: largest linear size; Col. 5: approximate distance from the cluster center, in the case of relics; Col. 6: monochromatic radio power at 1.4 GHz. \end{table} \section {Individual sources} In the following, we give comments on the individual sources and discuss their reliability. \par\noindent {\it A~13}. The diffuse radio source is not much displaced from the cluster center, however it is elongated in shape and does not include the central cluster galaxies. According to the data reported by Slee {\it et al.~} (1996), the radio sources identified with cluster galaxies account for a total flux of 3.9 mJy, confirming the presence of extended radio emission. \par\noindent {\it A~2744}. Beside the centrally located radio emission, there is also extended structure toward NE which could be either a relic or still related to the halo. \par\noindent {\it A~22}. The diffuse source could be the due to a Wide Angle Tailed radio galaxy plus a Narrow Angle Tailed radio galaxy. It is located at about 14$^{\prime}$~ from the cluster center, corresponding to $\sim$2.5 Mpc and would therefore be a relic. \par\noindent {\it A~85}. The relic in this cluster has been recently studied by Bagchi {\it et al.~} (1998). \par\noindent {\it A~115}. This cluster shows a double morphology in X-ray. The relic belongs to the northern clump. \par\noindent {\it A~133}. Komissarov and Gubanov (1994) report the existence of a very steep spectrum radio source ($\alpha >$ 2) coincident with the first ranked galaxy of this cluster. The high resolution image published by Slee {\it et al.~} (1994) shows a northern diffuse component of $\sim$100 kpc, probably not related to the radio galaxy. The emission detected in the NVSS extends to the South where another discrete radio source is present (see also the image by Owen {\it et al.~} 1993). The existence of a cluster-wide diffuse emission is uncertain, as well as its connection to the steep spectrum northern diffuse component. \par\noindent {\it A~209}. The presence of extended emission is uncertain due to the existence of strong discrete sources. \par\noindent {\it A~401}. The diffuse emission is very faint and located around the central cD galaxy, unlike the previous images by Harris {\it et al.~} (1980a) and Roland {\it et al.~} (1981). \par\noindent {\it A~520}. The diffuse emission is centrally located, but of irregular shape with the present sensitivity. \par\noindent {\it A~545}. Despite of the presence of a strong discrete source at the cluster center, the diffuse emission is easily visible. It is rather symmetric and centrally located. \par\noindent {\it A~548b}. This cluster consists of many X-ray subclumps (Davis {\it et al.~} 1995). We detect a diffuse source classified as a relic in the NW cluster region, but also extended emission is present around a radio galaxy to the North. In the following analysis, we only consider the NW relic. Further more sensitive observations should clarify whether the two extended features are bright regions of the same source. \par\noindent {\it A~665}. The presence of a halo was first reported by Moffet \& Birkinshaw (1989) and confirmed by Jones \& Saunders (1996). \par\noindent {\it A~754}. The existence of a halo in the center of this cluster was suggested by Harris {\it et al.~} (1980b). At the location of the previously reported halo, there are many cluster galaxies, which could account for the emission. Some diffuse emission of size $\sim$250 kpc is also detected in the NVSS in the eastern peripheral region. The overall cluster-wide extended emission in this cluster is considered uncertain. \par\noindent {\it A~773}. The diffuse emission is rather regular in shape and centrally located. The FIRST image shows 3 discrete radio sources whose total flux density is lower that that detected in the NVSS, confirming the presence of diffuse structure. \par\noindent {\it A~1300}. The presence of a central radio halo is reported by Lmonon {\it et al.~} (1998), who also classify the SW emission as a relic. Since from the NVSS map the relic is poorly resolved, we only consider in this paper the central radio halo. \par\noindent {\it A~1664}. The diffuse radio emission is located in the SW peripheral region of the cluster and shows a regular structure, unlike the relics in Coma and A~3667, which are generally elongated. The X-ray brightness distribution in this cluster is centrally peaked, with an asymmetric extension in the direction of the diffuse radio source (Allen {\it et al.~} 1995). \par\noindent {\it A~1758a}. The strongest source to the S is identified with a cluster galaxy. It shows a Narrow Angle Tailed structure (O'Dea \& Owen 1985) with the tail oriented to the SE. In the FIRST image, the easternmost structure is resolved in two compact sources, while the extended emission in between is not detected. It could be a faint halo or the blend of individual radio sources. \par\noindent {\it A~1914}. A very steep spectrum radio source ($\alpha >$2) is reported by Komissarov \& Gubanov (1994). From higher resolution images (Roland {\it et al.~} 1985 and the image retrieved from the FIRST survey) it is evident that the discrete sources cannot account for the extended emission. \par\noindent {\it A~2069}. The diffuse emission is located to the SE with respect to the cluster center at a distance of $\sim$6 Mpc, therefore its classification as a cluster relic is uncertain. The image of this region retrieved from the FIRST survey shows a faint point-like source coincident with the southernmost peak, while the whole structure is resolved out. \par\noindent {\it A~2142}. The presence of a halo in this cluster was suggested by Harris {\it et al.~} (1977). The image here shows that the diffuse radio emission is $\sim$350 kpc in size and is located around a cluster galaxy. Therefore, we still consider uncertain the classification of this source as a cluster-wide radio halo. \par\noindent {\it A~2163}. The presence of a powerful and very extended radio halo has been reported by Herbig \& Birkinshaw (1994). \par\noindent {\it A~2218}. The existence of a small radio halo was reported by Moffet \& Birkinshaw (1989). Here, the diffuse source is barely visible to the SW of the strongest radio source. \par\noindent {\it A~2219}. The strong radio source at the cluster center has a tailed structure (as detected in the FIRST survey). It is difficult to safely establish the existence of a diffuse radio emission. \par\noindent {\it A~2256}. The extended emission detected here is the brightest region of the complex diffuse radio source in A~2256. This region studied by R\"ottgering {\it et al.~} (1994) is at the cluster periphery, while a very low brightness emission, not visible here, permeates the cluster center as detected at 610 MHz by Bridle \& Fomalont (1976). \par\noindent {\it A~2255}. This cluster contains a central radio halo and a peripheral relic (Burns {\it et al.~} 1995, Feretti {\it et al.~} 1997a). Both features are visible in the NVSS, although very faint and at the limit of significance. Based only on the NVSS image, the existence of diffuse sources in this cluster would be considered very uncertain. We note, however, that the radio halo in A~2255 is best imaged at 90 cm, while at 20 cm the halo structure is spotty and irregular also with higher sensitivity observations (Feretti {\it et al.~} 1997). \par\noindent {\it A~2254}. The extended emission is centrally located and shows a regular structure. \par\noindent {\it A~2319}. This radio halo has been studied in detail by Feretti {\it et al.~} (1997b). \par\noindent {\it A~2345}. There are two peripheral extended sources in this cluster located approximately on opposite sides with respect to the cluster center, at distance of $\sim$0.9 and $\sim$2 Mpc. If both sources will be confirmed as diffuse relics, this cluster will be quite peculiar. The only known cluster with 2 relics is A~3667 (R\"ottgering {\it et al.~} 1997). \par\noindent {\it A~2390}. The extended emission in this cluster is uncertain, because of the presence of a discrete strong source (Owen {\it et al.~} 1993). \section {Discussion} The number of new diffuse halos and relics detected in the NVSS is rather high, especially considering that these sources are characterized by steep spectrum and therefore they are better imaged at frequencies lower than 1.4 GHz, and also taking into account the limited surface brightness sensitivity. \begin{figure} \special{psfile=fig4.ps hoffset=-5 voffset=-325 hscale=60 vscale =60} \vspace{8 cm} \caption{Distribution of the X-ray luminosity of the clusters searched for the presence of a diffuse source. The black squares indicate the halo and relic candidates, while the dashed squares indicate the uncertain sources (see Table 1). } \label{} \end{figure} The percentage of clusters showing diffuse sources is higher in clusters with high X-ray luminosity as can be deduced from Fig. 4 and Table 3. This trend can also be inferred from Fig. 2, due to the fact that the XBACs sample is flux limited. The percentage of clusters with diffuse sources is in the range 6\% to 9\% in the clusters with L$_X \le$ 10$^{45}$ erg s$^{-1}$ and becomes 27\% to 44\% in clusters with L$_X >$ 10$^{45}$ erg s$^{-1}$. This effect could partly reflect the mentioned bias in the sample against the detection of very extended diffuse sources in the nearby clusters. However, given that a lower limit to the redshift is assumed, the lack of diffuse sources in low X-ray luminosity clusters is real. This result is in agreement with previous findings that cluster-wide halos are present in massive clusters with high X-ray luminosity and high temperature (see e.g. Feretti \& Giovannini 1996). Therefore, the diffuse cluster sources are not a rare phenomenon if X-ray luminous clusters are considered. \begin{table} \caption{Occurrence of diffuse cluster sources} \begin{flushleft} \begin{tabular}{llll} \hline \noalign{\smallskip} Lum. range & N$_s$ & N$_f$ & N$_u$ \\ 10$^{44}$ erg s$^{-1}$ \\ \noalign{\smallskip} \hline \noalign{\smallskip} L$_X \le$ 10 & 173 & 11 & 4 \\ 10 $<$ L$_X \le$ 20 & 23 & 6 & 3 \\ 20 $<$ L$_X \le$ 30 & 7 & 2 & 2 \\ L$_X >$30 & 2 & 1 & 0 \\ \noalign{\smallskip} \hline \label{olog} \end{tabular} \end{flushleft} Caption. Col 1: X-ray luminosity range; Col 2: number of searched clusters; Col 3: number of clusters with diffuse halos or relics; Col 4: number of clusters with uncertain diffuse sources. \end{table} \begin{figure} \special{psfile=fig5.ps hoffset=-5 voffset=-325 hscale=60 vscale =60} \vspace{8 cm} \caption{Plot of monochromatic radio power at 1.4 GHz versus the cluster X-ray luminosity, for the halo and relic candidates. } \label{} \end{figure} We have looked for a possible correlation between the radio power of diffuse sources and the cluster X-ray luminosity. The plot of these parameters is given in Fig. 5. Although the data show a large scatter, the lack of high power radio sources in clusters of low X-ray luminosity is clear, while highly luminous X-ray clusters may have extended sources of either high or low radio power. Since the radio powers computed here may be largely underestimated because of missing flux, due to the steep spectrum of halos and relics and the poor sampling of short spacings, further investigation of this correlation is needed. A complete analysis of the connection between cluster properties and presence of diffuse sources is beyond the scope of this paper. We only note that the clusters A~2744, A~520, A~545, A~773, and A~2254, which are found here for the first time to host a central radio halo, are all non-cooling flow clusters (White {\it et al.~} 1997, Allen \& Fabian 1998). Likewise, all the Abell clusters listed by Allen \& Fabian (1998) as non-cooling flow clusters are found to host a radio halo, with the exception of A~2219, where the presence of a halo in considered uncertain but is possible. These results reinforce the strong correlation between the absence of a cooling flow and the presence of a radio halo at the cluster center. \section {Conclusions} We have presented in this paper new halo and relic candidates, found in the NVSS after inspection of a sample of clusters from the XBACs catalogue. We found 29 candidates, which can be divided as follows: \par\noindent $\bullet$ 11 clusters were already known from the literature. Out of them, 7 contain radio halos (A~401, A~665, A~1300, A~2163, A~2218, A~2255, A~2319), 2 contain relics (A~85, A~2256), while the remaining 2 (A~754, A~2142) cannot be confirmed on the basis of the NVSS images, and are therefore still considered uncertain. We note that in the clusters A~1300 and A~2255, also the existence of a peripheral relic is reported in the literature. \par\noindent $\bullet$ in 18 clusters, this is the first indication of the existence of a diffuse extended source. Among them, we found 6 clusters with halos (A~2744, A~520, A~545, A~773, A~1914, A~2254), and 5 clusters with relics (A~13, A~115, A~548b, A1664, A~2345). The cluster A~2345 contains actually 2 relics. The 7 clusters A~22, A~133, A~209, A~1758a, A~2069, A~2219 and A~2390 contain possible diffuse sources, which we indicate as uncertain. We found that the percentage of clusters showing diffuse sources is higher in clusters with high X-ray luminosity, being 6\%-9\% in clusters with L$_X\le$ 10$^{45}$ erg s$^{-1}$ and 27\%-44\% in clusters with L$_X >$ 10$^{45}$ erg s$^{-1}$. Therefore, the diffuse cluster sources are not a rare phenomenon if X-ray luminous clusters are considered. We have found no correlation between the radio power of diffuse sources and the cluster X-ray luminosity. However, we note that the radio powers may be largely underestimated because of missing flux in the NVSS images. Finally, we note that the large majority of clusters hosting a radio halo do not contain a cooling flow. This reinforces the strong correlation between the absence of a cooling flow and the presence of a radio halo in clusters. \begin{ack} We thank Dr. W.D. Cotton for discussions and advice on the NVSS. LF and GG acknowledge partial financial support from the Italian Space Agency (ASI) and from the Italian MURST. This research has made use of the NASA/IPAC Extragalactic Database (NED) which is operated by the Jet Propulsion Laboratory, Caltech, under contract with the National Aeronautics and Space Administration. \end{ack} \vskip 1 truecm {\bf References} \par\medskip\noindent Abell, G.O., Corwin, H.G., Olowin, R.P., 1989, ApJS 70, 1 \par\medskip\noindent Allen, S.W., Fabian, A.C., Edge, A.C., B\"ohringer, H., White, D.A., 1995, MNRAS 275, 741 \par\medskip\noindent Allen, S.W., Fabian, A.C., 1998, MNRAS 297, L57 \par\medskip\noindent Bagchi, J., Pislar, V., Lima Neto, G.B, 1998, MNRAS 296, L23 \par\medskip\noindent Becker, R.H., White, R.L., Helfand, D.J., 1995, ApJ 450, 559 \par\medskip\noindent Bridle, A.H., Fomalont, E.B., 1976, A\&A 52, 107 \par\medskip\noindent Burns, J.O., Roettiger, K., Pinkney, {\it et al.~}, 1995, ApJ 446, 583 \par\medskip\noindent Burns, J.O., Sulkanen, M.E., Gisler, G.R., Perley, R.A., 1992, ApJ 388, L49 \par\medskip\noindent Condon, J.J., Cotton, W.D., Greisen, E.W. Yin, Q.F., Perley, R.A., Taylor, G.B., Broderick, J.J., 1998, AJ 115, 1693 \par\medskip\noindent Davis, D.S. Bird, C.M., Mushotzky, R.F. Odewahn, S.C., 1995, ApJ 440, 48 \par\medskip\noindent Den Hartog, R., Katgert, P., 1996, MNRAS 279, 349 \par\medskip\noindent Ebeling, H., Voges, W., B\"ohringer, H., Edge, A.C., Huchra, J.P., Briel, U.G., 1996, MNRAS 281, 799 \par\medskip\noindent Feretti, L., Giovannini, G., 1996, In: Extragalactic Radio Sources, IAU Symp. 175, Eds. R. Ekers, C. Fanti \& L. Padrielli, Kluwer Academic Publisher, p. 333 \par\medskip\noindent Feretti, L., B\"ohringer, H., Giovannini, G., Neumann D., 1997a, AA 317, 432 \par\medskip\noindent Feretti, L., Giovannini, G. B\"ohringer, H., 1997b, New Astronomy, 2, 501 \par\medskip\noindent Feretti, L., Giovannini, G., 1998, in {\it Untangling Coma Berenices: A new view of an old Cluster}, Word Scientific Publishing Co Pte Ltd, p 123 \par\medskip\noindent Giovannini, G., Feretti, L., Venturi, T., Kim, K.-T., Kronberg, P.P., 1993, ApJ, 406, 399 \par\medskip\noindent Harris, D.E., Bahcall, N.A., Strom, R.G., 1977, A\&A 60, 27 \par\medskip\noindent Harris, D.E., Miley, G.K., 1978, A\&AS 34, 117 \par\medskip\noindent Harris, D.E., Kapahi, V.K., Ekers, R.D., 1980a, A\&AS 39, 215 \par\medskip\noindent Harris, D. E., Pineda, F. J., Delvaille, J. P., Schnopper, H. W., Costain, C. H., Strom, R. G., 1980b, A\&A 90, 283 \par\medskip\noindent Herbig, T., Birkinshaw, M., 1994, BAAS, Vol. 26, No 4, 1403 \par\medskip\noindent Jones, M., Saunders, R., 1996, in {\it R\"ontgenstrahlung from the Universe}, H.U. Zimmermann, J.E. Tr\"umper \& H. Yorke Eds., MPE Report 263, p. 553 \par\medskip\noindent Komissarov, S.S., Gubanov, A.G., 1994, A\&A 285, 27 \par\medskip\noindent L\'emonon, L., Pierre, M., Hunstead, R., {\it et al.~}, 1997, A\&A 326, 34 \par\medskip\noindent Lmonon, {\it et al.~}, 1998, astro-ph 9811467 \par\medskip\noindent Mazure, A., Katgert, P, Den Hartog, R., {\it et al.~}, 1996, A\&A 310, 31 \par\medskip\noindent Moffet, A. T., Birkinshaw, M., 1989, AJ 98, 1148 \par\medskip\noindent O'Dea, C.P., Owen, F.N., 1985, AJ 90, 927 \par\medskip\noindent Owen, F.N., White, R.A., Ge, J-P., 1993, ApJS 87, 135 \par\medskip\noindent Pislar, V., Durret, F., Gerbal, D., Lima Neto, G.B., Slezak, E., 1997, A\&A 322, 53 \par\medskip\noindent Roland, J. Sol, H. Pauliny-Toth, I. Witzel, A. 1981, A\&A 100, 7 \par\medskip\noindent Roland, J., Hanish, R.J., V\'eron, P. Fomalont, E., 1985, A\&A 148, 323 \par\medskip\noindent R\"ottgering, H., Snellen, I., Miley, G., {\it et al.~}, 1994, ApJ 436, 654 \par\medskip\noindent R\"ottgering, H.J.A., Wieringa, M.H., Hunstead, R.W, Ekers, R.D., 1997, MNRAS 290, 577 \par\medskip\noindent Slee, O.B., Roy, A.L., Savage, A, 1994, Aust. J. Phys. 47, 145 \par\medskip\noindent Slee, O.B., Roy, A.L., Andernach, H., 1996, Aust. J. Phys. 49, 977 \par\medskip\noindent White, D.A., Jones, C., Forman, W., 1997, MNRAS 292, 419 \end{document}
\section{Introduction} Standard explanations of the observed atmospheric and solar neutrino anomalies~\cite{SK} require neutrino oscillations between different species, which imply that neutrinos are massive, with mass-squared differences of at most $10^{-2}\ {\rm eV}^2$. On the other hand, if neutrinos are to play an essential role in the large scale structure of the universe, the sum of their masses must be a few eV, and therefore they must be almost degenerate. This scenario has recently attracted much attention~\cite{GG}-\cite{degenerofilia}. In this paper we will analyze under which circumstances the ``observed'' mass differences arise naturally (or not), as a radiative effect, in agreement with all the available experimental data. \vspace{0.2cm} Let us briefly review the current relevant experimental constraints on neutrino masses and mixing angles. Observations of atmospheric neutrinos require $\nu_\mu-\nu_\tau$ oscillations driven by a mass splitting and a mixing angle in the range \cite{range} \begin{Eqnarray} 5\times10^{-4}\ {\rm eV}^2 < &\Delta m^2_{at}& < 10^{-2}\ {\rm eV}^2\ , \nonumber\\ \sin^22\theta_{at}&>&0.82\ . \label{atm} \end{Eqnarray} On the other hand, there are three main explanations of the solar neutrino flux deficits, requiring oscillations of electron neutrinos into other species. The associated mass splittings and mixing angles depend on the type of solution: \noindent \vspace{0.2cm} Small angle MSW (SAMSW) solution: \begin{Eqnarray} 3\times10^{-6}\ {\rm eV}^2 < &\Delta m^2_{sol}& < 10^{-5}\ {\rm eV}^2, \nonumber\\ 4\times10^{-3} < &\sin^22\theta_{sol}& < 1.3\times10^{-2}. \label{SAMSW} \end{Eqnarray} \vspace{0.2cm} Large angle MSW (LAMSW) solution: \begin{Eqnarray} 10^{-5}\ {\rm eV}^2 < &\Delta m^2_{sol}& < 2\times 10^{-4}\ {\rm eV}^2, \nonumber\\ 0.5 < &\sin^22\theta_{sol}& < 1. \label{LAMSW} \end{Eqnarray} \vspace{0.2cm} Vacuum oscillations (VO) solution: \begin{Eqnarray} 5\times10^{-11}\ {\rm eV}^2 < &\Delta m^2_{sol}& < 1.1\times10^{-10}\ {\rm eV}^2, \nonumber\\ \sin^22\theta_{sol} &>& 0.67\, . \label{VO} \end{Eqnarray} Let us remark the hierarchy of mass splittings between the different species of neutrinos, $\Delta m^2_{sol}\ll\Delta m^2_{at}$, which is apparent from eqs.(\ref{atm}, \ref{SAMSW}--\ref{VO}). This hierarchy should be reproduced by any natural explanation of those splittings. As it has been shown in ref.~\cite{GG} the small angle MSW solution is unplausible in a scenario of nearly degenerate neutrinos, so we are left with the LAMSW and VO possibilities. An important point concerns the upper limit on $\sin^22\theta_{sol}$ in the LAMSW solution. According to ref.\cite{Giunti} the absolute limit $\sin^22\theta_{sol}=1$ is forbidden at 99.8\%, but with no indication about the tolerable upper limit. In this sense a conservative upper bound $\sin^22\theta_{sol}<0.99$ can be adopted. However, recent combined analysis of data, including the day--night effect \cite{upper}, indicate that even the $\sin^22\theta_{sol}=1$ possibility is allowed at 99\% for $2\times 10^{-5}\ {\rm eV}^2 < \Delta m^2_{sol} < 1.7\times 10^{-4}\ {\rm eV}^2$, although there may be problems to define the upper limit on $\sin^22\theta_{sol}$ in a precise sense \cite{pc}. For the moment we will not adopt an upper bound on $\sin^22\theta_{sol}$, but later on we will show the effect of such a bound on the results, which turns out to be dramatic. Finally, the non-observation of neutrinoless double $\beta$-decay puts an upper bound on the $ee$ element of the Majorana neutrino mass matrix, namely ~\cite{Baudis} \begin{Eqnarray} {\cal M}_{ee}<B=0.2\ {\rm eV}. \label{B} \end{Eqnarray} Concerning the cosmological relevance of neutrinos, as mentioned above, we will assume $\sum m_{\nu_i} = O({\rm eV})$. In particular, we will take $\sum m_{\nu_i} = 6$ eV as a typical possibility, although as we will see, the results do not depend essentially on the particular value. It should be mentioned here that Tritium $\beta$-decay experiments indicate $m_{\nu_i} < 2.5$ eV for any mass eigenstate with a significant $\nu_e$ component \cite{triti}. Using standard notation, we define the effective mass term for the three light (left-handed) neutrinos in the flavour basis as \begin{Eqnarray} {\cal L}=-\frac{1}{2} \nu^T {\cal M_\nu} \nu\;+\;{\rm h.c.} \label{Mnu} \end{Eqnarray} where ${\cal M_\nu}$ is the neutrino mass matrix. This is diagonalized in the usual way \begin{Eqnarray} {\cal M_\nu} = V^* D\, V^\dagger, \label{V} \end{Eqnarray} where $V$ is a unitary `CKM' matrix, relating the flavor eigenstates to the mass eigenstates \begin{Eqnarray} \pmatrix{\nu_e \cr \nu_\mu\cr \nu_\tau\cr}= \pmatrix{c_2c_3 & c_2s_3 & s_2e^{-i\delta}\cr -c_1s_3-s_1s_2c_3e^{i\delta} & c_1c_3-s_1s_2s_3e^{i\delta} & s_1c_2\cr s_1s_3-c_1s_2c_3e^{i\delta} & -s_1c_3-c_1s_2s_3e^{i\delta} & c_1c_2\cr}\, \pmatrix{\nu_1\cr \nu_2\cr \nu_3\cr}\,, \label{CKM} \end{Eqnarray} where $s_i$ and $c_i$ denote $\sin\theta_i$ and $\cos\theta_i$ respectively. $D$ may be written as \begin{Eqnarray} D=\pmatrix{m_1e^{i\phi}&0&0\cr 0&m_2e^{i\phi'}&0\cr 0&0&m_3\cr}\, . \end{Eqnarray} It should be mentioned here that, for a given mass matrix, the $\theta_i$ angles are not uniquely defined, unless one gives a criterion to order the mass eigenvectors $\nu_i$ in eq.(\ref{CKM}) (e.g. $m_{\nu_1}^2<m_{\nu_2}^2<m_{\nu_3}^2$). Of course, the corresponding $V$ matrices differ just in the ordering of the columns, and thus are physically equivalent. In this notation, constraint (\ref{B}) reads \begin{Eqnarray} {\cal M}_{ee} \equiv \big\vert m_{\nu_1}\,c_2^2c_3^2e^{i\phi} + m_{\nu_2}\,c_2^2s_3^2e^{i\phi'} + m_{\nu_3}\, s_2^2\,e^{i2\delta} \big\vert < B \,. \label{doublebeta} \end{Eqnarray} As it has been put forward by Georgi and Glashow in ref.~\cite{GG}, a scenario of nearly degenerate neutrinos plausibly leads to $\theta_2\simeq 0$. In that case, and for $m_\nu=2$ eV, eq.(\ref{doublebeta}) yields $\sin^2 2\theta_3\geq 0.99$, which, as discussed above, might be in conflict with the LAMSW solution to the solar neutrino anomaly. However, according to some fits, $\theta_2$ could be as large as $27^0$ or even larger \cite{range}. Therefore, although small values of $\theta_2$ are clearly preferred, in some acceptable cases a non-negligible contribution of $\sin^2 \theta_2$ in eq.(\ref{doublebeta}) is enough to relax the above mentioned stringent bound on $\sin^2 2\theta_3$, and in fact we will see some examples of this feature later on. Our criterion throughout the paper will be to keep eq.(\ref{B}) [or the complete eq.(\ref{doublebeta})] as the neutrinoless double $\beta$-decay constraint, without demanding any extra condition on $\theta_3$, $\theta_2$ \cite{vissani}. In any case, we will see that in all viable cases $\sin^2 2\theta_3\geq 0.99$ and $\theta_2$ is very small. \vspace{0.2cm} Let us now discuss the strategy we have followed to analyze if the required mass splittings and mixing angles can or cannot arise in a natural way through radiative corrections. Following ref.~\cite{GG} (some of their arguments have been discussed above), the scenario of nearly degenerate neutrinos should be close to a bimaximal mixing, which constrains the texture of the mass matrix ${\cal M}_\nu$ to be~\cite{GG,barger} \begin{Eqnarray} {\cal M}_{b} = m_\nu \,\pmatrix{ 0 & {\displaystyle{1\over\sqrt2}} & {\displaystyle{1\over\sqrt2}}\cr {\displaystyle{1\over\sqrt2}} & {\displaystyle{1\over2}} & - {\displaystyle{1\over2}}\cr {\displaystyle{1\over\sqrt2}} & - {\displaystyle{1\over2}} & {\displaystyle{1\over2}}\cr }\;, \label{MGG} \end{Eqnarray} where $m_\nu$ is a general mass scale. ${\cal M}_{b}$ can be diagonalized by a $V$ matrix \begin{Eqnarray} {V}_{b} = \,\pmatrix{ {\displaystyle{-1\over\sqrt2}} & {\displaystyle{1\over\sqrt2}} & 0 \cr {\displaystyle{1\over 2}} & {\displaystyle{1\over2}} & {\displaystyle{-1\over\sqrt2}} \cr {\displaystyle{1\over2}} & {\displaystyle{1\over2}} & {\displaystyle{1\over\sqrt2}} \cr }\;, \label{VGG} \end{Eqnarray} leading to exactly degenerate neutrinos: $D=m_\nu\ {\rm diag}(-1, 1, 1)$ and $\theta_2=0$, $\sin^2 2\theta_3=\sin^2 2\theta_1=1$. Let us remind that in this scenario only the LAMSW and VO solutions to the solar anomaly are acceptable. The nice aspect of ${\cal M}_{b}$ suggests that it could be generated at some scale by interactions obeying appropriate continuous or discrete symmetries. This is an interesting issue \cite{sym}, which we will not address in this paper. On the other hand, in order to be realistic, the effective matrix ${\cal M}_\nu$ at low energy should be certainly close to ${\cal M}_{b}$, but it must be {\em slightly} different in order to account for the mass splittings given in eqs.(\ref{atm},\ref{LAMSW},\ref{VO}). The main goal of this paper is to explore whether the appropriate splittings (and mixing angles) can be generated or not through radiative corrections; more precisely, through the running of the renormalization group equations (RGEs) between the scale at which ${\cal M_\nu}$ is generated and low energy. The output of this analysis can be of three types: \begin{description} \item[{\em i)}] All the mass splittings and mixing angles obtained from the RG running are in agreement with all experimental limits and constraints. \item[{\em ii)}] Some (or all) mass splittings are much larger than the acceptable ranges. \item[{\em iii)}] Some (or all) mass splittings are smaller than the acceptable ranges, and the rest is within. \end{description} Case {\em (i)} is obviously fine. Case {\em (ii)} is disastrous. The only way-out would be an extremely artificial fine-tuning between the initial form of ${\cal M}_\nu$ and the effect of the RG running. Hence we consider this possibility unacceptable. Finally, case {\em (iii)} is not fine, in the sense that the RGEs fail to explain the required modifications of ${\cal M}_\nu$. However, it leaves the door open to the possibility that other (not specified) effects could be responsible for them. In that case, the RGE would not spoil a fine pre-existing structure. Consequently, we consider this possibility as undecidable. We will see along the paper different scenarios corresponding to the three possibilities. Concerning the mixing angles, it is worth stressing that, due to the two degenerate eigenvalues of ${\cal M}_{b}$, ${V}_{b}$ is not uniquely defined (${V}_{b}$ times any rotation on the plane of degenerate eigenvalues is equally fine). Hence, once the ambiguity is removed thanks to the small splittings coming from the RG running, the mixing angles may be very different from the desired ones. However, if those cases correspond to the previous {\em (iii)} possibility, they are still of the ``undecidable'' type with respect to the mixing angles, since the modifications of ${\cal M}$ (of non-specified origin) needed to reproduce the correct mass splittings will also change dramatically the mixing angles. In section 2, we first examine the general case in which the neutrino masses arise from an effective operator, remnant from new physics entering at a scale $\Lambda$. In this framework, we take a low energy point of view, assuming a bimaximal-mixing mass structure at the scale $\Lambda$ as an initial condition. In this section we do not consider possible perturbations of that initial condition coming from high-energy effects. We find this case to be of the undecidable type [possibility {\em (iii)} above], except for the VO solution, which is excluded. In section 3 we go one step further and consider in detail a particularly well motivated example of the previous case: the see-saw scenario. We include here the high energy effects of the new degrees of freedom above the scale $\Lambda$ (identified now with the mass of the right-handed neutrinos). We find regions of parameter space where the neutrino spectrum and mixing angles fall naturally in the pattern required to explain solar (LAMSW solution) and atmospheric neutrino anomalies, which we find remarkable. We complement the numerical results with compact analytical formulas which give a good description of them, and allow to understand the pattern of mass splittings and mixing angles obtained. We also present plausible textures for the neutrino Yukawa couplings leading to a good fit of the oscillation data. Finally we draw some conclusions. \section{${\cal M}_\nu$ as an effective operator} In this section we consider the simplest possibility that the Majorana mass matrix for the left-handed neutrinos, ${\cal M}_\nu$, is generated at some high energy scale, $\Lambda$, by some unspecified mechanism (we allow $\Lambda$ to vary from $M_p$ to $M_Z$). Assuming that the only light fields below $\Lambda$ are the Standard Model (SM) ones, the lowest dimension operator producing a mass term of this kind is uniquely given by \cite{eff} \begin{Eqnarray} -\frac{1}{4}\kappa \nu^T \nu H H \;+\;{\rm h.c.} \label{kappa} \end{Eqnarray} where $\kappa$ is a matricial coupling and $H$ is the ordinary (neutral) Higgs. Obviously, ${\cal M}_\nu=\frac{1}{2}\kappa \langle H\rangle^2$. The effective coupling $\kappa$ runs with the scale below $\Lambda$, with a RGE given by \cite{Babu} \begin{Eqnarray} 16\pi^2 \frac{d \kappa}{dt}= \left[-3g_2^2+2\lambda+6Y_t^2+2 {\rm Tr}{\bf Y_e^\dagger Y_e} \right]\kappa -\frac{1}{2}\left[\kappa{\bf Y_e^\dagger Y_e} + ({\bf Y_e^\dagger Y_e})^T\kappa\right]\ , \label{rg1} \end{Eqnarray} where $t=\log \mu$, and $g_2,\lambda, Y_t, {\bf Y_e}$ are the $SU(2)$ gauge coupling, the quartic Higgs coupling, the top Yukawa coupling and the matrix of Yukawa couplings for the charged leptons, respectively. Let us note that the RGE depends on $\lambda$, and thus on the value of the Higgs mass, $m_H$. We have taken $m_H=150$ GeV throughout the paper, but in any case the dependence is very small (it slightly affects the overall neutrino mass scale but not the relative splittings). In the scenario in which we are interested (almost degenerate neutrinos), the simplest assumption about the form of $\kappa$ is that the interactions responsible for it produce the bimaximal mixing texture of eq.(\ref{MGG}). Hence \begin{Eqnarray} {\cal M}_\nu(\Lambda)=\frac{1}{2}\kappa(\Lambda) \langle H\rangle^2={\cal M}_{b}\, . \label{kappaboundary} \end{Eqnarray} Clearly, the last term of the RGE (\ref{rg1}) will slightly perturb the initial form of ${\cal M}_{\nu}$, so we expect at low energy small mass splittings, and mixing angles different from the bimaximal case. In order to gain intuition on the final (low energy) form of ${\cal M}_{\nu}$, and the corresponding mass splittings and mixing angles, it is convenient to neglect for a while all the charged lepton Yukawa couplings but $Y_\tau$. Then $\kappa$ maintains its form along the running, except for the third row and column: \begin{Eqnarray} {\cal M}_\nu(\mu_0)\propto\pmatrix{ 0 & {\displaystyle{1\over\sqrt2}} & {\displaystyle{1\over\sqrt2}}(1+\epsilon)\cr {\displaystyle{1\over\sqrt2}} & {\displaystyle{1\over2}} & - {\displaystyle{1\over2}}(1+\epsilon)\cr {\displaystyle{1\over\sqrt2}}(1+\epsilon) & - {\displaystyle{1\over2}}(1+\epsilon) & {\displaystyle{1\over2}}(1+2 \epsilon)\cr }\, \label{Mpert} \end{Eqnarray} where $\mu_0$ is the low-energy scale, which can be identified with $M_Z$ and $\epsilon$ has positive sign. Therefore, the mass eigenvalues are proportional to $1, \ -1-{\epsilon}/{2}, \ 1+{3\epsilon}/{2}$, and the corresponding splittings (adopting the convention $m_{\nu_1}^2<m_{\nu_2}^2<m_{\nu_3}^2$) are \begin{Eqnarray} \Delta m^2_{12}\ =\ \frac{1}{2}\Delta m^2_{23}\ =\ \frac{1}{3}\Delta m^2_{13} \ \simeq\ m_\nu^2 {\epsilon}. \label{Deltaskappa} \end{Eqnarray} \begin{figure}[t] \centerline{ \psfig{figure=cein1.ps,height=12cm,width=10cm,bbllx=8.cm,% bblly=5.cm,bburx=23.cm,bbury=19.cm}} \caption {\footnotesize Dependence of neutrino mass splittings at low energy ($\Delta m_{ij}^2$ in ${\mathrm eV}^2$) with the cut-off scale $\Lambda ({\rm GeV}$). } \end{figure} Clearly, these mass splittings are incompatible with the required hierarchy $\Delta m^2_{sol}\ll\Delta m^2_{at}$, of eqs.(\ref{atm}, \ref{LAMSW}, \ref{VO}). We will discuss the size of these splittings shortly. Equation (\ref{Mpert}) is also useful to get an approximate form of the $V$ matrix responsible for the diagonalization of ${\cal M}_{\nu}$ \begin{Eqnarray} V\simeq\pmatrix{ {\displaystyle{1\over\sqrt3}} & -{\displaystyle{1\over\sqrt2}} & {\displaystyle{1\over\sqrt6}}\cr {\displaystyle{\sqrt{2\over3}}} & {\displaystyle{1\over2}} & - {\displaystyle{1\over2\sqrt3}}\cr 0 & {\displaystyle{1\over2}} & {\displaystyle {\sqrt3\over2}}\cr }\, , \label{Vkappa} \end{Eqnarray} which leads to mixing angles \begin{Eqnarray} \sin^2 2 \theta_1 = {9\over25},\;\;\; \sin^2 2 \theta_2 = {5\over9},\;\;\; \sin^2 2 \theta_3 = {24\over25}. \label{mixingsskappa} \end{Eqnarray} Clearly, these values are far away from the bimaximal mixing ones. In consequence, they are not acceptable. The previous failures of the scenario considered in this section in order to reproduce the mass splittings and the mixing angles indicate that we are in one of the two possibilities {\em (ii)} or {\em (iii)} discussed in the Introduction. To go further, we need a numerical evaluation of the mass splittings, and thus of $\epsilon$. Solving the RGE (\ref{rg1}) at lowest order, we simply find \begin{Eqnarray} \epsilon =\frac{Y_\tau^2}{32\pi^2} \log\frac{\Lambda}{\mu_0}\, . \label{epsilonkappa} \end{Eqnarray} So, from eq.(\ref{Deltaskappa}) the mass splittings are typically of order $10^{-5}$ eV$^2$. This is confirmed by the complete numerical evaluation of the RGE , which gives the mass splittings shown in Fig.1. The first conclusion is that for any $\Lambda$ (even very close to $\mu_0$) the mass splittings are much larger than those required for the VO solution to the solar neutrino problem, $\Delta m^2_{sol}\sim 10^{-10}$ eV$^2$. Therefore, the effect of the RGE for this scenario is disastrous in the sense discussed in the Introduction for the possibility {\em (ii)}. In consequence the VO solution to the solar neutrino problem is excluded. For the LAMSW solution to the solar neutrino problem, things are a bit different. Clearly, the mass splittings obtained from the RGE analysis are within or below the required range, eq.(\ref{LAMSW}), for almost any value of the cut-off scale $\Lambda$. In addition, the mass splittings are clearly below the atmospheric range, eq.(\ref{atm}). Therefore, we are in the case {\em (ii)} discussed in the Introduction: The radiative corrections fail to provide an origin to the required mass splittings and mixing angles, but they will not destroy a suitable initial modification (from unspecified origin) of ${\cal M}_\nu$ at the $\Lambda$ scale. It is anyway remarkable that the radiative corrections place $\Delta m^2$ just in the right magnitude for the LAMSW solution. In the next section we will discuss a natural source for the initial modification of ${\cal M}_\nu$, which leads naturally to completely satisfactory (atmospheric plus LAMSW) scenarios. \section{${\cal M}_\nu$ from the see-saw mechanism} A natural way of obtaining small neutrino masses is the so-called see-saw mechanism \cite{seesaw} in which the particle content of the Standard Model is enlarged by one additional neutrino field (not charged under the SM group) per generation, $\nu_{\alpha,R}$ ($\alpha=e,\mu,\tau$). The Lagrangian reads \begin{Eqnarray} {\cal L}= -\bar{\nu}_R m_D \nu + \frac{1}{2} \bar{\nu}_R {\cal M} \bar{\nu}_R^T + {\mathrm h.c.} \end{Eqnarray} Here $m_D$ is a $3\times 3$ Dirac mass matrix with magnitude determined by the electroweak breaking scale, $m_D={\bf Y_\nu}\langle H\rangle$, where ${\bf Y_\nu}$ is the matrix of neutrino Yukawa couplings and $\langle H\rangle=246/\sqrt{2}$ GeV; and ${\cal M}$ is a $3\times 3$ Majorana mass matrix which does not break the SM gauge symmetry. Thus, the overall scale of ${\cal M}$, which we will denote by $M$, can be naturally many orders of magnitude higher than the electroweak scale. In that case, the low-energy effective theory, after integrating out the heavy $\nu_{\alpha,R}$ fields [whose masses are $O(M)$], is just the SM with left-handed neutrino masses given by \begin{Eqnarray} {\cal M}_\nu= m_D^T {\cal M}^{-1} m_D = {\bf Y_\nu}^T {\cal M}^{-1} {\bf Y_\nu} \langle H\rangle^2, \end{Eqnarray} suppressed with respect to the typical fermion masses by the inverse power of the large scale $M$. In this appealing framework, the degeneracy in neutrino masses can come about as a result of some symmetry (at some high energy, e.g. $M_p$) in the textures of ${\cal M}$ and ${\bf Y_\nu}$ such that ${\cal M}_\nu(M_p)={\cal M}_{b}$. Starting from this very symmetric condition at $M_p$ we make contact with the low energy neutrino mass matrix using the renormalization group. First we run ${\cal M}$ and ${\bf Y_\nu}$ from $M_p$ down to $M$, where the right-handed neutrinos are decoupled. Below that scale and all the way down to low energy we run ${\cal M}_\nu$ as an effective operator, in the same way as in the previous section. We can also think of the first stage of running between $M_p$ and $M$ as the high energy effects providing an starting form of ${\cal M_\nu}$ for the second stage of running, i.e. the only one considered in the previous section (the role of $\Lambda$ is played by $M$). This form, being slightly different from ${\cal M}_{b}$, may rescue the previous ``undecidable'' cases. In pursuing this idea, it is natural to assume \cite{altfe} that the structure leading to the relation ${\cal M}_\nu(M_p)={\cal M}_{b}$ occurs either in the Dirac matrix $m_D$ [while the Majorana matrix ${\cal M}$ is simply diagonal with equal eigenvalues (up to a sign)\footnote{Actually, the only condition we are imposing is that the eigenvalues are equal (up to a sign). A suitable transformation will diagonalize ${\cal M}$ to the form we assume.}] or in the Majorana matrix [with a simple diagonal Dirac matrix (again with eigenvalues equal up to a sign)]. We cannot expect a conspiracy between both matrices (which come from totally different physics) leading to ${\cal M}_\nu(M_p)={\cal M}_{b}$. \subsection{Textures of ${\cal M}$ and ${\bf Y_\nu}$ leading to bimaximal mixing} In the analysis of what structures for ${\cal M}$ or ${\bf Y_\nu}$ lead to ${\cal M}_\nu(M_p)={\cal M}_{b}$ for simplicity we only consider real matrices (thus avoiding potential problems with CP violation). If we assume that ${\bf Y_\nu}$ is proportional to the identity and all the structure arises from the Majorana mass matrix, then we simply need to impose ${\cal M}(M_p)\propto {\cal M}_{b}$ (note that ${\cal M}_{b}^{-1}\propto {\cal M}_{b}$ ). The alternative case is less trivial but equally simple. If ${\cal M}$ is of the form\footnote{If ${\cal M}$ is taken proportional to the identity, the Yukawa matrix ${\bf Y_\nu}$ must be chosen complex. Our choice of ${\cal M}$ avoids that complication and is physically equivalent.} \begin{Eqnarray} {\cal M}=M\pmatrix{-1 & 0 & 0\cr 0 & 1 & 0\cr 0& 0& 1}\, , \label{Mmaj} \end{Eqnarray} then it is easy to see that the most general form of ${\bf Y_\nu}$ satisfying ${\bf Y_\nu}^T {\cal M}^{-1} {\bf Y_\nu} \propto {\cal M}_{b}$ is \begin{Eqnarray} {\bf Y_\nu}= Y_\nu B V_{b}^T \, . \label{Ynu} \end{Eqnarray} Here $Y_\nu$ is the overall magnitude of ${\bf Y_\nu}$ and $B$ is a combination of two `boosts' \begin{Eqnarray} B=\pmatrix{ \cosh a & 0 & \sinh a \cr 0 & 1 & 0 \cr \sinh a & 0 & \cosh a \cr} \pmatrix{ \cosh b & \sinh b & 0 \cr \sinh b & \cosh b & 0 \cr 0 & 0 & 1 \cr} \, , \label{boosts} \end{Eqnarray} with two free parameters $a,b$. Actually, one could also take ${\bf Y_\nu}= Y_\nu RB V_{b}^T$, where $R$ is a rotation in the $(\mu,\tau)$ plane, but such rotations can be absorbed into a change of the $(\nu_R)_\alpha$ basis, $\nu_R\rightarrow R\nu_R$, with no modification in ${\cal M}$ and thus give physically equivalent results. Let us notice that the former case, where all the structure is in the ${\cal M}$ matrix, is equivalent to the latter case (all the structure in ${\bf Y_\nu}$) if we set $a=b=0$. To see this, note that a redefinition of the $\nu_R$ fields as $\nu_R\rightarrow V_b\nu_R$, would leave ${\cal M}\propto{\rm diag}(-1,1,1)$ and ${\bf Y_\nu}=Y_\nu V_{b}^T$. In consequence, it is enough to study the case where all the structure is in ${\bf Y_\nu}$, given by eq.(\ref{Ynu}). \subsection{Running ${\cal M}_\nu$ from $M_p$ to low energy} From $M_p$ to $M$ the evolution of the relevant matrices is governed by the following renormalization group equations~\cite{RGE}: \begin{Eqnarray} \label{rg2} \frac{d {\bf Y_\nu}}{dt}= -\frac{1}{16\pi^2} {\bf Y_\nu}\left[\left( \frac{9}{4}g_2^2+\frac{3}{4}g_1^2-{\mathrm T} \right){\bf I_3}-\frac{3}{2}\left( {\bf Y_\nu}^\dagger {\bf Y_\nu}-{\bf Y_e^\dagger Y_e}\right) \right], \end{Eqnarray} \begin{Eqnarray} \label{rg3} \frac{d {\bf Y_e}}{dt}= -\frac{1}{16\pi^2} {\bf Y_e} \left[\left( \frac{9}{4}g_2^2+\frac{15}{4}g_1^2-{\mathrm T} \right){\bf I_3}+\frac{3}{2}\left( {\bf Y_\nu}^\dagger {\bf Y_\nu}- {\bf Y_e^\dagger Y_e}\right) \right], \end{Eqnarray} where \begin{Eqnarray} {\mathrm T}={\mathrm Tr}(3{\bf Y_U^\dagger Y_U} +3 {\bf Y_D^\dagger Y_D}+{\bf Y_\nu^\dagger Y_\nu}+{\bf Y_e^\dagger Y_e}), \end{Eqnarray} and \begin{Eqnarray} \label{rg4} \frac{d {\cal M}}{dt}=\frac{1}{16\pi^2}\left[{\cal M} ({\bf Y_\nu} {\bf Y_\nu}^\dagger)^T+{\bf Y_\nu} {\bf Y_\nu}^\dagger {\cal M}\right] \end{Eqnarray} (not yet given in the literature). Here $g_2$ and $g_1$ are the $SU(2)_L$ and $U(1)_Y$ gauge coupling constants, and ${\bf Y_{U,D,e}}$ are the Yukawa matrices for up quarks, down quarks and charged leptons. At $M$, $\nu_R$ decouple, and ${\bf Y_e}$ must be diagonalized to redefine the flavour basis of leptons [note that the last term in (\ref{rg3}) produces non-diagonal contributions to ${\bf Y_e}$] affecting the form of the ${\bf Y_\nu}$ matrix. Then the effective mass matrix for the light neutrinos is ${\cal M}_\nu\simeq {\bf Y_\nu}^T {\cal M}^{-1} {\bf Y_\nu} \langle H\rangle^2$. From $M$ to $M_Z$, the effective mass matrix ${\cal M}_\nu$ is run down in energy exactly as described in section 2. The renormalization group equations are integrated with the following boundary conditions: ${\cal M}$ and ${\bf Y_\nu}$ are chosen at $M_p$ so as to satisfy \begin{Eqnarray} {\cal M}_\nu(M_p)={\cal M}_{b}, \end{Eqnarray} with the overall magnitude of ${\bf Y_\nu}$ fixed, for a given value of the Majorana mass $M$, by the requirement $m_{\nu}\sim O({\rm eV})$. The boundary conditions for the other Yukawa couplings are also fixed at the low energy side to give the observed fermion masses. The free parameters are therefore $M, a$ and $b$. \subsection{Limits on the parameter space} We discuss here the limits on the parameter space of our study, which is expanded by $M, a$ and $b$. The parameters $a$ and $b$, which define the texture of the ${\bf Y_\nu}$ matrix through eqs.(\ref{Ynu}, \ref{boosts}), can be in principle any real numbers. However, it is apparent from eqs.(\ref{Ynu}, \ref{boosts}) that if $a$ or $b$ are large, the entries of ${\bf Y_\nu}$ are extremely fine-tuned. Notice that the relative difference between $\cosh a$ and $\sinh a$ factors is $\sim 2e^{-2|a|}$ (and the analogue for $\cosh b$ and $\sinh b$). Therefore, the ${\bf Y_\nu}$ matrix is fine-tuned as $\sim 2e^{-2(|a|+|b|)}$ parts in one. In particular, if $|a|$ or $|b|$ are $>1.5$ the matrix elements are fine-tuned at least in a 10\%. In consequence, we will demand \begin{Eqnarray} |a|, |b| \leq 1.5 \, .\label{ablimits} \end{Eqnarray} Let us remark that the previous limits are based on a criterion of naturality for the ${\bf Y_\nu}$ matrix. However, let us mention that if we relax these limits, the final results (to be presented in the next subsection) are basically unchanged, since the allowed regions for $a,b$ remain in the ``natural'' region (\ref{ablimits}) in most of the cases. Concerning the remaining parameter, $M$, there is an upper bound on it coming from the fact that for large values of $M$, the neutrino Yukawa couplings, ${\bf Y_\nu}$, develop Landau poles below $M_p$, spoiling the perturbativity of the theory \cite{cdiq}. This occurs for $M(M_p)\sim 4.3\times10^{13}$ GeV. Actually, there is an additional effect, namely the closing of the allowed Higgs window, which occurs approximately for the same value of $M(M_p)$. Consequently, this sets the upper bound on $M$. It is interesting to note that this effect also restricts the values of $a, b$: for a given $M$, the larger $a, b$, the larger the entries of ${\bf Y_\nu}$, and thus the lower the scale at which the Landau pole appears. In general, this restriction on $a, b$ is less severe than eq.(\ref{ablimits}). Regarding the lower bound on $M$, there is no physical criterion for it, as the actual origin of the right-handed Majorana matrix is unknown. Since $M$ can be written in terms of $m_\nu$ and ${\bf Y_\nu}$ (roughly speaking $M \simeq Y_\nu^2 \langle H\rangle^2/m_\nu$), we can adopt the sensible criterion that $Y_\nu$ is at least as large as the smallest Yukawa coupling so far known, i.e. the electron one. This precisely corresponds to $M\sim 100$ GeV, below which is unplausible to descend. In consequence, our limits for $M$ are \begin{Eqnarray} 10^2 {\rm GeV}\ \stackrel{<}{{}_\sim} M\ \stackrel{<}{{}_\sim} 4.3\times 10^{13} {\rm GeV}. \label{Mlimits} \end{Eqnarray} On the other hand, as we will see in the next subsection, there are no physically viable scenarios unless \begin{Eqnarray} M\ \stackrel{>}{{}_\sim} 10^{8} {\rm GeV} , \label{Mlimit} \end{Eqnarray} which sets an operating lower limit on $M$. \begin{figure}[t] \centerline{\hbox{ \psfig{figure=cein2a.ps,height=10cm,width=10cm,bbllx=1.cm,% bblly=0.cm,bburx=12.cm,bbury=9.cm} \psfig{figure=cein2b.ps,height=10cm,width=10cm,bbllx=4.cm,% bblly=0.cm,bburx=15.cm,bbury=9.cm}}} \caption{\footnotesize Left plot: contours of $\Delta m_{12}^2/{\mathrm eV}^2$ in the $(b,a)$ plane from less than $10^{-5}$ (black area), through $2\times 10^{-5}$ and $10^{-4}$ (lines) to more than $2\times 10^{-4}$ (grey). Right plot: same for $\Delta m_{23}^2/{\mathrm eV}^2$, from $5\times 10^{-4}$ (black) to $10^{-2}$ (grey). The Majorana mass is $8\times 10^9$ GeV.} \end{figure} \subsection{Results} Figures 2 to 6 present our results for the mass splittings and mixing angles at low energy, after numerical integration of the RGEs from $M_p$ to low energy as described in subsection 3.2 (we follow the convention $m_{\nu_1}^2<m_{\nu_2}^2<m_{\nu_3}^2$ in all the figures). We have chosen $M=8\times 10^9$ GeV and $m_\nu\simeq 2$ eV as a typical example; the dependence of the results with $M$ and $m_\nu$ is discussed later on. \begin{figure}[t] \centerline{\hbox{ \psfig{figure=cein3a.ps,height=10cm,width=10cm,bbllx=1.cm,% bblly=0.cm,bburx=12.cm,bbury=9.cm} \psfig{figure=cein3b.ps,height=10cm,width=10cm,bbllx=4.cm,% bblly=0.cm,bburx=15.cm,bbury=9.cm}}} \caption{\footnotesize Left plot: Contours of $\sin^22\theta_{2}$ in the $(b,a)$ plane. The grey area marks the $\sin^22\theta_{2}>0.64$ region. The line singled-out corresponds to $\sin^22\theta_{2}=0.36$. Right plot: Contours of $\sin^22\theta_{1}$ in the $(b,a)$ plane. In the grey area $\sin^22\theta_{1}$ is smaller than 0.82, and the line corresponds to $\sin^22\theta_{1}=0.9$. The Majorana mass is $8\times 10^9$ GeV.} \end{figure} Figure 2, left plot, shows contour lines of constant $\Delta m_{12}^2$ (the squared mass difference between the lightest neutrinos) in the plane $(b,a)$. The black (grey) region is excluded because there $\Delta m_{12}^2<10^{-5}\, {\mathrm eV}^2$ ($\Delta m_{12}^2>2\times 10^{-4}\, {\mathrm eV}^2$), which is too small (large) to account for the oscillations of solar neutrinos (LAMSW solution). The white area is thus the allowed region. The lines in it correspond to $\Delta m_{12}^2=10^{-4}\, {\mathrm eV}^2$ and $2\times 10^{-5}\, {\mathrm eV}^2$. \begin{figure}[t] \centerline{\hbox{ \psfig{figure=cein4a.ps,height=10cm,width=10cm,bbllx=1.cm,% bblly=0.cm,bburx=12.cm,bbury=9.cm} \psfig{figure=cein4b.ps,height=10cm,width=10cm,bbllx=4.cm,% bblly=0.cm,bburx=15.cm,bbury=9.cm}}} \caption{\footnotesize Left plot: Same as figure 3 for $\sin^22\theta_{3}$. The grey area corresponds to values above 0.99. The curve gives $\sin^22\theta_{3}=0.95$. Right plot: The grey area corresponds to $\cos 2\theta_{3}<0$ } \end{figure} \begin{figure}[t] \centerline{ \psfig{figure=cein5.ps,height=10cm,width=10cm,bbllx=2.cm,% bblly=0.cm,bburx=13.cm,bbury=9.cm}} \caption{\footnotesize Region (two disconnected parts) in the $(b,a)$ parameter space for $M=8\times 10^9$ GeV where all mass splittings and mixing angles satisfy experimental constraints. (See text for qualifications). } \end{figure} Figure 2, right plot, gives contour lines of constant $\Delta m_{23}^2$. The black (grey) region is excluded because there $\Delta m_{23}^2<5\times 10^{-4}\, {\mathrm eV}^2$ ($\Delta m_{23}^2> 10^{-2}\, {\mathrm eV}^2$), which is too small (large) to account for the oscillations of atmospheric neutrinos. Again, the white area is allowed. (The black area corresponds in fact to the ``undecidable'' case discussed in the Introduction: it might be rescatable by unspecified extra effects.) We do not plot $\Delta m_{13}^2$ because it can always be inferred from $\Delta m_{23}^2$ and $\Delta m_{12}^2$. Moreover, in the interesting case, $\Delta m_{12}^2\ll \Delta m_{23}^2$, one has $\Delta m_{13}^2 \simeq\Delta m_{23}^2$. The intersection of the white areas in both plots is non-zero and would give the allowed area concerning mass splittings. It is always the case that the area surrounding the origin is excluded. There, the mass differences are always of the same order, and follow the same pattern discussed in subsection 2 ($\Delta m_{23}^2=2 \Delta m_{12}^2$). In any case we conclude that, away from the origin, there is a non negligible region of parameter space where it is natural to have $\Delta m_{23}^2\gg \Delta m_{12}^2$ and in accordance with the values required to explain the solar and atmospheric neutrino anomalies. In the following subsection we explain the origin of this hierarchy of mass differences. \begin{figure} \centerline{\hbox{ \psfig{figure=cein6b.ps,height=10cm,width=10cm,bbllx=2.5cm bblly=0.cm,bburx=13.5cm,bbury=9.cm}}} \centerline{\hbox{ \psfig{figure=cein6c.ps,height=10cm,width=10cm,bbllx=1.cm,% bblly=0.cm,bburx=12.cm,bbury=9.cm} \psfig{figure=cein6d.ps,height=10cm,width=10cm,bbllx=4.cm,% bblly=0.cm,bburx=15.cm,bbury=9.cm}}} \caption \noindent{\footnotesize Same as figure 5 for different values of the Majorana mass: Upper: $10^{11}$ GeV; Lower left: $10^{10}$ GeV; Lower right: $10^{9}$ GeV. } \end{figure} Next, we need to ensure that the mixing angles are the correct ones to give a good fit to atmospheric and solar neutrino data (as summarized by the ranges given in the Introduction). Figure 3, left plot, gives contours of constant $\sin^22\theta_{2}$ (one of the mixing angles relevant for atmospheric neutrino oscillations). The grey (white) area has $\sin^22\theta_{2}$ larger (smaller) than 0.64 and is disfavored (favored) by the data (SK $+$ CHOOZ) at 99\% C.L. according to the most recent analysis (last paper of ref.~\cite{range}). The line singled-out corresponds to $\sin^22\theta_{2}=0.36$ (maximum allowed value at 90\% C.L. according to the same reference). Figure 3, right plot, shows contours of constant $\sin^22\theta_{1}$ (the other mixing angle relevant for atmospheric neutrinos). The grey (white) area corresponds to $\sin^22\theta_{1}$ smaller (larger) than 0.82, and is thus disallowed (allowed). The additional line included has $\sin^22\theta_{1}=0.9$. Finally, figure 4, left plot, presents contours of constant $\sin^22\theta_{3}$ which is relevant for oscillations of solar neutrinos. The grey (white) region has $\sin^22\theta_{3}$ larger (smaller) than 0.99. If one is willing to interpret the existing data as impliying an upper bound of 0.99 on $\sin^22\theta_{3}$, then the grey region would be excluded. The plotted curve gives $\sin^22\theta_{3}=0.95$ ($\sin^22\theta_{3}>0.8$ in all the region shown). Figure 4, right plot, shows the region of the parameter space accomplishing the resonance condition ($\cos2\theta_{3}>0$), which is required for an efficient MSW solution of the solar anomaly (see however the first paper of ref.~\cite{range} for caveats on this issue). The region of parameter space where all constraints on mixing angles and mass splittings are satisfied is given by the intersection of all white areas in figures 2, 3 and 4 (right plot). If $\sin^22\theta_{3}<0.99$ is imposed, then that intersection region, including now figure 4 (left plot), is empty and no allowed region remains. It should be noticed that this fact does not come from an incompatibility between the previous constraint and the $\sin^22\theta_{3}>0.99$ obtained from neutrinoless double $\beta$-decay limits, eq. (\ref{doublebeta}), in the $\theta_2=0$ approximation. If this were the case, it could be easily solved by decreasing the overall size of the neutrino masses, $m_\nu$, in eq.(\ref{doublebeta}), and this is not the case. Indeed, eq.(\ref{doublebeta}) is satisfied in nearly all the parameter space. Even where $\sin^22\theta_{3}<0.99$, this is still true thanks to the contribution of $\theta_2$. What actually forbids the whole parameter space if $\sin^22\theta_{3}<0.99$ is imposed, is the incompatibility between acceptable $\theta_1$, $\theta_2$ and $\theta_3$ angles to fit simultaneously all the neutrino oscillation data, as can be seen from the figures. This fact remains when $m_\nu$ is decreased. In fact, the effect of decreasing $m_\nu$ is essentially an amplification of the figures shown here, which comes from the fact that for a given Majorana mass, the neutrino Yukawa couplings become smaller (the effect is similar to decrease $M$, which is discussed below). If the $\sin^22\theta_{3}<0.99$ condition is relaxed (as discussed in the Introduction), then the allowed region is given by the two islands in figure 5, which is non-negligible. If we now vary $M$, this allowed region will move in parameter space as indicated in figure 6, where we show the allowed regions for a sequence of Majorana masses that range from $10^9$ to $10^{11}$ GeV. As is apparent from the figure, lowering $M$ has the effect of enlarging the allowed region which flies away from the origin, leaving at some point the region of naturalness for $a$ and $b$. At $M=10^8$ GeV there is no allowed region inside the natural range for $(a,b)$. Conversely, increasing $M$ reduces the allowed region, which gets closer to the origin (at $M=10^{12}$ GeV the allowed region becomes extinct). Let us remark again that in the allowed region one fits the observed atmospheric and solar anomalies, while part of the disallowed region corresponds in fact to the undecidable case (more precisely the region near the $a=b=0$ origin), in which some additional physics could be invocated to explain the same data. In contrast, the region away from the origin becomes excluded (the mass splittings are too large, even for atmospheric neutrinos). Let us also notice that if the $\sin^22\theta_{3}<0.99$ condition is imposed, the whole parameter space becomes disallowed for any value of $M$. We also find that, whenever there is a hierarchy in the mass splittings, the two lightest eigenvalues have opposite signs. This is just what is needed to have a cancellation occurring in the neutrinoless double $\beta$-decay constraint (\ref{doublebeta}). This constraint is satisfied in almost the whole parameter space for any $M$. This and other features of our results are discussed further in the next subsection. \subsection{Analytical understanding of results} It is simple and very illuminating to derive analytical approximations for the results presented in the previous subsection. The renormalization group equations (\ref{rg1},\ref{rg2},\ref{rg3},\ref{rg4}) we integrated numerically can also be integrated analytically in the approximation of constant right hand side. In this approximation (which works very well for our analysis), the effective neutrino mass matrix at low-energy is simply ${\cal M}_{b}$ plus some small perturbation. It is straightforward to obtain how the degenerate eigenvalues of ${\cal M}_{b}$ get split by this perturbation. Neglecting the $Y_e$, $Y_\mu$ Yukawa couplings, we get the following analytical expressions: \be\begin{array}{cl} m_{\nu_1}\simeq &m_\nu \left[-1+ (2 c_a^2 c^2_b-1)\epsilon_\nu-2\epsilon_\tau \right],\vspace{0.2cm}\\ m_{\nu_{2,3}}\simeq &m_\nu \left[1+3 \epsilon_\tau - c^2_a c^2_b \epsilon_\nu\pm \left\{\left[\epsilon_\tau+(c^2_a c^2_b-c_{2a})\epsilon_\nu \right]^2 +\left[s_{2a} s_b \epsilon_\nu - 2\sqrt{2}\epsilon_\tau \right]^2 \right\}^{1/2} \right] \label{mnu123} \end{array}\ee where $c_a=\cosh a$, $s_{2a}=\sinh 2a$, etc, and \begin{Eqnarray} \epsilon_\tau =\frac{Y_\tau^2}{128\pi^2}\left[\ \log\frac{M}{M_Z}+3\log\frac{M_p}{M}\right], \end{Eqnarray} \begin{Eqnarray} \epsilon_\nu=\frac{Y_\nu^2}{16\pi^2}\log\frac{M_p}{M}. \end{Eqnarray} (The labelling of mass eigenvalues in eq.(\ref{mnu123}) may not always correspond to the conventional order $m_{\nu_1}^2<m_{\nu_2}^2<m_{\nu_3}^2$.) It can be checked that, for $a=b=0$, the ${\bf Y}_{\nu}$ couplings play no role in the mass splittings. This is not surprising since, as was mentioned in subsection 3.1, this case is equivalent to having all the structure in ${\cal M}$, while ${\bf Y}_{\nu}$ is proportional to the identity. Then, it can be seen from the RGEs that all the non-universal modifications on ${\cal M}_{\nu}$ come from the ${\bf Y}_{e}$ matrix, and has a form similar to the one found in section 2 [see eq.(\ref{Mpert})]. So this scenario gives similar (not satisfactory) results to those found in that section. As $\epsilon_\nu\gg \epsilon_\tau$ (which occurs as soon as ${Y}_{\nu}>Y_\tau$, i.e for $M\stackrel{>}{{}_\sim} 10^9$ GeV), a further expansion in powers of $\epsilon_\tau/\epsilon_\nu$ of the square root is possible in most of the parameter space (except where the coefficient of $\epsilon_\nu^2$ inside that square-root becomes very small). The mass eigenvalues then read \be\begin{array}{cl} m_{\nu_1}\simeq &m_\nu \left[-1+ (2 c_a^2 c^2_b-1)\epsilon_\nu-2\epsilon_\tau \right],\vspace{0.2cm}\\ m_{\nu_2}\simeq &m_\nu \left[1-(2 c_a^2 c^2_b-1)\epsilon_\nu+\left(2-{\displaystyle\frac{1-c_{2a}-2\sqrt{2} s_{2a}s_b}{c_a^2 c^2_b-1}} \right)\epsilon_\tau\right] ,\vspace{0.2cm}\\ m_{\nu_3}\simeq &m_\nu \left[1-\epsilon_\nu+\left(4+{\displaystyle\frac{1-c_{2a}-2\sqrt{2} s_{2a}s_b}{c_a^2 c^2_b-1}} \right)\epsilon_\tau\right]. \label{mnu1232} \end{array}\ee These expressions show clearly the origin of the neutrino mass splittings. The splitting between the first two neutrinos is controlled by the small parameter $\epsilon_\tau$, proportional to the squared Yukawa couplings of the charged leptons, and is insensitive to the parameter $\epsilon_\nu$ (proportional to the square of the larger Yukawa coupling $Y_\nu$) which is responsible for the mass difference of the third neutrino. From the previous expressions we can extract the following remarkable conclusions. If the neutrino Yukawa couplings are sizeable (i.e. bigger than $Y_\tau$), we will automatically obtain a hierarchy of mass splittings $\Delta m_{12}^2\ll \Delta m_{23}^2\sim \Delta m_{13}^2$. This is exactly what is needed for a simultaneous solution of the atmospheric and solar neutrino anomalies, and thus represents a natural mechanism for the $\Delta m_{sol}^2\ll \Delta m_{at}^2$ hierarchy. Thus, if the mixing angle $\theta_2$ is small, as it turns out to be in most of the parameter space, $\Delta m_{12}^2$ is to be correctly identified with $\Delta m_{sol}^2$ and $\Delta m_{23}^2$ with $\Delta m_{at}^2$. The two mass eigenvalues which are more degenerate correspond to the lighter states, i.e. $m_{\nu_1}^2\sim m_{\nu_2}^2<m_{\nu_3}^2$. Moreover, $m_{\nu_1}$ and $m_{\nu_2}$ have opposite signs in the diagonalized mass matrix, which is exactly what is needed to fulfill the neutrinoless double $\beta$-decay condition, eq.(\ref{doublebeta}). All these nice features are illustrated by the explicit results presented in the previous subsection. Concerning the mixing angles, it is straightforward to check that, working in the ${Y}_{\nu}>Y_\tau$ approximation, the eigenstates of the perturbed ${\cal M}_\nu$ matrix, corresponding to the previous eigenvalues, are \begin{Eqnarray} V_1' = V_1,\;\;\;V_2' = \frac{1}{\sqrt{\alpha^2+\beta^2}} (\alpha V_2+\beta V_3),\;\;\; V_3' = \frac{1}{\sqrt{\alpha^2+\beta^2}}(-\beta V_2+\alpha V_3), \end{Eqnarray} where $V_i$ are the eigenstates corresponding to the bimaximal mixing matrix, $V_b$ [see eq.(\ref{VGG})] \begin{Eqnarray} V_1 = \left ( \begin{array}{c} {\displaystyle \frac{-1}{\sqrt{2}}} \vspace{.1cm}\\ {\displaystyle\frac{1}{2}} \vspace{.1cm}\\ {\displaystyle\frac{1}{2}} \end{array} \right ), ~~ V_2 = \left ( \begin{array}{c} {\displaystyle{1 \over \sqrt{2}}} \vspace{.1cm}\\ {\displaystyle{1 \over 2}} \vspace{.1cm}\\ {\displaystyle{1 \over 2}} \end{array} \right ), ~~ V_3 = \left ( \begin{array}{c} 0 \vspace{.1cm}\\ {\displaystyle\frac{-1}{\sqrt{2}}} \vspace{.1cm}\\ {\displaystyle\frac{1}{\sqrt{2}}} \end{array} \right ), \label{vecbim} \end{Eqnarray} and $\alpha, \beta$ are given by \begin{Eqnarray} \alpha=c_{a}s_b + O({\epsilon_\tau}/{\epsilon_\nu}),\;\;\;\;\beta=s_{a} + O({\epsilon_\tau}/{\epsilon_\nu}). \label{alfabeta} \end{Eqnarray} The $V_i'$ vectors define the new ``CKM'' matrix $V'$ from which the mixing angles are extracted. Clearly, if just one of the two $(a,b)$ parameters is vanishing, then $V'=V_b$, i.e. exactly the bimaximal mixing case. Also, whenever $c_a$, $c_b$ are sizeable (i.e. away from $a=b=0$), $|\alpha|\gg |\beta|$, and thus we are close to the bimaximal case. Therefore, it is not surprising that in most of the parameter space shown in the previous section, this was in fact the case. This is remarkable, because it gives a natural origin for the bimaximal mixing, which was not guaranteed a priori due to the ambiguity in the diagonalization of the initial ${\cal M}_{\nu}(M_p)={\cal M}_{b}$ matrix, as was explained in the Introduction. On the other hand, the MSW condition $\cos2\theta_{3}>0$ (written using the conventional order $m_{\nu_1}^2<m_{\nu_2}^2$) will be clearly satisfied as long as $V_1$ corresponds to the lightest mass eigenvalue. In other words, this condition requires that the negative mass eigenvalue, see eqs.(\ref{mnu123}, \ref{mnu1232}), corresponds to the lightest neutrino. \subsection{Examples of acceptable ans\"atze} At a generic point in the allowed regions we have found, the form of the matrix ${\bf Y_\nu}(M_p)$ would look rather ad-hoc: different elements in that matrix seem to conspire to give the correct neutrino mass texture. However, there are particular cases in which this matrix has a plausible structure. We give an example of such a texture for ${\bf Y_\nu}(M_p)$ which for the case $M\simeq 8\times 10^{9}$ GeV studied in previous sections, would give mass splittings and mixing angles in agreement with observation: \begin{Eqnarray} {\bf Y_\nu}(M_p) = Y_\nu\,\pmatrix{ - {\displaystyle{1\over 2\sqrt2}} & 1 & 1 \cr {\displaystyle{1\over 2\sqrt2}} & 1 & 1 \cr 0 & - {\displaystyle{1\over\sqrt2}} & {\displaystyle{1\over\sqrt2}} \cr }. \label{ansatz} \end{Eqnarray} It corresponds to $a=0$ and $b=\sinh^{-1}(3/4)\simeq 0.69$, value which falls in the allowed region plotted in figure 5. More precisely, the mass splittings are \begin{Eqnarray} \Delta m_{12}^2\simeq 2\times 10^{-5}\, {\mathrm eV}^2,\;\;\; \Delta m_{13}^2\simeq 1\times 10^{-3}\, {\mathrm eV}^2,\;\;\; \Delta m_{23}^2\simeq 1\times 10^{-3}\, {\mathrm eV}^2, \end{Eqnarray} and the mixing angles \begin{Eqnarray} \sin^22\theta_{2}=0.04560,\;\;\; \sin^22\theta_{1}=0.95405,\;\;\; \sin^22\theta_{3}=0.99986. \end{Eqnarray} Concerning the resonance condition for the MSW mechanism (see Fig.4), this ansatz lies precisely at the border of the allowed area. Another examples of working ans\"atze can be obtained. For instance, the following ansatz (corresponding to $a=\cosh^{-1}(\sqrt{5}/2) \simeq 0.48$, $b=\log(\sqrt{10}/2)\simeq 1.15$ ) \begin{Eqnarray} {\bf Y_\nu}(M_p) = Y_\nu\,\pmatrix{ - {\displaystyle {1\over 4}} &{\displaystyle{ \sqrt2}} & {\displaystyle{3\over \sqrt2}} \cr {\displaystyle{1\over 2\sqrt5}} & {\displaystyle{\sqrt5\over 2}} & {\displaystyle{\sqrt5\over 2}} \cr - {\displaystyle{1\over 4\sqrt5}} & 0 & {\displaystyle{\sqrt5\over 2}} \cr }. \label{ansatz2} \end{Eqnarray} works correctly for a wider range of right-handed Majorana masses (e.g. for $M=10^{9}$ GeV and $M=10^{10}$ GeV, as can be seen from Figure 6). It would be interesting to explore the possibility of finding a symmetry that could be responsible for the form of these ans\"atze and to analyze their implications for future long baseline experiments \cite{derujula}. \section{Conclusions} We have performed an exhaustive study of the possibility that radiative corrections are responsible for the small mass splittings in the (cosmologically relevant) scenario of nearly degenerate neutrinos. To do that, we assume that the initial form of the neutrino mass matrix (generated at high energy by unspecified interactions) has the bimaximal mixing form, and run down to low energy. We then examine the form of the low-energy neutrino mass matrix, checking its consistency with all the available experimental data, including atmospheric and solar neutrino anomalies. We find cases where the radiative corrections produce mass splittings that are $i)$ just fine $ii)$ too large or $iii)$ too small. The vacuum oscillations solution to the solar neutrino problem always falls in the situation $(ii)$, and it is therefore excluded. On the contrary, if the initial bimaximal mass matrix is produced by a see-saw mechanism (a possibility that we analyze in a detailed way), there are large regions of the parameter space consistent with the large angle MSW solution, providing a natural origin for the $\Delta m^2_{sol} \ll \Delta m^2_{atm}$ hierarchy. Concerning the mixing angles, they are remarkably stable and close to the bimaximal mixing form (something that is not guaranteed a priory, due to an ambiguity in the diagonalization of the initial matrix). We have explained analytically the origin of these remarkable features, giving explicit expressions for the mass splittings and the mixing angles. In addition, we have presented particularly simple see-saw ans\"atze consistent with all the observations. Finally, we have noted that the scenario is very sensitive to a possible upper bound on $\sin^2 2\theta_3$ (the angle responsible for the solar neutrino oscillations). An upper bound such as $\sin^2 2\theta_3<0.99$ would disallow completely the scenario of nearly degenerate neutrinos due to the incompatibility between acceptable mixing angles to fit simultaneously all the neutrino oscillation data. \section*{Addendum} Shortly after completion of this work, a paper by J. Ellis and S. Lola on the same subject appeared \cite{EL}. In it, the scenario of our section~2 is also studied and similar (negative) conclusions reached. However, as we show in our section 3, positive results are obtained when the general see-saw scenario as the mechanism responsible for the effective neutrino mass matrix is studied. Also, the treatment by these authors of the constraints on mixing angles from neutrinoless double $\beta$-decay and LAMSW fits to solar neutrino data is more pessimistic than ours. \section*{Acknowledgements} We thank D. Casper and E. Kearns for clarification of the results of Super-Kamiokande fits. This research was supported in part by the CICYT (contract AEN95-0195) and the European Union (contract CHRX-CT92-0004) (JAC). J.R.E. thanks the I.E.M. (CSIC, Spain) and A.I, I.N the CERN Theory Division, for hospitality during the final stages of this work.
\section{Introduction} By making use of the idea that a quantum system of gravity may possess far less degrees of freedom than usually expected for a $3+1$ dimensional field theory\cite{01}\cite{02}, it was suggested recently that there is an AdS/CFT correspondence\cite{03}-- \cite{05}: the string theory (M theory) on background of the form AdS$_d\times M_{D-d}$ is dual to a conformal field theory living on the spacetime boundary. Here AdS$_d$ is an Anti-de Sitter (AdS) space of spacetime dimension $d$, and $M_{D-d}$ is a certain compactification space of dimension $D-d$ with $D=10$ for string theory ($D=11$ for M theory). A strong support for the proposal comes from comparing spectra of Type IIB string theory on the background of AdS$_5\times S^5$ and low-order correlation functions of the $3+1$ dimensional ${\cal N}=4$ $SU(N)$ super Yang-Mills theory. The dual super Yang-Mills theory lives on the boundary of the AdS space. To each field $\Phi_i$ there is a corresponding local operator ${\cal O}_i$ in the conformal field theory. The precise relation between string theory in the bulk and field theory on the boundary is \begin{equation} Z_{\rm eff}(\Phi_i)=e^{iS_{\rm eff}(\Phi_i)}=\langle Te^{i\int_{\cal B} \Phi_{b,i}{\cal O}^i}\rangle~, \end{equation} where $S_{\rm eff}$ is the effective action in the bulk and $\Phi_{b,i}$ is the field $\Phi_i$ restricted to the boundary. In the large $N$ limit, the string theory is weakly coupled and supergravity is a good approximation to it. Thus it is possible to describe a precise recipe expressing correlation functions of the ${\cal N}=4$ super Yang-Mills theory in four dimensions in terms of a calculations of tree approximation to supergravity performed in the bulk. At first sight, it seems very strange for us that quantum theories in different spacetime dimensions ever could be equivalent in any sense. The key to understand it is the fact that the theory in the larger dimension is always a quantum theory of gravity. For such theories the concepts of holography has been introduced by 't Hooft\cite{01} based on phenomenological study of the black hole theory and was referred as a generic properties of quantum system of gravity. The AdS/CFT correspondence is just an example of the realization of the holography on quantum theory of gravity. From point of view of general relativity, gravity is nothing but spacetime geometry. To check the Maldacena conjecture, one has to start from investigating of adequate geometry description for quantum theory of gravity. Because the holography is a generic properties of the quantum gravity, in principle, it should be deduced naturally from geometric properties of spacetime. 't Hooft showed that simple regularization of spacetime can not give correct account of observable degrees of freedom for quantum theory of gravity. On the other hand, the poor understanding of physics at Planck scale also indicates that the small scale structure of spacetime might not be adequately described by classical continuum geometry. Thus, new geometry should be introduced for quantum gravity. It has long time been suspected that the noncommutitive spacetime might be a realistic picture of how spacetime behaves near the Planck scale\cite{06}\cite{07}. Strong quantum fluctuations of gravity may make points in space fuzzy. The noncommutative geometry description\cite{08}\cite{09} is a strong candidate for quantum theory of gravity. We wish the holography can be obtained explicitly based on noncommutative geometry picture of quantum gravity and show the AdS/CFT correspondence by demonstrating conformally invariant symmetry on the boundary surface of noncommutative AdS space. In this paper, we present a kind of special regularization with exponentially increasing spacetime cutoff for both orthogonal and AdS space based on noncommutative geometry. It is used to argue that the same minimal cutoff of any geometry is the Planck scale $l_p$, that the most direct and obvious physical cutoff is from the formation of microscopic black holes as soon as too much energy would be accumulated into too small a region\cite{01}. We show that an adequately adopted noncommutative deformation of geometry makes the holography of higher dimensional quantum system of gravity and lower dimensional theory possible. As an example, detail calculations are carried out for the counting of observable degrees of freedom of quantum gravity in the bulk of noncommutative space $SO_q(3)$ and the classical limit of its boundary surface $S^3$. Results show that a very small (may be $<10^{-15}$) displacement of the noncommutative deformation parameter from its classical value $1$ reduces sharply the entropy of the quantum system of gravity. The desired entropy expression $S=4\pi M^2 +C$ of the universe can be deduced naturally. Conformally invariant symmetry is obtained for the equivalent theory of the quantum gravity, which lives on the boundary of the noncommutative AdS space. This is the bases of the AdS /CFT correspondence in nonperturbative string theory and M theory. This paper is organized as follows. In Section 2, we discuss the noncommutative orthogonal space $SO_q(3)$. The algebra satisfied by the coordinates and derivatives is decoupled into three independent subalgebras by introducing of a new set of variables. Quantum coherent states are constructed as reference ones for investigating representations of the spacetime algebra. Forms of Hilbert space show clearly that the noncommutative spacetime is discretely latticed with exponentially increasing space distances. The noncommutative deformation parameter is determined by an algebraic equation. The noncommutative space $SO_q(3)$ has a same entropy or observable degrees of freedom with the classical $S^3$ surface. Section 3 is devoted to the study of noncommutative AdS$_{2n}$ space. Conjugate operation is set up for the noncommutative AdS space. This conjugation has a induced counterpart for the set of decoupled coordinates and derivatives. Hilbert space of the noncommutative AdS$_{2n}$ space is constructed based on quantum coherent states. Discrete lattice structure of the noncommutative AdS$_{2n}$ with exponentially increasing space distances is obtained. Holography makes the quantum system of gravity on the noncommutative AdS$_{2n}$ space equivalent to the conformally invariant quantum theory living on the classical limit of its boundary surface. This is crucial for the AdS/CFT correspondence of string theory and M theory. By almost same procedures, properties of the noncommutative AdS$_{2n-1}$ space are shown in Section 4. In Section 5, some concluding remarks are given. \section{Noncommutative geometry and holography} We begin by discussing the quantum space $SO_q(3)$ with coordinates $x^i$ ($i=-,~0,~+$). The commutation relations\cite{10} among coordinates $x^i$ are \begin{equation} \begin{array}{l} x^-x^0=qx^0x^-~,~~~~~x^0x^+=qx^+x^0~,\\ x^{+}x^{-}-x^{-}x^{+}=(q^{1/2}-q^{-1/2})x^0x^0~. \end{array} \end{equation} The algebra satisfied by derivatives is of the form \begin{equation}\label{diff} \begin{array}{l} \partial_-\partial_0=q^{-1}\partial_0\partial_-~,~~~~~ \partial_0\partial_+=q^{-1}\partial_+\partial_0~,\\ \partial_-\partial_+-\partial_+\partial_-=(q^{1/2}-q^{-1/2})\partial_0 \partial_0~. \end{array} \end{equation} The action of derivatives on the coordinates is \begin{equation} \begin{array}{l} \partial_-x^-=1+q^2x^-\partial_-+\lambda q x^0\partial_0 +\lambda (q-1) x^+\partial_+~,\\ \partial_-x^0=qx^0\partial_--q^{1/2}\lambda x^+\partial_0~,~~~~~ \partial_-x^+=x^+\partial_-~,\\ \partial_0x^-=qx^-\partial_0-q^{1/2}\lambda x^0\partial_+~,\\ \partial_0x^0=1+qx^0\partial_0+q\lambda x^+\partial_+~,~~~~~ \partial_0x^+=qx^+\partial_0~,\\ \partial_+x^-=x^-\partial_+~,~~~~~\partial_+x^0=qx^0\partial_+~,~~~~~ \partial_+x^+=1+q^2x^+\partial_+~. \end{array} \end{equation} It is convenient to introduce the dilatation operators \begin{equation} \begin{array}{l} \mu_+=1+q\lambda x^+\partial_+ ~,\\ \Lambda=1+q\lambda\displaystyle\sum_{j=0,\pm}x^j\partial_j+q^3\lambda^2 \left(q^{-1/2}x^-x^++\frac{q}{1+q}x^0x^0\right)\left(q^{-1/2}\partial_+ \partial_-+\frac{q}{1+q}\partial_0\partial_0\right)~. \end{array} \end{equation} The dilatation operators obey \begin{equation} \begin{array}{l} \mu_+ x^+=q^2x^+\mu_+~,~~~~~\mu_+\partial_+=q^{-2}\partial_+ \mu_+~,\\ \Lambda x^k=q^2 x^k \Lambda~,~~~~~\Lambda\partial_k=q^{-2}\partial_k\Lambda~, ~~~{\rm for}~k=0,\pm~. \end{array} \end{equation} The real form $SO_q(3,R)$ (noted simply as $SO_q(3)$ whenever no confusion rised) of the noncommutative space $SO_q(3)$ is obtained by a consistent conjugation \begin{equation}\label{conjugation} \begin{array}{l} \overline{x^i}=C_{ji}x^j~,\\ \overline{\partial_i}=-q^{-2}C_{ij}\Lambda^{-1}\displaystyle\left(q^{-1/2} [\partial_+\partial_-,x^j]+\frac{q}{1+q}[\partial_0\partial_0,x^j]\right)~, \end{array} \end{equation} where the metric $C_{ij}$ is of the form $$C=\left(\begin{array}{ccc} & &q^{-1/2}\\ &1& \\ q^{1/2}& &\end{array}\right)~.$$ It is should be noticed that throughout this paper we limit us at the case of $q$ being real. By making use of the dilatation operators $\mu_+$ and $\Lambda$, we introduce a new set of coordinates and derivatives \begin{equation} \begin{array}{l} {\cal X}^-=\Lambda^{-1/2}\mu_+^{-1/2}\displaystyle\left(x^-+q^{3/2}\lambda \left(q^{-1/2}x^-x^++\frac{q}{1+q}x^0x^0\right)\partial_+\right)~,\\ {\cal D}_-=q^{-1}\Lambda^{-1/2}\mu_+^{-1/2}\displaystyle\left(\partial_-+q^{3/2}\lambda \left(q^{-1/2}\partial_+\partial_-+\frac{q}{1+q}\partial_0\partial_0\right) x^+\right)~,\\ {\cal X}^0=\mu_+^{-1/2}x^0~,~~~~~{\cal D}_0=\mu_+^{-1/2}\partial_0~,\\ {\cal X}^+=x^+~,~~~~~{\cal D}_+=\partial_+~. \end{array} \end{equation} In terms of the new variables, the commutation relations among coordinates and derivatives of the noncommutative space $SO_q(3)$ are transformed as \begin{equation} \begin{array}{l} {\cal D}_-{\cal X}^--{\cal X}^-{\cal D}_-=\mu_-^{-1}~,~~~~~~ \mu_- {\cal X}^-=q^2{\cal X}^-\mu_-~,\\ \mu_-{\cal D}_-=q^{-2}{\cal D}_-\mu_-~,~~~~~~\mu_-^{-1}\equiv 1+(q^{-2}-1){\cal X}^-{\cal D}_-~;\\ {\cal D}_0{\cal X}^0-{\cal X}^0{\cal D}_0=\mu_0^{1/2}~,~~~~~~ \mu_0 {\cal X}^0=q^2{\cal X}^0\mu_0~,\\ \mu_0{\cal D}_0=q^{-2}{\cal D}_0\mu_0~,~~~~~~\mu_0^{1/2}\equiv 1+(q-1){\cal X}^0{\cal D}_0~;\\ {\cal D}_+{\cal X}^+-{\cal X}^+{\cal D}_+=\mu_+~,~~~~~~ \mu_+ {\cal X}^+=q^2{\cal X}^+\mu_+~,\\ \mu_+{\cal D}_+=q^{-2}{\cal D}_+\mu_+~,~~~~~~\mu_+^{1/2}= 1+(q^2-1){\cal X}^+{\cal D}_+~;\\[1mm] [{\cal X}^i,{\cal X}^j]=0~,~~~~[{\cal D}_i,{\cal D}_j]=0~,~~~~ [{\cal D}_i,{\cal X}_j]=0~,\\[1mm] [\mu_i,{\cal X}^j]=0~,~~~~[\mu_i,{\cal D}_j]=0~,~~~~~~{\rm for}~ i\not= j~. \end{array} \end{equation} The noncommutative surface $S_q^3$ in terms of the set of independent operators ${\cal X}^j$ and ${\cal D}_j$ is of the form \begin{equation} \frac{q^{-1}}{1+q}{\cal X}^0{\cal X}^0+q^{-1/2}\Lambda^{1/2}\mu_+^{-1/2} {\cal X}^+{\cal X}^-=R^2~. \end{equation} At the $q\rightarrow 1$ limit, $S_q^3$ reduces to the familiar $S^3$ surface with radius $R$. The conjugate operation on ${\cal X}^j$ and ${\cal D}_j$ is induced by what on $x^j$ and $\partial_j$ (Eq.(\ref{conjugation})), \begin{equation} \begin{array}{l} {\overline{\cal X}^-}=\displaystyle\left({\overline x^-}+q^{3/2}\lambda {\overline\partial_+} \left(q^{-1/2}{\overline x^+}{\overline x^-}+\frac{q}{1+q}{\overline x^0} {\overline x^0}\right)\right){\overline \mu_+}^{-1/2} {\overline\Lambda}^{-1/2}~,\\ {\overline{\cal D}_-}=q^{-1}\displaystyle\left({\overline\partial_-}+ q^{3/2}\lambda {\overline x^+} \left(q^{-1/2}{\overline\partial_-}{\overline\partial_+}-\frac{q}{1+q} {\overline\partial_0}{\overline \partial_0}\right) \right){\overline \mu_+}^{-1/2}{\overline \Lambda}^{-1/2}~,\\ {\overline {\cal X}^0}={\overline x^0}{\overline \mu_+}^{-1/2}~,~~~~~ {\overline {\cal D}_0}={\overline \partial_0}{\overline \mu_+}^{-1/2}~,\\ {\overline {\cal X}^+}={\overline x^+}~,~~~~~{\overline {\cal D}_+}= {\overline \partial_+}~. \end{array} \end{equation} Thus, we conclude that the Quantum Heisenberg-Weyl algebra corresponds to the noncommutative space $SO_q(3)$ can be decoupled into three independent subalgebras. And then, one can investigate properties of the noncommutative space by constructing Hilbert spaces of the three quantum subalgebras. For the quantum algebra ${\cal A}_-$ \begin{equation} \begin{array}{l} {\cal D}_-{\cal X}^--{\cal X}^-{\cal D}_-=\mu_-^{-1}~,~~~~~~ \mu_- {\cal X}^-=q^2{\cal X}^-\mu_-~,\\ \mu_-{\cal D}_-=q^{-2}{\cal D}_-\mu_-~,~~~~~~\mu_-^{-1}\equiv 1+(q^{-2}-1){\cal X}^-{\cal D}_-~, \end{array} \end{equation} we construct a quantum coherent state $\vert z\rangle_-$ as \begin{equation} \vert z\rangle_-=\exp_{q^2}(-\frac{1}{2}|q^{-2}z|)\sum_{m=0}^\infty \frac{(-q^{-2}z)^m}{[m]_{q^2}!}({\cal D}_-)^m\vert 0\rangle_-~, \end{equation} where the notation of $q$-exponential $$\exp_q(x)\equiv\sum_{n=0}^\infty \frac{x^n}{[n]_q!}~,~~~~ [n]_q!=[n]_q[n-1]_q\cdots[1]_q~,~~~~[n]_q=\frac{q^n-1} {q-1}$$ has been used and the reference state $\vert 0\rangle_-$ was chosen such that ${\cal X}^-\vert 0\rangle_- =0$. As in the classical case, the coordinate ${\cal X}^-$ is diagonal in the quantum coherent state basis \begin{equation} {\cal X}^-\vert z\rangle_- =z\vert z\rangle_-~. \end{equation} The real value of parameter $z$ may be interpreted as position of an indispersive wave-pocket\cite{11}. Here we should notice that $z$ can be any complex number because of us working on a general quantum orthogonal space. The complex values of the quantum coherent state parameter are consistent with conjugate operation on the noncommutative space. Denote the quantum coherent state as $$\vert 0,z\rangle_-\equiv \vert z\rangle_-~,$$ we can construct a representation for the quantum algebra ${\cal A}_-$ based on the quantum coherent state as \begin{equation} \begin{array}{l} {\cal X}^-\vert n,z\rangle_-=q^{2n}z\vert n,z\rangle_-~,\\ {\cal D}_-\vert n,z\rangle_-=-q^{-1-2n}\lambda^{-1}z^{-1}\vert n+1,z\rangle_-~,\\ \mu_-\vert n,z\rangle_-=\vert n-1,z\rangle_-~. \end{array} \end{equation} It is clear from the Hilbert space representation of the quantum algebra ${\cal A}_-$ that the coordinates of the noncommutative orthogonal space is discretely latticed with exponentially increasing space distances. In fact, this is in agreement with the discrete difference representation of quantum derivatives, $${\cal D}f({\cal X})=\frac{f(q^2{\cal X})-f({\cal X})}{(q^2-1){\cal X}}~.$$ Similarly, we can construct the reference states $\vert 0,z\rangle_0$ and $\vert 0,z\rangle_+$ as \begin{equation} \begin{array}{l} \vert 0, z\rangle_0=\displaystyle\exp_{q^{-1}}(-\frac{1}{2}|qz|)\sum_{m=0}^\infty \frac{(-qz)^m}{[m]_{q^{-1}}!}({\cal D}_0)^m\vert 0 \rangle_0~,~~~~~{\cal X}^0\vert 0\rangle_0=0~,\\ \vert 0, z\rangle_+=\displaystyle\exp_{q^{-2}}(-\frac{1}{2}|q^2z|)\sum_{m=0}^\infty \frac{(-q^2z)^m}{[m]_{q^{-2}}!}({\cal D}_+)^m\vert 0 \rangle_+~,~~~~~{\cal X}^+\vert 0\rangle_+=0~. \end{array} \end{equation} The corresponding representations of the quantum algebras ${\cal A}_0$ \begin{equation} \begin{array}{l} {\cal D}_0{\cal X}^0-{\cal X}^0{\cal D}_0=\mu_0^{1/2}~,~~~~~~ \mu_0 {\cal X}^0=q^2{\cal X}^0\mu_0~,\\ \mu_0{\cal D}_0=q^{-2}{\cal D}_0\mu_0~,~~~~~~\mu_0^{1/2}\equiv 1+(q-1){\cal X}^0{\cal D}_0~, \end{array} \end{equation} and ${\cal A}_+$ \begin{equation} \begin{array}{l} {\cal D}_+{\cal X}^+-{\cal X}^+{\cal D}_+=\mu_+~,~~~~~~ \mu_+ {\cal X}^+=q^2{\cal X}^+\mu_+~,\\ \mu_+{\cal D}_+=q^{-2}{\cal D}_+\mu_+~,~~~~~~\mu_+\equiv 1+(q^2-1){\cal X}^+{\cal D}_+~, \end{array} \end{equation} are of the forms \begin{equation} \begin{array}{l} {\cal X}^0\vert n,z\rangle_0=q^{n}z\vert n,z\rangle_0~,\\ {\cal D}_0\vert n,z\rangle_0=\displaystyle\frac{q^{1-n}}{q-1}z^{-1}\vert n-1,z\rangle_0~,\\ \mu_0\vert n,z\rangle_0=\vert n-1,z\rangle_0~, \end{array} \end{equation} and \begin{equation} \begin{array}{l} {\cal X}^+\vert n,z\rangle_+=q^{2n}z\vert n,z\rangle_+~,\\ {\cal D}_+\vert n,z\rangle_+=q^{1-2n}\lambda^{-1}z^{-1}\vert n-1,z\rangle_+~,\\ \mu_+\vert n,z\rangle_+=\vert n-1,z\rangle_+~. \end{array} \end{equation} It has been strongly argued that the most direct and obvious physical cutoff of spacetime is from the formation of microscopic black holes as soon as too much energy would be accumulated into too small a region. Thus, from a physical point of view, the black holes should provide for a natural cutoff all by themselves. The cutoff distance scale is suspected to be the Planck scale. Because of this origin of spacetime cutoff, any geometry we working on there should be a same minimal cutoff $l_p$. For the classical geometry, the spacetime regularization is equal distance and thus there is one degree of freedom per Planck area. However, from the above discuss of noncommutative geometry, the spacetime is discretely latticed with exponentially increasing space distances. Thus, much less information can be stored in the noncommutative geometry. In fact, this may be the origin of the holography for quantum system of gravity. For the noncommutative space $SO_q(3)$ with radius $R$, if the assumed minimal cutoff induced by the black holes themselves is the Planck scale $l_p$, it is not difficult to count the degrees of freedom ${\cal N}_{\rm bulk}$, \begin{equation} \begin{array}{rcl} {\cal N}_{\rm bulk}&\approx&\displaystyle\sum_{i=1}^N\frac{(q^{2i}l_p)^2}{(q^{2i}l_p-q^{2(i-1)} l_p)^2}~,~~~~~~~~q^{2N}=R~,\\ &=&\displaystyle\frac{q^4\ln\left(\frac{R}{l_p}\right)}{2(q^2-1)^2\ln q}~. \end{array} \end{equation} By taking the deformation parameter of the noncommutative space value to be determined by the algebraic equation \begin{equation} q^{-4}(q^2-1)^2\ln q=\frac{\ln\left(\frac{R}{l_p}\right)}{8\pi\left(\frac{R}{l_p} \right)^2}~, \end{equation} we can check that ${\cal N}_{\rm bulk}$ is exact equal to the degrees of freedom on the classical limit of its boundary surface with radius $R$, ${\cal N}_{\rm boundary}$, \begin{equation} {\cal N}_{\rm boundary}=\frac{4\pi R^2}{l_p^2}~. \end{equation} And then one can write down the entropy of our world at Planck scale, \begin{equation} S=4\pi M^2+C~, \end{equation} where $M$ is the mass of the world (black hole) in natural units and $C$ a constant entropy can not determined. This just is what was called holography or dimension reduction in quantum theory of gravity by 't Hooft. At Planck scale, our world is not $3+1$ dimensional. Rather, the observable degrees of freedom can best be described as if they were Boolean variables defined on a $2$ dimensional lattice, evolving with time. It is clear now that the exact meaning of the holography can interpreted as that the quantum theory of gravity in higher dimensional noncommutative space is equivalent to the theory living on the classical limit of spacetime boundary. This supplies a reasonable picture for the 't Hooft's holography. \section{Noncommutative AdS$_{2n}$ space and exponential regularization} The noncommutative AdS$_{2n}$ space can be defined as the $2n$-dimensional noncommutative real hyperboloid embedded in a ($2n+1$)-dimensional space with coordinates $x^i$ ($i=-n,~-n+1,~\cdots,~-1,~0,~1,~\cdots,~n$), \begin{equation} \begin{array}{l} \displaystyle\frac{1}{1+q^{2n-1}}C_{ij}x^ix^j=-\frac{1}{a^2}~,\\ {\overline x^i}=C_{ji}M_{jk}x^k~, \end{array} \end{equation} where $\rho_{-i}=i-\frac{1}{2}$, $\rho_0=0$, $\rho_i=-i+\frac{1}{2}$, and the metric $C_{ij}=q^{-\rho_i}\delta_{i,-j}$ and $$M=\left(\begin{array}{ccc} 1& & \\ &\underbrace{\begin{array}{ccccc} -1& & & &\\ &-1& & &\\ & &\ldots& &\\ & & &-1& \\ & & & &-1 \end{array}}_{2n-1}&\\ & & 1\end{array}\right)~.$$ It is easy to check that, at the $q\longrightarrow 1$ limit, the noncommutative AdS$_{2n}$ space deduces to the familiar AdS$_{2n}$ space $$\eta_{ab}x^ax^b=-\frac{1}{a^2}$$ in $R^{2n+1}$ with Cartesian coordinates $x^a$, where $\eta_{ab}= {\rm diag}(-1,~\underbrace{1,~\cdots,~1}_{2n-1},~-1)$. For the noncommutative AdS$_{2n}$ space, in components, the commutation relations among coordinates are \begin{equation} \begin{array}{rcl} x^ix^j&=&qx^jx^i~,~~~~~{\rm for}~i<j~ ~{\rm and}~~i\not=-j~,\\ x^ix^{-i}&=&x^{-i}x^i+\lambda q^{i-3/2}L_{i-1}\\ &=&q^{-2}x^{-i}x^i+\lambda q^{i-3/2}L_i~,~~~~~{\rm for}~i>0~, \end{array} \end{equation} where we have used the notation for intermediate lengths $$L_i=\sum_{k=1}^i q^{\rho_k} x^{-k}x^k+\frac{q}{1+q}x^0x^0~.$$ By making use of the intermediate Laplacians $$\Delta_i=\sum_{k=1}^iq^{\rho_k}\partial_k\partial_{-k}+\frac{q}{1+q} \partial_0\partial_0~,$$ the algebra satisfied by the derivatives can be written compactly as \begin{equation} \begin{array}{rcl} \partial_i\partial_j&=&q^{-1}\partial_j\partial_i~,~~~~~{\rm for}~i<j~ ~{\rm and}~~i\not=-j~,\\ \partial_{-i}\partial_i&=&\partial_i\partial_{-i}+\lambda q^{i-3/2}\Delta_{i-1}\\ &=&q^{-2}\partial_i\partial_{-i}+\lambda q^{i-3/2}\Delta_i~,~~~~~{\rm for}~i>0~. \end{array} \end{equation} The commutation relations among the coordinates and derivatives are as follows \begin{equation} \begin{array}{l} \partial_{-i}x^i=x^i\partial_{-i}~,~~~~~{\rm for}~i\not=0~,\\ \partial_ix^j=qx^j\partial_i~,~~~~~{\rm for}~j>-i,~{\rm and}~j\not=i~,\\ \partial_ix^j=qx^j\partial_i-q\lambda q^{-\rho_j-\rho_k}x^{-i}\partial_{-j}~, ~~~~~{\rm for}~j<-i~,~~{\rm and}~i\not=j~,\\ \partial_jx^j=1+q^2x^j\partial_j+q\lambda\displaystyle\sum_{k>j}x^k\partial_k -q^{1-2\rho_j}\lambda x^{-j}\partial_{-j}~,~~~~~{\rm for}~j<0~,\\ \partial_0x^0=1+qx^0\partial_0+q\lambda\displaystyle\sum_{k>0}x^k\partial_k~,\\ \partial_jx^j=1+q^2x^j\partial_j+q\lambda\displaystyle\sum_{k>j}x^k\partial_k~, ~~~~~{\rm for}~j>0~. \end{array} \end{equation} For convenient, we introduce the dilatation operator $\Lambda_m$ ($0<m\leq n$) as $$ \displaystyle \Lambda_m=1+q\lambda E_m+q^{2m+1}\lambda^2L_m\Delta_m~,~~~~~ \displaystyle E_m=\sum_{j=-m}^mx^j\partial_j~.$$ The dilatation operator satisfies \begin{equation} \Lambda_m x^k=q^2x^k \Lambda_m~,~~~~~~~\Lambda_m\partial _k=q^{-2}\partial_k\Lambda_m~,~~~~{\rm for}~ \vert k\vert\leq m~. \end{equation} The noncommutative AdS$_{2n}$ space is accompanied with the conjugation\cite{12} \begin{equation}\label{ccc} \begin{array}{l} {\overline x^i}=C_{ji}M_{jk}x^k~,\\ {\overline \partial_i}=-q^{-2}\Lambda_{n}^{-1}C_{ij}M_{jk}[\Delta_n,x^k]~. \end{array} \end{equation} For $k>0$, using the notations $$\begin{array}{l} y^{-k}=x^{-k}+q^{k+1/2}\lambda L_k\partial_{+k}~,\\ \delta_{-k}=\partial_{-k}+q^{k+1/2}\lambda \Delta_kx^{+k}~, \end{array}$$ we can construct a set of independent basis\cite{13} on the noncommutative AdS$_{2n}$ space, \begin{equation} \begin{array}{l} {\cal X}^n=x^n~,\\ {\cal D}_n=\partial_n~,\\ {\cal X}^{+j}=\mu_{n}^{-1/2}\mu_{n-1}^{-1/2}\cdots \mu_{j+1}^{-1/2}x^{+j}~,~~~~~{\rm for}~n>j\geq 0~,\\ {\cal D}_{+j}=\mu_{n}^{-1/2}\mu_{n-1}^{-1/2}\cdots \mu_{j+1}^{-1/2}\partial_{+j}~,\\ {\cal X}^{-j}=\mu_{n}^{-1/2}\mu_{n-1}^{-1/2}\cdots\mu_{j+1}^{-1/2} \Lambda_{+j}^{-1/2}\mu_{+j}^{-1/2}y^{-j}~,\\ {\cal D}_{-j}=q^{-1}\mu_{n}^{-1/2}\mu_{n-1}^{-1/2}\cdots\mu_{j+1}^{-1/2} \Lambda_{+j}^{-1/2}\mu_{+j}^{-1/2}\delta_{-j}~,\\ {\cal X}^{-n}=\Lambda_{+n}^{-1/2}\mu_{+n}^{-1/2}y^{-n}~,\\ {\cal D}_{-j}=q^{-1}\Lambda_{+n}^{-1/2}\mu_{+n}^{-1/2}\delta_{-n}~, \end{array} \end{equation} where $(\mu_{\pm i})^{\pm 1}={\cal D}_{\pm i}{\cal X}^{\pm i}- {\cal X}^{\pm i}{\cal D}_{\pm i}$ and $\mu_0^{1/2}={\cal D}_0{\cal X}^0- {\cal X}^0{\cal D}_0$.\\ We note that the $\mu_i$'s satisfy simple commutation relations with the new variables ${\cal X}^j$ and ${\cal D}_j$, $$[\mu_i,\mu_j]=0~,~~~~~~ \mu_i{\cal X}^j={\cal X}^j\mu_i\cdot\left\{\begin{array}{l} q^2~,~~~{\rm for}~i=j\\ 1~,~~~{\rm for}~i\not=j \end{array}\right.~,~~~~~~ \mu_i{\cal D}^j={\cal D}^j\mu_i\cdot\left\{\begin{array}{l} q^{-2}~,~~~{\rm for}~i=j\\ 1~,~~~{\rm for}~i\not=j \end{array}\right.~.$$ For the new basis of coordinates and derivatives on the noncommutative AdS$_{2n}$ space, it is not difficult to show that \begin{equation} \begin{array}{l} {\cal D}_{-k}{\cal X}^{-k}=1+q^{-2}{\cal X}^{-k}{\cal D}_{-k}~,~~~~~{\rm for} ~k>0~,\\ {\cal D}_{0}{\cal X}^{0}=1+q{\cal X}^{0}{\cal D}_{0}~,\\ {\cal D}_{+k}{\cal X}^{+k}=1+q^{2}{\cal X}^{+k}{\cal D}_{+k}~,\\[2mm] [{\cal D}_i,{\cal D}_j]=0~,~~~~~[{\cal X}^i,{\cal X}^j]=0~,\\[2mm] {\cal D}_i{\cal X}^j={\cal X}^j{\cal D}_i~,~~~~~{\rm for}~ i\not=j~. \end{array} \end{equation} The noncommutative AdS$_{2n}$ space in terms of the ${\cal X}^i$ and ${\cal D}_i$ is of the form \begin{equation} \sum_{j=1}^n q^{\rho_j-2(n-j)}\Lambda_j^{1/2}\mu_j^{-1/2} {\cal X}^j{\cal X}^{-j} +\frac{q^{-2n+1}}{1+q}{\cal X}^0{\cal X}^0=-\frac{1}{a^2}~. \end{equation} The conjugate operation on the independent set of operators ${\cal X}^j$ and ${\cal D}_j$ is induced from what on $x^i$ and $\partial_i$ (Eq.(\ref{ccc})), \begin{equation} \begin{array}{l} {\overline {\cal X}^n}={\overline x^n}~,\\ {\overline {\cal D}_n}={\overline \partial_n}~,\\ {\overline {\cal X}^{+j}}={\overline x^{+j}}{\overline \mu_{j+1}}^{-1/2} {\overline \mu_{j+2}}^{-1/2}\cdots {\overline \mu_{n}}^{-1/2}~,~~~~~{\rm for}~n>j\geq 0~,\\ {\overline {\cal D}_{+j}}={\overline \partial_{+j}}{\overline \mu_{j+1}}^{-1/2} {\overline \mu_{j+2}}^{-1/2}\cdots {\overline \mu_{n}}^{-1/2}~,\\ {\overline {\cal X}^{-j}}={\overline y^{-j}}{\overline \mu_{+j}}^{-1/2} {\overline \Lambda_{+j}}^{-1/2}{\overline \mu_{j+1}}^{-1/2} {\overline \mu_{j+2}}^{-1/2}\cdots {\overline \mu_{n}}^{-1/2}~,\\ {\overline {\cal D}_{-j}}=q^{-1}{\overline \delta_{-j}}{\overline \mu_{+j}}^{-1/2}{\overline \Lambda_{+j}}^{-1/2} {\overline \mu_{j+1}}^{-1/2}{\overline \mu_{j+2}}^{-1/2}\cdots {\overline \mu_{n}}^{-1/2}~,\\ {\overline {\cal X}^{-n}}={\overline y^{-n}}{\overline \mu_{+n}}^{-1/2} {\overline \Lambda_{+n}}^{-1/2}~,\\ {\overline {\cal D}_{-j}}=q^{-1}{\overline \delta_{-n}} {\overline \mu_{+n}}^{-1/2}{\overline \Lambda_{+n}}^{-1/2}~. \end{array} \end{equation} Thus, the Quantum Heisenberg-Weyl algebra corresponds to the noncommutative AdS$_{2n}$ space can be decoupled into $(2n+1)$ independent subalgebras. For the quantum algebra ${\cal A}_{-k}$ ($ n\geq k>0$) \begin{equation} \begin{array}{l} {\cal D}_{-k}{\cal X}^{-k}-{\cal X}^{-k}{\cal D}_{-k}=\mu_{-k}^{-1}~,~~~~~~ \mu_{-k} {\cal X}^{-k}=q^2{\cal X}^{-k}\mu_{-k}~,\\ \mu_-{\cal D}_{-k}=q^{-2}{\cal D}_{-k}\mu_{-k}~,~~~~~~\mu_-^{-1}\equiv 1+(q^{-2}-1){\cal X}^{-k}{\cal D}_{-k}~, \end{array} \end{equation} we construct the quantum coherent state $\vert 0, z\rangle_{-k}$ as \begin{equation} \begin{array}{l} \vert 0,z\rangle_{-k}=\displaystyle\exp_{q^2}(-\frac{1}{2}|q^{-2}z|)\sum_{m=0}^\infty \frac{(-q^{-2}z)^m}{[m]_{q^2}!}({\cal D}_{-k})^m\vert 0\rangle_{-k}~,\\ {\cal X}^{-k}\vert 0,z\rangle_{-k}=z\vert 0,z\rangle_{-k}~, \end{array} \end{equation} where the reference state $\vert 0\rangle_{-k}$ was chosen such that ${\cal X}^{-k}\vert 0\rangle_{-k} =0$. From the coherent state $\vert 0,z\rangle_{-k}$, we can construct a representation for the quantum algebra ${\cal A}_{-k}$ as \begin{equation} \begin{array}{l} {\cal X}^{-k}\vert n,z\rangle_-=q^{2n}z\vert n,z\rangle_{-k}~,\\ {\cal D}_{-k}\vert n,z\rangle_-=-q^{-1-2n}\lambda^{-1}z^{-1}\vert n+1, z\rangle_{-k}~,\\ \mu_{-k}\vert n,z\rangle_-=\vert n-1,z\rangle_{-k}~. \end{array} \end{equation} The quantum coherent state corresponds to the quantum algebras ${\cal A}_0$ \begin{equation} \begin{array}{l} {\cal D}_0{\cal X}^0-{\cal X}^0{\cal D}_0=\mu_0^{1/2}~,~~~~~~ \mu_0 {\cal X}^0=q^2{\cal X}^0\mu_0~,\\ \mu_0{\cal D}_0=q^{-2}{\cal D}_0\mu_0~,~~~~~~\mu_0^{1/2}\equiv 1+(q-1){\cal X}^0{\cal D}_0~, \end{array} \end{equation} is \begin{equation} \begin{array}{l} \vert 0, z\rangle_0=\displaystyle\exp_{q^{-1}}(-\frac{1}{2}|qz|)\sum_{m=0}^\infty \frac{(-qz)^m}{[m]_{q^{-1}}!}({\cal D}_0)^m\vert 0 \rangle_0~,~~~~~{\cal X}^0\vert 0\rangle_0=0~,\\ {\cal X}^{0}\vert 0,z\rangle_{0}=z\vert 0,z\rangle_{0}~. \end{array} \end{equation} Representation of the quantum algebra ${\cal A}_0$ based on the quantum coherent state $\vert 0,z\rangle_0$ is as \begin{equation} \begin{array}{l} {\cal X}^0\vert n,z\rangle_0=q^{n}z\vert n,z\rangle_0~,\\ {\cal D}_0\vert n,z\rangle_0=\displaystyle\frac{q^{1-n}}{q-1}z^{-1} \vert n-1,z\rangle_0~,\\ \mu_0\vert n,z\rangle_0=\vert n-1,z\rangle_0~. \end{array} \end{equation} The quantum coherent state corresponds to the quantum algebra ${\cal A}_{+k}$ \begin{equation} \begin{array}{l} {\cal D}_{+k}{\cal X}^{+k}-{\cal X}^{+k}{\cal D}_{+k}=\mu_{+k}~,~~~~~~ \mu_{+k} {\cal X}^{+k}=q^2{\cal X}^{+k}\mu_{+k}~,\\ \mu_{+k}{\cal D}_{+k}=q^{-2}{\cal D}_{+k}\mu_{+k}~,~~~~~~\mu_{+k}\equiv 1+(q^2-1){\cal X}^{+k}{\cal D}_{+k}~, \end{array} \end{equation} is of the form \begin{equation} \begin{array}{l} \vert 0, z\rangle_{+k}=\displaystyle\exp_{q^{-2}}(-\frac{1}{2}|q^2z|)\sum_{m=0}^\infty \frac{(-q^2z)^m}{[m]_{q^{-2}}!}({\cal D}_{+k})^m\vert 0 \rangle_{+k}~,~~~~~{\cal X}^{+k}\vert 0\rangle_{+k}=0~,\\ {\cal X}^{+k}\vert 0,z\rangle_{+k}=z\vert 0,z\rangle_{+k}~. \end{array} \end{equation} Then, by writing down the representation of the quantum algebra ${\cal A}_{+k}$ \begin{equation} \begin{array}{l} {\cal X}^{+k}\vert n,z\rangle_{+k}=q^{2n}z\vert n,z\rangle_{+k}~,\\ {\cal D}_{+k}\vert n,z\rangle_{+k}=q^{1-2n}\lambda^{-1}z^{-1}\vert n-1,z\rangle_{+k}~,\\ \Lambda_{+k}\vert n,z\rangle_{+k}=\vert n-1,z\rangle_{+k}~, \end{array} \end{equation} we complete the Hilbert space constructure of the noncommutative AdS$_{2n}$ space. This shows that the noncommutative AdS$_{2n}$ space is discretely latticed with exponentially increasing space distances. The minimal cutoff induced by the quantum gravity itself is the Planck scale $l_p$. As in the case of noncommutative orthogonal space, the exponential regularization may effectively reduce the amount of observable degrees of freedom of the noncommutative AdS$_{2n}$ space, even one can not enumerate it exactly because of being infinite\cite{14}. An adequately adopted noncommutative deformation parameter $q$ (it may be even more closer to $1$ than the case of limited geometry) can equate the entropy of the quantum system of gravity in the bulk of noncommutative AdS$_{2n}$ space and what on the classical limit of its boundary. The commutative boundary is equal distance regularized and possesses conformally invariant symmetry. This is crucial for the AdS/CFT correspondence. \section{Noncommutative AdS$_{2n-1}$ space with exponential regularization} For the noncommutative AdS$_{2n-1}$ space, commutation relations among coordinates $x^i$ ($i=-n,~-n+1,~\cdots,~-2,~-1,~+1,~+2,~\cdots,~n$) are \begin{equation} \begin{array}{rcl} x^ix^j&=&qx^jx^i~,~~~~~{\rm for}~i<j~ ~{\rm and}~~i\not=-j~,\\ x^ix^{-i}&=&x^{-i}x^i+\lambda q^{i-2}L_{i-1}\\ &=&q^{-2}x^{-i}x^i+\lambda q^{i-2}L_i~,~~~~~{\rm for}~i>0~, \end{array} \end{equation} where we have used the notation for intermediate lengths $$L_i=\sum_{k=1}^i q^{\rho_k} x^{-k}x^k~,~~~~\rho_{-k}=k-1~,~~~~\rho_k=-k+1~.$$ By making use of the intermediate Laplacians $$\Delta_i=\sum_{k=1}^iq^{\rho_k}\partial_k\partial_{-k},$$ the algebra satisfied by the derivatives can be written compactly as \begin{equation} \begin{array}{rcl} \partial_i\partial_j&=&q^{-1}\partial_j\partial_i~,~~~~~{\rm for}~i<j~ ~{\rm and}~~i\not=-j~,\\ \partial_{-i}\partial_i&=&\partial_i\partial_{-i}+\lambda q^{i-2}\Delta_{i-1}\\ &=&q^{-2}\partial_i\partial_{-i}+\lambda q^{i-2}\Delta_i~,~~~~~{\rm for}~i>0~. \end{array} \end{equation} The commutation relations among the coordinates and derivatives are as follows \begin{equation} \begin{array}{l} \partial_{-i}x^i=x^i\partial_{-i}~,~~~~~{\rm for}~i\not=0~,\\ \partial_ix^j=qx^j\partial_i~,~~~~~{\rm for}~j>-i,~{\rm and}~j\not=i~,\\ \partial_ix^j=qx^j\partial_i-q\lambda q^{-\rho_j-\rho_k}x^{-i}\partial_{-j}~, ~~~~~{\rm for}~j<-i~,~~{\rm and}~i\not=j~,\\ \partial_jx^j=1+q^2x^j\partial_j+q\lambda\displaystyle\sum_{k>j}x^k\partial_k -q^{1-2\rho_j}\lambda x^{-j}\partial_{-j}~,~~~~~{\rm for}~j<0~,\\ \partial_jx^j=1+q^2x^j\partial_j+q\lambda\displaystyle\sum_{k>j}x^k\partial_k~, ~~~~~{\rm for}~j>0~. \end{array} \end{equation} For convenient, we introduce the dilatation operator $\Lambda_m$ ($0<m\leq n$) as $$\Lambda_m=1+q\lambda E_m+q^{2m}\lambda^2(1+q^{2n-2})L_m\Delta_m~,~~~~~ E_m=\sum_{j=-m}^mx^j\partial_j~.$$ The dilatation operator satisfies \begin{equation} \Lambda_m x^k=q^2x^k \Lambda_m~,~~~~~~~\Lambda_m\partial _k=q^{-2}\partial_k\Lambda_m~,~~~~{\rm for}~ <0\vert k\vert\leq m~. \end{equation} The noncommutative AdS$_{2n-1}$ space is accompanied with the conjugation \begin{equation} \begin{array}{l} {\overline x^i}=C_{ji}N_{jk}x^k~,\\ {\overline \partial_i}=-q^{-2}\Lambda_{n}^{-1}C_{ij}N_{jk}[\Delta_n,x^k]~, \end{array} \end{equation} where $$N=\left(\begin{array}{ccc} 1& & \\ &\underbrace{\begin{array}{ccccc} -1& & & &\\ &-1& & &\\ & &\ldots& &\\ & & &-1& \\ & & & &-1 \end{array}}_{2n-2}&\\ & & 1\end{array}\right)~.$$ For $k>0$, using the notations $$\begin{array}{l} y^{-k}=x^{-k}+q^{k+1/2}\lambda (1+q^{2n-2})L_k\partial_{+k}~,\\ \delta_{-k}=\partial_{-k}+q^{k+1/2}\lambda \Delta_kx^{+k}~, \end{array}$$ we can construct a new set of variables for the coordinates and derivatives, \begin{equation} \begin{array}{l} {\cal X}^n=x^n~,\\ {\cal D}_n=\partial_n~,\\ {\cal X}^{+j}=\mu_{n}^{-1/2}\mu_{n-1}^{-1/2}\cdots \mu_{j+1}^{-1/2}x^{+j}~,~~~~~{\rm for}~n>j> 0~,\\ {\cal D}_{+j}=\mu_{n}^{-1/2}\mu_{n-1}^{-1/2}\cdots \mu_{j+1}^{-1/2}\partial_{+j}~,\\ {\cal X}^{-j}=\mu_{n}^{-1/2}\mu_{n-1}^{-1/2}\cdots\mu_{j+1}^{-1/2} \Lambda_{+j}^{-1/2}\mu_{+j}^{-1/2}y^{-j}~,\\ {\cal D}_{-j}=q^{-1}\mu_{n}^{-1/2}\mu_{n-1}^{-1/2}\cdots\mu_{j+1}^{-1/2} \Lambda_{+j}^{-1/2}\mu_{+j}^{-1/2}\delta_{-j}~,\\ {\cal X}^{-n}=\Lambda_{+n}^{-1/2}\mu_{+n}^{-1/2}y^{-n}~,\\ {\cal D}_{-j}=q^{-1}\Lambda_{+n}^{-1/2}\mu_{+n}^{-1/2}\delta_{-n}~, \end{array} \end{equation} where $(\mu_{\pm i})^{\pm 1}={\cal D}_{\pm i}{\cal X}^{\pm i}- {\cal X}^{\pm i}{\cal D}_{\pm i}$.\\ Commutation relations among the new basis of coordinates and derivatives on the noncommutative AdS$_{2n-1}$ space are \begin{equation} \begin{array}{l} {\cal D}_{-k}{\cal X}^{-k}=1+q^{-2}{\cal X}^{-k}{\cal D}_{-k}~,~~~~~{\rm for} ~k>0~,\\ {\cal D}_{+k}{\cal X}^{+k}=1+q^{2}{\cal X}^{+k}{\cal D}_{+k}~,\\[2mm] [{\cal D}_i,{\cal D}_j]=0~,~~~~~[{\cal X}^i,{\cal X}^j]=0~,\\[2mm] {\cal D}_i{\cal X}^j={\cal X}^j{\cal D}_i~,~~~~~{\rm for}~ i\not=j~. \end{array} \end{equation} The noncommutative AdS$_{2n-1}$ space in terms of the ${\cal X}^i$ and ${\cal D}_i$ is of the form \begin{equation} \sum_{j=1}^n q^{\rho_j-2(n-j)}\Lambda_j^{1/2}\mu_j^{-1/2} {\cal X}^j{\cal X}^{-j}=-\frac{1}{a^2}~. \end{equation} The conjugate operation on the operators ${\cal X}^j$ and ${\cal D}_j$ is induced by the conjugation on $x^i$ and $\partial_i$, \begin{equation} \begin{array}{l} {\overline {\cal X}^n}={\overline x^n}~,\\ {\overline {\cal D}_n}={\overline \partial_n}~,\\ {\overline {\cal X}^{+j}}={\overline x^{+j}}{\overline \mu_{j+1}}^{-1/2} {\overline \mu_{j+2}}^{-1/2}\cdots {\overline \mu_{n}}^{-1/2}~,~~~~~{\rm for}~n>j> 0~,\\ {\overline {\cal D}_{+j}}={\overline \partial_{+j}}{\overline \mu_{j+1}}^{-1/2} {\overline \mu_{j+2}}^{-1/2}\cdots {\overline \mu_{n}}^{-1/2}~,\\ {\overline {\cal X}^{-j}}={\overline y^{-j}}{\overline \mu_{+j}}^{-1/2} {\overline \Lambda_{+j}}^{-1/2}{\overline \mu_{j+1}}^{-1/2} {\overline \mu_{j+2}}^{-1/2}\cdots {\overline \mu_{n}}^{-1/2}~,\\ {\overline {\cal D}_{-j}}=q^{-1}{\overline \delta_{-j}}{\overline \mu_{+j}}^{-1/2}{\overline \Lambda_{+j}}^{-1/2} {\overline \mu_{j+1}}^{-1/2}{\overline \mu_{j+2}}^{-1/2}\cdots {\overline \mu_{n}}^{-1/2}~,\\ {\overline {\cal X}^{-n}}={\overline y^{-n}}{\overline \mu_{+n}}^{-1/2} {\overline \Lambda_{+n}}^{-1/2}~,\\ {\overline {\cal D}_{-j}}=q^{-1}{\overline \delta_{-n}} {\overline \mu_{+n}}^{-1/2}{\overline \Lambda_{+n}}^{-1/2}~. \end{array} \end{equation} Then, the Quantum Heisenberg-Weyl algebra corresponds to the noncommutative AdS$_{2n-1}$ space is decoupled into $2n$ independent subalgebras. For the quantum algebra ${\cal A}_{-k}$ ($ n\geq k>0$) \begin{equation} \begin{array}{l} {\cal D}_{-k}{\cal X}^{-k}-{\cal X}^{-k}{\cal D}_{-k}=\mu_{-k}^{-1}~,~~~~~~ \mu_{-k} {\cal X}^{-k}=q^2{\cal X}^{-k}\mu_{-k}~,\\ \mu_-{\cal D}_{-k}=q^{-2}{\cal D}_{-k}\mu_{-k}~,~~~~~~\mu_-^{-1}\equiv 1+(q^{-2}-1){\cal X}^{-k}{\cal D}_{-k}~, \end{array} \end{equation} we have the quantum coherent state $\vert 0,z\rangle_{-k}$ as \begin{equation} \begin{array}{l} \vert 0,z\rangle_{-k}=\displaystyle\exp_{q^2}(-\frac{1}{2}|q^{-2}z|)\sum_{m=0}^\infty \frac{(-q^{-2}z)^m}{[m]_{q^2}!}({\cal D}_{-k})^m\vert 0\rangle_{-k}~,~~~~~{\cal X}^{-k}\vert 0\rangle_{-k} =0~,\\ {\cal X}^{-k}\vert 0,z\rangle_{-k} =z\vert 0,z\rangle_{-k}~. \end{array} \end{equation} From the coherent state $\vert 0,z\rangle_{-k}$, we can construct a representation for the quantum algebra ${\cal A}_{-k}$ as \begin{equation} \begin{array}{l} {\cal X}^{-k}\vert n,z\rangle_-=q^{2n}z\vert n,z\rangle_{-k}~,\\ {\cal D}_{-k}\vert n,z\rangle_-=-q^{-1-2n}\lambda^{-1}z^{-1}\vert n+1, z\rangle_{-k}~,\\ \mu_{-k}\vert n,z\rangle_-=\vert n-1,z\rangle_{-k}~. \end{array} \end{equation} Similarly, we can construct reference state$\vert 0,z\rangle_{+k}$ ($n\geq k>0$) as \begin{equation} \begin{array}{l} \vert 0, z\rangle_{+k}=\displaystyle\exp_{q^{-2}}(-\frac{1}{2}|q^2z|)\sum_{m=0}^\infty \frac{(-q^2z)^m}{[m]_{q^{-2}}!}({\cal D}_{+k})^m\vert 0 \rangle_{+k}~,~~~~~{\cal X}^{+k}\vert 0\rangle_{+k}=0~,\\ {\cal X}^{+k}\vert 0,z\rangle_{+k} =z\vert 0,z\rangle_{+k}~. \end{array} \end{equation} The corresponding representation of the quantum algebras ${\cal A}_{+k}$ \begin{equation} \begin{array}{l} {\cal D}_{+k}{\cal X}^{+k}-{\cal X}^{+k}{\cal D}_{+k}=\mu_{+k}~,~~~~~~ \mu_{+k} {\cal X}^{+k}=q^2{\cal X}^{+k}\mu_{+k}~,\\ \mu_{+k}{\cal D}_{+k}=q^{-2}{\cal D}_{+k}\mu_{+k}~,~~~~~~\mu_{+k}\equiv 1+(q^2-1){\cal X}^{+k}{\cal D}_{+k}~, \end{array} \end{equation} is of the form \begin{equation} \begin{array}{l} {\cal X}^{+k}\vert n,z\rangle_{+k}=q^{2n}z\vert n,z\rangle_{+k}~,\\ {\cal D}_{+k}\vert n,z\rangle_{+k}=q^{1-2n}\lambda^{-1}z^{-1}\vert n-1,z\rangle_{+k}~,\\ \Lambda_{+k}\vert n,z\rangle_{+k}=\vert n-1,z\rangle_{+k}~. \end{array} \end{equation} As the noncommutative AdS$_{2n}$ space did, the noncommutative AdS$_{2n-1}$ space is also discretely latticed with exponentially increasing space distances. The minimal cutoff induced by the quantum gravity itself is the Planck scale $l_p$. The exponential regularization effectively reduces degrees of freedom in the noncommutative AdS$_{2n-1}$ space. A very small amount of displacement of the noncommutative deformation parameter $q$ from unity equates the entropy of the quantum theory of gravity in the bulk of noncommutative AdS$_{2n-1}$ space and what on the commutative limit of its boundary surface. The commutative boundary is equal distance regularized and possesses a conformally invariant symmetry. Thus, the equivalent theory living on the spacetime boundary of the quantum system of gravity on the background of noncommutative AdS space is a conformal field theory. This is the bases for the AdS/CFT correspondence. \section{Concluding remarks} In this paper, by constructing Hilbert space with quantum coherent state as reference one, we have presented a kind of special regularization with exponentially increasing spacetime cutoff for both orthogonal and AdS space based on noncommutative geometry. It has been used to argue that the same minimal cutoff of any geometry is the Planck scale $l_p$, that the most direct and obvious physical cutoff is from the formation of microscopic black holes as soon as too much energy would be accumulated into too small a region. We have obtained results which show a very small ($<10^{-15}$) displacement of the noncommutative deformation parameter from its classical value ($1$) reduces sharply the entropy of quantum system of gravity. The noncommutative deformation parameter was determined by a well-defined algebraic equation. The noncommutative space $SO_q(3)$ with such a deformation parameter have the same entropy or degrees of freedom with classical $S^3$ surface. This is just the so called holography for quantum theory of gravity. The holography makes the quantum theory of gravity on the noncommutative AdS$_{d}$ space equivalent to the conformally invariant quantum field theory living on the classical limit of its boundary. This is the bases of the AdS/CFT correspondence of string theory and M theory. Here we should stress that the proper geometry for quantum gravity may be noncommutative but not classical continuum geometry. This is in agreement with the long time suspecting that the small scale structure of spacetime might not be adequately described by classical continuum geometry and the noncommutative spacetime might be a realistic picture of how spacetime behaves near the Planck scale. Strong quantum fluctuations of gravity at this spacetime scale may make points in space fuzzy. All of the strangeness' for the quantum theories in different spacetime dimensions ever could be equivalent are come from the noncommutative geometry description of the quantum gravity. The exact form of the AdS/CFT is concerned with noncommutative AdS space and the classical limit of its boundary surface (on which the conformal field theory lives). This suggests that the gravity-gauge theory connection should be of the form: the string theory or M theory on the noncommutative background of the form AdS$_d^q\times M^q_{D-d}$ is dual to a conformal field theory living on the classical limit of spacetime boundary. For the Type IIB string theory on the noncommutative background AdS$^q_5\times S_q^5$ spectra can compare with low-order correlation functions of the $3+1$ dimensional ${\cal N}=4$ $SU(N)$ super Yang-Mills theory. The dual super Yang-Mills theory lives on the classical limit of noncommutative AdS spacetime boundary. Only both in the large $N$ and commutative limit, the string theory is weakly coupled and the supergravity is a good approximation to it. Therefore, in fact as 't Hooft suspected\cite{01}, the nature is much more crazy at the Planck scale than even string theorists could have imagined. Formalisms of string theory (gravity) on noncommutative geometry have to be constructed to gain sights of unification of gravity and quantum mechanics. \bigskip \bigskip \centerline{\large\bf Acknowledgments } I would like to thank Prof. J. Wess for introducing the problem to me and for enlightening discussions. I am grateful to H. Steinacker for valuable discussions. The work was supported in part by the National Science Foundation of China under Grant 19625512.
\section{Introduction} \label{sec:introduction} Quite an industry has developed around the study of nonlinear phenomena occurring at the threshold of black hole formation first discovered by Choptuik~\cite{matt}. A number of reviews detail the substantial progress in understanding these phenomena~\cite{matt_rev,carsten_rev}. By specifying a form of initial data parameterized by some parameter $p$, numerical evolutions can determine whether the value of $p$ is sufficient to form a black hole. Such a solution is considered super-critical, while for small $p$, no black hole forms and the solution is sub-critical. Separating super-critical solutions from sub-critical solutions is the critical solution parameterized by the critical value $p^*$. In analogy with statistical mechanics where the black hole mass represents the order parameter, cases where infinitesimal mass black holes are formed are called Type~II, and cases in which only finite mass black holes are formed are called Type~I. Here, I present a new Type~II critical solution occurring within a model with a triplet scalar field. By assuming a hedgehog ansatz for the triplet, I find a critical solution which is, in a strict sense, non-spherically symmetric. Adding a potential allows the study of critical phenomena within a global monopole. No Type~I critical behavior around the static monopole solutions is observed. As a generalization of Choptuik's original model, the triplet scalar field contains the single scalar field model within a certain region of parameter space. However because of the different assumed symmetries of the fields, the hedgehog ansatz excludes this region. Hence, it is not surprising that I find a new solution, one which is discretely self-similar~(DSS) with echoing period $\Delta=0.46$, much smaller than that for either the original DSS~($\Delta=3.44$) or that of the Einstein Yang-Mills~(EYM) model ($\Delta=0.74$)~\cite{matt_eym}. An ansatz similar to this one (an $SU(2)$ valued hedgehog ansatz) was used to study the collapse of Skyrmions~\cite{bizon1,bizon2,bizon3}. Both these models allow the consideration of non-spherically symmetric initial data, albeit quite special data, within the comparatively simple framework of one spatial dimension, namely a radius $r$. In Section~\ref{sec:model}, I present the most general form of the model under consideration here. First, I choose a triplet scalar field $\Phi^a$ accompanied by the usual symmetry-breaking potential with vacuum value $|\Phi^a|=\eta$ and coupling $\lambda$. For certain evolutions discussed later, I also introduce a free, massless scalar field $\psi(r,t)$ which couples to $\Phi^a$ only through gravity. With the most general form of the equations of motion, Section~\ref{sec:results} then discusses various aspects of regions of the parameter space. In particular, Section~\ref{sec:newdss} presents a new critical solution obtained with the hedgehog ansatz, and Section~\ref{sec:addpsi} discusses the stability of this solution with respect to the addition of $\psi$, a free scalar field. Section~\ref{sec:inside} addresses the results obtained in the interior of a monopole. I then conclude in Section~\ref{sec:conclusion}. \section{The Model} \label{sec:model} Letting $\Phi^a$ represent an $SO(3)$-valued triplet scalar field, and $\psi$ a single-valued, free, massless scalar field, the Lagrangian for the model is \begin{equation} L = - \frac{1}{2} \psi_{;\mu} \psi^{;\mu} - \frac{1}{2} \Phi^a{}^{;\mu} \Phi^a{}_{;\mu} - \frac{1}{4}\lambda \left[ \left( \Phi^a \right)^2 - \eta^2 \right]^2, \label{eq:lagrangian} \end{equation} where $\lambda$ is the coupling to the symmetry-breaking potential, and $\eta$ is the scale of symmetry-breaking. The stress-energy tensor is then \begin{eqnarray} T_{\mu \nu} & = & \psi_{;\mu} \psi_{;\nu} - \frac{1}{2} g_{\mu \nu} \psi^{;\rho} \psi_{;\rho} \nonumber \\ & & + \Phi^a{}_{;\mu} \Phi^a{}_{;\nu} - \frac{g_{\mu \nu}}{2} \left( \Phi^a{}^{;\rho} \Phi^a{}_{;\rho} + \frac{1}{2}\lambda \left[ \left( \Phi^a \right)^2 - \eta^2\right]^2 \right) \end{eqnarray} and the equations of motion for the matter fields are \begin{eqnarray} \Box \psi & = & 0 \\ \Box \Phi^a & = & \lambda \Phi^a \left[ \left( \Phi^a\right)^2 - \eta^2 \right]. \end{eqnarray} The hedgehog ansatz is assumed for the triplet scalar field \begin{equation} \Phi^a = f(r,t)~\hat r = f(r,t) \left( \begin{array}{d} \sin \theta \cos \varphi \\ \sin \theta \sin \varphi \\ \cos \theta \end{array} \right), \end{equation} such that the triplet field is defined to be a vector in internal space of magnitude $f(r,t)$ pointing everywhere in the radial direction, $\hat r$. None of the fields $\Phi^a$ is a function only of $r$ and $t$, but the magnitude $f$ is. Thus, while highly symmetric, the model is not spherically symmetric in the strictest sense. With this hedgehog ansatz, a monopole of unit charge is obtained by enforcing the boundary condition $f(\infty,t)=\eta$. When $f(\infty,t)=0$, no monopole charge exists. I choose the spherically symmetric metric \begin{equation} ds^2 = - \alpha^2(r,t)~dt^2 + a^2(r,t)~dr^2 + r^2~d\Omega^2, \label{eq:metric} \end{equation} and introduce auxiliary variables \begin{eqnarray} \Pi_\psi \equiv \frac{a}{\alpha} \dot \psi, & ~~~~ & \Phi_\psi \equiv \psi',\\ \Pi_f \equiv \frac{a}{\alpha} \dot f, & ~~~~ & \Phi_f \equiv f', \end{eqnarray} where an overdot denotes $\partial/\partial t$ and a prime denotes $\partial/\partial r$. The full system of equations then becomes \begin{eqnarray} \dot \psi & = & \frac{\alpha}{a} \Pi_\psi ~~~~~~~~~~~~~~~~ \dot \Phi_\psi = \left( \frac{\alpha}{a} \Pi_\psi \right)' \label{eq:eom1} \\ \dot \Pi_\psi & = & \frac{1}{r^2} \left( \frac{\alpha}{a} r^2 \Phi_\psi \right)'\\ \dot f & = & \frac{\alpha}{a} \Pi_f ~~~~~~~~~~~~~~~~ \dot \Phi_f = \left( \frac{\alpha}{a} \Pi_f \right)' \\ \dot \Pi_f & = & \frac{1}{r^2} \left( \frac{\alpha}{a} r^2 \Phi_f \right)' -\alpha a \lambda f \Bigl( f^2 -\eta^2\Bigr) -\frac{2\alpha a f}{r^2}\\ \frac{a'}{a} & = & \frac{1 - a^2}{2r} + 2 \pi r \Bigl( \Pi_f^2 + \Phi_f^2 + \Pi_\psi^2 + \Phi_\psi^2 \Bigr) \nonumber \\ & & + 4 \pi a^2 \left[ \frac{f^2}{r} + \frac{r}{4}\lambda \bigl( f^2 - \eta^2 \bigr)^2 \right]\\ \frac{\alpha'}{\alpha} & = & \frac{a^2-1}{2r} +2 \pi r \Bigl( \Pi_f^2 + \Phi_f^2 + \Pi_\psi^2 + \Phi_\psi^2 \Bigr) \nonumber \\ & & -4 \pi a^2 \left[ \frac{f^2}{r} + \frac{r}{4} \lambda \bigl( f^2 - \eta^2 \bigr)^2 \right]\\ \dot a & = & 4 \pi r \alpha \Bigl( \Phi_f \Pi_f + \Phi_\psi \Pi_\psi \Bigr) . \label{eq:eom2} \end{eqnarray} In addition to these equations of motion, appropriate boundary conditions are needed. At the origin of coordinates, regularity is enforced with the conditions \begin{eqnarray} \Phi_\psi(0,t) & = &0 \cr f(0,t) & = &0 ~~~~~~~~~~ \Phi_f'(0,t) = 0 \cr a(0,t) & = &1 ~~~~~~~~~~~\hspace{0.02in} a'(0,t) = 0 \\ \alpha'(0,t) & = &0. \nonumber \end{eqnarray} Further, a mass aspect ratio $m(r,t)$ is defined by associating the $g_{rr}$ metric components of~(\ref{eq:metric}) and Schwarzschild \begin{equation} a^2 = \left( 1-\frac{2m}{r}\right)^{-1}. \end{equation} The quantity $m(r,t)$ then describes the amount of mass contained within the radius $r$ at time $t$. At large radius, an outgoing radiation condition is applied to the scalar field $\psi$. For non-vanishing potential, the field $f(r,t)$ has an effective mass which makes finding a good outer boundary condition difficult. While the use of an outgoing radiation boundary condition worked well in some situations, in general the value of $f$ at the outer boundary was simply held constant which sufficed whether or not a monopole was present. Reflection off this boundary was often significant so that the maximum radius was kept relatively large. This prevented reflections from affecting the dynamics near the origin. \section{Results} \label{sec:results} The field equations (\ref{eq:eom1}-\ref{eq:eom2}) are then integrated with a second-order accurate, Crank-Nicholson finite difference code. The generated solutions are found to be second-order convergent, stable, and mass conserving. Because either $\lambda$ or $\eta$ can be rescaled away, I choose either $\lambda=0.1$ or $\lambda=0$ for all that follows. As almost all other studies of Type~II critical behavior have, I have made use of an adaptive mesh. In particular, I incorporated finite difference approximations to the equations~(\ref{eq:eom1}-\ref{eq:eom2}) into the adaptive infrastructure developed by Choptuik~\cite{matt}. \subsection{A New DSS Critical Solution} \label{sec:newdss} As the Lagrangian presented in~(\ref{eq:lagrangian}) is quite general, I first discuss results obtained with vanishing scalar field $\psi$. Specifically, if $\psi(r,t)$ is initialized initially to zero with vanishing time derivative it will remain vanishing. This simplification allows the study of the collapse of the triplet field alone. I also postpone discussion of collapse within a monopole until Section~\ref{sec:inside}. Hence, in this section no monopole is present and $f(\infty,t)=0$. I find as the threshold of black hole formation is approached, a discretely self-similar solution is found with period $\Delta = 0.46$. A solution which approaches this precisely critical solution is shown in Fig.~\ref{fig:crit}. In the limit of precise criticality, the solution is expected to exhibit an infinite number of echoes and to be singular at the origin at the time of collapse, occurring at a finite proper time. The discrete nature of the self-similarity is demonstrated in Fig.~\ref{fig:echoing}. This solution also exhibits the usual mass scaling relationship in the super-critical regime, demonstrated in Fig.~\ref{fig:mass}. \begin{figure} \epsfxsize=8cm \centerline{\epsffile{crit.ps}} \caption{ Marginally sub-critical solution at late time for $\eta=0$. The top frame displays the hedgehog profile $f$, the middle frame shows the quantity $dm/dr$ explained in the text, and the last frame shows the value of $2m/r$. } \label{fig:crit} \end{figure} \begin{figure} \epsfxsize=9cm \centerline{\epsffile{echoing.ps}} \caption{ Demonstration of scale-periodicity of the new critical solution with $\Delta=0.46$. Here, the field $f$ is shown at four different times for a marginally sub-critical solution. The times and the value of $\Delta$ were chosen to minimize the difference $f(\ln(r),T_0)-f(\ln(r)+\Delta,T_1)$. Independently, the value of $\Delta$ was also computed by $\Delta = \ln((T^*-T_n)/(T^*-T_{n+1}))$ where $T$ represents the central proper time and $T^*$ the estimated critical time of collapse. The other two profiles are then shown shifted in $\ln(r)$ by $n\Delta$. The agreement demonstrates the scale periodicity. } \label{fig:echoing} \end{figure} \begin{figure} \epsfxsize=9cm \centerline{\epsffile{mass.ps}} \caption{ Demonstration of black hole mass scaling for $\eta=0$. The circles indicate the mass of black holes formed versus a normalized distance from criticality. The line indicates a least-squares fit for slope $\gamma=0.119$ where $M_{\rm BH} \propto |p-p^*|^\gamma$. } \label{fig:mass} \end{figure} Numerical results indicate that variations of $\eta$ in the interval $[0,0.5]$ do not alter the nature of the critical solution. In other words, the discrete nature remains with $\Delta=0.46$. The fact that the triplet field's value for large radius vanishes while the true vacuum lies at $\eta$ implies that the mass of the spacetime is divergent with increasing $r$. Hence, the profiles of $m'$ and $2m/r$ as shown in Fig.~\ref{fig:crit} are not the same as when $\eta=0$, differing for large $r$. However, the phenomena associated with approaching criticality occur for ever decreasing $r$, and in this regime, the solution remains discretely self-similar with period $\Delta=0.46$. This result is consistent with the potential not playing a role in the kinetic-energy dominated critical behavior. In the language of~\cite{carsten_rev}, the potential becomes ``asymptotically irrelevant.'' This divergence of the mass is not as problematic as it might initially appear because the numerical grid is artificially limited to some finite value of $r$. The mass is therefore limited, but this cut-off should not affect results for large $r$ because of the distance needed for any effect to travel from the outer boundary. By comparing otherwise identical evolutions with different maximum radius, no such effect was seen. It is interesting to note however that when $\eta \ne 0$ and $f(\infty,t)=0$, the value of $2m/r$ was seen to be divergent. This situation represents an infinite mass spacetime (essentially an infinity of false vacuum energy) and should have a horizon at some radius where $2m/r\rightarrow 1$. Hence, it would seem that the critical behavior observed here is occurring within this horizon. \subsection{Addition of a regular scalar field} \label{sec:addpsi} An interesting result found by Choptuik in his study of collapse of a Yang-Mills field, is that the presence of even an infinitesimal amount of scalar field added to the Einstein Yang-Mills model drives the solution to the scalar field's critical solution at criticality~\cite{matt_per}. Thus, the Yang-Mills field is unstable to the growth of the scalar field at criticality. To test for a similar effect here, I allow $\psi$ to be non-vanishing and look for the critical solution. Specifically, allowing the fields $f(r,t)$ and $\psi(r,t)$ to be arbitrary Gaussian pulses, I can ``mix'' the amplitudes with a parameter $\sigma$ such that for $\sigma=0$, the field $\psi$ vanishes and for $\sigma=1$, the field $f$ vanishes. \begin{figure} \epsfxsize=9cm \centerline{\epsffile{sigma.ps}} \caption{ Transition of critical solution from $\Delta=0.46$ to $\Delta=3.44$. A series of near-critical solutions are shown for $\sigma=0.4$ (solid), $\sigma=0.5$ (dotted), and $\sigma=0.6$ (dashed). The critical solution for small $\sigma$ is the new DSS, and as $\sigma$ increases, the critical solution becomes the original DSS. In the transition region, near-critical solutions appear to demonstrate structure from both critical solutions. } \label{fig:sigma} \end{figure} For the EYM case, any initial data for which $\sigma\ne0$ results in the scalar field critical solution being found~\cite{matt_per}. Here, no such sensitivity is observed. Instead, within some finite range of $\sigma$ the critical solution is observed to transition between the DSS of the scalar field ($\Delta=3.44$) and the new DSS presented here ($\Delta=0.46$). As $\sigma$ is varied across the transition from the original DSS to the new one, the fields $f$ and $\psi$ exchange dominance, both echoing at their respective value of $\Delta$. In the transition region, the near-critical solution, as represented by the field $m'(r,t)$, displays this dual echoing and hence has a non-trivial structure. This transition is shown in Fig.~\ref{fig:sigma}. \subsection{Critical Phenomena Inside a Monopole} \label{sec:inside} In contrast to the cases presented above in which $f(\infty,t)=0$, I consider initial data which have unit monopole charge as indicated by $f(r,t)$ asymptoting to $\eta$. Here, collapse within a monopole is modeled by first solving for the static monopole and then adding an in-going Gaussian perturbation to the profile $f(r,t)$. This solution has the benefit of not being dynamic near the outer boundary. The Gaussian pulse is then parameterized by an amplitude $p$, and a search is conducted in $p$ for a near critical solution. As stated, these evolutions should have the same critical solution as those lacking the potential. This is verified numerically for $\eta \in [0,0.2]$. For $\eta$ large ($\eta \agt 0.20$), no horizon-less, static solutions exist~\cite{steve_future} and hence the code is able to model these monopoles only within their horizons. A near-critical solution for $\eta=0.15$ is shown in Fig.~\ref{fig:mono}. \begin{figure} \epsfxsize=8cm \centerline{\epsffile{mono.ps}} \caption{ Near-critical solution obtained within a monopole for $\eta=0.15$. Notice that the field $f$ asymptotes to $\eta$ for large $r$ indicating the unit charge of the configuration. The discrete nature with $\Delta=0.46$ is observed to be the same as in the vanishing potential case. In comparison to the $\eta=0$ case (see Fig.~\ref{fig:crit}), in the scaling regime (small $r$) the echoing behavior is evident, and in the large $r$ regime the solutions are different. } \label{fig:mono} \end{figure} That static monopoles exist in this model naturally leads one to question whether any Type~I critical phenomena exists. However, evolutions appear to indicate that the static monopoles with no horizon are not unstable to collapse to a black hole. Hence, no such Type~I critical phenomena has been observed. \section{Conclusion} \label{sec:conclusion} A new critical solution is found by assuming a hedgehog ansatz for the triplet scalar field. The critical solution exhibits discrete self-similarity with echoing period $\Delta=0.46$. The solution also exhibits power-law mass scaling with exponent $\gamma=0.119$. Results verify that the potential does not affect the critical behavior as expected. Furthermore, the threshold of black hole formation is examined within a global monopole. Though the potential is expected to play no role in Type~II critical behavior, it does provide for the existence of static solutions, namely global monopoles. These static solutions then provide for the possibility that Type~I behavior is present. The static solutions would have to be unstable to black hole formation for this Type~I behavior to exist. However, evolutions do not show any instability of the monopole to collapse, and no Type~I critical behavior is seen. A linear perturbation analysis should be sufficient to settle the question of their stability. If the static solutions are indeed stable, then no Type~I behavior would be expected. Finally, as discussed in~\cite{liebling} symmetries of the initial data can determine which critical solution is attracting. For the spherically symmetric complex scalar field, in addition to Choptuik's original critical solution, another critical solution exhibiting a continuous self-similarity~(CSS) exists~\cite{hirschmann}. An arbitrary family of initial data will find Choptuik's solution, not the CSS. However, certain types of initial data which maximize the $U(1)$ charge density are unable to shed their charge density and find instead the CSS. Here, instead of a complex scalar field (i.e. a doublet field), the single scalar field is extended to a triplet field. Within the triplet scalar field model before making any assumptions on the triplet, at least two critical solutions exist, both DSS. The appropriate question is then will arbitrary families of initial data find Choptuik's original solution, the one presented here, or yet some other solution. To answer this with dynamical evolutions would require a general, three-dimensional collapse code. As a first step towards the goal of answering this question, Martin-Garcia and Gundlach~\cite{gundlach} have shown that non-spherical perturbations to Choptuik's DSS decay and hence that at least some families of non-symmetric initial data will find Choptuik's DSS. However this work considered only a single scalar field, and does not address the case where a triplet is present. The question of the relative stability of the original DSS to this new DSS essentially reduces to determining whether a family of generic, three-dimensional initial data for a triplet scalar field will shed its ``in-phase'' component to become the hedgehog DSS, or instead shed all its ``out-of-phase'' (i.e. non-zero monopole charge density) component to become the scalar field DSS. \section*{Acknowledgments} \label{sec:ack} I would like to thank Matthew Choptuik for providing me with his adaptive mesh framework. I am also thankful for the support of the Southampton College Research \& Awards Committee.
\section{Introduction} \setcounter{equation}{0} The shape of the universe is presently under investigation, and cosmic crystallography (CC) is one of the various methods proposed to determine it \cite{LeLaLu96}. The idea which supports the CC method is that if the universe is multiply connected then multiple images of a same cosmic object (a given quasar, say) may be observed in the sky. The separations between pairs of these images are correlated by the geometry and the topology of the spacetime; so if one selects a catalogue of observed images of cosmic objects and performs a histogram of the separations $l$ between the images, then these correlations manifest either as spikes (associated with Clifford translations) or as slight deformations of the histogram of the corresponding simply connected manifold \cite{spikes99}. A variant method was recently proposed by Fagundes and Gausmann ~\cite{FagundesGausmann98} , of subtracting from a histogram $ \phi (l_{i}) $ of a multiply connected space a histogram of the corresponding simply connected space. The reported result was a plot with much oscillations in small scales. In the present paper we propose an alternative to Fagundes-Gausmann method: we derive continuous probability density functions $ {\cal F}(l) $ to be subtracted from the histogram $ \phi (l_{i}) $ , and thereby obtain a histogram sensibly more suitable for analysis. In section 2 we derive the functions $ {\cal F}(a,l) $ for the Euclidean, hyperbolic and elliptic geometries, and in section 3 we make a few comments. \section{Probability densities} \setcounter{equation}{0} We consider one of the simply connected spaces $ E^3, H^3 $ or $ S^3 $. In that space, a spherical solid ball $ {\cal B}_a $ is taken , with radius $a$. The ball is assumed to contain an infinite number of pointlike objects, spatially distributed as uniformly as possible. We next select pseudorandomly two points of $ {\cal B}_{a} $ and ask for the probability $ {\cal F}(a,l)dl $ that the separation between the points lie between $l$ and $l + dl$. \vspace{5mm} {\bf a. Euclidean geometry} In the Euclidean three-space take a ball $ {\cal B}_{a} $ centred at the origin $ O $ and select two points $ P , Q $ in the ball; let $ r\in [0 ,a] $ be the radial position of $P$ and let $ l\leq 2a $ be the distance from $P$ to $Q$ (see Figure 1). Clearly the probability density $ {\cal F}_E(a,r,l)$ of this configuration is proportional both to the area $ {\cal S}_E(r)= 4\pi r^2 $ of the locus of $P$ well as to the area of the locus of $Q$. When $r + l < a$ this locus is a sphere $S^2$ with area ${\cal S}_E(l)$, while when $r+l>a$ the locus is a spherical disk $D^2$ in $E^3$ with area $D_E(a,r,l) = (\pi l/r)[a^2-(l-r)^2]$. The probability density of the configuration is then \begin{equation} \label{euc1} {\cal F}_E(a,r,l)=k{\cal S}_E(r)\{{\cal S}_E(l)\times\Theta (a-l-r)+D_E(a,r,l) \times\Theta (l+r-a)\}, \end{equation} where $k$ is a normalization constant and $\Theta$ is the Heaviside function. We integrate eq. (\ref{euc1}) for $r\in [0,a]$, and finally obtain, for $l\in (0, 2a]$, \begin{equation} \label{euc2} {\cal F}_E(a,l)=\frac{3 l^2}{16 a^6} (2a-l)^2 (l+4a), \end{equation} where the value $k=9/(16\pi ^2 a^6)$ was set to satisfy the normalization condition \begin{equation} \label{euc3} \int_{0}^{2a} {\cal F}_E(a,l)dl = 1. \end{equation} One often encounters in the literature reference to the probability density ${\cal P}(s)$ that the {\em squared} separation be $s$; since $s=l^2$ and ${\cal P}(s)ds = {\cal F}(l)dl$, then \begin{equation} \label{euc4} {\cal P}_E(a,s) = \frac{3\sqrt{s}}{32 a^6}(2a-\sqrt{s})^2 (\sqrt{s}+4a), \end{equation} valid for $s\in (0, 4a^2] $. In Figure 2 we reproduce a typical mean pair separation histogram (MPSH) for pseudorandomly distributed objects in an Euclidean solid sphere ${\cal B}_a$ with arbitrary radius, together with the corresponding probability density ${\cal F}_E(a,l)$. It should be stressed that, differently from the hyperbolic and elliptical cases, the shape of the function ${\cal F}_E(a,l)$ does not depend on the value of the radius $a$. \vspace{5mm} {\bf b. Hyperbolic geometry} To obtain the probability density ${\cal F}_H(a,l)$ for the hyperbolic geometry we follow the same lines as before. The area of a sphere with radius $r$ is now $S_H(r)=4\pi R^2 \sinh^2r/R$, where $R$ is the radius of curvature of the geometry; without loss of generality we henceforth set $R=1$. On the other hand, the area of a spherical disk $D^2$ in $H^3$ is (see Figure 1) \begin{equation} \label{hip1} D_H(a,r,l)= 2\pi \sinh\! l [\:\sinh\! l -\cosh\! l \coth\! r + \cosh\! a \: {\rm csch} r ], \end{equation} to be considered whenever $r+l>a$. The probability density of the configuration is then \begin{equation} \label{hip2} {\cal F}_H(a,r,l)=kS_H(r)[S_H(l)\times \Theta(a-l-r) + D_H(a,r,l)\times \Theta(l+r-a)], \end{equation} which upon integration for $r\in [0,a]$ and normalization gives \begin{equation} \label{hip3} {\cal F}_H(a,l)=\frac{8 \sinh^2 l}{(\sinh 2a -2a)^2}[\cosh\! a \:{\rm sech}(l/2)\sinh(a-l/2)-(a-l/2)], \end{equation} valid for $l\in (0,2a]$. In Figure 3 we reproduce graphs of ${\cal F}_H(a,l)$ for three values of the radius $a$. For $a<\!<1$ the function tends to the Euclidean one given in Figure 2, as expected. For increasing values of $a$ the function shifts towards the large values of $l$, and for $a>\!>1$ a strong concentration of ${\cal F}_H(a,l)$ is found near the extreme value $l=2a$. Figure 4 shows a typical MPSH in the hyperbolic three-space, together with the corresponding probability density ${\cal F}_H(a,l)$. \vspace{5mm} {\bf c. Elliptic geometry} The basic strategy to obtain the probability density ${\cal F}_S(a,l)$ is the same as before, and the calculations are similar whenever the diameter $2a$ of the ball is less than the separation $\pi R$ between antipodal points in the three-sphere; we then find for ${\cal F}_S(a,l)$ the expression, valid for $l\in (0,2a]$, \begin{equation} \label{eli1} {\cal F}_S(a<\pi/2,l)=\frac{8 \sin^2 l}{(2a-\sin 2a)^2}[(a-l/2)-\cos \!a\:\sec (l/2)\:\sin(a-l/2)], \end{equation} where we have taken $R=1$ without loss of generality. However, the cases where $a>\pi /2$ are considerably trickier to deal with, due to the treacherous connectivity of the spherical space $S^3$ and the new requirement that $l$ must not exceed $\pi$. A much larger quantity of trivial integrations now comes into scene, before the following expression is eventually obtained: \begin{eqnarray} {\cal F}_S(a,l) & = & \frac{8\sin^2 l}{[2a-\sin 2a]^2}\{[2a-\sin 2a-\pi] + \label{eli2} \\ & & \Theta(2\pi-2a-l)\times [\:\sin\! 2a+\pi-a-l/2-\cos\! a\: \sec(l/2)\:\sin(a-l/2)]\}\nonumber, \end{eqnarray} valid for all $a\in (0,\pi]$ and $l\in (0,{\rm min}(2a,\pi)]$. In Figure 5 four graphs of ${\cal F}_S(a,l)$ are shown, for different values of the radius $a$ of the ball. For $a$ increasing from $0$ to $\pi$ the function continuously shifts towards the smaller values of $l$. In particular, when $a=\pi/2$ we have \begin{equation} \label{eli3} {\cal F}_S(\pi/2,l)=\frac{4}{\pi}(1-l/\pi)\sin^2\! l\; , \end{equation} while when $a=\pi$ we have the harmonic, symmetric probability density \begin{equation} \label{eli4} {\cal F}_S(\pi,l)=\frac{2}{\pi}\sin^2\! l\; . \end{equation} In Figure 6 a sample MPSH in the spherical space is reproduced, together with the corresponding probability density ${\cal F}_S(a,l)$. \section{Comments} It is perhaps worth clarifying the meaning of the functions ${\cal F}(a,r,l)$: if we pseudorandomly choose two points $P$ and $Q$ in a solid sphere with radius $a$ (see Figure 1) then ${\cal F}(a,r,l)\:dr\:dl$ is the probability that $P$ lies between the radial positions $r$ and $r+dr$, times the probability that the separation from $Q$ to $P$ lies between the values $l$ and $l+dl$. The form of ${\cal F}(a,r,l)$ clearly depends on the geometry one is concerned with. Each histogram in Figures 2, 4, and 6 has $m=100$ subintervals and is a MPSH -- a mean pair separation histogram comprising $K=10$ comparable catalogues with $N=50$ objects each \cite{spikes99}. All computer-generated histograms assume a homogeneous distribution, as described by Lehoucq, Luminet and Uzan \cite{LeLuUz98}. A close inspection of Figure 2 shows that the most probable separation between two arbitrarily chosen points in an Euclidean solid ball is slightly greater than the radius of the ball; also the maximum of ${\cal P}_E(a,s)$ in eq.(\ref{euc4}) occurs when $s/4a^2=0.134$, in agreement with plots of Fagundes and Gausmann \cite{FagundesGausmann98b}. A characteristic feature of the hyperbolic geometries is that at large distances there is more space than in the Euclidean geometries; this fact is clearly exhibited in Figure 3, which shows a strong predominance of large separations $l$ when the radius $a$ of the solid sphere is large. It is worth noting that Fagundes and Gausmann \cite{FagundesGausmann98} obtained histograms with $a=2.34$ which closely resemble ours with $a=2$ in Figure 4. In contrast, the hyperbolic histograms given by Lehoucq {\em et al.} \cite{LeLuUz98} use radius $a$ nearly $0.6$, so they are similar to our Euclidean one. Oppositely to the hyperbolic case, in distant places endowed with the {\em elliptic} geometry there is {\em less} space than in the Euclidean geometry; this is nicely illustrated in Figure 5. Indeed, with increasing $a$ ( increasing solid ball) the probability of finding small distances $l$ (in comparison with $a$) also increases. We further note that when $a$ increases from $\pi/2$ to $\pi$ the ratio $l_{max}/2a$ recedes from $1$ to $1/2$.
\section{Introduction} In the old inflation scenario\cite{1}, it was assumed that the Universe underwent isentropic expansion during the stage of rapid growth of the scale factor. However, this scenario predicts an inhomogeneous Universe due to the fact the temperature is larger than the critical one $T_c \sim 10^{15}$ GeV. The chaotic inflation scenario describes a quasi - de Sitter expansion in a supercooled scenario. The entropy required to make the post-inflationary Universe consistent with observation is assumed to be generated in a short-time reheating period\cite{2}. However, the fluctuations of temperature should be very large and many domains should surpass the critical temperature $T_c$. This fact would lead to a very inhomogeneous Universe. Recently, Berera and Fang\cite{3} showed how thermal fluctuations may play the dominant role in producing the initial perturbations during inflation. They invoked slow - roll conditions. This ingenious idea was extended in some papers\cite{4} into the warm inflation scenario. This scenario served as an explicit demonstration that inflation can occur in presence of a thermal component. However, the radiation energy density $\rho_r$ must be small with respect to the matter energy density $\rho_{\varphi}$. More exactly, the kinetic component of the energy density ($\rho_{kin}$) must be small with respect to the vacuum energy density. This condition is satisfied if \begin{equation} \rho_{\varphi} \sim V(\varphi) \gg \rho_r \gg \rho_{kin}, \end{equation} where $V(\varphi)$ is the potential associated with the scalar field $\varphi$. A scenario of this kind provides a rapid expansion of the Universe in presence of a thermal component with small fluctuations of temperature compatible with the COBE data\cite{5}. The thermal equilibrium is reached near the minimum of the potential $V(\varphi)$. Particles are created during this expansion and it is not necesary to have a further reheating era (like in standard inflation). More recently\cite{YL}, J. Yokoyama and A. Linde showed that the solutions of warm inflation violate the adiabatic condition that the scalar field should not change significantly in the relaxation time of particles interacting with it. They claim that if the energy released by the interaction of the field with the created particles is small, then the total number of particles in the warm universe must be very small and their interaction with the scalar field may be too small to keep it from rapid falling down. According to the inflationary scenario, the inhomogeneities in the very early Universe are of genuine quantum origin\cite{6}. But, at end of inflation the Universe becomes classical. The quantum to classical transition is due to accelerated expansion of the Universe and loss of coherence of the quantum fluctuations. This last is produced by the increment of the degrees of freedom of the coarse - grained field which characterizes the infrared sector, jointly with the dissipation produced by the interaction of the inflaton with the thermal bath\cite{8,9,10}. In this paper, I extend the formalism of warm inflation recently introduced\cite{9}. Exact solutions for the power - law expansion of the Universe are calculated for this formalism. The paper is organized as follow: in section II, I develope the formalism. In section III the potential inflation model is studied, and finally in section IV, the final comments and conclusions are presented. \section{The Formalism} We consider a scalar field $\varphi$ coupled minimally to a classical gravitational one with a Lagrangian: \begin{equation} {\cal L}(\varphi,\varphi_{,\mu}) = - \sqrt{-g}\left[\frac{R}{16 \pi}+ \frac{1}{2} g^{\mu \nu} \varphi_{,\mu} \varphi_{,\nu}+V(\varphi) \right]+ {\cal L}_{int}, \end{equation} where $R$ is the scalar curvature, $g^{\mu \nu}$ the metric tensor, $V(\varphi)$ the scalar potential and $g$ is the metric. The Lagrangian ${\cal L}_{int}$ takes into account the interaction of $\varphi$ with other fields (i.e. bosons $X$,$Y$ or fermions) of the thermal bath. The spacetime will be considered as isotropic and homogeneous, and characterized by a Friedmann - Robertson - Walker (FRW) metric \begin{equation} ds^2 = - dt^2 + a^2(t) \ dr^2. \end{equation} The scale factor $a(t)$, is a time dependent function which grows with time. The equation of motion for the operator $\varphi$ is \begin{equation}\label{1} \ddot\varphi - \frac{1}{a^2} \nabla^2 \varphi + \left( 3 H(\varphi)+\tau(\varphi)\right) \dot\varphi+V'(\varphi) =0, \end{equation} where $H$ is the Hubble parameter (the dot denotes the time derivative), $\tau(\varphi)\dot\varphi $ describes the density energy dissipated by the field $\varphi$ into a thermalized bath, and the prime denotes the derivative with respect to $\varphi$. The imposibility to solve the equation (\ref{1}) becomes from our unknowledge of the Hilbert's space over which acts the field $\varphi$. For our proposal, will be sufficient to write the Friedmann equations in the semiclassical form \begin{equation} H^2 = \frac{8 \pi}{3}G \left< \rho_{\varphi}+\rho_{r}\right>, \end{equation} where $G=M^{-2}_p$ is the gravitational constant, $M_p$ is the Planckian mass ($M_p = 1.2 \ 10^{19}$ GeV), and $\rho_{r}$ and $ \rho_{\varphi}$ are the radiation and matter energy densities \begin{eqnarray} && \rho_{\varphi}= \frac{\dot\varphi^2}{2}+\frac{1}{a^2}\left( \vec{\nabla}\varphi\right)^2+ V(\varphi), \\ && \rho_{r} = \frac{\tau}{8 H} \dot\varphi^2. \end{eqnarray} In the conventional approach to the inflaton dynamics, the field $\varphi$ is split into a spatially homogeneous classical piece plus a spatially inhomogeneous quantum piece that represents quantum fluctuations of the field\footnote{These fluctuations will be consider as very small.} \begin{equation} \varphi(\vec{x},t)= \phi_c(t) + \phi(\vec{x},t). \end{equation} We require that $<E|\varphi|E>=\phi_c(t)$ and $<E|\phi|E>=0$, where $|E>$ is an arbitrary state. \subsection{Dynamics of the classical field $\phi_c$} We define the classical field as a solution of the equation \begin{equation} \ddot\phi_c+ (3 H_c + \tau_c) \dot\phi_c+ V'(\phi_c)=0. \end{equation} The Hubble parameter is expanded as: \begin{equation} H= H_c \left[ 1 + \frac{1}{2 H^2_c} \left< \left(1+ \frac{\tau_c}{4 H_c}\right) \dot\phi^2+ \frac{1}{a^2}(\vec{\nabla}\phi)^2 + \sum_n \frac{1}{n!} V^{(n+1)} (\phi_c) \ \phi^n \right> \right], \end{equation} where the classical Hubble parameter is \begin{equation} H^2_c = \frac{4 \pi}{3 M^2_p} \left[\left( 1+ \frac{\tau_c}{4 H_c}\right) \dot\phi^2_c+ 2 V(\phi_c)\right]. \end{equation} The classical dynamics of $\phi_c$ and $H_c$ are characterized by the following equations \begin{eqnarray} \dot\phi_c &=& - \frac{M^2_p}{4\pi} H'_c \left( 1 + \frac{\tau_c}{3 H_c}\right)^{-1}, \\ \dot H_c &=& H'_c \ \dot\phi_c = - \frac{M^2_p}{4\pi} (H'_c)^2 \left( 1+ \frac{\tau_c}{3 H_c}\right)^{-1}, \end{eqnarray} and the potential is \begin{equation} V(\phi_c)= \frac{3 M^2_p}{8\pi} \left[ H^2_c - \frac{M^2_p}{12\pi} (H'_c)^2 \left( 1+ \frac{\tau_c}{4 H_c}\right) \left( 1+ \frac{\tau_c}{3 H_c}\right)^{-2} \right], \end{equation} where we have assumed $H(\varphi)=H(\phi_c)\equiv H_c$ and $\tau(\varphi)=\tau(\phi_c)\equiv \tau_c$. The expression for radiation energy density is \begin{equation}\label{ro} \rho_r \simeq \frac{\tau_c}{8 H_c} \left(\frac{M^2_p}{4\pi}\right)^2 (H'_c)^2 \left( 1+ \frac{\tau_c}{3 H_c}\right)^{-2}. \end{equation} Near the minimum of the potential, the thermal equilibrium holds and the temperature is \begin{equation} < T_r > \sim \left( \frac{\tau_c(0)}{4 H(\phi_c=0)} \dot\phi^2_c\right)^{1/4}. \end{equation} In the inflationary epoch the kinetic energy is much smaller than the vacuum energy density $\rho_{\varphi}(\phi_c) \sim V(\phi_c)$. As in a previous work\cite{9}, I will consider the following relation between the Hubble parameter $H_c(\phi_c)=\dot a/a$ and the friction one $\tau_c(\phi_c)$: \begin{equation}\label{ga} \tau_c(\phi_c) = \gamma \ H_c(\phi_c). \end{equation} This expression takes into account that the particlelike are dispersed during the expansion of the Universe. The rate of expansion of the Universe is given by the Hubble parameter $H_c(\phi_c)$ and the increment of the rate of expansion lead to an increment of the rate of friction. The dimensionless constant $\gamma = \tau_c / H_c$ in the equation (\ref{ga}) is a parameter of the theory which takes into account the intensity of friction due to the interaction between the inflaton and the fields of the thermal bath. Note that in a the de Sitter expansion $H_c(\phi_c)$ is constant (i.e., $H_c(\phi_c)=H_o$), and thus $\tau_c$ and $\rho_r$ become zero ($\tau_c=\rho_r=0$). In this case one recovers the standard inflation model\cite{10}. The dynamics for the expansion and friction parameters is governed by the evolution of the classical field $\phi_c$. \subsection{Dynamics of the Quantum Fluctuations} The equation of motion for the operator $\phi$ is \begin{equation} \ddot\phi - \frac{1}{a^2} \nabla^2 \phi + (3 H_c+\tau_c)\dot\phi+ \sum_n \frac{1}{n!} V^{(n+1)}(\phi_c) \phi^n =0, \end{equation} where $V^{(n+1)}(\phi_c)$ denotes the $n+1$-th derivative. We assume that the quantum fluctuations are small and the Taylor expansion of $V'(\varphi)$ can be truncated at first order in $\phi$ \begin{equation} \ddot\phi - \frac{1}{a^2} \nabla^2 \phi + (3 H_c+\tau_c)\dot\phi+ V''(\phi_c) \phi =0. \end{equation} For our propose, will be useful to consider the redefined field $\chi = e^{3/2 \int (H_c+\tau_c/3) \ dt} \ \phi$. The equation of motion for this field is \begin{equation} \ddot\chi- a^{-2} \nabla^2 \chi - \frac{k^2_o}{a^2} \chi =0, \end{equation} where \begin{equation} k^2_o= a^2 \left[ \frac{9}{4} \left( H_c+\frac{\tau_c}{3}\right)^2 - V''(\phi_c) + \frac{3}{2}\left( \dot H_c + \frac{\dot\tau_c}{3}\right) \right], \end{equation} is the squared time dependent wavenumber. The field $\chi$ can be written as a Fourier expansion of the modes $\xi_k(t) \ e^{i \vec{k}.\vec{r}}$ \begin{equation} \chi = \frac{1}{(2\pi)^{3/2}} \int d^3 k \left[ a_k e^{i \vec{k}.\vec{r}} \xi_k(t) + h.c. \right], \end{equation} where $\xi_k(t)$ are the time dependent modes. The annihilation and creation operators $a_k$ and $a^{\dagger}_k$ satisfy the commutation relations \begin{eqnarray} && [a_k,a^{\dagger}_{k'}] = \delta (\vec{k} - \vec{k'}), \\ && [ a_k, a_{k'} ] = [a^{\dagger}_k,a^{\dagger}_{k'}] = 0. \end{eqnarray} The equation of motion for $\xi_k$ is \begin{equation} \ddot\xi_k(t)+ \omega^2_k \ \xi_k(t) =0, \end{equation} where $\omega^2_k = a^{-2}(k^2-k^2_o)$ is the squared frecuency of each mode with a given wavenumber $k$. The function $k_o$ separates both, the unstable $(k^2 \ll k^2_o)$ and the stable $(k^2 \gg k^2_o)$ sectors. The quantum fluctuations with wavenumbers below $k_o$ are interpreted as inhomogeneities superimposed on the classical field $\phi_c$. These fluctuations are responsible for the density inhomogeneities generated during the inflation. The quantum field theory imposes the commutation relation between $\chi$ and $\dot\chi$ \begin{equation}\label{2} [\chi(\vec{r},t),\dot\chi(\vec{r'},t)] = {\rm i} \delta^{(3)} \ (\vec{r}-\vec{r'}). \end{equation} The equation (\ref{2}) implies that \begin{equation} \xi_k \dot\xi^*_k - \dot\xi_k \xi^*_k = {\rm i}. \end{equation} To study the Universe on a scale greater than the scale of the observable Universe, we define the coarse - grained field that takes into account the long - wavelength modes. The wavelengths of such modes are \begin{equation} l \ge \frac{1}{\epsilon k_o}, \end{equation} where $\epsilon \ll 1$ is a dimensionless constant. The coarse - grained field, so defined, is \begin{equation} \chi_{cg}(\vec{r},t) = \frac{1}{(2\pi)^{3/2}} \int d^3 k \ \theta (\epsilon k_o -k) \left[ a_k e^{i\vec{k}.\vec{r}} \xi_k(t) + h.c. \right], \end{equation} which satisfies the following operatorial stochastic equation \begin{equation}\label{p1} \ddot \chi_{cg} - \frac{k^2_o}{a^2} \chi_{cg} = \epsilon \left(\frac{d}{dt}(\dot k_o \eta) + 2 \dot k_o \kappa\right), \end{equation} with the operatorial noises \begin{eqnarray} \eta(\vec{r},t) &=& \frac{1}{(2\pi)^{3/2}} \int d^3k \ \delta(\epsilon k_o - k) \quad \left[a_k e^{i\vec{k}.\vec{r}} \xi_k(t) + h.c \right], \\ \kappa(\vec{r},t) &=& \frac{1}{(2\pi)^{3/2}} \int d^3k \ \delta(\epsilon k_o - k) \quad\left[a_k e^{i\vec{k}.\vec{r}} \dot\xi_k(t) + h.c \right]. \end{eqnarray} The commutation relations are \begin{eqnarray} && [\chi_{cg}(t), \eta(t)] = 0, \label{c2}\\ && [\chi_{cg}(t), \kappa(t)] = \frac{2 \epsilon k_o}{(2\pi)^{3}} \int d^3 k \quad \theta(\epsilon k_o - k) \delta(\epsilon k_o - k) \quad \left( \xi_k \dot\xi^*_k - \xi^*_k \dot\xi_k\right), \label{c3}\\ && [\eta(t), \kappa(t)] = \frac{2 \epsilon k_o}{(2\pi)^{3}} \int d^3 k \quad \delta(\epsilon k_o - k) \delta(\epsilon k_o - k) \quad \left( \xi_k \dot\xi^*_k - \xi^*_k\dot\xi_k \right).\label{c4} \end{eqnarray} The correlations between the noises are \begin{eqnarray} &&<\kappa(t) \kappa(t') > =\frac{1}{(2\pi)^{3}} \int d^3 k \quad \delta(\epsilon k_o(t) - k) \quad \delta(\epsilon k_o(t') - k) \quad \left( \dot\xi_k(t) \dot\xi^*_k(t') \right),\label{a} \\ &&<\eta(t) \eta(t') > =\frac{1}{(2\pi)^{3}} \int d^3 k \quad \delta(\epsilon k_o(t) - k) \quad \delta(\epsilon k_o(t') - k) \quad\left( \xi_k(t) \xi^*_k(t') \right), \label{b} \\ && <\eta(t)\kappa(t')+\kappa(t) \eta(t') > = \frac{2}{(2\pi)^{3}} \int d^3 k \quad \delta(\epsilon k_o(t) - k) \quad \delta(\epsilon k_o(t') - k) \quad \left( \dot\xi_k(t) \xi^*_k(t') \right).\label{c} \end{eqnarray} Quantum to classical transition occurs due to the rapid expansion of the Universe during inflation. This transition is satisfied when the commutators (\ref{2}), (\ref{c2}), (\ref{c3}) and (\ref{c4}) are nearly zero. We can write the time dependent modes as \begin{equation} \xi_k(t)= u_k(t)+ {\rm i} \ v_k(t), \end{equation} and the condition to obtain the complex to real transition of these modes is \begin{equation} \left|\frac{v_k(t)}{u_k(t)}\right| \ll 1. \end{equation} We can define the quantum to classical transition function $\alpha_k(t)=\left|{v_k(t)\over u_k(t)}\right|$. The modes are real when this function becomes nearly zero. The condition for the coarse - grained field $\chi_{cg}(\vec{r},t)$ to be classical is \begin{equation} \frac{1}{N(t)}\sum_{k=0}^{k= \epsilon k_o} \alpha_k(t) \ll 1, \end{equation} where $N(t)$ is time dependent number of degrees of freedom of the infrared sector. During inflation, this number increases with time. This effect is due to the temporal evolution of both, the scale factor $a$ and the superhorizon (with size $l \sim {1 \over \epsilon k_o}$). Quantum to classical transition of the coarse-grained field holds when $\alpha_{k=\epsilon k_o}(t)\rightarrow 0$. Making ${\rm tan}[\Theta_k(t)]={v_k(t)\over u_k(t)}$, we can write the modes as \begin{equation} \xi_k(t)= e^{i \Theta_k(t)} \quad \left|\xi_k(t)\right|, \end{equation} where $\left|\xi_k(t)\right|= \sqrt{v^2_k+u^2_k}$. When the quantum fluctuations become classical, the correlations (\ref{a}), (\ref{b}) and (\ref{c}) are \begin{eqnarray} &&<\kappa(t) \kappa(t') > =\frac{1}{2\pi^{2}} \frac{\epsilon k^2_o}{\dot k_o} \xi_k(t)\xi_k(t') \quad\delta(t - t') \\ &&<\eta(t) \eta(t') > = \frac{1}{2\pi^{2}} \frac{\epsilon k^2_o}{\dot k_o} \dot\xi_k(t)\dot\xi_k(t') \quad\delta(t - t') \\ && <\eta(t)\kappa(t')+\kappa(t) \eta(t') > = \frac{1}{2\pi^{2}} \frac{\epsilon k^2_o}{\dot k_o} \left(\xi_k(t)\dot\xi_k(t')+\dot\xi_k(t)\xi_k(t')\right)\quad \delta(t - t'). \end{eqnarray} \subsection{Coarse-grained field correlations and power spectral density} Now we consider the correlation of $\chi_{cg}$ for different times $t$ and $t'$ ($t'> t$), once the modes are classical [$\xi_k(t)\xi^*_k(t') \simeq \xi_k(t)\xi_k(t')$] \begin{equation} < \chi_{cg}(t) \chi_{cg}(t')> = \frac{1}{(2\pi)^3} \int^{\epsilon k_o(t')}_{\epsilon k_o(t)} d^3 k \quad \xi_k(t)\xi_k(t'). \end{equation} The Fourier transform of $< \chi_{cg}(t) \chi_{cg}(t')>$ gives the power spectral density for $\chi_{cg}$ in the infrared sector \begin{equation} S[\chi_{cg};\omega_k]= 4 \int^{\infty}_{0} cos[\omega t''] \quad \left|< \chi_{cg}(t) \chi_{cg}(t+t'')>\right|_{t\gg 1} \quad dt'', \end{equation} where $t''=t'-t$. This expression, written explicitly, becomes \begin{equation}\label{S} S[\chi_{cg};\omega_k]=\frac{1}{\pi}\int^{\infty}_{0} dt''\quad cos[\omega t''] \quad \int^{\epsilon k_o(t+t'')}_{\epsilon k_o(t)} dk \ k^2 \ \xi_k(t)\xi_k(t+t''), \end{equation} where $\omega_k$ is the oscillation frecuency for a given wavenumber $k$. The equation (\ref{S}) is only valid when the thermal equilibrium holds. The fluctuations of radiation energy density are \begin{equation} \frac{\delta\rho_r}{\rho_r} \simeq \left| 2(H'_c)^{-1} \ H''_c \right| \quad <\phi^2_{cg}>^{1/2}, \end{equation} where $\phi_{cg}(\vec{r},t)=a^{-1/2(3+\gamma)} \ \chi_{cg}(\vec{r},t)$. The condition for the thermal equilibrium holds is ${d\over dt}\left({\delta\rho_r \over \rho_r}\right)< 0$. \section{Power - law Inflation} This model of inflation is characterized by $a(t)$ and $H_c$ given by \begin{eqnarray} && a(t)= H^{-1}_o \ \left(\frac{t}{t_o}\right)^p, \\ && H_c(t)= \frac{p}{t}. \end{eqnarray} The temporal evolution of the classical field is \begin{equation} \phi_c(t)= \phi_o- m {\rm ln} \ \left[\frac{H_o}{p}t\right]. \end{equation} The solution of the equation (\ref{ro}) for the radiation energy density is \begin{equation} \rho_r = \frac{\gamma M^4_p \ p^2}{128 \pi^2 m^2 (1+\gamma/3)^2} \ t^{-2}. \end{equation} The radiation temperature is \begin{equation} <T_r> \propto M_p m^{-1/2} \ t^{-1/2}, \end{equation} where $m$ is the mass of the scalar field. The potential for this model is \begin{eqnarray} && V(\phi_c)= \frac{3 M^2_p}{8 \pi} H^2_c \left\{ 1- \frac{M^2_p}{48 \pi} m^{-2} \left[\left( 1+\gamma /4 \right)\left( 1+\gamma/3\right)^{-2} \right] \right\}. \end{eqnarray} The equation of motion for the modes is \begin{equation}\label{mo} \ddot \xi_k(t) + \left[\frac{H^2_o t^{2p}_o k^2}{t^{2p}}-\mu^2(t)\right] \quad \xi_k(t)=0, \end{equation} with \begin{equation} \mu^2(t)=t^{-2} K^2, \end{equation} and \begin{equation}\label{k} K^2=\left[ \frac{9}{4} p^2(1+\gamma/3)^2- \frac{3}{2}p(1+\gamma/3)+\frac{3 p^2M^2_p}{2\pi m^2} \left(1- \frac{M^2_p}{48 \pi^2 m^2} (1+\gamma/4)(1+\gamma/3)^{-2}\right) \right]. \end{equation} For inflation takes place, we must take $k^2_o>0$, and the expansion must be sufficiently rapid such that $p> {3 \over 2 A} (1+\gamma/3)$, with $A=9/4(1+\gamma/3)^2+ {3 M^2_p \over 2 \pi m^2} \left( 1 - {M^2_p \over 48 \pi m^2} (1+\gamma/4)(1+\gamma/3)^{-2}\right)$. The general solution for the equation (\ref{mo}) is \begin{equation} \xi_k(t)= A_1 \sqrt{t} \ H^{(1)}_{\nu}(t)\left[\frac{H_o k (t/t_o)^{1-p}}{p-1}\right] + A_2 \sqrt{t} \ H^{(2)}_{\nu}(t)\left[\frac{H_o k (t/t_o)^{1-p}}{p-1}\right], \end{equation} with $\nu = {1\over 2(p-1)} \sqrt{1+4 K^2}$. When $\gamma=0$ (i.e., for $\tau_c=0$), we recover the results of standard inflation\cite{10}. We choose the Bunch - Davis vacuum ($A_1=0$), but with imaginary constant $A_2={\rm i} |A_2|$ that can be written in terms of the Bessel functions \begin{equation} \xi_k(t)= \frac{\sqrt{\pi}}{2}\frac{\sqrt{t/t_o}}{p-1} \left[{\cal Y}_{\nu}[x(t)]+{\rm i} {\cal J}_{\nu}[x(t)]\right] ={\rm i} \ H^{(2)}_{\nu}[x(t)], \end{equation} where $H^{(2)}_{\nu}[x(t)]$ is the Hankel function and $x(t)= {H_o k \left(t/t_o\right)^{1-p}\over p-1}$. Observe that $x(t)$ tends to zero when $t$ is sufficiently large. In this case one obtains $x(t)\ll 1$, and the asymptotic modes are real \begin{equation}\label{xi} \xi_k(t) \simeq \frac{\sqrt{\pi}}{2} \frac{\sqrt{t/t_o}}{p-1} {\cal Y_{\nu}}\left[ x(t)\right], \end{equation} which is due to \begin{equation} \left|\frac{{\cal J}_{\nu}[x(t)]}{{\cal Y}_{\nu}[x(t)]}\right| \ll 1. \end{equation} So, for $\left(t/t_o\right)^{p-1}\gg \frac{H_o k}{p-1}$ the modes with the wavenumber $k$ are real. For $x(t)\ll 1$ the function ${\cal Y}_{\nu}$ becomes \begin{equation} {\cal Y}_{\nu} \simeq - \frac{{\rm i}}{\pi} \Gamma(\nu) \left( \frac{H_o k (t/t_o)^{1-p}}{2 (p-1)}\right)^{-\nu}, \end{equation} where $\Gamma(\nu)$ is the gamma function with argument $\nu$. The correlation of the coarse - grained field, for $t\gg 1$, is \begin{equation}\label{corr} \left.<\phi_{cg}(t=t_o) \phi_{cg}(t_1)> \right|_{p\gg 1} \simeq \frac{2\pi 4^{\nu} \Gamma^2(\nu)}{(p-1)^4} \left[ \epsilon K \right]^{2(1-\nu)+1} \ \left(\frac{t_1}{t_o}\right)^{4\nu p(1-\nu/p-p/\nu)+3-\gamma}, \end{equation} where $K$ is given by the equation (\ref{k}). The power spectral density of the fluctuations (when the thermal equilibrium holds) is the Fourier transform of the equation (\ref{corr}), which becomes [see eqs. (\ref{S}) and (\ref{xi})] \begin{eqnarray} S[\chi_{cg}; \omega_k=\frac{k H_o }{(t_1/t_o)^p}] & \simeq & \frac{8\pi 4^{\nu} \Gamma^2(\nu)}{(p-1)^4} \left[ \epsilon K \right]^{2(1-\nu)+1} \cos\left[\frac{1}{2}\pi\left(4\nu(p-1)-2p+4\right)\right]\nonumber \\ &\times & \Gamma\left[\left(4\nu(p-1)-2p+4\right)\right] \left|\omega_k\right|^{-\left( 4\nu (p-1)-2p+4 \right) }. \end{eqnarray} For $p \gg 1$ one obtains $S[\chi_{cg}, \omega_k] \propto \left|\omega_k\right|^n$, with $n= -(4\nu(p-1)-2p+4)$. The fluctuations of the radiation energy density are \begin{equation} \left.\frac{\delta\rho_r}{\rho_r}\right|_{t\gg 1} \sim t^{2\nu p(1-\nu/p-\frac{5}{4}p/\nu) +\frac{1}{2}-\frac{\gamma}{2}}. \end{equation} Note that both, $<\phi_{cg}(t_o) \phi_{cg}(t_1)> $ and ${\delta\rho_r\over\rho_r}$ decreases with time for sufficiently large $p$. \section{Conclusions} In this work, I extended the formalism for warm inflation introduced in a previous paper. In the model the rapid expansion of the Universe is produced with a temperature $T_r$ smaller than $T_c$. The dynamics of the fluctuations of the matter field generates thermal fluctuations which decrease with time. So, at the end of inflation, the thermal equilibrium holds. On the other hand, the classical field $\phi_c(t)$ is responsible for the expansion of the Universe and the dissipation of energy. The dissipation of energy is due to the interaction between the inflaton and the particles in the thermal bath (with mean temperature $<T_r> \ < T_c$). The dynamics of the interaction is represented by the classical parameter $\tau_c(\phi_c)$. This parameter is considered as proportional to the rate of expansion of the Universe, which is given by the Hubble parameter $H_c(\phi_c)$. This model has four foundamental aspects: {\bf 1)} The classical one: characterized by the size of the superhorizon $l(t)\sim {1\over \epsilon k_o}$. The quantum to classical transition of the fluctuations in the infrared sector is due to the complex to real transition of the modes $\xi_k$ of $\chi_{cg}$. The dynamics of this transition is described by the sum over $k$ in the infrared sector, for the function $\alpha_k(t)$. {\bf 2)} The stochastic ingredient of the coarse - grained field: the fluctuations of the matter field $\chi_{cg}$ describes the infrared sector and and its evolution is governed by a classical stochastic equation. {\bf 3)} The thermal ingredient: the dissipation parameter $\tau_c$ represents the interaction of the inflaton with a thermal bath. In this model, when the inflation starts the Universe is with a mean temperature smaller than the GUT's temperature ($T_c \sim 10^{15}$ GeV). In the example for power - law inflation here studied, the mean temperature decreases as $t^{-1/2}$ and the thermal fluctuations also decrease for sufficiently large $p$. So, at the end of inflation the Universe is more cold than when inflation starts. The thermal equilibrium holds due to the decreasing with time of the fluctuations for the radiation energy density. Due to this fact, it is possible to calculate the power spectral density for $\chi_{cg}$. This spectrum depends on the rate of expansion of the Universe and the interaction of the inflaton with the thermal bath. {\bf 4)} The quantum aspect: relies in the time dependent modes $\xi_k(t)$ of the coarse - grained field $\chi_{cg}$. These modes must be real in the infrared sector to give classical fluctuations. Initially (i.e., when the inflation starts), the fluctuations $\phi(\vec x, t)$ [and also $\chi(\vec x,t)$] are quantized, since the energy density of the false vacuum is of the order (but less) of the Planckian scale ($\rho_{\varphi} \sim V(\varphi) < M^4_p$). Furthermore the amplification of the modes that enters in the infrared sector generates loss of coherence of the field $\chi_{cg}$. More rapid is the expansion of the Universe, more rapid is the complex to real transition of each mode with a wavenumber $k$. The modes $\xi_k$ with smaller wavenumber will become real before the bigger ones. So, the Universe becomes classical in the infrared sector. \vskip 1cm \centerline{\bf Acknowlekgements} \vskip 1cm I thank Adnan Bashir for your careful reading of the manuscript. \vskip 1cm
\section{#1}} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}} \def\lap{\Delta} \def\lambda{\lambda} \def\begin{equation}{\begin{equation}} \def\end{equation}{\end{equation}} \def\begin{eqnarray}{\begin{eqnarray}} \def\end{eqnarray}{\end{eqnarray}} \def\begin{quote}{\begin{quote}} \def\end{quote}{\end{quote}} \def\noindent{\noindent} \def{ \it Phys. Lett.} {{ \it Phys. Lett.} } \def{\it Phys. Rev. Lett.} {{\it Phys. Rev. Lett.} } \def{\it Nucl. Phys.} {{\it Nucl. Phys.} } \def{\it Phys. Rev.} {{\it Phys. Rev.} } \def{\it Mod. Phys. Lett.} {{\it Mod. Phys. Lett.} } \def{\it Int. J. Mod .Phys.} {{\it Int. J. Mod .Phys.} } \def{\it Class. Quant. Grav.} {{\it Class. Quant. Grav.} } \def{\it Phys. Rep.} {{\it Phys. Rep.} } \newcommand{\labell}[1]{\label{#1}\qquad_{#1}} \newcommand{\labels}[1]{\vskip-2ex$_{#1}$\label{#1}} \parskip 0.3cm \begin{document} \thispagestyle{empty} \begin{flushright} Stanford SU-ITP-99-17 \\ April 1999 \end{flushright} \vspace*{1cm} \begin{center} {\Large \bf Vacuum polarization in Schwarzschild space-time}\\ {\Large \bf by anomaly induced effective actions }\\ \vspace*{2cm} R. Balbinot, \footnote{E-mail: <EMAIL>}\\ \vspace*{0.2cm} {\it Dipartimento di Fisica dell' Universit\`a di Bologna}\\ {\it and INFN sezione di Bologna}\\ {\it Via Irnerio 46, 40126 Bologna, Italy }\\ \vspace*{0.5cm} A. Fabbri, \footnote{E-mail: <EMAIL>}\\ \vspace*{0.2cm} {\it Department of Physics, Stanford University}\\ {\it Stanford, CA 94305-4060, USA}\\ \vspace*{0.5cm} I. Shapiro \footnote{E-mail: <EMAIL> }\\ \vspace*{0.2cm} {\it Departamento de Fisica, ICE, Universidade Federal de Juiz de Fora}\\ {\it 36036-330, Juiz de Fora -- MG, Brazil } \\ \vspace*{0.2cm} {\it Tomsk Pedagogical University, Tomsk, Russia} \vspace*{2cm} { \bf Abstract} \end{center} The characteristic features of $\langle T_{\mu\nu}\rangle$ in the Boulware, Unruh and Hartle-Hawking states for a conformal massless scalar field propagating in the Schwarzschild space-time are obtained by means of effective actions deduced by the trace anomaly. The actions are made local by the introduction of auxiliary fields and boundary conditions are carefully imposed on them in order to select the different quantum states. \vfill \setcounter{page}{0} \setcounter{footnote}{0} \newpage \sect{Introduction} One of the most interesting aspects of quantum matter fields propagating in curved space-time \cite{birdav} is the existence of a trace anomaly discovered by Capper and Duff \cite{duff} in 1974 and its deep connection with Hawking's black hole evaporation \cite{Hawking}, \cite{duff-77}, \cite{chfu} . \\ At the classical level if the action describing the matter fields is conformally invariant the trace $T$ of the corresponding stress energy tensor vanishes. However, when the fields are quantized as a result of the renormalization procedure the expectation value of the trace $\langle T \rangle$ differs from zero when space-time dimension is even. We then have an ``anomaly''. This anomaly depends on geometrical quantities only: it is expressed in terms of the Riemann tensor of the space-time and its contractions. Furthemore being an ultraviolet effect, the anomaly is general in the sense that its value does not depend on the quantum state in which the expectation value is taken. In four dimensions (4D) we have \begin{equation} \langle T^{\alpha}_{\alpha} \rangle \equiv \langle T \rangle = -\frac{1}{(4\pi)^2}(aC^2 +bE +c\Box R ) \label{trac} \end{equation} where $C^2\equiv C_{\alpha\beta\gamma\delta}C^{\alpha\beta\gamma\delta} $ is the square of the Weyl tensor and $E$ is an integrand of the Gauss-Bonnet topological term \begin{equation} E\equiv R_{\mu\nu\alpha\beta}R^{\mu\nu\alpha\beta} -4R_{\alpha\beta}R^{\alpha\beta} +R^2 \label{gabo} \end{equation} and $a,b,c$ are known numerical coefficients that depend on the spin of the matter fields considered \cite{birdav}. \\ In 2D we have the much simpler form \begin{equation} \langle T^{\alpha}_{\alpha}\rangle \equiv \langle T\rangle = g R \label{tradu} \end{equation} where as before $g$ depends on the matter species. \\ The knowledge of the trace anomaly allows some information on the complete quantum stress tensor expectation values $\langle T_{\mu\nu}\rangle$ to be extracted. \\ In 2D the conservation equations $\nabla^a \langle T_a^{\ b} \rangle =0$ and the trace anomaly, eq. (\ref{tradu}), determine almost completely the stress tensor. Parametrizing the 2D spacetime metric $g_{ab}(x)$ as \begin{equation} ds^2 =-e^{2\rho}dx^+dx^- \label{tuco} \end{equation} in the conformal gauge $\{ x^+, x^- \}$ one finds \cite{davies} \begin{eqnarray} \langle x^{\pm} |T_{\pm\pm}|x^{\pm}\rangle &=& - \frac{1}{12\pi} (\partial_{\pm}\rho\partial_{\pm}\rho - \partial_{\pm}^2\rho) + t_{\pm}(x^{\pm}), \nonumber \\ \langle x^{\pm}|T_{+-}|x^{\pm}\rangle &=& -\frac{1}{12\pi} \partial_+\partial_-\rho \ , \label{renst} \end{eqnarray} where $t_{\pm}(x^{\pm})$ are arbitrary functions of their arguments which depend on the choice of quantum state in which the expectation values are taken. They represent conserved radiation which, being traceless, cannot be determined by the trace anomaly alone. These functions are fixed by boundary conditions. The above expressions for $\langle T_{ab}\rangle$ can also be obtained by rigorous canonical quantization and point splitting regularization of the divergences \cite{dafu}. \\ In 4D the situation is much more fuzzy since the four conservations equations and the trace anomaly eq. (\ref{trac}) are clearly insufficient to determine the ten components of $\langle T_{\mu\nu}\rangle$. Things improve if the background space-time on which the quantum matter fields propagate enjoys some symmetries which reduce the number of independent components of $\langle T_{\mu\nu}\rangle$ \cite{chfu}. \\ Unfortunately in 4D it appears impossible (at least for the moment) to find a nice formula analogous to eqs. (\ref{renst}) giving $\langle T_{\mu\nu}\rangle $ as a function of a general metric $g_{\alpha\beta}$. Such an expression would allow us the study of the evolution of the background geometry driven by the quantum fluctuation of the matter fields propagating on it. This is the so called backreaction, governed by the semiclassical Einstein equations \begin{equation} G_{\mu\nu} (g_{\alpha\beta})=8\pi \langle T_{\mu\nu}(g_{\alpha\beta})\rangle \ , \label{i} \end{equation} where $G_{\mu\nu}$ is the Einstein's tensor. \\ In principle $\langle T_{\mu\nu}(g_{\alpha\beta})\rangle$ can be derived from an effective action $S_{eff}(g_{\alpha\beta})$ describing the quantum matter fields propagating on $g_{\alpha\beta}$. Although $S_{eff}(g_{\alpha\beta})$ is unknown, the trace anomaly $\langle T\rangle$ can be used to reconstruct at least a part of it, i.e. the so called ``anomaly induced effective action''. $S_{eff}^{an}(g_{\alpha\beta})$ is then defined as following \begin{equation} \frac{2}{\sqrt{-g}}g^{\mu\nu}\frac{\delta S_{eff}^{an}}{\delta g^{\mu\nu}} =\langle T\rangle \ . \label{llll} \end{equation} By functional integration of this equation one obtains the explicit form of $S_{eff}^{an}$. \\ The complete effective action $S_{eff}(g_{\alpha\beta})$ can then be written as \begin{equation} S_{eff}(g_{\alpha\beta})=S_{eff}^{an}(g_{\alpha\beta})+ S^W (g_{\alpha\beta})\ , \label{coac} \end{equation} where $S^W$ is a conformal invariant functional which can be regarded as an integration constant of eq. (\ref{llll}). As expected, the exact form of $S^W$ is not known. \\ Taking the functional derivation of $S_{eff}^{an}(g_{\alpha\beta})$ with respect to $g_{\alpha\beta}$ we define the ``anomaly induced stress tensor'' $S_{\mu\nu}$ as follows \begin{equation} \frac{2}{\sqrt{-g}}\frac{\delta S_{eff}^{an}}{\delta g^{\mu\nu}} =\langle S_{\mu\nu} \rangle \label{stan} \end{equation} where of course $\langle S\rangle = \langle T\rangle$. It is tempting, given our ignorance on $\langle T_{\mu\nu}\rangle$, to use $\langle S_{\mu\nu} \rangle$ in the semiclassical Einstein equations (\ref{i}) to get at least a feeling on the backreaction problem. \\ The basic question is: how close $\langle S_{\mu\nu}\rangle$ is to $\langle T_{\mu\nu}\rangle $? In other words, can we ignore the Weyl invariant part $S^W$ to get correctly at least the qualitative features of the backreaction? It is hard to find answers to the above questions since we do not know in general $\langle T_{\mu\nu}(g_{\alpha\beta})\rangle$ . However, for some particular background, say $\bar g_{\alpha\beta}$, $\langle T_{\mu\nu} (\bar g_{\alpha\beta})\rangle$ is known and hence a check of the above conjecture can be given by direct comparison of $\langle T_{\mu\nu}(\bar g_{\alpha\beta})\rangle$ and the corresponding $\langle S_{\mu\nu} (\bar g_{\alpha\beta})\rangle$. \\ The background $\bar g_{\alpha\beta}$ we have in mind is the Schwarzschild black hole geometry \begin{equation} ds^2=-(1-2M/r)dt^2 + (1-2M/r)^{-1}dr^2 +r^2 (d\theta^2 +\sin^2\theta d\varphi^2) \label{sci} \end{equation} or \begin{equation} ds^2 =-(1-2M/r) dudv + r^2 d\Omega^2\ , \label{edfi} \end{equation} where \begin{eqnarray} u &=& t-r-2M\ln |r/2M -1|, \nonumber \\ v &=& t+r+2M\ln|r/2M -1| \label{coed} \end{eqnarray} and $d\Omega^2$ is the metric for the unit two-sphere. This, because of its physical implication, most notably Hawking black hole evaporation, is one of the most studied space-times in quantum field theory in curved space. \\ Detailed analytical and numerical investigations , by means of modes sum and point-splitting regularization or by axiomatic principles, of $\langle T_{\mu\nu}\rangle$ in this background have been performed \cite{molti}. We are rather confident on these results on which, we stress, our understanding of all quantum black hole physics is based. \\ In the Schwarzschild space-time three different quantum states (``vacua'') are defined by expanding field operators in suitable basis: \\ \\ \noindent i) The Boulware vacuum $|B\rangle$ \cite{bula}. \\ This is obtained by choosing {\it in} and {\it out} modes to be positive frequency with respect to the Killing vector $\partial_t$ of the metric (\ref{sci}). This state most closely reproduces the familiar notion of Minkowski vacuum $|M\rangle$ asymptotically. One finds, as $r \to \infty$, $\langle B|T_{\mu\nu}|B\rangle \to 0 $ sufficiently rapidly, i.e. $O(r^{-6})$. Unfortunately, the behaviour on the horizon $r=2M$ is rather pathological, being $\langle B|T_{\mu\nu}|B\rangle$ divergent there in a free falling frame. \\ This state is appropriate for the description of vacuum polarization around a static star whose radius is necessary bigger than the horizon. \\ \\\ \noindent ii) The Unruh vacuum $|U\rangle$ \cite{uuru}. \\ {\it in} modes are chosen as before to be positive frequency with respect to $\partial_t$. This ensures that asymptotically in the past $|U\rangle \sim |M\rangle$. On the other hand, {\it out} modes are taken to be positive frequency with respect to Kruskal's $U=-4Me^{-u/4M}$, the affine parameter along the past horizon. This mimics the late time behaviour of modes coming out of a collapsing star as its surface approaches the horizon. By this choice $\langle U|T_{\mu\nu}|U\rangle$ is regular on the future (but not on the past) event horizon and most remarkably asymptotically in the future $\langle U|T_{\mu\nu}|U\rangle$ has the form of a flux of radiation at the Hawking temperature $T_H=\frac{1}{8\pi M}$. \\ This state is the most appropriate to discuss evaporation of black holes formed by gravitational collapse of matter. In that case the divergence on the past horizon is spurious. \\ \\ \noindent iii) The Israel-Hartle-Hawking state $|H\rangle$ \cite{ishh}. \\ This state is obtained by choosing {\it in} modes to be positive frequency with respect to Kruskal's $V=4Me^{v/4M}$, the affine parameter on the future horizon, whereas outgoing modes are positive frequency with respect to $U$. By this choice asymptotically both in the future and the past $|H\rangle \neq |M\rangle$. \\ $\langle H|T_{\mu\nu}|H\rangle$ for $r\to\infty$ describes in fact a thermal bath of radiation at the Hawking temperature $T_H$. By construction this state is regular on both the future and the past horizon. $|H\rangle$ is used to describe a black hole in thermal equilibrium with a surrounding bath. \\ \\ We summarize here, for later comparison, the analytic expression of the stress tensor for a conformally invariant scalar field in the three vacua defined above. \\ In 2D Schwarzschild spacetime (neglect the angular part in eq. (\ref{edfi}) ) we have \cite{full} \begin{eqnarray} \langle B |T_{tt}|B\rangle &=& \frac{1}{12\pi} \left[ -\frac{2M}{r^3} +\frac{7 M^2}{2 r^4}\right] , \nonumber \\ \langle B |T_{rr}|B\rangle &=& -\frac{1}{48\pi}\frac{2M^2}{r^4} (1-2M/r)^{-2}\ , \nonumber \\ \langle B |T_{rt}|B\rangle &=& 0 \label{boli} \end{eqnarray} and \begin{eqnarray} \langle U |T_{tt}|U\rangle &=& \frac{1}{12\pi} \left[ -\frac{2M}{r^3} +\frac{7M^2}{2r^4}\right] + (768\pi M^2)^{-1} , \nonumber \\ \langle U |T_{rr}|U\rangle &=& -\frac{1}{48\pi} (1-2M/r)^{-2}\frac{2M^2}{r^4} + (768\pi M^2)^{-1} , \nonumber \\ \langle U |T_{rt}|U\rangle &=& -(1-2M/r)^{-1}(768\pi M^2)^{-1} \label{unli} \end{eqnarray} and finally \begin{eqnarray} \langle H |T_{tt}|H\rangle &=& \frac{1}{12\pi} \left[ -\frac{2M}{r^3} +\frac{7 M^2}{2 r^4}\right] + (384\pi M^2)^{-1} , \nonumber \\ \langle H |T_{rr}|H\rangle &=& -\frac{1}{48\pi}\frac{2M^2}{r^4} (1-2M/r)^{-2} + (384\pi M^2)^{-1} , \nonumber \\ \langle H |T_{rt}|H\rangle &=& 0 \ . \label{hhli} \end{eqnarray} Note that $\langle B|T^a_a|B\rangle=\langle U|T^a_a|U\rangle= \langle H|T^a_a|H\rangle$ as a consequence of the state independence of the trace anomaly. \\ In 4D only approximate expressions of $\langle T_{\mu\nu}\rangle$ can be given \cite{molti}. Here we just report the leading behaviour at infinity and on the horizon. For the Boulware vacuum $|B\rangle$ in Schwarzschild coordinates $(t,r,\theta,\varphi)$ \begin{equation} \langle B| T_{\mu}^{\ \nu}|B\rangle \to O(r^{-6}) \ ,\ r\to\infty\ , \label{boin} \end{equation} \begin{equation} \langle B| T_{\mu}^{\ \nu}|B\rangle \sim - \frac{1}{30\ 2^{12}\pi^2 M^4f^2} \pmatrix{ -1 & 0 & 0 & 0 \cr 0 & 1/3 & 0 & 0 \cr 0 & 0 & 1/3 & 0 \cr 0 & 0 & 0 & 1/3 \cr}\ , \ r\to 2M\ , \label{boho} \end{equation} where $f\equiv 1-2M/r$. \\ For the Unruh vacuum $|U\rangle$ \begin{equation} \langle U| T^{\mu}_{\nu} |U\rangle \to \frac{L}{4\pi r^2} \pmatrix{ -1 & -1 & 0 & 0 \cr 1 & 1 & 0 & 0 \cr 0 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 \cr},\ r\to\infty \ , \label {unin} \end{equation} describing an outgoing flux, and \begin{equation} \langle U| T_{a}^{\ b} |U\rangle \sim \frac{L}{4\pi (2M)^2} \pmatrix{ 1/f & -1 \cr 1/f^2 & -1/f \cr}, \ r\to 2M\ , \label {unho} \end{equation} $a,b=r,t$, which describes a negative energy flux of radiation going down the hole compensating the escaping flux at infinity. $L$ is the luminosity. In geometric optics approximation $L= \frac{2.197\ 10^{-4}}{\pi M^2} $ \cite{dewitt}. Numerical extimates set $L=\frac{2.337\ 10^{-4}}{\pi M^2} $ \cite{elster}. \\ Finally for the $|H\rangle$ state \begin{equation} \langle H| T^{\mu}_{\nu} |H\rangle \to \frac{1}{30\ 2^{12}\pi^2 M^4} \pmatrix{ -1 & 0 & 0 & 0 \cr 0 & 1/3 & 0 & 0 \cr 0 & 0 & 1/3 & 0 \cr 0 & 0 & 0 & 1/3 \cr}\ , \ r\to \infty\ , \label{hhin} \end{equation} describing a thermal bath at the temperature $T_H$. Furthemore $\langle H|T_{\mu}^{\ \nu}|H\rangle$ is regular on the horizons . \\ \\ \noindent In this paper, starting from the trace anomaly, we shall consider $S_{eff}^{an}$ for a massless scalar field propagating in the Schwarzschild space-time. From this, by functional derivation, we obtain the corresponding stress tensor $\langle S_{\mu\nu}\rangle$ which should then be compared with the expected results of $\langle T_{\mu\nu}\rangle$ given by eqs. (\ref{boli})-(\ref{hhin}). \sect{Polyakov's action} As a warm up exercise we start with the 2D Schwarzschild space-time and a conformal massless scalar field propagating on it. The purpose of it is purely pedagogical. The results will not be new, although their nonstandard derivation is new and illustrates in a simple way our strategy, the computational techniques and the basic physical ingredients we shall use to deal the most interesting 4D case (see next section). \\ The classical action for the massless field $f$ reads \begin{equation} S=\int d^2x\sqrt{-g^{(2)}} \partial_{\mu} f\partial^{\mu}f\ , \label{scaca} \end{equation} where $g^{(2)}$ is the determinant of the 2D Schwarzschild metric (see eqs. (\ref{sci}) and (\ref{edfi}) with angular part omitted). By canonical quantization and renormalization the corresponding $\langle T_{ab}\rangle$ can be computed in the three vacua and the results are given in eqs. (\ref{boli})-(\ref{hhli}). \\ Our procedure starts with the expression of the trace anomaly, which in arbitrary 2D spacetime reads \cite{duff-77} \begin{equation} \langle T\rangle = (24\pi)^{-1} R\ . \label{antr} \end{equation} The anomaly induced effective action is obtained by functional integration of this anomaly (see eq. (\ref{llll}) ). It is the well known Polyakov action \cite{polyak} \begin{equation} S_{eff}^{an}=S_P=-\frac{1}{96\pi } \int d^2 x \sqrt{-g} R \frac{1}{\Box} R \ , \label{viii} \end{equation} where $\Box$ is the d'Alembertian. \\ This action is nonlocal, but it can be made local either by choosing a conformal gauge or by introducing an auxiliary field $\psi$ (and keeping the gauge arbitrary). We can write \begin{equation} S_P=-\frac{1}{96\pi} \int d^2x\sqrt{-g}\, (-\psi\Box\psi +2\psi R)\ . \label{polo} \end{equation} The equation of motion for $\psi$ is \begin{equation} \Box\psi=R \ . \label{empsi} \end{equation} Once this is substituted back in eq. (\ref{polo}) we reobtain the nonlocal form of eq. (\ref{viii}). The anomaly induced stress tensor $\langle S_{ab}\rangle$ is defined as \begin{equation} \frac {2}{\sqrt{-g}}\frac{\delta S_P}{\delta g^{ab}}\equiv \langle S_{ab}\rangle = -\frac{1}{48\pi} \left[ 2\nabla_a\nabla_b\psi - \nabla_a\psi \nabla_b\psi - g_{ab}\left(2R -\frac{1}{2} (\nabla\psi)^2\right)\right] \ . \label{stlo} \end{equation} Our strategy is to solve eq. (\ref{empsi}) for the auxiliary field $\psi$ by assuming the 2D Schwarzschild spacetime as background, substitute the result in eqs. (\ref{stlo}) to find $\langle S_{ab}\rangle$ and compare with eqs. (\ref{boli})-(\ref{hhli}). \\ Before we start the program we have to solve a basic problem, how to implement state dependence in our $\langle S_{ab}\rangle$. Being $\langle T\rangle$ state independent, $S_{eff}^{an}$ makes no apparent reference to a particular quantum state. \\ Note however that eq. (\ref{empsi}) determines the solution for $\psi$ modulo arbitrary solution of the homogeneous equation $\Box\psi=0$. It is through the choice of the homogeneous solution that the state dependence will be encoded in our framework. The homogeneous solution will be chosen by imposing appropriate boundary conditions on $\psi$ that reflect the physical features of the states $|B\rangle$, $|U\rangle$ and $|H\rangle$.\\ Since by eq. (\ref{empsi}) we have \begin{equation} \psi= \frac{1}{\Box}R \ , \label{mmmm} \end{equation} by this procedure we impose boundary conditions on the operator $1/\Box$ characterizing the particular quantum state we are working with. \\ For the 2D Schwarzschild spacetime eq. (\ref{empsi}) becomes \begin{equation} \partial_{a}[\partial^{a}\psi]=\frac{4M}{r^3}\ , \label{emca} \end{equation} whose general solution we can write as \begin{equation} \psi= at -\ln(1-2M/r) + A\left[ r+2M\ln(r-2M)\right] +B \label{soge} \end{equation} with \begin{equation} \frac{\partial \psi}{\partial r} = (1-2M/r)^{-1} (-\frac{2M}{r^2} +A)\ , \label{depi} \end{equation} where $a,A$ and $B$ are arbitrary constants which once fixed determine the solution of the homogeneous equation. \\ The choice of a linear dependence in $t$ is simple. Given the structure of $\langle S_{ab}\rangle$ in terms of $\psi$ (see eq. (\ref{stlo}) ), it is clear that any $t$ dependence different from the linear would imply $\partial_t \langle S_{ab}\rangle \neq 0$ , which contradicts the exact result $\partial_t \langle T_{ab}\rangle =0$. The presence of such a linear term allows, as we shall see, that $\langle S_{rt}\rangle \neq 0$ which is indeed needed in the Unruh state. \\ Inserting the solution for $\psi$ in eq. (\ref{stlo}) we obtain \begin{eqnarray} \langle S_{tt}\rangle &=& \frac{1}{12\pi} \left[ -\frac{2M}{r^3} +\frac{7 M^2}{2 r^4}\right] + (48\pi)^{-1} \frac{A^2 +a^2}{2}\ , \nonumber \\ \langle S_{rr}\rangle &=& -\frac{1}{48\pi} (1-2M/r)^{-2}\left(\frac{2M^2}{r^4} - \frac{A^2 +a^2}{2}\right)\ , \nonumber \\ \langle S_{rt} \rangle &=& (48\pi)^{-1}(1-2M/r)^{-1}Aa \ . \label{solro} \end{eqnarray} Now we have to fix the arbitrary constants according to our choice of quantum state. \\ \\ \noindent i) Boulware vacuum $|B\rangle$. \\ This state by construction reduces as $r\to \infty$ to Minkowski vacuum and there $\langle B|T_{ab}|B\rangle$ should vanish. It is clear from the above equations that in order to fullfill the asymptotic requirement we have to impose $a=A=0$ and we find \begin{eqnarray} \langle B|S_{tt}|B\rangle &=& \frac{1}{12\pi} \left[ -\frac{2M}{r^3} +\frac{7M^2}{2r^4}\right] \ , \nonumber \\ \langle B| S_{rr}|B\rangle &=& -\frac{1}{48\pi}\frac{2M^2}{r^4} (1-2M/r)^{-2} \ , \nonumber \\ \langle B| S_{rt}|B \rangle &=& 0 \ . \label{bolio} \end{eqnarray} We see the exact agreement of our $\langle B|S_{ab}|B\rangle$ with eqs. (\ref{boli}), namely \begin{equation} \langle B|S_{ab}|B\rangle = \langle B|T_{ab}|B\rangle \ . \label{exag} \end{equation} With the above choice of integration constants the auxiliary field $\psi$ becomes \begin{equation} \psi=-\ln(1-2M/r) +B \label{sobo} \end{equation} which vanishes for $r\to\infty$ if we further set $B=0$. Note the singularity of $\psi $ for $r=2M$ which causes the divergence of $\langle B|S_{ab}|B\rangle $ on the horizon. \\ \\ \noindent ii) Unruh state $|U\rangle$. \\ By assumption there is no incoming flux (i.e. $\langle U|S_{vv}|U\rangle=0$ as $r\to\infty$) and $\langle U|S_{ab}|U\rangle$ has to be regular on the future horizon (i.e. $\langle U|S_{uu}|U\rangle \to 0$ at least as $(r-2M)^2$, $\langle U|S_{vv}|U\rangle \sim reg.$ and $\langle U|S_{uv}|U\rangle \sim (r-2M)$ for $r\to 2M$ ). The asymptotic condition \begin{equation} \langle U|S_{vv}|U\rangle = \frac{1}{4}\left( \langle U|S_{tt}|U\rangle + (1-2M/r)^2\langle U|S_{rr}|U\rangle +2(1-2M/r)\langle U|S_{rt}|U\rangle \right) \to 0 \label{noin} \end{equation} requires $A=-a$. Furthemore $\langle U|S_{uu}|U\rangle$ vanishes like $(r-2M)^2$ for $r\to 2M$ if $A=(4M)^{-1}$. Summarizing we have \begin{eqnarray} \langle U |S_{tt}|U\rangle &=& \frac{1}{12\pi} \left[ -\frac{2M}{r^3} +\frac{7M^2}{2r^4}\right] + (768\pi M^2)^{-1} , \nonumber \\ \langle U |S_{rr}|U\rangle &=& - (1-2M/r)^{-2}\left[ \frac{M^2}{24\pi r^4} - (768\pi M^2)^{-1}\right] , \nonumber \\ \langle U |S_{rt}|U\rangle &=& -(1-2M/r)^{-1}(768\pi M^2)^{-1} \ . \label{anin} \end{eqnarray} Comparison with eqs. (\ref{unli}) reveals \begin{equation} \langle U|S_{ab}|U\rangle = \langle U|T_{ab}|U\rangle\ . \label{coun} \end{equation} It is quite interesting to examine the behaviour of the auxiliary field $\psi$. From eq. (\ref{soge}) we see that the asymptotic requirement $A=-a$ yields $\psi \sim -Au $ as $r\to \infty$, whereas the regularity condition on the horizon $A=(4M)^{-1}$ reveals that $\psi \sim -v/4M + const.$ as $r\to 2M$. We have therefore the nice connection: \\ \noindent a) for $r\to\infty$ (we set $B=0$) $\psi \sim -u/4M \ \Leftrightarrow \langle U |S_{ab}|U\rangle $ describes outgoing radiation; \\ \noindent b) for $r\to 2M$ $\psi \sim -v/4M +const. \ \Leftrightarrow \langle U |S_{ab}|U\rangle$ describes an ingoing flux of negative energy radiation; $\psi$ and $\partial_a\psi$, $a,b=v,r$ regular on the future horizon $\Leftrightarrow \ \langle U |S_{ab}|U\rangle$ regular on the future horizon. \\ \\ iii) Israel-Hartle-Hawking state $|H\rangle$. \\ This state corresponds to thermal equilibrium and thus we demand $\psi$ to be time independent, i.e. $a=0$ which implies $\langle H|S_{rt}|H\rangle=0$. Furthemore $|H\rangle$ is constructed in a way that makes the stress tensor regular both on the future and past horizon, which is achieved by requiring that as $r\to 2M$ $\langle H|S_{uu}|H\rangle=\langle H|S_{vv}|H\rangle$ vanish like $(r-2M)^2$. Inspection of eqs. (\ref{solro}) reveals that we can fullfill the above requirement by demanding \begin{equation} \frac{1}{48\pi} \frac{A^2}{2} = \frac{1}{384\pi M^2} \ , \label{prul} \end{equation} i.e. \begin{equation} A=\frac{1}{2M}\ . \label{afiss} \end{equation} With the above choice $\psi$ is regular as $r \to 2M$, since by eq. (\ref{soge}) we have that for the state $|H\rangle$ \begin{equation} \psi= \ln r + \frac{r}{2M} +B \label{psihh} \end{equation} which is indeed regular. \footnote{The other solution of eq. (\ref{prul}), namely $A=-1/2M$, makes $\psi$ singular on the horizon.} We can make $\psi$ vanishing on the horizon by choosing the arbitrary constant $B$ such that \begin{equation} B=-(\ln 2M +1)\ , \label{bico} \end{equation} i.e. for the $|H\rangle$ state we have \begin{equation} \psi= \ln r +\frac{r}{2M} - (\ln2M +1)\ . \label{pire} \end{equation} Note that this expression could have been derived by looking at the static solution of the auxiliary field equation (\ref{emca}) of the form \begin{equation} \psi(r)= \int_{2M}^r dr_1 (1-2M/r_1)^{-1}\int _{2M}^{r_1} dz \frac{4M}{z^3} \ . \label{conint} \end{equation} Inserting now $a=0$ and $A=1/2M$ in eqs. (\ref{solro}) we find \begin{eqnarray} \langle H|S_{tt}|H\rangle &=& (12\pi)^{-1}\left[-\frac{2M}{r^3} +\frac{7M^2}{2r^4}\right] + (384\pi M^2)^{-1}\ , \nonumber \\ \langle H|S_{rr}|H\rangle &=& -(48\pi)^{-1} (1-2M/r)^2 \left[ \frac{2M}{r^4} - \frac{1}{8M^2} \right]\, \nonumber \\ \langle H|S_{tr}|H\rangle &=& 0 \ . \label{ricu} \end{eqnarray} Direct check with eqs. (\ref{hhli}) reveals again the identity \begin{equation} \langle H|S_{ab}|H\rangle= \langle H|T_{ab}|H\rangle\ . \label{ulid} \end{equation} We have therefore shown in this simple 2D example the perfect agreement of our $\langle S_{ab}\rangle$ and $\langle T_{ab}\rangle$. This is not surprising at all; it is a general feature not at all limited to the 2D Schwarzschild background. It is well known that the auxiliary field equation can be integrated without making reference to any particular background. An arbitrary 2D spacetime \begin{equation} ds^2 =-g_{ab}dx^adx^b \label{geme} \end{equation} can always be parametrized as follows (conformal gauge) \begin{equation} ds^2=-e^{2\rho}dx^+dx^- \ . \label{conga}\end{equation} In that gauge the general solution of the equation for $\psi$ (eq. (\ref{empsi}) ) reads \begin{equation} \psi= -2\rho + F(x^+) + G(x^-)\ , \label{roso} \end{equation} where $F(x^+)$ and $G(x^-)$ are arbitrary functions of their arguments. Inserting this in eq. (\ref{stlo}) we obtain $\langle S_{ab}\rangle$. This has exactly the same form as $\langle T_{ab}\rangle$ in eqs. (\ref{renst}) which, as already said, can be derived by canonical quantization. \\ The motivation for our nonstandard derivation of the stress tensor for a conformal scalar in 2D Schwarzschild space-time was purely pedagogical\footnote{More detailed consideration, including critical analysis of the current literature, have been reported separately in \cite{NEW}.}. It has shown explicitly how to implement state dependence in the framework of the anomaly induced effective action by imposing appropriate boundary conditions on the auxiliary field $\psi$ (or equivalently on $\langle S_{ab}\rangle$) which reflect the physical properties of the quantum state. \sect{Four dimensions: Reigert's action} Having set the basis of our formalism we are ready to attack the physically more interesting example: a conformal invariant scalar field $f$ in the 4D Schwarzschild black hole space-time. The classical action describing massless conformal invariant scalar field $f$ is \begin{equation} S=\int d^4x\sqrt{-g}\left( \partial_{\mu}f \partial^{\mu}f -\frac{1}{6} Rf^2\right) + S_{vacuum}\, , \label{quca} \end{equation} where the $S_{vacuum}$ is necessary for the renormalizability of the theory (see \cite{book} for the general introduction to the renormalization in curved space-time). Performing the path integration over the scalar field, one meets divergences and the renormalization is necessary. The trace anomaly results from the renormalization of the vacuum action in (\ref{quca}). The vanishing of the classical trace has its quantum counterpart in the presence of a trace anomaly, composed by the terms coming from the vacuum counterterms \cite{duff-77} \begin{equation} \label{6} \langle T^{\alpha}_{\alpha}\rangle \equiv \langle T \rangle =- \frac{1}{(4\pi)^2}\left( aC^2 + bE +c\Box R \right) \ . \end{equation} For the free conformal invariant matter fields the three structures presented in (\ref{6}) are sufficient, and the $\,\sqrt{-g}R^2$-type counterterm does not arise. That is why the corresponding term is absent in (\ref{6}) \footnote{Also we notice that that there is no expression, local or nonlocal, in the effective action, which could produce the term $\,\sqrt{-g}R^2$ in the trace anomaly \cite{reigert}.}. The anomaly induced effective action which yields the anomaly (\ref{6}) has been independently derived by Reigert \cite{reigert} and Fradkin and Tseytlin \cite{frts} (see also \cite{book,antmot,cosh,anju} for further applications and references about the Reigert's action). The expression derived in \cite{reigert} has the form: \begin{eqnarray} S_{eff}^{an}= \frac{1}{(4\pi)^2} \int d^4 x \int d^4 y \left\{\sqrt{-g} \left[ E-\frac{2}{3}\Box R\right]\right\}_x G(x,y) \nonumber \\ \left\{ \sqrt{-g}\left[ \frac{a}{4}C^2 +\frac{b}{8}(E-\frac{2}{3}\Box R)\right]\right\}_y - \frac{c+\frac{2}{3}b}{12 (4\pi)^2} \int d^4 x\sqrt{-g(x)}R^2\ , \label{reiac} \end{eqnarray} where $a= 1/120$, $b=-1/360$, $c = 1/180$. $G(x,y)$ is the inverse of the fourth order operator $\Delta_4$ \begin{equation} \label{10} \Delta_4=\Box^2 - 2R^{\mu\nu}\nabla_{\mu}\nabla_{\nu}+ \frac{2}{3}R\Box -\frac{1}{3}(\nabla^{\mu}R)\nabla_{\mu} \ , \end{equation} i.e. \begin{equation} (\sqrt{-g}\Delta_4)_x G(x,y)=\delta^4(x-y) \ . \label{11} \end{equation} Note that $\Delta_4$ acting on a scalar field of zero conformal weight is the unique self-adjoint conformal invariant differential operator in $4D$ \cite{pan}. This also implies the conformal invariance of $G(x,y)$. Let us rewrite the Reigert's action in a more symmetric way \begin{eqnarray} S_{eff}^{an} &=&- \frac{c+\frac{2}{3}b}{12(4\pi)^2}\int d^4 x \sqrt{-g}R^2 -\frac{1}{2(4\pi)^2}\int d^4 x \sqrt{-g(x)} \int d^4 y \sqrt{-g(y)} \nonumber \\ &\ &\frac{\sqrt{-b}}{2} \left[ (E-\frac{2}{3}\Box R)+\frac{a}{b}C^2\right]_x G(x,y) \frac{\sqrt{-b}}{2}\left[ (E-\frac{2}{3}\Box R) + \frac{a}{b}C^2\right]_y \nonumber \\ &+& \frac{1}{2(4\pi)^2}\int d^4 x\sqrt{-g(x)} \int d^4 y \sqrt{-g(y)} \left( \frac{a}{2\sqrt{-b}}C^2\right)_x G(x,y)\left(\frac{a}{2\sqrt{-b}}C^2\right)_y \ . \label{azsim} \end{eqnarray} This nonlocal action can be made local by the introduction of two auxiliary fields $\phi$ and $\psi$ \cite{reigert,shapiro} \begin{eqnarray} \label{14} S_{eff}^{an}&=& +\int d^4 x\sqrt{-g}\left[ \frac{1}{2}\phi\Delta_4\phi + \phi\left( \frac{\sqrt{-b}}{8\pi}(E-\frac{2}{3}\Box R)-\frac{a}{8\pi\sqrt{-b}}C^2 \right) \right] \nonumber \\ &+&\int d^4x\sqrt{g}\left(-\frac{1}{2}\psi\Delta_4\psi +\frac{a}{8\pi\sqrt{-b}} C^2 \psi \right) -\frac{c+\frac{2}{3}b}{12(4\pi)^2}\int d^4x \sqrt{-g}R^2\ . \end{eqnarray} Once the equations of motion for $\phi$ and $\psi$ are used, eq. (\ref{14}) reduces to the nonlocal form (\ref{reiac}). We remark that the path integration over the auxiliary fields would give an extra contribution to the anomaly (\ref{6}), modifying the values of the constants $a,b,c$ in (\ref{6}). Here we consider the introduction of the auxiliary fields $\psi,\phi$ as a method of working with the nonlocalities of the expression (\ref{azsim}) and hence require the correspondence to hold only on classical level. \\ Note that the last nonlocal term in eq. (\ref{azsim}) (or equivalently the second integrand in eq. (\ref{14}) ) is conformally invariant and hence could be removed and included in the Weyl invariant part of the effective action $S^W$ (see eq. (\ref{coac}) ). For later use we shall keep this term. However its coefficient $a/(8\pi\sqrt{-b})$ can be replaced by an arbitrary parameter, say $l_1$, since a change in this parameter can be reabsorbed by the Weyl invariant part $S^W(g_{\alpha\beta})$ of $S_{eff}$ which is not determined by the trace anomaly. \\ So we shall work with the following anomaly induced effective action \begin{eqnarray} \label{elle} S_{eff}^{an}&=&-\frac{k_3}{12}\int d^4 x\sqrt{-g} R^2 +\int d^4 x\sqrt{-g}\left[ \frac{1}{2}\phi\Delta_4\phi + \phi\left(k_1C^2 + k_2(E-\frac{2}{3}\Box R) \right) \right] \nonumber \\ &+&\int d^4x\sqrt{g}\left(-\frac{1}{2}\psi\Delta_4\psi +l_1 C^2 \psi \right) \ , \end{eqnarray} where we set $k_1=-a/(8\pi\sqrt{-b})$, $k_2=\sqrt{-b}/8\pi$ and $k_3=(c+\frac{2b}{3})/(4\pi)^2 $, bearing in mind the values of $a,b,c$ specific for a single conformal scalar. \\ After eliminating the auxiliary fields $\phi$ and $\psi$ eq. (\ref{elle}) does not reduce to Reigert's nonlocal action of eq. (\ref{reiac}) (unless $l_1=a/(8\pi\sqrt{-b})$). The difference, as we just remarked, is a conformally invariant functional, hence of the form $S^W$, on which we have no control. \\ From eq. (\ref{elle}) the equations of motion for the auxiliary fields follow \begin{equation} \label{15} \frac{1}{\sqrt{-g}}\frac{\delta S_{eff}^{an}}{\delta\phi}= \Delta_4\phi + k_1 C^2 + k_2(E-\frac{2}{3}\Box R)=0\ , \end{equation} \begin{equation} \label{16} \frac{1}{\sqrt{-g}}\frac{\delta S_{eff}^{an}}{\delta\psi}= -\Delta_4\psi +l_1 C^2 =0\ . \end{equation} Introducing now the traceless tensor $K_{\mu\nu}(\phi)$ \begin{eqnarray} \label{17} &\ & K_{\mu\nu}(\phi) = \frac{1}{\sqrt{-g}}\,\frac{\delta }{\delta g^{\mu\nu}} \int d^4x\sqrt{-g}\,\left\{ \phi\Delta_4\phi \right\} = -\frac12\,g_{\mu\nu}\,\left( (\Box \phi)^2 - 2R^{\alpha\beta}\phi_\alpha\phi_\beta + \frac23\,(\nabla\phi)^2 R\right) \nonumber \\ &-& \phi_\mu(\nabla_\nu {\Box}\phi) - \phi_\nu(\nabla_\mu {\Box}\phi) + 2\phi_{\mu\nu}({\Box}\phi) + 2\phi^\lambda\phi_{\mu\nu\lambda} - \frac23\,\phi^\lambda \,(\phi_{\lambda\mu\nu} + \phi_{\lambda\nu\mu}) + \frac13\,g_{\mu\nu}\,\phi^{\lambda\tau}\,\phi_{\lambda\tau} \nonumber \\ &+& \frac13\,g_{\mu\nu}\,\phi^{\lambda}\,({\Box}\phi_{\lambda}) - \frac43\,\phi^{\lambda}_\mu\,\phi_{\lambda\nu} - 2\,\phi_\lambda\,(R^\lambda_\mu\phi_\nu + R^\lambda_\nu\phi_\mu) + \frac23\,R\,\phi_\mu\,\phi_\nu + \frac23\,R_{\mu\nu} \,\phi_\lambda\,\phi^\lambda \end{eqnarray} we can write the anomaly induced stress tensor $\langle S_{\mu\nu}\rangle$ as \begin{eqnarray} \label{18} \frac{2}{\sqrt{-g}}\frac{\delta S^{eff}_{an}}{\delta g^{\mu\nu}} &=& \langle S_{\mu\nu}\rangle = K_{\mu\nu}(\phi)-K_{\mu\nu}(\psi) + 8\nabla^{\lambda}\nabla^{\tau}Z R_{\mu\lambda\nu\tau} -\,g_{\mu\nu}\, Z R^2_{\rho\sigma\alpha\beta} \nonumber \\ &+& 4Z\,R_{\mu\rho\lambda\tau}\,{R_{\nu }}^{\rho\lambda\tau} +\frac{4k_2}{3}\,\left[ (\nabla_\mu\nabla_\nu {\Box} \phi) - g_{\mu\nu} ({\Box}^2 \phi) \right] +... \end{eqnarray} where we have defined \begin{equation} \label{19} Z\equiv (k_1+k_2)\phi +l_1\psi \end{equation} and $\phi_{\lambda\tau}\equiv \nabla_{\lambda}\nabla_{\tau}\phi$, etc. . The dots in eq. (\ref{18}) indicate terms containing either the Ricci tensor $R_{\mu\nu}$ or its contraction $R$. In our subsequent analysis the background space-time chosen will be Schwarzschild and therefore these terms, together with the sum of the two terms linear in $Z$ in eq. (\ref{18}), vanish identically. \\ The equation of motion for $\phi$ in the Schwarzschild spacetime (see eq. (\ref{sci})) reads \begin{equation} \Box_s^2\phi=\alpha \frac{M^2}{r^6}\ , \label{emsb} \end{equation} where $\alpha \equiv -48(k_1+k_2)$ and $\Box_s$ is the Delambertian written in Schwarzschild coordinates. The solution for $\phi$ can be given in the following form \begin{equation} \phi(r,t)=d\cdot t + w(r)\ , \label{capo} \end{equation} where $w(r)$ is such that \begin{eqnarray} \label{22} \frac{dw}{dr} &=& \frac{B}{3}r +\frac{2}{3}MB-\frac{A}{6}-\frac{\alpha}{72M} + \left(\frac{4}{3}BM^2 + \frac{C}{2M}-AM -\frac{\alpha}{24}\right) \frac{1}{r-2M} -\frac{C}{2M}\frac{1}{r} \nonumber \\ &+&\ln r \left[ -\frac{\alpha M}{18}\frac{1}{r(r-2M)} -\left(\frac{A}{2M}-\frac{\alpha}{48M^2}\right)\frac{r^2}{3(r-2M)} \right] \nonumber \\ &+& \ln(r-2M) \left[ \left(\frac{A}{2M}-\frac{\alpha}{48M^2}\right) \frac{r^3 -(2M)^3}{3r(r-2M)} \right] \end{eqnarray} and $(d,A,B,C)$ are constants that specify the homogeneous solution $\Box_s^2\phi=0$ and hence the quantum state. The presence of a linear term in time in eq. (\ref{capo}) has the same explanation given in section II (see the discussion following eq. (\ref{depi}); it allows $\langle S_{rt}\rangle \neq 0$ but keeping $\partial_t \langle S_{\mu\nu}\rangle=0$). One can give solution for the equation of motion for the other auxiliary field $\psi$ in the same form with obvious replacements $$ \alpha \equiv - 48 (k_1+k_2) \longrightarrow \beta\equiv 48l_1 \,\,\,\,\,\,\,\,\, {\rm and} \ (d,A,B,C)\longrightarrow (d',A',B',C')\,. $$ Substituting the solutions for the auxiliary fields in eqs. (\ref{18}) we obtain the effective stress tensor for the Schwarzschild space-time in the form \begin{equation} \langle S_{\mu\nu}\rangle = \langle S_{\mu\nu}(\phi)\rangle + \langle S_{\mu\nu}(\psi)\rangle \label{seva} \end{equation} separating the contribution of each individual field to $\langle S_{\mu\nu}\rangle$. Both $\phi$ and $\psi$ are related to the inverse of the fourth order differential operator $\Delta_4$, but they are independently defined, and therefore the boundary conditions have to be imposed on them individually. For later use we reproduce the limiting behaviour of $\partial_r\phi$ as \begin{equation} \label{25} \partial_r\phi \sim \frac{E}{r-2M} +2M (\frac{A}{2M}-\frac{\alpha}{48M^2})\ln(r-2M) + reg.\ terms \ , \end{equation} where \begin{equation} \label{26} E=-\frac{\alpha}{12}+\frac{4}{3}BM^2 +\frac{C}{2M}-2M^2 (\frac{A}{2M}-\frac{\alpha}{48M^2})-\frac{2}{3}AM \ln2M \ , \end{equation} in the limit $r\to 2M$, and \begin{equation} \label{29} \partial_r\phi \to \frac{B}{3}r + \frac{2}{3}MB -\frac{A}{2} \end{equation} at infinity. \\ \\ \noindent i) Boulware state $|B\rangle$. \\ \noindent For flat space and Minkowski vacuum $\phi=const.$ (which we can set to zero without any consequences). This implies $\langle M|S_{\mu\nu}|M\rangle=0$ as it should be. Since as $r\to \infty$ $|B\rangle \to |M\rangle$, the correct asymptotic limit on $\phi$ is obtained in eqs. (\ref{capo}), (\ref{22}) by setting all constants to zero. So for the state $|B\rangle$ we have \begin{eqnarray} \label{ccoo} \frac{dw}{dr}&=& -\frac{\alpha}{72M} -\frac{\alpha}{24}\frac{1}{r-2M} +\ln r \left[ -\frac{\alpha M}{18}\frac{1}{r(r-2M)} +\frac{\alpha}{48M^2}\frac{r^2}{3(r-2M)} \right] \nonumber \\ &+&\ln(r-2M) \left[ -\frac{\alpha}{48M^2}\frac{r^3 -(2M)^3}{3r(r-2M)} \right] \end{eqnarray} and an analogous expression for $\psi$ ($\alpha\to \beta$). \\ Starting from these expressions one can obtain $\langle B|S_{\mu\nu}|B\rangle$. Being the calculations rather long and boring, we report just the limiting behaviours \begin{equation} \label{ciuco} \langle B| S_{\mu}^{\nu} |B \rangle \to \frac{1}{2} \frac{\alpha^2-\beta^2}{(24)^2}\frac{1}{(2M)^4f^2} \pmatrix{ -1 & 0 & 0 & 0 \cr 0 & 1/3 & 0 & 0 \cr 0 & 0 & 1/3 & 0 \cr 0 & 0 & 0 & 1/3 \cr} \end{equation} for $r\to 2M$ ($f=1-2M/r$) and \begin{equation} \langle B| S_{\mu}^{\nu} |B \rangle \to O(\frac{1}{r^6}) \label{calo} \end{equation} as $r\to \infty$. \\ The qualitative agreement of our $\langle B|S_{\mu\nu}|B\rangle$ and $\langle B|T_{\mu\nu}|B\rangle$ of eqs. (\ref{boin}), (\ref{boho}) is rather nice. We see the expected $r^{-6}$ fall off and the $1/f^2$ divergence on the horizon. \\ \\ \noindent ii) Unruh state $|U\rangle$. \\ \noindent The Unruh state $|U\rangle$ agrees with Minkowski vacum on past null infinity, i.e. no incoming radiation. This requires $\phi\sim u$ as $r\to\infty$, which implies $\langle U|S_{vv}|U\rangle \to 0$ for $r\to \infty$. This asymptotic behaviour of $\phi$ is achieved by requiring $B=0$ and $d=A/2$. Regularity along the future horizon requires $\phi \sim v$ and $\partial_a\phi\sim reg.$, $a=v,r$ (see the discussion following eq. (\ref{coun})). This is achieved by fixing $E=2dM$ ($\phi\sim v$) and $A=\alpha/24M$ ($\partial_a\phi$ finite). That this is the correct choice it can be seen by direct evaluation of $\langle U|S_{\mu\nu}|U\rangle$ \cite{balfabsha}. \\ Near the horizon to the order $f^{-2}$ we have \begin{equation} \label{27} \langle U| S_{\mu}^{\ \nu}(\phi)|U\rangle \sim \frac{(E^2-4d^2M^2)}{32M^4f^2} \pmatrix{ -1 & 0 & 0 & 0 \cr 0 & 1/3 & 0 & 0 \cr 0 & 0 & 1/3 & 0 \cr 0 & 0 & 0 & 1/3 \cr}\ , \end{equation} which indeed vanishes for $E=2dM$. Logarithmic divergence in the pressure $\langle T_{\theta}^{\ \theta}\rangle$ are eliminated by $A=\alpha/24M$. Repeating the same arguments for $\psi$ gives $E'=2d'M, A'=\beta /24M, B'=0$. The following leading behaviour of $\langle S_{ab}\rangle$ emerges \begin{equation} \label{28} \langle U| S_{a}^{\ b} |U \rangle \sim \frac{\alpha^2-\beta^2}{2(48M^2)^2} \pmatrix{ 1/f & -1 \cr 1/f^2 & -1/f \cr}\ ,\ r\to 2M\ , \end{equation} which is indeed regular on the future horizon. By the above choice of the constants, we can find the asymptotic form of $\langle U|S_{\mu\nu}|U\rangle$ \begin{equation} \label{giango} \langle U| S_{\mu}^{\ \nu} |U \rangle \to \frac{\alpha^2-\beta^2}{2r^2(24M)^2} \pmatrix{ -1 & -1 & 0 & 0 \cr 1 & 1 & 0 & 0 \cr 0 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 \cr}\ , \ r\to \infty \ . \end{equation} Our results eqs. (\ref{28}), (\ref{giango}) are in exact agreement with $\langle U|T_{\mu\nu}|U\rangle$ given by eqs. (\ref{unin}), (\ref{unho}) once the luminosity $L$ is identified with \begin{equation} \frac{L}{4\pi} =\frac{(\alpha^2-\beta^2)}{2(24M)^2}\ . \label{lume} \end{equation} \\ \noindent iii) Israel-Hartle-Hawking state $|H\rangle$. \\ \noindent This state is an equilibrium state regular both on the future and the past horizons. For an equilibrium state $d=0$ which implies no net fluxes ($\langle H|S_{rt}(\phi)|H\rangle =0$). Inspection of eq. (\ref{25}) reveals that $\phi$ and $\partial_r\phi$ are regular both on the future and the past horizons by imposing $E=0$ and $A=\alpha/24M$. This leaves the parameter $B$ free. However, if the solution for $\phi$ is obtained by an integral formula like eq. (\ref{conint}), which sets the lower limit of integration to be $r=2M$, we would obtain \begin{equation} A=\frac{\alpha}{24M} \ ,\ \ \ B=\frac{11\alpha}{288M^2}\ ,\ \ \ C=\frac{7\alpha M}{108} +\frac{\alpha M}{18}\ln 2M \label{prpr} \end{equation} which implies $E=0$. By this choice, which we think characterizes the $|H\rangle$ state, we eventually arrive at the following expression \begin{equation} \label{lola} \langle H| S_{\mu}^{\ \nu} |H \rangle \to - \frac{k_2 \alpha}{96M^4} \pmatrix{ 1 & 0 & 0 & 0 \cr 0 & 1 & 0 & 0 \cr 0 & 0 & 2 & 0 \cr 0 & 0 & 0 & 2 \cr} \end{equation} in the limit $r\to 2M$ and \begin{equation} \label{lopp} \langle H| S_{\mu}^{\ \nu} |H \rangle \to \frac{7}{2}\left(\frac{11}{864M^2}\right)^2(\alpha^2-\beta^2) \pmatrix{ -1 & 0 & 0 & 0 \cr 0 & 1/3 & 0 & 0 \cr 0 & 0 & 1/3 & 0 \cr 0 & 0 & 0 & 1/3 \cr} \end{equation} as $r\to\infty$. The asymptotic limit of eq. (\ref{lopp}) describes indeed a thermal bath as expected (compare to eq. (\ref{hhin})). Note the appearance of the factor $(\alpha^2-\beta^2)$ just as in eq. (\ref{giango}). \\ Putting some numbers, if we choose $l_1=\frac{a}{8\pi\sqrt{-b}}$ which reconduces our $S_{eff}^{an}$ to Reigert's one, we have the disappointing result $(\alpha^2-\beta^2)<0$ that is physically meaningless. A similar situation was found in investigation of the anomaly induced effective action representing 4D minimally coupled scalar fields spherically reduced to 2D at the classical level ( see \cite{treref}). For $l_1=0$ (i.e. complete absorbtion of the conformally invariant part of our $S_{eff}^{an}$ in $S^W$) we find that the overall coefficient of eq. (\ref{lopp}) is $\sim \frac{2.2\ 10^{-4}}{\pi^2M^4} $ , which is much bigger than the correct one $\sim \frac{8\ 10^{-6}}{\pi^2 M^4} $. Exact agreement of our result eq. (\ref{lopp}) and eq. (\ref{hhin}) requires a fine tuning of the parameter $\beta$, namely $\beta\sim\frac{6.2 \ 10^{-1}}{\pi} $. For this value of $\beta $ the luminosity of an evaporating black hole (see eq. (\ref{lume})) turns out to be $L= \frac{4.9\ 10^{-5}}{\pi M^2}$, which is roughly four times smaller than the value given in \cite{dewitt}. \\ \\ \noindent Despite many nice features, there are some disappointing aspects of our $\langle S_{\mu\nu}\rangle$. The overall coefficient of $\langle B|T_{\mu\nu}|B\rangle$ on the horizon is expected to be the same as the asymptotic limit of $\langle H|T_{\mu\nu}|H\rangle$ up to a minus sign (see eqs. (\ref{boho}) and (\ref{hhin}) ). This is not the case for $\langle S_{\mu\nu}\rangle$ (compare eq. (\ref{ciuco}) and eq. (\ref{lopp}) ). \\ Unfortunately this is not all. If we go beyond the leading terms there is no agreement of our $\langle S_{\mu\nu}\rangle$ and $\langle T_{\mu\nu}\rangle$. Just to mention some negative aspects: $\langle U|S_{\theta}^{\ \theta}|U\rangle \sim r^{-3}$ in the limit $r\to \infty$ while $\langle U|T_{\theta}^{\ \theta}|U\rangle$ is expected to fall off as $r^{-4}$. Furthemore in the same limit the coefficients in $r^{-1}$ and $r^{-2}$ of our $\langle H|S_{\mu\nu}|H\rangle$ do not correspond to redshifted thermal radiation as we have for $\langle H|T_{\mu\nu}|H\rangle$. One can guess that this discrepancy may be removed by a careful modelling of the conformal part in (\ref{coac}) through the introduction of other conformal structures \cite{alter}, or, hopefully, by using the results of some of the direct approximate calculations of the effective action (see, for example, \cite{vilk}). \sect{Conclusions} In a field theory the knowledge of the trace anomaly allows part of the effective action to be reconstructed; this is the so called anomaly induced effective action, i.e. $S_{eff}^{an}$. The basic features of this object are the nonlocality and the apparent absence of any reference to a particular quantum state. Auxiliary fields have been introduced to make $S_{eff}^{an}$ local and boundary conditions have been imposed on them to select the appropriate quantum states. \\ This procedure has been positively tested in a simple 2D example where $S_{eff}^{an}$ is the well known Polyakov action. We found exact agreement of the stress tensor calculated from this effective action (i.e. $\langle S_{ab}\rangle$) and the one ($\langle T_{ab}\rangle$) resulting from standard canonical quantization. \\ The expertise gained from the analysis of this simple 2D model has allowed us to attack by similar methods the physically much more interesting example of 4D Schwarzschild black hole. We have been able to construct explicitly $\langle S_{\mu\nu}\rangle$ for the three quantum states relevant for the discussion of quantum matter fields in a black hole spacetime, namely Boulware state (vacuum polarization around a static star), Unruh state (evaporation of a black hole formed by gravitational collapse) and Israel-Hartle-Hawking state (equilibrium of a black hole and a thermal bath). \\ By appropriate choice of the arbitrary parameter $l_1$ of our model (or, equivalently, of $\beta$), we were able to show the qualitative agreement of our $\langle S_{\mu\nu}\rangle $ and the canonically computed $\langle T_{\mu\nu}\rangle$ at infinity and near the horizon. Unfortunately these nice features cannot be extended beyond leading terms. This is expected since the fondamental brick of our 4D investigation, namely $S_{eff}^{an}$, up to a Weyl invariant term coincides with Reigert's action. It is well known \cite{erosbo} that this action is rather deficitary if considered as full effective action. Unlike Polyakov's action Reigert's one is unable even to correctly reproduce the three-point correlation function of the theory on the flat background. In view of this fact, the discrepancies found in our $\langle S_{\mu\nu}\rangle$ are not surprising at all. \\ What in our opinion is however remarkable is that, neverthless, our $S_{eff}^{an}$ does indeed reproduce the expected behaviour of the stress tensor near the horizon and at infinity upon which our understanding of black hole evaporation is based. \\ In view of our investigation, a careful use (not a straightforward one) of $\langle S_{\mu\nu}\rangle$ in the semiclassical Einstein equations to get at least a feeling of the backreaction in an evaporating black hole can be made. We hope to come back to this point in a future work. \vskip 4mm \noindent {\bf Acknowledgements} R.B. deeply thanks V. Frolov and R. Zucchini for stimulating discussions. A.F. is supported by an INFN fellowship. I.Sh. is grateful to the Physics Department of UFJF for warm hospitality, to CNPq for the fellowship and to FAPEMIG (MG) for the travel grant. His work was partially supported by RFFI (project 99-02-16617).
\section{introduction} Let ${\bold C}$ be a tensor category over a field ${\Bbb K}$ and $\Omega$ an ``embedding" of ${\bold C}$ into the category $\bold{mod}({\Bbb K})$ of all finite-dimensional vector spaces over ${\Bbb K}$. A version of the Tannaka-Krein duality theorem says that there exists a natural way to construct a bialgebra $C (\Omega)$ such that the category $\bold{com} (C (\Omega))$ of all finite-dimensional right $C (\Omega)$-comodules is equivalent to ${\bold C}$. After the discovery of the quantum groups and Jones' index theory, it is recognized that there exist natural interesting examples of tensor categories which have no embeddings into $\bold{mod}({\Bbb K})$ (see e.g. \cite{GelfandKazhdan}, \cite{Ocneanu}, \cite{Turaev}). In this letter, we show that there exists a natural simple way to construct an embedding $\Omega_0$ of each tensor category ${\bold C}$ into a certain bimodule category ${{\bold{bmd}}} ({\it R})$, provided that ${\bold C}$ is finite semisimple. Here $R$ denotes a suitable finite direct product of copies of ${\Bbb K}$. Also, we give a generalization of the Tannaka-Krein duality to tensor categories ${\bold C}$ equipped with such embeddings $\Omega\!: {\bold C} \to {{\bold{bmd}}} ({\it R})$, using a generalization of the notion of a bialgebra, which is called a {\it face algebra} (cf. \cite{subf}-\cite{fb}, and also \cite{BS}, \cite{JurcoSchupp}, \cite{Schauenburgface}). Combining these, we obtain a natural way to construct a face algebra $C(\Omega_0)$ such that $\bold{com} (C (\Omega_0))$ is equivalent to ${\bold C}$, where ${\bold C}$ is an arbitrary finite semisimple tensor category over ${\Bbb K}$. In Section 2, we recall the definition of the ${\EuScript V}$-face algebra and that of the ${\EuScript V}$-{\it dressed coalgebra} (cf. \cite{fb}). The latter enables us to imitate Schauenburg's formulation of the Tannaka-Krein duality (cf. \cite{SchauenburgTD}). In Section 3, we give a generalization of the Tannaka-Krein duality, which we call ${\EuScript V}$-{\it Tannaka duality}. In Section 4, we state the main result. Throughout this letter, we denote by $\bold{Mod} (A)$ and $\bold{mod} (A)$ the category of all left $A$-modules and the category of all finite-dimensional $A$-modules of a ${\Bbb K}$-algebra $A$, respectively, where ${\Bbb K}$ denotes the ground field. In particular, $\bold{Mod} ({\Bbb K})$ denotes the category of all vector spaces over ${\Bbb K}$. Also we denote by $\bold{bmd} (A)$ and $\bold{bmd} (A)$ the category of all $A$-bimodules and the category of all ${\Bbb K}$-finite-dimensional $A$-bimodules, respectively. For a coalgebra $C$ and its right comodule $M$, we use Sweedler's sigma notation, such as $\Delta (c)$ $=$ $\sum_{(c)} c_{(1)} \otimes c_{(2)}$ $(c \in C)$ and $(\mathrm{id} \otimes \Delta) \circ \delta (m)$ $=$ $\sum_{(m)} m_{(0)} \otimes m_{(1)} \otimes m_{(2)}$ $(m \in M)$, where $\Delta$ and $\delta$ denotes the coproduct of $C$ and the coaction of $M$, respectively (cf \cite{Sweedler}). We denote by $\bold{com} (C)$ the category of all finite-dimensional right $C$-comodules. We wish to thank Prof. S. Yamagami for valuable discussions. \section{Preliminary} Let ${\Bbb K}$ be a field and ${\EuScript V}$ a finite non-empty set. We denote by $R = R_{{\EuScript V}}$ the ${\Bbb K}$-linear span of ${\EuScript V}$ equipped with an algebra structure given by $\lambda \mu = \delta_{\lambda \mu} \lambda$ $(\lambda, \mu \in {\EuScript V})$. Let ${\stackrel{\scriptscriptstyle\circ}{R}} = \bigoplus_{\lambda \in {\EuScript V}} {\Bbb K} {\stackrel{\scriptscriptstyle\circ}{\lambda}}$ be a copy of $R$ and $E_{{\EuScript V}} = E = {\stackrel{\scriptscriptstyle\circ}{R}} \otimes R$ the tensor product algebra equipped with a coalgebra structure given by setting \begin{equation} \Delta({\stackrel{\scriptscriptstyle\circ}{\lambda}} \mu) = \sum_{\nu \in {\EuScript V}} {\stackrel{\scriptscriptstyle\circ}{\lambda}} \nu \otimes {\stackrel{\scriptscriptstyle\circ}{\nu}} \mu, \quad \varepsilon ({\stackrel{\scriptscriptstyle\circ}{\lambda}} \mu) = \delta_{\lambda \mu} \label{D(ee)} \end{equation} for each $\lambda, \mu \in {\EuScript V}$. Here we write $\lambda$ and ${\stackrel{\scriptscriptstyle\circ}{\lambda}}$ instead of $\mathrm{id} \otimes \lambda$ and ${\stackrel{\scriptscriptstyle\circ}{\lambda}} \otimes \mathrm{id}$, respectively. Let $\frak H$ be a ${\Bbb K}$-algebra equipped with a coalgebra structure $(\frak H,\Delta,\varepsilon)$ together with an algebra-coalgebra map $E \to \frak H$. We say that $\frak H$ 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} \varepsilon (ab) = \sum_{\nu \in {\EuScript V}} \varepsilon (a\nu) \varepsilon ({\stackrel{\scriptscriptstyle\circ}{\nu}} b) \label{e(ab)} \end{equation} for each $a, b \in \frak H$. Here we denote the images of $\nu$ and ${\stackrel{\scriptscriptstyle\circ}{\nu}}$ of $E$ via the map $E \to \frak H$ simply by $\nu$ and ${\stackrel{\scriptscriptstyle\circ}{\nu}}$, respectively. We call the elements $\nu, {\stackrel{\scriptscriptstyle\circ}{\nu}} \in \frak H$ {\it face idempotents} of $\frak H$. It is known that a bialgebra is an equivalent notion of a ${\EuScript V}$-face algebra with $\mathrm{card}({\EuScript V}) =$ $1$. We say that a linear map $S\!: {\frak H} \to {\frak H}$ is an {\em antipode} of ${\frak H}$, or $({\frak H}, S)$ is a {\em Hopf} ${{\EuScript V}}$-{\em face algebra} if \begin{equation} \sum_{(a)} S(a_{(1)})a_{(2)} = \sum_{\nu \in {\EuScript V}} \varepsilon (a \nu) \nu, \quad \sum_{(a)} a_{(1)}S(a_{(2)}) = \sum_{\nu \in {\EuScript V}} \varepsilon (\nu a){{\stackrel{\scriptscriptstyle\circ}{\nu}}}, \label{S(a)a} \end{equation} \begin{equation} \sum_{(a)} S(a_{(1)})a_{(2)}S(a_{(3)}) = S (a) \label{S(a)aS(a)} \end{equation} for each $a \in \frak H$. An antipode of a ${\EuScript V}$-face algebra is an antialgebra-anticoalgebra map, which satisfies \begin{equation} S({\stackrel{\scriptscriptstyle\circ}{\lambda}} \mu) = {\stackrel{\scriptscriptstyle\circ}{\mu}} \lambda \quad (\lambda, \mu \in {{\EuScript V}}). \label{S(ee)} \end{equation} The antipode of a ${\EuScript V}$-face algebra is unique if it exists. Let $M$ be a right comodule of a ${\EuScript V}$-face algebra $\frak H$. We define an $R$-bimodule structure on $M$ by \begin{equation} \lambda m \mu = \sum_{(m)} m_{(0)} \varepsilon (\lambda m_{(1)} \mu) \quad (m \in M,\, \lambda, \mu \in {\EuScript V}). \label{} \end{equation} Let $N$ be another $\frak H$-comodule. We define an $\frak H$-comodule structure on $M \otimes_{R} N$ by \begin{equation} m \otimes_R n \mapsto \sum_{(m), (n)} ( m_{(0)} \otimes_R n_{(0)} ) \otimes m_{(1)} n_{(1)} \quad (m \in M,\, n \in N). \label{} \end{equation} The category $\bold{com} (\frak H)$ becomes a monoidal category via this operation. If, in addition, $\frak H$ has a bijective antipode, then $\bold{com} (\frak H)$ is rigid and the left dual object of $M \in \mathrm{ob} \bold{com} (\frak H)$ is given by \begin{gather} M^{\vee} = \mathrm{Hom}_{{\Bbb K}} (M, {\Bbb K}), \nonumber\\ \sum_{(m^{\vee})} \langle m^{\vee}_{(0)},\, m \rangle m^{\vee}_{(1)} = \sum_{(m)} \langle m^{\vee},\, m_{(0)} \rangle S(m_{(1)}) \nonumber\\ (m \in M,\, m^{\vee} \in M^{\vee}). \label{Mveedef} \end{gather} Let $C$ be a coalgebra equipped with $E$-bimodule structure. We say that $C$ is a ${\EuScript V}$-{\it dressed coalgebra} if $C$ satisfies \begin{equation} \label{D(eeaee)} \Delta ({\stackrel{\scriptscriptstyle\circ}{\lambda}} \mu c {\stackrel{\scriptscriptstyle\circ}{\lambda}}{}^{\prime} \mu{}^{\prime}) = \sum_{(c)} {\stackrel{\scriptscriptstyle\circ}{\lambda}} c_{(1)} {\stackrel{\scriptscriptstyle\circ}{\lambda}}{}^{\prime} \otimes \mu c_{(2)} \mu{}^{\prime}, \end{equation} \begin{equation} \sum_{(c)} \lambda c_{(1)} \mu \otimes c_{(2)} = \sum_{(c)} c_{(1)} \otimes {\stackrel{\scriptscriptstyle\circ}{\lambda}} c_{(2)} {\stackrel{\scriptscriptstyle\circ}{\mu}}, \label{eae*a} \end{equation} \begin{equation} \varepsilon ({\stackrel{\scriptscriptstyle\circ}{\lambda}} c {\stackrel{\scriptscriptstyle\circ}{\mu}}) = \varepsilon (\lambda c \mu) \label{e(ee)} \end{equation} for each $c \in C$ and $\lambda, \mu, \lambda', \mu' \in {\EuScript V}$. For a ${\EuScript V}$-dressed coalgebra $C$, we define another ${\EuScript V}$-dressed coalgebra $C^{\mathrm{bop}}$ to be the opposite coalgebra of $C$ equipped with $E$-bimodule structure given by ${\stackrel{\scriptscriptstyle\circ}{\lambda}} \mu \otimes c \otimes {\stackrel{\scriptscriptstyle\circ}{\lambda}} {}^{\prime} \mu^{\prime}$ $\mapsto$ ${\stackrel{\scriptscriptstyle\circ}{\mu}}{}^{\prime} {\lambda}^{\prime} c {\stackrel{\scriptscriptstyle\circ}{\mu}} \lambda$ $(\lambda, \mu, \lambda {}^{\prime}, \mu^{\prime} \in {\EuScript V})$. Let $D$ be another ${\EuScript V}$-dressed coalgebra. A coalgebra map $f$ from $C$ to $D$ is called a {\it map of ${\EuScript V}$-dressed coalgebras} if it is an $E$-bimodule map. The tensor product $E$-bimodule $C {\otimes}_{E} D$ becomes a ${\EuScript V}$-dressed coalgebra via \begin{equation} \label{D(ctensEd)} \Delta (c {\otimes}_{E} d) = \sum_{(c), (d)} (c_{(1)} {\otimes}_{E} d_{(1)}) \otimes (c_{(2)} {\otimes}_{E} d_{(2)}), \end{equation} \begin{equation} \label{cu(ctensEd)} \varepsilon (c {\otimes}_{E} d) = \sum_{\nu \in {\EuScript V}} \varepsilon (c \nu) \varepsilon (\nu d). \quad \end{equation} The category $\bold{DrCoalg}_{{\EuScript V}}$ of all ${\EuScript V}$-dressed coalgebras becomes a monoidal category via this operation with unit object $E$. It is easy to see that a monoid in this category is simply a ${\EuScript V}$-face algebra. \section{The ${\EuScript V}$-Tannaka duality} For each $M \in \mathrm{ob} \bold{bmd} ({\it R})$ and $N \in \mathrm{ob} {\bold{bmd}$, we regard $N \otimes M$ as an object of $\bold{Bmd} (E)$ via \begin{equation} {\stackrel{\scriptscriptstyle\circ}{\lambda}} \mu (n \otimes m) {\stackrel{\scriptscriptstyle\circ}{\lambda}}{}^{\prime} \mu{}^{\prime} = {\stackrel{\scriptscriptstyle\circ}{\lambda}} n {\stackrel{\scriptscriptstyle\circ}{\lambda}} {}^{\prime} \otimes \mu m \mu{}^{\prime} \quad (n \in N, m \in M, \lambda, \mu, \lambda {}^{\prime}, \mu{}^{\prime} \in {\EuScript V}). \end{equation} For each $M \in \mathrm{ob} {{\bold{bmd}}} ({\it R})$, we define $M^*$ to be the ${\Bbb K}$-linear dual of $M$ equipped with ${\stackrel{\scriptscriptstyle\circ}{R}}$-bimodule structure given by \begin{equation} \langle {\stackrel{\scriptscriptstyle\circ}{\lambda}} m^* {\stackrel{\scriptscriptstyle\circ}{\mu}},\, m \rangle = \langle m^*,\, \lambda m \mu \rangle \quad (m \in M, m^* \in M^*, \lambda, \mu \in {\EuScript V}). \end{equation} Also, we define $M^{\vee}$ to be the ${\Bbb K}$-linear dual of $M$ equipped with $R$-bimodule structure given by \begin{equation} \langle \lambda m^{\vee} \mu,\, m \rangle = \langle m^{\vee},\, \mu m \lambda \rangle \quad (m^{\vee} \in M^{\vee},\, m \in M,\, \lambda, \mu \in {\EuScript V}). \end{equation} We say that a pair $({\bold C}, \Omega)$ is a {\it category with ${\EuScript V}$-face} if ${\bold C}$ is an essentially small category and $\Omega$ is a functor from ${\bold C}$ to ${{\bold{bmd}}} ({\it R})$. We say that a pair $(F, \zeta)$ is a {\it map of categories with ${\EuScript V}$-faces} from $({\bold C}, \Omega)$ to $({\bold C}^{\prime}, \Omega^{\prime})$ if $F\!: {\bold C} \to {\bold C}^{\prime}$ is a functor and $\zeta\!:\Omega \cong \Omega^{\prime} \circ F$ is a natural isomorphism. For categories with ${\EuScript V}$-faces $({\bold C}_1, \Omega_1)$, $({\bold C}_2, \Omega_2)$, we denote the category ${\bold C}_1 \times {\bold C}_2$ with ${\EuScript V}$-face $\Omega_1 \otimes_R \Omega_2\!:$ $(X_1, X_2)$ $\mapsto $ $\Omega_1 (X_1) \otimes_R \Omega_2 (X_2)$ $(X_1 \in \mathrm{ob} {\bold C}_1, X_2 \in \mathrm{ob} {\bold C}_2)$ by $({\bold C}_1, \Omega_1) \times ({\bold C}_2, \Omega_2)$. For a category with ${\EuScript V}$-face $({\bold C}, \Omega)$, we denote by $C ( \Omega ) = C({\bold C}, \Omega)$ the {\it coend} of $\Omega^* \otimes \Omega$, where $\Omega^* \otimes \Omega$ is viewed as a functor from ${\bold C}^{\mathrm{op}} \times {\bold C}$ to $\bold{Bmd} (E)$ (cf. \cite{Maclane}). By definition, these exists a unique family of maps $\kappa_X = \kappa_X^{\Omega}\!:$ $\Omega (X)^* \otimes \Omega (X) \to C(\Omega)$ $(X \in \mathrm{ob} {\bold C})$ in $\bold{Bmd} (E)$ such that (i) $\{\kappa_X \}_X$ is a dinatural transformation, that is, the diagram \begin{equation} \begin{CD} \Omega (Y)^* \otimes \Omega (X) @> \mathrm{id} \otimes \Omega (f)>> \Omega (Y)^* \otimes \Omega (Y)\\ @V\Omega (f)^* \otimes \mathrm{id}VV @VV\kappa_YV \\ \Omega (X)^* \otimes \Omega (X) @>\kappa_{X}>> C (\Omega) \end{CD} \end{equation} is commutative for each $f\!: X \to Y$ and that (ii) for another dinatural transformation $\{ \iota_X\!:$ $\Omega (X)^* \otimes \Omega (X) \to M \}_X$, there exists a unique map $h\!: C(\Omega) \to M$ such that $\iota_X = h \circ \kappa_X$ for each $X \in \mathrm{ob} {\bold C}$. By the explicit description of the coend via direct sums and a coequalizer, $C (\Omega)$ agrees with the coend of $\bar{\Omega}^* \otimes \bar{\Omega}\!:$ ${\bold C}^{\mathrm{op}} \times {\bold C}$ $\to$ $\bold{Mod} ({\Bbb K})$ as a vector space, where $\bar{\Omega}$ denotes the composition of $\Omega$ with the forgetful functor ${{\bold{bmd}}} ({\it R}) \to \bold{mod} ({\Bbb K})$. Hence $C ( \Omega)$ becomes a coalgebra (see e.g. Schauenburg \cite{SchauenburgTD} p. 29). For $X \in \mathrm{ob} {\bold C}$, we define the map $\delta_X = \delta_X^{\Omega}\!:$ $\Omega (X) \to \Omega (X) \otimes C (\Omega)$ by \begin{equation} \delta_X (m) = \sum_i m_i \otimes \kappa_X (m^i \otimes m) \quad (m \in \Omega (X)), \end{equation} where $\{ m_i \}$ denotes a basis of $\Omega (X)$ and $\{ m^i \}$ denotes its dual basis. Then $\delta_X$ gives a right $C(\Omega)$-comodule structure on $\Omega (X)$, and satisfies \begin{gather} \delta_X (\lambda m \mu) = \sum_{(m)} m_{(0)} \otimes \lambda m_{(1)} \mu,\\ \sum_{(m)} \lambda m_{(0)} \mu \otimes m_{(1)} = \sum_{(m)} m_{(0)} \otimes {\stackrel{\scriptscriptstyle\circ}{\lambda}} m_{(1)} {\stackrel{\scriptscriptstyle\circ}{\mu}} \end{gather} for each $m \in \Omega (X)$ and $\lambda, \mu \in {\EuScript V}$. Using these, we see that $C(\Omega)$ is a ${\EuScript V}$-dressed coalgebra. For example, if ${\bold C} = {\bold 1}$ has the only one object $I$ and the only one morphism, and if $\Omega = R$ sends $I$ to $R$, then, $C(\Omega)$ is isomorphic to $E$ as a ${\EuScript V}$-dressed coalgebra. For categories with ${\EuScript V}$-faces $({\bold C}_1, \Omega_1)$ and $({\bold C}_2, \Omega_2)$, there exists a unique isomorphism $\phi_2\!: C({\bold C}_1, \Omega_1) {\otimes}_{E} C({\bold C}_2, \Omega_2)$ $\cong$ $C(({\bold C}_1, \Omega_1) \times ({\bold C}_2, \Omega_2))$ of ${\EuScript V}$-dressed coalgebras such that \begin{equation*} \begin{CD} (\Omega_1 (X_1)^* \otimes \Omega_1 (X_1)) \otimes_E (\Omega_2 (X_2)^* \otimes \Omega_2 (X_2)) @>\kappa^{\Omega_1}_{X_1} \otimes_E \kappa^{\Omega_2}_{X_2}>> C (\Omega_1) {\otimes}_{E} C (\Omega_2) \\ % @V{\cong}VV @VV{\phi_2}V\\ % (\Omega_1 (X_1) \otimes_R \Omega_2 (X_2))^* \otimes (\Omega_1 (X_1) \otimes_R \Omega_2 (X_2)) @>{(\kappa^{\Omega_1 \otimes_R \Omega_2})_{(X_1, X_2)}}>> C(\Omega_1 \otimes_R \Omega_2) \end{CD} \end{equation*} is commutative for each $X_1 \in \mathrm{ob} {\bold C}_1$ and $X_2 \in \mathrm{ob} {\bold C}_2$. Suppose ${\bold C}$ is a ${\Bbb K}$-linear abelian category and $\Omega$ is a faithful ${\Bbb K}$-linear exact functor. Then applying a primitive version of the Tannaka-Krein duality theorem to $\bar{\Omega}$, we see that $\Omega$ gives a categorical equivalence from ${\bold C}$ onto $\bold{com} (C(\Omega))$ (see e.g. Schauenburg \cite{SchauenburgTD} p. 39). For a map $(F, \zeta)\!: ({\bold C}, \Omega) \to ({\bold C}^{\prime}, \Omega^{\prime})$ of categories with ${\EuScript V}$-faces, there exists a unique map $C (F, \zeta)\!: C (\Omega) \to C (\Omega^{\prime})$ of ${\EuScript V}$-dressed coalgebras such that the following diagram is commutative for each $X \in \mathrm{ob} {\bold C}$: \begin{equation} \begin{CD} \Omega (X) @>\delta^{\Omega}_X>> \Omega (X) \otimes C (\Omega) \\ @V\zeta_XVV @VV\zeta_X \otimes C(F, \zeta)V \\ \Omega^{\prime} (F(X)) @>\delta^{\Omega^{\prime}}_{F(X)}>> \Omega^{\prime} (F(X)) \otimes C (\Omega^{\prime}). \end{CD} \end{equation} For another map $(F^{\prime}, \zeta^{\prime})\!: ({\bold C}^{\prime}, \Omega^{\prime})$ $\to$ $({\bold C}^{\prime \prime}, \Omega^{\prime \prime})$ of categories with ${\EuScript V}$-face, we have $C((F^{\prime}, \zeta^{\prime}) \circ (F, \zeta)) =$ $C(F^{\prime}, \zeta^{\prime}) \circ C(F, \zeta)$, where $(F^{\prime}, \zeta^{\prime}) \circ (F, \zeta)$ denotes the composition $(F^{\prime} \circ F, \{ \zeta^{\prime}_{F(X)} \circ \zeta_X \}_X))$. Let $({\bold C}, \Omega)$ be a category with ${\EuScript V}$-face such that ${\bold C}$ $=$ $({\bold C}, \otimes)$ is a monoidal category with unit object $I$ and that $\Omega$ is a monoidal functor with natural isomorphism $(\varphi_2)_{XY}\!: \Omega(X) \otimes_R \Omega(Y) \cong \Omega(X \otimes Y)$ and with isomorphism $\varphi_0\!: R \cong \Omega (I)$. Since $(\otimes, \varphi_2)\!:$ $({\bold C}, \Omega) \times ({\bold C}, \Omega) \to ({\bold C}, \Omega)$ is a map of categories with ${\EuScript V}$-faces, we obtain a map $m = C (\otimes, \varphi_2) \circ \phi_2\!: C(\Omega) \otimes_E C(\Omega) \to C(\Omega)$ of ${\EuScript V}$-dressed coalgebras. Similarly, considering $(I, \varphi_0)$ as a map of categories with ${\EuScript V}$-faces from $({\bold 1}, R)$ to $({\bold C}, \Omega)$, we obtain a map $\eta = C (I, \varphi_0)$ of ${\EuScript V}$-dressed coalgebras from $E$ to $C(\Omega)$. The triple $(C(\Omega), m, \eta)$ becomes a monoid in $\bold{DrCoalg_{{\EuScript V}}}$, and therefore it gives a ${\EuScript V}$-face algebra. Moreover, $\Omega$ gives a monoidal functor from ${\bold C}$ to $\bold{com}(C(\Omega))$. Next, suppose that each object $X$ of ${\bold C}$ has a left dual $X^{\vee}$. Using the uniqueness of the left dual of $\Omega (X)$ in ${{\bold{bmd}}} ({\it R})$, we obtain a natural isomorphism $j_X\!: \Omega (X)^{\vee} \to \Omega (X^{\vee})$ in a natural manner. Since $({}^{\vee}, j)\!: ({\bold C}^{\mathrm{op}}, \Omega^{\vee}) \to ({\bold C}, \Omega)$ is a map of categories with ${\EuScript V}$-faces, there exists a unique map $S\!: C (\Omega)^{\mathrm{bop}} \to C (\Omega)$ of ${\EuScript V}$-dressed coalgebras such that \begin{equation} \begin{CD} \Omega (X)^{\vee} @>{\mathrm{tw} \circ (\delta^{\Omega}_X)^{\flat}}>> \Omega (X)^{\vee} \otimes C (\Omega)^{\mathrm{bop}} \\ % @V{j_X}VV @VV{j_X \otimes S}V \\ % \Omega (X^{\vee}) @>\delta^{\Omega}_{X^{\vee}}>> \Omega (X^{\vee}) \otimes C (\Omega) \end{CD} \end{equation} is commutative for each $X \in \mathrm{ob} {\bold C}$, where \begin{gather} (\delta^{\Omega}_X)^{\flat}(m^{\vee}) = \sum_i \sum_{(m_i)} \langle m^{\vee},\, {m_i}_{(0)} \rangle {m_i}_{(1)} \otimes m^i,\\ \mathrm{tw} (c \otimes m^{\vee}) = m^{\vee} \otimes c \quad (m^{\vee} \in \Omega (X)^{\vee}, c \in C(\Omega)^{\mathrm{bop}}). \end{gather} The map $S$ gives an antipode of $C(\Omega)$. Let $({\bold C}, \otimes)$ be a monoidal category with unit object $I$. We say that ${\bold C}$ is a {\it tensor category} over a field ${\Bbb K}$ if ${\bold C}$ is a ${\Bbb K}$-linear abelian category and $\otimes$ is a bi-additive functor. As the way stated above, the Tannaka-Krein duality is generalized as follows. \begin{thm} Let ${\bold C}$ be an essential small tensor category over ${\Bbb K}$ and $\Omega\!: {\bold C} \to \bold{bmd}(R_{{\EuScript V}})$ a faithful ${\Bbb K}$-linear exact monoidal functor. Then the coend $C (\Omega)$ becomes a ${\EuScript V}$-face algebra and $\Omega$ gives an equivalence ${\bold C}$ $\cong$ $\bold{com}(C(\Omega))$ of tensor categories. If ,in addition, ${\bold C}$ is rigid, then $C (\Omega)$ has a bijective antipode. \end{thm} \noindent {\it Remark}. $\!$ In \cite{fa}, we introduced the notion of an $R$-{\it face algebra} for each commutative separable algebra $R$ over a field ${\Bbb K}$. Similarly to the``${\EuScript V}$-Tannaka'' duality stated above, we also have an $R$-Tannaka duality for arbitrary $R$. \section{The canonical fiber functor} A tensor category ${\bold C}$ is called a {\it split finite semisimple category} if there exists a finite set of objects $\{ L_{\lambda} | \lambda \in {\EuScript V}\}$ such that each object of ${\bold C}$ is isomorphic to a finite direct sum of $L_{\lambda}$'s and that these objects satisfy the following Schur's lemma; \begin{equation} {\bold C} (L_{\lambda}, L_{\mu}) \cong {\Bbb K} \delta_{\lambda \mu} \mathrm{id}_{L_{\lambda}}. \label{SchurLem} \end{equation} Let ${\bold C}$ be a split finite semisimple tensor category. For each object $X$ of ${\bold C}$, we define a bimodule $\Omega_0 (X)$ of $R = R_{{\EuScript V}}$ by setting \begin{equation} \label{Omega0Xdef} \lambda \Omega_0 (X) \mu = {\bold C} (L_{\mu}, L_{\lambda} \otimes X) \quad (\lambda, \mu \in {\EuScript V}). \end{equation} Then $\Omega_0$ gives a faithful ${\Bbb K}$-linear exact functor from ${\bold C}$ to ${{\bold{bmd}}} ({\it R})$ by setting \begin{equation} \label{Omega0fdef} \Omega_0 (f) (k) = ({\mathrm{id}}_{L_{\lambda}} \otimes f) \circ k \end{equation} for each $f \in {\bold C} (X, Y)$ and $k \in \lambda \Omega_0 (X) \mu$ $=$ ${\bold C} (L_{\mu}, L_{\lambda} \otimes X)$. Moreover $\Omega_0$ gives a ${\Bbb K}$-linear exact faithful monoidal functor by setting \begin{equation*} (\varphi_2)_{X,Y}:\! \Omega_0 (X) \otimes_R \Omega_0 (Y) \to \Omega_0 (X \otimes Y) ; \end{equation*} \begin{gather} \label{phi2def} f \otimes g \mapsto a_{L_{\lambda} X Y} \circ (f \otimes \mathrm{id}_Y) \circ g \nonumber\\ (f \in \lambda \Omega_0 (X) \nu,\, g \in \nu \Omega_0 (Y) \mu,\, \lambda, \mu, \nu \in {\EuScript V}), \end{gather} \begin{equation} \label{phi0def} \varphi_0:\! R \to \Omega_0 (I); \quad \lambda \mapsto r_{L_{\lambda}}^{-1} \quad (\lambda \in {\EuScript V}), \end{equation} where $a_{XYZ}:\! (X \otimes Y) \otimes Z$ $\to$ $X \otimes (Y \otimes Z)$ and $r_X:\! X \otimes I \to X$ denote the associativity constraint and the right unit constraint of ${\bold C}$, respectively. Thus, applying the ${\EuScript V}$-Tannaka duality to $({\bold C}, \Omega_0)$, we obtain the following {\it canonical Tannaka duality}. \begin{thm} For each split finite semisimple category ${\bold C}$, there exists a natural way to construct a ${\EuScript V}$-face algebra $C(\Omega_0)$ and an equivalence $\Omega_0\!:{\bold C} \cong \bold{com} (C(\Omega_0))$ of tensor categories. \end{thm} \noindent {\it Example}. $\!$ Let $G$ be a finite group. The tensor category $\bold{com} ({\Bbb C} [G])$ is finite semisimple and $\{ {\Bbb C} g | g \in G \}$ gives a complete set of representatives of simple objects of $\bold{com} ({\Bbb C} [G])$. The canonical Tannaka dual of this category has a basis $\{ e^a_b [g] | a,b,g \in G \}$ which satisfies the following relations: \begin{equation} \label{eijpeklq} e^a_b [g] e^c_d [h] = \delta_{ag, c} \delta_{bg, d} e^a_b [gh], \end{equation} \begin{equation} \Delta( e^a_b [g] ) = \sum_{c \in G} e^a_c [g] \otimes e^c_b [g], \quad \varepsilon (e^a_b [g]) = \delta_{ab}, \end{equation} \begin{equation} \stackrel{\scriptscriptstyle\circ}{a}\, =\, \sum_{b \in G} e^a_b [1], \quad b\, =\, \sum_{a \in G} e^a_b [1], \end{equation} \begin{equation} S (e^a_b [g]) = e^{bg}_{ag} [g^{-1}], \end{equation} where $a,b,c,d,g,h \in G$.
\section{Introduction} The investigations of nuclear parity violation, both theoretical and experimental, have already a long history. New light on this problem is shed by observation of the nuclear anapole moment (AM) of $^{133}$Cs in atomic experiment~\cite{wi}. The result of this experiment is in a reasonable quantitative agreement with the theoretical predictions, starting with~\cite{fk,fks}, if the so-called ``best values''~\cite{ddh} are chosen for the parameters of P odd nuclear forces. The AM is a rather peculiar multipole in the following sense (for a more detailed discussion see, for instance,~\cite{kh}). The interaction of a charged probe particle with an anapole moment is of a contact nature. Therefore, for instance, the interaction of the electron with the nucleon AM, being on the order of $\mbox{${\alpha}$} G$, cannot be distinguished in general case from other electromagnetic radiative corrections to the weak electron-nucleon interaction due to the neutral currents. And in a gauge theory of electroweak interactions only the total scattering amplitude, i.e., the sum of all diagrams on the order of $\mbox{${\alpha}$} G$, is gauge-invariant, independent of the gauge choice for the Green's functions of heavy vector bosons (here $\mbox{${\alpha}$}$ is the fine-structure constant, $G$ is the Fermi weak interaction constant). No wonder that, generally speaking, the AM of an elementary particle or a nucleus is not gauge-invariant, i.e., physically well-defined, quantity. However, there is a special case where the anapole moment has a real independent physical meaning. In heavy nuclei, of course $^{133}$Cs included, the AM is enhanced $\sim A^{2/3}$~\cite{fks} ($A$ is the atomic number), as distinct from common radiative corrections. By the way, it means that there is an intrinsic limit for the relative accuracy, $\sim A^{-2/3}$, with which the AM of a heavy nucleus can be defined at all. For $^{133}$Cs this limiting accuracy is about 4\%. There is one more object, the deuteron, whose anapole moment could make sense for a sufficiently large P odd $\pi$NN constant~\cite{fk}. The deuteron is a loosely bound system of a relatively simple structure. Therefore, there are all the reasons to believe that its AM is induced mainly by the P odd $\pi$-meson exchange, pion being the lightest possible mediator of the nucleon-nucleon weak interaction. The problem of the deuteron AM was discussed phenomenologically in~[2,6-8]. In the present work the deuteron AM, as induced by the P odd $\pi$-meson exchange, is explicitly expressed through the P odd $\pi$NN coupling constant. The same problem was considered in recent paper~\cite{ss}. The result of the original version of~\cite{ss} was much smaller than ours because a leading contribution, that of the isovector magnetic moment of the nucleon, was omitted in it. After acquaintance with the preprint of the present work, the authors of~\cite{ss} corrected their result (see their revised preprint~\cite{ss}, Section VI. Erratum and Addendum). On the other hand, under the influence of~\cite{ss}, we have refined our own calculations with the results presented below. The obtained result for the deuteron AM is singular as $1/m_{\pi}$ in the limit $m_{\pi} \to 0$ (of course, when going over to this limit, one should keep the deuteron radius larger than the Compton wave length of the pion). Being combined with the radiative corrections to the weak electron -- deuteron scattering amplitude~\cite{ms}, which are regular in $m_{\pi}$, our calculations result in a sufficiently accurate value for the corresponding effective constant $C_{2d}$. We also calculate here the P odd and T odd electromagnetic moments of the deuteron. \bigskip \section{The deuteron anapole moment} It is convenient to start the discussion with the nucleon AM in the chiral limit. It was shown in 1980 by A.I. Vainshtein and one of the authors (I.Kh.) to be given in this limit by the diagrams 1 and 2. \begin{figure}[h] \unitlength 1cm \begin{picture}(15,5) \epsfxsize 20cm \put(-2,-18.5){\epsfbox{ptfig.eps}} \end{picture} \end{figure} The circle on the nucleon lines refers to the usual strong interaction $\pi$NN vertex (coupling constant $g\sqrt{2}$), the cross describes the P odd weak $\pi$NN interaction (coupling constant $\bar{g}\sqrt{2}$). The result for the nucleon AM is \begin{equation}\label{nam} \mbox{${\bf a}$}_N\,=\,\mbox{${\bf a}$}_p\,=\,\mbox{${\bf a}$}_n\,=\,-\,{e g \bar{g} \over 12 m_p m_{\pi}}\, \left(1-\,{6 \over \pi}\,{m_{\pi} \over m_p}\ln{m_p \over m_{\pi}}\right) \,\mbox{\boldmath $\sigma$}. \end{equation} The diagrams discussed lead to the same result for a proton and neutron since under the permutation $p \leftrightarrow n$ the strong coupling constant $g$ does not change, and the weak one $\bar{g}$ changes sign together with the charge $e$ of the $\pi$-meson (we assume $e>0$, exact definitions of the strong and weak interaction Lagrangians and coupling constants $g$ and $\bar{g}$ are given below). Being the only contribution to the nucleon AM, which is singular in $m_{\pi}$, the result (\ref{nam}) is gauge-invariant. In this respect, it has a physical meaning. Unfortunately, in spite of the singularity in $m_{\pi}$, the corresponding contribution to the electron-nucleon scattering amplitude is small numerically as compared to other radiative corrections to the weak scattering amplitude. Indeed, the radiative corrections to the effective constants $C_{2p,n}$ of the proton and neutron axial neutral-current operators $G / \sqrt 2\,C_{2p,n}\mbox{\boldmath $\sigma$}_{p,n}$ are~\cite{ms} \begin{equation}\label{cpn} C^r_{2p}\,=\,0.032\pm 0.030\,, \quad C^r_{2n}\,=\,-0.018\pm 0.030\,. \end{equation} In the same units $G/\sqrt 2$, the effective axial constants induced by the electromagnetic interaction with the proton and neutron anapole moments (\ref{nam}), is \[ C^a_{p,n}\,=\,-\,\mbox{${\alpha}$} a_N\,(|e|G/\sqrt 2)^{-1}\,= \,0.07\times 10^5 \bar{g}. \] At the ``best value'' $\bar{g}= 3.3\times 10^{-7}$ (strongly supported by the experimental result for the $^{133}$Cs anapole moment) we obtain \begin{equation}\label{cna} C^a_{p,n}\,=\, 0.002. \end{equation} With this value being much less than both central points and error bars in (\ref{cpn}), the notion of the nucleon AM practically has no physical meaning. This is why the result (\ref{nam}) was never published by the authors. It is quoted in book~\cite{kh} (without the logarithmic term) just as a theoretical curiosity. The logarithmic term in the nucleon AM is discussed in~\cite{hhm}. However, the situation with the deuteron AM is quite different. Not only the proton and neutron AMs add up here. The isovector part of the radiative corrections is much smaller than the individual contributions $C^r_{2p}$ and $C^r_{2n}$, and is calculated with much better accuracy~\cite{ms}: \begin{equation}\label{cdr} C^r_{2d}\,=\,C^r_{2p}+C^r_{2n}\,=\,0.014\pm 0.003\,. \end{equation} Moreover, there is the already mentioned, qualitatively new contribution, due to the isovector magnetic moment of the nucleon, which dominates numerically the deuteron AM. Thus $a_d$ acquires a real physical meaning. Let us go over now to the problem itself. The Lagrangians of the strong $\pi$NN interaction and of the weak P odd one, $L_s$ and $L_w$, respectively, are well-known: \begin{equation}\label{s} L_s\,=\,g\,[\,\sqrt{2}\,(\overline{p}i\gamma_{5}n \,\pi^+ +\overline{n}i\gamma_{5}p\, \pi^-)\,+(\,\overline{p}i\gamma_{5}p -\overline{n}i\gamma_{5}n)\,\pi^0]; \end{equation} \begin{equation}\label{w} L_w\,=\,\bar{g}\,\sqrt{2}\,i\,(\,\overline{p}n \,\pi^+ -\overline{n}p \,\pi^-). \end{equation} Our convention for $\gamma_5$ is \begin{equation}\label{g5} \gamma_5 = \left(\begin{array}{rr}0 & -I \\ -I & 0 \end{array}\right); \end{equation} the relation between our P odd $\pi$NN constant $\bar{g}$ and the common one $h^{(1)}_{\pi NN}$ is $\bar{g}\sqrt{2}=h^{(1)}_{\pi NN}$. The effective nonrelativistic Hamiltonian of the P odd nucleon-nucleon interaction due to the pion exchange is in the momentum representation \begin{equation}\label{q} V(\mbox{${\bf q}$})\,=\,{2 g \bar{g} \over m_p}\,{(\mbox{${\bf I}$} \mbox{${\bf q}$}) \over m_{\pi}^2 + \mbox{${\bf q}$}^2} (N_1^{\dagger}\tau_{1-}N_1)\,(N_2^{\dagger}\tau_{2+}N_2). \end{equation} Here \[ \mbox{${\bf I}$} = {1 \over 2} (\mbox{\boldmath $\sigma$}_p +\mbox{\boldmath $\sigma$}_n) \] is the deuteron spin, $\mbox{${\bf q}$} =\mbox{${\bf p}$}_1^{\prime}-\mbox{${\bf p}$}_1= -(\mbox{${\bf p}$}_2^{\prime}-\mbox{${\bf p}$}_2) =\mbox{${\bf p}$}_n^{\prime}-\mbox{${\bf p}$}_p= -(\mbox{${\bf p}$}_p^{\prime}-\mbox{${\bf p}$}_n).$ Let us note that the P odd interaction (\ref{q}) which interchanges the proton and neutron, when applied to the initial state $a^{\dagger}_p(\mbox{${\bf r}$}_1)a^{\dagger}_n(\mbox{${\bf r}$}_2)|0\rangle$ transforms it into $a^{\dagger}_n(\mbox{${\bf r}$}_1)a^{\dagger}_p(\mbox{${\bf r}$}_2)|0\rangle =-a^{\dagger}_p(\mbox{${\bf r}$}_2)a^{\dagger}_n(\mbox{${\bf r}$}_1)|0\rangle$. On the other hand, the coordinate wave function of the admixed $^3P_1$ state is proportional to the relative coordinate $\mbox{${\bf r}$}$, which we define as $\mbox{${\bf r}$}_p-\mbox{${\bf r}$}_n$. Therefore, it also changes sign under the permutation $p \leftrightarrow n$. Thus, for the deuteron the P odd potential can be written in the coordinate representation as a simple function of $\mbox{${\bf r}$}=\mbox{${\bf r}$}_p-\mbox{${\bf r}$}_n$ without any indication of the isotopic variables: \begin{equation}\label{r} V(\mbox{${\bf r}$})\,=\,{g \bar{g} \over 2\pi m_p}\,(-i\mbox{${\bf I}$} \cdot \mbox{\boldmath $\nabla$}) \,{\exp{(-m_{\pi}r)} \over r} \end{equation} The above expressions are rather standard. As standard is our sign convention for the coupling constants: $g=13.45$, and $\bar{g} > 0$ for the range of values discussed in~\cite{ddh}. The discussed P odd interaction $V$ generates a contact current $\mbox{${\bf j}$}^c$. To obtain the explicit expression for it, we have to consider $V$ in the presence of the electromagnetic field. Its including modifies the proton momentum: $\mbox{${\bf p}$} \to \mbox{${\bf p}$} - e \mbox{${\bf A}$}$, which results in the shift $\mbox{${\bf q}$} \to \mbox{${\bf q}$} + e\mbox{${\bf A}$}$ in the interaction (\ref{q}). Then in the momentum representation the contact current is \[ \mbox{${\bf j}$}^c(\mbox{${\bf q}$})\,=\,-\,{\partial V(\mbox{${\bf q}$}) \over \partial \mbox{${\bf A}$}}\,=\, -\,{\partial \over \partial \mbox{${\bf A}$}}\;{2g\bar{g} \over m_p}\; {\mbox{${\bf I}$} (\mbox{${\bf q}$} + e \mbox{${\bf A}$}) \over m_{\pi}^2 + (\mbox{${\bf q}$}+ e \mbox{${\bf A}$})^2}\, \] \begin{equation}\label{jc} =\, -\,{2 e g\bar{g} \over m_p}\;\left\{{\mbox{${\bf I}$} \over m_{\pi}^2 + \mbox{${\bf q}$}^2}\, -\,{2 \mbox{${\bf q}$} (\mbox{${\bf I}$} \mbox{${\bf q}$}) \over (m_{\pi}^2 + \mbox{${\bf q}$}^2)^2}\right\}. \end{equation} In the last expression we have neglected the dependence of the contact current on $\mbox{${\bf A}$}$. In the coordinate representation it equals \begin{equation}\label{jcc} \mbox{${\bf j}$}^c(\mbox{${\bf r}$})\,=\,{e g\bar{g} \over 2 \pi m_p}\,\mbox{${\bf r}$}(\mbox{${\bf I}$}\mbox{\boldmath $\nabla$}) {e^{-m_{\pi}r} \over r}\,. \end{equation} Let us derive at first a general structure of the deuteron AM generated by a P odd $np$ interaction, assuming only that the deuteron is a pure $^3S_1$ state, bound by a spherically symmetric potential. We follow here essentially the line of reasoning applied in~\cite{fk} (see also book~\cite{kh}) to the problem of a single proton in a spherically symmetric potential. In this case the formula for the AM operator is~\cite{fk} \begin{equation}\label{sao} \mbox{${\bf a}$}\,=\,{\pi e \over m_p}\,\{\mu_p \mbox{${\bf r}$} \times \mbox{\boldmath $\sigma$}\, -\,{i \over 3}\,[{\bf l}^2, \mbox{${\bf r}$}]\}\,+\,{2\pi \over 3}\,\mbox{${\bf r}$} \times[\mbox{${\bf r}$} \times \mbox{${\bf j}$}^c], \end{equation} with the proton magnetic moment $\mu_p=2.79$. In the case of the deuteron this formula generalizes to \begin{equation}\label{dao} \mbox{${\bf a}$}_d\,=\,{\pi e \over 2 m_p}\, \{\mbox{${\bf r}$} \times (\mu_p\mbox{\boldmath $\sigma$}_p\, -\,\mu_n\mbox{\boldmath $\sigma$}_n)\,-\, {i \over 6}\,[{\bf l}^2, \mbox{${\bf r}$}]\}\,+\,{\pi \over 6}\,\mbox{${\bf r}$} \times[\mbox{${\bf r}$} \times \mbox{${\bf j}$}^c], \end{equation} $\mu_n=-1.91$ is the neutron magnetic moment. Both AM operators (\ref{sao}) and (\ref{dao}) are orthogonal to $\mbox{${\bf r}$}$ (neither of them commutes with $\mbox{${\bf r}$}$, so the orthogonality means here that $\mbox{${\bf a}$}\mbox{${\bf r}$}+\mbox{${\bf r}$}\mbox{${\bf a}$}=0$). Therefore, the contact current (\ref{jcc}) generated by the P odd pion exchange and directed along $\mbox{${\bf r}$}$, does not contribute to the nuclear AM. Let us present now the wave function of the deuteron $^3S_1$ state as $\psi_0(r)\chi$, where $\chi$ is the spin wave function for $I=1$ (we neglect here and below a small $^3D_1$ admixture in the deuteron). If the P odd interaction conserves the total spin $\mbox{${\bf I}$}$ of the deuteron, the $^3P_1$ state admixed by it can be written as $i(\mbox{${\bf I}$} \mbox{${\bf r}$}/r)\psi_1(r)$ (both radial wave functions, $\psi_0(r)$ and $\psi_1(r)$, are spherically symmetric). Simple calculations demonstrate that the deuteron AM, as induced by the operator (\ref{dao}), is in the absence of the contact contribution \begin{equation}\label{da} \mbox{${\bf a}$}_d\,=\,{\pi e \over 3 m_p}\,\left(\mu_p\,-\,\mu_n \,-\,{1 \over 3}\right)\, \int d\mbox{${\bf r}$} r \psi_0(r) \psi_1(r). \end{equation} So, under the assumptions made, the deuteron AM should depend on the universal combination $\;(\mu_p-\mu_n-\,1/3)$. We confine mainly in our calculation to the na\"{\i}ve zero-range approximation (ZRA) for the deuteron wave function: \begin{equation}\label{p00} \psi_0^{(0)}(r)\,=\,\sqrt{{\kappa \over 2\pi}}\;{\exp{(-\kappa r)} \over r}. \end{equation} Here $\kappa=\sqrt{m_p \mbox{${\varepsilon}$}}$; $\mbox{${\varepsilon}$} = 2.23$ MeV is the deuteron binding energy. The P odd correction to the deuteron wave function due to $V(\mbox{${\bf r}$})$ will be found in the common stationary perturbation theory. In the same ZRA the admixed $^3P_1$ states of the continuous spectrum are free. Moreover, we can choose plane waves as the intermediate states since the perturbation $V(\mbox{${\bf r}$})$ selects by itself the $P$-state from the plane wave. Thus obtained first-order correction to the wave function is \begin{equation}\label{p1} \psi_1(\mbox{${\bf r}$})\,=\,\int {d \mbox{${\bf k}$} \over (2\pi)^3}\; {e^{i \mbox{${\bf k}$} \mbox{${\bf r}$}} \over -\mbox{${\varepsilon}$} - k^2/m_p}\; \int d \mbox{${\bf r}$}^{\prime}\,e^{- i \mbox{${\bf k}$} \mbox{${\bf r}$}^{\prime}} V(\mbox{${\bf r}$}^{\prime})\, \psi_0(r^{\prime}). \end{equation} Rather lengthy calculation leads to the following expression for the matrix element of the radius-vector: \begin{equation}\label{vr} \int d \mbox{${\bf r}$} \psi_0(r) \mbox{${\bf r}$} \psi_1(\mbox{${\bf r}$})\, =\,-\,{i \mbox{${\bf I}$} g \bar{g} \over 6 \pi m_{\pi}}\,{1+\xi \over (1+2\xi)^2}, \end{equation} where $\xi = \kappa/m_{\pi}\,=\,0.32$. With this matrix element and the operator (\ref{dao}) one obtains easily the following result for the deuteron AM: \begin{equation}\label{pad} \mbox{${\bf a}$}_d^{(0)}\,=\,-\,{e g\bar{g} \over 6 m_p m_{\pi}} \,{1+\xi \over (1+2\xi)^2}\,\left(\mu_p-\mu_n-\,{1 \over 3}\right)\,\mbox{${\bf I}$}, \end{equation} in accordance with the general formula (\ref{da}). Our overall factor at the structure $\;(\mu_p-\mu_n-1/3)$ is the same as that at $\;(\mu_p-\mu_n)$ in the revised version of~\cite{ss}. However, the corresponding total result obtained in~\cite{ss}, even in it revised version, is not proportional to the universal combination $\;(\mu_p-\mu_n-\,1/3)$. In fact, the range $1/m_{\pi}$ of the P odd interaction (\ref{r}) is quite comparable to the range of the usual nuclear forces. Therefore, it is, strictly speaking, inconsistent to use the zero-range approximation for calculating effects induced by the perturbation $V(\mbox{${\bf r}$})$. Still, numerical estimates made with a model deuteron wave function which has somewhat more realistic properties, indicate that the error introduced by using the ZRA does not exceed 20\%. As to other sources of P violation, different from the pion exchange, there are no reasons to expect that in the case of deuteron their neglect creates a serious error if the P odd $\pi NN$ coupling constant $\bar{g}$ is at least comparable to its ``best value''. It looks reasonable to combine the potential contribution (\ref{pad}) with the additive contribution of the nucleon anapole moments, which according to (\ref{nam}) is \begin{equation}\label{nad} \mbox{${\bf a}$}_d^N\,=\,\mbox{${\bf a}$}_p\,+\,\mbox{${\bf a}$}_n\,=\,-\,{e g\bar{g} \over 6 m_p m_{\pi}}\, \left(1-\,{6 \over \pi}\,{m_{\pi} \over m_p}\ln{m_p \over m_{\pi}}\right) \mbox{${\bf I}$}. \end{equation} In this way we arrive at the final result for the deuteron AM in the chiral limit: \begin{equation}\label{fff} \mbox{${\bf a}$}_d\,=\,-\,{e g\bar{g} \over 6 m_p m_{\pi}}\, \left[\,0.49\,(\mu_p-\mu_n -{1 \over 3})\,+\,0.46\right]\,\mbox{${\bf I}$} \, =\,-2.60\,\,{e g\bar{g} \over 6 m_p m_{\pi}}\,\mbox{${\bf I}$} \,. \end{equation} This result includes all contributions to the P odd amplitude of $ed$-scattering, which are singular in $m_{\pi}$, and thus is gauge-invariant, independent of the gauge choice for the Green's functions of heavy vector bosons. Finally, let us compare the contribution of (\ref{fff}) to the P odd $ed$ scattering amplitude which is due to the usual radiative corrections, nonsingular in $m_{\pi}$. For the deuteron the axial operator looks as follows: \[ {G \over \sqrt 2}\,C_{2d}\,\mbox{${\bf I}$}\,. \] The contributions to the isoscalar axial constant $C_{2d}$ originate from the anapole moment, from usual radiative corrections nonsingular in $m_{\pi}$, and from the admixture of strange quarks in nucleons~\cite{ef}. The magnitude of the $s$-quarks contribution is extremely interesting, but highly uncertain. As to the usual radiative corrections, their contribution to this constant is found in~\cite{ms} with good accuracy (see (\ref{cdr}): $C^r_{2d}\,=\,0.014\pm 0.003$. In the same units $G/\sqrt 2$ the effective axial constant induced by the electromagnetic interaction with the deuteron AM (\ref{fff}) is \begin{equation}\label{cda} C^a_{2d}\,=\,\mbox{${\alpha}$} a_d\,(e G/\sqrt 2)^{-1}\,=\,0.44\times 10^5 \bar{g}. \end{equation} At the ``best value'' $\bar{g}= 3.3\times 10^{-7}$ (strongly supported by the experimental result for the $^{133}$Cs anapole moment) we obtain \begin{equation}\label{ca} C^a_{2d}\,=\,0.014 \pm 0.003. \end{equation} We use here the above estimate of 20\% for the accuracy of our calculation (for given $\bar{g}$). The numbers in (\ref{cdr}) and (\ref{ca}) are quite comparable, and taken together result in the following value of the total effective constant: \begin{equation}\label{caf} C_{2d}\,=\,C^r_{2d}+C^a_{2d}\,=\,0.028 \pm 0.005. \end{equation} We are fully aware of the extreme difficulty of the experimental measurement of the constant $C_{2d}$. However, with such a good accuracy of the theoretical prediction (\ref{caf}), this experiment becomes a source of the valuable information on the P odd $\pi$NN constant and on the $s$-quark content of nucleons. As it was 15 years ago, now again ``$C_{2d}$ seems to be the most interesting parity-violating parameter accessible to atomic-physics experiments''~\cite{ms}, although by rather different reasons. Of course, if necessary the accuracy of our prediction (\ref{ca}) can be improved by using a more detailed and realistic description of the deuteron. On the other hand, the accuracy of radiative corrections (\ref{cdr}) can be also improved, at least by using much more precise modern experimental values of the parameters of the electroweak theory. \bigskip \section{The deuteron P odd, T odd moments} The problem of the deuteron P odd, T odd multipoles: electric dipole, magnetic quadrupole, and the so-called Schiff moment, was treated phenomenologically in~\cite{sfk}. Now we will calculate the electric dipole and magnetic quadrupole moments within the approach applied above to the anapole. As to the Schiff moment, strong cancellations occur when calculating its value for the deuteron~\cite{sfk}. Therefore, one cannot expect reasonable accuracy for it with our ZRA deuteron wave function, and we will not consider here this problem. As distinct from the P odd, T even interaction, there are three independent P odd, T odd effective $\pi$NN Lagrangians. They are conveniently classified by their isotopic properties:\\ \begin{equation}\label{0} \Delta T =0. \quad L_0\,=\,g_0\,[\,\sqrt{2}\,(\overline{p}n \,\pi^+ +\overline{n}p\, \pi^-)\,+(\,\overline{p}p -\overline{n}n)\,\pi^0]; \end{equation} \begin{equation}\label{1} |\Delta T| =1. \quad L_1\,=\,g_1\,\bar{N}N\,\pi^0\,=\,g_1\,(\,\overline{p}p \,+ \,\overline{n}n \,)\,\pi^0);\quad \quad \quad \quad \end{equation} \[ |\Delta T| = 2. \quad L_2\,=\,g_2\,( \,\bar{N}\mbox{\boldmath $\tau$}N\,\mbox{\boldmath $\pi$}\,- \,3\bar{N}\tau^3 N\,\pi^0\,) \quad \quad \quad \quad \quad \quad \] \begin{equation}\label{2} \quad \quad \quad \quad \quad \quad \quad \quad \quad \;\;\; =\,g_2\,[\,\sqrt{2}\,(\overline{p}n \,\pi^+\,+\,\overline{n}p\,\pi^-)\, -\,2\,(\,\overline{p}p \,-\,\overline{n}n \,)\,\pi^0)\,]. \end{equation} Since the possible values of the isotopic spin for two nucleons is $T=0,\,1$ only, the last interaction, with $|\Delta T| = 2$, is not operative in our approach. The effective P odd, T odd proton -- neutron interaction is derived in the same way as in the AM problem. In the momentum representation it looks as follows: \begin{equation}\label{wq} W(\mbox{${\bf q}$})\,=\,{g \over 2 m_p}\,{i\mbox{${\bf q}$} \over m_{\pi}^2 + \mbox{${\bf q}$}^2}\, [\,(\,3g_0 - g_1\,)\,\mbox{\boldmath $\sigma$}_p\,-\, (\,3g_0 + g_1\,)\,\mbox{\boldmath $\sigma$}_n\,]. \end{equation} In the coordinate representation it is \begin{equation}\label{ww} W(\mbox{${\bf r}$})\,=\,{g \over 8\pi m_p}\, [\,(\,3g_0 - g_1\,)\,\mbox{\boldmath $\sigma$}_p\,-\, (\,3g_0 + g_1\,)\,\mbox{\boldmath $\sigma$}_n\,] \,\mbox{\boldmath $\nabla$}\,{e^{-m_{\pi}r} \over r}\,. \end{equation} The calculation of the deuteron EDM ${\bf d}_d$, i.e., of the $e\mbox{${\bf r}$}_p=e\mbox{${\bf r}$}/2$ expectation value, goes along the same lines as that for the anapole moment and results in \begin{equation} {\bf d}\,=\,-\,{egg_1 \over 12\pi m_{\pi}}\, {1+\xi \over (1+2\xi)^2}\,\mbox{${\bf I}$}. \end{equation} The magnetic quadrupole moment (MQM) operator is expressed through the current density $\mbox{${\bf j}$}$ as follows (see, for instance,~\cite{kh,kl}): \begin{equation}\label{m} M_{mn}\,=\,(r_m\mbox{${\varepsilon}$}_{nrs}+ r_n\mbox{${\varepsilon}$}_{mrs})r_r j_s. \end{equation} This expression transforms to \begin{equation}\label{mm} M_{mn}\,=\,{e \over 2m}\left\{\,3\mu\left[\,r_m\sigma_n + r_n\sigma_m -\,{2 \over 3}\,(\mbox{\boldmath $\sigma$}\mbox{${\bf r}$})\right]\,+ 2q (r_m l_n + r_n l_m)\right\}. \end{equation} Here $\mu$ is the total magnetic moment of the particle, $q$ is its charge in the units of $e$. The magnetic quadrupole moment is the expectation value ${\cal M}$ of the operator $M_{zz}$ in the state with the maximum total angular momentum projection $I_z=I$. In our case, due to the spherical symmetry of the deuteron nonperturbed wave function, the orbital contribution to $M_{mn}$ vanishes. The contact current generated by the P and T odd charged pion exchange, here is also directed along $\mbox{${\bf r}$}$, and thus does not contribute to MQM. So, the deuteron magnetic quadrupole moment originates from the spin term in (\ref{mm}). It equals \begin{equation} {\cal M}\,=\,-\,{e g \over 12\pi m_p m_{\pi}}\,{1+\xi \over (1+2\xi)^2}\, [\,(\,3g_0+g_1\,)\,\mu_p + (\,3g_0-g_1\,)\,\mu_n\,)\,]. \end{equation} \bigskip \begin{center} *** \end{center} \bigskip We very much appreciate numerous helpful discussions with V.B. Telitsin. We are grateful also to V.F. Dmitriev and V.V. Sokolov for the interest to the work. The work was supported by the Russian Foundation for Basic Research through Grant No. 98-02-17797, by the Ministry of Education through Grant No. 3N-224-98, and by the Federal Program Integration-1998 through Project No. 274. \newpage
\section{Introduction} Cygnus X-2 is one of the brightest persistent low-mass X-ray binaries. It varies on time scales from milliseconds to months (e.g.\ Kuulkers, van der Klis \&\ Vaughan 1996, Wijnands, Kuulkers \&\ Smale 1996, Wijnands et al.\ 1997a, 1998a). The primary is a neutron star, while the donor star is an A9 subgiant. They orbit each other with a period of $\sim$9.8~days (Cowley, Crampton \&\ Hutchings 1979; Casares, Charles \&\ Kuulkers 1998). The mass accretion rate is high (\.M$\sim$10$^{18}$\,g\,s$^{-1}$), giving rise to near-Eddington X-ray luminosities (see Smale 1998). The source shows the Z source behaviour in X-ray colour-colour diagrams and hardness-intensity diagrams, and associated fast timing ($\mathrel{\hbox{\rlap{\lower.55ex \hbox {$\sim$}$100\,sec) properties that are characteristic for such high X-ray luminosity. It shows type~I X-ray bursts (see Smale 1998, and references therein) and kiloHertz (kHz) quasi-periodic oscillations (QPO; Wijnands et al.\ 1998a). Five other persistent high luminosity neutron stars show similar X-ray behaviour. They are referred to as ``Z'' sources, because of the Z shape of the tracks they trace out in the colour-colour diagram (Hasinger \&\ van der Klis 1989). The limbs of the Z are, from top to bottom, called horizontal branch, normal branch and flaring branch. It is thought that mass-accretion rate increases from the horizontal branch, through the normal branch, to the flaring branch. In the horizontal branch and upper part of the normal branch QPO are present with frequencies varying between $\sim$15 and $\sim$60\,Hz (called horizontal branch QPO or HBO) together with a noise component below $\la$20\,Hz (called low-frequency noise or LFN). On the normal branch different QPO (called normal branch QPO or NBO) are present with frequencies of 5--7\,Hz. In the Z sources Sco\,X-1 and GX\,17+2 the normal branch QPO merge smoothly into flaring branch QPO (called FBO) with frequencies of up to $\sim$20\,Hz on the lower part of the flaring branch . No such flaring branch QPO have been reported in Cyg\,X-2, although $\sim$26\,Hz QPO were seen when Cyg\,X-2 was in the upper part of the flaring branch, during an intensity `dip' (Kuulkers \&\ van der Klis 1995). The Rossi X-ray Timing Explorer (RXTE) has opened up a new window on low-mass X-ray binaries in the millisecond regime. KHz QPO (for a review see e.g.\ van der Klis 1998) have been detected in all 6 Z sources (Van der Klis et al.\ 1996, 1997, Wijnands et al.\ 1997b, 1998a, 1998b; Jonker et al.\ 1998; Zhang, Strohmayer \&\ Swank 1998). The frequency of the kHz QPO increases with increasing mass-accretion rate. Of the Z sources, Cyg\,X-2 displays the most noticeable variations in the X-ray intensity on long time scales (days to months, e.g.\ Smale \&\ Lochner 1992; Wijnands et al.\ 1996; see also Kong, Charles \&\ Kuulkers 1998). These so-called secular variations (which have been empirically divided into three intensity intervals, called low, medium and high intensity state) recur on a time scale of $\sim$78~days and are associated with systematic changes in position and shape of the Z track in the colour-colour and hardness-intensity diagrams (Kuulkers et al.\ 1996; Wijnands et al.\ 1996; 1997a). They are clearly distinct from the process by which the source traces out the Z track itself on a time scale of hours to a day. As the source goes from the medium intensity to high intensity state (or vice versa) the fast timing properties in at least the normal branch change (Wijnands et al.\ 1997a). On one occasion when Cyg\,X-2 was very faint in X-rays, no clear Z pattern was seen in the colour-colour and hardness-intensity diagram, but only a long diagonal branch associated with flaring behaviour which is stronger at higher energies (see Kuulkers et al.\ 1996). Since the fast timing properties of the source in the low intensity state were unknown we proposed to observe the source in this rare state by using the RXTE All Sky Monitor (ASM) to trigger pointed observations. In this paper we report on the results of these observations. \section{Observations and analysis} The RXTE Proportional Counter Array (PCA, Bradt, Rothschild \&\ Swank 1993) obtained Target of Opportunity observations of Cyg\,X-2 on 1996 October 31 05:29--08:00~UTC (orbital phase according to Casares et al.\ 1998: $\phi_{\rm orb}$$\sim$0.48--0.49, where phase zero corresponds to X-ray source superior conjunction), 1997 September 28 09:24--18:13~UTC ($\phi_{\rm orb}$$\sim$0.22--0.26) and 1997 September 29 04:36--13:42~UTC ($\phi_{\rm orb}$$\sim$0.30--0.34), when the ASM rate dropped below $\sim$20\,counts\,s$^{-1}$\,SCC$^{-1}$. Most of the data were collected with all five proportional counter units (PCUs) on, simultaneously with a time resolution of 16\,s (129 photon energy channels, effectively covering 2--60\,keV) and down to 16\,$\mu$s using various timing (event, binned and single-bit) modes covering the 2--60\,keV range. We constructed colour-colour and hardness-intensity diagrams from the 16\,s data using the same energy ranges as Wijnands et al.\ (1998a). The intensity is defined as the 3-PCU count rate in the energy band 2.0--16.0\,keV, whereas the soft and hard colours are defined as the logarithm of the count rate ratios between 3.5--6.4\,keV and 2.0--3.5\,keV and between 9.7--16.0\,keV and 6.4--9.7\,keV, respectively. All count rates were corrected for background. Power density spectra were made from the high time resolution data using 16\,s data stretches, also in the same energy range as Wijnands et al.\ (1998a), i.e., 5.0--60\,keV. In order to study the low-frequency ($\la$100\,Hz) behaviour we fitted the 0.125--256\,Hz power spectra with a constant representing the dead time modified Poisson noise, Lorentzians or exponentially cut-off power laws to describe peaked noise components, and a power law describing the underlying continuum (called very-low-frequency noise or VLFN). To search for kHz QPO we fitted the 256--2048\,Hz power spectra with a function described by a constant and a Lorentzian to describe any QPO. Errors quoted for the power spectral parameters were determined using $\Delta\chi^2$=1. Upper limits were determined using $\Delta\chi^2$=2.71, corresponding to 95\%\ confidence levels. \section{Results} \subsection{Colour-colour and hardness-intensity diagrams} In Fig.~1 we show the colour-colour diagram and in Figs.~2 and 3 the hardness-intensity diagrams of the individual observations (a--c) and combined (d) together with the data points of Wijnands et al.\ (1998a). All our data points correspond to X-ray intensities of $\sim$2000--2500\,counts\,s$^{-1}$ (3 PCUs), so we succeeded in catching the source at its lowest intensity levels. The observations obtained in 1996 October seem to be extensions towards lower intensity on the horizontal branch. This is apparent in both the hard hardness-intensity diagram and colour-colour diagram by comparing with the data of Wijnands et al.\ (1998a). In the soft hardness-intensity diagram, however, the 1996 October data fall slightly below their horizontal branch. The observations obtained in 1997 September, however, can not be immediately placed within the general Z pattern behaviour of the source as defined by the Wijnands et al.\ (1998a) data. In both hardness-intensity diagrams the 1997 September data describe a slightly curved branch, which does not fall on top the earlier data points. In the colour-colour diagram the September 29 observations trace out a curved track; the September 28 observations fall on top of the September 29 data points and on top of the upper part of the normal branch of Wijnands et al.\ (1998a). It looks as if the curved track respresents the lower part of the normal branch and flaring branch but shifted to higher soft and hard colours. However, in the hardness-intensity diagrams no clear indications of ``flaring'' or ``dipping'' behaviour (see Kuulkers et al.\ 1996) can be found. \subsection{Power spectra} \subsubsection{1996 October} The mean power spectrum (5--60\,keV) of the 1996 October data (Fig.~4a; total of $\sim$6.3\,ksec) clearly showed horizontal branch QPO near $\sim$19\,Hz together with a higher harmonic near $\sim$38\,Hz on top of a low-frequency noise component, confirming that the source is in the left end of the horizontal branch. A fit to this power spectrum (LFN + 2 QPO) resulted in a reduced $\chi^2$ of 1.72 for 114 degrees of freedom (dof). This is not a good fit; in fact, a close inspection of the power spectrum reveals that the harmonic is not well fitted with this model. We therefore added another cut-off power-law component with a cut-off near 20\,Hz, i.e.\ a so-called high-frequency noise (HFN) component, which significantly improved the fit at high frequencies: reduced $\chi^2=1.42$ for 112 dof. Moreover, the resulting fit seems to describe the second harmonic much better; the centroid frequency of the harmonic is 37.6$\pm$0.2, compared to 36.5$\pm$0.3 without the high-frequency noise component. The ratio of the harmonic frequency to the QPO frequency is 1.96$\pm$0.01 (compare this with 1.90$\pm$0.01 without the high-frequency noise component). The full-width-at-half-maximum (FWHM) is $\sim$10\,Hz, compared to $\sim$20\,Hz without the high-frequency noise component. The resulting fit to the power spectrum including the high-frequency noise component is shown in Fig.~4a. The QPO and noise components are significant enough to divide the data up into two parts. We computed the S$_{\rm Z}$ values which measure the position along the Z in the hard hardness-intensity diagram, using the Z track of Wijnands et al.\ (1998a). In Table 1 we give the results of fits to the power spectra corresponding to the two selected regions. As Cyg\,X-2 moves further onto the horizontal branch (to lower inferred mass-accretion rate), the frequencies of the horizontal branch QPO and the harmonic decrease, as expected. We found no evidence for kHz QPO with upper limits of $\sim$3.4\%, when fixing the FWHM at 150\,Hz. This is significantly different from earlier observations in the same part of the horizontal branch (see Section 4.3). \subsubsection{1997 September} The variability during both the September 28 and 29 observations was low. In Table 2 and Figs.~4b--d we give the results of the fits to the power spectra for the September 28 and 29 observations. The mean power spectrum of the September 28 observations (total of $\sim$22.6\,ksec) showed weak power-law noise ($\sim$1.3\%, Fig.~4b). Weak QPO near $\sim$40\,Hz are, however, discernable (see inset in Fig.~4b). We fitted these QPO and found that they were significant at the $\sim$4$\sigma$ level as estimated from an F-test for the inclusion of QPO [$\chi^2$/dof=104/97 vs.\ $\chi^2$/dof=130/100] and from the 68\%\ confidence error-scan of the integral power in the $\chi^2$-space, i.e.\ $\Delta\chi^2$=1. Taking into account the number of trials decreases the significance to $\sim$3$\sigma$. A subdivision in the colour-colour and hardness-intensity diagram tracks of this observation did not show significant differences in the power spectral shapes. Since the colour-colour diagram of the September 29 observations (total of $\sim$21.0\,ksec) indicates the presence of two different branches we decided to investigate the power spectra by selecting these branches as indicated by the two regions denoted `A' and `B' in Fig.~1b. The power spectra are displayed in Fig.~4c and Fig.~4d, for regions `A' and `B', respectively. The mean power spectrum for region `A' shows only a power-law noise component (rms $\sim$1.4\%\/), whereas the mean power spectrum for region `B' shows a weak peaked-noise component between $\sim$2--20\,Hz peaking near 6--7\,Hz (rms $\sim$3\%\/), on top of a power-law noise component (rms $\sim$1\%\/). Since the September 28 observation and part of the September 29 observations are parallel to the normal branch of the Wijnands et al.\ (1998a) observations, we investigated the power spectra for normal branch QPO. None were found with upper limits of $\sim$1.3\%\ and $\sim$0.8\%, for the September 28 and 29 observations, respectively, with typical values for the frequency and FWHM of 5.5\,Hz and 2.5\,Hz, respectively. The upper limits on normal branch QPO in regions `A' and `B' of the September 29 observation are $\sim$0.8\%\ and $\sim$1.1\%, respectively. Upper limits on the strength of QPO in the September 29 observations, similar to that found in the Feb 28 observations near $\sim$40\,Hz, are $\sim$1.3\%. We also searched the September 28 and 29 power spectra for the presence of kHz QPO but found none. Upper limits are $\sim$3\%\ and $\sim$2.8\%, for the September 28 and 29 observations, respectively. \section{Discussion} We performed RXTE observations when the ASM indicated that Cyg\,X-2 was at overall low intensities. These were sucessfully performed in October 1996 and September 1997. It appears that we obtained data in two different kinds of low intensity ``states'' on the two occasions. In the next subsections we discuss the two observations separately, and investigate the kHz QPO properties. \subsection{October 1996} During our first observation in October 1996 we found the source in the left part of the horizontal branch, based on the place of the source in the colour-colour diagram and hardness-intensity diagrams and the presence of horizontal branch QPO at $\sim$19\,Hz and their harmonic at twice this frequency. Previous EXOSAT (Hasinger 1987) and RXTE (Focke 1996; Smale 1998) observations already showed horizontal branch QPO in the same frequency range. Assuming that the horizontal/normal-branch vertex is at the same location in the hard hardness-intensity diagram as derived by Wijnands et al.\ (1998a), the horizontal branch QPO frequencies found during the October 1996 observations are lower than expected at the same position in the Z. This most likely indicates that the Z track of our observation was shifted with respect to that of Wijnands et al.\ (1998a). This is supported by the fact that our observation is located below the horizontal branch of Wijnands et al.\ (1998a) in the soft hardness-intensity diagram, similar to what was seen in EXOSAT data by Kuulkers et al.\ (1996). We found evidence for the presence of a cut-off power-law component in the power with a cut-off frequency near 20\,Hz. A similar component has been observed previously in Cyg X-2 (Hasinger \&\ van der Klis 1989, Wijnands et al.\ 1997a) and in other Z sources (Hasinger \&\ van der Klis 1989, Hertz et al.\ 1992, Kuulkers et al.\ 1994, 1997, Kamado, Kitamoto \&\ Miyamoto 1997), and is mostly referred to as high-frequency noise. High-frequency noise is strongest in the horizontal branch. We note, however, that our observed high-frequency noise strength is higher than that reported previously for Cyg\,X-2. This may be due to the higher energy range we investigated compared to that of Hasinger \&\ van der Klis (1989) and Wijnands et al.\ (1997a). The high-frequency noise has been observed to become stronger at higher energies (e.g.\ Dieters \&\ van der Klis 1999). \subsection{September 1997} The September 1997 observations do not show clear Z behaviour in the colour-colour diagram and hardness-intensity diagrams, although we cannot rule out the possibility that a ``complete'' Z was traced out on a longer timescale. However, the September 29 observations show a curved branch in the colour-colour diagram, which might be part of the Z, i.e.\ the lower normal branch and the lower flaring branch, but shifted to higher colour values. Moreover, the September 28 observation is aligned with the normal branch, suggesting it to be the same branch. It is known that the source hardens, i.e.\ the Z-pattern shifts to higher colour values, when it is at overall lower intensities (Kuulkers et al.\ 1996, Wijnands et al.\ 1996). The situation is less clear for the hardness-intensity diagrams. The hardness-intensity diagrams of the September observations are more reminiscent of those reported by Vrtilek et al.\ (1986). Such shapes are seen when the source intensity is at an overall low level (see Kuulkers et al.\ 1996). Both the hardness-intensity diagrams and colour-colour diagram of the September observations do not resemble, however, the large diagonal branch seen with EXOSAT in 1983 (see Kuulkers et al.\ 1996), which also occurred during a low intensity state. For the first time we have been able to examine the rapid variability at low overall intensities. We find that the very-low frequency variability during the RXTE September observations is low, i.e.\ $\sim$1.0--1.4\%\ (0.1--1\,Hz, 5--60\,keV). Such low variability was also found for the very-low-frequency noise in the medium intensity level (1--20\,keV; Wijnands et al.\ 1997a). Our observed very-low-frequency-noise component is unusually flat. Its index is $\alpha$$\sim$0.6--0.7), consistent with extrapolating the observed decrease in index in the normal branch (Wijnands et al.\ 1997a) from the high ($\alpha$$\sim$1.5--1.7) to medium ($\alpha$$\sim$1) intensity level down to the low intensity level. In order to compare our RXTE observations with the 1983 "diagonal branch" observations of Cyg\,X-2 we calculated power spectra of the EXOSAT data using 64-s data stretches. These data were obtained with a 0.25\,s time resolution and no energy information (1--20\,keV; so-called ``I3''-data from the HER3 mode, see e.g.\ Kuulkers 1995). We used all data during which the collimator response was 100\%\ and all detectors were on source (total of $\sim$16.5\,ksec). The resulting 0.02--2\,Hz average power spectrum corrected for instrumental noise, see Berger \&\ van der Klis 1998) can be well ($\chi^2_{\rm red}$ of 1.06 for 42 dof) described by a steep power law ($\alpha$=2.0$\pm$0.2) with 1.9$\pm$0.1\%\ rms (0.01--1\,Hz). Clearly, during the 1983 observations the very-low-frequency noise was much steeper than that during the September 1997 observations. We found evidence for weak ($\sim$2\%, 5--60\,keV) QPO at $\sim$40\,Hz during the September 28 observations. Since it has been observed in Cyg\,X-2 (Wijnands et al.\ 1997a) that the horizontal branch QPO frequency (and rms amplitude) {\em decreases} from the horizontal/normal-branch connection ($\sim$55\,Hz) down the normal branch (down to $\sim$45\,Hz), we can interpret our observed QPO as horizontal branch QPO occurring in the lower/middle part of the normal branch. The fact that the horizontal branch QPO on the normal branch has a similar width (i.e.\ 10--20\,Hz FWHM; e.g.\ Wijnands et al.\ 1997a, 1998a) as we see in our observations ($\sim$15\,Hz) supports this identification. Since the normal branch QPO become more prominent when going from the high to the medium intensity level, we searched for normal branch QPO in our data. None were seen with upper limits of $\sim$1\%\ (5--60\,keV), which is below that seen in the normal branch of the medium intensity level ($\sim$1--2.5\%, 1--20\,keV; Wijnands et al.\ 1997a). However, when during the September 29 observations the source went from the inferred normal branch to the inferred flaring branch, a broad ($\sim$13\,Hz) noise component appeared, which peaked near 6--7\,Hz. Interestingly, similar broad noise components have been reported in the lower part of the flaring branch of other observations, but with somewhat lower strength, i.e.\ $\sim$2\%\ (1--20\,keV; Hasinger \&\ van der Klis 1989; Hasinger et al.\ 1990; Kuulkers \&\ van der Klis 1995; Wijnands et al.\ 1997a) compared to $\sim$3\%\ (5--60\,keV). It is apparant from Wijnands et al.\ (1997a) that the strength of these ``flaring branch QPO'' becomes stronger from the high to medium intensity level. Our observations extend this trend to lower overall intensities. \subsection{KiloHertz QPO} No kHz QPO were found during the October 1996 observations with upper limits which are significantly lower ($\sim$3\%, 5--60\,keV) than previously observed by Wijnands et al.\ (1998a) in the same part of the horizontal branch as inferred from the horizontal branch QPO frequency (4--5\%, 5--60\,keV). It is, however, consistent with the upper limits quoted by Smale (1998) when the source was also in the horizontal branch ($\sim$1\%, 4--11\,keV), but at higher overall intensities. As noted by Smale (1998), this may indicate that the strength of kHz QPO (at the same position in the Z) changes as a function of the overall intensity level. Unfortunately, for our October 1996 observations we can not infer to which overall intensity level it corresponds. During the September 1997 observations we found no indication for kHz QPO with upper limits of $\sim$3\%\ (5--60\,keV). This is consistent with the upper limits reported previously in the normal/flaring-branch region (2--4\%, 5--60\,keV; Wijnands et al.\ 1998a). \section{Conclusion} Using RXTE we observed Cyg\,X-2 at low overall intensities, for the first time with sufficient time resolution. In October 1996 we found the source in the leftmost part of the horizontal branch. Our observations show horizontal branch QPO properties which are generally consistent with earlier observations in this part of the Z track, but also indicate significant variations in the strength of the kHz QPO there. We conclude that we have seen parts of the normal branch and flaring branch during our September 1997 observations, when the source was seen at low overall intensities. They do not, however, resemble the behaviour seen during a rare low intensity state in 1983. Such a rare state may be observed when the overall intensity is even lower than during our observations. The properties of the very-low-frequency noise during our September low-intensity observations (low amplitude, flat power law slope) are consistent with extrapolation from those seen in previous observations at higher intensity. However, the lack of normal branch QPO during our observations is {\it not} consistent with the observed trends, and suggests that the normal branch QPO amplitude is either non-monotonically related to intensity or varies independently from this parameter. It has been suggested that obscuration by the outer accretion disk of the inner accretion disk regions and neutron star causes the low overall observed intensities during certain times and the high to medium to low intensity level variations. Such a configuration might be due to the precession of a warped accretion disk, mainly based on the rather strict periodicity of the overall intensity variations on times scales of months (see e.g.\ Wijnands et al.\ 1996, Wijers \&\ Pringle 1999). We note that obscuration effectively hardens the spectrum which leads to the changes in the position of the Z in the colour-colour diagram (see Kuulkers et al.\ 1996, Wijnands et al.\ 1996). Scattering in the outer disk would affect the variability amplitudes by light travel time smearing down to a frequency of order 0.01\,Hz. While this picture would explain the monotonic decrease in very-low-frequency noise amplitude with decreasing intensity, it seems inconsistent with the flattening of its power law index and the non-monotonic dependence of normal branch QPO amplitude on intensity. A model where the low intensity states are associated with changes in the character of the inner accretion flow itself seems therefore favoured. \section*{acknowledgements} This work was supported in part by the Netherlands Organization for Scientific Research (NWO) and by the Netherlands Foundation for Research in Astronomy (ASTRON) under grants PGS 78-277 and 781-76-017, respectively. EK thanks the Astronomical Institute ``Anton Pannekoek'', where part of the analysis was done, for its hospitality.
\section{Introduction} The generally accepted idea that supermassive accreting black holes power the highly energetic phenomena in active galactic nuclei has motivated a great deal of effort to gather information about these extraordinary objects. Speculation about the presence of a black hole at the center of our own Galaxy has been ongoing for over 20 years [see Genzel et al. (1996) or Kormendy \& Richstone (1995) and references therein for recent summaries]. Although the Galactic center appears to be nearly radio dormant, there is ample evidence for the presence of a supermassive black hole. [See, however, Tsiklauri \& Viollier (1998) and Munyaneza, Tsiklauri, \& Viollier (1998) for discussions of other possible interpretations.] One of the most efficient ways to constrain the possible existence of a massive object in the Galactic core is to carefully observe the orbital motions of the stars and gas closest to the Galactic center. Furthermore, deviations from ideal orbits may provide a probe of the distribution of dark matter around the Galactic center. Progress toward the accumulation of the relevant observational data has recently been made by Eckart et al. (1995), Genzel et al. (1996), Eckart \& Genzel (1997), Genzel et al. (1997), and Ghez et al. (1998) who have obtained high angular resolution K-band images of the Galaxy's central stellar cluster. Here we report on an attempt to reconstruct ideal orbits for the innermost 8 stars in the Ghez et al. (1998) survey using relativistic equations of motion. There is an obvious difficulty in determining orbital parameters for stars with periods much longer than the observational baseline, particularly when the uncertainties in the locations of the stars at each epoch are comparable to the apparent motion of the stars between observing epochs. Nonetheless, orbit reconstruction is possible in principle, since orbital mechanics is deterministic once the position and velocity vectors for the relevant masses are given at one instant of time. The 8 stars studied here are the ones for which the possible general relativistic effects should be greatest. The high velocities of these stars, and as we shall see, the possibility of significant periapse precession could provide a groundwork for interesting tests of general relativity as future observations of these stars are made. We also discuss other possible detectable relativistic effects which could occur if the stars pass near the massive object. Jaroszy\'nski (1998) has presented Monte Carlo simulations to demonstrate the feasibility of measuring such effects. Jaroszy\'nski (1999) and Salim \& Gould (1999) also demonstrate how accurate determinations of the orbit parameters can be used to constrain distance and mass estimates for the Galactic center. Here, we assess what can and has been learned from the currently available data. This work can also serve as a foundation for future, more-detailed probes of the mass distribution of the Galactic core. It is important to initially compute the two-body orbits of these stars around the central massive object, without considering stellar interactions or mass distribution effects. These two-body orbits provide a well-defined dynamical model against which to compare the actual astrometry and velocities of the objects. Any discrepancy or lack thereof bears directly on the mass distribution around the central black hole and could serve as an indirect means to detect the dark matter distribution in the Galactic center. For completeness, we also consider the case in which a few percent of the mass is distributed on scales comparable to the semi-major axes of the orbits being considered. As will be seen, there are distinctly different observational consequences in each case. In this paper, we present a summary of the available published astrometric data in {\ms \symbol{'170}} 2. Section 3 describes our orbital solution technique, and {\ms \symbol{'170}} 4 gives the constraints on the orbit parameters. In {\ms \symbol{'170}} 5 we discuss the possible relativistic and hydrodynamic effects these orbits may display. In {\ms \symbol{'170}} 6 we discuss which observations would be most beneficial in future studies. \section{Astrometric Data} Genzel et al. (1997) (hereafter referred to as GEOE) have presented astrometric $K$-band maps of the central $3\times3$ arcsec$^2$ of the Galaxy's central star cluster for five epochs between 1992 and 1996. These images were taken using the 3.5-m New Technology Telescope of the European Southern Observatory. Ghez et al. (1998) (hereafter referred to as GKMB) have presented $K$-band maps of the central $1\times1$ arcsec$^2$ for three epochs between 1995 and 1997. Their data was taken using the W. M. Keck 10-m telescope. Each group reported on the RA and DEC separation from Sgr A$^*$ as well as the angular proper motions for the stars. Genzel et al. (2000) presents an updated tabular summary of the proper motions obtained to date. Fig. 1 shows the combined data for the 8 stars studied here. The error bars show the uncertainty in the centroid positions for each data point. The reported uncertainties were typically on the order of $0.008$ arcsec for GEOE and $0.002$ arcsec for GKMB. We followed the naming convention of GKMB, so the 8 stars studied here are given the labels S0-1 through S0-8. We have ignored the first epoch of data from GEOE (1992.65) because of the difficulty they had in resolving any of the individual stars in the central-most region during that observation. Also, star S0-8 was not evident in the 1994.27 epoch of the GEOE data. Both groups used the results from Menten et al. (1997) to determine the position of Sgr A$^*$. This means that the RA and DEC positions from both studies are subject to the inherent uncertainty of that work. Although Menten et al. (1997) give an error estimate of $0.03$ arcsec, the GKMB group was able to link the infrared and radio reference frames and use 2 SiO maser sources to identify the position of Sgr A$^*$ to within $0.01$ arcsec. \section{Fitting Method} Because these stars have high velocities and may pass quite close to the supermassive object at the Galactic center, we choose to evolve the orbits using relativistic equations of motion for a test object of negligible mass orbiting a Schwarzschild or maximal Kerr black hole. Given the extremely large mass of the central object ($2.6\pm0.2\times10^6M_\odot$), the approximation that the stars move in a fixed background metric is certainly acceptable. Jaroszy\'nski (1998) has studied possible orbits of these stars in a Kerr background and concluded that the effects of black hole angular momentum are probably negligible. We also have made a study of orbits around a maximal Kerr black hole. As we will describe below, we found no detectable difference between the Kerr and Schwarzschild dynamics. On the other hand, Munyaneza et al. (1998) have demonstrated that noticeable differences in orbit characteristics are possible if an extended mass distribution is present instead of a single compact object. However, GKMB have placed significant constraints on the possibilities of an extended mass distribution. For this reason, we consider two separate cases: one in which all of the mass is contained in the central black hole and one in which roughly 5\% of the central mass is contained in an ideal ($\gamma = 5/3)$ gas cloud in hydrostatic equilibrium, distributed on a scale equal to the best-fit semi-major axis of star S0-2. The remaining 95\% of the mass is still in a central black hole. The gravitational potentials for these two cases are plotted in Fig. 2. As more and better data become available, refinements to the mass distribution may be detectable. For illustration, the equations of motion in Schwarzschild coordinates with the origin fixed at Sgr A$^*$ are written (Weinberg 1972, Sec. 8.4 \& 11.1) \begin{eqnarray} {d^2r \over d\tau^2}&=&{-1 \over 2A(r)}{dA(r) \over dr}\left({dr \over d\tau}\right)^2 + {r \over A(r)}\left({d\theta \over d\tau}\right)^2 \nonumber \\ & + & {r \sin^2\theta \over A(r)} \left({d\phi \over d\tau}\right)^2 - {1 \over 2A(r)}\left({dB(r) \over dr}\right) \left({dt \over d\tau}\right)^2, \end{eqnarray} \begin{equation} {d^2\theta \over d\tau^2}=-{2 \over r}{d\theta \over d\tau}{dr \over d\tau} + \sin\theta \cos\theta \left({d\phi \over d\tau}\right)^2, \end{equation} \begin{equation} {d^2\phi \over d\tau^2}= -{2 \over r}{d\phi \over d\tau}{dr \over d\tau} - 2 \cot\theta\left({d\phi \over d\tau}\right)\left({d\theta \over d\tau}\right), \end{equation} and \begin{equation} {d^2t \over d\tau^2}=-{d\ln B(r) \over dr}\left({dt \over d\tau}\right)\left({dr \over d\tau}\right), \end{equation} where the Schwarzschild metric parameters inside the mass distribution are \begin{equation} A(r)=\left[1-{2\mathcal{M}(r) \over r}\right]^{-1} \end{equation} and \begin{equation} B(r)=\exp \left\{-\int^\infty_r {2 \over {r'}^2}[\mathcal{M}(r')+4\pi {r'}^3 P(r')]\left[1-{2\mathcal{M}(r') \over r'}\right]^{-1} dr' \right\} \end{equation} where \begin{equation} \mathcal{M}(r)\equiv\int^r_04\pi r'^2 \rho(r') dr' ~. \end{equation} Outside the mass distribution, \begin{equation} B(r)=A^{-1}(r)=1-{2\mathcal{M}(R) \over r} ~. \end{equation} Here $r$, $\theta$, and $\phi$ have their usual meanings, $\tau$ is the proper time, $t$ is the coordinate time, $\mathcal{M}(R)=M$ is the mass of the central object, $P$ is the proper pressure, and $\rho$ is the proper total energy density. Here and throughout we use the convention of Weinberg (1972), i.e. a time-like metric, a negative Riemann tensor, and a negative sign in the Einstein equation [cf. Misner, Thorne, \& Wheeler (1973), to convert to other conventions]. Also, we use geometrized units ($G=c=1$) except where otherwise noted. Setting up the problem thus means that the parameters which specify a model orbit are $\{r_0,\theta_0,\phi_0,\dot{r}_0,\dot{\theta}_0,\dot{\phi}_0\}$, i.e.~the position and velocity components at some instant in time, $t_0$. We find these parameters using a least square minimization of the expression: \begin{equation} \chi^2=\sum_{i=1}^N{(\mathrm{RA}(t_i;r_0,\theta_0,\phi_0,\dot{r}_0,\dot{\theta}_0,\dot{\phi}_0)-\mathrm{RA}_i)^2+(\mathrm{DEC}(t_i;r_0,\theta_0,\phi_0,\dot{r}_0,\dot{\theta}_0,\dot{\phi}_0)-\mathrm{DEC}_i)^2 \over \sigma_i^2} \end{equation} where RA$_i$ and DEC$_i$ are the measured RA and DEC offsets at epoch $t_i$ and RA$(t_i;r_0,\theta_0,\phi_0,\dot{r}_0,\dot{\theta}_0,\dot{\phi}_0)$ and DEC$(t_i;r_0,\theta_0,\phi_0,\dot{r}_0,\dot{\theta}_0,\dot{\phi}_0)$ are the corresponding coordinate offsets for that same epoch, calculated using the model parameters. The error bars in the measured RA and DEC offsets are generally unequal. Therefore the best weighting factor for the goodness of fit is the radius of the error ellipse along the line formed by joining the observed position at an epoch with the modeled location of the star at that epoch. Hence, we write the weighting factor for the orbit fits as \begin{equation} \sigma_i^2 = (\sigma_{RA_i}\cos \Phi_i)^2+(\sigma_{DEC_i}\sin \Phi_i)^2 \end{equation} where $\Phi_i$ is the angle between the RA-axis and the line drawn between the observed and model-orbit positions. To convert from angular separations to physical distances, we adopt 8.0($\pm0.5$) kpc as the distance to the Galactic center (Reid, 1993). The uncertainty in this distance is included in our error estimates. It is important to note that the orbits thus determined are degenerate in the parameters $\theta_0$ and $\dot{\theta}_0$. There is no way to determine from the available data whether or not a particular star is currently in front of or behind Sgr A*. By choice, we have assumed that all of these stars currently lie in front. This choice affects the orbit parameters given in the next section, but has no effect on the conclusions of this paper. Refined spectroscopic studies will help to break this degeneracy. \section{Computed Orbits} Having determined the model orbits as described above, we then convert the orbit parameters to their more familiar form $\{a,e,i,\Omega,\omega,T\}$, where $a$ is the semi-major axis, $e$ is the eccentricity, $i$ is the inclination, $\Omega$ is the longitude of the ascending node, $\omega$ is the argument of periapse, and $T$ is the time of last periapse passage [cf. Taff (1985), for the analytic expressions needed to perform this conversion]. Table 1 gives the orbit parameters for stars S0-1 through S0-3. Stars S0-4 through S0-8, though analyzed, are not listed in Table 1 because the orbit parameter $\theta_0$ for these stars is not constrained by the observations, i.e. all orbits with $0\le\theta_0<\pi$ are within $1\sigma$ of the optimum fit. For the convenience of the readers, we have included the period, $P$, for each of the orbits in Table 1. The errors given in the table are the statistical $1\sigma$ uncertainties for each parameter. Star S0-1 is an interesting case because its ``best-fit'' orbit is actually an unbound, hyperbolic trajectory. Fig. 2 shows the computed, optimum least $\chi^2$ orbits overlaid on the observed data. \section{Relativistic and Hydrodynamic effects} The most relevant relativistic effect in a proper motion study is the precession of periapse. The angular advance of the periapse, evaluated numerically here, is given approximately by \begin{equation} |\Delta\omega|=6\pi{GM \over a(1-e^2)c^2} \end{equation} in units of radians per revolution [cf. Ohanian \& Ruffini (1994)]. It is a simple matter to transform the angular advance of periapse in the plane of the orbit into the more easily measured angular shift of apoapse in the plane of the sky. Table 2 gives the periapse precession and the angular shift of apoapse for stars S0-2 and S0-3, assuming no mass distribution. For both stars, values near the $1\sigma$ upper limits are in principle measurable with the current observational position accuracies. Fig. 4 illustrates possible precession effects over multiple revolutions of a model orbit for star S0-2. For this illustration, we show orbits confined to the plane of the sky ($i=0$), although the effect looks similar for any inclination in the range, $-40^\circ\lesssim i\lesssim40^\circ$. The $\chi^2$ for this orbit differs from the ``best-fit'' orbit by only 0.4. For comparison, we have shown a Newtonian orbit for two point-masses ({\it left panel}), as well as relativistic orbits with and without distributed mass ({\it right panel} and {\it middle panel}, respectively). As expected, there is no precession in the Newtonian, point-mass limit. For the relativistic cases, there is a very important difference in the resulting precession. Relativistic periapse precession, by itself, causes an angular {\bf advance} of periapse ({\it middle panel}). With the orbital inclination confined to small values, as in the model orbit illustrated here, the apoapse shift is $0.001$ arcsec per revolution and the orbital period is 10 years. Using the current measured angular separation accuracy of $\pm0.002$ arcsec, the apoapse shift in this case could in principle be detectable with a baseline of $\approx20$ years of observation. The periapse precession due to a possible mass distribution, on the other hand, normally results in an angular {\bf regression} ({\it right panel}). Therefore, a measured advance of periapse for these stars is most likely a relativistic effect. Such a measurement would also place strict limits on any mass distribution. Clearly this is a difficult measurement and one must carefully exclude competing effects from other masses around the central black hole. Nevertheless, it is worth searching for, since even determining the sign of the precession will reveal a great deal. We also considered the effect of black hole angular momentum on this orbit. Any effect will be most pronounced when the angular momentum axis of the orbit is aligned (or anti-aligned) with the angular momentum axis of the black hole. In this case, both were chosen to lie along the line of sight. For the model orbit shown in Fig. 4, the effect of the black hole angular momentum amounts to a 0.0001 arcsec shift of the apoapse after 3 revolutions ($\approx 30$ years). Thus it represents a 3\% effect on top of the relativistic periapse precession. Since our $1 \sigma$ upper limit on the relativistic apoapse shift for S0-2 and S0-3 is 0.001 arcsec per revolution, the best-case $1 \sigma$ upper limit on the apoapse shift due to black hole angular momentum is 0.00003 arcsec per revolution. Measuring this effect would require observations on a time scale of hundreds of years at current astrometric accuracies. Hence, we conclude that the orbits of these stars can probably not be used to detect black hole angular momentum. Another potentially measurable effect is the gravitational redshift ($\Delta\nu/\nu=-GM/c^2R$) of the light emitted from these stars as they proceed through their orbits. Note, however, that the expected special relativistic red- and blueshifts for these stars can be much larger that any anticipated gravitational redshift and must therefore be precisely accounted for before any gravitational redshift will be discernible. Table 2 contains a summary of both of these effects. If these stars are sufficiently massive, it is possible that gravitational radiation from their final inspiral and plunge may be detectable by space-based interferometry such as the proposed LISA mission (P. Bender 1999, private communication). Nevertheless, the time scale for angular momentum loss by gravitational radiation for the orbits considered here is greater than a Hubble time ($\sim 10^{15}$ yr). The most likely cause of the final plunge will be scattering by a 2-body interaction somewhere along the orbit. This raises the issue as to whether the time scale for these stars to suffer small angle deflections from other stars around the black hole is less than or comparable to the time scale necessary to carry out a measurement of the periapse precession. The magnitude of the acceleration on star $i$ due to the other stars in the field \begin{equation} a = \sum_j {m_j \over r_{ij}^2} \end{equation} will be dominated at any time by the nearest stars. For the observed stars, the typical nearest neighbor is separated by about 0.2 arcsec on the sky or about 0.008 pc at 8 kpc. Allowing for eight times as many other (non K-luminous) stars to be present, the typical separation might be 0.004 pc with an associated mean acceleration due to nearest neighbors of $\langle a \rangle \sim 9 \times 10^{-9}$ m/sec. Assuming the most pessimistic case that there is always a 1 M$_\odot$ star at this closest distance with the same direction, then the time scale for a star to be deflected by a minimally detectable 0.002 arcsec ($\approx 16$ AU) shift is then \begin{equation} t = \sqrt{2 \Delta x/ \langle a \rangle} \approx {\rm~700~yrs} ~. \end{equation} Also, it should be noted that due to the motion around the black hole, these stars are probably moving with rapid relative velocities $v \sim 0.01~c$. Therefore, the radius of gravitational influence $G M_\odot/\langle v^2 \rangle$ is quite small ($\sim 10^4$ km). Hence, unless there are extremely close encounters, we do not expect the mutual interactions to dominate over the relativistic periapse precession of interest here. On the other hand, any such deviation, if detected, would most likely be distinguishable from periapse precession in both direction and magnitude and would serve as a means to detect the presence of unseen matter. The possibility of gravitational lensing of stars near the the Galactic center has been discussed elsewhere (Alexander \& Sternberg 1998; Salim \& Gould 1999; Jaroszy\'nski 1999; and Capozziello \& Iovane 1999). It has been concluded that this effect is probably negligible (Jaroszy\'nski 1999). As another point of possible interest we consider the hydrodynamic effects of these stars passing close to the central object. For a star of mass $M_*\ll M_{BH}$, the Roche-lobe radius, $R$, is given by \begin{equation} {R \over r}=0.49\left({M_* \over M_{BH}}\right)^{1/3}, \end{equation} where $r$ is the radial distance from the star to the central object and $M_{BH}$ is the mass of the central object (King \& Done 1993). This effect is strongly dependent on the periapse separation of these stars. It is reassuring that none of the optimum fits in the present study are passing so close that they might have experienced Roche-lobe overflow. Also, since it is now evident (Figer et al. 2000; Eckart, Ott, \& Genzel 1999) that these stars are probably OB stars and not extended K-giants, Roche overflow seems unlikely. Nevertheless, within the $1\sigma$ uncertainties deduced here, it is at least possible that one or more of these stars might approach or temporarily exceed their Roche limit during the next periapse passage. This might be interesting to watch for, and if it occurred, it might lead to a burst of activity in the Galactic center. As another possible observable effect, we consider whether any of these stars could pass close enough to the central black hole to experience hydrodynamic distortion due to tidal interactions and whether these distortions could lead to observable oscillations in surface temperature or luminosity. For distortions from an equilibrium spherical shape of the form \begin{equation} r = R + \epsilon Y_l^m(\phi,\theta) ~, \end{equation} where $Y_l^m(\phi,\theta)$ are spherical harmonics, both the oscillatory period and the excitation time scale for these Kelvin modes (Lang 1999) should be of order \begin{equation} P = { 2 \pi \over [ 2 l(l-1) G M_* / (2 l + 1) R^3]^{1/2}} ~. \end{equation} We estimate $ P \lesssim 3$ days, which is too short compared to the time of periapse passage for such oscillations to be excited. \section{Discussion} Given the interesting physics that might be gleaned from extended observations of these orbits, it is useful to summarize what steps must be taken to reduce the large statistical uncertainties in the orbital parameters. Of course, a great deal is to be learned by continued astrometry. This is ultimately the only way in which the periapse precessions and/or effects of the mass distribution can be determined. In addition, it is crucial to obtain accurate radial line-of sight velocities as well as the radial distance to Sgr A$^*$. Current studies have provided fairly accurate measurements of RA, DEC, $V_{RA}$, and $V_{DEC}$, however, these must be complemented by K-band spectroscopy. Eckart et al. (1999) and Figer et al. (2000) have reported on high resolution infrared spectroscopy in the vicinity of Sgr A$^*$. The spectra are consistent with all of the stars in this sample being OB stars. The strongest feature observed from this region is a weak Br$\gamma$ emission line. However, as of yet no individual redshifts for these stars have been identified. What is needed are long integrations with high spatial resolution. Clearly, it should be a high priority to obtain such redshifts. If all six coordinates could be measured to comparable precision (e.g. $\sigma_x=\sigma_y=\sigma_z=0.002$ arcsec $\approx20$ AU and $\sigma_{V_x}=\sigma_{V_y}=\sigma_{V_z}\approx100$ km s$^{-1}$) then the $1\sigma$ uncertainties in the orbit parameters for star S0-2 would reduce to about 25\% of the best-fit values derived in this work. If the velocity uncertainties were reduced to 50 km s$^{-1}$ then the uncertainties in orbit parameters for S0-2 would be reduced to about 20\%. If the line-of-sight velocity for each star could be accurately obtained directly from spectroscopic studies then the change in radial velocity, $dV_z/dt=-GMz/(x^2+y^2+z^2)^{3/2}$, over time might be used to find a value for the final unknown parameter - the line-of-sight separation between the star and Sgr A$^*$ (labeled $z$). Fig. 4 illustrates graphically the relation between $dV_z/dt$ and $z$ for each of the stars. From this we see that, in order to measure $z$ to an accuracy of $20$ AU, $dV_z/dt$ must be measured to an accuracy of $\sim7.5$ km s$^{-1}$ yr$^{-1}$, in the case of the star S0-2. This would require, for instance, two measurements of $V_z$ at accuracies of 50 km s$^{-1}$ separated by about 9.5 years. This time scale for accurate orbit determination is similar to the time scale derived by Salim \& Gould (1999) to use these orbits to better constrain the distance to Sgr A$^*$. \section{Conclusion} We have explored families of possible orbits for the 8 known stars located within $0.5$ arcsec of the Galactic center. Because line-of-sight velocities have not yet been obtained and only positions through the 1998 observing epoch are available, orbital parameters could only be constrained for 3 of the stars. The ideal orbits of these stars could be much better constrained from relatively short baseline studies of line-of-sight velocities, although continued observations of angular positions and angular proper motions are crucial if relativistic or mass-distribution effects are to be identified. These orbits as they now stand have at least the potential to display some extraordinary properties, including periods of less than 10 years and very high eccentricities. The current range of orbit parameters also allow for the possibility of interesting close encounters with the $2.6\times10^6M_\odot$ object at the Galactic core. It is clear from Eckart et al. (1999) and Figer et al. (2000) that these stars are probably blue, luminous, and very young OB stars most likely formed from material infalling toward the black hole. A likely explanation is tidal disruption of infalling rich OB associations by the central black hole. Hence, it is reasonable for the stars to have the highly eccentric orbits deduced here. The possibility of these eccentric orbits passing close to the black hole opens up the possibility for detection of relativistic effects such as periapse precession and/or gravitational redshift. At the same time, continued observations may also reveal the presence and distribution of unseen matter at the Galactic core. Hydrodynamic effects may be evident as well, if the stars pass close enough to be tidally distorted. Also, the star currently passing closest to Sgr A$^*$ in the plane of the sky may not be in a bound orbit around the central mass. This introduces the intriguing questions as to whether other members of this association are unbound. If so, then the inferred virial mass of the central object may need revision. In light of the many interesting puzzles highlighted in this paper, it is imperative that priority be given to continued astrometric, as well as spectroscopic, study of these stars. Such studies will provide valuable knowledge about the distribution and the dynamical evolution of mass within the central core of the Galaxy. \begin{acknowledgements} The authors wish to acknowledge useful discussion with A. Ghez. We would also like to thank an anonymous referee for numerous helpful comments, particularly in regards to the importance of mass distribution effects. This work was supported by the National Science Foundation under grant PHY-97-22086. \end{acknowledgements} \clearpage
\section*{I. Introduction} The violations of the three discrete symmetries C, P, T have changed our ideas about the physical world. The discovery of CP violation in $K^0- \bar{K^0}$ complex was made long time ago. Recently, The direct T violation \cite{CPLEAR1} and direct CP violation ($Re(\epsilon'/\epsilon)\neq 0$) \cite{KTeV} are found experimentally. Only the combined CPT symmetry is left unbroken. The CPT theorem is the general result of the local, relativistic field theory. If it is violated, it will cause the fundamental crisis of our present particle theory. The recent tests of CPT violating effects give the bounds \cite{CPLEAR2} $r_{_K}\equiv \mid \frac{m_{\bar{K^0}}-m_{K^0}}{m_{K^0}}\mid \leq 10^{-18}$ and the very uncertain values $Re(\delta)=( 3.0\pm 3.3_{stat}\pm 0.6_{syst})\times 10^{-4}$, $Im(\delta)=(-1.5\pm 2.3_{stat}\pm 0.3_{syst})\times 10^{-2}$. The tests of CPT symmetry in B system have been suggested from the theory \cite{Kostel} and phenomenology \cite{Sanda1} \cite{Xing}. In \cite{Sanda1}, the authors point out that CPT violation in $B^0\bar B^0$ mixing can lead to a dilepton asymmetry of neutral-B decays. They also discuss some general effects caused by CPT violation. Neutral-meson interferometry is a powerful tool for investigating discrete symmetry. The CP violation ($\epsilon$), direct T and direct CP violation are all observed in $K^0 \bar{K^0}$ complex. In the decay chain $B\to J/\psi+K \to J/\psi+[f]$, neutral K mixing follows on the heels of B mixing. This mixing is called "cascade mixing" and the decay is "cascade decay". Cascade mixing has been got some theorists' interests \cite{Azimov} \cite{Kayser1}. The extension to the decay chain $B\to D \to [f]_D$ for exploring new physics can be found in \cite{Meca}. The advantage of cascade decay is that we can use the known K mixing parameter to determine the B mixing parameter. In \cite{Kayser1}, Kayser shows that the cascade decay contains more information than the usually discussed processes. So, cascade decay provides a complex and elegant window for the detail of the $B^0-\bar{B^0}$ mixing. In this work, we shall make a detailed study of the CPT violating effects caused by $B^0-\bar{B^0}$ mixing in cascade decays. The $B^0-\bar{B^0}$ mixing is described by two complex phases $\theta$, $\phi$. We first take an approximation method to treat the parameters $\theta$ and $\phi$ and give the general formulas of the direct CPT and T asymmetry. Then we investigate how to extract the $B^0-\bar{B^0}$ mixing parameters. The feasibility of exploring the CPT violation in B-factories and LHC-B is discussed. Finally, we suggest a procedure to determine the two complex phases. \section*{II. $B^0-\bar{B^0}$ mixing and CP, T, CPT Asymmetries} Weak interaction can cause oscillation between $B^0$ and $\bar{B^0}$. The eigenstates of weak decays are not $B^0$ and $\bar{B^0}$ but their superpositions which have the simple exponential laws. The two eigenstates of $B^0 \bar B^0$ mesons are given by \begin{eqnarray} |B_L>=\frac{1}{\sqrt{|p_1^2|+|q_1|^2}}[p_1|B^0>+q_1|\bar{B^0}>] \nonumber\\ |B_H>=\frac{1}{\sqrt{|p_2^2|+|q_2|^2}}[p_2|B^0>-q_2|\bar{B^0}>] \end{eqnarray} and their eigenvalues are \begin{eqnarray} \mu_L=m_L-\frac{i}{2}\Gamma_L =m_B-\frac{\Delta m_B}{2}-\frac{i}{2}(\Gamma_B+\frac{\Delta \Gamma_B}{2}) =m_B-\frac{i}{2}\Gamma_B-\frac{\Delta m_B}{2}-\frac{i}{2}y\Gamma_B \nonumber\\ \mu_H=m_H-\frac{i}{2}\Gamma_H =m_B+\frac{\Delta m_B}{2}-\frac{i}{2}(\Gamma_B-\frac{\Delta \Gamma_B}{2}) =m_B-\frac{i}{2}\Gamma_B+\frac{\Delta m_B}{2}+\frac{i}{2}y\Gamma_B \end{eqnarray} From PDG98\cite{PDG98}, $x\equiv\frac{\Delta m_B}{\Gamma_B}\sim 0.7$, while $y\equiv\frac{\Delta \Gamma_B}{2\Gamma_B}\leq 10^{-2}$ is theoretically expected \cite{Nir1}. The mixing parameter $p_i$, $q_i$ are related by\cite{Lee} \begin{eqnarray} \frac{q_1}{p_1}= tg\frac{\theta}{2}e^{i\phi}, ~~~~ \frac{q_2}{p_2}=ctg\frac{\theta}{2}e^{i\phi} \end{eqnarray} where $\theta$ and $\phi$ are complex phases in general. For real $\theta$ and $\phi$, $0<\theta<\pi$, $0\leq\phi<2\pi$. From Eq.(1), The mass difference between $B^0$ and $\bar{B^0}$ is: \begin{equation} M_{B^0}-M_{\bar{B^0}}=(\mu_L-\mu_H)\frac{p_1q_2-p_2q_1}{p_1q_2+p_2q_1} =(\mu_L-\mu_H)cos\theta \end{equation} where $M_{\stackrel{(-)}{B^0}}=m_{\stackrel{(-)}{B^0}} -\frac{i}{2}\Gamma_{\stackrel{(-)}{B^0}}$. Using the Bell-Steinberg unitarity relation \cite{Bell}, \begin{equation} |<B_H|B_L>|\leq \frac{\Gamma_L \Gamma_H}{|\mu_L^*-\mu_H|} =\frac{\sqrt{1-y^2}\Gamma_B}{|i-x|\Gamma_B}\simeq 0.8 \end{equation} This constraint is more relaxed than that in $K^0-\bar{K^0}$ complex where $|<K_L|K_S>|\leq 0.06$. So, $|B_H>$ and $|B_L>$ are likely unorthogonal. The more relaxed constraint of Eq.(5) perhaps indicates the large CP or CPT violation in $B^0$ system. The initially $|B^0>$ or $|\bar{B^0}>$ will evolve after a proper time $t$ to \begin{eqnarray} |B^0(t)>=g_+(t)|B^0>+\bar g_+(t)|\bar{B^0}> \nonumber\\ |\bar B^0(t)>=g_-(t)|\bar{B^0}>+\bar g_-(t)|B^0> \end{eqnarray} where \begin{eqnarray} g_+(t)=f_+(t)+cos\theta f_-(t), ~~~&g_-(t)=f_+(t)-cos\theta f_-(t) \nonumber\\ \bar{g}_+(t)=sin\theta e^{i\phi}f_-(t), ~~~~& \bar{g}_-(t)=sin\theta e^{-i\phi}f_-(t) \end{eqnarray} and \begin{eqnarray} f_+(t)=\frac{1}{2}(e^{-i\mu_L t}+e^{-i\mu_H t}) =e^{-im_B t-\frac{1}{2}\Gamma_B t}ch(\frac{ix-y}{2}\Gamma_Bt) \nonumber\\ f_-(t)=\frac{1}{2}(e^{-i\mu_L t}-e^{-i\mu_H t}) =e^{-im_B t-\frac{1}{2}\Gamma_B t}sh(\frac{ix-y}{2}\Gamma_Bt) \end{eqnarray} The probability for $|B^0>$ in a proper time $t$ to transform into $|B^0>$ is: \begin{equation} P_{B^0(t)\to B^0}=|<B^0|B^0(t)>|^2=|g_+(t)|^2 \end{equation} Similarly, we can define $P_{B^0(t)\to \bar B^0}$, $P_{\bar B^0(t)\to B^0}$, and $P_{\bar B^0(t)\to \bar B^0}$. So, the mixing-induced CPT and T asymmetries can be defined as \begin{equation} A_{CPT}(t)\equiv \frac{P_{B^0(t)\to B^0}-P_{\bar B^0(t)\to \bar B^0}} {P_{B^0(t)\to B^0}+P_{\bar B^0(t)\to \bar B^0}} =\frac{2Re[cos\theta sh(\frac{ix-y}{2}\Gamma_Bt) (ch(\frac{ix-y}{2}\Gamma_Bt))^*]} {|ch(\frac{ix-y}{2}\Gamma_Bt)|^2+ |cos\theta|^2 |sh(\frac{ix-y}{2}\Gamma_Bt)|^2}, \end{equation} and \begin{equation} A_T(t)\equiv \frac{P_{B^0(t)\to \bar B^0}-P_{\bar B^0(t)\to B^0}} {P_{B^0(t)\to \bar B^0}+P_{\bar B^0(t)\to B^0}} =\frac{|e^{i\phi}|^2-|e^{-i\phi}|^2} {|e^{i\phi}|^2+|e^{-i\phi}|^2} \end{equation} Some analysis can lead to the following conclusions \cite{Sanda2}: \\ (1) CPT invariance requires $cos\theta=0$, and thus $\theta=\frac{\pi}{2}$. \\ (2) T invariance requires $\phi=0$; \\ (3) CP conservation requires $cos\theta=0$ ( and thus $\theta=\frac{\pi}{2}$ ) and $\phi=0$. In Standard Model, CPT is conserved and $|\frac{q}{p}|-1=|e^{i\phi}|-1=\frac{1}{2} Im\frac{\Gamma_{12}}{M_{12}}\sim {\cal O}(10^{-3})$ \cite{Nir2}. Thus, the direct T violating asymmetry is about the order of ${\cal O}(10^{-3})$. From the experience in K system, we guess that the CPT violating effects may be smaller than the CP violating effects. Under the above assumption, it is convenient to introduce a complex $\theta'$ and two real $\phi_0$, $\phi'$ by \begin{eqnarray} \theta=\frac{\pi}{2}+\theta', ~~~&\theta'=Re\theta'+iIm\theta' \nonumber\\ \phi=\phi_0+i\phi', ~~~&\phi_0=Re\phi, ~\phi'=Im\phi \end{eqnarray} where $Re\theta'$, $Im\theta'$, $\phi_0$, $\phi'$ are all real, and $|\theta'|<<1$, ~$|\phi'|<<1$. The relation between $\phi_0$ and CKM phase $\beta$ is $\phi_0=-2\beta$. Then, we have a very simple relation \begin{eqnarray} cos\theta=-\theta', ~~~~sin\theta=1 \nonumber\\ e^{i\phi}=e^{i\phi_0}(1-\phi')~~~~~~~~ \end{eqnarray} Here we only keep terms up to the first order of $\theta'$ and $\phi'$. From PDG98 \cite{PDG98}, The mixing parameter $\frac{\Delta \Gamma_B} {\Gamma_B}$ has not been measured experimentally up to now. We further assume $y=0$ in order to simplify the formulations below. Thus, \begin{equation} ch(\frac{ix-y}{2}\Gamma_Bt)=cos\frac{\Delta m_B t}{2}, ~~~~~~~~~ sh(\frac{ix-y}{2}\Gamma_Bt)=isin\frac{\Delta m_B t}{2} \end{equation} From Eq.(10), Eq.(11), Eq.(13) and Eq.(14), we obtain \begin{eqnarray} A_{CPT}(t)&=&\frac{2Im\theta' sin\Delta m_B t}{1+cos\Delta m_B t} \nonumber\\ A_{T}(t)&=&-2\phi' \end{eqnarray} So, the mixing-induced CPT asymmetry is proportional to $Im\theta'$, and the mixing-induced T asymmetry is proportional to $Im\phi=\phi'$. Now, we discuss two cases: (1) Final state is not CP eigenstate We study the decays $B^0\to Xl^+\nu$, $\bar{B^0}\to\bar Xl^-\nu$. From $\Delta B=\Delta Q$ rule, the decays of $B^0\to\bar Xl^-\nu$, $\bar{B^0}\to Xl^+\nu$ are forbidden. For the allowed processes, we define the amplitude: \begin{eqnarray} <Xl^+\nu|H|B^0>=A, ~~~~~~~<\bar X l^-\nu|H|\bar{B^0}>=A^* \nonumber \end{eqnarray} So the asymmetry of semileptonic decay rates are \begin{eqnarray} D_1(f,t)&\equiv& \frac {\Gamma(B^0(t)\to Xl^+\nu)-\Gamma(\bar{B^0}(t)\to \bar X l^-\nu)} {\Gamma(B^0(t)\to Xl^+\nu)+\Gamma(\bar{B^0}(t)\to \bar X l^-\nu)} =\frac {P_{B^0(t)\to B^0}-P_{\bar{B^0}(t)\to \bar{B^0}}} {P_{B^0(t)\to B^0}+P_{\bar{B^0}(t)\to \bar{B^0}}} =A_{CPT}(t) \nonumber\\ D_2(f,t)&\equiv& \frac {\Gamma(B^0(t)\to \bar X l^-\nu)-\Gamma(\bar{B^0}(t)\to Xl^+\nu)} {\Gamma(B^0(t)\to \bar X l^-\nu)+\Gamma(\bar{B^0}(t)\to Xl^+\nu)} =\frac {P_{B^0(t)\to \bar{B^0}}-P_{\bar{B^0}(t)\to B^0}} {P_{B^0(t)\to \bar{B^0}}+P_{\bar{B^0}(t)\to B^0}} =A_{T}(t) \end{eqnarray} Eq.(16) shows that CPT and T asymmetry can lead to CP asymmetry. One can use the CP asymmetry of semileptonic B decays to measure the CPT and T violation parameter. (2) Final state is CP eigenstate The decay rate for an initial $B^0$ or $\bar{B^0}$ transform into a CP eigenstate $f$ is \begin{eqnarray} \Gamma(B^0(t)\to f)=e^{-\Gamma_B t}|A|^2 \{\frac{1+cos\Delta m_Bt}{2}+Im\theta'sin\Delta m_Bt+ |\frac{\bar A}{A}|^2(1-2\phi')\frac{1-cos\Delta m_Bt}{2} \nonumber\\ -Im[e^{i\phi_0}\frac{\bar A}{A} (sin\Delta m_Bt-\phi' sin\Delta m_Bt +i\theta'^*(1-cos\Delta m_Bt))]\}~~~~ \\ \Gamma(\bar{B^0}(t)\to f)=e^{-\Gamma_B t}|A|^2 \{(1+2\phi')\frac{1-cos\Delta m_Bt}{2}+|\frac{\bar A}{A}|^2 (\frac{1+cos\Delta m_Bt}{2}-Im\theta'sin\Delta m_Bt) \nonumber\\ +Im[e^{i\phi_0}\frac{\bar A}{A} (sin\Delta m_Bt+\phi' sin\Delta m_Bt +i\theta'(1-cos\Delta m_Bt))]\}~~~~ \nonumber \end{eqnarray} where $A\equiv A(B^0\to f)$ and $\bar A\equiv A(\bar{B^0}\to f)$. For $f=J/\psi K_S$, $\frac{\bar A}{A}=-1$, the CP asymmetry is \begin{eqnarray} D(J/\psi K_S,t)&=&\frac {\Gamma(B^0(t)\to J/\psi K_S)-\Gamma(\bar{B^0}(t)\to J/\psi K_S)} {\Gamma(B^0(t)\to J/\psi K_S)+\Gamma(\bar{B^0}(t)\to J/\psi K_S)} \nonumber\\ &=&sin\phi_0sin\Delta m_Bt+Re\theta'cos\phi_0(1-cos\Delta m_Bt) +Im\theta'sin\Delta m_Bt \nonumber\\ &&-Im\theta'sin\phi_0(1-cos\Delta m_Bt) -\phi'(1-cos\Delta m_Bt)+\phi'sin\phi_0sin\Delta m_Bt \end{eqnarray} For $f=J/\psi K_L$, $\frac{\bar A}{A}=+1$, the CP asymmetry is \begin{eqnarray} D(J/\psi K_L,t)&=&\frac {\Gamma(B^0(t)\to J/\psi K_L)-\Gamma(\bar{B^0}(t)\to J/\psi K_L)} {\Gamma(B^0(t)\to J/\psi K_L)+\Gamma(\bar{B^0}(t)\to J/\psi K_L)} \nonumber\\ &=&-sin\phi_0sin\Delta m_Bt-Re\theta'cos\phi_0(1-cos\Delta m_Bt) +Im\theta'sin\Delta m_Bt \nonumber\\ &&+Im\theta'sin\phi_0(1-cos\Delta m_Bt) -\phi'(1-cos\Delta m_Bt)-\phi'sin\phi_0sin\Delta m_Bt \end{eqnarray} \section*{III. Cascade decays} We have discussed the $B^0-\bar{B^0}$ mixing in the previous section. Now we turn to the cascade mixing involving both neutral B and neutral K systems in succession. Neglecting CPT violating effects in the neutral K system, the weak eigenstates of the neutral K mesons can be represented by the usual form: \begin{eqnarray} |K_S>=\frac{1}{\sqrt{|p_K^2|+|q_K|^2}}[p_K|B^0>+q_K|\bar{B^0}>] \nonumber\\ |K_L>=\frac{1}{\sqrt{|p_K^2|+|q_K|^2}}[p_K|B^0>-q_K|\bar{B^0}>] \end{eqnarray} and their eigenvalues are \begin{eqnarray} \mu_{S(L)}=m_K{\stackrel{(+)}{-}}\frac{\Delta m_K}{2} -i\frac{\Gamma_{S(L)}}{2} \end{eqnarray} where $m_K$ is the average of the $K_S$ and $K_L$ masses, $\Gamma_{S,L}$ are the $K_{S,L}$ widths. The time evolution of the neutral K mesons can be easily obtained \begin{eqnarray} |K^0(t)>=e_+(t)|K^0>+\frac{q_K}{p_K}e_-(t)|\bar{K^0}> \nonumber\\ |\bar{K^0}(t)>=\frac{p_K}{q_K}e_-(t)|K^0>+e_+(t)|\bar{K^0}> \end{eqnarray} where \begin{eqnarray} e_{\pm}(t)=\frac{1}{2}(e^{-i\mu_S}\pm e^{-i\mu_L}) \end{eqnarray} Consider the decay chain $B\to J/\psi+K \to J/\psi+[f]$ where $f$ can be $2\pi$, $3\pi$ and $\pi l\nu$ as shown in Fig.1. Other decay modes of the neutral K mesons are neglected because of either very small fractions or less physical interest for this paper. We first give the formulations of the most complicated case where the final state $f=\pi l\nu$. According to \cite{Kayser2}, the decay amplitude of the cascade decay $B^0{\stackrel{t_1}{\to}} J/\psi+K{\stackrel{t_2}{\to}} J/\psi+[\pi^-l^+\nu]$ is \begin{eqnarray} && A(B^0{\stackrel{t_1}{\to}} J/\psi+K{\stackrel{t_2}{\to}} J/\psi+[\pi^-l^+\nu])= g_+(t_1)A(B^0\to J/\psi K^0)e_+(t_2)A(K^0\to \pi^-l^+\nu) \nonumber\\ && ~~~~~~~ +\bar g_+(t_1)A(\bar{B^0}\to J/\psi \bar{K^0}) \frac{p_K}{q_K}e_-(t_2)A(K^0\to \pi^-l^+\nu) \end{eqnarray} We assume that the transition amplitude for B and K decays satisfy the $\Delta S=\Delta Q$ rule and CP, CPT invariance. There is no experimental signal of the violation $\Delta S=\Delta Q$ rule. The assumption of CP conservation in the transition amplitude for B and K decays is valid because $\frac{A(\bar{B^0}\to J/\psi \bar{K^0})}{A(B^0\to J/\psi K^0)}=-1$ to a very high degree and the direct CP violation in $K^0-\bar{K^0}$ system is very small ($Re(\epsilon'/\epsilon)\sim 10^{-3}$). We further neglect the small CP violations in $ K^0-\bar{K^0}$ mixing, thus $\frac{q_K}{p_K}=1$. Under the above assumptions, the decay rate of the cascade decay $B^0{\stackrel{t_1}{\to}} J/\psi+K{\stackrel{t_2}{\to}} J/\psi+[\pi^{\mp}l^{\pm}\nu]$ is \begin{eqnarray} &&\Gamma(B^0, J/\psi+[\pi^{\mp}l^{\pm}\nu])\equiv \Gamma(B^0{\stackrel{t_1}{\to}} J/\psi+K{\stackrel{t_2}{\to}} J/\psi+[\pi^{\mp}l^{\pm}\nu]) \nonumber\\ &&\propto e^{-\Gamma_Bt_1}\{e^{-\Gamma_St_2} [1+sin\phi_0sin\Delta m_Bt_1+Re\theta'cos\phi_0(1-cos\Delta m_Bt_1) +Im\theta'sin\Delta m_Bt_1 \nonumber\\ && +Im\theta'sin\phi_0(1-cos\Delta m_Bt_1) -\phi'(1-cos\Delta m_Bt_1)-\phi'sin\phi_0sin\Delta m_Bt_1] \nonumber\\ &&+e^{-\Gamma_Lt_2} [1-sin\phi_0sin\Delta m_Bt_1-Re\theta'cos\phi_0(1-cos\Delta m_Bt_1) +Im\theta'sin\Delta m_Bt_1 \nonumber\\ && -Im\theta'sin\phi_0(1-cos\Delta m_Bt_1) -\phi'(1-cos\Delta m_Bt_1)+\phi'sin\phi_0sin\Delta m_Bt_1] \nonumber\\ && \pm 2e^{-\frac{1}{2}(\Gamma_S+\Gamma_L)t_2} [cos\Delta m_Bt_1cos\Delta m_Kt_2+cos\phi_0sin\Delta m_Bt_1sin\Delta m_Kt_2 \nonumber\\ && -Re\theta'sin\phi_0(1-cos\Delta m_Bt_1) +Im\theta'sin\Delta m_Bt_1cos\Delta m_Kt_2 \nonumber\\ && +Im\theta'cos\phi_0(1-cos\Delta m_Bt_1)sin\Delta m_Kt_2 +\phi'(1-cos\Delta m_Bt_1)cos\Delta m_Kt_2 \nonumber\\ && -\phi'cos\phi_0sin\Delta m_Bt_1sin\Delta m_Kt_2] \} \end{eqnarray} Similarly, the decay rate of the cascade decay $\bar{B^0}{\stackrel{t_1}{\to}} J/\psi+K{\stackrel{t_2}{\to}} J/\psi+[\pi^{\pm}l^{\mp}\nu]$ is \begin{eqnarray} &&\Gamma(\bar B^0, J/\psi+[\pi^{\pm}l^{\mp}\nu])\equiv \Gamma(\bar{B^0}{\stackrel{t_1}{\to}} J/\psi+K{\stackrel{t_2}{\to}} J/\psi+[\pi^{\pm}l^{\mp}\nu]) \nonumber\\ && =\Gamma(B^0{\stackrel{t_1}{\to}} J/\psi+K{\stackrel{t_2}{\to}} J/\psi+[\pi^{\mp}l^{\pm}\nu]) (\theta'\to -\theta', \phi_0\to -\phi_0, \phi'\to -\phi') \end{eqnarray} Because we have neglected the small CP-violating effects in K system, only $K_S\to 2\pi$ and $K_L\to 3\pi$ are possible. The decay rate for the cascade decays of the $f=2\pi$ and $f=3\pi$ are: \begin{eqnarray} &&\Gamma({\stackrel{(-)}{B^0}}, J/\psi+[2\pi])\equiv \Gamma({\stackrel{(-)}{B^0}}{\stackrel{t_1}{\to}} J/\psi+K_S {\stackrel{t_2}{\to}}J/\psi+[2\pi]) \nonumber\\ &&\propto 4e^{-\Gamma_Bt_1}\{e^{-\Gamma_St_2} [1{\stackrel{(-)}{+}}sin\phi_0sin\Delta m_Bt_1 {\stackrel{(-)}{+}}Re\theta'cos\phi_0(1-cos\Delta m_Bt_1) {\stackrel{(-)}{+}}Im\theta'sin\Delta m_Bt_1 \nonumber\\ &&~~~~~ +Im\theta'sin\phi_0(1-cos\Delta m_Bt_1) {\stackrel{(+)}{-}}\phi'(1-cos\Delta m_Bt_1) -\phi'sin\phi_0sin\Delta m_Bt_1] \} \end{eqnarray} and \begin{eqnarray} &&\Gamma({\stackrel{(-)}{B^0}}, J/\psi+[3\pi])\equiv \Gamma({\stackrel{(-)}{B^0}}{\stackrel{t_1}{\to}} J/\psi+K_L {\stackrel{t_2}{\to}}J/\psi+[3\pi]) \nonumber\\ &&\propto 4e^{-\Gamma_Bt_1}\{e^{-\Gamma_St_2} [1{\stackrel{(+)}{-}}sin\phi_0sin\Delta m_Bt_1 {\stackrel{(+)}{-}}Re\theta'cos\phi_0(1-cos\Delta m_Bt_1) {\stackrel{(-)}{+}}Im\theta'sin\Delta m_Bt_1 \nonumber\\ &&~~~~~ +Im\theta'sin\phi_0(1-cos\Delta m_Bt_1) {\stackrel{(+)}{-}}\phi'(1-cos\Delta m_Bt_1) +\phi'sin\phi_0sin\Delta m_Bt_1] \} \end{eqnarray} \section*{IV. The determination of the parameter $\theta$ and $\phi$} 1. $\phi'$ From Eq.(16) and \cite{Sanda1}, \begin{eqnarray} A_{T}(t)=\frac{\Gamma(B^0(t)\to \bar Xl^-\nu)- \Gamma(\bar{B^0}(t)\to Xl^+\nu)} {\Gamma(B^0(t)\to \bar Xl^-\nu)+ \Gamma(\bar{B^0}(t)\to Xl^+\nu)} =\frac{N^{++}-N^{--}}{N^{++}+N^{--}} =-2\phi' \end{eqnarray} where $N^{++}$, $N^{--}$ are the same-sign dilepton events. $\phi'$ can be measured by the semileptonic decays of the B mesons or by the same-sign dilepton ratios suggested in \cite{Sanda1}. \noindent 2. $Im\theta'$ From Eq.(16) and \cite{Sanda1}, \begin{eqnarray} A_{CPT}(t)=\frac{\Gamma(B^0(t)\to Xl^+\nu)- \Gamma(\bar{B^0}(t)\to \bar Xl^-\nu)} {\Gamma(B^0(t)\to Xl^+\nu)+ \Gamma(\bar{B^0}(t)\to \bar Xl^-\nu)} =\frac{N^{+-}-N^{-+}}{N^{+-}+N^{-+}} =\frac{2Im\theta'sin\Delta m_Bt}{1+cos\Delta m_Bt} \end{eqnarray} where $N^{+-}$, $N^{-+}$ are opposite-sign dilepton events. Thus, $Im\theta'$ can be measured by the semileptonic decays of the B mesons or by the opposite-sign dilepton ratios. For the dileptonic decays in Eq.(29) and Eq.(30), they correspond to the case of $C=-1$ where $C$ is the charge conjugation number of the $B^0\bar{B^0}$ pair. Another method is: from Eq.(27) and Eq.(28), \begin{eqnarray} &&\frac{\Gamma(B^0, J/\psi+[2\pi])-\Gamma(\bar{B^0}, J/\psi+[2\pi])} {\Gamma(B^0, J/\psi+[2\pi])+\Gamma(\bar{B^0}, J/\psi+[2\pi])} +\frac{\Gamma(B^0, J/\psi+[3\pi])-\Gamma(\bar{B^0}, J/\psi+[3\pi])} {\Gamma(B^0, J/\psi+[3\pi])+\Gamma(\bar{B^0}, J/\psi+[3\pi])} \nonumber\\ &&=2[Im\theta'sin\Delta m_Bt_1-\phi'(1-cos\Delta m_Bt_1)] \end{eqnarray} Using the known $\phi'$ value from the semileptonic decays or the dilepton ratios, $Im\theta'$ can be determined by Eq.(31). But this method is not good for experiment because their branching ratios are smaller than the semileptonic decays. \noindent 3. $sin\phi_0$ and the absolute value of the $cos\phi_0$ and $Re\theta'$ From Eq.(27) and (28), we can obtain the time-dependent asymmetry of the decay rates \begin{eqnarray} &&\frac {[\Gamma(B^0, J/\psi+[2\pi])-\Gamma(\bar{B^0}, J/\psi+[2\pi])]} {[\Gamma(B^0, J/\psi+[2\pi])+\Gamma(\bar{B^0}, J/\psi+[2\pi])]} -\frac {[\Gamma(B^0, J/\psi+[3\pi])-\Gamma(\bar{B^0},J/\psi+[3\pi])]} {[\Gamma(B^0, J/\psi+[3\pi])+\Gamma(\bar{B^0},J/\psi+[3\pi])]} \nonumber\\ &&=2[sin\phi_0sin\Delta m_Bt_1+ Re\theta'cos\phi_0(1-cos\Delta m_Bt_1)] \end{eqnarray} The asymmetry of Eq.(32) is twice as much as the usual CP asymmetry in the decay of $B\to J/\psi K_S$ because we have used the decay mode of $B\to J/\psi K_L$ to double the asymmetry. We will discuss the problem caused by the detection of $K_L$ later. There are two contributions in the asymmetry of Eq.(32). The $sin\Delta m_Bt_1$ term is an odd function of time while the $(1-cos\Delta m_Bt_1)$ term is an even function. These two terms can be distinguished experimentally by measuring the decay time order of $B^0$ abd $\bar{B^0}$ decays. The detail of this method is given in \cite{Gronau}. Here we only use this method to distinguish the $sin\Delta m_Bt_1$ and $(1-cos\Delta m_Bt_1)$ terms. Like \cite{Gronau}, define two asymmetries \begin{eqnarray*} a_{-}(f,t)=\frac {(\bar \Gamma+{\stackrel{\sim }{\Gamma}}) -(\Gamma+{\stackrel{\sim }{\bar \Gamma}})} {(\bar \Gamma+{\stackrel{\sim }{\Gamma}}) +(\Gamma+{\stackrel{\sim }{\bar \Gamma}})} ,~~~~ a_{+}(f,t)=\frac {(\bar \Gamma+{\stackrel{\sim }{\bar \Gamma}}) -(\Gamma+\bar \Gamma)} {(\bar \Gamma+{\stackrel{\sim }{\bar \Gamma}}) +(\Gamma+\bar\Gamma)} \end{eqnarray*} where $\Gamma$, $\bar \Gamma$, ${\stackrel{\sim }{\Gamma}}$ and ${\stackrel{\sim }{\bar \Gamma}}$ are defined in \cite{Gronau}, and the subscript (-) or (+) denotes an odd or even function of time. The measurement of asymmetry $a_{-}(f,t)$ requires measuring the decay time order. Thus the two terms of Eq.(32) can be distinguished by \begin{eqnarray} A_{-}(t_1)=a_{-}(f_1,t_1)-a_{-}(f_2,t_1)=2sin\phi_0sin\Delta m_Bt_1 \end{eqnarray} and \begin{eqnarray} A_{+}(t_1)=a_{+}(f_1,t_1)-a_{+}(f_2,t_1) =2Re\theta'cos\phi_0(1-cos\Delta m_Bt_1) \end{eqnarray} where $f_1$ and $f_2$ represent the final states $J/\psi+[2\pi]$ and $J/\psi+[3\pi]$. In Eq.(34), the asymmetry is often used to measure the direct CP violation when CPT invariance holds. The direct CP violation in $B\to J/\psi K$ decays is at the order of ${\cal O}(10^{-3})$. So the measurement of the CPT violation in Eq.(34) can only reach the accuracy up to $10^{-2}$ because of the dilution of direct CP violation in $B\to J/\psi K$ decay and the CP violation in $K^0-\bar{K^0}$ mixing. From the Eq.(33) and Eq.(34), the values of $sin\phi_0$ and $Re\theta'cos\phi_0$ can be obtained. So $cos\phi_0$ and $Re\theta'$ can be determined except for the ambiguity of their sign. This ambiguity can be solved by the cascade decays where $f=\pi l\nu$. \noindent 4. The sign of $cos\phi_0$ and $Re\theta'$ From Eq.(25) and Eq.(26), \begin{eqnarray} &&\frac {[\Gamma( B^0, J/\psi+[\pi^- l^+\nu]) -\Gamma( B^0, J/\psi+[\pi^+ l^-\nu]) +\Gamma(\bar{B^0}, J/\psi+[\pi^- l^+\nu]) -\Gamma(\bar{B^0}, J/\psi+[\pi^+ l^-\nu])} {[\Gamma( B^0, J/\psi+[\pi^- l^+\nu]) +\Gamma( B^0, J/\psi+[\pi^+ l^-\nu]) +\Gamma(\bar{B^0}, J/\psi+[\pi^- l^+\nu]) +\Gamma(\bar{B^0}, J/\psi+[\pi^+ l^-\nu])} \nonumber\\ &&=\frac{-2e^{-\frac{1}{2}(\Gamma_S+\Gamma_L)t_2} Re\theta'sin\phi_0(1-cos\Delta m_Bt_1)} {\{e^{-\Gamma_St_2}[1+Im\theta'sin\phi_0(1-cos\Delta m_Bt_1) -\phi'sin\phi_0sin\Delta m_Bt_1]} \nonumber\\ &&~~ +e^{-\Gamma_Lt_2}[1-Im\theta'sin\phi_0(1-cos\Delta m_Bt_1) +\phi'sin\phi_0sin\Delta m_Bt_1]\} \\ && \sim \frac{-2e^{-\frac{1}{2}(\Gamma_S+\Gamma_L)t_2} Re\theta'sin\phi_0(1-cos\Delta m_Bt_1)} {e^{-\Gamma_St_2}+e^{-\Gamma_Lt_2}} \nonumber \end{eqnarray} and \begin{eqnarray} &&\frac {[\Gamma( B^0, J/\psi+[\pi^- l^+\nu]) -\Gamma( B^0, J/\psi+[\pi^+ l^-\nu]) -\Gamma(\bar{B^0}, J/\psi+[\pi^- l^+\nu]) +\Gamma(\bar{B^0}, J/\psi+[\pi^+ l^-\nu])} {[\Gamma( B^0, J/\psi+[\pi^- l^+\nu]) +\Gamma( B^0, J/\psi+[\pi^+ l^-\nu]) +\Gamma(\bar{B^0}, J/\psi+[\pi^- l^+\nu]) +\Gamma(\bar{B^0}, J/\psi+[\pi^+ l^-\nu])} \nonumber\\ &&=\frac {A}{B} \sim \frac { 2e^{-\frac{1}{2}(\Gamma_S+\Gamma_L)t_2} [cos\Delta m_Bt_1cos\Delta m_Kt_2 +cos\phi_0sin\Delta m_Kt_2sin\Delta m_Bt_1]} {e^{-\Gamma_St_2}+e^{-\Gamma_Lt_2}} \end{eqnarray} where \begin{eqnarray} A&=& 2e^{-\frac{1}{2}(\Gamma_S+\Gamma_L)t_2} [cos\Delta m_Bt_1cos\Delta m_Kt_2 +cos\phi_0sin\Delta m_Kt_2(sin\Delta m_Bt_1 \nonumber\\ && -Im\theta'(1-cos\Delta m_Bt_1) +\phi'sin\Delta m_Bt_1) \nonumber\\ && -Im\theta'sin\Delta m_Bt_1cos\Delta m_Kt_2 -\phi'(1-cos\Delta m_Bt_1)] \nonumber\\ B&=& e^{-\Gamma_St_2}[1+Im\theta'sin\phi_0(1-cos\Delta m_Bt_1) -\phi'sin\phi_0sin\Delta m_Bt_1] \nonumber\\ && +e^{-\Gamma_Lt_2}[1-Im\theta'sin\phi_0(1-cos\Delta m_Bt_1) +\phi'sin\phi_0sin\Delta m_Bt_1] \nonumber \end{eqnarray} From Eq.(35) and Eq.(36), the sign of $cos\phi_0$ and $Re\theta'$ can be measured. Actually, we do not need to use the Eq.(35) since we have known the value of $Re\theta'cos\phi_0$. \section*{V. Feasibility and discussions} In order to meet the goal of a three standard deviation measurement for the $B^0-\bar{B^0}$ mixing parameter $\theta$ and $\phi$, the relation between the number of $B^0\bar{B^0}$ pairs and the asymmetry is \cite{SLAC89}: \begin{eqnarray} N_{B^0\bar{B^0}}=\frac{1}{B\epsilon_r\epsilon_t [(1-2W)d\cdot\delta A]^2} \end{eqnarray} where $\delta A=\frac{A}{3}$; $A$ is the asymmetry of the decay ratios; $B$ is the branching ratio of the decay; $\epsilon_r$ is the reconstruction efficiency of the final state $f$; $\epsilon_t$ is the tagging efficiency; $W$ is the fraction of incorrect tags; $d$ is the dilution factor which takes into account the loss in asymmetry due to fitting, time integration, and/or the mixing of the tagged decay. Sometimes, another relation is used: \begin{eqnarray} N_{eff}=\frac{1}{[(1-2W)d\cdot\delta A]^2} \end{eqnarray} where $N_{eff}$ is the effective number of the decay events. The B-factories can accumulate $1.8\times 10^8$ $B\bar B$ pairs every year \cite{PEP}, and the effective event number of Eq.(38) for $B\to J/\psi K_S$ is taken to be $2.7\times 10^{5}$ in LHC-B\cite{LHC}. Table 1 and Table 2 give some parameters \cite{SLAC89} and the minimum asymmetries (lower bound) which can be achieved in B-factory and LHC-B. \begin{table}[hbt] \begin{center} Table 1. Branching Ratios and Reconstruction efficiencies \\ for the cascade decays and semileptonic decays. \vspace{0.6cm} \begin{tabular}{|c|c|c|} \hline \hline Decay mode & Braching Ratio $B$ & Reconstruction effeciency $\epsilon_r$ \\ \hline $B\to J/\psi K_S \to J/\psi+[2\pi]$ & $5\times 10^{-4}$ &0.61 \\ \hline $B\to J/\psi K_L \to J/\psi+[3\pi]$ & $5\times 10^{-4}\times \frac{1}{3}$ & 0.4 \\ \hline $B\to J/\psi K_S \to J/\psi+[\pi l\nu]$ & $5\times 10^{-4} \cdot 1.2\times 10^{-3}$ & 0.61\\ \hline $B\to J/\psi K_L \to J/\psi+[\pi l\nu]$ & $5\times 10^{-4} \times \frac{2}{3}$ & 0.4 \\ \hline $B\to l\nu+X$ & 0.1 & 1 \\ \hline $J/\psi\to l^+l^-$ & 0.14 & - \\ \hline \end{tabular} \vspace{0.8cm} Tag efficiencies and Asymmetry dilution at B-factory and LHC-B \vspace{0.5cm} \begin{tabular}{|c|c|c|} \hline \hline & at B-factory & at LHC-B \\ \hline Tag efficiency $\epsilon_t$ & 0.48 & 0.61 \\ \hline Asymmetry dilution $d$ & 0.61(for $B_d$) &0.61 \\ \hline \end{tabular} \end{center} \end{table} \begin{table}[hbt] \begin{center} Table 2. Comparison between the Minimum of Asymmetries \\ with $3\sigma$ standard deviation at B-factory and LHC-B \vspace{0.5cm} \begin{tabular}{|c|c|c|} \hline \hline Decay mode & \multicolumn{2}{|c|} {Asymmetry $A$} \\\cline{2-3} & at B-factory & at LHC-B \\ \hline $B\to J/\psi K_S \to J/\psi+[2\pi]$ & 0.08 & $1.2\times 10^{-2}$ \\ \hline $B\to J/\psi K_L \to J/\psi+[3\pi]$ & 0.17 & $2.6\times 10^{-2}$ \\ \hline $B\to J/\psi K_S \to J/\psi+[\pi l\nu]$ & 1.44 & 0.27\\ \hline $B\to J/\psi K_L \to J/\psi+[\pi l\nu]$ & 0.06 & 0.01 \\ \hline $B\to l\nu+X$ & $1.7\times10^{-3}$ & $1\times10^{-5}$ \\ \hline $B\bar B\to l^+l^-$ & $7\times10^{-3}$ & $4\times10^{-5}$ \\ \hline \end{tabular} \end{center} \end{table} Now we discuss how to determine the two complex phase $\theta$ and $\phi$. (1) $\phi'$ and $Im\theta'$ can be measured in semileptonic decays and dileptonic decays given in Eq.(29) and Eq.(30). Semileptonic decays have larger branching ratio but smaller detection efficiency, while dileptonic decays have larger detection efficiency but smaller branching ratio than semileptonic decays. In B-factory (with $1.8\times 10^8$ $B^0\bar{B^0}$ pairs per year), the $\phi'$ and $Im\theta'$ can be measured to an accuracy of $2 \times 10^{-3}$ for semileptonic decays and $7\times 10^{-3}$ for dileptonic decays at $3\sigma$ level. In LHC-B (with $4\times 10^{12}$ $b\bar b$ pairs every snow mass year), the $\phi'$ and $Im\theta'$ can be measured to an accuracy of $10^{-5}$ for semileptonic decays and $4\times 10^{-5}$ for dileptonic decays with $3\sigma$ standard deviation. (2) $Re\theta'cos\phi_0$ can be determined clearly from Eq.(34). In Eq.(32) the decay of $K_L\to 3\pi$ is used in order to increase the asymmetry factor and reduce the ambiguity or error caused by the unknown $\phi'$ and $Im\theta'$. The $K_L$ detection is a challenge for experiment. In \cite{SLAC92}, one idea to catch $K_L$ by using Fe sampling after the electromagnetic calorimeter is suggested. In B-factory, the accuracy of measuring $Re\theta'cos\phi_0$ can reach $0.07$. So, it is likely that $Re\theta'cos\phi_0$ can not be measured in B-factory. In LHC-B, the accuracy of measuring $Re\theta'cos\phi_0$ can reach $0.01$. We have no confidence that $Re\theta'cos\phi_0$ can be measured clearly with $3\sigma$ standard deviation in LHC-B. If it can be measured, $Re\theta'$ must be larger than $0.01$. This will be the largest CPT violation effects. (3) $sin\phi_0$ is usually suggested to be measured in $B\to J/\psi K_S(\to\pi^+\pi^-)$ which gives the $A(t)=sin\phi_0sin\Delta m_Bt$ in Standard Model. If the CPT violating effects are considered, the asymmetry is modified to Eq.(18). In order to cancel the errors caused by $\phi'$ and $Im\theta'$, we use the Eq.(33) to determine $sin\phi_0$. (4) $sin\phi_0$ can be measured in B-factory and LHC-B as discussed above, but the value of $\phi_0$ has two ambiguity. If $Re\theta'cos\phi_0$ can be measured, the only ambiguity is the sign of $cos\phi_0$. This can be solved by measuring the ratios of the cascade decays $B\to J/\psi+K\to J/\psi+[\pi l\nu]$ given in Eq.(36). Because the very small branching ratio of decay $B\to J/\psi+K\to J/\psi+[\pi l\nu]$ in $t_2 \le 2\tau_S$, the determination of the sign of $cos\phi_0$ can only be done in LHC-B. Table 2 gives the lower bound of measuring for asymmetry in cascade decay is 0.2. This is possible because $|cos\phi_0|>0.5$, we have taken $0.3\leq sin\phi_0\leq 0.88$ \cite{Nir2} for our estimation. In conclusion, the cascade decays provide an elegant and beautiful place to study the CPT violation caused by the $B^0-\bar{B^0}$ mixing. \section*{Acknowledgment} This work is supported in part by National Natural Science Foundation of China and the Grant of State Commission of Science and Technology of China.
\section{Introduction} The presence of large amount of ``cold" (i.e. not much ionized, and substantially opaque in X--rays) matter around Active Galactic Nuclei is now a well established fact. For all Seyfert 2 galaxies observed in X--rays so far there is evidence for absorption in excess of the Galactic one. In a significant fraction of them, the column density of the absorbing matter exceeds 10$^{24}$ cm$^{-2}$ (Maiolino et al. 1998), and the matter is therefore optically thick to Compton scattering. In a few objects, like NGC~1068 (Matt et al. 1997), the column density is so high that the X--ray photons cannot escape even in hard X--rays, being trapped in the matter, downscattered to energies where photoelectric absorption dominates, and eventually destroyed. In other cases, like NGC~4945 (Iwasawa et al. 1993; Done et al. 1996), Mrk~3 (Cappi et al. 1999) and the Circinus Galaxy (Matt et al. 1999), the column density is a few$\times$10$^{24}$ cm$^{-2}$, so permitting the transmission of a significant fraction of X--ray photons, many of them escaping after one or more scatterings. To properly model transmission through absorbers with these intermediate column densities, is therefore necessary to fully take into account Compton scattering. This has been done in an analytical, approximated way by Yaqoob (1997), but his model is valid only below $\sim$15 keV, a painful limitation after the launch of BeppoSAX, which carries a sensitive hard X--ray (15-200 keV) instrument, and in view of future missions like Astro-E and Constellation-X. We have therefore calculated transmitted spectra by means of Monte Carlo simulations. The code is, as far as physical processes are concerned, basically that described in Matt, Perola \& Piro (1991). A spherical geometry, with the X--ray source in the centre, has been assumed, while the element abundances are those tabulated in Morrison \& McCammon (1983). Photoelectric absorption, Compton scattering (in a fully relativistic treatment) and fluorescence (for iron atoms only) are included in the code. Photon's path are followed until either the photon is photoabsorbed (and not re--emitted as iron fluorescence) or escapes from the cloud. Spectra have been calculated for 31 different column densities, ranging from 10$^{22}$ to 4$\times10^{24}$ cm$^{-2}$. In order to be independent of the shape of the primary radiation, transmitted spectra for monochromatic emission have been calculated, with a step of 0.1 keV below 20 keV, and 1 keV above. A grid has then been constructed, which can be folded with the chosen spectral shape. \section{Transmitted spectra. Comparison with simple absorption models} To illustrate the effects of including the Compton scattering in the transmission spectrum, we show in Figs 1-5 (which refer to column densities of 10$^{23}$, 3$\times10^{23}$, 10$^{24}$, 3$\times10^{24}$ and 10$^{25}$ cm$^{-2}$, respectively), the results of the MonteCarlo simulations (solid lines), along with the results when only photoelectric absorption (dotted lines) or both photoelectric and Compton absorption without scattering (dashed lines) are included. The first case is unphysical, and it is shown here only for the sake of illustration; the second case would correspond to absorption by matter with a negligible covering factor to the primary source (i.e. a small cloud on the line of sight), a physically possible but highly unlikely situation, at least for Seyfert galaxies (the fraction of Compton--thick sources is estimated to be at least 30\%, Maiolino et al. 1998, and then the covering fraction of the matter must be significant). The injected spectrum is a power law with a photon index of 2 and an exponential cut--off at 500 keV, as typical for Seyfert galaxies (even if the latter parameter is at present poorly known). As expected, our curves lie below the dotted curves (because of the larger absorption especially above $\sim$10 keV, where the Compton cross section starts dominating over the photoelectric cross section), and above the dashed curves (because of the extra radiation provided by the scattering of photons into the line of sight). The effect is dramatic, especially for large column densities and high energies. \begin{figure} \epsfig{file=123.ps, height=12.cm, width=12.cm, angle=0.} \caption{Transmitted spectra (in $EF(E)$) for a column density N$_{\rm H}$=$10^{23}$ cm$^{-2}$. The solid line refer to the Monte Carlo results discussed in this paper. The dotted line is for photoelectric absorption alone (an unphysical situation, presented here only for the sake of comparison); the dashed line for complete (photoelectric and Compton scattering) absorption, as for a small cloud on the line of sight.} \end{figure} \begin{figure} \epsfig{file=323.ps, height=12.cm} \caption{ As for Figure 1, but for N$_{\rm H}$=$3\times10^{23}$ cm$^{-2}$.} \end{figure} \begin{figure} \epsfig{file=124.ps, height=12.cm} \caption{ As for Figure 1, but for N$_{\rm H}$=$10^{24}$ cm$^{-2}$.} \end{figure} \begin{figure} \epsfig{file=324.ps, height=12.cm} \caption{ As for Figure 1, but for N$_{\rm H}$=$3\times10^{24}$ cm$^{-2}$.} \end{figure} \begin{figure} \epsfig{file=125.ps, height=12.cm} \caption{ As for Figure 1, but for N$_{\rm H}$=$10^{25}$ cm$^{-2}$.} \end{figure} \section{Applications. I. The Circinus Galaxy} As a first application, let us discuss the case of the Circinus Galaxy. Matt et al. (1999) analyzed the BeppoSAX observation of this source and found a clear excess in hard X--rays (i.e. in the PDS instrument) with respect to the best fit medium energy (i.e. LECS and MECS) spectrum (which, in turn, was in good agreement with the ASCA result, Matt et al. 1996). The excess is best explained assuming that the nuclear emission is piercing through material with moderate Compton--thick (i.e. between $10^{24}$ and $10^{25}$ cm$^{-2}$, see previous section) column density. Compton scattering must therefore be taken into account in modeling the emerging spectrum, not only in absorption but also in emission, as there is clear evidence of large amount of reflection too, suggesting a fairly large solid angle subtended by the cold matter to the primary source. The fit with the model described here yields the parameters of the transmitted component reported in Table~1 (model 1). Model 2 in the same table refers to the fit with a pure absorption model (photelectric plus Compton). Both models are statistically acceptable (reduced $\chi^2\sim$1), but the differences in the best fit parameters are significant, leading to dramatically different (i.e. two orders of magnitude) X--ray nuclear luminosities. \begin{table} \vspace{0.15in} \caption{Parameters of the best fit models. Model 1 includes Compton scattering, model 2 only absorption. The primary spectrum is a power law with photon index $\Gamma$ and a high energy cut--off; E$_{\rm C}$ is the corresponding e--folding energy. A is the density flux at 1 keV.} \begin{tabular}{lcc} \hline \hline ~ & 1 & 2 \cr \hline ~ & ~ & ~\cr N$_{\rm H,1}$~(10$^{24}$~cm$^{-2}$) & 4.3 & 6.9 \cr A~(ph~cm$^{-2}$~s$^{-1}$~keV$^{-1}$~at~1~keV) & 0.11 & 15.4 \cr $\Gamma$ & 1.56 & 1.58 \cr E$_{\rm C}$~(keV) & 56 & 38 \cr L(2-10~keV)~(erg~s$^{-1}$) & 10$^{42}$ & 1.5$\times$10$^{44}$ \cr ~ & ~ & ~\cr \hline \hline \end{tabular} \vspace{0.15in} \end{table} \section{Applications. II. The hard X--ray Background} The origin of the thermal--like, $\sim$40 keV spectrum (Marshall et al. 1980) of the hard Cosmic X--ray background (XRB) has remained unexplained for many years. In 1989, Setti \& Woltjer proposed an explanation in terms of a mixture of obscured (i.e. Seyfert 2s) and unobscured (i.e. Seyfert 1s) AGN. Following this idea, many authors developed synthesis models for the XRB (e.g. Madau, Ghisellini \& Fabian 1993, 1994; Matt \& Fabian 1994; Comastri et al. 1995), and nowadays this explanation is widely considered as basically correct. To model the spectrum of the XRB it is necessary to include all the relevant ingredients, and a correct transmission spectrum is one of them because, as remarked above, Compton--thick sources are a significant fraction of all Seyfert 2s. To our knowledge, out of the many papers devoted to fitting the XRB, the transmission component has been properly included only by Madau, Ghisellini \& Fabian (1994). Here we do it again, to highlight and discuss the differences with models in which only absorption is included. In Fig.~\ref{localsp}, we show the integrated local spectrum of Seyfert 1 galaxies (dotted curve), of Seyfert 2 galaxies (lower dashed and solid curves) and of the sum of Seyfert 1 and 2 galaxies (upper dashed and solid curves). The spectrum of Seyfert 1 galaxies is described by a power law with a photon spectral index of 1.9 and an exponential cut--off with $e$-folding energy of 400 keV; a Compton reflection component, corresponding to an isotropically illuminated accretion disk observed at an inclination angle of 60$^{\circ}$, is also included. According to unification models, the spectrum of Seyfert 2 galaxies is assumed to be intrinsically identical to that of Seyfert 1s, but seen through obscuring matter. The solid lines in the figure refer to a synthesis model in which the transmitted component is included, while in the dashed ones only absorption is considered. Type 2 sources are assumed to outnumber type 1 sources by a factor of 4, independently of the luminosity. The adopted distribution of column densities of the absorbing matter for the Seyfert 2s is: ${dN \over dLog(N_{\rm H})} \propto Log(N_{\rm H})$, from 10$^{21}$ to 4$\times$10$^{25}$ cm$^{-2}$. The fraction of Compton--thick sources is then about 1/3, in agreement with the estimate of Maiolino et al. (1998). The two total spectra differ significantly above 10 keV, the spectrum including the transmission component being about 20\% higher at 30 keV. The best fit spectrum to the XRB (HEAO-1 data, Marshall et al. 1980), obtained after evolving the local spectrum of Seyfert galaxies to cosmological distances, following Boyle et al. (1994), is shown in Fig.\ref{fitsp}. Different descriptions of the pure luminosity evolution scenario do not change significantly the results. The study of both the spectral shape of the XRB and the source counts in different scenarios, including e.g. density evolution, is beyond the scope of the present work, and is deferred to a forthcoming paper (Pompilio et al., in preparation). Apart from the highest energy part of the spectrum, where the fit is not very good (suggesting either that an exponential cut--off is not a good description of the spectrum of Seyfert galaxies, or that there is not a universal value of such a parameter, as actually is emerging from BeppoSAX observations: see e.g. Matt 1998), and the lowest part (where contributions from other classes of sources, like Clusters of Galaxies, may be relevant), the agreement between the data and the model is acceptable. The soft X--ray source counts are also well reproduced, while the hard (5--10 keV) counts (Fiore et al., in preparation; Comastri et al. 1999) are somewhat underestimated, but still marginally consistent with the data. The complete model and the detailed analysis will be discussed elsewhere (Pompilio 1999; Pompilio et al., in preparation). \begin{figure} \epsfig{file=localsp.ps, height=12.cm} \caption{The integrated local spectrum of Seyfert 2s (with, solid line, and without, dashed line, the transmitted component), of Seyfert 1s (dotted line) and total. See text for detail.} \label{localsp} \end{figure} \begin{figure} \epsfig{file=fitxrb.ps, height=12.cm} \caption{ Best fit spectrum to the XRB (data from HEAO-1, Marshall et al. 1980), using our model including the transmitted component.} \label{fitsp} \end{figure}
\section{Introduction} The chiral phase transition in QCD separating high temperature (high density) quark-gluon plasma phase from low temperature (low density) hadronic phase has been studied intensively in the last decade. At temperatures of the order of the pion mass, $T \approx 140$ MeV or densities a few times that of ordinary nuclear matter, $ n \approx (3 - 5)n_0$ with $n_0 \approx 0.15 $ fm$^{-3}$, conditions are reached where hadrons start overlapping, only the short distance interaction among the internal parton degrees of freedom described by QCD is important and dynamical breaking of chiral symmetry no longer occurs. Understanding the properties of this transition is becoming increasingly important in view of recent experimental efforts to create and detect the quark-gluon plasma in the relativistic heavy-ion collisions at BNL and CERN. Since the problem of chiral symmetry breaking and its restoration is intrinsically non-perturbative, the number of available techniques is limited and most of our knowledge about the phenomenon comes from lattice simulations. Due to the complexity of QCD, studies have so far been done on lattices of modest size and have been unable to yield quantitative results as far as the universality class of the finite temperature phase transition is concerned. Also, finite density QCD, which is more relevant to real physics, is a field where lattice technology is still at an early stage \cite{barbour98}. The complex nature of the determinant of the Dirac operator at finite chemical potential makes it impossible to use standard simulation algorithms based on positive definite probability functions. This manuscript addresses the problem of chiral symmetry restoration at non-zero temperature and non-zero chemical potential in the three-dimensional four-fermion model \cite{rosen91,kogut93} with a $Z_2$ chiral symmetry in order to understand what ingredients might play a decisive role in more complex systems like gauge theories. The model has been simplified as much as possible in order to produce data of the highest quality and get some insight into the range of parameters we need for studies of more realistic problems. The large $N$ description (which is equivalent to a mean field approximation) of the phase structure of the three-dimensional four-fermion model at non-zero chemical potential \cite{hands_kocic_kogut.93} predicts a first order transition for $T=0 $ and a continuous transition for $T>0$. The critical value of the chemical potential $\mu_c$ is equal to the value of the fermion mass at $\mu=0$. Interactions as expected decrease $\mu_c$ below the mean field result \cite{kim95}. A recent numerical study of the $T \neq 0, \mu = 0$ transition favors the two-dimensional Ising model exponents for $N<\infty$ \cite{kogut_98}. In this manuscript we present the results of Monte Carlo simulations with a small number of fermion species, i.e., $N=4$ in order to test the mean field theory predictions and look for possible next-to-the-leading order corrections. We show in this paper by employing finite size scaling techniques that the $T=0$, $\mu \neq 0$ phase transition is first order. We also show that the second order nature of the $T \neq 0$, $\mu=0$ transition remains stable down to low temparatures and large values of the chemical potential. A simple continuity argument suggests that a tricritical point in the $(\mu, T)$ plane must lie on the low temperature and high density section of the critical line. More specifically, our simulations show that the tricritical point lies on the section of the phase boundary defined by: $T/T_c \leq 0.23$, $\mu/\mu_c \geq 0.97$, where $T_c$ is the zero density critical temperature and $\mu_c$ is the zero temperature critical chemical potential. Our simulations did not provide any evidence for non-zero fermion number density at $T=0$ and $\mu<\mu_c$. This implies that the nuclear liquid-gas transition is either extremely weak or very close to the chiral transition or it does not exist in this model. Our results show that higher order corrections in the $1/N$ expansion for the nature and the location of the transition points in the phase diagram are very small for this model. \section{The model} The continuum spacetime (we work in Euclidean space throughout) Lagrangian of the model is \begin{equation} \label{lagr1.ff3d} {\cal L}= \bar{\psi}_i(\partial\hskip -.5em / +\mu \gamma_0 +m_q) \psi_i - \frac{g^{2}}{2 N} (\bar{\psi}_i \psi_i)^{2}. \end{equation} Here $\psi_i$ is a four-component spinor and the index $i$ runs over $N$ fermion species. $\mu$ is the chemical potential and $m_q$ is the fermion bare mass. For zero fermion bare mass, ${\cal L}$ has a $Z_2$ chiral symmetry \begin{equation} \psi \rightarrow \gamma_5 \psi \ ; \ \bar{\psi} \rightarrow - \bar{\psi} \gamma_5 \end{equation} which is spontaneously broken whenever a non-vanishing condensate $ \langle \bar{\psi} \psi \rangle$ is generated. Both analytical and numerical work in this model are aided by the introduction of an auxiliary field $\sigma$, so Eq.~(\ref{lagr1.ff3d}) becomes \begin{equation} \label{lagr2.ff3d} {\cal L}= \bar{\psi}_i(\partial\hskip -.5em / + \sigma)\psi_i + \frac{N}{2 g^{2}} \sigma^{2}. \end{equation} At tree level, the field $\sigma$ has no dynamics; it is truly an auxiliary field. However, it acquires dynamical content by dint of quantum effects arising from integrating out the fermions. In the chiral limit the vacuum expectation value of $\sigma$ becomes a dynamical fermion mass and it is a convenient order parameter for any chiral phase transition in the theory. There are several motivations for studying such a simple model: (i) for sufficiently strong coupling it exhibits chiral symmetry breaking at zero temperature and density; (ii) the spectrum of excitations contains both baryons and mesons, i.e. the elementary fermions and the composite fermion$-$anti-fermion states; (iii) the model has an interacting continuum limit; (iv) when formulated on the lattice it has a real Euclidean action even for non-zero chemical potential; (v) the Yukawa-like coupling makes the inverse Dirac operator non-singular and the simulations can be done in the chiral limit directly, whereas in gauge theories the matrix to be inverted becomes ill-conditioned and the simulations have to be done using a fermion bare mass; (vi) the model has fewer degrees of freedom than QCD, and can be studied with greater precision on much bigger lattices than the lattices presently used for QCD thermodynamics. The chiral symmetry must be discrete in order to observe a phase transition at non-zero temperature in $d=2+1$, otherwise according to Coleman$-$Mermin$-$Wagner theorem \cite{coleman} a continuous chiral symmetry must be manifested algebraically for all $T>0$. In four dimensions the four-fermion model is a trivial theory \cite{hands98} and is believed to be an effective theory of quarks and gluons at intermediate energies. Unfortunately, the order of the chiral phase transition at non-zero chemical potential in this model depends on the value of the cutoff \cite{klevansky}. In two dimensions the discrete chiral symmetry restoration is dominated by the materialization of kink$-$anti-kink states. The zero temperature and finite chemical potential transition is first order, whereas for any finite number of fermion species $N$ the chiral symmetry is restored at any non-zero temperature due to the condensation of the kinks \cite{karsch}. The four-fermion model in its bosonized form may be formulated on a spacetime lattice using the following action: \begin{equation} S_{lat}=\sum_{i=1}^{N/2} \left( \sum_{x,y} \bar{\chi}_{i}(x) M_{x,y} \chi_{i}(y) +\frac{1}{8} \sum_{x} \bar{\chi}_{i}(x) \chi_{i}(x) \sum_{ \langle \tilde{x},x \rangle} \sigma(\tilde{x}) \right) +\frac{N}{4g^{2}} \sum_{\tilde{x}} \sigma^{2}(\tilde{x}), \label{gnaction_lattice} \end{equation} where $\chi_{i}$ and $\bar{\chi}_{i}$ are Grassmann-valued staggered fermion fields defined on the lattice sites, the auxiliary field $\sigma$ is defined on the dual lattice sites \cite{cohen}, and the symbol $\langle \tilde{x},x \rangle$ denotes the set of 8 dual lattice sites $\tilde{x}$ surrounding the direct lattice site $x$. The lattice spacing $a$ has been set to one for convenience. Details and motivation of this particular scheme can be found in \cite{kogut93}. The fermion kinetic operator $ M $ is given by \begin{equation} M_{x,y} = \frac{1}{2} \left[ e^{\mu} \delta_{y,x+\hat{0}} - e^{-\mu} \delta_{y,x-\hat{0}} \right] + \frac{1}{2} \sum_{\nu=1,2} \eta_{\nu}(x) \left[ \delta_{y,x+\hat{\nu}} - \delta_{y,x-\hat{\nu}} \right], \end{equation} where $\eta_{\nu}(x)$ are the Kawamoto-Smit phases $(-1)^{x_0+...+x_{\nu-1}}$. The influence of the chemical potential $\mu$ is manifested through the timelinks following \cite{chem.latt}. Only fermion loops which wrap around the timelike direction are affected by its inclusion. The energy density $\epsilon$, and the fermion number density $n$, are defined by \begin{equation} E = - \frac{1}{V_{s}} \frac{ \partial \ln Z}{\partial \beta} = \frac{1}{V} \mbox{tr} \partial_{o} \gamma_{o} S_{F} = \frac{1}{2 V} \langle \sum_{x} e^{\mu} M^{-1}_{x,x+\hat{0}} - e^{-\mu} M^{-1}_{x,x-\hat{0}} \rangle, \end{equation} \begin{equation} n = - \frac{1}{V_{s} \beta} \frac {\partial \ln Z}{\partial \mu} = \frac{1}{V} \mbox{tr}\gamma_{o} S_{F} = \frac{1}{2V} \langle \sum_{x} e^{\mu} M^{-1}_{x,x+\hat{0}} + e^{-\mu} M^{-1}_{x,x-\hat{0}} \rangle. \end{equation} Here $V_s$ is the spatial volume and $V=V_s \beta$ the overall volume of spacetime. The final expression in each case is the quantity measured in the simulation, using a noisy estimator to calculate the matrix inverses. The cubic lattice has $L_{s}$ lattice spacings $a$ in spatial directions and $L_t$ lattice spacings in the temporal direction. The momentum cutoff scale can be defined as $\Lambda= 1/a$, the temperature is given by $T=1/(L_ta)$ and the mass of the sigma meson is $m=1/(\xi a)$, where $\xi$ is the correlation length of the sigma field. To reach a continuum limit one has to satisfy the following two conditions: $\Lambda\gg T$ and $\Lambda\gg m$. The condition $\Lambda\gg T$ requires a lattice with sufficiently large $L_t\gg 1$. The parameters of the action should then be tuned towards their critical values where the correlation length $\xi \gg 1$. This satisfies the condition $\Lambda\gg m$. \section{Monte Carlo simulations} We performed Monte Carlo simulations of the four-fermion model in $d=2+1$ dimensions to study its phase structure in the $(\mu, T)$ plane. We used a hybrid Monte Carlo method described in \cite{kogut93}, which proved to be very efficient for our purposes. Since the chiral symmetry is discrete we were able to simulate the model directly in the chiral limit, i.e., setting the fermion bare mass to zero. The three-dimensional character of the model and its relatively simple form allowed us to perform simulations on large lattices and generate a large number of trajectories relative to state-of-the-art lattice QCD simulations. This allowed a particularly accurate determination of the order of the phase transition at various points in the $(\mu, T)$ plane. In order to check whether it is possible to detect next-to-the-leading order corrections to the large $N$ calculation we set $N=4$. This is the smallest number of fermion species allowed by the hybrid Monte Carlo algorithm, because in order to have a positive semi-definite fermionic determinant, $N$ must be a multiple of four. We used the following two methods to optimize the performance of the hybrid Monte Carlo procedure. The first method consisted of tuning the effective number of fermion flavors $N^{'}$, which is used during the integration of the equations of motion along a microcanonical trajectory, so as to maximize the acceptance rate of the Monte Carlo procedure for a fixed microcanonical time-step $d\tau$. As the lattice size was increased, the time step $d\tau$ had to be taken smaller and the optimal $N'$ approached $N$. For example, for the $N=4$ theory on a $16 \times 32^{2}$ lattice the choices $d\tau=0.12$ and $N'=4.026$ gave acceptance rates greater than $90\%$ for all couplings of interest. To maintain this acceptance rate on a $16 \times 64^{2}$ lattice we used $d\tau=0.10$ and $N'=4.018$. The Monte Carlo procedure was also optimized by choosing the trajectory length $\tau$ at random from a Poisson distribution with the mean equal to $\bar{\tau}$. This method of optimization, which guarantees ergodicity, was found to decrease autocorrelation times dramatically \cite{HaKoKoRe94}. For most of our runs we used the average trajectory length $\bar{\tau} \simeq 2.0$. As usual, the errors were calculated by the jackknife blocking, which accounts for correlations in a raw data set. \subsection{$\mu = 0$, $T=0$} \label{m=0.t=0} In this section we present the results of the simulations at zero temperature and zero chemical potential. The purpose of these simulations is to determine the scaling window of the theory at $T=0$ and $\mu=0$. The knowledge of the scaling window is necessary in order to ensure that the value of the lattice coupling $\beta \equiv 1/g^2$ used in the simulations at non-zero temperature and density is sufficiently close to the bulk critical coupling $\beta_c$, which means that lattice artifacts are negligible. In other words, we verified that another important physical parameter, the zero-temperature mass, $m_0$, is sufficiently smaller than the cutoff $\Lambda$. We ran on large symmetric lattices, such as $20^3$ with $N=4$ for $\beta=[0.500, 0.750]$ and determined the magnetic critical exponent $\beta_{mag}$ by fitting the values of the order parameter $\Sigma_0 \equiv \langle \sigma \rangle$ to the scaling relation $\Sigma_0 = A(\beta_c - \beta)^{\beta_{mag}}$. We found that $\beta_c \approx 0.845$ and $\beta_{mag}=1.00(2)$ which is in agreement with the analytical prediction $\beta_{mag}=1+{\cal O}(1/N^2)$ for the $T=0$, $\mu=0$ scaling \cite{kogut93}. According to the graph in Fig.~\ref{fig:sigma_t=0_mu=0}, the range of $\beta$ for which the $(\Sigma_0, \beta)$ points fall on the `scaling' line is $\beta \approx [0.55,0.75]$. This confirms that for these values of the coupling the lattice theory remains in the scaling window for all range of temperatures down to $T=0$ and effects of the lattice are negligible. \subsection{$\mu \neq 0$, $T = 0$} \label{m.neq.0_t=0} In this section we discuss the results from numerical simulations of the model at non-zero chemical potential and zero temperature. The mean field theory approximation \cite{hands_kocic_kogut.93} predicts that the finite $\mu$ phase transition is everywhere second order except at the isolated point $T=0$. The issue raised by our lattice simulations in this section is whether the first order nature of the transition predicted by mean field theory at $T=0$ persists, when the theory has a small number of fermion species. In addition these simulations are motivated by our interest to study the difficulties in determining the order of the transition using detailed finite size scaling methods. We simulated the system at $\beta=0.625$ because this value of $\beta$ is deep in the broken phase and is also in the scaling window of the bulk critical point which means we can extract continuum physics from these simulations by using standard methods. The order parameter for these simulations is defined by $\Sigma \equiv \langle |\sigma| \rangle$. The absolute value of $\sigma$ is necessary on a finite size lattice in order to take into account the tunneling events between the degenerate $Z_2$-symmetric vacua when the parameters are tuned close to the transition. In the thermodynamic limit $\Sigma \rightarrow \langle \sigma \rangle$. Figure~\ref{fig:sigma_t=0_mu.nonzero} shows the order parameter $\Sigma$ versus $\mu$ for fixed $\beta=0.625$ on $16^{3}$, $22^{3}$, $30^{3}$ and $40^{3}$ lattices. The values of $\Sigma$ on the bigger lattices show no $\mu$ dependence until $\mu$ reaches the vicinity of the critical chemical potential $\mu_c$. The variation of $\Sigma$ becomes more abrupt as the volume of the system increases, indicating that a step-function type singularity will develop in the thermodynamic limit. The data on the $40{^3}$ lattice show a jump discontinuity in $\Sigma$ as $\mu$ varies from $0.3850$ to $0.3875$. The fermion number density $n$ is plotted in Fig.~\ref{fig:density_t=0_mu.nonzero} as a function of $\mu$ for various lattice sizes. The transition at $\mu_c \approx 0.3850$ is seen equally well in this quantity. Futhermore, the variation of $n$ near the transition becomes more abrupt as the volume increases, indicating again that a step-function singularity will develop in the thermodynamic limit. It is clear from the behavior of the various observables that relatively large lattices are required to simulate the theory's actual critical behavior. The discontinuous transition seen on a $40^{3}$ lattice is replaced by a relatively smooth curve on the $16^{3}$ lattice. In order to compare the fermion dynamical mass $m_f$ at $\mu=0$ with with the critical value of the chemical potential $\mu_c$ we calculated the fermion propagator $G({\bf x}, t)$ on the $40^{3}$ lattice. Then, we formed a zero momentum fermion propagator \begin{equation} C_f(t)=\sum_{\bf x} G({\bf x}, t). \end{equation} After calculating the average and the covariance matrix of $C_f(t)$ we fitted it to the following functional form: \begin{equation} C_f(t)=A[e^{m_f t}-(-1)^{t}e^{-m_f(L_t-t)}]. \end{equation} The value of the fermion dynamical mass is $m_f=0.403(1)$, which is according to our expectation very close to $\mu_c$. The small discrepancy is due to interaction and finite size effects. Another interesting indicator which is often used to determine the order of the transition was proposed by Binder and Landau \cite{binder_84}, namely the reduced cumulant of the bosonic action $S$ defined by \begin{equation} B_S=1- \frac{\langle S^{4} \rangle}{3 \langle S^{2} \rangle^2}. \end{equation} This quantity should approach $\frac{2}{3}$, if the distribution of the bosonic action is described by a single Gaussian form with the width vanishing at infinite volume, such as in the case of a second order transition; otherwise it deviates from this value. In practice we calculate $B_S$ for several values of the chemical potential near the transition for each of several lattice sizes. Then a plot of $B_S$ versus chemical potential is made for each lattice size. An accurate value of $B_{S,min}$ for each lattice size is determined and then $B_{S,min}$ is plotted versus $\frac{1}{V}$. If the transition is second order, this plot should have a $y$-intercept (representing the infinite volume limit) approaching the value of two-thirds. In order to locate $B_{S,min}$ accurately we performed simulations at very small intervals of $\mu$ near the transition. We accumulated approximately 30,000 to 40,000 trajectories for the various lattice sizes at each value of $\mu$ near $\mu_c$. The cumulant as a function of $\mu$ is plotted in Fig.~\ref{fig:cumulant_t=0_mu.nonzero} for $L=16,18,22,30$. In Fig.~\ref{fig:cumulant_min.t=0_mu.nonzero} we plot $B_{S,min}$ versus inverse volume $\frac{1}{V}$. The straight line in Fig.~\ref{fig:cumulant_min.t=0_mu.nonzero} represents the best linear fit to the data and its $y$-intercept is equal to 0.66475(14), which is distinctly different from the value of $\frac{2}{3}$, indicating clearly that the model has a first order finite density and zero temperature phase transition. A similar result was obtained from the analysis of the reduced cumulant behavior of the energy density. Our simulations were unable to distinguish a possible nuclear matter liquid-gas transition from the chiral transition. In the large $N$ approximation the two transitions coincide \cite{hands_kocic_kogut.93}. The data in Fig.~\ref{fig:density_t=0_mu.nonzero} for $n$ versus $\mu$ show only one phase transition at $\mu_c$. The smooth tail in the density before the transition is a finite size effect which dies out as the volume of the system increases. We extended our search for a non-zero $n$ at $\mu<\mu_c$ by performing new simulations on large lattices ($30^3, 40^3, 54^3$) at $\beta=0.45$ which is far away from the continuum limit. In this case the correlation length is smaller than before which implies smaller finite size (temperature) effects. The chiral phase transition at this value of the coupling takes place at $\mu_c=0.692(1)$. Again, we did not detect a size-independent saturation in the fermion number density at $\mu<\mu_c$. At $\mu=0.690$ the fermion number density $n$ is $0.0118(4)$ on the $30^3$ lattice, $0.0040(2)$ on the $40^3$ lattice and $0.00126(5)$ on the $54^3$ lattice. This implies that for $N=4$ the next-to-the-leading order corrections are very small if not zero, in the sense that either the two transitions are very close or the nuclear matter liquid-gas transition is so weak that is smoothed out by finite size effects or it doesn't exist in this model. \subsection{$\mu \neq 0$, $T \neq 0$} \label{m.neq.0_t.neq.0} In this section we present the results from the simulations of the four-fermion model at non-zero temperature and chemical potential. The purpose of these simulations is to study the phase structure of the model in the $(\mu, T)$ plane. The large $N$ result predicts a second order phase transition for $T>0$. We varied $T$ by changing either the lattice temporal extend $L_t$ or the lattice spacing $a$ which according to the discussion in Sec.~\ref{m=0.t=0} vanishes when the coupling $\beta$ is tuned to its $T=\mu=0$ critical value $\beta_c \approx 0.845$. In Fig.~\ref{fig:sigma_t_mu_normalized} we plot the normalized order parameter $\Sigma(T,\mu)/\Sigma(T=0,\mu)$ as a function of $\mu/m_f$ at different values of $T$. This normalization is essential in order to convert the order parameter and the chemical potential into physical units by getting rid of the lattice spacing. It therefore enables us to compare the behavior of the order parameter as a function of the chemical potential at different temperatures on lattices with different lattice spacings. According to the discussion in Sec.~\ref{m=0.t=0} the lattice discretization effects cannot be neglected when $\beta$ is smaller than $0.55$. The parameters for the curves in Fig.~\ref{fig:sigma_t_mu_normalized} from left to right in terms of $(\beta$, lattice size, $T/T_c)$ are: $(0.65, 8\times 48^{2}, 0.71)$, $(0.65, 16\times48^{2}, 0.36)$, $(0.55, 16\times48^{2}, 0.23)$, $(0.55, 24\times48^{2}, 0.16)$, $(0.625, 40^{3}, 0.0)$. It is clear from the shapes of these curves that the transition becomes sharper as we decrease the temperature. In Fig.~\ref{fig:phase.dgr} we map out the phase diagram of the model in the $(\mu/m_f, T/m_f)$ plane. As the temperature increases from zero, the chiral condensate `melts' at a smaller value of the critical chemical potential $\mu_c(T)$ which suppresses energetically the quark$-$anti-quark pairing. After having determined the order of the phase transition at zero temperature (see previous section), we extended our work to determine the order of the transition at the three points of Fig.~\ref{fig:phase.dgr} with the lowest non-zero values of $T/T_c$. At these points the variation of $\Sigma(T,\mu)/\Sigma(T=0, \mu)$ as a function of $\mu/m_f$ is sharp near the transition. We first present the results for the order parameter $\Sigma$ and the fermion number density $n$ versus $\mu$ in Figs.~\ref{fig:sigma_mu.non0_t.non0} and \ref{fig:density_mu.non0_t.non0} respectively. In each case we plot the values of the observables for various values of the lattice spatial extend $L_s$. For $T/T_c=0.36$, $L_s=32, 48, 64$; for $T/T_c=0.23$, $L_s=32, 48, 64$ and for $T/T_c=0.16$, $L_s=48, 72$. These figures show a marked contrast with the corresponding ones at $T=0$, $\mu \neq 0$ discussed in the previous section. In each case an asymptotic behavior is approached without mutual crossings of curves for different $L_s$, in contrast to Figs.~\ref{fig:sigma_t=0_mu.nonzero} and \ref{fig:density_t=0_mu.nonzero}. However, as the temperature is lowered the variations of the thermodynamic observables become sharper near the transition and it is therefore very difficult to distinguish between a weak first order and a second order transition. It should also ne noted that in Fig.~\ref{fig:density_mu.non0_t.non0} the values of the non-zero $n$ before the transition do not depend on $L_s$. This implies that the tail in $n$ for $\mu < \mu_c$ is not a finite $L_s$ effect; it is instead a finite $L_t$ or non-zero temperature effect. For a more accurate determination of the order of the transition we performed again a more detailed finite size scaling analysis. We estimated the reduced cumulants $B_S$ for the bosonic action and $B_{E}$ for the internal energy density. For a precise calculation of these quantities we accumulated $30,000 - 50,000$ trajectories for the various lattices near $\mu_c$ at $T/T_c=0.36$ and $90,000 - 120,000$ trajectories at $T/T_c=0.23$ and $T/T_c=0.16$. We plot $B_{E}$ and $B_S$ versus $\mu$ in Figs.~[10-14]. Figure~\ref{fig:cum.all.min} shows the minimum values of $B_{E}$ and $B_S$ as functions of the inverse volume. The straight lines represent the best linear fits to the data. Their $y$-intercepts for the three different parameter sets are : $B_S=0.66667(3)$, $B_{E}=0.66668(3)$ for $T/T_c=0.36$; $B_S=0.66628(13)$, $B_{E}=0.66641(13)$ for $T/T_c=0.23$ and $B_S=0.6660(3)$, $B_{E}=0.6658(4)$ for $T/T_c=0.16$. The results of this analysis confirm that the transition at $T/T_c=0.36$ is a second order transition. They also suggest that at the two lower temperatures the transitions are weak first order, although the possibility that these might also be second order transitions cannot be excluded decisively. Studies on much larger lattices are required in order to detect the order of these transitions conclusively. Nevertheless, the steep variations of the thermodynamic quantities near the transitions at $T/T_c=0.23, 0.16$ together with the deviations of the thermodynamic limit of the cumulants' minima from the $\frac{2}{3}$ value support the fact that the system is close to its tricritical point. \section{Conclusions} In this manuscript we presented the results of Monte Carlo simulations of the three-dimensional four-fermion model with $N=4$ fermion species at non-zero temperature and density. With the exception of confinement this model incorporates most of the essential properties of QCD. The great advantage of this model is that it is possible to perform lattice simulations at non-zero chemical potential because its fermionic determinant remains real at $\mu \neq 0$. We demonstrated using finite size scaling techniques that the $\mu \neq 0$, $T=0$ phase transition is a first order transition. Our analysis did not succeed to distinguish the nuclear matter liquid-gas transition from the chiral phase transition. Either the two transitions are extremely close or the nuclear matter liquid-gas transition which is expected to occur at $\mu < \mu_c$ is very weak and is smoothed out by finite size effects or it doesn't exist in this model. We have also shown that the second order nature of the $T \neq 0$, $\mu=0$ phase transition remains stable for $T/T_c>0.23$ and $\mu/\mu_c<0.97$. This imples that the large $N$ approximation (or mean field theory description of the transition) which predicts a second order phase transition everywhere except at the isolated point $T=0$ \cite{hands_kocic_kogut.93} describes the order of the transition on the largest part of the critical line in the $(T, \mu)$ plane. In this sense the model has a ``soft'' behavior on the biggest section of the critical line. It should be noted that the mean field approximation does not predict the correct two-dimensional Ising universality class \cite{kogut_98} of the second order phase transitions line. In that case non-perturbative effects are crucial in the critical region and the mean field results are modified by $1/N$ corrections. However, the results of this study show that corrections to the large $N$ limit for other characteristics of the model, including the position of the tricritical point and the position of the nuclear matter liquid-gas transition, are very small indeed. A simple continuity argument implies that the model has a tricritical point on the critical line at a low temparature and very high density. Although our simulations suggest that at low $T$ and high $\mu$ the phase transition is a weak first order transition, studies on much larger lattices would be required to ``locate'' the tricritical point unambiguously. Various approaches to QCD with two massless quarks at finite temperature and density suggest the existence of a tricritical point on the boundary of the phase with spontaneously broken chiral symmetry. The position of the tricritical point in two-flavor QCD was estimated recently using a random matrix model \cite{shuryak} and a Nambu$-$Jona-Lasinio model \cite{berges} as $T_{tr} \approx 100$ MeV and $\mu_{tr} \approx 600-700$ MeV. Possible experimental signatures suggested in \cite{stephanov99} should allow future experiments at RHIC and LHC to provide information about the location and properties of this point. Another conclusion drawn from our work is that lattices with very large spatial extends are needed in order to observe the order of the transition. Our studies on small volumes highlight the difficulties that must be faced when trying to understand the critical behavior of the thermodynamic limit by extrapolation from systems away from this limit. It was also shown that qualitative signatures of a first order phase transition such as the variation of various thermodynamic quantities near the transition can be very misleading unless they are accompanied by a detailed finite size scaling analysis. This should be a warning to the lattice community that a quantitative understanding of the critical behavior of QCD at non-zero chemical potential lies along a difficult road. \section*{Acknowledgements} Discussions with Simon Hands, Misha Stephanov and Pavlos Vranas are greatly appreciated. This work was supported in part by NSF grant PHY96-05199. CGS is supported by a Leverhulme Trust grant. The computer simulations were done on the Cray C90's and J90's at NERSC and on the NOW at SDSC. \newpage
\subsubsection*{\it I--I\,$^\prime$ Duality} We shall now describe the sequence of isomorphisms \eqn{botthomKO} in the language of brane configurations, as we did in the type-II case. For definiteness we shall take $l=2k$ to be even. The isomorphism in the first line of \eqn{botthomKO} arises in a similar way as in the type-II case by representing a $p-1$-brane in codimension $2k+1$ in terms of $2N$ unstable 9-branes. The induced charge takes values in the higher KO-group \begin{equation} \widetilde{KO}^{-1}(S^{l+1})=\pi_l\Bigl(O(2N)/[O(N)\times O(N)]\Bigr) \label{KO1groups}\end{equation} and the KO-class is determined by the rank $2N$ spinor bundle ${\cal S}_{\frac N2}\oplus{\cal S}_{\frac N2}$ with tachyon field \begin{equation} T_N^{({\rm I}\,')}=T_N^{\rm(I)}\oplus T_N^{\rm(I)}+T^{(1)}\,\sigma_3\otimes I_N \label{T2NIprime}\end{equation} As in \eqn{TNI}, eq. \eqn{T2NIprime} shows that the given set of spacetime filling 9-branes interact via tachyon fields only among themselves and not with their mirror images. There is, however, a subtlety with this transformation which is related to the doubling \eqn{T2NIprime} in rank of the spinor modules due to the mirror IIB 9-branes (representing again a K-theoretic stabilization). Under a $T$-duality transformation on $S^1$ the operator $\Omega$ is mapped to $\Omega\cdot{\cal I}_1$, where ${\cal I}_1$ is the geometrical operator which reflects the compactified target space direction. The $T$-dual of the type-I theory is type-I$\,^\prime$ superstring theory which is obtained as the $\zed_2$ orientifold projection of the type-IIA theory by $\Omega\cdot{\cal I}_1$. The relevant K-theory is given by the Real K-group $KR(X)$ \cite{witten,atiyahreal}, i.e. the group of virtual bundles $[(E,F)]$ with an antilinear involution on both $E$ and $F$ which commutes with the induced action of ${\cal I}_1$ by pullback. As the type-I$\,^\prime$ theory contains $2N$ unstable 9-branes, its D-brane charges live in the Real K-group $\widetilde{KR}^{-1}(X)$ \cite{horava}. The relation to the present homotopy analysis comes from the fact that if the KR-involution is taken to act trivially on $X$, then there is a natural isomorphism \begin{equation} KR^{-n}(X)=KO^{-n}(X) \label{KRKOiso}\end{equation} The precise equivariant homotopy properties inherent in Real K-theory will be discussed in more detail in the next section. We see that the periodicity theorem \eqn{botthomKO} will in fact go beyond describing just toroidal compactifications of the type-I theories. A $T$-duality transformation of type-I superstring theory on $T^m$ gives a type-II orientifold on $T^m/\zed_2$ by the operator $\Omega\cdot{\cal I}_m$ (${\cal I}_m$, along with appropriate factors of $\klein$, reflects all of the coordinates of the compactification torus $T^m$). In the following we shall also describe the structure of these orientifold theories using K-theoretic properties. A general orientifold has $2^m$ orientifold planes which are the fixed points of the $\Omega\cdot{\cal I}_m$ involution and which fill all of the non-compact spacetime. They have dimension $9-m$ and carry $-2^{4-m}$ units of $9-m$-brane RR-charge. We will denote them by O$(9-m)^-$. Far away from the orientifold planes, the spacetime looks like the original $X$ along with its mirror image under the action of the KR-involution which interchanges the two copies. An elementary result of K-theory shows that for a trivial $\zed_2$ action on the space $X$ \cite{atiyahreal} \begin{equation} KR^{-n}(X\amalg X)=KR^{-n}(X\times S^{0,1})=K^{-n}(X) \label{KRKiso}\end{equation} where $S^{p,q}$ denotes the $p+q-1$ dimensional unit sphere in $\real^{p,q}$. The theory far away from the singularities is therefore equivalent to the original type-II theory. On the other hand, branes on top of an orientifold plane have their generic unitary gauge symmetry broken to the real subgroup, in accordance with the first part of the sequence \eqn{botthomKO}. As we did in the type-II cases, we shall follow the orbit of the original stable 9-$\overline{9}$ brane pairs in the homotopy sequence, and in addition keep track of the structure of the orientifold planes since their RR-charges will be measured by the relevant K-theory groups. At the level of (\ref{KO1groups},\ref{T2NIprime}), there are $N$ stable 8-$\overline{8}$ brane pairs and two O$8^-$ planes which each carry $-8$ units of RR-charge. Tadpole anomaly cancellation requires that the 8-branes be arranged so as to cancel out the orientifold charge in the supersymmetric vacuum configuration. Note that the K-group mapping between \eqn{KOSgroups} and \eqn{KO1groups}, i.e. the basic I-I$\,^\prime$ $T$-duality, involves the homotopy groups of real symmetric spaces. \subsubsection*{\it m = 2} Next we come to the first isomorphism in the second line of \eqn{botthomKO}. This corresponds to the toroidal compactification of the type-I theory which is $T$-dual to type-IIB superstring theory on the $T^2/\zed_2$ orientifold. There are now four O$7^-$ planes, each with $-4$ units of 7-brane RR-charge, and $N$ 7-$\overline{7}$ brane pairs which are described in terms of $2N$ 9-$\overline{9}$ brane pairs, as in \eqn{tachyonN+1}. The appearence of the unitary group in \eqn{botthomKO} is due to the following facts. The chiral spinor modules $\Delta_2^\pm$ are complex, so that the desired map must be taken with respect to the real spinor module $\Delta_2^+\oplus\Delta_2^-$, as in \eqn{tachyonN+1} (see also \eqn{Bottspinmap}). This means that the relevant homotopy is defined with respect to a unitary symmetric space. Physically, the appearence of a unitary gauge symmetry can be understood from the analysis of \cite{gimonpol} (see also \cite{quivers}) where the requirement of closure of the worldsheet operator product expansion was shown to put stringent restrictions on the actions of discrete gauge symmetries on Chan-Paton bundles. In particular, the square of the worldsheet parity operator $\Omega$ acts on Chan-Paton indices as \begin{equation} \Omega^2\,:\,~~~~|{\rm D}p;ab\rangle~\mapsto~(\Lambda_\Omega^2)^{a'}_a\,|{\rm D}p;a'b'\rangle\,(\Lambda_\Omega^{-2})^{b'}_b=(\pm i)^{(9-p)/2}\,|{\rm D}p;ab\rangle \label{Om2action}\end{equation} where $a,b$ are the open string endpoint Chan-Paton labels of a D$p$-brane state of the IIB theory, and $\Lambda$ denotes the (adjoint) representation of the orientifold group in the Chan-Paton gauge group. While the 9-branes have the standard orthogonal subgroup projection (as required by tadpole anomaly cancellation), \eqn{Om2action} leads to an inconsistency on 7-branes which are therefore quantized using the unprojected unitary gauge bundles. Thus, while the naive gauge group on the spacetime filling 9-branes is $O(2N)\times O(2N)\subset O(4N)$, the inconsistent $\Omega$-projection on IIB 7-branes enhances the orientifold symmetry to $U(2N)$ because from the point of view of worldvolume gauge fields the 9-branes are indistinguishable from $\overline{9}$-branes due to the reality property of the relevant spinor bundles. The requisite tachyon field $T_{2N}^{\rm(I)}$ of the corresponding virtual bundle $[(E,F)]$ can be regarded as a section of $E\otimes\overline{F}$ and so is required to be $\zed_2$-equivariant with respect to the orientifold projection (this ensures that the resulting lower dimensional brane configurations are invariant under the $\zed_2$-action), i.e. it transforms under the orientifold group as \begin{equation} T_{2N}^{\rm(I)}\to\Lambda_\Omega\,T_{2N}^{\rm(I)}\,\Lambda_\Omega^{-1} \label{tachyonOmtransf}\end{equation} As shown in \cite{witten}, the tachyon vertex operator for a $p$-$\overline{p}$ brane pair acquires the phase $(\pm i)^{7-p}$ under the action of $\Omega^2$. For the 7-branes this operator is even under $\Omega^2$, and so the eigenvalues of the vacuum expectation value $T_{2N}^{\rm(I)0}$ are real. Thus the tachyon field breaks the $U(2N)$ gauge symmetry down to its orthogonal subgroup $O(2N)$, and the induced brane charge is given by the winding numbers around the vacuum manifold $U(2N)/O(2N)$ of the IIB orientifold on $T^2/\zed_2$. What is particularly interesting about this orientifold is that it is equivalent to a more general, dual description in terms of $F$-theory \cite{Ftheory}. In this description, $F$-theory on the third complex Kummer surface $K3$ is the compactification of type-IIB superstring theory on the base space of the elliptic fibration $K3\to\complex P^1$ with complex structure modulus of the fiber given by the axion-dilaton modular parameter of the IIB theory. At the special point of the $K3$ moduli space where it can be identified with the orbifold $T^4/\zed_2$ (i.e. the degenerate $K3$ with eight $D_2$ singularities), the fiber acquires singularities and the base space is the orbifold limit $T^2/\zed_2$ of $\complex P^1$. The orientifold planes split into 7-branes in this description and have some dynamics (although they cannot support perturbative gauge fields). The equivalence between $F$-theory on $K3$ and type-I superstring theory on $T^2$ is then a starting point for a K-theory description of $F$-theory. In this particular compactification of $F$-theory, the relevant vacuum manifold for tachyon condensation is $U(2N)/O(2N)$ and a K-theory subgroup of the full compactification is $\widetilde{KR}^{-2}(Y\times T^{1,2})$ (or equivalently $\widetilde{KO}(Y\times T^2)$ \cite{bgh} as we show in the next subsection). Here we have used the definition $T^{1,m}=(S^{1,1})^m$. The relevant K-theory incorporates the charges of the orientifold planes, as we will discuss further in the following. \subsubsection*{\it m = 3} The second isomorphism in the second line of \eqn{botthomKO} describes the mapping onto the I$\,^\prime$ theory on $T^3$ which is $T$-dual to the $T^3/\zed_2$ orientifold of the type-IIA string. There are eight O$6^-$ planes each with $-2$ units of 6-brane RR-charge. The appearence of a symplectic gauge group follows from the mathematical fact that the complex spinor module $\Delta_3$ is the restriction of a quaternionic Clifford module, so that the appropriate augmentation of the spin bundles on the 9-branes is taken with respect to the rank 4 real representation $\Delta_3\oplus\Delta_3$. This means that there are now $4N$ unstable 9-branes which have an $Sp(2N)$ worldvolume gauge symmetry and whose KO-theory class is represented by the spin bundle ${\cal S}_{N}\oplus{\cal S}_{N}$ with tachyon field \begin{equation} T_{2N}^{({\rm I}\,')}=T_{2N}^{\rm(I)}\oplus T_{2N}^{\rm(I)}+T^{(1)}\,\sigma_3\otimes I_{4N} \label{tachyonm3}\end{equation} This enhanced $Sp(2N)$ gauge symmetry comes from the intermediate representation of a given type-I$\,^\prime$ $p-3$-brane in terms of 6-$\overline{6}$ brane pairs and will be more easily understood below at the next isomorphism in terms of 5-branes. Again by $\zed_2$ equivariance the tachyon field breaks this symmetry to its complex subgroup $U(2N)$, so that the vacuum manifold is $Sp(2N)/U(2N)$. As shown in \cite{horava}, the codimension 3 tachyonic soliton in \eqn{tachyonm3} coincides with the usual 't Hooft-Polyakov magnetic monopole. Since $M$-theory compactified on a circle is the type-IIA string, a careful analysis \cite{Mtheory} reveals that this orientifold model is equivalent to $M$-theory defined on the orbifold limit $T^4/\zed_2$ of the 4-manifold $K3$. This lends another clue to the puzzle of how to describe $M$-theory in K-theoretic terms \cite{witten,horava}. The equivalence with the IIA theory when compactified on a circle shows that its description must then reduce to a vacuum manifold $U(2N)/[U(N)\times U(N)]$ describing the stable soliton configurations which represent the branes with K-group $K^{-1}(X)$, while the equivalence of the type-IIA orientifold on $S^1/\zed_2$ to the $M$-theory orientifold on $S^1/\zed_2$ is described in terms of the vacuum manifold $O(2N)/[O(N)\times O(N)]$ and the KR-group $\widetilde{KR}^{-1}(Y'\times S^1)$. Now we see that the duality between $M$-theory compactified on a $K3$ surface and type-I$\,^\prime$ superstring theory on $T^3$ implies that this 11-dimensional compactification has vacuum manifold $Sp(2N)/U(2N)$ and K-theory group $\widetilde{KR}^{-3}(Y\times T^{1,3})$ (or equivalently $\widetilde{KO}(Y\times T^3)$ \cite{bgh} as we discuss in the next subsection). \subsubsection*{\it m = 4} Now we come to the isomorphisms in the third line of \eqn{botthomKO}. The superstring theory is type-I on $T^4$ which is $T$-dual to the orientifold of type-IIB on $T^4/\zed_2$ which has 16 O$5^-$ planes each with a negative unit of 5-brane RR charge. The rank 2 spin modules $\Delta_4^\pm$ are again the restrictions of quaternionic Clifford modules and yield the real chiral spinor representations $\Delta_4^\pm\oplus\Delta_4^\pm$. There are now $4N$ 9-$\overline{9}$ brane pairs which determine the KO-theory class $[({\cal S}_{2N}^+\oplus{\cal S}_{2N}^+,{\cal S}_{2N}^-\oplus{\cal S}_{2N}^-)]$ with tachyon field \begin{equation} T_{4N}^{\rm(I)}=T_{2N}^{\rm(I)}\otimes(\sigma_3\oplus\sigma_3)+ T_2^{\rm(I)}\otimes I_{4N} \label{TI4N}\end{equation} where the codimension 2 tachyon field $T_2^{\rm(I)}$ is defined as in \eqn{tachyonN+1}. The quaternionic gauge symmetry is naturally explained by \eqn{Om2action} which shows that $\Omega^2=-1$ when acting on the $N$ 5-brane-antibrane states (and also on the corresponding tachyon vertex operator). The 5-branes must therefore be quantized using pseudo-real gauge bundles, i.e. Chan-Paton bundles with structure group $Sp(2N)$ on the 9-branes and on the $\overline{9}$-branes. This fact explains, via $T$-duality transformations, the appearence of symplectic gauge groups in the $m=3$ case above and in the cases to follow. An alternative explanation \cite{witten} utilizes the fact that a type-I 5-brane is equivalent to an instanton on the spacetime filling 9-branes \cite{witteninst}. The tachyon field breaks the $SO(4N)\times SO(4N)$ gauge symmetry of the 9-$\overline{9}$ brane configuration to the diagonal subgroup $SO(4N)_{\rm diag}$, which is then further broken down to $Sp(2N)$ by the instanton field. K-theory defined with pseudo-real bundles is denoted $KSp(X)$. The induced $p-4$-brane charge of type-I superstring theory on $T^4$ is labelled by the homotopy groups of the vacuum manifold $Sp(2N)\times Sp(2N)/Sp(2N)_{\rm diag}$ which yield the KSp-group \begin{equation} \widetilde{KSp}(S^{l+4})=\pi_{l+3}\Bigl(Sp(2N)\Bigr) \label{KSpSl4}\end{equation} In the language of the orientifold theory, the replacement of the KR-involution by a $\zed_4$ generator defines the pseudo-Real K-group $KH(X)$ \cite{kh} appropriate to orientifold theories with symplectic Chan-Paton bundles.\footnote{Strictly speaking, the $\Omega\cdot{\cal I}_m$ orientifold projection should be accompanied by the action of the operator $(-1)^{\frac12(9-p)(8-p)F_{\rm L}}$ to preserve the $\zeds_2$-equivariant structure in the homotopy sequence. The usage of the KH-theory generalization of the orientifold models has been discussed in \cite{gukov,hori}.} The present model is dual to a conventional orbifold of the type-IIA theory on $T^4/\zed_2$ and its moduli space coincides with that of IIA strings on $K3$ \cite{bsv}. \subsubsection*{\it m = 5} The situation for type-I$\,^\prime$ theory on $T^5$ introduces a new chain of dualities which may be attributed to the fact that the orientifold planes now begin acquiring fractional RR-charges. This theory is $T$-dual to the $T^5/\zed_2$ orientifold of type-IIA superstring theory with 32 O$4^-$ planes each of fractional magnetic charge $-\frac12$. This new duality chain can also be understood in the K-theory language from the change of nature of the Chan-Paton spinor bundles on the 9-$\overline{9}$ branes at the $m=4$ case above. Again the spinor module $\Delta_5$ is quaternionic in origin and yields the eight dimensional real spinor representation $\Delta_5\oplus\Delta_5$. This is consistent with the physical expectations from $T$-duality of the pseudo-real vacuum manifold associated with $8N$ unstable 9-branes carrying the Chan-Paton bundle ${\cal S}_{2N}\oplus{\cal S}_{2N}$ (with tachyon field $T_{4N}^{({\rm I}\,')}$ defined analogously to \eqn{tachyonIIA}) which appears in the second isomorphism of the third line of \eqn{botthomKO}. A candidate dual theory is the $K3\times S^1$ compactification of type-IIB superstring theory, which may then be conjectured to be dual to a $T^5/\zed_2$ orientifold of $M$-theory \cite{IIBK3}. This chiral $M$-theory orientifold has no D-branes but rather M5-branes on which M2-branes can end. The chiral lift of the $m=5$ string orientifold should then be related to various equivalent K-groups given by $\widetilde{KR}^{-5}(Y\times T^{1,5})$, $\widetilde{KO}(Y\times T^5)$, $\widetilde{KSp}(Y'\times S^1)$ and $\widetilde{KH}^{-1}(Y'\times S^{1,1})$. Along with the recent construction \cite{yi} of M2-branes as tachyonic-type solitons in the worldvolume of M5-$\overline{\rm M5}$ brane configurations, this could shed more light on the interpretation of $M$-theory using K-theory (some related results can also be found in \cite{matrix}). \subsubsection*{\it m = 6} The type-I superstring on $T^6$ is $T$-dual to the IIB string on $T^6/\zed_2$ which has 64 O$3^-$ planes each of fractional RR-charge $-\frac14$ filling the non-compact space. The $N$ 3-$\overline{3}$ brane pairs, which are described as the tachyonic kinks of $8N$ 9-$\overline{9}$ brane pairs, have the same fate according to \eqn{Om2action} as the 7-branes, and hence the 9-brane gauge symmetry is the unprojected $U(8N)$. Algebraically this owes to the fact that the spinor modules $\Delta_6^\pm$ are complex and thus form the real spinor representation $\Delta_6^+\oplus\Delta_6^-$. The corresponding KO-group element is thus $[({\cal S}_{4N}^+\oplus\overline{\cal S}_{4N}^-,\overline{\cal S}_{4N}^+\oplus{\cal S}_{4N}^-)]$, and $\zed_2$ equivariance of the tachyon field implies that it breaks the $U(8N)$ gauge symmetry to its pseudo-real subgroup $Sp(4N)$. Again this model can be described as the limit of an $F$-theory compactification on the orbifold $T^8/\zed_2$ which has terminal singularities and is not the limit of any suitable smooth 8-manifold. This $F$-theory compactification should thus reduce to the vacuum manifold $U(8N)/Sp(4N)$ around which the tachyonic windings give rise to K-theory classes in $\widetilde{KR}^{-6}(Y\times T^{1,6})$ (or the various other equivalent K-groups). \subsubsection*{\it m = 7} The $T^7/\zed_2$ orientifold of the type-IIA theory has 128 O$2^-$ planes each of electric charge $-\frac18$. The spinor module $\Delta_7$ is real, so that the $16N$ unstable 9-branes support the spinor bundle ${\cal S}_{4N}\oplus{\cal S}_{4N}$ and have a real $O(16N)$ gauge symmetry. This reality condition arises from the intermediate 2-$\overline{2}$ brane pairs and again will be explained at the next isomorphism in terms of D-strings. The tachyon field breaks the $O(16N)$ gauge symmetry to the unitary subgroup $U(8N)$, as follows from equivariance with respect to the orientifold group. This model can be mapped to the compactification of $M$-theory on $T^8/\zed_2$, from which we can identify another limiting vacuum manifold $O(16N)/U(8N)$ and K-group $\widetilde{KR}^{-7}(Y\times T^{1,7})$. \subsubsection*{\it m = 8} The final isomorphism in the sequence \eqn{botthomKO} maps us into the type-I theory on $T^8$ which is $T$-dual to the type-IIB orientifold on $T^8/\zed_2$ that has 256 O$1^-$ planes each of RR-charge $-\frac1{16}$. The vacuum manifold is now the real orthogonal group $O(16N)$ because the intermediate IIB D-strings have the usual orthogonal projection according to \eqn{Om2action}. This is consistent with the fact that the rank 8 chiral spinor modules $\Delta_8^\pm$ are real, and thus the $16N$ 9-$\overline{9}$ brane pairs support Chan-Paton bundles which each transform under the $\Delta_{2k+8}^+\oplus\Delta_{2k+8}^-$ spinor representation of the transverse rotation group, similarly to the case we started with. This orientifold is equivalent to the ordinary orbifold compactification of type-IIA superstring theory on $T^8/\zed_2$, which has fundamental strings condensed in the supersymmetric vacuum. This latter property and its K-theory origin may again be combined with the analysis of \cite{yi} which shows that the solitonic description of an M2-brane from the annihilation of a coincident pair of M5-$\overline{\rm M5}$ branes reduces (upon compactification on $S^1$) to a fundamental string stretched between an annihilating pair of D4-$\overline{\rm D4}$ branes in the IIA theory. The closure of the Bott periodicity sequence \eqn{botthomKO} at this stage implies that there is no new physics at lower compactifications, as is indeed precisely the case. \subsubsection*{\it Hopf Maps} The strong Bott periodicity isomorphisms for real and pseudo-real K-theory take the form \begin{eqnarray} \widetilde{KO}^{-n-8}(X)=\widetilde{KO}^{-n}(X)~~~~~~&, &~~~~~~\widetilde{KR}^{-n-8}(X)=\widetilde{KR}^{-n}(X) \label{bottRealKn}\\\widetilde{KSp}^{-n-4}(X)=\widetilde{KO}^{-n}(X) {}~~~~~~&,&~~~~~~\widetilde{KH}^{-n-4}(X)=\widetilde{KR}^{-n}(X) \label{bottrealKn}\end{eqnarray} They can again be deduced from the periodicity relations $C_{l+8}=C_l\otimes\real(16)$ and $C_{l+4}=C_l\otimes_\reals\quater(2)$ of the corresponding Clifford algebras \cite{spingeom}. As mappings on K-theory classes they can be represented via cup products involving spinor bundles, as in \eqn{Bottspinmap}, or equivalently by using Hopf fibrations as in (\ref{alphaiso},\ref{altBottmap}). The isomorphism of KO-groups in \eqn{bottRealKn} comes from taking the cup product of an element of $\widetilde{KO}^{-n}(X)$ with the generator $[{\cal N}_\reals]-[I^7]$ of $\widetilde{KO}(S^8)=\zed$, where ${\cal N}_\reals$ is the rank 7 Hopf bundle over $\real P^8$ associated with the real Hopf fibration $S^{15}\to S^8$. The corresponding mapping on the Real K-theory groups is obtained via the natural Real structure on the complex Hopf bundle over $\complex P^7$ \cite{atiyahreal}. This shows that the construction of a $p$-brane in terms of $p+8$-branes (e.g. a type-I D-particle from 8-$\overline{8}$ brane pairs) is determined by a D-string solitonic configuration which gives another explicit physical realization of the $spin(8)$ instanton.\footnote{Note that for the orientifold models, the equivariance condition \eqn{tachyonOmtransf} on the tachyon field and a similar one on the worldvolume gauge fields implies that the topological defects arising from the Hopf fibrations are always equivariant versions of these solitons.} The corresponding eight dimensional non-trivial gauge connections, and the associated spinor structures, may be found in \cite{hopf8}. This minimizing solution of the eight dimensional Euclidean Yang-Mills equations satisfies a generalized duality condition with respect to the topological 4-form $F\wedge F$ constructed from the associated field strength. This identifies the explicit form of the worldvolume gauge fields living on the $p+8$-$\overline{p+8}$ brane pair required to produce the finite energy solitonic $p$-brane configuration as \cite{hopf8} \begin{equation} A_i(x)=-2i\,\sum_{j=1}^8\Gamma_{ij}\,\frac{x^j}{(1+|x|^2)^2} \label{8gaugefield}\end{equation} where $\Gamma_{ij}$ are the generators of $spin(8)$. These gauge field configurations are $spin(9)$ symmetric (thereby preserving the manifest spacetime symmetries) and carry unit topological charge. Similarly, the isomorphism of pseudo-real K-groups in \eqn{bottrealKn} comes from taking the cup product with the class of the rank 2 instanton bundle ${\cal N}_\quaters$ associated with the pseudo-real Hopf fibration $S^7\to S^4$ \cite{trautman}, i.e. the holomorphic vector bundle of rank 2 over $\complex P^3$. Thus the relationship between a BPS $p$-brane and a BPS $p+4$-brane is a 5-brane soliton which may be identified with an $SU(2)$ Yang-Mills instanton field. This descent relation was noted in \cite{seninst} in the case of a type-I D-string in the worldvolume of a 5-$\overline{5}$ brane pair. The worldvolume gauge symmetry is $SU(2)\times SU(2)$ and the tachyon field transforms in its ${\bf2}\otimes{\bf\overline{2}}$ representation. The $\Omega$-projection identifies the vacuum manifold of the 5-brane configuration as $SU(2)=Sp(1)$, and the finite energy static string solution in the corresponding 5+1 dimensional field theory has asymptotic boundary $S^3$. By choosing the asymptotic form of the tachyon field as in \eqn{puregauge}, the topological stability of the string is guaranteed by the homotopy group $\pi_3(SU(2))=\zed$. Arranging the asymptotic form of the gauge field on the 5-brane as in \eqn{puregauge} (and that on the $\overline{5}$-brane to vanish), the string soliton carries 1 unit of instanton number which is a source of D-string charge in type-I string theory \cite{witteninst}. These arguments agree precisely with the general homotopy analysis of the $m=4$ case above. We see that the descent relations in type-II and type-I superstring theory give natural realizations of all three higher Hopf fibrations. The elementary Hopf fibration $S^1\to S^1$ with discrete fiber $\zed_2$ arises in the construction of a type-I non-BPS brane as a codimension 1 kink of a brane-antibrane pair, e.g. the type-I D-particle from a D-string anti-D-string pair \cite{witten,seninst}. The double cover of $S^1$ corresponds to the pair of D-strings, and the winding number of the tachyon field is labelled by the homotopy group $\pi_0(\zed_2)=\zed_2$ of the fiber corresponding to the discrete gauge transformation $T^{(1)}\to-T^{(1)}$. This agrees with the degenerate situation of describing an 8-brane in terms of a single 9-$\overline{9}$ brane pair \cite{horava}, whereby the vacuum manifold is $O(1)=\{\pm\,T^{(1)0}\}$ and one of the 9-branes carries a $\zed_2$ Wilson line. The cup product with the generator of $\widetilde{KO}(S^1)=\zed_2$ then achieves the mapping of $\zed_2$-valued KO-theory classes. Therefore, the four fundamental Hopf fibrations are responsible for the complete spectrum of D-brane charges in type-II and type-I superstring theory. These cup products also shed light on the appearence of the rich spectrum of brane charges in the type-I theories as compared to the type-II theories. Since any real vector bundle may be regarded as complex, and any complex one as quaternionic, there is a natural homomorphism between the K-groups of the type-I and type-II theories, \begin{equation} \widetilde{KO}^{-n}(X)~\longrightarrow~\widetilde{K}^{-n}(X)~ \longrightarrow~\widetilde{KSp}^{-n}(X) \label{KSPOmap}\end{equation} As an example of this map, consider the generator of $\widetilde{K}(S^4)=\zed$, which is the pseudo-real instanton bundle described above. To realize it as a generator of $\widetilde{KO}(S^4)=\zed$, which labels type-I 5-brane charge, it must be embedded in the orthogonal structure group as $SO(4)=SU(2)\times SU(2)$ which then doubles the 5-brane charge \cite{witten}. Thus the natural map between the integer-valued K-groups $\widetilde{K}(S^4)$ and $\widetilde{KO}(S^4)$ is multiplication by 2, $\zed\to2\zed$. Upon forming cosets in the ring structure of the KO-groups (see \eqn{absisoKO}), we see how extra $\zed_2$ charged objects arise in the spectrum of the type-I theories (see \cite{spingeom} for the mathematical details). \newsubsection{Duality Transformations} We shall now describe the various transformations among the dual theories described in the previous section. We will first discuss the duality between the type-I and type-I$\,^\prime$ theories. As in subsection 2.2 we may readily compute \begin{eqnarray} \widetilde{KO}(Y\times S^1)&=&\left(\widetilde{KO}^{-1}(Y)\oplus\zed_2\right) \oplus\widetilde{KO}(Y)\label{KOYS1}\\\widetilde{KR}^{-1} (Y\times S^{1,1})&=&\left(\widetilde{KO}(Y)\oplus\zed\right) \oplus\widetilde{KO}^{-1}(Y) \label{KRYS1}\end{eqnarray} where we have used \eqn{KRKOiso} and the fact that the orientifold group does not act on the non-compact space $Y$. Modulo the usual discrete groups of gauge transformations, the K-groups \eqn{KOYS1} and \eqn{KRYS1} of the type-I and type-I$\,^\prime$ theories are the same. The cyclic subgroup in \eqn{KOYS1} comes from \eqn{absisoKO}, while the integer subgroup in \eqn{KRYS1} comes from the identity \cite{atiyahreal} \begin{equation} \widetilde{KR}^{-n}(S^{p,q})= \widetilde{KR}^{q+1-n-p}(S^{1,0})=\widetilde{KO}^{q+1-n-p}(S^0) \label{KRSpq}\end{equation} It arises in exactly the same way as in the type-II case, i.e. from the tachyon field \eqn{TNI} which is the large gauge transformation of the $O(N)$ gauge bundle over the sphere $S^l$. The cyclic subgroup of \eqn{KOYS1} comes from the $\zed_2$ gauge transformation in the degenerate codimension 1 vacuum manifold, as explained above. Thus the duality transformations between the type-I and I$\,^\prime$ theories correctly account for the relevant changes of gauge field configurations living on the brane worldvolumes, just as in the type-II theories.\footnote{For an alternative construction of the $\zeds_2$ Wilson lines, see \cite{bgh}.} Again at the level of {\it unreduced} K-theory, the two K-groups coincide. By iterating \eqn{KOYS1} and \eqn{KRYS1}, we find that the same is true of the higher dimensional toroidal compactifications demonstrating the explicit mod $\zed$ equivalence of the type-I and type-I$\,^\prime$ models, or equivalently of their $T$-dual type-II orientifold theories, \begin{eqnarray} KO(Y\times T^m)&=&\bigoplus_{n=0}^mKO^{-n}(Y)^{\oplus{m\choose n}}\label{KOTm}\\KR^{-m}(Y\times T^{1,m})&=&\bigoplus_{n=0}^mKO^{n-m}(Y)^{\oplus{m\choose n}} \label{KRTm}\end{eqnarray} The decompositions of K-groups in (\ref{KOTm},\ref{KRTm}) contain the relevant degeneracies of wrapped branes around the cycles of $T^m$, which in the case of \eqn{KRTm} identifies the distribution of brane charges over the $2^m$ orientifold planes. As in subsection 2.2, each subgroup $KO^{n-m}(Y)^{\oplus{m\choose n}}$ is generated by a descendent tachyon field as one cycles through the periodicity maps in \eqn{botthomKO}. Upon writing the isomorphism between \eqn{KOTm} and \eqn{KRTm} in terms of reduced K-groups, we obtain winding numbers of the tachyon fields corresponding to large gauge transformations around the various cycles. In the type-II cases, these winding numbers indicated the precise degeneracy $2^{m-1}$ associated with the fact that the branes transformed under the spinor representation of the target space duality group. In the present case, we do not know the full duality group of a generic type-II orientifold, much less the representation that the brane charges carry. The decompositions into reduced groups in \eqn{KRTm} should be a clue as to what the appropriate group theoretic properties are of the target space dualities in this case. The results are summarized in table 1. \begin{table} \begin{center} \begin{tabular}{|c|c|c|l|} \hline \ $m$ \ & \ Dual theory\ & \ Vacuum manifold\ & $\widetilde{KR}^{-m}(Y\times T^{1,m})$\\ \hline\hline 0 & Type-I & $O(N)$ & $\widetilde{KO}(X)$\\ \hline 1 & Type-I$\,^\prime$ & $\frac{O(2N)}{O(N)\times O(N)}$ & $\widetilde{KO}(Y)\oplus\widetilde{KO}^{-1}(Y)\oplus\zed$\\ \hline & & & $\widetilde{KO}(Y)\oplus\widetilde{ KO}^{-1}(Y)^{\oplus 2}$\\\raisebox{1.5ex}[0pt]{2} & \raisebox{1.5ex}[0pt]{$F$-theory on $K3$} & \raisebox{1.5ex}[0pt]{$\frac{U(2N)}{O(2N)}$} & $\oplus\,\widetilde{KO}^{-2}(Y)\oplus\zed\oplus(\zed_2)^{\oplus 2}$\\ \hline{} & {} & {} & $\widetilde{KO}(Y)\oplus \widetilde{KO}^{-1}(Y)^{\oplus 3}$\\ 3 & $M$-theory on $K3$ & $\frac{Sp(2N)}{U(2N)}$ & $\oplus\, \widetilde{KO}^{-2}(Y)^{\oplus 3}\oplus \widetilde{KO}^{-3}(Y)$\\ {} & {} & {} & $\oplus\,\zed\oplus(\zed_2)^{\oplus 6}$\\ \hline {} & {} & {} & $\widetilde{KO}(Y)\oplus \widetilde{KO}^{-1}(Y)^{\oplus 4}$\\ 4 & IIA on $K3$ & $Sp(2N)$ & $\oplus\,\widetilde{KO}^{-2}(Y)^{\oplus 6}\oplus \widetilde{KO}^{-3}(Y)^{\oplus 4}$\\ {} & {} & {} & $\oplus\,\widetilde{KO}^{-4}(Y)\oplus\zed\oplus (\zed_2)^{\oplus 10}$\\ \hline{} & {} & {} & $\widetilde{KO}(Y)\oplus \widetilde{KO}^{-1}(Y)^{\oplus 5}$\\ & & & $\oplus\,\widetilde{KO}^{-2}(Y)^{\oplus 10}\oplus \widetilde{KO}^{-3}(Y)^{\oplus 10}$\\ \raisebox{1.5ex}[0pt]{5} & \raisebox{1.5ex}[0pt]{IIB on $K3\times S^1$} & \raisebox{1.5ex}[0pt]{$\frac{Sp(4N)}{Sp(2N)\times Sp(2N)}$} & $\oplus\,\widetilde{KO}^{-4}(Y)^{\oplus 5}\oplus \widetilde{KO}^{-5}(Y)$\\ {} & {} & {} & $\oplus\,\zed^{\oplus 6}\oplus(\zed_2)^{\oplus 15}$\\ \hline {} & {} & {} & $\widetilde{KO}(Y)\oplus \widetilde{KO}^{-1}(Y)^{\oplus 6}$\\ & & & $\oplus\,\widetilde{KO}^{-2}(Y)^{\oplus 15}\oplus \widetilde{KO}^{-3}(Y)^{\oplus 20}$\\ \raisebox{1.5ex}[0pt]{6} & \raisebox{1.5ex}[0pt]{$F$-theory on $T^8/\zed_2$} & \raisebox{1.5ex}[0pt]{$\frac{U(8N)}{Sp(4N)}$} & $\oplus\,\widetilde{KO}^{-4}(Y)^{\oplus 15}\oplus \widetilde{KO}^{-5}(Y)^{\oplus 6}$\\ {} & {} & {} & $\oplus\,\widetilde{KO}^{-6}(Y)\oplus\zed^{\oplus 16}\oplus (\zed_2)^{\oplus 21}$\\ \hline{} & {} & {}& $\widetilde{KO}(Y)\oplus \widetilde{KO}^{-1}(Y)^{\oplus 7}$\\ & & & $\oplus\, \widetilde{KO}^{-2}(Y)^{\oplus 21}\oplus \widetilde{KO}^{-3}(Y)^{\oplus35}$\\ 7 & $M$-theory on $T^8/\zed_2$ & $\frac{O(16N)}{U(8N)}$ & $\oplus\,\widetilde{KO}^{-4}(Y)^{\oplus 35}\oplus \widetilde{KO}^{-5}(Y)^{\oplus 21}$\\ {} & {} & {} & $\oplus\,\widetilde{KO}^{-6}(Y)^{\oplus 7}\oplus \widetilde{KO}^{-7}(Y)$\\{} & {} & {} & $\oplus\,\zed^{\oplus 36}\oplus(\zed_2)^{\oplus 28}$\\ \hline{} & {} & {} & $\widetilde{KO}(Y)\oplus \widetilde{KO}^{-1}(Y)^{\oplus 8}$\\ & & & $\oplus\,\widetilde{KO}^{-2}(Y)^{\oplus 28}\oplus \widetilde{KO}^{-3}(Y)^{\oplus 56}$\\ 8 & IIA on $T^8/\zed_2$ & $O(16N)$ & $\oplus\,\widetilde{KO}^{-4}(Y)^{\oplus 70}\oplus \widetilde{KO}^{-5}(Y)^{\oplus 56}$\\ {} & {} & {} & $\oplus\,\widetilde{KO}^{-6}(Y)^{\oplus 28}\oplus \widetilde{KO}^{-7}(Y)^{\oplus 8}$\\{} & {} & {} & $\oplus\,\widetilde{KO}^{-8}(Y)\oplus\zed^{\oplus 71}\oplus(\zed_2)^{\oplus 36}$\\ \hline \end{tabular} \end{center} \caption{\baselineskip=12pt {\it Type-II orientifolds on spacetimes $X=Y\times T^{1,m}$. The general dual orbifold model in each case is listed along with the corresponding vacuum manifold for tachyon condensation in the worldvolume of $2^{[\frac m2]+1}N$ spacetime filling 9-branes. The last column represents the distribution of brane charges over the various orientifold planes and their relevant multiplicities according to some representation of the target space duality group of the orientifold theory.}} \end{table} \subsubsection*{\it Orbifold Dualities} As discussed in the previous subsection, all of the type-II orientifold string theories are conjectured to be dual to more general conventional orbifold theories (see the second column of table 1). Although we cannot test these relations in general, we can make a heuristic analysis for the $m=4,5,8$ cases in table 1. The relevant K-groups for the type-IIA orbifolds on $T^m/\zed_2$ are given by the equivariant cohomology $K^{-1}_{\zeds_2}(Y\times T^{1,m})$. These groups can be computed as in subsection 2.3 using the six term exact sequence \eqn{sixterm}, with the results \begin{eqnarray} K^{-1}_{\zeds_2}(Y\times T^{1,4})&=& \left( K^{-1}(Y)^{\oplus 16} \otimes R[\zed_2]\right)\oplus K^{-1}(Y)\label{IIAT4orb}\\ K^{-1}_{\zeds_2}(Y\times T^{1,8})&=& \left( K^{-1}(Y)^{\oplus 256} \otimes R[\zed_2]\right)\oplus K^{-1}(Y) \label{IIAT8orb}\end{eqnarray} For the type-IIB orbifold on $S^1\times T^4/\zed_2$, we use in addition the decompositions (\ref{wKXciso},\ref{K1Xciso}) to get \begin{eqnarray} \widetilde{K}_{\zeds_2}(Y\times S^1\times T^{1,4})&=&\left(\widetilde{K}(Y)^{\oplus16}\oplus K^{-1}(Y)^{\oplus16}\oplus\zed^{\oplus15}\right)\otimes R[\zed_2]\nonumber\\& &\oplus\,K^{-1}(Y)\oplus\widetilde{K}(Y) \label{IIBT4orb}\end{eqnarray} Let us compare the $m=4$ line of table 1 with the complex K-group \eqn{IIAT4orb} which represents the spectrum of D-brane charges in the orbifold limit of the IIA compactification on $K3$. Both decompositions contain the multiplicities of brane charges localized on the 16 fixed point 5-planes of the given involution. The orientifold K-groups \eqn{KRTm} contain in addition the tachyon field winding numbers around cycles of $T^4$ as well as the appropriate windings around the various vacuum manifolds as dictated by the general homotopy analysis of subsection 3.1. On the other hand, the orbifold K-group \eqn{IIAT4orb} contains the mirror image brane charges on the orbifold planes along with the contribution from the unwrapped IIA brane configurations (represented by the second $K^{-1}(Y)$ direct summand). The natural map \eqn{KSPOmap} between real and complex K-groups shows how the IIA orbifold charges correspond to precisely the orientifold charges from a given tachyon configuration, along with the appropriate multiplicity of 2 as discussed at the end of the previous subsection. The remnant large gauge symmetry of the orientifold charges are then represented by unwrapped orbifold charges. This provides a new relationship between the given type-II orientifold theory and its dual. Of course, this heuristic comparison only holds at the level of the orbifold limit of the $K3$ moduli space of the IIA orbifold theory. A more precise analysis of this duality should make a comparison with the full $K3$ compactification. The group \eqn{IIAT4orb} can be seen to account for the usual BPS branes of the IIA orbifold theory \cite{sendescent,bg}. This spectrum contains fractional D-particles of unit charge with respect to the twisted $U(1)$ RR gauge fields at the orbifold planes, with the correct multiplicity of 4 arising from the possible bulk and twisted charges of the states at each plane giving a total of 64 such states. The spectrum also contains wrapped D2-branes around non-vanishing supersymmetric cycles, and D4-branes which wrap around the entire compact space. However, it is less clear how to identify the spectrum of stable non-BPS configurations directly from the decomposition \eqn{IIAT4orb}. For instance, the IIA theory on $T^4/\zed_2$ should contain a $\zed_2$-charged non-BPS 4-brane which comes from the D-particle in the dual type-I string theory. The spectrum of $\zed_2$-charges in general can only be identified under the natural homomorphism between \eqn{IIAT4orb} and table 1, so that we can take the latter groups to represent the full brane spectrum of the orbifold theory. The problem can be traced back to the fact that the duality map in the present case involves an intermediate $S$-duality transformation \cite{bg}, whose description in the K-theoretic formalism is at present not known \cite{witten,gukov}. Other non-BPS states in the spectrum of the present model include D-particles stuck at the fixed point planes, and D-strings stretched between pairs of orbifold fixed points with the same magnitude of charge as those of the fractional BPS D-particles \cite{bg}. A similar comparison holds between the last line of table 1 and the complex K-group \eqn{IIAT8orb} which represents the type-IIA orbifold compactification on $T^8/\zed_2$. The duality between the $m=5$ line of table 1 and \eqn{IIBT4orb}, which represents the orbifold limit of the IIB compactification on $K3\times S^1$, is more involved because the latter group contains extra unwrapped and wrapped brane charges. The 32 O$4^-$ planes of the IIB orientifold have brane charges distributed according to table 1. On the other hand, the brane charges localized at the 16 fixed point 5-planes of the IIB orbifold (along with the mirror images) split evenly into two sets of type-II charges corresponding to wrapped and unwrapped configurations around the extra $S^1$. Now the mapping between the two K-groups according to \eqn{KSPOmap} matches the symmetrical splitting of the KO-group decomposition for $m=5$, and one may take the K-groups of table 1 to represent the non-BPS configurations of these dual models. It would be interesting to test this matching by explicit string theoretical constructions. \newsubsection{$(-1)^{F_{\rm L}}$ Transformations} The generalization of the action of the Klein operator $\klein$ on the type-I theories and on the type-IIB orientifolds can be deduced from their relationships with the type-II theories. Let us start from the type-I theory, regarded as the quotient of type-IIB superstring theory by worldsheet parity $\Omega$. Regarding the type-IIA theory as the quotient of IIB by $\klein$, the type-I$\,^\prime$ theory may then be obtained as the quotient of IIB by the involution $\Omega\cdot\klein\cdot{\cal I}_1$. The Grothendieck group $KR_\pm(X)$ of virtual complex vector bundles with such an involution is a generalization of the Hopkins K-groups to the category of Real vector bundles over $X$ \cite{witten,gukov}. Thus the action on K-theory by the $\klein$ projection on the type-I theory should induce a map \begin{equation} \widetilde{KO}(X)~\longrightarrow~\widetilde{KR}_\pm(X) \label{KOKRpmmap}\end{equation} The problem we encounter at this stage is a mathematical one. The theory of Hopkins groups $K_\pm(X)$ has not been investigated much in the literature, much less its Real generalization. In particular, a product formula such as that given by the right-hand side of \eqn{Kequivmap} is not known in this case. In \cite{gukov} it was suggested that an analog of the Hopkins formula could be \begin{equation} KR_\pm(X)=KR_{\zeds_2}(X\times\real^{1,1}) \label{KRpmprod}\end{equation} where the cyclic group acts as the product of the action of ${\cal I}_1$ on $X$ and an orientation reversing symmetry of the real space $\real^{1,1}$. This relationship exemplifies the fact that duality transformations and orbifold operations do not always commute \cite{senorb}, in this case within the various interrelationships between the type-I and type-II theories. Instead of the mapping \eqn{KOKRpmmap}, there is a more natural candidate which comes about in analogy with the $T$-duality transformations of the type-I theories. One could consider the action of $\klein$ directly on the type-I$\,^\prime$ theory, thereby obtaining the map \eqn{Kequivmap} into Real virtual bundles. The relevant K-group for the operation of modding out the type-I$\,^\prime$ theory $m$ times by $\klein$ is then \begin{equation} \widetilde{KR}^{-m}_{\zeds_2}(X\times\real^{0,m})= \left(\widetilde{KR}^{-m}(X)\otimes R[\zed_2]\right)\oplus\widetilde{KR}(X) \label{KRkleinm}\end{equation} with a trivial $\zed_2$ action on the spacetime $X$. The decomposition \eqn{KRkleinm} follows using the methods of subsection 2.3 applied to the KR-groups and the suspension isomorphism \cite{spingeom} \begin{equation} KR^{-n}(X\times\real^{p,q})=KR^{q-p-n}(X) \label{KRsusp}\end{equation} for Real K-theory. The projection onto the first direct summand in \eqn{KRkleinm}, representing the states which survive the $\klein$ projections, can be carried out explicitly on the invariant brane-antibrane configurations as in subsection 2.3. The other summand $\widetilde{KR}(X)$ then always represents the orientifold states projected out by the $\klein$ involution. The subsequent $\klein$ projections now take us through the entire sequence of type-II orientifolds described in subsection 3.1. It would be interesting to test the mappings described here directly using explicit string theoretical constructions (for example, a boundary state calculation along the lines of \cite{daspark}). \newsubsection{Summary} Again we may succinctly summarize the descent relations among type-I branes and those of type-II orientifolds by a diagram representing the various mappings on K-groups: \begin{equation}\oldfalse{\begin{array}{ccllllc} KO(X)&{\buildrel\beta\over\longrightarrow}&KO^{-1}(X)&{\buildrel\beta \over\longrightarrow}&\!\!\!\!\!\!\!\!\cdots&\!\!\!\!\!\!\!\! {\buildrel\beta\over\longrightarrow}&\!\!\!\!\!\!\!\! KO^{-8}(X)\\& & &\!\!\!\!\!\!\!\!&\!\!\!\!\!\!\!\!&\!\!\!\!\!\!\!\!&\!\! \!\!\!\!\!\!\parallel\\{\scriptstyle\klein}\downarrow&\searrow&\!\!\!\!\!\!\! {\scriptstyle KO(Y\times S^1)}&\!\!\!\!\!\!\!\!&\!\!\!\!\!\!\!\!& \!\!\!\!\!\!\!\!&\!\!\!\!\!\!\!\!KO(X)\\& &{\scriptstyle=KR^{-1}(Y\times S^{1,1})}& & & &\\& & & & & &\\KR_{\zeds _2}^{-1}(X\times\real^{0,1})&{\buildrel{\beta\circ\Pi_1}\over\longrightarrow} &KR^{-1}(X)& & & &\\& & & & & &\\& &{\scriptstyle KO(Y\times T^2)}& & & &\\ {\scriptstyle\klein}\downarrow& &{\scriptstyle=KR^{-2}(Y\times T^{1,2})} &\!\!\!\!\!\!\!\searrow& & &\\& & & & & &\\KR_{\zeds_2}^{-2}(X\times\real^{0,2})& &{\buildrel {\beta\circ\Pi_1}\over\longrightarrow}&KR^{-2}(X)& & &\\& & & & & &\\& & &{\scriptstyle KO(Y\times T^3)}& & &\\{\scriptstyle\klein}\downarrow& & &{\scriptstyle=KR^{-3}(Y\times T^{1,3})}&\!\!\!\! \!\!\!\searrow& &\\& & & & & &\\\vdots& & & &\ddots& &\\& & & & {\scriptstyle KO(Y\times T^8)}& &\\{\scriptstyle\klein}\downarrow& & & &{\scriptstyle=KR(Y\times T^{1,8})}&\!\!\!\!\!\!\!\searrow& \\& & & & & & \\KR_{\zeds_2}^{-8}(X\times\real^{0,8})& & &{\buildrel\Pi_1\over\longrightarrow}& &KR(X)&\end{array}} \label{KOKRdiag}\end{equation} Again $\beta$ is the Bott periodicity isomorphism representing tachyon condensation in codimension 1, except that now in general it leads to a stable brane configuration (non-trivial K-theory class). In particular, $\beta$ realizes a type-I non-BPS $p-1$-brane as a kink in the worldvolume of a $p$-$\overline{p}$ pair. The map $\beta^4$ acts on KO-groups according to \begin{equation} \beta^4\,KO(X)=KSp(X) \label{beta4KOX}\end{equation} and it realizes a BPS $p-4$-brane as an $SU(2)$ Yang-Mills instanton in the worldvolume of four $p$-$\overline{p}$ pairs. The map $\beta^8$ realizes a $p-8$-brane as a $spin(8)$ instanton in the worldvolume of 16 $p$-$\overline{p}$ pairs. The projections $\Pi_1$ are as before. Note that the statements about KR-groups above can be translated into ones about KO-groups upon taking the KR-involution to act trivially on the spacetime $X$. \newsection{Orientifold Symmetries} One subtlety in our description of type-II orientifolds in the previous section is that the homotopy classification of KR-theory requires a slightly refined definition. This is in turn related to a more general periodicity property of Real K-theory. In this final section we will describe these general symmetries of the orientifold models in some detail, thereby exposing the rich internal symmetries predicted by the K-theory formalism. \newsubsection{Periodicity in Real K-theory} The definition of the higher KR-groups that we have used thus far has been done with respect to suspensions whereby the KR-involution acts trivially on the sphere. There is a more general class of Real K-groups which are defined by the double index suspensions \begin{equation} KR^{p,q}(X)=KR(\Sigma^{p,q}X)~~~~~~{\rm with}~~\Sigma^{p,q}X=X\wedge\real_+^{p,q} \label{KRpqdef}\end{equation} With this definition we have \begin{equation} KR^{-n}(X)=KR^{n,0}(X) \label{KRdiff}\end{equation} Taking the cup product with the class of the complex Hopf line bundle ${\cal N}_\complexs$ over $\complex P^1$ with its natural Real structure (given by the anti-linear complex conjugation involution) induces the strong Bott periodicity isomorphism \cite{atiyahreal} \begin{equation} KR^{p+1,q+1}(X)=KR^{p,q}(X) \label{KRbottper}\end{equation} showing that in fact $KR^{p,q}(X)=KR^{q-p}(X)$. Identifying $\real^{1,1}$ as the space $\complex$ with complex conjugation, this $(1,1)$ periodicity may be cast in the form of a suspension isomorphism \begin{equation} KR(X)=KR(X\times\complex) \label{KRsuspiso}\end{equation} Thus the K-theory of the orientifold models described in the previous section is naturally contained within this two-index set of KR-groups. To understand what the periodicity \eqn{KRbottper} means physically in terms of brane charges, we appeal to the ABS construction for Real K-theory. For this, we define a two-parameter set of Clifford algebras $C_{n,m}$ of the real space $\real^{n,m}$ as the usual algebra associated with $\real^{n+m}$ together with an involution generated by the ${\cal I}_m$ involution acting on $\real^{n,m}$. A Real module over $C_{n,m}$ is then a finite-dimensional representation together with a $\complex$-antilinear involution which preserves the Clifford multiplication. The corresponding representation ring $R[spin(n,m)]$ is naturally isomorphic to the Grothendieck group generated by the irreducible $\real$-modules $\Delta_{n,m}$ of the Clifford algebra of the space $\real^n\oplus\real^m$ with quadratic form of Lorentzian signature $(n,m)$. The ABS map is now the graded ring isomorphism \cite{spingeom,atiyahreal} \begin{equation} KR(\real^{n,m})~\cong~R\Bigl[spin(n,m)\Bigr]\,/\,R\Bigl[spin(n+1,m)\Bigr]~=~KO^{n-m}(S^0) \label{absKR}\end{equation} This isomorphism relates the groups on the left-hand side of \eqn{absKR} to the Clifford algebras $C_{n,m}$, so that the $(1,1)$ periodicity \eqn{KRbottper} is reflected in the $(1,1)$ periodicity of the Clifford algebras, $C_{n+1,m+1}=C_{n,m}\otimes\real(2)$ \cite{spingeom}. Let us now consider a $p$-brane of codimension $l=n+m$ in a type-II orientifold by $\Omega\cdot{\cal I}_m$. The $p$-brane charge is induced by the tachyon field which is given by Clifford multiplication on the transverse space $\real^{n,m}$, i.e. $T(x)=\sum_i\Gamma_i\,x^i$ where $\Gamma_i$ are the generators of the spinor module $\Delta_{n,m}$, and which generates $\widetilde{KR}(\real^{n,m})$. Under the ABS isomorphism above, this KR-theory class is multiplied, via the cup product, by the Hopf generator of $\widetilde{KR}(\complex P^1)=\zed$, or equivalently by the spin bundles which carry the spinor representation $\Delta_{1,1}$. This gives a class with tachyon field that generates the KR-group of the new transverse space $\real^{n+1,m+1}$. This class represents a $p-2$-brane of the type-II orientifold by $\Omega\cdot{\cal I}_{m+1}$. From this mathematical fact we may deduce a new descent relation for type-II orientifold theories. A $p-2$-brane localized at an O$(8-m)^-$-plane in a type-II $\Omega\cdot{\cal I}_{m+1}$ orientifold may be constructed as the tachyonic soliton of a bound state of a $p$-$\overline{p}$ pair located on top of an O$(9-m)^-$-plane in a type-II $\Omega\cdot{\cal I}_m$ orientifold. This realizes the branes of a type-II orientifold as equivariant magnetic monopoles in the worldvolumes of brane-antibrane pairs of an orientifold with fixed point planes of one higher dimension. The former orientifold has $2^{m+1}$ O$(8-m)^-$-planes each carrying RR-charge $-2^{3-m}$, while the latter one has $2^m$ O$(9-m)^-$-planes of charge $-2^{4-m}$. In a sense, in this process of tachyon condensation the number of fixed point planes is doubled while their charges are lowered by a factor of 2 by a combined operation of charge transfer and dimensional reduction through the orientifold planes. An example is the non-BPS state consisting of a D5-brane on top of an orientifold 5-plane in the type-IIB theory \cite{senbps}, which may be constructed via a tachyon condensate from a pair of D7-$\overline{\rm D7}$ branes on an orientifold 6-plane in the type-IIA theory. The 8 O$6^-$-planes each carrying charge $-2$ are transfered to the 16 O$5^-$-planes of charge $-1$. This new sort of internal symmetry among the type-II orientifolds may be thought of as a type of $T$-duality symmetry acting on the RR-charges of the orientifold planes, in analogy with the transformations described in subsection 3.1. By repeated iteration as before, we obtain an entire hierarchy of novel bound state constructions of D-branes in various higher dimensions. To describe the field content, however, we must be careful about identifying the appropriate homotopy of the relevant vacuum manifolds. The classifying spaces for Real vector bundles are described in \cite{krclass}. Consider an orientifold of the type-IIB theory, and a set of brane-antibrane pairs with worldvolume gauge symmetry $U(N)\times U(N)$. The $U(N)$ gauge group is endowed with its complex conjugation involution whose fixed point set is the real subgroup $O(N)$. The tachyon field $T$ is equivariant with respect to the orientifold group, so that \begin{equation} T(x,-y)=T(x,y)^* \label{tachyonKR}\end{equation} where $(x,y)\in\real^n\oplus\real^m$. It breaks the worldvolume gauge symmetry down to $U(N)_{\rm diag}$. The relevant homotopy group generated by \eqn{tachyonKR} comes from decomposing the one-point compactification of $\real^{n,m}$ into upper and lower hemispheres as described in subsection 2.2, such that the tachyon field is the transition function on the overlap. The D-brane charges thereby reside in the KR-group of the transverse space which is given by \begin{equation} \widetilde{KR}(\real^{n,m})=\pi_{n,m}\Bigl(U(N)\Bigr)_R \label{KRhomotopy}\end{equation} where the equivariant homotopy group is defined by the maps $S^{n,m}\to U(N)$ which obey the Real equivariance condition \eqn{tachyonKR}. The refined weak Bott periodicity theorem for stable equivariant homotopy in KR-theory then reads \begin{equation} \pi_{n,m}\Bigl(U(N)\Bigr)_R=\pi_{n+1,m+1}\Bigl(U(2N)\Bigr)_R \label{KRbotthom}\end{equation} In a similar way one may relate the Real K-groups $\widetilde{KR}^{-1}(\real^{n,m})=\widetilde{KR}(\real^{n+1,m})$ to the stable equivariant homotopy of the complex Grassmanian manifold $U(2N)/[U(N)\times U(N)]$. In this way we arrive at the Real version of the homotopy sequence \eqn{botthomotopy}. The first isomorphism from a type-IIB orientifold to a type-IIA orientifold preserves the structure of the orientifold planes, while the second step of the sequence decreases the fixed point plane dimension by 1. This gives a novel generalization of the $T$-duality and descent relations between IIA and IIB orientifolds, which would be very interesting to reproduce using string theoretic arguments, as in \cite{senbps}. Note that the relevance of the Dirac monopole in the orientifolds is not surprising, given its prominent role in the type-II theories of section 2. Having identified these orientifold symmetries, we may now restrict our attention to D-brane charges living in the groups \eqn{KRdiff}, and hence proceed to the real Bott periodicity relations, as in subsection 3.1. The sequence of homotopy groups in \eqn{botthomKO} now arises from the fact \cite{krclass} that if the KR-involution acts freely, then the classifying space for stable equivariant homotopy reduces to the coset space $U(2N)/O(2N)$. The rest of the isomorphisms now follow as before. Note that the gauge fields living on the brane worldvolumes in these cases must also satisfy an equivariance condition like \eqn{tachyonKR}. A similar analysis can be done for the $(1,1)$ peridicity of KH-theory \cite{kh} which gives internal relations among branes localized on O$(9-m)^+$-planes. \newsubsection{Internal Symmetries} The Real K-groups are in a certain sense ``universal'' as they contain all of the other generalized cohomology theories. This feature follows from the many further internal symmetries present in KR-theory. In particular, there are the natural periodicity isomorphisms \cite{atiyahreal} \begin{equation} KR(X\times S^{0,m})=KR^{-2m}(X\times S^{0,m}) \label{KRSper}\end{equation} for $m=1,2,4$. The periodicity of KR-theory with coefficients in $S^{0,m}$ follows by using the multiplication in the fields $\real,\complex,\quater$, respectively, and $(1,1)$ periodicity. Note that for $m=1$ we have the isomorphism \eqn{KRKiso}, so that \eqn{KRSper} then reduces to the complex Bott periodicity theorem \eqn{bottKn}, while the $m=4$ case leads immediately to the real periodicity theorem \eqn{bottRealKn}. In fact, there is the usual decomposition \cite{atiyahreal} \begin{equation} KR^{-n}(X\times S^{0,m})=KR^{-n}(X)\oplus KR^{-n+m+1}(X)~~~~~~\forall m\geq3 \label{KRm3split}\end{equation} The case $m=2$ is special and the corresponding KR-groups are isomorphic to the Grothendieck groups generated by the category of self-conjugate vector bundles over $X$ \cite{atiyahreal}. This self-conjugate K-theory has Bott periodicity 4, which from \eqn{bottrealKn} we see is associated with symplectic Chan-Paton gauge bundles. This fact has been exploited in \cite{bgh} to relate this type of K-theory to certain orientifold compactifications of type-II theories without vector structure \cite{wittenvec}, i.e. those with the same number of both O$(9-m)^-$ and O$(9-m)^+$ planes of cancelling charge that require no D-branes in the vacuum configuration, and hence have no gauge group. The $T$-duality transformations in these theories is also explained in \cite{bgh}. These transformations, as well as the brane descent relations in these models, are now natural consequences of the periodicity and homotopy properties of the KR-groups explained in the previous subsection. It would be interesting to construct these transformations more explicitly, thus verifying the results of \cite{bgh}. It would also be interesting to see if the various duality properties of orientifold compactifications without vector structure \cite{wittenvec}, which differ somewhat from those described in section 3, can be deduced from the internal symmetries described in this section. Thus the internal symmetries of KR-theory encompass most of the brane descent relations which are based on periodicity theorems, and they further emphasize the role played by Hopf fiber bundles in the topological classification of D-branes. In addition, the further symmetry relations in orbifold models may be thought of as originating within the equivariant structure of KR-theory, and from the usual Bott periodicities of the given equivariant K-groups, with solitonic configurations provided by equivariant versions of the four canonical solitons coming from the Hopf fibrations. It would also be interesting to study more closely the internal symmetries of type-II orientifolds arising from quotients by the operator $\klein\cdot{\cal I}_m$, which induce the Hopkins K-groups $K_\pm(X)$. For instance, as discussed in \cite{gukov}, the products with the Thom spaces of $\complex$ and $\complex/\zed_2$ induce, respectively, the periodicity isomorphisms which identify the Hopkins K-groups of $\real^{n,m}$ with $\real^{n+2,m}$, and of $\real^{n,m}$ with $\real^{n,m+2}$. The first periodicity builds a $p-2$-brane from a $p$-$\overline{p}$ brane pair through the usual monopole and leaves the orientifold plane structure unchanged, while the second one realizes a $p-2$-brane inside a $p$-$\overline{p}$ pair through a $\zed_2$-equivariant monopole and lowers the dimension of the orientifold planes by 2. This phenomenon is similar to that described above for the $\Omega\cdot{\cal I}_m$ orientifolds, and the monopole symmetries can be seen to naturally arise from the definition of the Hopkins K-functor as the usual equivariant K-functor (see the right-hand side of \eqn{Kequivmap}). Again it would be most interesting to carry out these constructions from a more physical standpoint. \bigskip \noindent {\bf Acknowledgements:} We thank P. Di Vecchia, F. Lizzi, J. M\o ller, R. Nest, N. Obers, S. Schwede, G. Semenoff and G. Sparano for helpful discussions, and P. Ho\v rava and G. Landi for comments on the manuscript. \newpage
\section{#1}\smallskip} \newcommand{\subsect}[1]{\medskip\subsection{#1}\smallskip} \newcommand{\subsubsect}[1]{\medskip\subsubsection{#1}\smallskip} \renewcommand{\theequation}{\thesection.\arabic{equation}} \def\begin{eqnarray}{\begin{eqnarray}} \def\end{eqnarray}{\end{eqnarray}} \def\begin{eqnarray*}{\begin{eqnarray*}} \def\end{eqnarray*}{\end{eqnarray*}} \def\end{eqnarray}{\end{eqnarray}} \def&\!\!\!\!\!=\!\!\!\!\!&{&\!\!\!\!\!=\!\!\!\!\!&} \def&\!\!\!\!\!\!\!\!\!\!&{&\!\!\!\!\!\!\!\!\!\!&} \def\vartheta{\vartheta} \def\tilde{\vartheta}{\tilde{\vartheta}} \def\varphi{\varphi} \def\mathcal{L}{{\cal{L}}} \def\kappa{\kappa} \def\alpha{\alpha} \def\beta{\beta} \def\gamma{\gamma} \def\delta{\delta} \def\gamma{\gamma} \def\mu{\mu} \def\nu{\nu} \def\lambda{\lambda} \def\mathcal{L}{\mathcal{L}} \def\nabla^2{\nabla^2} \newcommand{\bi}[1]{\bibitem{#1}} \title{{\bf On variations in teleparallelism theories}} \author{ \thanks {\quad email <EMAIL>} Yakov Itin \\ \small {Institute of Mathematics}\\ \small {Hebrew University of Jerusalem}\\ \small {Givat Ram, Jerusalem 91904, Israel}\\ } \begin{document} \maketitle \begin{abstract} \sf The variation procedure on a teleparallel manifold is studied. The main problem is the non-commutativity of the variation with the Hodge dual map. We establish certain useful formulas for variations and restate the master formula due to Hehl and his collaborates. Our approach is different and sometimes easier for applications. By introducing the technique of the variational matrix we find necessary and sufficient conditions for commutativity (anti-commutativity) of the variation derivative with the Hodge dual operator. A general formula for the variation of the quadratic-type expression is obtained. The described variational technique are used in two viable field theories: the electro-magnetic Lagrangian on a curved manifold and the Rumpf Lagrangian of the translation invariant gravity. \end{abstract} \sf \sect{Introduction} One of the biggest challenges of the theoretical physics is to unify the standard model of particles interactions with Einstein's theory of gravity. Both these theories are in a good according with all known observable phenomena. But their fundamental concept are rather differ. Standard model is a quantum field theory on a flat Minkowskian space-time. In contrast to that the Einstein's description of gravity is a classical field theory essential connected with the geometrical properties of the pseudo-Riemannian manifold. One can expect the combination of these different theories only after principal modification of every one of them. Great attempts are making recently in modernization the standard model in accordance with the idea of superstrings. The resent description of the situation in this area can be found in \cite{S-S}. The second partner of the future couple should also make some necessary preparations. The Einsteinian theory of general relativity (GR) is almost a unique theory of gravity that can be formulated in the framework of the pseudo-Riemannian manifold. Thus an essential modification can be made only on a basis of some generalized geometrical construction. The study of various geometries relevant to gravity began at once after Einstein had proposed his general relativity. The most general geometrical theory of gravity is the metric-affine theory \cite{hehl95}. In the present paper we study a rather simpler generalization of the pseudo-Riemann geometry - the theories of teleparallelism. The teleparallel space was introduced for the first time by Cartan \cite{Ca} and used by Einstein \cite{E1}, \cite{EM} in a certain variant of his unified theory of gravity and electromagnetism. The work of Weitzenb\"{o}ck \cite{We1} was the first denoted to the investigation of the geometric structure of teleparallel spaces. The theories based on this geometrical structure appear in physics time to time in order to give an alternative model of gravity or to describe the spin properties of matter. For the resent investigations in this area see Ref. \cite{Kop}, \cite{M-H}, \cite {Maluf2}, \cite {Maluf3}, \cite {T-N}, \cite {N-Y}.\\ It is convenient and useful to have a Lagrangian formulation of a field theory. A designated functional, denoted by action, is the basis for the Lagrangian formalism. The field equations of the classical field theory are the representation of the critical point condition for the prescribed action functional.\\ Let us take a brief look on the main properties of the action functionals in the classical relativistic field theory. \begin{itemize} \item The action functional is an integral taken on the whole 4-dimensional differential manifold. \item The integrand is a certain differential 4-form - \textsl{ Lagrangian density}\sf. \item The Lagrangian density is a scalar invariant under the group of transformations of the considered system. \item The Lagrangian density incorporates only the squares of the first-order derivatives of the field variables in the same point - local densities. \end{itemize} We consider the last condition as a necessary one, because almost all the physically meaningful field equations are second-order partial differential equations of hyperbolic type. The applying of the variational principle on the teleparallel framework is connected with various problems, as it emphasized in \cite{Hehl}. The main source of this problems connects with the fact that the operators of a certain type, namely the Hodge dual operator, the interior product and the coderivative operator, non-commute with the variational derivative. In the present paper we study the the variation procedure (free and constrained) on the teleparallel manifold. The overview of the work is as following. We begin with a brief overview of the variational procedure in the general relativity. In the second section we describe the preferences that bring the consideration of a teleparallel space instead of a pseudo-Riemannian manifold. The most advantage of such generalization is a possibility to describe the gravity by the quadratic Lagrangians, similar to the other field theories. Variational procedure on the teleparallel spaces is described in the third section. We prove the commutative relation which coincides in the case of pseudo-orthonormal coframe with the master formula of Muench, Gronwald and Hehl \cite{Hehl}. The analogous relation described also in \cite{BFK}. Using the commutative relation we derive the formula for variation of the general quadratic Lagrangian. The fourth section is devoted for describing the various types of the covariant constrains on a teleparallel structure. We study the relation of such constrains with the commutativity and anti-commutativity of the variational derivative with the Hodge dual map. The last two sections are devoted to the application of our variational formula for viable field theories. We consider the Maxwell Lagrangian on a teleparallel space and the Rumpf Lagrangian for a translation invariant gravity. \sect{Actions on a teleparallel manifold.} The teleparallel space was considered originally as a 4D-space endowed with a smooth field of frame (an ordered set of 4 independent vectors). Thus one can give a sense to parallelism of two vectors taken in different points of the space. In order to furnish the action with a differential 4-form and to apply the technique of the differential forms we consider a coframe field instead of the frame field. Thus \begin{definition} A teleparallel manifold is a pair $(M,\Gamma)$, where $M$ is a differential 4D-manifold and $\Gamma=\{\vartheta^a(x^\mu\}, \ a=0,1,2,3$ is a fixed smooth cross-section of the coframe bundle $FM$. \vrule height7pt width4pt depth1pt\end{definition} The field variable is a 4-tuple of 1-forms $\vartheta^a$ and they constitute a basis of the covector space $T^*_xM$ at every point $x$ of $M$. We endow this vector space with the Lorentzian metric $\eta^{ab}=\eta_{ab}=diag(-1,1,1,1)$ and require the 1-forms $\vartheta^a$ to be pseudo-orthonormal relative to this metric. Let $C^\infty$ be a set of smooth real valued functions on $M$. Denote by $\Omega^p$ the $C^\infty$-modulo of differential $p$-forms on $M$.\\ The following algebraic operations are defined on a $n$-dimensional teleparallel manifold: \begin{itemize} \item[\bf{1}] The exterior (wedge) product $\wedge:\Omega^p\times\Omega^q\to\Omega^{p+q}$ of two differential forms $\alpha\in\Omega^p$ and $\beta\in\Omega^q$, which is associative, $C^\infty$-bilinear and in general not commutative \begin{equation} \alpha\wedge\beta=(-1)^{pq}\beta\wedge\alpha. \label{2-1} \end{equation} \item[\bf{2}] The Hodge dual map $*:\Omega^p\to\Omega^{n-p}$. This is a $C^\infty$-linear map $*:\Omega^p\to\Omega^{n-p}$, which acts on the wedge product monomials of the basis 1-forms as \footnote{We use the abbreviated notations for the wedge product monomials: $\vartheta^{ab\cdots}:=\vartheta^a\wedge\vartheta^b\wedge\cdots$} \begin{equation} *(\vartheta^{a_1\cdots a_p})=\epsilon^{a_1\cdots a_n}\vartheta_{a_{p+1}\cdots{a_n}}, \label{2-2} \end{equation} where $\vartheta_a$ is the down indexed 1-forms $\vartheta_a=\eta_{ab}\vartheta^b$ and $\epsilon^{a_1...a_n}$ is the total antisymmetric pseudo-tensor. \item[\bf{3}] The interior product $v\hook \alpha$ of an arbitrary differential form $\alpha\in \Omega^p $ with an arbitrary vector $v\in T_xM$. This is a $C^\infty$-bilinear map $\hook:\Omega^p\to\Omega^{p-1}$, which satisfies the following properties: \begin{equation} v\hook (\alpha\wedge\beta)=(v\hook \alpha)\wedge\beta+(-1)^{deg\alpha}\alpha\wedge(v\hook \beta), \label{2-3} \end{equation} \begin{equation} e_a\hook \vartheta^b=\delta^b_a, \label{2-4} \end{equation} where $e_a$ is a basis vector of $T_xM$ and $\vartheta^a$ is a basis 1-form of $T^*_xM$. \end{itemize} Define a product for an arbitrary 1-form $w$ and an arbitrary $p$-form as \begin{equation} w\hook \alpha:=*(w\wedge*\alpha). \label{2-5} \end{equation} It is easy to see that this operation is a $C^\infty$-bilinear map $\Omega^p\to\Omega^{p-1}$, which satisfies the properties (\ref{2-3}, \ \ref{2-4}). In particular, we have \begin{equation} \vartheta_a\hook \vartheta^b:=*(\vartheta_a \wedge *\vartheta^b)=\delta^b_a. \label{2-6} \end{equation} Note that the product defined by (\ref{2-5}) is really a multiplication of two exterior forms thus this type of definition justifies the term ``interior product''. We use in this paper the definition of the interior product in the form (\ref{2-5}).\\ The following first-order differential operators are defined on a teleparallel manifold: \begin{itemize} \item[\bf{1}] The exterior derivative operator $d:\Omega^p\to\Omega^{p+1}$ \begin{equation} d\alpha=d\Big(\alpha_{a_1\cdots a_p}dx^{a_1}\wedge\cdots\wedge dx^{a_p}\Big)= d\alpha_{a_1\cdots a_p}\wedge dx^{a_1}\wedge\cdots\wedge dx^{a_p}. \label{2-7} \end{equation} This is an anti-derivative relative to the wedge product of forms \begin{equation} d(\alpha\wedge\beta)=d\alpha\wedge\beta+(-1)^{deg\alpha}\alpha\wedge d\beta. \label{2-7a} \end{equation} \item[\bf{2}] The coderivative operator $d:\Omega^p\to\Omega^{p-1}$ defined by \begin{equation} d^+\alpha=*d*\alpha. \label{2-8} \end{equation} \end{itemize} Using the operators described above the action functional can be accepted on the teleparallel manifold in the quadratic (Dirichlet) form. The simplest choice of the Lagrangian density can be made in the form of the Yang-Mills Lagrangian: \begin{equation} S[\vartheta^a] = \int_M d\vartheta^a \wedge *d\vartheta_a. \label{2-9} \end{equation} Observe that the action (\ref{2-9}) satisfies the following conditions: \begin{itemize} \item[\bf{1}] It is independent on a particular choice of a local coordinate system. \item[\bf{2}] It includes only the first order derivatives of the field variables (coframe 1-forms) so the corresponding field equation is at most of the second order. \item[\bf{3}] It is a functional of a Dirichlet-type so it provides an equation of a harmonic type. \item[\bf{4}] It is invariant under the global group of $SO(1,3)$ transformations of the coframe field $\vartheta^a$. \end{itemize} Therefore the generalization of the pseudo-Riemannian structure to a teleparallel structure allows to consider general Lagrangians for gravity similar to the quadratic Lagrangians of the other (already quantized) field theories. In this way one can hope to define the energy of gravity field in a local, covariant form. \sect{Variational procedure} The variational procedure on a teleparallel manifold for the action functional $S=S[\vartheta^a]$ can be described as follows. \\ Let $\vartheta^a(\lambda)$ be a smooth 1-parametric family of cross-sections of $FM$, with the following initial conditions \begin{equation} \vartheta^a(\lambda=0)= \vartheta^a\hbox{\qquad and\qquad}{\frac{\partial\vartheta^a}{\partial\lambda}}\bigg|_{\lambda=0}=\delta\vartheta^a. \label{3-1}\end{equation} The critical points of the functional $S[\vartheta^a]$ are defined by the condition: \begin{equation} \delta S={\frac{dS}{d\lambda}} \bigg |_{\lambda=0} d\lambda=0. \label{3-2}\end{equation} The Hamilton principle of the least action postulates that the field equation of the physical system coincides with the critical point condition for an appropriative action functional. \\ The variation $\delta$ ($\lambda$-differential) is independent on the space-time coordinates so it satisfies the following rules. \begin{itemize} \item[\bf{1}]The ordinary Leibniz rule for the wedge product \begin{equation} \delta(\alpha \wedge \beta)=\delta\alpha \wedge \beta+\alpha \wedge \delta\beta. \label{3-3}\end{equation} \item[\bf{2}] The commutativity with the exterior derivative \begin{equation} \delta \circ d=d \circ \delta \label{3-4}\end{equation} \item[\bf{3}] The non-commutativity (in general) with the Hodge star-operator \begin{equation} \delta \circ *\ne * \circ \delta \label{3-5}\end{equation} \item [\bf{4}] The non-commutativity (in general) with the coderivative operator \begin{equation} \delta \circ d^+\ne d^+ \circ \delta \label{3-6}\end{equation} \end{itemize} The following lemma is useful for the actual calculations of the variations. \begin{lemma} A variation of the wedge product monomials satisfies \begin{equation} \delta\vartheta^{a_1...a_p} = \delta\vartheta^m \wedge (\vartheta_m \hook \vartheta^{a_1...a_p}), \label{3-7}\end{equation} and for the Hodge dual monomials \begin{equation} \delta*\vartheta^{a_1...a_p} = \delta\vartheta^m \wedge (\vartheta_m \hook *\vartheta^{a_1...a_p}). \label{3-8}\end{equation} \end{lemma} \noindent{\bf Proof:} Using the Leibniz rule (\ref{3-3}) we obtain \begin{equation} \delta\vartheta^{a_1...a_p} = \delta\vartheta^{a_1} \wedge \vartheta^{a_2...a_p}+\vartheta^a_1 \wedge \delta\vartheta^{a_2} \wedge \vartheta^{a_3...a_p}+...+\vartheta^{a_1...a_{p-1}} \wedge \delta\vartheta^{a_p}. \label{3-9}\end{equation} The property of the interior product yields \begin{equation} \vartheta_m\hook\vartheta^{a_1\cdots a_p}=\delta^{a_1}_m\vartheta^{a_2\cdots a_p}-\delta^{a_2}_m\vartheta^{a_1a_3\cdots a_p}+\cdots. \label{3-10}\end{equation} Thus \begin{equation} \delta\vartheta^m\wedge\Big(\vartheta_m\hook\vartheta^{a_1\cdots a_p}\Big)= \delta\vartheta^{a_1} \wedge \vartheta^{a_2...a_p}-\delta\vartheta^{a_2}\wedge\vartheta^{a_1a_3\cdots a_p}+\cdots. \label{3-11}\end{equation} Comparing (\ref{3-9}) and (\ref{3-11}) we obtain the required formula (\ref{3-7}). The second formula (\ref{3-8}) is the consequence of the previous one, because the Hodge dual of a wedge product monomial is also a monomial. \vrule height7pt width4pt depth1pt \\ Therefore, we have the following formula for variation of the dual form \begin{equation} \delta*\vartheta^a=\delta\vartheta_m\wedge *(\vartheta^m\wedge\vartheta^a) \label{3-12}\end{equation} Thus the variation of the dual form is essentially differ from the variation $\delta\vartheta^a$ of the basis form $\vartheta^a$.\\ For the variation of the volume element the formula (\ref{3-8}) yields \begin{equation} \delta*1=\delta\vartheta^m \wedge (e_m \hook *1)=\delta\vartheta^m \wedge *(\vartheta_m\wedge *^2 1)= -\delta\vartheta^m \wedge *\vartheta_m. \label{3-13}\end{equation} \begin{proposition} For an arbitrary differential form $\alpha \in \Omega^p$ on a $n$-dimensional manifold the following commutative relation holds \begin{equation} \delta*\alpha - *\delta\alpha= *\Big(\delta\vartheta_m\hook(\vartheta^m\wedge \alpha)-\delta\vartheta_m\wedge(\vartheta^m\hook\alpha)\Big). \label{3-14}\end{equation} \end{proposition} \noindent{\bf Proof:} Evaluating the $p$-form $\alpha$ in the basis component forms $$ \alpha=\alpha_{a_1...a_p}\vartheta^{a_1...a_p} $$ and using the relation (\ref{3-8}) we derive \begin{eqnarray*} \delta*\alpha&=& \delta(\alpha_{a_1...a_p})*\vartheta^{a_1...a_p}+\alpha_{a_1...a_p}\delta*\vartheta^{a_1...a_p}\\ &=&\delta(\alpha_{a_1...a_p})*\vartheta^{a_1...a_p}+ \alpha_{a_1...a_p}\delta\vartheta^m\wedge(e_m\hook *\vartheta^{a_1...a_p})\\ &=&\delta(\alpha_{a_1...a_p})*\vartheta^{a_1...a_p}+\delta\vartheta^m\wedge(\vartheta_m\hook *\alpha). \end{eqnarray*} On the other hand \begin{eqnarray*} \delta\alpha&=& \delta(\alpha_{a_1...a_p})\vartheta^{a_1...a_p}+\alpha_{a_1...a_p}\delta(\vartheta^{a_1...a_p})\\ &=&\delta(\alpha_{a_1...a_p})\vartheta^{a_1...a_p}+\delta\vartheta^m\wedge(\vartheta_m\hook *\alpha). \end{eqnarray*} Therefore \begin{eqnarray*} \delta*\alpha-*\delta\alpha&=& \delta\vartheta_m\wedge(\vartheta^m\hook *\alpha)-*\Big(\delta\vartheta_m\wedge(\vartheta^m\hook *\alpha)\Big)\\ &=&\Big((-1)^{p(n-p)+1}\delta\vartheta_m\wedge*(\vartheta^m\wedge\alpha)- *(\delta\vartheta_m\wedge(\vartheta^m\hook\alpha))\Big)\\ &=&*\Big[\delta\vartheta_m\hook(\vartheta^m\wedge\alpha)-\delta\vartheta_m\wedge(\vartheta^m\hook\alpha)\Big] \end{eqnarray*} This proves the relation (\ref{3-14}). \vrule height7pt width4pt depth1pt\\ It is easy to see from the proof of the proposition above that in a particular case of basis forms the following (anti)commutativity relations hold \begin{equation} \delta*\vartheta^{a_1...a_p} \pm *\delta\vartheta{a_1...a_p} = *\Big(\delta\vartheta_m\hook\vartheta^{ma_1...a_p}\pm \delta\vartheta_m\wedge(\vartheta^m\hook\vartheta^{a_1...a_p})\Big). \label{3-15} \end{equation} The commutative relation similar to (\ref{3-14}) was proposed for the first time in \cite{Hehl}. The slightly different form is exhibited in \cite{BFK}. The relation (\ref{3-14}) coincides with the master formula of \cite{Hehl} in a case of a pseudo-orthonormal coframe. \sect{Variations and constrains.} The ordinary variational problem is used to study the critical points of a real-valued functional. The variations of the field variables are considered to be independent - free variations. In order to study a constraint physical system, for instant a motion of a particle on a fixed surface, one consider a critical point of a functional restricted by appropriative constrains. This restriction is usually incorporated by the method of Lagrangian multipliers. Another view on the constrained variational problem can be proposed by consideration of constraint variations. These variations satisfy some additional conditions. In the framework of a global $SO(1,3)$ invariant and diffeomorphic covariant theory we are interesting in covariant constraint conditions. Denote the variation of the basis 1-form $\vartheta^a$ by \begin{equation} \delta\vartheta^a=\delta\vartheta^a={\epsilon^a}_b\vartheta^b. \label{4-1}\end{equation} We will refer to the matrix $\epsilon_{ab}=\eta_{ac}{\epsilon^c}_b$ as the \textsl{ variational matrix.}\sf\\ The variational matrix admits a natural covariant decomposition \begin{equation} \epsilon_{ab}=\epsilon^{(1)}_{ab}+\epsilon^{(2)}_{ab}+\epsilon^{(3)}_{ab}, \label{4-2}\end{equation} where \begin{eqnarray} \epsilon^{(1)}_{ab}&=&{\frac{1}{2}}(\epsilon_{ab}-\epsilon_{ba}) \qquad \qquad \textsl{ - antisymmetric variation,}\sf\label{4-3}\\ \epsilon^{(2)}_{ab}&=&{\frac{1}{2}}(\epsilon_{ab}+\epsilon_{ba}) - \epsilon \eta_{ab} \qquad \textsl{- traceless symmetric variation,}\sf\label{4-4}\\ \epsilon^{(2)}_{ab}&=&\epsilon \eta_{ab} \qquad \qquad \qquad \qquad \textsl{ {- trace variation.}}\sf\label{4-5} \end{eqnarray} We use here and later the notation $\epsilon={\epsilon^a}_a=\epsilon_{ab}\eta^{ab}$ for a trace of the variation matrix. We can study now a free variation with an arbitrary variational matrix as well as various constrain variations, determined by a special choice of the variational matrix $\epsilon_{ab}$.\\ Let us examine the conditions given by the different constrains.\\ The question is: What condition should be imposed on the variations of $\{\vartheta^a\}$ in order to obtain some commutativity conditions with the Hodge star operator?\\ Consider two following cases \begin{equation} \delta*\vartheta^a=*\delta\vartheta^a \qquad \qquad\quad \textsl{{- commuatativity}} \sf\label{4-6}\end{equation} \begin{equation} \delta*\vartheta^a=-*\delta\vartheta^a \qquad \qquad \textsl{{- anticommuatativity}} \sf \label{4-7}\end{equation} Using (\ref{3-12}) in the right hand sides of (\ref{4-6},\ref{4-7}) and considering they together we obtain \begin{equation} *\delta\vartheta^a=\pm\delta\vartheta_m\wedge *(\vartheta^m\wedge\vartheta^a). \label{4-8}\end{equation} Hence the (anti)commutativity conditions take the form \begin{equation} *\delta\vartheta^a=\pm\delta\vartheta_m\hook\vartheta^{ma}. \label{4-9}\end{equation} Inserting the variational matrix notation we derive \begin{eqnarray*} {\epsilon^a}_b\vartheta^b&=&\pm{\epsilon_m}^k\vartheta_k\hook\vartheta^{ma} =\pm{\epsilon_m}^k(\delta^m_k\vartheta^a-\delta^a_k\vartheta^m)\\ &=&\pm({\epsilon_m}^m\vartheta^a-{\epsilon_m}^a\vartheta^m)= \pm(\epsilon d^a_b-{\epsilon_b}^a)\vartheta^b. \end{eqnarray*} Therefore in the notations of the variational matrix the (anti)commutativity conditions (\ref{4-9}) take the form \begin{equation} {\epsilon^a}_b = \pm(\epsilon\delta^a_b-{\epsilon_b}^a). \label{4-10}\end{equation} Taking the trace in two sides of this matrix equation we obtain $$ \epsilon=\pm 3\epsilon. $$ Thus the first necessary condition for (anti)commutativity of the variation with the Hodge dual is \begin{equation} \epsilon=0. \label{4-11} \end{equation} The equation (\ref{4-10}) gives the second condition for (anti)commutativity \begin{equation} \epsilon_{ab}\pm\epsilon_{ba}=0. \label{4-12}\end{equation} The consideration above can be summarized in the following \begin{proposition}\\ The variation commutes with the Hodge dual of a basis 1-form \begin{equation} \delta*\vartheta^a=*\delta\vartheta^a \label{4-13}\end{equation} if and only if the variational matrix $\epsilon^{ab}$ is antisymmetric.\\ The variation anticommutes with the Hodge dual of a basis 1-form \begin{equation} \delta*\vartheta^a=-*\delta\vartheta^a \label{4-14}\end{equation} if and only if the variational matrix $\epsilon^{ab}$ is traceless and symmetric. \end{proposition} Consider now the variation of the metric tensor $$ \delta g_{\mu \nu}=\delta(\eta_{ab}\vartheta^a_\mu\vartheta^b_\nu)= \eta_{ab}(\delta\vartheta^a_\mu\vartheta^b_\nu +\vartheta^a_\mu \delta\vartheta^b_\nu). $$ But $$ \delta\vartheta^a={\epsilon^a}_b\vartheta^b=\delta(\vartheta^a_\mu dx^\mu)= \delta\vartheta^a_\mu dx^\mu=\vartheta^\mu_b\delta\vartheta^a_\mu\vartheta^b. $$ Therefore $$\vartheta^\mu_b\delta\vartheta^a_\mu={\epsilon^a}_b,$$ or $$\delta\vartheta^a_\mu={\epsilon^a}_b\vartheta^b_\mu.$$ Hence the variation of the metric tensor is \begin{equation} \delta g_{\mu \nu}=\eta_{ab} ({\epsilon^a}_c\vartheta^c_\mu\vartheta^b_\nu+{\epsilon^b}_c\vartheta^a_\mu\vartheta^c_\nu)= \epsilon_{bc}(\vartheta^c_\mu\vartheta^b_\nu+\vartheta^b_\mu\vartheta^c_\nu). \label{4-15}\end{equation} Note that the expression in the brackets is symmetric under the permutation of the indices $c$ and $b$. Thus the variation of the metric tensor $\delta g_{\mu \nu}$ is identically equal to zero if and only if the matrix $\epsilon_{bc}$ is antisymmetric. Hence the commutativity of the variational operator with the Hodge dual map results in a zero variation of the metric tensor.\\ In the second (anti-commutativity) case the variation of the metric tensor is generally differ from zero. This variation is traceless and symmetric so it represents a 9-parametric family of transformations. \\ The traceless variations preserve the determinant of the metric tensor $g^{\mu\nu}$. Indeed the relation (\ref{4-15}) yields: \begin{equation} g^{\mu\nu}\delta g_{\mu\nu}=2\epsilon_{bc}\eta^{bc}=2\epsilon. \label{4-16}\end{equation} The variation of the determinant of the metric tensor can be written as \begin{equation} \delta g=gg^{\mu\nu}\delta g_{\mu\nu}. \label{4-17}\end{equation} Therefore in the case of the traceless variations \begin{equation} \delta g=2\epsilon g=0. \label{4-18}\end{equation} Note a connected result. Since $\delta*1=\epsilon *1$ it follows that the volume element is preserved if and only if $\epsilon=0$. Summarize the conclusions in the following table: \vspace{0.3cm} \newpage \qquad {\bf{ Constrained variations and commutativity rules.}} \vspace{0.4cm} \begin{tabular}{|c|c|c|c|c|} \hline Type of&Variational&Metric &Variation&Commutativity \\ variation & matrix &tensor&of volume&rules \\ \hline & & & & \\ free & $\epsilon_{ab}$ &$\delta g^{\mu\nu}\ne 0$ &$ \delta*1 \ne 0$& \\ \hline & & & & \\ antisymmetric&$\epsilon^{(1)}_{ab}$&$\delta g^{\mu\nu}=0$&$\delta*1=0$&$\delta*\vartheta^a=*\delta\vartheta^a$ \\ \hline & & & & \\ symmetric&$\epsilon^{(2)}_{ab}+\epsilon^{(3)}_{ab}$&$\delta g^{\mu\nu}\ne 0$&$\delta*1\ne0$&\\ \hline & & & &\\ traceless&$\epsilon^{(1)}_{ab}+\epsilon^{(2)}_{ab}$&$\delta g^{\mu\nu}\ne 0$&$\delta*1=0$&\\ \hline traceless& & & &\\ symmetric&$\epsilon^{(2)}_{ab}$&$\delta g^{\mu\nu}\ne 0$&$\delta*1=0$&$\delta*\vartheta^a=-*\delta\vartheta^a$\\ \hline \end{tabular} \vspace{0.3cm} We have established the commutativity conditions only for the variations of the basis 1-forms, but they can be extended straightforward for basis forms of an arbitrary degree. For instant, for a form $\alpha$ of an arbitrary degree the formula (\ref{3-14}) yields \begin{eqnarray*} \delta*\alpha-*\delta\alpha&=&*{\epsilon_m}^k\Big(\vartheta_k\hook(\vartheta^m\wedge\alpha)- \vartheta_k\wedge(\vartheta^m\hook\alpha)\Big)\\ &=&*{\epsilon_m}^k\Big(\delta^m_k\alpha-\vartheta^m\wedge(\vartheta_k\hook\alpha)- \vartheta_k\wedge(\vartheta^m\hook\alpha)\Big)\\ &=&*\epsilon\alpha-*\epsilon_{mk}\Big(\vartheta^k\hook(\vartheta^m\wedge\alpha)+ \vartheta^k\wedge(\vartheta^m\hook\alpha)\Big) \end{eqnarray*} The first term in the last expression vanish in the case of a traceless variation. The expression in the brackets is symmetric under the underchange of the indices $k$ and $m$. Thus in the case of an antisymmetric variational matrix $\epsilon_{mk}$ the variation commutes with the Hodge dual. In the case of a basic form of an arbitrary degree the relation (\ref{3-15}) yields to the anti-commutative condition for traceless symmetric variation \begin{equation} \delta*\vartheta^{a_1...a_p}=-*\delta\vartheta^{a_1...a_p}. \end{equation} We see that the commutativity (anti-commutativity) of the variation with the Hodge dual operator is connected with a certain special case of constrained variation of the coframe field. The special case of commutativity was proved for the first time in \cite{Hehl}. \sect{General quadratic Lagrangian} A typical Lagrangian used in the field theory is quadratic. It is useful to have a formula for a variational derivative of a such type expression. \begin{proposition} For $\alpha,\beta \in \Omega^p$ on a $n$-dimensional manifold the variation of the square-norm expression $\mathcal{L}=\alpha\wedge*\beta$ takes a following form \begin{equation} \delta(\alpha\wedge*\beta)=\delta\alpha \wedge*\beta +\alpha\wedge*\delta\beta-\delta\vartheta_m \wedge J^m, \label{5-1}\end{equation} where \begin{equation} J^m=(\vartheta^m\hook\beta)\wedge *\alpha-(-1)^p\alpha\wedge (\vartheta^m\hook*\beta). \label{5-2}\end{equation} \end{proposition} \noindent{\bf Proof:} Using the Leibniz rule for the variation (\ref{3-3}) we write \begin{equation} \delta(\alpha\wedge*\beta)=\delta\alpha \wedge*\beta +\alpha\wedge \delta*\beta. \label{3-17}\end{equation} The formula for commutator (\ref{3-14}) yields \begin{equation} \delta(\alpha\wedge*\beta)=\delta\alpha \wedge*\beta +\alpha\wedge*\delta\beta + \alpha\wedge *\Big(\delta\vartheta_m\hook(\vartheta^m\wedge \beta)-\delta\vartheta_m\wedge(\vartheta^m\hook\beta)\Big). \label{3-18}\end{equation} Compute the third term on the right hand side of (\ref{3-18}) \begin{eqnarray*} \alpha\wedge *\Big(\delta\vartheta_m\hook(\vartheta^m\wedge \beta)\Big) &=&\alpha\wedge *^2\Big(\delta\vartheta_m\wedge *(\vartheta^m\wedge \beta)\Big)\\ &=&(-1)^{p(n-p)+1}\alpha\wedge\delta\vartheta_m\wedge *(\vartheta^m\wedge \beta)\\ &=&(-1)^p\delta\vartheta_m\wedge\alpha\wedge (\vartheta^m\hook*\beta). \end{eqnarray*} The fourth term on the right hand side of (\ref{3-18}) is \begin{eqnarray*} \alpha\wedge *\Big(\delta\vartheta_m\wedge(\vartheta^m\hook\beta)\Big) &=&\delta\vartheta_m\wedge(\vartheta^m\hook\beta)\wedge *\alpha. \end{eqnarray*} These prove the statement. \vrule height7pt width4pt depth1pt\\ The $n-1$ form $J^m$ represents a certain type of a field current. Because of a symmetry of the expression (\ref{5-1}) under the permutation of $\alpha$ and $\beta$ we can rewrite the value $J^m$ in an explicitly symmetric form: \begin{eqnarray} J^m&=&\frac 12\Big([(\vartheta^m\hook\beta)\wedge *\alpha+(\vartheta^m\hook\alpha)\wedge *\beta]\nonumber\\ && \qquad -(-1)^p[\alpha\wedge (\vartheta^m\hook*\beta)+\beta\wedge (\vartheta^m\hook*\alpha)]\Big). \label{3-20}\end{eqnarray} In the special case $\alpha=\beta$ we obtain \begin{equation} J^m=(\vartheta^m\hook\alpha)\wedge *\alpha-(-1)^p\alpha\wedge (\vartheta^m\hook*\alpha). \label{3-21}\end{equation} The $n-1$ form of current $J^m$ can be written component-wisely as \begin{equation} J^m=J^{mn}*\vartheta_n. \label{3-22}\end{equation} Using \begin{equation} *(\vartheta^k\wedge J^m)=J^{mn}*(\vartheta^k\wedge*\vartheta_n) \label{3-23}\end{equation} we obtain the explicit expression for the matrix $J^{mn}$ \begin{equation} J^{mn}=*(\vartheta^n\wedge J^m). \label{3-24}\end{equation} The two indexed object $J^{mn}$ can be considered as an analog of the energy-momentum tensor. \begin{proposition} The matrix $J^{mn}$ is symmetric \begin{equation} J^{mn}=J^{nm}. \label{3-25}\end{equation} \end{proposition} \noindent{\bf Proof:} For the proof we use a simpler expression (\ref{5-2}) for $J^{mn}$, but take in account that this object is actually symmetric for a permutation of the forms $\alpha$ and $\beta$. Using (\ref{5-2},\ref{3-24}) we derive \begin{eqnarray*} J^{mn}&=&*(\vartheta^n\wedge J^m)=*\Big[\vartheta^n\wedge\Big((\vartheta^m\hook\beta)\wedge *\alpha- (-1)^p\alpha\wedge (\vartheta^m\hook*\beta)\Big)\Big]\\ &=&(-1)^{p-1}*\Big((\vartheta^m\hook\beta)\wedge(\vartheta^n\wedge*\alpha)+ (\vartheta^n\wedge\alpha)\wedge (\vartheta^m\hook*\beta)\Big). \end{eqnarray*} Write \begin{equation} \vartheta^n\wedge*\alpha=(-1)^{\sigma_1}*^2\vartheta^n\wedge*\alpha=(-1)^{\sigma_1}*(\vartheta^n\hook\alpha), \label{3-26}\end{equation} \begin{equation} \vartheta^m\hook*\beta=*(\vartheta^m\wedge *^2\beta)=(-1)^{\sigma_2}*(\vartheta^m\wedge \beta), \label{3-27}\end{equation} where $\sigma_1,\sigma_2$ are two integers. Thus we obtain \begin{eqnarray} J^{mn}&=&(-1)^{p-1}*\Big((-1)^{\sigma_1}(\vartheta^m\hook\beta)\wedge*(\vartheta^n\hook\alpha)+\nonumber\\ &&(-1)^{\sigma_2}(\vartheta^n\wedge\alpha)\wedge *(\vartheta^m\wedge \beta)\Big). \label{3-28}\end{eqnarray} Now the symmetry of $J^{mn}$ in the form (\ref{3-28}) under the permutation of the forms $\alpha$ and $\beta$ yields the symmetry under the permutation of the indices $m$ and $n$. \vrule height7pt width4pt depth1pt\\ Calculate the trace of the matrix $J^{mn}$. \begin{eqnarray} {J^m}_m&=&*(\vartheta_m\wedge J^m)=*\Big[\vartheta_m\wedge \Big((\vartheta^m\hook\beta)\wedge *\alpha- (-1)^p\alpha\wedge(\vartheta^m\hook *\beta)\Big)\Big]\nonumber\\ &=&* \Big(\vartheta_m\wedge(\vartheta^m\hook\beta)\wedge *\alpha-\alpha\wedge\vartheta_m\wedge(\vartheta^m\hook *\beta)\Big). \end{eqnarray} Use the formula \begin{equation} \vartheta_m\wedge(\vartheta^m\hook w)=p w, \label{3-29}\end{equation} where $w$ is an arbitrary $p$-form. The trace takes the form \begin{equation} {J^m}_m=p\beta\wedge *\alpha-(n-p)\alpha\wedge *\beta=-(n-2p)*(\alpha\wedge *\beta). \label{3-30}\end{equation} Thus we obtain \begin{proposition} For a quadratic Lagrangian \begin{equation} \mathcal{L}=\alpha\wedge*\beta \qquad deg\alpha=deg\beta=p \label{3-31}\end{equation} on a manifold $M$ the matrix $J^{mn}$ is traceless ${J^m}_m=0$ if and only if the dimensional of the manifold $M$ is even. \end{proposition} The second condition, namely $deg\alpha=deg\beta=\frac n2$, does not so crucial as it can be supposed on a first look. Indeed, for a 4-dimensional manifold three different cases are possible \begin{eqnarray} \label{3-31a} deg\alpha&=&deg\beta=2,\\ \label{3-31b} deg\alpha&=&deg\beta=1,\\ \label{3-31c} deg\alpha&=&deg\beta=2. \end{eqnarray} In the second case we can rewrite the Lagrangian as \begin{equation} \mathcal{L}=\alpha\wedge*\beta=\vartheta^a\wedge\alpha\wedge(\vartheta_a\hook*\beta)= (\vartheta^a\wedge\alpha)\wedge*(\vartheta_a\wedge\beta). \end{equation} Now this is a product of two forms of a second degree. In the third case (\ref{3-31c}) we can apply the same transformation twice.\\ The variational procedure described above can be straightforward extended for a general situation with an external field on the manifold $\{M,\vartheta^a\}$. Consider, for instance, the Lagrangian \begin{equation} \mathcal{L}=\alpha(\phi,d\phi,\vartheta^a)\wedge*\beta(\phi,d\phi,\vartheta^a), \label{3-21a}\end{equation} where $\phi$ is a certain external field (possible multicomponent) - differential form of degree $q$. Write the variation of the forms $\alpha,\beta$ as \begin{equation} \delta\alpha=\delta\phi\wedge\frac{\delta\alpha}{\delta\phi}+\delta d\phi\wedge\frac{\delta\alpha}{\delta d\phi}+ \delta\vartheta^m\wedge\frac{\delta\alpha}{\delta\vartheta^m}, \label{3-22a}\end{equation} \begin{equation} \delta\beta=\delta\phi\wedge\frac{\delta\beta}{\delta\phi}+\delta d\phi\wedge\frac{\delta\beta}{\delta d\phi}+ \delta\vartheta^m\wedge\frac{\delta\beta}{\delta\vartheta^m}. \label{3-23a}\end{equation} The relations (\ref{5-1}) yields \begin{eqnarray} \delta\mathcal{L}&=&(\delta\phi\wedge\frac{\delta\alpha}{\delta\phi}+\delta d\phi\wedge\frac{\delta\alpha}{\delta d\phi}+ \delta\vartheta^m\wedge\frac{\delta\alpha}{\delta\vartheta^m}) \wedge*\beta +\nonumber\\ && (\delta\phi\wedge\frac{\delta\beta}{\delta\phi}+\delta d\phi\wedge\frac{\delta\beta}{\delta d\phi}+ \delta\vartheta^m\wedge\frac{\delta\beta}{\delta\vartheta^m})\wedge*\alpha -\delta\vartheta_m \wedge J^m. \label{3-24a}\end{eqnarray} Thus we obtain two Euler-Lagrange field equations \begin{equation} \frac{\delta\alpha}{\delta\phi}\wedge*\beta+\frac{\delta\beta}{\delta\phi}\wedge*\alpha- (-1)^q d\Big(\frac{\delta\alpha}{\delta d\phi}\wedge*\beta+ \frac{\delta\beta}{\delta d\phi}\wedge*\alpha\Big)=0, \label{3-26a}\end{equation} \begin{equation} \frac{\delta\alpha}{\delta\vartheta^m} \wedge*\beta +\frac{\delta\beta}{\delta\vartheta^m}\wedge*\alpha=J^m. \label{3-27a}\end{equation} In two following sections we apply the variational procedure described above for two important cases: the Maxwell Lagrangian for electromagnetism and the Rumpf Lagrangian for translation invariant gravity. \sect{Maxwell Lagrangian on a teleparallel space The Lagrangian of the Maxwell theory of electromagnetism can be written as \begin{equation} \mathcal{L}=dA\wedge*dA, \label{6-1}\end{equation} where $A$ is a 1-form of electro-magnetic potential. This form of Lagrangian represents the classical theory of electromagnetism on a flat Minkowskian space.\\ If one consider the field $A$ to be defined on a curved pseudo-Riemannian manifold the 4-form $\mathcal{L}$ is well defined (invariant under the transformation of coordinates). The Hodge star operator in this situation depends on the metric $g$ on the manifold. Thus it is better to write the electro-magnetic Lagrangian on a pseudo-Riemannian manifold as \begin{equation} \mathcal{L}=dA\wedge*_gdA. \label{6-2}\end{equation} The variation of the Lagrangian have to be applied by using a free variation of the electro-magnetic potential $A$ as well as a free variation of the metric $g$.\\ On a teleparallel manifold the Lagrangian can be taken in the same form, but now the Hodge star operator depends on the coframe field $\vartheta^a$ \begin{equation} \mathcal{L}=dA\wedge*_{\vartheta^a}dA. \label{6-3}\end{equation} The variation of the Lagrangian (\ref{6-3}) can be produced using the formulas (\ref{5-1}, \ref{5-2}). \begin{equation} \delta\mathcal{L}=\delta(dA)\wedge*dA+dA\wedge*\delta(dA)-\delta\vartheta_m \wedge J^m, \label{6-4}\end{equation} where \begin{equation} J^m=(\vartheta^m\hook dA)\wedge *dA-dA\wedge (\vartheta^m\hook*dA). \label{6-5}\end{equation} Extracting the total derivative in (\ref{6-4}) \begin{eqnarray*} \delta\mathcal{L}&=&2d(\delta A)\wedge*dA-\delta\vartheta_m \wedge J^m\\&=& 2d(\delta A\wedge*dA)+2\delta A\wedge d*dA-\delta\vartheta_m \wedge J^m \end{eqnarray*} we obtain two field equations \begin{equation} d*dA=0 \label{6-6}\end{equation} and \begin{equation} J^m=0, \label{6-6a}\end{equation} where \begin{equation} J^m=(\vartheta^m\hook dA)\wedge *dA-dA\wedge (\vartheta^m\hook*dA). \label{6-7}\end{equation} The equation (\ref{6-6}) has exactly the same form as the ordinary electro-magnetic field equation, but it is an equation on a curved manifold (operator $*$ depends on the coframe $\vartheta^a$). For interpretation the equations (\ref{6-6},\ref{6-6a}) introduce the strength notation for electro-magnetic field \begin{equation} dA=F. \label{6-8}\end{equation} The first field equation (\ref{6-6}) takes the form of ordinary Maxwell equations \begin{eqnarray} dF&=&0,\\ \label{6-9} d*F&=&0. \label{6-10} \end{eqnarray} The equation (\ref{6-6a}) gives an additional relation \begin{equation} J^m=(\vartheta^m\hook F)\wedge *F-F\wedge (\vartheta^m\hook*F). \label{6-10a}\end{equation} The 2-form $F$ can be explained component-wisely as \begin{equation} F=F_{mn}\vartheta^{mn}. \label{6-8a}\end{equation} Note that the coefficients of $F$ are antisymmetric by definition $F_{mn}=-F_{nm}$. Consider the first term of $J^m$ (\ref{6-10a}) \begin{eqnarray*} (\vartheta^m\hook F)\wedge *F&=& (\vartheta^m\hook F^{kn}\vartheta_{kn})\wedge *F_{pq}\vartheta^{pq}\\ &=&2F^{kn}F_{pq}\delta^m_k\vartheta_n\wedge *\vartheta^{pq}=4F^{mn}F_{nq}*\vartheta^q \end{eqnarray*} The second term of (\ref{6-10a}) takes the form \begin{eqnarray*} F\wedge (\vartheta^m\hook*F)&=&F^{kn}\vartheta_{kn}\wedge(\vartheta^m\hook*F_{pq}\vartheta^{pq}) =-F^{kn}F_{pq}\vartheta_{kn}\wedge*\vartheta^{mpq}\\&=& F^{kn}F_{pq}\vartheta_{k}\wedge*(\delta^n_m\vartheta^{pq}-2\delta^n_p\vartheta^{mq})\\&=& F^{kn}F_{pq}*(2\delta^n_m\delta^k_p\vartheta^{q}-2\delta^n_p\delta^k_m\vartheta^{q}+ 2\delta^n_p\delta^k_q\vartheta^{m})\\&=&4F^{pm}F_{pq}*\vartheta^{q}+2F^{pq}F_{pq}*\vartheta^{m} \end{eqnarray*} Thus the object $J^m$ takes the form \begin{equation} J^m=8(F^{mn}F_{nk}-\frac 14 F^{pq}F_{pq}\delta^m_k) *\vartheta^k. \label{6-11}\end{equation} or in a component-wise form \begin{equation} {J^m}_q=8(F^{mn}F_{nk}-\frac 14 F^{pq}F_{pq}\delta^m_k). \label{6-12}\end{equation} We see now what is the field equation (\ref{6-7}) mean. It is a vanishing condition for the trace of the matrix ${J^m}_q$. Note that due to the propositions (5.3-5.4) the traceless condition as well as the symmetric condition are the consequences of the definition of $J_{ab}$. Thus in the case of the Maxwell Lagrangian the object ${J^m}_n$ coincides (up to a numerical coefficient equal to $-\frac{1}{32\pi}$) with the classical expression of the energy-momentum tensor. Note that by the variational procedure described above we obtain it in a symmetric and a traceless form. \sect{Translational Lagrangian in gravity In the teleparallel approach to gravity the coframe field $\vartheta^a$ is the basic gravitational field variable. A general Lagrangian density for the coframe field $\vartheta^a$ (quadratic in the first order derivatives) described by the gauge invariant translation Lagrangian of Rumpf \cite{Rumpf}. Up to the $\Lambda$-term it can be written as \begin{equation} \mathcal{L}=\frac 1{2\ell^2}\sum^3_{I=1}\rho_I{}^{(I)}V, \label{7.1} \end{equation} where \begin{eqnarray} \label{7.1a} {}^{(1)}\mathcal{L}&=&d\vartheta^a\wedge*d\vartheta_a,\\ \label{7.2} {}^{(2)}\mathcal{L}&=&(d\vartheta_a \wedge \vartheta^a ) \wedge*( d\vartheta_b \wedge \vartheta^b),\\ \label{7.3} {}^{(3)}\mathcal{L}&=&(d\vartheta_a \wedge\vartheta^b ) \wedge * \Big(d\vartheta_b \wedge \vartheta^a \Big). \end{eqnarray} and $\ell$ is the Plank length constant. \\ Note that the Lagrangian (\ref{7.1}) is a linear combination of three independent terms every one of which is of the prescribed form $\alpha\wedge *\beta$ and we can apply the procedure described above. Using the proposition (5.1) we obtain for the first Lagrangian (\ref{7.2}) \begin{eqnarray} \delta\Big({}^{(1)}\mathcal{L}\Big)&=&2\delta(d\vartheta_a)\wedge*d\vartheta^a-\delta(\vartheta_m)\wedge {}^{(1)}J^m \nonumber\\ &=&2d\Big(\delta(\vartheta_a)\wedge*d\vartheta_a\Big)+ 2\delta(\vartheta_a)\wedge d*d\vartheta^a-\delta(\vartheta_m)\wedge {}^{(1)}J^m, \label{7.4} \end{eqnarray} where \begin{eqnarray} {}^{(1)}J^m&=& (\vartheta^m\hook d\vartheta_a)\wedge *d\vartheta^a - d\vartheta^a\wedge (\vartheta^m\hook *d\vartheta_a)\nonumber\\ &=&2(\vartheta^m\hook d\vartheta_a)\wedge *d\vartheta^a-\vartheta^m\hook(d\vartheta_a\wedge*d\vartheta^a) \label{7.5} \end{eqnarray} Thus the contribution of the Lagrangian ${}^{(1)}\mathcal{L}$ in the field equation is \begin{equation} \boxed{2\rho_1\delta(\vartheta_a)\wedge d*d\vartheta^a+2\rho_1(\vartheta^a\hook d\vartheta_m)\wedge *d\vartheta^m- \rho_1\vartheta^a\hook(d\vartheta_m\wedge*d\vartheta^m)} \label{7.6} \end{equation} Variation of the second Lagrangian (\ref{7.3}) takes the form \begin{eqnarray} \delta\Big({}^{(2)}\mathcal{L}\Big)&=&2\delta(d\vartheta_a \wedge \vartheta^a)\wedge*(d\vartheta_b \wedge \vartheta^b)- \delta(\vartheta_m)\wedge {}^{(2)}J^m\nonumber\\ &=&2d(\delta\vartheta_a) \wedge \vartheta^a\wedge*(d\vartheta_b \wedge \vartheta^b) +2d\vartheta_a \wedge \delta\vartheta^a\wedge*(d\vartheta_b \wedge \vartheta^b)-\nonumber\\ &&\delta(\vartheta_m)\wedge {}^{(2)}J^m\nonumber\\ &=&2d\Big(\delta\vartheta_a \wedge \vartheta^a\wedge*(d\vartheta_b \wedge \vartheta^b)\Big) +2\delta\vartheta_a \wedge d\Big(\vartheta^a\wedge*(d\vartheta_b \wedge \vartheta^b)\Big)\nonumber\\ && +2\delta\vartheta^a\wedge d\vartheta_a \wedge*(d\vartheta_b \wedge \vartheta^b)- \delta(\vartheta_m)\wedge {}^{(2)}J^m\nonumber\\ &=&2d\Big(\delta\vartheta_a \wedge \vartheta^a\wedge*(d\vartheta_b \wedge \vartheta^b)\Big)- 2\delta\vartheta_a \wedge \vartheta^a\wedge d*(d\vartheta_b \wedge \vartheta^b)+\nonumber\\ && 4\delta\vartheta^a\wedge d\vartheta_a \wedge*(d\vartheta_b \wedge \vartheta^b)- \delta(\vartheta_m)\wedge {}^{(2)}J^m, \label{7.7} \end{eqnarray} where the current term \begin{eqnarray} {}^{(2)}J^m &\!\!\!\!\!=\!\!\!\!\!& \Big(\vartheta^m\hook (d\vartheta_b \wedge \vartheta^b)\Big)\wedge *(d\vartheta_a \wedge \vartheta^a)+ (d\vartheta_a \wedge \vartheta^a)\wedge \Big(\vartheta^m\hook*(d\vartheta_b \wedge \vartheta^b)\Big) \nonumber\\ &\!\!\!\!\!=\!\!\!\!\!& -\vartheta^m\hook\Big((d\vartheta^a\wedge\vartheta_a)\wedge*(d\vartheta^b\wedge\vartheta_b)\Big)+ 2\vartheta^m\hook(d\vartheta^a\wedge\vartheta_a)\wedge*(d\vartheta^b\wedge\vartheta_b) \nonumber\\ &\!\!\!\!\!=\!\!\!\!\!& -\vartheta^m\hook\Big((d\vartheta^a\wedge\vartheta_a)\wedge*(d\vartheta^b\wedge\vartheta_b)\Big)+ 2(\vartheta^m\hook d\vartheta^a)\wedge\vartheta_a\wedge*(d\vartheta^b\wedge\vartheta_b)\nonumber\\ &\!\!\!\!\!\!\!\!\!\!& +2d\vartheta^m\wedge*(d\vartheta^b\wedge\vartheta_b) \label{7.8} \end{eqnarray} Thus the contribution of the Lagrangian ${}^{(2)}\mathcal{L}$ in the field equation is \begin{equation} \boxed{ \begin{array}{ll} &-2\rho_2\vartheta^a\wedge d*(d\vartheta_b \wedge \vartheta^b) +\rho_2\vartheta^a\hook\Big((d\vartheta^m\wedge\vartheta_m)\wedge*(d\vartheta^n\wedge\vartheta_n)\Big)+\\ &2\rho_2 d\vartheta_a \wedge*(d\vartheta_b \wedge \vartheta^b)+ -2\rho_2(\vartheta^a\hook d\vartheta^m)\wedge\vartheta_m\wedge*(d\vartheta^n\wedge\vartheta_n) \end{array}} \label{7.9} \end{equation} As for the third laplacian (\ref{7.4}) by the proposition (5.1) we obtain \begin{eqnarray} \delta\Big({}^{(3)}\mathcal{L}\Big)\eqq2\delta(d\vartheta_a \wedge \vartheta^b)\wedge*(d\vartheta_b \wedge \vartheta^a)- \delta\vartheta_m\wedge {}^{(3)}J^m=- \delta\vartheta_m\wedge {}^{(3)}J^m+\nonumber\\ &\!\!\!\!\!\!\!\!\!\!& 2d(\delta\vartheta_a) \wedge \vartheta^b\wedge*(d\vartheta_b \wedge \vartheta^a)+ 2\delta\vartheta^b\wedge d\vartheta_a \wedge*(d\vartheta_b \wedge \vartheta^a)\nonumber\\ &\!\!\!\!\!=\!\!\!\!\!& 2d\Big(\delta\vartheta_a \wedge \vartheta^b\wedge*(d\vartheta_b \wedge \vartheta^a)\Big)- 2\delta\vartheta_a \wedge \vartheta^b\wedge d*(d\vartheta_b \wedge \vartheta^a)+\nonumber\\ &\!\!\!\!\!\!\!\!\!\!& 4\delta\vartheta^b\wedge d\vartheta_a \wedge*(d\vartheta_b \wedge \vartheta^a)- \delta\vartheta_m\wedge {}^{(3)}J^m. \label{7.10} \end{eqnarray} As for the current term \begin{eqnarray} {}^{(3)}J^m &\!\!\!\!\!=\!\!\!\!\!& \Big(\vartheta^m\hook(d\vartheta_a \wedge \vartheta^b)\Big)\wedge *(d\vartheta_b \wedge \vartheta^a)+ (d\vartheta_a \wedge \vartheta^b)\wedge \Big(\vartheta^m\hook*(d\vartheta_b \wedge \vartheta^a)\Big)\nonumber\\ &\!\!\!\!\!=\!\!\!\!\!& -\vartheta^m\hook\Big((d\vartheta_a \wedge \vartheta^b)\Big)\wedge *(d\vartheta_b \wedge \vartheta^a) +2\vartheta^m\hook(d\vartheta_a \wedge \vartheta^b)\wedge *(d\vartheta_b \wedge \vartheta^a)\nonumber\\ &\!\!\!\!\!=\!\!\!\!\!& -\vartheta^m\hook\Big((d\vartheta_a \wedge \vartheta^b)\Big)\wedge *(d\vartheta_b \wedge \vartheta^a) +2(\vartheta^m\hook d\vartheta_a) \wedge \vartheta^b\wedge *(d\vartheta_b \wedge \vartheta^a)+\nonumber\\ &\!\!\!\!\!=\!\!\!\!\!& 2d\vartheta_a\wedge *(d\vartheta^m \wedge \vartheta^a) \label{7-11} \end{eqnarray} Hence the contribution of the Lagrangian ${}^{(3)}\mathcal{L}$ in the field equation is \begin{equation} \boxed{ \begin{array}{ll} &-2\rho_3 \vartheta^b\wedge d*(d\vartheta_b \wedge \vartheta^a) +\rho_3\vartheta^a\hook\Big((d\vartheta_m \wedge \vartheta^n)\Big)\wedge *(d\vartheta_n \wedge \vartheta^m)\\ &-2\rho_3(\vartheta^a\hook d\vartheta_m) \wedge \vartheta^n\wedge *(d\vartheta_n \wedge \vartheta^m) +2\rho_3d\vartheta_b\wedge *(d\vartheta^a \wedge \vartheta^b) \end{array}} \label{7-12} \end{equation} The relations (\ref{7.6},\ref{7.9},\ref{7-12}) yield the field equation generated by free variations of the Rumpf Lagrangian cf. Kopczy\'nski \cite{Kop} in the form \cite{Hehl}. \begin{eqnarray}\label{fe} - 2\ell^2\Sigma_a &=& 2\rho_1 d*d\vartheta_a - 2\rho_2 \vartheta_a \wedge d *(d\vartheta^b\wedge \vartheta_b)- 2\rho_3\vartheta_b \wedge d*( \vartheta_a \wedge d \vartheta^b ) \nonumber \\ && + \rho_1 \Big[ e_a \hook( d\vartheta^b \wedge * d\vartheta_b) -2 ( e_a \hook d\vartheta^b ) \wedge *d\vartheta_b\Big] \nonumber \\ && + \rho_2\Big[2 d\vartheta_a \wedge * ( d\vartheta^b \wedge \vartheta_b) + e_a \hook ( d\vartheta^c \wedge\vartheta_c \wedge * ( d\vartheta^b \wedge\vartheta_b) )\nonumber\\ &&\qquad -2( e_a \hook d\vartheta^b) \wedge \vartheta_b \wedge * ( d\vartheta^c \wedge \vartheta_c)\Big] \nonumber \\ && +\rho_3\Big[2 d\vartheta_b \wedge * ( \vartheta_a\wedge d \vartheta^b) + e_a \hook( \vartheta_c \wedge d\vartheta^b \wedge *( d\vartheta^c \wedge \vartheta_b )) \nonumber \\ &&\qquad - 2( e_a \hook d\vartheta^b) \wedge \vartheta_c \wedge * ( d\vartheta^c \wedge\vartheta_b )\Big] , \end{eqnarray} where $\Sigma_a$ depends on matter fields. \sect{Conclusions} We discuss the variational procedure on a teleparallel manifold. The situation is rather differ from the variation on a pseudo-Riemannian manifold. We have reproved a part of the propositions exhibited in the paper \cite{Hehl} and have generalized some of them. The commutativity and anti-commutativity of the variation with the Hodge dual map are related with the special covariant types of the constraint variation. We derive a general relation for variation of the quadratic type Lagrangians and apply it to the viable cases of the electro-magnetic Lagrangian and to the translation invariant of Rumpf. The established formulas can be applied for study the various material fields on teleparallel background. We hope that these techniques will also be useful for Lagrangian formulation of the dynamic of $p$-branes. \vspace{1.3cm} \bf{Acknowledgments}:\\ \sf I am grateful to Prof. Shmuel Kaniel for constant support and interesting discussions. My gratitude to Prof. Friedrich W.~Hehl and to his collaborates for useful comments and helpful remarks. This work is considerably influenced by their paper \cite{Hehl}.
\section{Local $U(N)$: Yang-Mills integrals} {\it In this part we discuss some aspects of results obtained in collaboration with W.~Krauth and H.~Nicolai and published in} \cite{kns},\cite{ks1},\cite{ks2}. Consider $D$-dimensional pure $SU(N)$ Yang-Mills field theory and, inspired by the principle (\ref{equi}), reduce it by brute force to zero dimensions. The continuum path integral, involving traceless hermitian gauge connections $X_\mu$, becomes an ordinary matrix integral: \begin{equation} {\cal Z}_{D,N} = \int \prod_{A=1}^{N^2-1} \prod_{\mu=1}^{D} \frac{d X_{\mu}^{A}}{\sqrt{2 \pi}} \exp \bigg[ \frac{1}{2} {\rm Tr}\, [X_\mu,X_\nu] [X_\mu,X_\nu] \bigg]. \label{bosint} \end{equation} Note that gauge fixing is no longer required here, since the overcounting of gauge-equivalent configurations involves merely a factor of the compact, finite volume of the gauge group: space time has become a point (or more precisely, an infinitesimal torus, since the ``point'' still keeps a sense of the $D$ directions.). Now, as was explained in \cite{gk}, the integral eq.(\ref{bosint}) still ``knows'' something about $D$-dimensional space-time. Indeed, shifting \begin{equation} X_\mu \rightarrow P_\mu + X_\mu \label{shift} \end{equation} by diagonal matrices $P_\mu=$diag$(p^1_\mu, \ldots, p^N_\mu)$ we formally recover Feynman rules which look like the ordinary ones except that the momentum integrations are replaced by sums over discretized momenta $p^i_\mu-p^j_\mu$. As $N \rightarrow \infty$ one might hope that the sums turn back into loop integrals, motivating the correspondence (\ref{equi}). Now in \cite{gk} a somewhat complicated quenching and gauge fixing procedure was introduced in order to ensure the recovery of the field theory. Indeed it would seem at first sight that the integral eq.(\ref{bosint}) is meaningless without the procedure of \cite{gk} since there are unconstrained flat directions in integration space, due to mutually commuting matrices. However, the Monte Carlo results of \cite{ks1} suggest \vspace{0.5cm}\noindent {\bf Proposition 1a:} The Yang-Mills integrals ${\cal Z}_{D,N}$ exist iff $N> {D \over D-2}$. \vspace{0.5cm} \noindent It would be quite important to find methods enabling one to rigorously prove this statement, or even calculate the partition sums ${\cal Z}_{D,N}$. Some important analytic evidence comes from the perturbative calculations of \cite{nishi}. For $SU(2)$, a proof of the proposition, as well as an analytic expression for ${\cal Z}_{D,2}$, is known. The matrix integrals eq.(\ref{bosint}) have beautiful supersymmetric extensions in dimensions $D=4,6,10$. These read \begin{equation} {\cal Z}_{D,N}^{{\cal N}}:=\int \prod_{A=1}^{N^2-1} \Bigg( \prod_{\mu=1}^{D} \frac{d X_{\mu}^{A}}{\sqrt{2 \pi}} \Bigg) \Bigg( \prod_{\alpha=1}^{{\cal N}} d\Psi_{\alpha}^{A} \Bigg) \exp \bigg[ \frac{1}{2} {\rm Tr}\, [X_\mu,X_\nu] [X_\mu,X_\nu] + {\rm Tr}\, \Psi_{\alpha} [ \Gamma_{\alpha \beta}^{\mu} X_{\mu},\Psi_{\beta}] \bigg]. \label{susyint} \end{equation} where we have supersymmetrically added ${\cal N}=2 (D-2)$ hermitian fermionic matrices $\Psi_{\alpha}$ to the models. The $D=10$ model corresponds to the dimensional reduction of the maximally supersymmetric conformal $D=4,{\cal N}=4$ Yang-Mills field theory to zero dimensions. It is also the crucial ingredient in the IKKT model for IIB superstrings \cite{ikkt1}, which however, instead of taking the large $N$ limit, sums ${\cal Z}_{10,N}^{16}$ over all values of $N$. Following the $SU(2)$ calculations of \cite{sestern}, the perturbative arguments of \cite{ikkt2}, the arguments of \cite{greengut}, the calculations of \cite{moore}, and our Monte Carlo work, we are led to \vspace{0.5cm}\noindent {\bf Proposition 1b:} The susy Yang-Mills integrals ${\cal Z}_{4,N}^4,{\cal Z}_{6,N}^8,{\cal Z}_{10,N}^{16}$ exist iff $N \geq 2$. \vspace{0.5cm} \noindent The analytic results of these integrals are believed to be known, and a rigorous mathematical proof would be welcome. It is interesting to understand the similarities and differences of these little studied ``{\it new}'' matrix models eqs.(\ref{bosint}),(\ref{susyint}), whose existence has been missed until recently, in relation to the conventional ``{\it old}'' matrix models of Wigner type. A crucial quantity in the old matrix models is the distribution of eigenvalues of the random matrices. An interesting novel feature of the new matrix models is the fact that, at finite $N$, only a finite number of one-matrix moments exist. The numerical results agree with perturbative powercounting arguments, and one is led, for the bosonic models eq.(\ref{bosint}), to \vspace{0.5cm}\noindent {\bf Proposition 2a:} $\Big\langle {1 \over N}{\rm Tr}\, X_1^{2 k} \Big\rangle < \infty$ iff $k<N (D-2) - {3\over 2} D+2$, \vspace{0.5cm} \noindent while in the supersymmetric cases $D=4,6,10$ eq.(\ref{susyint}) one has \vspace{0.5cm}\noindent {\bf Proposition 2b:} $\Big\langle {1 \over N}{\rm Tr}\, X_1^{2 k} \Big\rangle < \infty$ iff $k<D-3$. \vspace{0.5cm} \noindent Once again, except for $SU(2)$, rigorous proofs of these conjectures are missing. These findings indicate that in the new matrix models the density of eigenvalues falls off much slower (powerlike) than in the old ones (exponential). As $N \rightarrow \infty$ the bosonic densities behave once again rather conservatively (infinitely many moments exist), while for the susy densities the behavior indicated in proposition 2b is {\it independent} of $N$. A much more difficult question is whether these models might lead to a ``self-quenching'' effect where a background $P_\mu$ (in eq.(\ref{shift})), bearing some resemblance to real Yang-Mills theory, is dynamically generated as $N \rightarrow \infty$. The above Yang-Mills integrals have many applications even at finite $N$ (for a recent unexpected one see \cite{hollo}); however, here we would like to stress that they constitute an ideal laboratory for developing new large $N$ techniques aimed at making progress with `t~Hooft's large $N$ QCD \cite{thooft}. \section{Global $U(N)$: Master partitions} The problem of finding the $N=\infty$ solution to matrix field theories has not even been solved in the presumably simpler case of models with a global $U(N)$ symmetry. The main obstacle has been that no systematic procedure was known to reduce the local number of degrees of freedom from ${\cal O} (N^2)$ to ${\cal O} (N)$. In \cite{master} we outlined a general approach for achieving such a reduction for any field theory with a global matrix symmetry. Let us sketch the idea in the specific example of an interacting $D=2$ hermitian scalar field theory. It is convenient to put the theory on a lattice: \begin{displaymath} {\cal Z}= \int \prod_x {\cal D} M(x)~e^{ -{\cal S}}, \end{displaymath} \begin{equation} {\cal S}= N {\rm Tr}\, \sum_x \bigg[ {1 \over 2} M(x)^2 + {g \over 4} M(x)^4 - {\beta \over 2} \sum_{\mu=1,2} [M(x) M(x+\hat{\mu}) + M(x) M(x-\hat{\mu})] \bigg], \label{mlattice} \end{equation} where the field variables are $N\times N$ hermitian matrices $M(x)$ defined on the square lattice sites $x$ and $\hat{\mu}$ denotes the unit vector in the $\mu$-direction. The measure is the usual flat measure on hermitian matrices. The first step consists in applying the reduction principle (\ref{equi}). Naively reducing the system as in the previous section down to a single point results in an ordinary one-matrix model where the information on the 2$D$ lattice is lost. A more careful reduction has to hide the propagation on the lattice in group space; here we will use the beautiful procedure of ``twisting'', see \cite{twisted} and references therein. Using the $N \times N$ Weyl-`t~Hooft matrices \begin{equation} P=\left( \begin{array}{cccccc} 0 & 1 & & & & \\ & 0 & 1 & & & \\ & & \ddots & \ddots & & \\ & & & & 0 & 1\\ 1 & & & & & 0 \end{array} \right), \quad Q=\left( \begin{array}{cccccc} 1 & & & & & \\ & \omega & & & & \\ & & \ddots & & & \\ & & & & \omega^{N-2} & \\ & & & & & \omega^{N-1} \end{array} \right), \label{pq} \end{equation} where $\omega=\exp {2 \pi i \over N}$ and $P Q=\omega Q P$, one can show by a Fourier transform in matrix index space that the one-matrix integral \begin{equation} Z= \int {\cal D} M \exp N {\rm Tr}\, \bigg[ -{1 \over 2} M^2 - {g \over 4} M^4 +\beta \Big( M P M P^{\dagger} + M Q M Q^{\dagger} \Big) \bigg], \label{mtwist} \end{equation} has the same vacuum energy as the path integral eq.(\ref{mlattice}). As a second step we need to reduce the number of variables from $N^2$ to $N$. The brute force approach would be to diagonalize the matrix $M$ and perform the integration over the unitary diagonalizing matrix. One would then obtain an effective action for the $N$ eigenvalues of $M$. However, calculations at small $N$ show that this effective action is extremely complicated in the case at hand. On the other hand, if we change variables from the eigenvalues to {\it partitions}, corresponding to a Fourier transform in group space, something very interesting happens. The $N$ variables dual to the $N$ eigenvalues are the Young weights $h_i=N-i+m_i$, $i=1, \ldots, N$, where the $m_i$ are the lengths of the $i$'th row in the Young diagram corresponding to the partition. Denoting the partitions through $h=(h_1, \ldots, h_N)$, the dual representation of the integral eq.(\ref{mtwist}) is found to be \begin{equation} Z=\sum_h {\cal I}_h~{\cal L}_h~\beta^{{|h| \over 2}}, \label{mfull} \end{equation} where instead of an integration over the $N\times N$ matrix $M$ we now have a sum over all partitions $h$ of the non-negative integer $|h|=0,1,2,\ldots$. Here ${\cal I}_h$ contains all the information on the interaction, and essentially requires the general correlation function of the $U(N)$-invariant one-matrix integral \begin{equation} {\cal I}_h=N^{|h|} \prod_{i=1}^N {(N-i)! \over h_i!}~ \int {\cal D} M~\exp~N {\rm Tr}\, \Big[ -{1 \over 2} M^2 -{g \over 4} M^4 \Big]~\chi_h(M), \label{hermitian} \end{equation} which are known. Here $\chi_h(M)$ are the Schur functions on $h$ which are nothing but a complete set of of class functions (non-abelian Fourier modes) on the group. The information on the lattice is contained in the {\it lattice polynomials} \begin{equation} {\cal L}_h=\exp{1 \over N} {\rm Tr}\, \Big(\partial P \partial P^{\dagger} + \partial Q \partial Q^{\dagger} \Big) \cdot \chi_h(J)~\Big|_{J=0}. \label{poly} \end{equation} Here $\partial$ denotes the $N \times N$ matrix differential operator whose matrix elements are $\partial_{j i}={\partial \over \partial J_{i j}}$. The ${\cal L}_h$ are easily shown to be polynomials in the variable ${1 \over N}$ of maximal degree ${1 \over 2} |h| -1$. Now the result of this harmonic analysis is that the terms to be summed over partitions in eq.(\ref{mfull}) {\it factorize} into a piece ${\cal I}_h$ containing the information on the local interaction and and the piece ${\cal L}_h$ containing the information on the space-time structure. Since there are only $N$ variables $h_i$ we expect the sum eq.(\ref{mfull}) to be dominated at $N=\infty$ by a saddle point, i.e.~an effective master partition. In a third and final step we will need to write the full system of bootstrap equations for the saddle point. This will require a deeper analysis of the lattice polynomials. But it should be clear that the problem of solving the large $N$ lattice field theory has been reformulated in a rather non-trivial way: In fact, the interacting theory (i.e. $g \neq 0$ in eq.(\ref{mlattice})) is no harder to solve in this dual space of Young weights than the free theory ($g=0$). \vskip0.5cm \noindent {\large \bf Acknowledgements} \smallskip \noindent I thank W.~Krauth and H.~Nicolai for fruitful collaboration, and J.~Hoppe, V.~A.~Kazakov, I.~K.~Kostov, and J.~Plefka for useful discussions. This work was supported in part by the EU under Contract FMRX-CT96-0012.
\section{ \Large $B$-Physics and the CKM Matrix}\label{subsec:Tev} A major objective in the study of bottom hadrons is determining the elements of the Cabbibo-Kobayashi-Maskawa (CKM) matrix~\cite{CKM}, and to stringently test its adequacy. This matrix transforms the flavor eigenstates of quarks into their mass eigenstates, which are not the same in the Standard Model (SM). A convenient parameterization in powers of the Cabibbo angle ($\lambda \equiv \sin \,\theta_C = |V_{us}|$) is due to Wolfenstein~\cite{Wolf}: $$ V_{CKM} \equiv \left( \begin{array}{ccc} V_{ud} & V_{us} & V_{ub} \\ V_{cd} & V_{cs} & V_{cb} \\ V_{td} & V_{ts} & V_{tb} \\ \end{array} \right) = \left( \begin{array}{ccc} 1-\frac{\lambda^2}{2} & \lambda & A\lambda^3(\rho-i\eta) \\ { -\lambda} & 1-\frac{\lambda^2}{2} & { A\lambda^2} \\ { A\lambda^3(1-\rho-i\eta)} & -A\lambda^2 & { 1} \end{array} \right) + O(\lambda^4). $$ The imaginary term $\eta$ was conjectured by Kobayashi and Maskawa to be the source of $CP$ violation, which has been an outstanding issue for the last 35 years. Constraints on the CKM matrix from the $b$-sector initially came from lifetime and branching ratio measurements in the early '80's. In 1986, a new window was opened by the observation of $B^0$-$\overline{B}{^0}$ mixing in an unresolved mixture of $B^0_d$ and $B^0_s$ by UA1~\cite{UA1} in $\bar{p}p$ collisions, and subsequently for pure $B^0_d$'s by ARGUS~\cite{ARGUS} at the $\Upsilon(4S)$. Through mixing, one gains access to the $t$ CKM elements, an important consideration given the limitations of direct top studies. Global fits to experimental data constrain the four parameters of the CKM~\cite{Mele,Ali}, with $\lambda$ and $A$ already known quite well. Constraining $\eta$ and $\rho$ has been the recent focus of $B$-physics. One of the unitarity constraints ($V_{tb}^*V^{\,}_{td} + V_{cb}^*V^{\,}_{cd}+V_{ub}^*V^{\,}_{ud} = 0$), is graphically represented in Fig.~\ref{fig:Unitarity1} as a triangle in the complex $\rho$-$\eta$ plane, with the apex at ($\rho$,$\eta$). Its base is of unit length, leaving three angles and two sides that may be measured. $B^0$-$\overline{B}{^0}$ mixing constrains the right leg ($\propto\!V_{td}/V_{ts}$), and $CP$ violation in $B^0/\overline{B}{^0} \rightarrow J/\psi K^0_S$ decays determines the angle $\beta$. Mixing studies of $B^0$ mesons have greatly advanced in the '90's, and we are entering a new stage at the close of the millennium with the advent of $CP$-violation measurements in $B^0$ mesons. The contributions from the Fermilab Tevatron Collider program to these efforts from Run I (1992-6) data are discussed. The two collider experiments, CDF~\cite{CDF} and D{\O}~\cite{D0}, are well known, and their descriptions are not repeated here. \begin{figure}[t] {\raggedright \centering \epsfxsize=20pc \epsfbox{hc_mele_triangle_shaded_bfrBd_beta.eps} \caption{ The CKM unitarity triangle along with constraints derived from $CP$ violation in $K^0$'s ($\varepsilon_K$), and the rate of charmless $B$ decays ($V_{ub}/V_{cb}$) (derived from Mele\protect~\protect\cite{Mele}). Bands indicate ``$1\sigma$'' allowed regions. The angle $\beta = arg(-V^{\,}_{cd}V^*_{cb}/V^{\,}_{td}V^*_{tb}) = \arctan(\eta/[1-\rho])$. } \label{fig:Unitarity1} } \end{figure} \section{ \Large $B^0$-$\overline{B}{^0}$ Mixing Measurements} \subsection{ \Large $B$ Mixing and Flavor Tagging} Like the $K^0$-$\overline{K}{^0}$ system, 2nd order weak ``box'' diagrams result in oscillations between $B^0$ and $\overline{B}{^0}$ mesons. The frequency is the mass difference ($\Delta m = m_H - m_L$) between the heavy/light mass eigenstates $B^0_H$ and $B^0_L$. The dominant effect arises from diagrams with virtual top quarks, with $\Delta m_q \sim m^2_t F(m^2_t/m^2_W) |V^*_{tb}V_{tq}|$ for $B^0_q$ mesons. Thus, $\Delta m_q$ constrains the CKM element relating transitions between top and the light quark $q$ composing the $B^0_q$, and relates to the right leg of the unitarity triangle (Fig.~\ref{fig:Unitarity1}). The probability that an initially pure $B^0$ state decays as a $\overline{B}{^0}$ (and {\it vice versa}) at proper time $t$ is ${\cal P}_{Mix}(t) = [1 -\cos(\Delta m t)]e^{-t/\tau}/2\tau$. The asymmetry between the mixed (${\cal P}_{Mix}$) and unmixed (${\cal P}_{Un}$) states is therefore \begin{equation} {\cal A}_0(t) \equiv \frac{P_{Un}(t)-P_{Mix}(t)}{P_{Un}(t)+P_{Mix}(t)} = \cos(\Delta m_q t). \label{eq:MixAsym} \label{eq:true_asym} \end{equation} Observing $B^0$ mixing is predicated upon determining the $b$ ``flavor''---whether the $B$ is composed of a $b$ or a $\bar{b}$ quark---at the times of production and decay. The decay flavor is usually known from the $B$ reconstruction. More problematic is tagging the initial flavor. If this is correct with probability $P_0$, then the observed asymmetry is attenuated by the ``dilution'' ${\cal D}_0 = 2P_0 -1$, {\it i.e.} ${\cal A}_{Obs} = {\cal D}_0 \cos(\Delta m_q t)$. A tagger with efficiency $\epsilon$ yields an error on the asymmetry which scales as $1/\sqrt{\epsilon {\cal D}^2_0 N}$ for $N$ background-free mesons, and thus $\epsilon {\cal D}^2_0$ measures the tagger's effective power. \subsection{ \Large Time-Integrated Mixing Measurements} \label{sec:timeint} It is not necessary to measure the proper decay time to observe mixing, since \begin{equation} \chi \equiv \int^\infty_0{\cal P}_{Mix}(t) dt = \frac{x^2}{2(1+x^2)}, \;\; {\rm with}\;\; x \equiv \frac{\Delta m}{\Gamma}, \label{eq:chi} \end{equation} is nonzero. At the Tevatron both $B^0_d$ and $B^0_s$ are produced, and unless one explicitly identifies the $B$-species one measures an average $\overline{\chi} \equiv f_d\chi_d + f_s\chi_s$, for fractions $f_d$ and $f_s$ of the $B^0_d$ and $B^0_s$ contributions. CDF and D{\O} have measured $\overline{\chi}$ using dileptons, where the leptons identify both the $B$ and tag its flavor. Like-sign pairs indicate that one $b$-hadron has mixed. In $10$ pb$^{-1}$ of dimuon triggers ($p_T^\mu>3$ GeV/$c$) D{\O} found 59 like-sign (LS) and 113 unlike (US) pairs. The ratio $LS/US = 0.43 \pm 0.07 (stat.)\pm 0.05(syst.)$ is used in conjunction with models of other processes ($b\rightarrow c \rightarrow \ell^+$ sequential decays, $c\bar{c}$, fake leptons,\ldots etc.) to extract $\overline{\chi} = 0.09 \pm 0.04 \pm 0.03$~\cite{D0Mix}. Similarly, CDF has used $20$ pb$^{-1}$ of dimuons to obtain $\overline{\chi} = 0.118 \pm 0.021 \pm 0.026$~\cite{CDFmumuMix}; and in $e$-$\mu$ events $\overline{\chi} = 0.130 \pm 0.010 \pm 0.009$~\cite{CDFemuMix}. All agree with $\overline{\chi} = 0.118 \pm 0.006$ from the PDG~\cite{NewPDG}. \subsection{ \Large Time-Dependent $B^0_d$-Mixing Measurements} \label{subsec:TimDep} With the advent of precision vertex detectors direct observation of the $B^0_d$ oscillation has overshadowed the $\overline{\chi}$ analyses. The D{\O} detector will have such tracking capability starting in Run II~\cite{D0up}, and such studies have been restricted to CDF in Run I. Six analyses have been reported ($\sim\!\! 100$ pb$^{-1}$), and are summarized~in~Fig.~\ref{fig:CDFdmdSum}. Two analyses are extensions of the time-integrated measurements described above: dilepton samples ($\sim\!6000\;\mu\mu$, $\sim\!\!10000\; e\mu$) are used where the leptons define both the $B$ signal and its flavor. In this case a secondary vertex associated to a lepton is sought to establish the $B$-decay vertex. Average $\beta \gamma$ corrections transform the observed momentum and decay length into the proper decay time. The inclusive nature of the selection allows other processes to contribute ($c\bar{c}$, fakes, etc.\ldots). Their contributions to the sample are constrained by kinematic variables, such as the relative $p_T$ of the lepton to other parts of the decaying $B$. The samples are more than 80\% pure $b\bar{b}$. The oscillation is revealed in the time variation of the like-sign dilepton fraction, and fits extract $\Delta m_d$ (see ``$\mu/\mu$''~\cite{CDFmumuOsc} and ``$e/\mu$'' in Fig.~\ref{fig:CDFdmdSum}). In another dilepton analysis~\cite{tomoko} the $B^0_d$ is more cleanly identified by reconstructing a $D^*$ near a lepton. From the lepton on the other side one infers the initial flavor of the $B \!\rightarrow\! \ell^+ D^{*-} X$ meson. The oscillation from a signal of $\sim\!500$ events is shown in Fig.~\ref{fig:CDFdmdSum} under ``$D^*$lep/lep,'' along with its $\Delta m_d$. The cleanliness of the sample results in a small systematic error ($^{+0.31}_{-0.38}$), but at the price of worse statistical precision. Inclusive lepton triggers ($e$ and $\mu$, $p_T > 8$ GeV/$c$) are used to obtain a tagged $B$ sample, where the $B$ flavor on the opposite side is identified by reconstructing a $D^{(*)-}$. The resulting oscillation for $\sim\!800$ events is labeled as ``$D^{(*)}$/lep'' in Fig.~\ref{fig:CDFdmdSum}~\cite{stephen}. \begin{figure}[t] \begin{center} \epsfxsize=15pc \epsfxsize=14pc \epsfig{file=mix_dmd_comp3.eps, width=12.5cm,clip=,angle=-90} \caption{ Summary of the six CDF $B^0_d$ oscillation measurements of $\Delta m_d$. In each case an oscillation is observed in a charge-correlation asymmetry as a function of proper decay length. The average CDF $\Delta m_d$ accounts for correlations between measurements. \label{fig:CDFdmdSum} } \end{center} \end{figure} We next consider the ``lep/$Q^{jet}$lep'' analysis~\cite{CDFjetqOsc}, which again uses a lepton+vertex to identify a $b$-hadron and a second lepton as a tag. It also uses the subtle technique of ``jet-charge'' tagging. The sample arises from inclusive lepton triggers, where we search for another lepton as a tag similar to the earlier analyses. We do not dwell on this aspect. More interesting is the use of ``jet-charge'' tagging, where an average charge of a jet opposite the $\ell$+vertex is used to infer the initial flavor of the $b$-hadron producing the trigger lepton. The charge of the away-side jet is defined as \begin{equation} Q_{jet} \equiv \frac{\Sigma_i q_i (\vec{p}_i \cdot \hat{a})} {\Sigma_i \vec{p}_i \cdot \hat{a} }, \label{eq:JetQ} \end{equation} where $q_i$ and $\vec{p}_i$ are the charge and momentum of the $i$-th track in the jet, and $\hat{a}$ is the unit vector pointing along the jet axis. A negative (positive) $Q_{jet}$ implies the jet contained a $b$ ($\bar{b}$). One finds that the tag purity is higher for larger $|Q_{jet}|$. The sample composition is again determined via kinematic variables like the $p_T^{rel}$ of the trigger lepton and the invariant mass of the secondary vertex. The fit results of the tagging asymmetry for $\sim\!\frac{1}{4}$ million events are given in Fig.~\ref{fig:CDFdmdSum} as ``lep/$Q^{jet}$lep.'' The sixth analysis uses ``Same-Side Tagging'' (SST), where a track near the reconstructed $B$ tags its flavor~\cite{Gronau}, rather than using the other $b$-hadron. The idea is simple. A $\bar{b}$ quark hadronizing into a $B^0_d$ picks up a $d$ in the fragmentation, leaving a $\bar{d}$. To make a charged pion, the $\bar{d}$ picks up a $u$ making a $\pi^+$. Conversely, a $\overline{B}{^0_d}$ will be associated with a $\pi^-$. Correlated pions also arise from $B^{**+} \rightarrow B^{(*)0} \pi^+$ decays. Both sources have the same correlation, and are not distinguished here.\footnote{ \normalsize A recent CDF $B^{**}$ analysis of a $\ell D^{(*)}$ sample found the fraction of $B$ mesons arising from $B^{**}$ states to be $0.28 \pm 0.06 \pm 0.03$~\protect\cite{Dejan}. $B^{**}$ mesons are indeed a significant source of correlated pions. } CDF adopted a ``$p_T^{rel}$'' algorithm, where the tag is the candidate track with the smallest momentum component transverse to the $B$+Track momentum. A charged particle is a valid SST candidate if it is reconstructed in the Si-$\mu$vertex detector (SVX), has $p_T > 400$ MeV/$c$, is within $\Delta R = \sqrt{(\Delta\eta)^2+(\Delta\phi)^2} \le 0.7$ of the $B$, and its impact parameter is within $3\sigma$ of the primary vertex. SST is applied to almost 10,000 $B \rightarrow \ell^+ D^{(*)}X$ events reconstructed via four $B^0_d$ decay signatures and one for $B^+$~\cite{SSTPRL,SSTPRD}. The sample composition is unraveled in the fit (including $B^0_d \leftrightarrow B^+$ cross-talk). The $\chi^2$ fit results are shown in Fig.~\ref{fig:CDFdmdSum} as ``$D^*$lep/SST,'' with the upper plot showing the $B^0_d \rightarrow \ell^+ {D}{^-}X$ reconstruction and the lower one is for $\ell^+ \overline{D}{^{*-}}X$. Along with $\Delta m_d$, one also obtains the dilution, ${\cal D}_0 = 0.181^{+0.036}_{-0.032}$, of this SST method. This analysis provides the dilution calibration for the SST analyses of Sec.~\ref{sec:cp}. The six $\Delta m_d$ results are combined, accounting for correlations, into a CDF average of $0.481 \pm 0.026\pm 0.026$ ps$^{-1}$. This is comparable to other experiments, and is in good agreement with the PDG value of $0.464 \pm 0.018$ ps$^{-1}$~\cite{NewPDG}. \subsection{ \Large Time-Dependent $B^0_s$-Mixing Measurements} $B^0_s$ mixing follows the same formalism as for $B^0_d$'s, except the relevant CKM element is $V_{ts}$ rather than $V_{td}$. Measurements of $\overline{\chi}$ (Sec~\ref{sec:timeint}), along with $\chi_d$ measured at the $\Upsilon(4S)$ and estimates for the species fractions $f_d$ and $f_s$, provided the original direct indication that $\chi_s$ is close to its asymptotic limit ($1/2$) and thereby insensitive to $\Delta m_s$. Further progress on $B^0_s$ mixing necessitates time-dependent methods. CDF has searched for $B^0_s$ oscillations~\cite{CDFBs} using dileptons (110 pb$^{-1}$ of $\mu\mu$ or $e \mu$) for $\ell^\pm\ell^+ \phi h^-$, where $h^-$ is a charged track associated to a $\phi \rightarrow K^+K^-$ decay vertex. This signature selects $B^0_s \rightarrow \ell^+ \nu D^-_s$, $D^-_s \rightarrow \phi \pi^- X$; exclusive reconstruction is not required to increase statistics. The resulting $K^+K^-$ mass distribution is shown in Fig.~\ref{fig:Bs}. The $B^0_s$ decay point is obtained by projecting the $\phi h^-$ decay back to the $\ell^+$. Monte Carlo corrections are applied to the $\beta\gamma$ factor for the proper time estimation. A sample of $1068 \pm 70$ $B^0_s$ candidates (purity of $61.0 ^{+4.4}_{-7.0}$\%) is obtained. \begin{figure}[t] {\centering \epsfxsize=13pc \epsfxsize=18pc \epsfbox{hc_bs_mix_phimass.eps} \epsfxsize=18pc \epsfbox{hc_bs_mix_amp_scan.eps} \caption{ \normalsize Left: The $K^+K^-$ mass distribution showing the $\phi$ peak. Right: The likelihood amplitude scan through $\Delta m_s$ (points), and the 95\% CL limits on the amplitude for statistical error (dashed curve) and statistical and systematic errors combined (solid curve). \label{fig:Bs} }} \end{figure} The $B^0_s$ flavor is inferred from the other trigger lepton, similar to the $\Delta m_d$ analyses. Events are again classified as ``unmixed'' ($\ell^\pm\ell^\mp_{tag}$) or ``mixed'' ($\ell^\pm\ell^\pm_{tag}$). Limits will be set rather than an observation of the oscillation, so one must know {\it a priori} the mistag rate $R_{m}$. This was found to be $R_{m} = \frac{1}{2}(1-{\cal D}_0) = 0.24 \pm 0.08$ from a likelihood fit of the mixed/unmixed fractions in the $\ell^+ \phi$ data. The $B^0_s$ oscillation is too rapid to influence the determination of $R_{m}$; rather it is governed by the sample contributions of $B^0_d$, $B^+$, $c\bar{c}$, sequential decays and fake background. The data are fit with an unbinned likelihood to describe the mixed versus unmixed components. The fit includes the various sources of events ($B^0_s$, $B^0_d$, $B^+$,\ldots) and $\Delta m_s$ is a free parameter. No oscillation is observed, and limits are set. The amplitude method~\cite{HGMoser} is adopted, whereby the functional form of $\cos(\Delta m_s t)$ is replaced by $A(\Delta m_s)\cos(\Delta m_s t)$, {\it i.e.} the amplitude is a free $\Delta m_s$-dependent parameter. For the true value $\Delta m_s^{true}$, $A(\Delta m_s^{true}) = 1$, and otherwise $A(\Delta m_s) = 0$. The result of the scan of $A(\Delta m_s)$ from the likelihood is shown in Fig.~\ref{fig:Bs}. The data fluctuate about zero, with no evidence for an oscillation. Values of $\Delta m_s$ are excluded at the 95\% CL if $A(\Delta m_s) + 1.645\sigma_A(\Delta m_s) \le 1$, and thus $\Delta m_s > 5.8$ ps$^{-1}$ at 95\% CL accounting for both statistical and systematic errors (Fig.~\ref{fig:Bs}, solid line). This result is competitive with other single tagging measurements. However, the world limit, $\Delta m_s > 12.4$ ps$^{-1}$ (95\% CL), is dominated by the multi-tag results from ALEPH and DELPHI~\cite{Parodi}. It is conceivable that $B^0_s$ oscillations are too rapid to be directly observed. If so, the width difference $\Delta \Gamma_s$ between the mass eigenstates is expected to be large. CDF has searched for two lifetime components in a $\ell D_s$ sample, and finds $\Delta \Gamma_s/\Gamma_s < 0.83$ at 95\% CL~\cite{CDFBsLife}. Given $\Delta \Gamma_s/\Delta m_s$ and the mean $B^0_s$ lifetime $\overline{\tau}_s$, this can be expressed as the upper bound $\Delta m_s < 96 \, {\rm ps}^{-1} \times [5.6 \times 10^{-3}/(\Delta \Gamma_s /\Delta m_s)] [1.55 {\rm ps}/\overline{\tau}_s]$ at 95\% CL., with $\Delta \Gamma_s /\Delta m_s = 5.6 \times 10^{-3}$ a recent estimate~\cite{Beneke}. The limit is weak, but with the increased statistics of Run II either $\Delta m_s$ or $\Delta \Gamma_s$ should be directly determined. \section{ \Large $CP$ Violation in $J/\psi K^0_S$} \label{sec:cp} The origin of $CP$ violation has been an outstanding question since its unexpected discovery in $K^0_L \rightarrow \pi^+\pi^-$ 35 years ago~\cite{KL_CP}. In 1972, before the discovery of charm, Kobayashi and Maskawa~\cite{CKM} proposed that this was the result of quark mixing with 3 (or more) generations. Unfortunately the $K^0$ has been the only place $CP$ violation has been observed. Despite precision $K^0$-studies, a complete picture of $CP$ violation is still lacking; and it is often argued that the CKM model can {\it not} be the full story~\cite{Cosmo}. $CP$-violation searches have encompassed $B$ mesons, but the effects in inclusive studies~\cite{CDFmumuMix,B_CP} are too small ($\sim\!\!10^{-3}$) to as yet detect. In the early '80's it was realized~\cite{CarterSanda} that the mixing {\it interference} of $B^0_d$ decays into the same $CP$ state could manifest large violations. Unfortunately these decays were, until recently, too rare to study. The ``golden'' mode for observing large $CP$ violation in $B$'s is $B^0_d/\overline{B}{^0_d} \rightarrow J/\psi K^0_S$,\footnote{ \normalsize Another mode of possible interest is $B^0_d \rightarrow J/\psi K^{*0}$. While this is not a $CP$ eigenstate, it can be decomposed into even and odd eigenstates by an angular analysis. This has been done by CDF for $B^0_d \rightarrow J/\psi K^{*0}$ (and $B^0_s \rightarrow J/\psi \phi$) to extract the decay matrix elements~\cite{Pappas}, but the Run~I statistics are insufficient to be of interest for measuring $CP$-violation parameters. } and, critically, it is related to the CKM matrix with little theoretical uncertainty. A $B^0_d$ may decay directly to $J/\psi K^0_S$, or the $B^0_d$ may oscillate into $\overline{B}{^0_d}$ and then decay to $J/\psi K^0_S$. These two paths have a phase difference, quantified by the angle $\beta$ of the unitarity triangle (Fig.~\ref{fig:Unitarity1}). This gives rise to a decay asymmetry \begin{equation} {\cal A}_{CP}(t) \equiv \frac{\overline{B}{^0_d}(t)-B^0_d(t) } {\overline{B}{^0_d}(t)+B^0_d(t) } = \sin(2\beta) \sin (\Delta m_d t), \label{eq:cp_asym} \end{equation} where $B^0_d(t)$ [$\overline{B}{^0_d}(t)$] is the number of decays to $J/\psi K^0_S$ at proper time $t$ given that the meson was a $B^0_d$ [$\overline{B}^0_d$] at $t=0$. OPAL investigated $CP$ violation with 24 $J/\psi K^0_S$ candidates (60\% purity), and obtained $\sin(2\beta) = 3.2^{+1.8}_{-2.0} \pm0.5$~\cite{OPALcp}. CDF has taken advantage of the large $B$ cross section at the Tevatron and obtained a sample of several hundred decays to $J/\psi K^0_S$ to measure $\sin(2\beta)$~\cite{CDFcp}. \subsection{ \Large Same Side Tagging Analysis of $J/\psi K^0_S$\protect} \label{sec:SSTcp} $J/\psi K^0_S$ candidates are selected from the $J/\psi \rightarrow \mu^+\mu^-$ sample (110 pb$^{-1}$). A time-dependent analysis demands that both muons are in the Si-$\mu$vertex detector (SVX), cutting away half of the $J/\psi$'s. The $K^0_S \rightarrow \pi^+\pi^-$ reconstruction tries all tracks, assumed to be pions. The $p_T(K^0_S)$ must be above 0.7 GeV/$c$, its decay vertex displaced from the $J/\psi$'s by more than $5\sigma$, and $p_T(B^0_d) > 4.5$ GeV/$c$. We construct $M_N \equiv (M_{FIT} - M_0)/\sigma_{FIT}$, where $M_{FIT}$ is the fitted $J/\psi K^0_S$ mass, $\sigma_{FIT}$ its error ($\sim\!\!9\;{\rm MeV}/c^2$), and $M_0$ is the central $B^0_d$ mass. The distribution for candidates with $ct >0$ is shown in Fig.~\ref{fig:JpsiKs}a. A likelihood fit yields (for all $ct$) $198 \pm 17$ $B^0_d/\overline{B}{^0_d}$'s. The initial flavor is tagged by the identical SST of the $\ell D^{(*)}$ $\Delta m_d$ measurement of Sec.~\ref{subsec:TimDep}. SST is independent of the $B^0_d$ decay mode, and therefore the dilution measurement can be transferred from one mode to another. However a small kinematic correction, determined from Monte Carlo, is made to translate the $\ell D^{(*)}$ dilution to the $J/\psi K^0_S$ sample due to the different $p_T(B)$ ranges~\cite{SSTPRD}. The appropriate ${\cal D}_0$ is $0.166 \pm 0.018 \pm 0.013$~\cite{CDFcp}, where the first error is due to the dilution measurements, and the second is due to the translation to the $J/\psi K^0_S$ sample. \begin{figure}[t] \begin{center} \epsfxsize=17.5pc \epsfxsize=22pc \epsfbox{ksh_m2_prl.eps} \epsfxsize=12.0pc \epsfxsize=16pc \epsfbox{hc_cp_ksh_vs_ct.eps} \caption{ \normalsize The normalized mass distribution of $J/\psi K^0_S$ events with $ct > 0$ (a) and $200 \, \mu$m (b). The curve is the Gaussian signal plus linear background from the likelihood fit (see text). The sideband-subtracted flavor asymmetry (c) as a function of the $J/\psi K^{0}_s$ proper decay length [Eq.~(\ref{eq:meas_asym})]: points are data, dashed curve is a simple $\chi^2$ fit, and the solid curve is the likelihood fit. The inset shows a scan of the log-likelihood function as ${\cal D}_0\sin(2\beta)$ is varied about the best fit value. \label{fig:JpsiKs} } \end{center} \end{figure} The SST method is applied to the $J/\psi K^0_S$ sample, with a resultant tagging efficiency of $\sim\!65\%$. Analogous to Eq.~(\ref{eq:cp_asym}), we compute the asymmetry \begin{equation} {\cal A}(ct) \equiv \frac{N^-(ct)-N^+(ct) } {N^-(ct)+N^+(ct) }, \label{eq:meas_asym} \end{equation} where $N^\pm(ct)$ are the numbers of positive and negative tags (implying $B^0_d$ and $\overline{B}{^0_d}$ respectively) in a given $ct$-bin. Signal and sideband regions are defined as $|M_N|<3$ and $3<|M_N|<20$, and the sideband-subtracted asymmetry of Eq.~(\ref{eq:meas_asym}) is plotted in Fig.~\ref{fig:JpsiKs}c. The dashed curve is a $\chi^2$ fit of ${\cal D}_0 \sin(2\beta)\sin(\Delta m_d t$) to the data, with $\Delta m_d$ fixed to $0.474 \;{\rm ps}^{-1}$~\cite{PDG1}. The amplitude, $0.36\pm 0.19$, measures $\sin(2\beta)$ attenuated by the dilution ${\cal D}_0$. Due to the $\sin(\Delta m_d t)$ shape the fit amplitude is driven by the asymmetries at larger $ct$'s, where backgrounds are small (see Fig.~\ref{fig:JpsiKs}b). The fit is refined using an unbinned likelihood fit. This makes optimal use of the low statistics by fitting the data in $M_N$ and $ct$, including sideband and $ct<0$ events which help constrain the background. The fit also incorporates resolutions and corrections for (small) systematic detector biases. The solid curves in Fig.~\ref{fig:JpsiKs}a,c are the result of the likelihood fit, which gives ${\cal D}_0\sin(2\beta)=0.31\pm 0.18$. As expected, both fits give similar values since the result is dominated by the sample size. The systematic uncertainty on ${\cal D}_0\sin(2\beta)$ is 0.03, dominated by the uncertainty on $\Delta m_d$ ($\pm 0.031$ ps$^{-1}$), but includes the effects from detector biases and the $B^0_d$ lifetime. To extract $\sin(2\beta)$ from the measured asymmetry the dilution must be divided out. However, as long as ${\cal D}_0 \!\not = \! 0$, the exclusion of $\sin(2\beta)\!=\! 0$ is {\it independent} of further knowledge of ${\cal D}_0$. Given ${\cal D}_0 > 0$, the unified frequentist approach of Feldman and Cousins~\cite{Feldman} yields a dilution-independent exclusion of $\sin(2\beta) \le 0$ at 90\% CL. From above, ${\cal D}_0 = 0.166 \pm 0.018 \pm 0.013$, which results in $\sin(2\beta)= 1.8 \pm 1.1 \pm 0.3$. The central value is unphysical since the amplitude of the raw asymmetry is larger than ${\cal D}_0$. This result corresponds to excluding $\sin(2\beta)<-0.20$ at a 95\% CL. \subsection{ \large Multi-Tagging Analysis of $J/\psi K^0_S$} Shortly after this conference CDF released an expanded $J/\psi K^0_S$ analysis using three taggers, as well as non-SVX data. These preliminary results are briefly summarized. The above SST result is statistically very limited. It can be improved by increasing the {\it effective} statistics by using lepton and jet-charge tagging to increase the total $\epsilon {\cal D}^2_0$. The raw statistics can also be increased by utilizing the $J/\psi$ candidates not reconstructed in the SVX. Precision lifetime information is lost, reducing the power of these events, but significant information remains. The $J/\psi K^0_S$ selection is otherwise similar to the SST-only analysis, and results in a total of $395 \pm 31$ $J/\psi K^0_S$ candidates ($202 \pm 18$ in the SVX, $193 \pm 26$ are non-SVX). SST, lepton, and jet-charge tagging are applied to this sample. Lepton tagging follows the $\Delta m_d$ analyses. The jet-charge algorithm is similar to the $\Delta m_d$ analysis but uses a ``mass'' jet algorithm rather than a ``cone'' based one. This improves the efficiency for identifying low-$p_T$ ``jets'' for the tag. A lepton tends to dominate the jet charge if a lepton tag is in the jet. Since lepton tagging has low efficiency but high dilution, the correlation between lepton and jet-charge tags is avoided by dropping the jet-charge tag if there is a lepton tag. The dilutions of these two methods are measured in $B^+ \rightarrow J/\psi K^+$ decays (Table~1), and are directly applicable to the $J/\psi K^0_S$ sample. The precision is not high, but this obviates the complex problem of translating dilutions from kinematically different samples. The tagged $J/\psi K^0_S$ events are fit in an unbinned likelihood fit ($\Delta m_d$ constrained to $0.464 \pm 0.018$ ps$^{-1}$~\cite{NewPDG}). The $\sin(2\beta)$ results for the individually tagged subsamples are listed in Table~1, \begin{table}[t] \caption{ \normalsize Multi-tag analysis of $J/\psi K^0_S$. There are two dilutions for SST, one for SVX events and another for non-SVX events.} \label{tab:multi} \begin{center} \begin{tabular}{|c|c|c|c|} \hline Tagger &Eff. (\%)&Dil. (\%)& $\sin(2\beta)$ \\ \hline \raisebox{0pt}[10pt][3pt]{$\!$SST$_{svx}\!$} & $\!35.5\pm 3.7\!$ & $\!16.6 \pm 2.2\!$ &\\ \raisebox{0pt}[3pt][5pt]{$\!$SST$_{non}\!$} & $\!38.1 \pm 3.9\!$ & $\!17.4 \pm 3.6\!$ & \raisebox{5pt}[7pt][0pt]{$\!\;2.03^{+0.84}_{-0.77}$}\\ \raisebox{0pt}[7pt][6pt]{$\!$Lepton$\!$} & $\!5.6 \pm 1.8\!$ &$\!62.5 \pm 14.6\!$ & $\!\;0.52^{+0.61}_{-0.75} \!$ \\ \raisebox{0pt}[7pt][7pt]{$\!$Jet-$Q\!$} & $\!40.2 \pm 3.9\!$ &$\!23.5 \pm 6.9\!$ & $\!\!-0.31^{+0.81}_{-0.85}\! $ \\ \hline \raisebox{0pt}[11pt][6pt]{$\!$Global$\!$} & \multicolumn{2}{|c|}{$\epsilon {\cal D}^2 = 6.3\pm1.7$} & $\!\;0.79 ^{+0.41}_{-0.44}\!$ \\ \hline \end{tabular} \end{center} \end{table} \begin{figure}[t] \centering \protect\epsfxsize=22pc \hspace*{-0.2cm}\protect\epsfbox{newsin2beta.eps} \protect\epsfxsize=12pc \caption{ \normalsize Multi-tag $\sin(2\beta)$ result. Left: time-dependent asym\-metry of SVX data; Right: time-integrated asym\-metry of non-SVX data. } \protect\end{figure} including systematic errors due to the dilutions, $\Delta m_d$, $\tau_{B^0}$, and $m_{B^0}$. The SST result is slightly larger than before with the inclusion of the non-SVX events, and the error has decreased by $\sim\!\!20$\%. The other two taggers fall in the physical range, one positive and the other negative. Rather than average these three results, the likelihood fitter is generalized to fit all three simultaneously while accounting for tag correlations. The global multi-tag result is $\sin(2\beta) = 0.79^{+0.41}_{-0.44}$ (Fig.~5), including the systematic uncertainties. This result corresponds to a unified frequentist confidence interval of $0.0 < \sin(2\beta) < 1$ at 93\% CL. Although the exclusion of zero has only slightly increased from 90\% for the (SVX) SST-only analysis, the uncertainty on $\sin(2\beta)$ is cut in half. \section{ \Large Summary and Prospects} Measurements of $\Delta m_{d}$ and $\Delta m_{s}$ provide important constraints on the CKM matrix. World averages constrain the triangle of Fig.~\ref{fig:Unitarity1} quite well, as shown in Fig.~\ref{fig:summary}. An indirect determination of $\sin(2\beta) = 0.77 ^{+0.09}_{-0.12}$ was reported at this conference~\cite{Ali}. This is much more precise than the direct CDF measurement; nevertheless, the agreement is an auspicious omen for the CKM model's account of $CP$ violation. Additional sources of $CP$ violation are, however, thought necessary to account for the baryon asymmetry in the universe~\cite{Cosmo}. Searching for physics beyond the CKM model demands stringent tests, and is the focus of dedicated $B$ factories. Both CDF and D{\O} will also be fully engaged in this effort by exploiting the rich $B$ harvest from Run II. Commencing in 2000, a two-year run will deliver $20\times$ the luminosity ($\sim\!\!2$~fb$^{-1}$), to be exploited by greatly enhanced detectors. With the ``baseline'' detector and trigger upgrades~\cite{D0up,CDFup} CDF projects 10,000 $J/\psi K^0_S$ dimuon triggers for a $\sin(2\beta)$ error of about $\pm 0.08$; and D{\O}, with a new precision tracking system, expects an error of 0.12-0.15 (Fig.~\ref{fig:summary}). Dielectron triggers may further increase the samples by $\sim\!50$\%. These errors are in the range projected for $B$ factories. The Run~II Tevatron will be competitive in many other areas of $B^0_d$ physics as well. $B^0_s$ oscillations have, so far, eluded all comers. The Tevatron should have a virtual monopoly on the $B^0_s$ after the closure of the $Z^0$ machines and before the start of the LHC. Run II baseline expectations are for CDF to reach $x_s$'s up to 30-40, and D{\O} up to 20-25. Although D{\O}'s reach is less, it is sufficient that failure to observe oscillations should critically challenge the CKM model. \begin{figure}[t] \begin{center} \epsfxsize=15.5pc \epsfxsize=20pc \epsfbox{manfred_mele_triangle.eps} \epsfxsize=13pc \epsfxsize=16pc \epsfbox{hc_my_sin2beta_newcdf_opal.eps} \vspace*{-0.4cm} \caption{ \normalsize Left: CKM constraints from $K^0_L$'s, charmless $B$ decays, $\Delta m_d$, and $\Delta m_s$ on ($\rho,\eta$). The ``$1\sigma$'' allowed regions of Fig.~\protect\ref{fig:Unitarity1} are collapsed to their central values, except for $\Delta m_s$ which is a 95\% CL limit boundary excluding the left region. The combined constraints on ($\rho,\eta$) are indicated by the 68 and 95\% CL contours, and translate to $\sin(2\beta) = 0.75 \pm 0.09$~\protect\cite{Mele}. Superimposed is the CDF $\sin(2\beta)$ measurement with two of the four solutions for $\beta$ shown (long dashed lines). The $1\sigma$ allowed range (light shaded region) is shown for the smaller $\beta$ solution. Right: Summary of $\sin(2\beta)$ measurements~\protect\cite{OPALcp,CDFcp}, Run II projections, and an indirect ``$1\sigma$'' range reported at this Conference~\protect\cite{Ali}. \label{fig:summary} } \end{center} \end{figure} In addition to the baseline upgrades, both experiments are aggressively pursuing further improvements. For example, CDF is working towards an additional Si-layer to improve vertex resolution and a Time-of-Flight system, which may push $x_s$ out to $\sim\!60$; and D{\O} is looking at a displaced-track trigger to greatly enhance $B$ triggering. We close with the CKM model unscathed, but look forward to an exciting future where, perhaps, some of the mystery surrounding $CP$ violation may be unveiled. \section*{ \Large Acknowledgments} I would like to thank my fellow collaborators, and colleagues across the ring, for the pleasure of representing them and their work. Assistance with this presentation from T.~Miao, M.~Paulini, C.~Paus, F.~Stichelbaut, and J.~Tseng is appreciated.
\section{Introduction}Most widely accepted emission models assume that pulsar radiation is emitted over a (hollow) cone centered around the magnetic dipole axis. The observed emission is generally highly linearly polarized with a systematic rotation of the position angle across the pulse profile. This behaviour, following Radhakrishnan \& Cooke (1969), is interpreted in terms of the radiation being along the cone of the dipolar open field-lines emerging from the polar cap, and the plane of the linear polarization is that containing the field line associated with the emission received at a given instant. During each rotation of the star, the emission beam crosses the observers line-of-sight resulting in a pulse of emission. The observed pulse profile thus corresponds to a {\it thin} cut across the beam at a fixed rotational latitude. The information on the beam shape as a function of latitude, although generally not measurable directly, may be forthcoming from observations at widely separated frequencies, as emission at different frequencies is believed to originate at different heights from the star leading to changes in beam size. For this, the dependence of the radiation frequency on the height, the so called {\it radius-to-frequency mapping}, should be known a priori. Alternatively, it is possible to use the data on an ensemble of pulsars sampling a range of impact parameters. However, it is important that all the pulsars in the sample form a homogeneous set in terms of the profile types etc. Several attempts to model the pulsar beam have used the latter approach. Based on their study, Narayan and Vivekanand (1983) concluded that the beam is elongated in the latitude. Lyne \& Manchester (1988), on the other hand, have argued that the beam is essentially circular (see also Gil \& Han 1996, Arendt \& Eilek 1998). Based on the dipole geometry of the cone of open field-lines, Biggs (1990) found that the beam shape is a function of the angle ($\alpha$) between the rotation and the magnetic axes. The reasons that all these analyses predict different results could be manifold. For example, Narayan \& Vivekanand used a data set consisting of only 16 pulsars and assessed the beam axial ratio on the basis of the total change in the position angle of the linear polarization across the pulse profile. Apart from poor statistics, their analysis suffered from the large uncertainties in the polarization measurements available then. Lyne \& Manchester (1988) used a much larger data set in comparison and examined the distribution of normalized impact parameter $\beta_{n}\equiv\beta_{90}/\rho_{90}$, where $\beta_{90}$ \& $\rho_{90}$ are the impact angle and the beam radius computed for $\alpha = 90^{\circ}$. Based on their observation that the distribution of $\beta_{n}$ is `essentially uniform', they concluded that the beams are circular in shape. The apparent deficit at large $\beta_{n}$ is attributed to a luminosity bias. It is worth noting that the deficit is seen despite the fact that $\beta_{n}$ overestimates the {\it true} $\beta/\rho$ (given that they disregarded the sign of $\beta$), this is particularly so at large $\beta$ values. Biggs (1990) used the same data set as well as the $\beta_{n}$ distribution as used by Lyne and Manchester (1988), but drew attention to a `peak' in the distribution at low $\beta_{n}$. The shapes of the polar cap defined by the region of open field lines, as derived by Biggs, show that the beam is circular for an aligned rotator, but undergoes compression along the latitudinal direction with increasing inclination $\alpha$. In this paper, we address this question within the basic framework advanced by Rankin (1993a) which, at the least, is qualitatively different from that of Lyne \& Manchester (1988). The classification scheme (Rankin, 1983a), based on the phenomenology of pulse profiles, polarization and other fluctuation properties etc., provides a sound basis for explicit distinction between the core and the conal components, with each of them following a predictable geometry (see also Oster \& Sieber 1976; Gil \& Krawczyk 1996 for {\it conal beams}). Lyne \& Manchester (1988), on the other hand, prefer to interpret the observed variety in pulse shape and other properties as a result of patchy illumination, rather than any particular pattern within the radiation cone. The observed differences in the properties of pulse components are then to be understood as gradual changes as a function of the distance from the center of the basic emission cone. Their analysis thus naturally disregards the possible existance of conal features. Assuming the possibly confined `conal-component' geometry and by accounting for all the relevant geometrical effects, we re-examine the shape of pulsar beams and their frequency dependence. Recently published multifrequency polarization data, at six frequencies in the range between 234-1642 MHz (Gould \& Lyne, 1998), has made this investigation possible. \section{Data set} For the present investigation requiring reliable estimates of $\alpha$ \& $\beta$, we use the data set comprised of only those pulsars whose pulse profiles are identified as `triple' ($\bf{T}$) or `multiple' ($\bf{M}$), as classified by Rankin (1993a, 1993b). The reason for the choice is that the $\bf{T}$ and $\bf{M}$ pulsars show a core component in addition to the conal components, so that a reliable estimation of the angle ($\alpha$) between the rotation axis and the magnetic axis is possible, using Rankin's (1990) method. In this method, the ratio of the observed core-width to the limiting width ($2.45^{\circ} P^{-0.5}$) is interpreted as the geometric factor $1/\sin(\alpha)$, providing by far the most reliable estimates of $\alpha$. For the conal doubles and conal singles, devoid of any core component, the estimates of $\alpha$ are less reliable. The core singles are naturally excluded from this analysis of the conal emission geometry. For each pulsar in our selected sample, we define the conal width as the separation between the peaks of the outermost conal components. It is important to note that the nominally `central' core component, which is argued to originate closer to the stellar surface, may not necessarily be along the cone axis. Such a possibility is clearly reflected in many pulse profiles where the core component is displaced from the `center' definable from the conal components. Hence, the location of the core component is disregarded in our estimation of the conal separation. Columns 1 and 2 of table 1 list the name and profile type of these pulsars. Columns 3 to 8 list the calculated widths of the pulsars at frequencies 234, 408, 610, 925, 1400, and 1642 MHz respectively. Column 9 gives the pulsar period in seconds. Columns 10 and 11 list the $\alpha$ and $\beta$ values of the pulsars taken from Rankin (1993b). \begin{figure} \epsfig{file=8545f1.ps,width=7.5cm,height=7cm,clip=true,angle=270} \caption[]{Schematic representation showing the geometry of the pulsar emission region.} \label{fig:fig1} \end{figure} Rankin (1990) has estimated the inclination angle $\alpha$ using the relation, $\sin(\alpha) = 2.45^{\circ} P^{-0.5}/W_{\rm{core}}$, where $W_{\rm{core}}$ is the half-power width of the core component (at a reference frequency 1 GHz) and the period $P$ is in seconds. The impact angle $\beta$ has been estimated based on the rotating vector model of Radhakrishnan \& Cooke (1969), using the relation $\sin(\beta) =(d\chi/d\phi)_{\rm{max}} /\sin(\alpha)$, where $(d\chi/d\phi)_{\rm{max}}$ is the maximum rate of change of the polarization angle $\chi$ with respect to the longitude $\phi$. In the following analysis, we treat the different frequency measurements on a given pulsar as `independent' inputs much the same way as the data on different pulsars, since the pulsar beam size is expected to evolve with frequency. Thus, at different frequencies one obtains independent cuts (at different $\beta/\rho$) across the beam, though $\beta$ remains constant for a given pulsar. This increases the number of independent constraints by a usefully large factor. In fact, we would like to contrast this approach with the one where, for each pulsar, one obtains a best fit frequency dependence of the observed widths and then uses the data to obtain the width at a chosen reference frequency. The latter approach fails to take into account the dependence of the observed widths on $\beta/\rho$ that is inherent for any non-rectangular shape of the beam. \begin{figure} \epsfig{file=8545f2.ps,height=6cm,width=4.2cm,clip=true,angle=270} \caption[]{ Schematic representation of an elliptic pulsar beam of axial ratio $R$ with the longitudinal and the latitudinal axis as $b$ and $Rb$ respectively. $\delta r$ is the width of the emission beam cone. See text for discussion on the connection between the `gap-angle' $\theta_g$ and $\delta r/r$.} \label{fig:fig2} \end{figure} \section{A direct test for the shape of beams} The Fig~\ref{fig:fig1} is a schematic diagram illustrating the geometry of pulsar emission cone. The emission cone, with half-opening angle $\rho_{\nu}$, sweeps across the observers line-of-sight with an impact parameter (distance of closest approach to the magnetic axis) $\beta$. The spherical triangle PQS (refer to Fig.~\ref{fig:fig1}) relates the angles $\alpha$, $\beta$ and the profile half-width $\phi_{\nu}$ to the beam radius $\rho_{\nu}$ by the following relation (Gil, Gronkowski \& Rudnicki 1984), \begin{equation} \sin^{2}(\rho_{\nu}/2)= \sin^{2}(\phi_{\nu}/2) \sin(\alpha) \sin(\alpha + \beta) + \sin^{2}(\beta/2) \label{eq:modeq} \end{equation} The subscript $\nu$ in $\rho_{\nu}$ and $\phi_{\nu}$ denotes that these quantities depend on frequency $\nu$. This equation assumes that the cone is circular, in which case $\rho_{\nu}$ becomes independent of $\beta$. \begin{figure} \epsfig{file=8545f3.ps,width=6 cm,height=6 cm,clip=true,angle=270} \caption[]{The above curves illustrate the normalized variation of r with $\theta$ (refer to figure~\ref{fig:fig2}) with three different values of $R$.} \label{fig:fig3} \end{figure} In reality, the beam may not be circular, but rather elliptical with, say, $R$ the axial ratio and b the longitudinal semi-axis of the ellipse as shown in Fig.~\ref{fig:fig2}. It is easy to see that the length of the radius vector r depends on the angle $\theta$ (with the longitudinal axis) when $R$ is not equal to 1. The variation of r as a function of $\theta$ for three different $R$ values (namely 1, 1.5 and 0.5) are shown as examples in Fig.~\ref{fig:fig3}. The $\rho_{\nu}$, determined assuming that the cone shape is circular (as in Rankin 1993b) is indeed a measure of the radius vector r, once the period and frequency dependences are corrected for. Such data on (r,$\theta$) spanning a wide enough range in $\theta$ can therefore be examined to seek a consistent value of the axial-ratio $R$. However, if $R$ is a function of $\alpha$, as suggested by Biggs (1990), then the (r,$\theta$) samples would show a spread bounded by the curves corresponding to the maximum and minimum values of $R$. Such an examination of the present data suggests a spread below the line for $R=1$, indicating that the beam deviates from circularity and that the spread could be due to the $\alpha$ dependence of $R$. However, this deviation from circularity is not very significant. We discuss this in detail later in section 5. We have also examined the $\rho_{\nu}$ values obtained by Rankin (1993b) through such a test. However, no significant deviation from circular beams was evident. We became aware of a similar study by C.-I. Bj$\ddot{o}$rnsson (1998), also with a similar conclusion. We note that the only difference between our estimates of $\rho_{\nu}$ and those of Rankin is in the definition of the conal widths. Rankin defines the width as the distance between the outer half-power points (rather than the peaks) of the two conal outriders, and the widths were then `interpolated' to a reference frequency of 1 GHz. Such estimates are prone to errors due to mode changes, differing component shapes etc., and to the effects of dispersion \& scattering (some of which she attempted to accommodate). We measure the widths as the peak-to-peak separations of the outer conal components, which are less sensitive to the sources of error mentioned above. We have also confirmed (in the PSRs 0301+19, 0525+21, 0751+32, 1133+16, 1737+13, 2122+13 and 2210+29 using the data from Blaskiewicz et al. 1991) that the `peaks' of the conal components are symmetrically placed with respect to the ``zero-longitude" (associated with the maximum rate of change of the position angle), which is not always true for the outer half-power points. \begin{table*} \begin{flushleft} \caption[]{ The table lists the pulsar name and the widths measured at 6 different frequencies from the observations of Gould \& Lyne (1998). In several cases the widths could not be estimated due either to poor quality profiles or to absence of data. The $\alpha$, $\beta$ values are taken from Rankin (1990, 1993b). $LM$ indicates that the $\beta$ value (for PSR 0656+14 and 1914+09) is taken from Lyne \& Manchester (1988).} \begin{tabular}{lccccccccccccc } \noalign{\smallskip} \hline {\bf Pulsar}&{\bf Profile}&\multicolumn{6}{c}{\bf Width in deg}&{\bf Period}&{\bf $\alpha$}& {\bf $\beta$}\\ \cline{3-8} {\bf Bname}&{\bf Class}&{\bf $W_{234}$}&{\bf $W_{408}$}&{\bf $W_{610}$}&{\bf $W_{925}$}&{\bf $W_{1400}$}& {\bf $W_{1642}$}&{\bf (sec)}&{\bf (deg)}& {\bf (deg)}\\ \noalign{\smallskip} \hline 0329+54& \bf{T}& 25.4 & 23.3& 21.8 & 21.8& 21.2& 20.7& 0.714518& 30& 2.1\\ 0450-18& \bf{T}& 16.6 & 14.5& 13.5 & 12.9& 12.4& 11.9& 0.548937& 24& 4\\ 0450+55& \bf{T}& 27.3 & 20.7& 20.7 & 24.6& 22.0& 22.0& 0.340729& 32& 3.3\\ 0656+14& \bf{T}& 27.9 & 21.7& 17.8 & 25.5& 20.1& 17.8& 0.384885& 30& 8.2 (LM)\\ 0919+06& \bf{T}&18.1 & 16.5& 14.6& 11.5& 10& 8.4 & 0.430619& 48& 4.8 \\ ~~~~~~~& ~~~~~~ &~~~~~ & ~~~~ ~~ & ~~~ ~~ & ~~ ~~~& ~~ ~~ & ~~~~~ &~~ & ~~~& \\ 1508+55& \bf{T}& - & 12.0& 8.57 & 11.6& 10.9& 10.5& 0.739681& 45& -2.7\\ 1541+09& \bf{T}& 126.5& 107.8& 105.4& 96.0& 91.4& 84.3& 0.748448& 5& 0.0 \\ 1738-08& \bf{T}&- & 14.6& 13.7& 13.6& 12.6& 12.1& 2.043082& 26& 1.7 \\ 1818-04& \bf{T}&- & 10.7& 8.2& 9.20& 8.8 & 8.5 & 0.598072& 65& 3.5 \\ 1821+05& \bf{T}& 36.2 & 32.1& 29.4 & 29.4& 26.6& 26.6& 0.752906& 32& 1.7 \\ & & & & & & & & & & \\ 1911+13& \bf{T}& - & 12.3& 10.7& 12.0& 11.6& 11.0& 0.521472& 52& 1.9 \\ 1914+09& \bf{T}& - & 10.9& 12.6& 8.9& 8.5& 8.1 & 0.270254& 52& 7.3 (LM) \\ 1917+00& \bf{T}& - & 8.3 & 8.1& 8.0& 7.2& 6.7& 1.272255& 81& 1.3 \\ 1918+19& \bf{T}&- & 49.1& 42.7& 41.3& 41.3& 38.7& 0.821034& 12& -4.6 \\ 1919+14& \bf{T}& - & 22.3& 20.7& 18.7& 19.7& 17.1& 0.618179& 26& -6.4 \\ & & & & & & & & & & \\ 1919+21& \bf{T}&- & 7.17& 6.7& 8.2& 7.6& 7.4 & 1.337301& 45& -3.7 \\ 1920+21& \bf{T}& - & 15.1& 10.1& 14.4& 14.0& 13.2& 1.077919& 44& 1.1 \\ 1944+17& \bf{T}&- & 25.2& 23.3& 33.0& 33.0& 31.0& 0.440618& 19& 6.1 \\ 2045-16& \bf{T}& - & 12.9& 12.3& 11.6& 11.0& 10.7& 1.961566& 36& -1.1 \\ 2111+46& \bf{T}& 69.8 & 63.3& 59.4& 55.6& 53.0& 49.1& 1.014684& 9& 1.4 \\ & & & & & & & & & & \\ 2224+65& \bf{T}&39.9 & 35.0& 31.1& 31.1& 31.1& 31.1& 0.682537& 16& 3.4 \\ 2319+60& \bf{T}&21.8 & 18.7& 17.1& 15.0& 13.5& 13.5& 2.256487& 18& 2.2 \\ 1804-08& \bf{M/T}&- & 28.5& 12.9 & 16.2& 15.5& 14.2& 0.163727& 63& 5.1 \\ 1910+20& \bf{M/T}&- & 12.8& 11.5& 11.2& 10.8& - & 2.232963& 29& 1.5 \\ 1952+29& \bf{M/T}&- & 22.7& 21.6& 22.2& 21.0& 19.2& 0.426676& 30& -7.2 \\ & & & & & & & & & & \\ 2020+28& \bf{M/T}&12.9& 10.9& 10.1& 10.1& 9.74& 9.3 & 0.343401& 72& 3.6 \\ 0138+59& \bf{M}&25.8 & 20 & 23.2& 20.6& 18.7& 17.4& 1.222948& 20& 2.2 \\ 0402+61& \bf{M}&14.2 & 14.6& 10.7& 10.3& 10& 9.6 & 0.594573& 83& 2.2 \\ 0523+11& \bf{M}&- & 12.4& 10.8& 12.0& 11.6& 10.8& 0.354437& 78& 5.9 \\ 0621-04& \bf{M}&18.5 & 21.2& 18.4& 18.0& 17.5& - & 1.039076& 32& 0.0 \\ & & & & & & & & & & \\ 1039-19& \bf{M}&15.4 & - & 11.5& 10.7& 10& 9.6 & 1.386368& 31& 1.7 \\ 1237+25& \bf{M}&10.0 & 10.3& 10.0& 9.3& 9.0& 10.0& 1.382449& 53& 0.0 \\ 1737+13& \bf{M}&- & 17.4& 17.0& 16.1& 15.2& 13.8& 0.803049& 41& 1.9 \\ 1831-04& \bf{M}&95.3 & 97.6& 95.3& 96.2& 93.0& 93.0& 0.290106& 10& 2.0 \\ 1857-26& \bf{M}&- & 32.5& 29.4& 26.3& 25.5& 24.8& 0.612209& 25& 2.2 \\ & & & & & & & & & & \\ 1905+39& \bf{M}&- & 15.1& 13.7& 13.1& 12.6& 11.7& 1.235757& 33& 2.1 \\ 2003-08& \bf{M}& 55.6 & 40.0& 38.7& 33.6& 32.3& 31.0& 0.580871& 13& 3.3 \\ \noalign{\smallskip} \hline \end{tabular} \end{flushleft} \label{table:tab1} \end{table*} \normalsize \section{The model of the pulsar beam} We model the pulsar beam shape as elliptical in general and express it analytically as, \begin{equation} \frac{\sin^{2}(\phi_{\nu}/2) \sin(\alpha) \sin(\alpha + \beta)}{\sin^{2}(\rho_{\nu}/2)} + \frac{\sin^{2}(\beta/2)}{\sin^{2}(R \rho_{\nu}/2)}=1 \label{eq:mod} \end{equation} \noindent While $\alpha$, $\beta$ and $\phi_{\nu}$ can be estimated, directly or indirectly, from observations, $R$ and $\rho_{\nu}$ are the two parameters which in turn define the beam shape and size--- and the available data set of $\bf{T}$ and $\bf{M}$-profiles is expected to sample most of the $\mid \beta/\rho_{\nu}\mid$ range (0--1) with reasonable uniformity. The implicit assumption in this statistical approach is that a common description for $R$ \& $\rho_{\nu}$ is valid for all pulsars. The common description should, however, account for relevant dependences on quantities, such as frequency, period, $\alpha$, etc. properly. \subsection{Frequency dependence of $\rho_{\nu}$} The radio emission at different frequencies is expected to originate at different altitudes above the stellar surface, with the higher frequency radiation associated with regions of lower altitude. This phenomenon known as {\em radius-to-frequency mapping}, finds overwhelming support from observations. Thorsett (1991) has suggested an empirical relation for the observed pulse width as a function of frequency, which seems to provide adequate description of the observed behaviour. We adopt a similar relation for the frequency evolution of the beam radius $\rho_{\nu}$ as follows \begin{equation} \rho_{\nu} = \hat{\rho}(1 + K \nu^{-\zeta}), \label{eq:rho} \end{equation} \noindent where $\hat{\rho}$ is the value of $\rho_{\nu}$ at infinite frequency, $\zeta$ the spectral index, and $K$ a constant. Note that both $\zeta$ \& $K$ are expected to have positive values, so that the minimum value of $\rho_{\nu}$ is $\hat{\rho}$, which should correspond to the angular size of the polar cap. \subsection{Period dependence on $\rho_{\nu}$} Rankin (1993a) has demonstrated (see also Gil, Kijak \& Seiradakis 1993; Kramer et al. 1994) that the beam radius $\hat{\rho}$ varies as $P^{-0.5}$ (where $P$ is the period of the pulsar), a result which is in excellent agreement with that expected from a dipole geometry (Gil 1981). Eq~\ref{eq:rho} thus takes the form \begin{equation} \rho_{\nu} = \rho_{\circ} (1 + K \nu^{-\zeta}) P^{-0.5}, \label{eq:rho1} \end{equation} \noindent where $\rho_{\circ}$ is the minimum beam radius for $P = 1$ sec. \subsection{Functional dependence of $R$ on $\alpha$} Biggs (1990) has suggested that $R$ should be a function of $\alpha$, such that the beam shape is circular for $\alpha=0$ and is increasingly compressed in the latitudinal direction as $\alpha$ increases to $90^{\circ}$. We therefore model the functional dependence of $R$ on $\alpha$ as $R = R_{\circ}\tau$, where $R_{\circ}$ is the axial ratio of the beam at $\alpha = 0$, and $\tau$ is a function of $\alpha$. According to Biggs (1990), $R_{\circ} = 1$ and $\tau$ is given by \begin{equation} \tau(\alpha)\;=\;1 - K_{1}\times10^{-4} \alpha - K_{2}\times10^{-5}\alpha^{2}, \label{eq:biggs} \end{equation} \noindent where $K_{1}$, $K_{2}$ are constants and $\alpha$ is in degrees. Biggs finds that $K_{1}$ and $K_{2}$ are 3.3 and 4.4, respectively. We, however, treat $K_{1,2}$ as free parameters in our model. \subsection{The number of hollow cones}Based on the study of conal components, Rankin (1993a) has argued for two nested hollow cones of emission-- namely, the outer and the inner cone. Assuming the beams to be circular in shape, opening half angles of the two cones at 1 GHz were found to be $4.3^{\circ}$ and $5.7^{\circ}$, respectively. During our preliminary examination of the present sample, we noticed a need to allow for three cones of emission. To incorporate this feature in our model, we introduce two ratios, $r1 < 1$ and $r2 > 1$, to define the size scaling of the inner-most and the outer-most cone, respectively, with reference to a `middle' cone, for which the detailed shape is defined. Using the model here defined, we need to solve for $R_{\circ}$, $\zeta$, $K$, $\rho_{\circ}$, $K_{1}$, $K_{2}$, $r1$ and $r2$ in this three-conal-ring model. The parameter set thus represents an `average' description of the beam. \begin{table*} \caption[]{The best-fit model parameters for the shape of conal beams. The error bars correspond to a 1$\sigma$ uncertainty.} \begin{flushleft} \begin{tabular}{cccccccc} \multicolumn{8}{c}{\bf Model parameters}\\ \hline $ R_{\circ}$ & $\rho_{\circ}$ (deg)& $K$ & $\zeta$ & $r1$ & $r2$ & $K_{1}$ ($\rm{deg^{-1}}$) & $K_{2}$ ($\rm{deg^{-2}}$)\\ \cline{1-8} $0.91 \pm^{0.2}_{0.1}$ &${4.8\pm 0.3}$ & $66\pm 10$ & $1 \pm 0.1$ & $0.8\pm 0.03$ & $1.3\pm 0.03$ & ${7.2\pm 0.2}$ & ${4.4\pm 0.3}$ \\ \hline \end{tabular} \end{flushleft} \end{table*} \section{Results and Discussion}An optimized grid search was performed for suitable ranges of the parameter values and in fine enough steps. For $\zeta$, the search range allowed for both +ve and -ve values. By definition, $r_{1}\leq 1$ and $r_{2}\geq 1$. The best fit was obtained by minimizing the standard deviation $\sigma_{\circ}$ defined by \begin{equation} \sigma_{\circ}\;=\;\sqrt{\frac{\sum_{i=1}^{n} D_{i}^{2}}{N_{dof}}} \times \frac{180^{\circ}}{\pi}, \label{sigma} \end{equation} \noindent where $D_{i}$ is the deviation of the $i^{th}$ data point from the nearest conal ring in the model and $N_{dof}$ denotes the degrees of freedom. \begin{figure} \epsfig{file=8545f4.ps,width=7cm,height=7cm,clip=true,angle=270} \caption[]{Distribution of the (x,y) locations of the conal components on a common scale. The three solid lines indicate the three emission cones in the quadrant shown.The circles with crosses refers to pulsars with $\alpha$ values less than $45^{\circ}$ and the filled circles with $\alpha$ greater than $45^{\circ}$.} \label{fig:fig4} \end{figure} The factor $180/\pi$ gives $\sigma_{\circ}$ in units of degrees under the small-angle approximation. Table 2 lists the parameter values which correspond to the best fit for the entire sample set. With these values, the eq.~\ref{eq:rho1} can now be rewritten as \begin{equation} \rho_{\nu} = 4.8^{\circ} (1 + 66\; \nu^{-1}_{\rm MHz}) P^{-0.5}, \label{eq:rho2} \end{equation} \noindent where $\rho_{\nu}$ is in degrees. This average description for the `middle' cone applies also to the other two cones when $\rho_{\nu}$ is scaled by the ratio $r1=0.8$ or $r2=1.3$ (for the inner and the outermost, respectively). Fig~\ref{fig:fig4} shows the data (plotted to a common scale) for one quadrant of the beam and the three solid curves corresponding to the best fit cones. The points in the figure, though corresponding to different pulsars and frequencies, are translated to a common reference scale appropriate for $P=1$ sec, $\alpha = 0$ and $\nu =\infty $. \begin{figure} \epsfig{file=8545f5.ps,width=7cm,height=7cm,clip=true,angle=270} \caption[]{Histogram of the distribution of effective $\frac{\beta}{\rho}$.} \label{fig:fig5} \end{figure} We have assumed the period dependence of $\rho_{\nu}$ as $P^{-0.5}$, whereas Lyne and Manchester (1988) found a dependence of $P^{-\frac{1}{3}}$. We have examined the latter possibility and found that the difference in the standard deviation is at the level of 2.5-3 $\sigma$ and we cannot rule out the $P^{-\frac{1}{3}}$ law with confidence. We have also checked for the dependence of $R$ on $\alpha$ by using 3 sub-sets, each of range $30^{\circ}$ in $\alpha$. The best fit values for $R$ in the different $\alpha$ segments are $1\pm^{0.4}_{0.2}$, $0.8\pm^{0.4}_{0.2}$ \& $0.5\pm^{0.4}_{0.2}$ for $\alpha$ ranges $0^{\circ}-30^{\circ}$, $30^{\circ}-60^{\circ}$ \& $60^{\circ}-90^{\circ}$, respectively. This dependence of $R_{\circ}$ on $\alpha$, even if it were significant, is quite consistent with our values of $K1$, $K2$ (Table 2) as well as with the results of Biggs (1990). However, given the uncertainties in the $R$ estimates for the three ranges, it is not possible presently to rule out a dependence of $R$ on $\alpha$. Indeed, this part of the goodness-of-fit is negligible, $\sigma_{\circ}$ (the standard deviation) is $0.18^{\circ}$ when $K1$ and $K2 \neq 0$ and $0.2^{\circ}$ when $K1, K2 = 0$. Earlier Narayan \& Vivekanand (1983) had argued that $R$ is a function of the pulsar period. To assess this claim, our sample was divided into three period ranges and the corresponding $R$ estimates compared. However, no period dependence was evident and it was possible to rule out such a dependence with high confidence. {\bf The number and thickness of conal rings:} As already noted and can be seen in Figure~\ref{fig:fig4}, we do see evidence for a possible cone outside the two cones discussed by Rankin (1993a). Also, presence of a `further inner' cone has been suggested by Rankin \& Rathnasree (1997) in the case of PSR 1929+10. The pulsars suggestive of this outer cone (refer Figure~\ref{fig:fig4}) are PSRs 0656+14, 1821+05, 1944+17 and 1952+29 (at frequencies 234 MHz and higher). We have examined the possibility that these cases really belong to the central-cone, but are well outside of it due to an error in the assumed values of $\alpha$. We rule out the possibility as the implied error in $\alpha$ turns out to be too high to be likely. It is important to point out that a noisy sample like the present one would appear increasingly consistent, judging by the best-fit criterion, with models that include more cones. The question, therefore is whether we can constrain the number of cones by some independent method. In this context, we wish to discuss {\it the noticeable deficit of points at high $\beta/\rho_{\circ}$}. Since the deficit reflects the absence of conal singles and conal doubles in our data set, the size of the related `gap' at large $\theta$ values, can be used to estimate the possible thickness of the conal rings. The absence of points at $\theta\raisebox{-0.3ex}{\mbox{$\stackrel{>}{_\sim} \,$}} 60^{\circ}$ (Figure~\ref{fig:fig4}) suggests that the conal rings are rather thin, since a radial thickness $\delta r$ comparable to the ring radius would imply a wider gap in $\theta$. To quantify this, we write the following relation, \begin{equation} \delta r= 2r \frac{(1-\sin\theta_{g})}{(1+\sin{\theta_{g}})}, \label{eq:deltar} \end{equation} \noindent where $\theta_{g}$ is the $\theta$ at the start of the gap (as illustrated in Fig.~\ref{fig:fig2}). With $\theta_{g} \sim 60^{\circ}$, $\delta r/r$ would be about 20\%. The presence of more than one distinguishable peak in the distribution of beam radii (shown in the bottom panel of Fig.~\ref{fig:fig4}) clearly indicates that the conal separation is larger than the cone width. This combined with our cone-width estimate suggests the number of cones is 3 (for the present range of radii), providing an independent support for our model. This picture is consistent with the estimates by Gil \& Krawczyk (1997) and Gil \& Cheng (1999). {\bf Component separation vs. frequency:} It is interesting to note that for certain pulsars the cone associated with the emission seems to change with frequency. For example, the conal emission in PSR 1920+21 appears to have `switched' at 610 MHz to the innermost cone while being associated with the central cone at other frequencies. Rankin (1983b), in a comprehensive study of the dependence of component separation with frequency, invokes deep `absorption' features to explain the apparent anomalous reduction in the component separation in certain frequency ranges. We suggest that such anomalous reduction in the separations could be due to switching of the emission to an inner cone at some frequencies. Observations at finely spaced frequencies in the relevant ranges would be helpful to study this effect in detail. The other pulsars which show similar trends are PSRs 1804-08, 2003-08, 1944+17 and 1831-04. It should be noted that such switching is possibly reflected, also, in mode changes. {\bf The deficit at low $\beta/\rho_{\circ}$:} The absence of points near $\beta = 0$ is clearly noticeable in Fig.~\ref{fig:fig4}. Such a `gap' is also apparent in the distribution of $\beta/\rho_{\circ}$ plotted in Fig.~\ref{fig:fig5}. The gap was already noted by Lyne \& Manchester (1988). They argued that it arises because the rapid position-angle swings (expected at small $\beta$'s) are difficult to resolve due to intrinsic or instrumental smearing, leading to underestimation of the sweep-rates. With the improved quality of data now available, the intrinsic smearing is likely to be the dominant cause for this circumstance. There are a number of clear instances among the general population of pulsars where the polarization angle traverse near the central core component is distorted. PSR 1237+25 provides an extreme examples of such distortion, and Ramachandran \& Deshpande (1997) report promising initial efforts to model its polarization-angle track as distorted by by a low-$\gamma$ core-beam. Another possibility for the low-$\beta/\rho_{\circ}$ gap is that it could simply be a selection effect caused by less intense emission in the cone center than at intermediate traverses. If so, the low frequency turn-overs in the energy spectra of pulsars may at least be partly due to this, since at lower radio frequencies the $\beta/\rho_{\circ}$ is relatively smaller. {\bf The sources of uncertainties in the present analysis:} The standard deviation $\sigma_{\circ}$ corresponding to the best-fit model amounts to about 15\% of the conal radius. This fractional deviation (comparable to the thickness of the cone) is too large to allow any more detailed description of the beam shape (such as dependence on $\alpha$, for example). We find it useful to assess and quantify the sources of error, partly to help possible refinement for future investigations. The three data inputs to our analysis are $\alpha, \beta$ and $\phi_{\nu}$, while the basic observables are the maximum polarization-angle sweep rate and core width, apart from the conal separation measured. It is easy to see that the errors in the core-widths will affect directly both $\alpha$ and $\beta$ estimates. Over the range of $\theta$s spanned by the present data set the errors in $\alpha$ are likely to dominate, since the x \& y (in figure~\ref{fig:fig4}) are almost linearly proportional to $\sin(\alpha)$. Hence, the fractional deviation may be nearly equal to (or define the upper limit of) the fractional error in $\sin(\alpha)$ and therefore in the core-width estimates. Rankin (1990, 1993b) notes that in several cases the apparent core-widths might suffer from `absorption' and the widths might be underestimated if the effect is not properly accounted for. Also, in some cases, the widths were extrapolated to a reference frequency of 1 GHz using a $\nu^{-0.25}$ dependence. There have been several suggestions regarding the `appropriate' frequency dependence which would give significantly different answers when used for width extrapolation. For example, if our best-fit dependence for conal width is used for the core-width extrapolation, the values would differ from Rankin's estimates (through extrapolation) by as much as 15\%, enough to explain the present deviation in some cases. Another possible source of error is the uncertainty in the sign of $\beta$ (important only for the $\sin(\alpha + \beta)$ term in equation~\ref{eq:mod} and hence for small $\alpha$). As Rankin points out, it is difficult to determine the sign unambiguously in most cases and hence the information is only available for a handful of pulsars. {\bf Evidence in favour of `conal' emission:} The significant implication of the gap at $\theta\raisebox{-0.3ex}{\mbox{$\stackrel{>}{_\sim} \,$}} 60^{\circ}$ (referred to earlier) deserves further discussion. If the `conal' components were results of a merely patchy (random) illumination across the beam area, (as argued by Lyne \& Manchester, 1988), then such a gap should not exist. If a single thick hollow cone were to be responsible for the conal components, a gap (corresponding to the conal-single types) would still be apparent but then it should be above a cut-off y value (refer figure 4) and not in a angular sector like that observed. On the other hand, if indeed the conal emission exists in the form of nested cones (as distinct from the core emission), then the shape of the gap is a natural consequence of our not including conal-single profiles in this analysis. This gap, therefore, should be treated as an important evidence for a pulsar beam form comprised, in general, of nested cones of emission. \section{Summary} Using the multifrequency pulse profiles of a large number of conal-triple and multiple pulsars we modelled the pulsar beam shape in an improved way. Our analysis benefits from the different frequency measurements being treated as independent samples, thus increasing the number of independent constrains. The main results are summarized below. 1) Our profile sample is consistent with a beam shape that is a function of $\alpha$, circular at $\alpha =0$ and increasingly compressed in the latitudinal direction as ${\alpha}$ increases, as suggested by Biggs (1990). However, the data is equally consistent with the possibility that the beam is circular for all values of ${\alpha}$. 2) We identify three nested cones of emission based on a normalized distribution of outer components. The observed gap ($\theta\raisebox{-0.3ex}{\mbox{$\stackrel{>}{_\sim} \,$}} 60^{\circ}$) in the distribution independently suggests three cones in the form of annular rings whose widths are typically about 20\% of the cone radii. We consider this circumstance as an important evidence for the nested-cone structure. Any further significant progress in such modelling would necessarily need refined estimates of the observables, particularly the core-widths. \section*{Acknowledgement} We thank V. Radhakrishnan, Rajaram Nityananda and Joanna Rankin for fruitful discussions and for several suggestions that have helped in improving the manuscript. We acknowledge Ashish Asgekar, D. Bhattacharya and R. Ramachandran for useful discussions and thank our referee, J. A. Gil, for critical comments and suggestions.
\section{Introduction} After reviewing the indirect information on the Higgs mass based on precise electroweak measurements performed at LEP1, SLD and at the TEVATRON, I will discuss the mechanisms of Higgs production and decay and the strategy adopted to search for the neutral Higgs boson (in the SM and in the MSSM) at LEP2~\cite{reviews}. I will summarise the results based on the analysis of approximately 170~$\pb^{-1}$ collected by each LEP experiment at $\roots=189$~\Gev updated to the more recent Winter Conferences numbers~\cite{felcini}. In the end I will briefly discuss the prospects for Higgs discovery at LEP2. \section{Higgs mass from precision electroweak measurements and from theoretical arguments} The aim of precision electroweak tests is to prove the SM beyond the tree level plus pure QED and QCD corrections and to derive constraints on its fundamental parameters. Through loop corrections, the SM predictions for the electroweak observables depend on the top mass via terms of order $\rm{G_F}/\Mt^2$ and on the Higgs mass via logarithmic terms. Therefore from a comparison of the theoretical predictions~\cite{pre_calc}, computed to a sufficient precision to match the experimental capabilities and the data for the numerous observables which have been measured, the consistency of the theory is checked and constraints on $\MH$ are placed, once the measurement of $\Mt$ from the TEVATRON is input. The present 95\%~C.L. upper limit on the Higgs mass in the SM is~\cite{mh_smfits,felcini} \begin{equation} \label{mh_up} \MH< 220\,\Gevcc\,, \end{equation} if one makes due allowance for unknown higher loop uncertainties in the analysis. The corresponding central value is still rather imprecise: \begin{equation} \MH= 71^{+75}_{-42}\pm5\,\Gevcc\,. \end{equation} The range given by Eq.\ref{mh_up} may be compared with the one derived from theoretical arguments~\cite{hambye}. It is well known that in the SM with only one Higgs doublet a lower limit on the Higgs mass $\MH$ can be derived from the requirement of vacuum stability. This limit is a function of the energy scale $\Lambda$ where the model breaks down and new physics appears. Similarly an upper bound on $\MH$ is obtained from the requirement that up to the scale $\Lambda$ no Landau pole appears. If, for example, the SM has to remain valid up to the scale $\Lambda\simeq{\rm M_{GUT}}$, then it is required that $135<\MH<180~\Gevcc$. In the MSSM two Higgs doublets are introduced, in order to give masses to the up-type quarks on the one hand and to the down-type quarks and charged leptons on the other. The Higgs particle spectrum therefore consists of five physical states: two CP-even neutral scalars (h,A), one CP-odd neutral pseudo-scalar (A) and a charged Higgs boson pair ($\rm{H}^{\pm}$). Of these, h and A could be detectable at LEP2~\cite{yellow}. In fact, at tree-level h is predicted to be lighter than the Z. However, radiative corrections to $\Mh$~\cite{ellis}, which are proportional to the fourth power of the top mass, shift the upper limit of $\Mh$ to approximately 135~\Gevcc, depending on the MSSM parameters. \section{Higgs production and decay} At LEP2, the dominant mechanism for producing the standard model Higgs boson is the so-called Higgs-strahlung process $\bha\to$~HZ~\cite{khoze,bjorken}, with smaller contributions from the WW and ZZ fusion processes leading to H$\nu_{\rm{e}}\bar{\nu}_{\rm{e}}$ and H$\bha$ final states, respectively. A sizeable cross section (few 0.1~pb) is obtained up to $\MH \sim \roots - \MZ$, so that an energy larger than 190~\Gev is needed to extend the search above $\MH \simeq \MZ$. For example the production cross section at $\roots=189$~GeV for $\MH=95$~\Gevcc is 0.18~pb, which for an integrated luminosity $\cal{L}$=170~$\pb^{-1}$/exp. gives 30 signal events per experiment. For the MSSM Higgs the main production mechanisms are the Higgs-strahlung process $\bha\to$~hZ, as for the SM Higgs, and the associated pair production $\bha\to$~hA~\cite{ha-prod}. The corresponding cross sections may be written in terms of the SM Higgs-strahlung cross section, $\sigma^{\rm{SM}}$, and of the cross section $\sigma^{\rm{SM}}_{\nunu}$ for the process $\rm{Z}^*\to\nunu$ as \begin{eqnarray} \label{Zh-hA} \sigma(\bha\to\rm{Zh}) = & \rm{sin}^2(\beta-\alpha)\,\sigma^{\rm{SM}} \\ \sigma(\bha\to\rm{hA}) \propto & \rm{cos}^2(\beta-\alpha)\,\sigma^{\rm{SM}}_{\nunu}. \nonumber \end{eqnarray} The parameter $\rm{tan}\beta$ gives the ratio of the vacuum expectation values of the two Higgs doublets and $\alpha$ is a mixing angle in the CP-even sector. The Higgs-strahlung hZ process occurs at large $\rm{sin}^2(\beta-\alpha)$, i.e., at small $\rm{tan}\beta$. Conversely, at small $\rm{sin}^2(\beta-\alpha)$, i.e., at large $\rm{tan}\beta$, when hZ production dies out, the associated hA production becomes the dominant mechanism with rates similar to the previous case. In this region the masses of h and A are approximately equal. For masses below $\sim 110~$~\Gevcc, the SM Higgs decays into $\bb$ in approximately 85\% of the cases and into $\tautau$ in approximately 8\% of the cases. Similar branching ratios (BR) are expected for the MSSM Higgs bosons. Above $\MH \sim 135$~\Gevcc, the BR into W and Z pairs becomes dominant. \section{Searches at LEP2} While at LEP1 energies the signal to noise ratio was as small as $10^{-6}$ due to the very high $\qq$ cross section, at LEP2 the signal to noise ratio is much more favourable, increasing to $\simeq1\%$. In order to reduce this background, mainly due to W pair production, $\qq$ (with two gluons or two additional photons in the final state) and ZZ events, use is made of b-tagging techniques which exploit the large BR of the Higgs into $\bb$. For $\MH\simeq\MZ$, as is the case for the expected experimental sensitivity, ZZ production represents an irreducible source of background since the Z decays into $\bb$ in 15\% of the cases. The following event topologies are studied: \begin{itemize} \item[$ i)$] The leptonic channel (Z$\to \bha, \mumu$, H$\to\bb$) which represents $7\%$ of the Higgs-strahlung cross section. These events are characterised by two energetic leptons with an invariant mass close to $\MZ$ and a recoil mass equal to $\MH$. Because of the clear experimental signature, no b-tag is necessary and therefore the signal efficiency is high, typically $\sim75\%$. \item[$ ii)$] The missing energy channel (Z$\to \nunu$, H$\to\bb$) comprising $\simeq20\%$ of the Higgs-strahlung cross section. This channel is characterised by a missing mass consistent with $\MZ$ and two b-jets. The selection efficiency is $\simeq35\%$. \item[$ iii)$] The four jet channel (Z$\to \qq$, H$\to\bb$) which is not as distinctive as the two previous topologies but compensates for this drawback with its large BR of $\simeq64\%$. The efficiency for this channel is typically $\simeq40\%$. \item[$ iv)$] The $\tautau \qq$ channel (Z$\to \tautau$, H$\to\qq$ and vice-versa) with a $\simeq9\%$ BR. The event topology includes two hadronic jets and two oppositely-charged, low multiplicity jets due to neutrinos from the $\tau$ decays. The signal efficiency is of the order of 25\%. \end{itemize} The b-tagging algorithms are based on the long lifetime of weakly decaying b-hadrons, on jet shape variables such as charged multiplicity or boosted sphericity and on high $p_t$ leptons from semileptonic b decays. The b-jet identification is improved by combining information from the different b-tagging algorithms with tools like neural-networks and likelihoods. Typically, for a 60\% signal efficiency, the WW background, which has no b-content, is suppressed by a factor over 100, and the $\qq$ and ZZ backgrounds by approximately a factor 10. With respect to the b-tagging algorithms developed for the measurement at LEP1 of $\rm{R_b}$, the b fraction of Z hadronic decays, the performances at LEP2 have improved by almost a factor of 2, due to vertex detectors with an extended solid angle coverage and to more efficient b-tagging techniques. All the analyses developed for the standard model Higgs produced via the Higgs-strahlung mechanism can be used with no modification for the supersymmetric case, provided that the Higgs decays to standard model particles ($\bb$, $\tautau$). The results can then be reinterpreted in the MSSM context, by simply rescaling the number of expected events by the factor $\rm{sin}^2(\beta-\alpha)$. For the pair production process, the signal consists of events with four b-quark jets or a $\tautau$ pair recoiling against a pair of b-quark jets. \section{Results and prospects} Table~\ref{tab:res} shows the number of selected events in the data for the SM Higgs search, the expected number of background events and the expected numbers of signal events assuming $\MH=95$~\Gevcc~\cite{felcini,al_moriond,del_moriond,l3_moriond,op_moriond}. \begin{table}[h] \caption{Standard Model Higgs search. Number of observed events in the data $n_{\rm obs}$, expected number of background events $n_{\rm back}$ and expected numbers of signal events $n_{\rm sig}$ assuming $\MH=95$~\Gevcc for the four LEP experiments and for their combination. Also shown are the number of events observed and expected by the four experiments combined in the mass window $\Delta\MH=92-96$~\Gevcc.} \label{tab:res} \begin{center} \footnotesize \begin{tabular}{|c|lll|} \hline & $n_{\rm obs}$ &$n_{\rm back}$ & $n_{\rm sig}$ \\ \hline ALEPH & 53 & 44.8 & 13.8 \\ DELPHI & 26 & 31.3 & 10.1 \\ L3 & 30 & 30.3 & 9.9 \\ OPAL & 50 & 43.9 & 12.6 \\ \hline Total & 159 & 150 & 46.4 \\ \hline $\Delta\MH=92-96$~\Gevcc & 47 & 37.5 & 24.6 \\ \hline \end{tabular} \end{center} \end{table} As can be observed from Table~\ref{tab:res}, an excess of events is observed by ALEPH~\cite{al_moriond} and OPAL~\cite{op_moriond} which, in the case of OPAL, is concentrated in the mass region around $\MH\simeq\MZ$, while for ALEPH it is distributed over higher masses, typically $\geq95$~\Gevcc. These results translate into the lower limits shown in Table~\ref{tab:lim}, together with the sensitivity (expected limit) of each experiment. \begin{table}[t] \caption{ Observed 95\% C.L. lower limits on $\MH$. Also shown are the limits predicted by the simulation if no signal were present. } \label{tab:lim} \begin{center} \footnotesize \begin{tabular}{|c|cc|} \hline & Observed & Expected \\ & limit (\Gevcc)& limit(\Gevcc)\\ \hline ALEPH & 90.2 & 95.7 \\ DELPHI & 95.2 & 94.8 \\ L3 & 95.2 & 94.4 \\ OPAL & 91.0 & 94.9 \\ \hline \end{tabular} \end{center} \end{table} Table~\ref{tab:MSSM_lim} shows the preliminary 95\% C.L. lower limits on $\Mh$ and $\MA$ for the four LEP experiments~\cite{felcini,al_moriond,del_moriond,l3_moriond,op_moriond}, as well as the derived excluded ranges of $\tan\beta$ for both no mixing and maximal mixing in the scalar-top sector. \begin{table}[b] \caption{ Observed 95\% C.L. lower limits on $\Mh$ and $\MA$. Also shown are the derived excluded ranges of $\tan\beta$. The mass limits are given for $\tan\beta>1$, except for those of DELPHI, given for $\tan\beta>0.5$. } \label{tab:MSSM_lim} \begin{center} \footnotesize \begin{tabular}{|c|cccc|} \hline & $\Mh$ (\Gevcc) & $\MA$ (\Gevcc) & $\tan\beta$ & $\tan\beta$ \\ & & & max. mixing & no mixing \\ \hline ALEPH & 80.8 & 81.2 & - & $1<\tan\beta<2.2$ \\ DELPHI & 83.5 & 84.5 & $0.9<\tan\beta<1.5$ & $0.6<\tan\beta<2.6$ \\ L3 & 77.0 & 78.0 & $1.<\tan\beta<1.5$ & $1.<\tan\beta<2.6$ \\ OPAL & 74.8 & 76.5 & - & $0.81<\tan\beta<2.19$ \\ \hline \end{tabular} \end{center} \end{table} In the years 1999 to 2000 LEP2 is expected to deliver a luminosity larger than 200~$\rm{pb}^{-1}$ per experiment at a centre-of-mass energy eventually as high as $\sim 200$~GeV. These data should allow to discover a SM Higgs of 107~\Gevcc or to exclude a Higgs lighter than $\sim$108~\Gevcc~\cite{lellouch,chamonix}. This is a particularly interesting region to explore, given the present indication for a light Higgs from the standard model fit of the electroweak precision data. The sensitivity to the Higgs in the MSSM will reach $\sim90$~\Gevcc for the high $\tan\beta$ region and $\sim108$~\Gevcc for $\tan\beta\simeq1$, therefore allowing good coverage of the MSSM plane. \section*{Acknowledgments} I would like to thank Cesareo Dominguez and Raul Viollier for their great hospitality and excellent organization of the Workshop and ``Maestro'' Patrick Janot for his precious advice and for carefully reading this manuscript.
\section{Introduction :} \label{intro} Random matrices are useful for a wide range of physical problems. In particular, they can be related to two-dimensional quantum gravity coupled to matter fields with a non-zero central charge $C$ \cite{refgen}. While $C \leq 1$ models are relatively well understood, the $C >1$ domain remains almost totally unknown : there is a $C=1$ ``barrier''. When studying $C \neq 0$ models, we are led to consider multi-matrix models \cite{staudacher,eynard} which are often non-trivial. One class of difficult matrix models corresponds to the $q$-state Potts model (in short: Potts-$q$) on a random surface. This model is a $q$-matrix model where all the matrices are coupled to each other, thus making difficult the use of usual techniques such as the saddle point or the orthogonal polynomials method. Moreover, the $q \rightarrow 4$ limit corresponds to $C \rightarrow 1$, thus, by solving Potts-$q$ models, we shall gain a new understanding of the $C=1$ barrier. In this letter, we show that, contrary to what was previously thought, one can use the loop equations to solve the Potts-3 random matrix model, and we find that the resolvent (which generates many of the operators of the problem) obeys an algebraic equation that we write explicitely. We also show that this method applies when one adds branching interactions (gluing of surfaces, also called ``branched polymers'') \cite{bp} and we derive the critical line of this extended model. The extension to the model with branching interactions and the study of its phase diagram is necessary to verify \cite{FD}'s conjecture about the $C=1$ transition. Finally, we apply the method to the Potts-$\infty$ matrix model, which corresponds to $C=\infty$. As this work was approaching its completion, a paper appeared on the dilute Potts model \cite{PZJ}, which partially overlaps our present work. In this article, the author also has an algebraic equation for the conventional Potts-3 model. Here, we go further as we consider the Potts-3 + branching interactions model. Moreover, his method is quite different : while he uses analytical considerations on the resolvents and large-$N$ techniques, we solve our model by the loop equations method, which can be extended to finite $N$ problems and is also more adapted to the use of renormalization group techniques \cite{higuchi,GBFD}. \section{The Potts-3 + branching interactions model :} Let us define : \begin{equation} Z=\int \, d\Phi \, e^{-N^2 V(\Phi)} \end{equation} \begin{equation} V(\Phi)=g {\makebox{tr} \Phi^3 \over 3 N}+ \psi({ \makebox{tr} \Phi^2 \over 2 N} , { \makebox{tr} \Phi \delta \Phi \delta \over 2 N} ) \end{equation} \begin{equation} \Phi= \left( \begin{array}{ccc} \Phi_1 & 0 & 0\\ 0 & \Phi_2 & 0\\ 0 & 0 & \Phi_3\\ \end{array} \right) , \, \delta_{1}= \left( \begin{array}{ccc} 0 & 1 & 0\\ 0 & 0 & 1\\ 1 & 0 & 0\\ \end{array} \right) , \, \delta_{-1}= \left( \begin{array}{ccc} 0 & 0 & 1\\ 1 & 0 & 0\\ 0 & 1 & 0\\ \end{array} \right) \end{equation} $\delta= \delta_{1} + \delta_{-1}$. We shall also use later the notation $\delta_{0} = Id .$\\\ \noindent $\Phi$, $\delta_0$, $\delta_1$ and $\delta_{-1}$ are $(3 N) \times (3 N)$ and $\Phi_1$, $\Phi_2$, $\Phi_3$ $N \times N$ hermitean matrices. $\psi$ is a general two-variable function, and will mainly appear through its partial derivatives $U$ and $c$ with respect to ${\makebox{tr} \Phi^2 \over 2 N}$ and ${\makebox{tr} \Phi \delta \Phi \delta \over 2 N}$ respectively. If these are constants, then we recover the conventional Potts-3 model (no branching interactions). This model was given partial solution by J.M. Daul \cite{daul}, by considering the analytic structure of the resolvents. He had its critical point and its associated critical exponent. He did not know, however, if the resolvent obeyed an algebraic equation. We shall give here the expression of this equation for the conventional and extended Potts-$3$ model. We also derive the critical line of the extended model and check it corresponds to Daul's result in the particular case of the conventional model. \\ Let us note, for convenience : \begin{equation} t_{i_1 i_2 \cdots i_n \Phi^k} ={1 \over 3 N} \langle \makebox{tr} \delta_{i_1} \Phi \delta_{i_2} \Phi \ldots \delta_{i_n} \Phi^k \rangle \end{equation} where $i_1$, $\ldots$, $i_n$ can be $+1$, $-1$ or $0$. This trace is non-zero if and only if $i_1+\ldots+i_n \equiv 0 \pmod{3}$. $\langle \ldots \rangle$ is the expectation value of $(\ldots)$ : \begin{equation} \langle \ldots \rangle= {1 \over Z} \int d\Phi \, (\ldots)\, e^{-N^2 V(\Phi)} \end{equation} A trace will be said to be ``of degree $m$'' if there are $m$ matrices $\Phi$ in it. For example, the above trace is of degree $k+n-1$.\\\ Let us now use the method of the equations of motion (or loop equations). If we make the infinitesimal change of variables in $Z$ : \begin{equation} \Phi \rightarrow \Phi + \epsilon \,\delta_{i_1} \Phi\, \delta_{i_2} \ldots \Phi \delta_{i_n} \end{equation} with $$i_1+i_2+\ldots +i_n \equiv 0 \pmod{3} $$ then we obtain the expression of the general equations of motion : \begin{equation} \label{eq7} g \, t_{i_1 \ldots i_n \Phi^2}+U \, t_{i_1 \ldots i_n \Phi} +c\, (t_{(i_1+1) i_2 \ldots (i_n-1) \Phi} + t_{(i_1-1) i_2 \ldots (i_n+1) \Phi}) - \sum_{j=1}^{n-1} t_{i_1 \ldots i_j} \, t_{i_{j+1} \ldots i_n} =0 \end{equation} The first three terms come from the transformation of $V(\Phi)$, and the last one, from the jacobian of the transformation. \eq{eq7} relates any expectation value of trace containing a quadratic term (i.e. a $\Phi^2$ term) to expectation values of traces of lower degrees. The problem is that we do not have any recursion relation for more general expectation values like $t_{i_1 \ldots i_n \Phi}$ where all the $i_k \neq 0$. Moreover, when one wants to compute even a very simple trace : for example $t_{\Phi^n}$ by using \eq{eq7}, one obtains, $[{n \over 2}]$ steps later, a $n - [{n \over 2}]$ degree complicated trace which does not contain quadratic terms any more. Thus, the recursion stops there. In fact, this problem can be overcome by a very simple idea : one uses the invariance of traces by cyclic permutations to get rid of the $n+1$ degree term in \eq{eq7}. Then, one obtains relations between general traces, and it is thus possible to compute the expectation values of any trace in function of the first ones. Let us see now how this idea applies to the computation of the resolvent. We denote : \begin{equation} \nonumber \omega_{i_1 i_2 \ldots i_n}={1 \over 3 N} \langle \makebox{tr} \delta_{i_1} \Phi \delta_{i_2} \Phi \ldots \delta_{i_n} {1 \over z-\Phi} \rangle \end{equation} $\omega_0 =\omega$ is the usual resolvent. Using the change of variables : \begin{equation} \nonumber \Phi \rightarrow \Phi + \epsilon {1 \over z-\Phi} \end{equation} we obtain the equation : \begin{equation} \label{e10} z\, (U+g z)\, \omega - U-g z-g \, t_{\Phi} -\omega^2 +2\, c \,\omega_{1-1}=0 \end{equation} Similarly, \begin{equation} \nonumber \Phi \rightarrow \Phi + \epsilon \delta_{1} \Phi \delta_{-1} {1 \over z-\Phi} \end{equation} yields : \begin{equation} \label{e12} z \,(U+g z)\, \omega_{1-1}-(U+g z) \, t_{\Phi} -g\, t_{1-1\Phi} -\omega \, \omega_{1-1} +c \,\omega_{1\ 0-1} + c\, \omega_{-1-1-1}=0 \end{equation} and, by the means of similar changes in variables, we have the equations : \begin{equation} z\, (U+g z)\, \omega_{-1-1-1}-(U+g z)\, t_{1-1\Phi}-g\, t_{-1-1-1\Phi} +c \,\omega_{-1\ 0-1-1}+c \,\omega_{1\ 1-1-1}- \omega \, \omega_{-1-1-1} =0 \end{equation} \begin{equation} z\, (U+g z)\, \omega_{1\ 1-1-1}-(U+g z)\, t_{-1-1-1\Phi} -g\, t_{1\ 1-1-1\Phi} + c\, \omega_{1\ 0\ 1-1-1}+c \, \omega_{-1-1\ 1-1-1}-\omega \, \omega_{1\ 1-1-1} =0 \end{equation} These equations alone are not sufficient to compute $\omega(z)$. Indeed, if we intend to calculate $\omega(z)$, we generate the function $\omega_{1-1}(z)$ (\eq{e10}). Then, in turn, we generate the function $\omega_{-1-1-1}(z)$ (\eq{e12}) and so on. As for $\omega$ functions containing a $0$ (i.e. a $\Phi^2$ term) such as $\omega_{1\ 0-1}$, they are easy to deal with : we know how to compute traces containing $\Phi^2$. $\omega_{1\,0-1}={1 \over 3 N}\makebox{tr} \delta_1 \Phi^2 \delta_{-1} {1 \over z -\Phi}$, will be seen as ${1 \over 3 N} \makebox{tr} \Phi^2 \delta_{-1} {1 \over z -\Phi} \delta_1$. Then the change in variables : \begin{equation} \Phi \rightarrow \Phi + \epsilon \delta_{-1} {1 \over z-\Phi} \delta_{1} \end{equation} yields ($\omega_{1-1} =\omega_{-1\ 1}$ for symmetry reasons) \begin{equation} g \, \omega_{1 0-1}+(U+c)\, \omega_{1-1} +c \, z \, \omega-c=0 \end{equation} and similar changes in variables lead to the equations : \begin{equation} g \, \omega_{-1\ 0-1-1}+U \, \omega_{-1-1-1}+c \, \omega_{1\ 0-1}+c \, z \, \omega_{1-1} -c \, t_{\Phi}=0 \end{equation} \begin{equation} g \, \omega_{1\ 0\ 1-1-1}+U \, \omega_{1\ 1-1-1} +c \, z \, \omega_{-1-1-1}-c \, t_{1-1\Phi} +c \, \omega_{-1\ 0-1-1} - t_{\Phi} \, \omega=0 \end{equation} But, to compute $\omega_{-1-1\ 1-1-1}$, as mentionned in the comments to \eq{eq7}, we have to substract two different changes in variables and use cyclicity of traces : \begin{equation} \Phi \rightarrow \Phi + \epsilon [\Phi \delta_{-1} \Phi \delta_{-1} (z-\Phi)^{-1} \delta_{-1} \Phi - \delta_{-1} \Phi \delta_{-1} (z-\Phi)^{-1} \delta_{-1} \Phi^2] \end{equation} yields : \begin{equation} c \, (\omega_{-1-1\ 1-1-1}+\omega_{-1\ 1-1-1-1}-\omega_{-1-1\ 0-1}-\omega_{-1\ 0\ 1\ 1-1})-\omega_{-1-1-1}+t_{\Phi} \, \omega_{1-1}=0 \end{equation} This equation, as we know how to compute $\omega_{1-1}$, $\omega_{-1-1-1}$, $\omega_{-1-1\ 0-1}$ and $\omega_{-1\ 0\ 1\ 1-1}$, relates $\omega_{-1-1\ 1-1-1}$ to $\omega_{-1\ 1-1-1-1}$. $$\Phi \rightarrow \Phi + \epsilon (\Phi \delta_{-1} \Phi \delta_{-1} \Phi \delta_{-1} (z-\Phi)^{-1} \Phi - \delta_{-1} \Phi \delta_{-1} \Phi \delta_{-1} (z-\Phi)^{-1} \Phi^2)$$ allows us to relate similarly $\omega_{-1\ 1-1-1-1}$ to $\omega_{1-1-1-1-1}$. Then \begin{equation} \Phi \rightarrow \Phi + \epsilon (\Phi \delta_{-1} (z-\Phi)^{-1} \delta_{1} \Phi \delta_{-1} \Phi \delta_{1} - \delta_{-1} (z-\Phi)^{-1} \delta_{1} \Phi \delta_{-1} \Phi \delta_{1} \Phi ) \end{equation} allows us to relate $ \omega_{1-1-1-1-1}$ to $ \omega_{1-1\,1\,1\,1}$, and we have $\omega_{1-1\ 1\ 1\ 1}=\omega_{-1\ 1-1-1-1}$ as the roles of $\delta_{1}$ and $\delta_{-1}$ are completely symmetric.\\ Finally, as a result of these operations, we have : \begin{equation} \omega_{-1\ 1-1-1-1}+\omega_{-1\ 1-1-1-1}=K(z) \end{equation} where $K(z)$ only contains easy to compute $\omega$ functions. We can then write an equation for $\omega_{-1\ 1-1-1-1}$ and thus for $\omega_{-1-1\ 1-1-1}$ which only involves $\omega$ functions that we either already know or are able to compute similarly as was done during the two first steps of the procedure.\\ That way, our set of equations is closed, and we obtain a degree five algebraic equation for $\omega(z)$. This expression applies to general expressions of $U$ and $c$. For the exact expression of this equation see Appendix A. The equation only contains four unknown parameters : \begin{equation} \label{parameters} t_{\Phi}={1 \over N} \langle \makebox{tr} \Phi_{1} \rangle, \, t_{1-1\Phi}= {1 \over N} \langle \makebox{tr} \Phi_1 \Phi_2 \rangle, \, t_{111\Phi}= {1 \over N} \langle \makebox{tr} \Phi_1 \Phi_2 \Phi_3 \rangle, \, t_{1\ 1-1-1\Phi} ={1 \over N} \langle \makebox{tr} \Phi_1 \Phi_2 \Phi_1 \Phi_3 \rangle \end{equation} These parameters are also those that would be involved if we used the renormalisation group method \cite{BrezinZinn,higuchi,GBFD} to compute the Potts-3 model. The renormalization group flows would relate the conventional Potts-3 to the Potts-3 + branching interactions model, with arbitrary $U$ and $c$; but the presence of $t_{111\Phi}$ shows us it would also be related to the dilute Potts-3 model, where one has a ${1 \over N} \makebox{tr} (\Phi_1+\Phi_2+\Phi_3)^3 $ term. Finally, the $t_{1\,1-1-1\Phi}$ term shows us it may also be related to more complicated quartic models. \\\ We are now going to derive from our equation the critical behaviour and critical line of the model when $U=1+h \makebox{tr} \Phi^2 /6$ and $c$ is a constant. This is the most common type of extension of a matrix model to branching interactions. The values of the unknown parameters given in \eq{parameters} are fixed by the physical constraint that the resolvent has only one physical cut which corresponds to the support of the eigenvalues of $\Phi$. Then, one can study the critical behaviour of the model. It is easy to look for the Potts critical line. Indeed, the scaling behaviour of the resolvent is then, if we denote the physical cut of $\omega$ as $[a,b]$ : \begin{equation} \qquad \omega(z) \sim (z-a)^{1 \over 2} \makebox{ when } z \sim a \makebox{ and } \omega(z) \sim (z-b)^{6 \over 5} \makebox{ when } z \sim b \end{equation} The corresponding exponent $\gamma_s$ is $-{1 \over 5}$, which corresponds to the $C={4 \over 5}$ central charge of the model. Rather than looking for the resolvent for any values of the coupling constants, it is easier to search for the resolvent only on this critical line where the presence of the $6 \over 5$ exponent leads to simple conditions on the derivatives of the algebraic equation. We obtain : \begin{equation} \begin{array}{l} 105 \, c^3 + 4 \,g^2 =0\\ 2480625 \, c^2 \, (-1-4 c + 43 c^2)\, +\, 296100\, c\, (15+113 c)\, h\, -\, 692968\, h^2 =0\\ \end{array} \end{equation} Let us note here that, when $h=0$ (no branching interactions) we recover the Potts-3 bicritical point which agrees with Daul's result \cite{daul} : $$c={2 - \sqrt{47} \over 43}\ ,\ g={\sqrt{105} \over 2} \left({-3+\sqrt{47} \over 41 - \sqrt{47}}\right)^{3 \over 2}$$ Thus, we have shown that the resolvent for the model of Potts-$3$ plus branching interactions obeys a degree five algebraic equation. We have found the critical line and exponent of this extended model. This extends the results of Daul \cite{daul} who had only derived the position of the critical point and exponent of the conventional model. Finally, let us recall that, in a recent paper \cite{PZJ}, P. Zinn-Justin obtains independently algebraic equations for similar problems. His method, though, does not involve loop equations, and is rather in the spirit of \cite{daul}. Moreover, it does not address the problem of branching interactions and thus overlaps our results only in the case of the conventional Potts-3 model. \section{The Potts-$\infty$ model :} We are now going to briefly derive the solution for the Potts-$\infty$ model, from the equations of motion point of view. The purpose of this part is mainly to show the efficiency of our method on this $c=\infty$ model. This model was previously studied by Wexler in \cite{wexler}. \noindent Let us denote $\Phi= \left( \begin{array}{ccc} \Phi_1 &\, & 0\\ \, & \ddots &\, \\ 0 &\, & \Phi_q\\ \end{array} \right)$,\ \ and \ \ $X={(\Phi_1 + \ldots +\Phi_q) \over N} \bigotimes {\bf 1}_{q \times q}$.\\ We shall define the Potts-q partition function as \begin{equation} Z=\int d\Phi e^{-N \, V(\Phi)} \qquad \makebox{where} \qquad V(\Phi)=g {\makebox{tr} \Phi^3 \over 3 N}+U {\makebox{tr} \Phi^2 \over 2 N}+c {\makebox{tr} X^2 \over 2 N} \end{equation} $V(\Phi)$ is of order $q$ when $q\rightarrow \infty$. First, let us use the equations of motion to relate \begin{equation} a(x)={1 \over q N} \langle \makebox{tr} {1 \over x-\Phi} \rangle \qquad \makebox{to} \qquad b(y)= {1 \over q N} \langle \makebox{tr} {1 \over y - X} \rangle \end{equation} Let us also denote : \begin{equation} d(x,y)={1 \over q N} \langle \makebox{tr} {1 \over x-\Phi } {1 \over y-X} \rangle \end{equation} $$\Phi \rightarrow \Phi + \epsilon {1 \over x-\Phi} {1 \over y -X} \qquad \makebox{yields }$$ \begin{equation} (x \, (g x + U)+ c \, y - a(x) - {b(y) \over q}) \ d(x,y)\, +\, g - g y \, b(y) - c \, a(x) - b(y) \, (g x + U ) \, = \, 0 \end{equation} We can get rid of $d(x,y)$ since, when $x\, (g x+U)+c \,y-a(x)-{b(y) \over q}=0$, $d(x,y)$ remains finite, thus $g-g y\, b(y) -c\, a(x)-b(y)\, (g x+U)=0$. This is sufficient to relate $a(x)$ to $b(y)$. Moreover, the value of $b(y)$ is easy to compute when $q=\infty$.\\ Let us briefly summarize this computation : we calculate the value of $\langle \makebox{tr} X^n \rangle$ in the $q \rightarrow \infty $ limit. First : \begin{equation} \langle \makebox{tr} X^n \rangle = \langle \makebox{tr} \Phi_1 \ldots \Phi_n \rangle + O({1 \over q}) \end{equation} (recall that all the $\Phi_i$ play the same role). If we now separate the first $n$ matrices from the remaining $q-n$ (with $q \gg n$), and suppose there is a saddle point for the eigenvalues of ${\Phi_{n+1} + \ldots +\Phi_q \over q}$, then this saddle point is (in the $q \rightarrow \infty$ limit) independent from the matrices $\Phi_1\, , \, \ldots \, , \, \Phi_n$. Then, in this limit, up to a change in variables : $\tilde{\Phi}_k= U \Phi_k U^{-1}$, we have $n$ independent matrice $\tilde{\Phi}_1 \ldots \tilde{\Phi}_n$. Each of them has the partition function \begin{equation} \label{vevc} Z_{\Lambda_{C}}=\int \ d\tilde{\Phi}_k \ e^{-N ({g \over 3}\makebox{tr} \tilde{\Phi}_k^3 +{U \over 2}\makebox{tr} \tilde{\Phi}_k^2 +c \makebox{tr} \tilde{\Phi}_k \Lambda_{C})} \end{equation} As $ \makebox{tr} \Phi_1 \ldots \Phi_n = \makebox{tr} \tilde{\Phi}_1 \ldots \tilde{\Phi}_n$, we have $\langle \makebox{tr} X^n \rangle = \makebox{tr} \langle \tilde{\Phi}_1 \rangle_{\Lambda_{C}} \ldots \langle \tilde{\Phi}_n \rangle_{\Lambda_{C}} $ where $\langle \ldots \rangle_{\Lambda_{C}}$ is the expectation value obtained with the partition function $Z_{\Lambda_{C}}$ (cf \eq{vevc} ).\\ The matrices $\tilde{\Phi}_k\, , \ k=1,\ldots n$ all play the same part, and $\langle \makebox{tr} X^n \rangle = \makebox{tr} \Lambda_{C}^n$, thus \begin{equation} \makebox{tr} \Lambda_{C}^n = \makebox{tr} \langle \tilde{\Phi}_1 \rangle^n \end{equation} This must give us $\Lambda_{C}$, provided we calculate $\langle \tilde{\Phi}_1 \rangle_{\Lambda_{C}}$ in function of $\Lambda_{C}$. This is a solvable problem, but it is much faster to note that \begin{equation} \Lambda_{C} = t_{\Phi} \, {\bf 1}_{q N \times q N} \end{equation} is solution. Thus, ${1 \over q N} \langle \makebox{tr} X^n \rangle = (t_{\Phi})^n$, and $b(y)$ is simply $(y-t_{\Phi})^{-1}$.\\ This gives us immediately the solution for $a(x)$ : it obeys a second order equation and reads : \begin{equation} \label{pottsinf} a(x)={1 \over 2} (x (U + gx)+c \, t_{\Phi} -\sqrt{(x (U+g x)+c t_{\Phi})^2 -4 (U+g x+g t_{\Phi})}) \end{equation} The Potts-$\infty$ plus branched polymers model is thus very similar to an ordinary pure gravity model. As previously, we compute the parameter $t_{\Phi}$ by imposing that the resolvent $a(x)$ has only one physical cut. The model is critical when $a(x)$ behaves as $(x-x_0)^{3 \over 2}$, $x_0$ being a constant, and the critical point verifies (as in \cite{wexler}) : \begin{equation} g_c={1 \over 4 \sqrt{2}} \qquad \makebox{and} \qquad c_c=-{1 \over 2} \end{equation} Let us note finally that the loop equations method used here is appropriate for the renormalization group method of \cite{BrezinZinn,higuchi,GBFD}. \section{Conclusion} In this letter, we have shown that it is possible to solve the Potts-3 and Potts-$\infty$ models on two-dimensional random lattices through the method of the equations of motion. We have obtained a closed set of loop equations for the Potts-3 model, which was thought to be impossible. We have shown that the Potts-3 resolvent obeys an order five equation, and this new knowledge opens the door to the calculation of expectation values of the operators of the model. We have extended the Potts-3 conventional model to Potts-3 plus branching interactions, and given the general algebraic equation and the Potts critical line of this model. Finally, we have shown our method also applies successfully to another Potts model : the Potts-$\infty$ model. \iffalse While we were working on these models with added branching interactions, though, we had a discussion with P. Zinn-Justin where it appeared he had another method to solve Potts-3 (and even Potts-4) matrix model \cite{PZJ} with dilution terms. He also obtains algebraic equations for the resolvents. However, our method is different and, since it relies on the use of the equations of motion, it seems more adapted to the renormalization group method \cite{higuchi,GBFD}. Indeed, we have treated Potts + branched polymers models in order to compare low $q$ to large $q$ critical lines and, if possible, renormalization group flows. This corresponds to $c < 1$ and $c >1$ (Potts-$\infty$ is a $c=+ \infty$ model) models, and we would like to verify if F. David's conjecture \cite{FD}, applied to such models, is right. This work is still in progress, in particular for large $q$ Potts + branched polymers models. \fi We hope to generalize soon our method to more general Potts-$q$ models, in particular for large-$q$ Potts + branching interactions models.
\section{Loop Space} Most of us have learned by experience to work with gauge theory using the gauge potentials $A_\mu(x)$ as variables. But suppose we were to approach gauge theory now for the first time, we would probably ask ourselves the question whether $A_\mu(x)$ are the right variables to use to describe gauge theory. After all, $A_\mu(x)$ are gauge-dependent and therefore physically unobservable. Would it not be wiser instead to describe a physical theory with measurable quantities? Indeed, in classical electrodynamics, the variables used by Faraday and Maxwell were not the gauge potentials $A_\mu(x)$ but the gauge invariant, measurable field strengths $F_{\mu\nu}(x)$. It was only when we started to deal with quantum mechanics that we were forced to turn to $A_\mu(x)$ as variables. That this is so is demonstrated by the famous Bohm-Aharonov experiment \cite{Bohmaha}, as illustrated in Figure \ref{Bohm}. \begin{figure} \centering \input{bohmaharo.pstex_t} \caption{Schematic Representation of the Bohm-Aharonov Experiment.} \label{Bohm} \end{figure} Although the field strength $F_{\mu\nu}(x)$ vanishes throughout the region traversed by the charged particle, there are observable effects of the magnetic field $H$ in the form of diffraction patterns on the screen, showing that $F_{\mu\nu}(x)$ itself is inadequate to describe completely the physical conditions. The Bohm-Aharonov experiment shows that to describe the quantum mechanics of a charged particle interacting with an electromagnetic field, the field strengths $F_{\mu\nu}(x)$ are inadequate and the gauge potentials $A_\mu(x)$ are sufficient. But the potentials $A_\mu(x)$ actually give us more information than we need. To describe the diffraction pattern on the screen, it is already sufficient to know the loop integrals: \begin{equation} \alpha_C = e \oint_C A_\mu(x)\, dx^\mu, \label{alphaC} \end{equation} over closed paths $C$, not necessarily the $A_\mu(x)$'s themselves. Indeed, even $\alpha_C$ are more than necessary, for if all $\alpha_C$ change by integral multiples of $2\pi$, the diffraction pattern will not be affected. Thus, what we need are only the phase factors: \begin{equation} \Phi_C = \exp ie \oint_C A_\mu(x)\, dx^\mu. \label{phiC} \end{equation} Hence, we conclude, in the words of Wu and Yang, ``The field strength $F_{\mu\nu}$ underdescribes electromagnetism, $\ldots$the phase ($\alpha_C$) overdescribes electro\-mag\-net\-ism$\ldots$. What provides a complete description that is neither too much nor too little is the phase factor ($\Phi_C$).''\cite{Wuyang1} By `underdescription' here, we mean that different physical conditions may correspond to the same values of the variables, while by `overdescription', that different values of the variables may correspond to the same physical condition. Hence to have a unique labelling for the physical conditions in terms of $A_\mu(x)$, one will need to factor out the classes of $A_\mu(x)$ which are physically equivalent. In contrast, what is nice about the phase factors $\Phi(C)$ is that the same physical condition corresponds to the same values of $\Phi(C)$, and different conditions to different values, although any given set of values of $\Phi(C)$ need not necessarily correspond to a physical condition. The situation in nonabelian Yang-Mills theories is similar, except that here, even in the classical theory, the field strengths $F_{\mu\nu}(x)$ no longer offer a sufficient description \cite{Wuyang2}. For the quantum theory, the gauge potentials $A_\mu(x)$ are again adequate but overdescribe the theory, and what provides a complete description for the theory yet not an over-description are the path-ordered phase factors (Wilson loops): \begin{equation} \Phi_C = P \exp ig \oint_C A_\mu(x)\,dx^\mu. \label{PhiC} \end{equation} Why then do we not use $\Phi_C$ as variables to describe gauge theory? The reason is that $\Phi_C$ is labelled by the loops $C$ in space-time which are infinitely more numerous than the points $x$ in space-time. Since the gauge potentials $A_\mu(x)$ labelled by $x$ are already sufficient to describe gauge theory, a description in terms of $\Phi_C$ must therefore be highly redundant. By `redundant' here, we mean that not all points in the space spanned by the variables $\Phi(C)$ correspond to actual physical conditions, but only a subset of it which we may think of as a constraint surface in that space. The `redundancy' being infinite, the constraint required is bound to be complicated, which makes the description in terms of loop variables extremely clumsy. Hence, in usual circumstances, one would much rather deal with the vagaries of the gauge-dependent $A_\mu$ than with the redundancy of $\Phi_C$. But there are situations some of which we shall discuss, where a description in terms of $\Phi_C$ is preferable, indeed may even be necessary. In that case we shall need to face the complications and develop the formalism for dealing with gauge theory in terms of loop quantities. To effect a loop space formulation of gauge theory, our first task would be to label the loops in space-time, or in other words to introduce some sort of co-ordinates in loop space. (It will be seen that it is sufficient to consider only those loops passing through some fixed reference point $P_0 = \xi_0^\mu$.) An obvious possibility is to label a loop by the space-time co-ordinates of the points on it, thus: \begin{equation} C: \ \ \ \{\xi^\mu(s) \colon s = 0 \rightarrow 2\pi,\ \xi(0) = \xi(2\pi) = \xi_0 \}, \label{ximus} \end{equation} so that $\Phi_C$ in (\ref{phiC}) or (\ref{PhiC}) can be rewritten as: \begin{equation} \Phi[\xi] = P_s \exp ig \int_0^{2\pi} ds\,A_\mu(\xi(s)) \dot{\xi}^\mu(s), \label{Phixi} \end{equation} where a dot denotes differentiation with respect to the parameter $s$. This labelling, however, is again redundant in that if one replaces $s$ by another parametrization $s'= f(s)$, it would leave the phase factor $\Phi_C$ invariant. To effect a unique labelling of $C$, these reparametrizations should in principle be factored out. Nevertheless, this quotient space is so complicated that most people would rather live with the redundancy of the space of the loops parametrized by the functions $\xi$. This is the attitude that we shall adopt. In {\it parametrized loop space}, loop quantities such as $\Phi[\xi]$ are just functionals of the continuous (piece-wise smooth) functions $\xi$ of $s$, which are relatively easy to handle, although care has always to be taken in removing the additional redundancy introduced by the parametrization. Thus, for example, a derivative can be introduced in (parametrized) loop space just as the functional derivative with respect to $\xi(s)$. To be specific, we shall define the derivative as: \begin{equation} \delta_\mu(s) \Psi[\xi] = \lim_{\Delta \rightarrow 0} \frac{1}{\Delta} \{\Psi[\xi'] - \Psi[\xi]\}, \label{loopderiv} \end{equation} with: \begin{equation} \xi'^\alpha(s') = \xi^\alpha(s') + \Delta \delta_\mu^\alpha\, \delta(s-s'), \label{xiprime} \end{equation} meaning that the loop is `plucked' by a delta function in the direction $\mu$ at the point on the loop corresponding to the value $s$ of the parameter. This definition has to be interpreted with some care, especially when approximating the continuum by a discretized space as in lattice theories, where a careless handling may easily lead, for example, to asymmetric second derivatives \cite{Corhass, loop2,Book}. Our next task in a loop space formulation is to select the variables for describing the gauge field. In doing so, we have to bear in mind a major problem already mentioned before in connection with the high degree of redundancy in loop variables. For example, suppose we choose the phase factors $\Phi[\xi]$. If we allow all of these $\Phi$'s to take any value in the gauge group $G$, then clearly not all of them will be expressible in terms of a local gauge potential $A_\mu(x)$ via (\ref{Phixi}), there not being enough freedom in $A_\mu(x)$ to satisfy all the conditions thereby imposed. In other words, there are certain sets of values of the variables $\Phi[\xi]$ in $G$ which are unphysical. Thus, in order to ensure that in changing to a loop description we are still dealing with the same theory though in a different language, we have to impose contraints on the values that these loop variables can take so as to guarantee that one can recover a local gauge potential $A_\mu(x)$ from them. The ability to write down the appropriate constraints for doing so is thus one of the first consideration in any loop formulation of gauge theory. For this reason, instead of the seemingly more natural choice of $\Phi[\xi]$ as variables, we choose rather to work with the quantities $F_\mu[\xi|s]$ first introduced by Polyakov \cite{Polyakov} as the logarithmic loop derivatives of the phase factors $\Phi[\xi]$ in (\ref{Phixi}), namely \begin{equation} F_\mu[\xi|s] = \frac{i}{g} \Phi^{-1}[\xi]\, \delta_\mu(s) \Phi[\xi], \label{Fmuxis} \end{equation} Its meaning in space-time is illustrated in Figure \ref{Fmufig}, \begin{figure} \centering \input{pluckloop.pstex_t} \caption{Illustration for the quantity $F_\mu[\xi|s]$} \label{Fmufig} \end{figure} where it can be seen that $F_\mu[\xi|s]$ depends only on that part of the loop before the point labelled by $s$ (hence the special notation $[\xi|s]$ for its argument). The virtue of the quantity $F_\mu[\xi|s]$ lies in the fact that in loop space it plays the role of a sort of `connection' similar to that played by the gauge potential $A_\mu(x)$ in space-time. By (\ref{Fmuxis}) it tells us how the phase of $\Phi[\xi]$ changes as one moves from one loop to a neighbouring loop, i.e. from point to neighbouring point in loop space. Of course, when the phase factor $\Phi[\xi]$ exists as a single-valued function in loop space, then $F_\mu[\xi|s]$ as given in (\ref{Fmuxis}) is trivial as a `connection', corresponding just to what is known in gauge theory language as `pure gauge'. But for $F_\mu[\xi|s]$ as variables taking arbitrary values the corresponding connection is not in general trivial. Because of this geometrical significance, the redundancy-removing contraints take on a particularly elegant and physically lucid form, which is intimately related to the concept of monopoles as topological obstructions in gauge theories. The explicit formulation of these constraints has thus to be postponed to Section 3 after the concept of monopoles has been introduced. For the moment, we must first prepare some necessary tools. Given the concept of $F_\mu[\xi|s]$ as a `connection' in loop space, one can proceed as usual to define a loop space curvature as: \begin{equation} G_{\mu\nu}[\xi|s] = \delta_\nu(s) F_\mu[\xi|s]- \delta_\mu(s) F_\nu[\xi|s] + ig [F_\mu[\xi|s], F_\nu[\xi|s]]. \label{Gmunu} \end{equation} This is the exact parallel of the familiar formula for the field strength $F_{\mu\nu}(x)$ in terms of the gauge potential $A_\mu(x)$, and has the same geometric significance of a parallel phase transport around an infinitesmal circuit, but now in loop space. Its meaning in space-time, however, is as illustrated in Figure \ref{skiprope} where the loop `skips' over a small 3-volume in space. For the pure gauge connection in \begin{figure} \centering \input{loopcurv.pstex_t} \caption{Illustration for $G_{\mu\nu}[\xi|s]$.} \label{skiprope} \end{figure} (\ref{Fmuxis}), the curvature $G_{\mu\nu}[\xi|s]$ is of course zero. But in general the curvature need not vanish, in which case, as will be shown in the next section, it means physically that there is a monopole charge enclosed inside the small 3-volume `skipped' over by the loop in Figure \ref{skiprope}. Further, just as one has constructed from the potential $A_\mu(x)$ the phase factor $\Phi[\xi]$ which, as the `holonomy', is an extension of the concept of curvature to a finite-sized loop, so a holonomy in loop space can also be constructed from the `connection' $F_\mu[\xi|s]$ as \cite{loop1}: \begin{equation} \Theta_\Sigma = P_t \exp ig \int_0^{2\pi} dt \int_0^{2\pi} ds\, F_\mu[\xi_t|s] \frac{\partial \xi_t^\mu(s)}{\partial t}, \label{Thetasigma} \end{equation} where $\Sigma$ denotes the parametrized surface: \begin{equation} \Sigma: \{ \xi_t^\mu(s) \colon s =0 \rightarrow 2 \pi, t = 0 \rightarrow 2 \pi \}, \label{Sigma} \end{equation} with: \begin{equation} \xi_t^\mu(0) = \xi_t^\mu(2 \pi) = \xi_0^\mu,\ t = 0 \rightarrow 2 \pi, \label{xit0} \end{equation} \begin{equation} \xi_0^\mu(s) = \xi_{2\pi}^\mu(s) = \xi_0^\mu,\ s = 0 \rightarrow 2 \pi. \label{xis0} \end{equation} The closed surface $\Sigma$ swept out by the one-parameter family of loops $\xi_t$ is illustrated in Figure \ref{looptheloop}, which may be considered also as a loop in loop space. Again, for the `pure gauge' connection (\ref{Fmuxis}), $\Theta_\Sigma$ is trivial and equals the group identity. However, it will not be so when the volume enclosed by $\Sigma$ contains monopole charges, as we shall see later. \begin{figure} \centering \includegraphics{loopsph.eps} \caption{Surface swept out by the one-parameter family of loops $\xi_t$.} \label{looptheloop} \end{figure} To round off this section, we mention some facts which we shall find useful later. From Figure \ref{Fmufig}, it can be seen that in terms of ordinary field variables, $F_\mu[\xi|s]$ is also expressible as \cite{Polyakov,Zois}: \begin{equation} F_\mu[\xi|s] = \Phi_\xi^{-1}(s,0) F_{\mu\nu}(\xi(s)) \Phi_\xi(s,0) \dot{\xi}^\nu(s), \label{Fmuinx} \end{equation} where $\Phi_\xi(s_2,s_1)$ is the parallel phase transport: \begin{equation} \Phi_\xi(s_2,s_1) = P_s \exp ig \int_{s_1}^{s_2} ds\, A_\mu(\xi(s)) \dot{\xi}^\mu (s) \label{Phis2s1} \end{equation} from $s_1$ to $s_2$ along the loop $\xi$. This formula (\ref{Fmuinx}) allows us to translate from loop space language back to local field language. For example, by substitution of (\ref{Fmuinx}) into the expression: \begin{equation} {\cal A}_F^0 = -\frac{1}{4 \pi \bar{N}} \int \delta \xi \int_0^{2\pi} ds \,{\rm Tr}\{F_\mu[\xi|s] F^\mu[\xi|s] \} |\dot{\xi}(s)|^{-2}, \label{CalA0F} \end{equation} and performing the functional integral, one obtains the standard pure Yang-Mills action: \begin{equation} -\frac{1}{16 \pi} \int d^4x\, {\rm Tr}\{ F_{\mu\nu}(x) F^{\mu\nu}(x) \} \label{calA0F} \end{equation} in terms of local field variables, if we define the normalization factor $\bar{N}$ as: \begin{equation} \bar{N} = \int_0^{2\pi} ds \int \prod_{s' \neq s} d^4 \xi(s'). \label{Nbar} \end{equation} The expression (\ref{CalA0F}) will serve as the loop space field action in terms of $F_\mu[\xi|s]$ as variables. \setcounter{equation}{0} \section{Monopoles} Monopoles occur as topological obstructions in gauge theories with compact multiply-connected gauge groups. They may be defined as follows \cite{Lubkin,Wuyang1,Coleman}. Take a 1-parameter family of closed loops $\{C_t\}$, as that parametrized by $\xi_t$ in (\ref{Sigma}) above, sweeping out a 2-dimensional surface $\Sigma$. For each $t$ we can then associate a phase factor $\Phi(C_t)$ which is an element of the gauge group $G$. As $t$ varies from $0$ to $2\pi$, $\Phi(C_t)$ traces out a closed curve, say $\Gamma_\Sigma$, in $G$. If $G$ is multiply-connected, then $\Gamma_\Sigma$ will belong to one or other of the homotopy classes $\pi_0(G)$ of closed curves in $G$, where members of different classes cannot be continuously deformed into one another. The homotopy class to which $\Gamma_\Sigma$ belongs is defined as the monopole charge enclosed inside the surface $\Sigma$. At first sight, this might seem a rather abstruse definition for a monopole for which, after all, the primary example is just the magnetic charge of electromagnetism, which can be represented simply by a novanishing divergence of the magnetic field, or in relativistic notation by a nonvanishing $\partial_\mu {}^*\!F^{\mu\nu}(x)$. On closer examination, however, it is easily seen first, that the above definition reduces in the abelian theory back to the usual interpretation of the monopole as a source of the field ${}^*\!F$, and secondly, that for a nonabelian theory this latter interpretation no longer works and cannot be used to define the monopole \cite{Book}. Indeed, the above definition, or its equivalent, is the only known valid extension of Dirac's magnetic monopole to nonabelian Yang-Mills theory. This definition exhibits the essentially topological nature of the monopole charge which is by definition discrete, and since invariant under continuous deformations, also conserved. The values that this charge can take depend on the topological property of the gauge group. Thus, for (compact) electrodynamics, the gauge group is $U(1)$ which has the topology of a circle, on which the homotopy classes of closed curves are labelled by their winding numbers. As a result, the magnetic charge is quantized, meaning that it takes integral values in ${\bf Z}$, as first noted by Dirac \cite{Dirac}. For simply-connected gauge groups such as $SU(N)$, there can be no monopoles, there being only one homotopy class of closed curves which contains the vacuum. However, this does not mean that there can be no nonabelian monopoles in gauge theories with $su(N)$ symmetries. By an $su(N)$ theory, one usually means a theory invariant under the gauge Lie algebra $su(N)$. This by itself does not specify the gauge group, since different Lie groups can correspond to the same Lie algebra. But it is the the global structure of the gauge group which determines whether a theory can have monopoles. Thus, for the pure $su(N)$ Yang-Mills theory containing only gauge bosons in the adjoint representation and nothing else, the gauge group is $SU(N)/{\bf Z}_N$ and not $SU(N)$, since two elements in $SU(N)$ differing by only a factor $\exp 2i\pi/N$ have the same effect on the gauge boson field and should thus be considered as identical elements of the gauge group. That being the case, and $SU(N)/{\bf Z}_N$ being $N$-tuply connected, the pure $su(N)$ Yang-Mills theory can have monopoles with charges labelled by elements of ${\bf Z}_N$. In particular, pure $su(2)$ Yang-Mills theory has gauge group $SU(2)/{\bf Z}_2 = SO(3)$ and monopole charges labelled by a sign $\pm$, with $+$ corresponding to the vacuum and the charge $-$ being its own conjugate. Similarly, pure $su(3)$ Yang-Mills theory has gauge group $SU(3)/{\bf Z}_3$ and monopole charges labelled by the cube roots of unity $1, \omega, \omega^2$, with $\omega = \exp 2\pi i/3$. To determine the gauge group and hence whether a theory has monopoles, we need to examine the gauge transformation properties of all fields present in the theory \cite{Yang,monocharge}. Take for example the electroweak theory as we know it today which has either: \begin{itemize} \item[(i)] $SU(2)$ doublets with half-integral hypercharges, e.g. $(\nu, e)_L$ with $Y = 1/2$; or else: \item[(ii)] $SU(2)$ singlets or triplets with integral hypercharges, e.g. $e_R$ with $Y = 1$, $A_\mu$ with $Y = 0$. \end{itemize} Hence, if we put: \begin{equation} \tilde{f} = \exp 2\pi i T_3 \ \ \ \in SU(2)_f, \label{ftilde} \end{equation} and: \begin{equation} \tilde{y} = \exp 2\pi i Y \ \ \ \in U(1)_Y, \label{ytilde} \end{equation} the couple $(f \tilde{f}, y \tilde{y})$ in $SU(2)_f \times U(1)_Y$ has exactly the same physical effect as $(f, y)$ and therefore has to be identified with the latter. As a result, the gauge group is not $SU(2) \times U(1)$ but $[SU(2) \times U(1)]/{\bf Z}_2 = U(2)$. Now, in contrast to $SU(2)$, the group $U(2)$ can have monpoles \cite{monocharge}. Indeed, as seen in Figure \ref{U2wind}, $AB$ is a closed curve in $U(2)$ which cannot be continuously deformed to zero. It winds half-way round each of the $SU(2)_f$ and $U(1)_Y$ subgroups. Hence a $U(2)$-monopole of unit charge can be thought of as carrying an $SO(3)$ monopole charge $\eta = -$, as well as a $U(1)_Y$ monopole charge of half the Dirac value. In general, the monopoles of the electroweak theory are labelled by an integer $n$, where a charge $n$ monopole can be thought of as carrying simultaneously: \begin{eqnarray} \eta & = & (-1)^n \ \ \ SO(3)\ {\rm monopole\ charge}, \nonumber \\ \tilde{y} & = & \pi n/g_1 \ \ \ U(1)_Y\ {\rm monopole\ charge}. \label{U2charge} \end{eqnarray} \begin{figure} \centering \includegraphics{u2torus.eps} \caption{Monopoles in the electroweak theory} \label{U2wind} \end{figure} A similar analysis carried out for the full Standard Model given the presently known spectrum of charges reveals that the gauge group is $[SU(3) \times SU(2) \times U(1)]/ {\bf Z}_6$, and that its monopoles, also labelled by an integer $n$, can be thought of as carrying simultaneously the following charges \cite{monocharge}: \begin{eqnarray} \zeta & = & \exp 2\pi i n/3 \ \ \ SU(3)/{\bf Z}_3\ {\rm monopole\ charge}, \nonumber \\ \eta & = & (-1)^n \ \ \ SO(3)\ {\rm monopole\ charge}, \nonumber \\ \tilde{y} & = & 2 \pi n/(3g_1) \ \ \ U(1)_Y\ {\rm monopole\ charge}. \label{SMcharge} \end{eqnarray} This fact will be of use in the physical applications of nonabelian duality \cite{DSMrph98}. The lists given in (\ref{U2charge}) and (\ref{SMcharge}) of available monopole charges in respectively the electroweak theory and the Standard Model are the equivalents of the statement in electromagnetism that the magnetic charge is quantized. The well-known Dirac quantization condition for the abelian theory, however, contains more information, for it says not only that the magnetic charge $\tilde{e}$ is quantized but that it is quantized in units of $1/2e$, which can be thought of as a relation \begin{equation} 2 e \tilde{e} = 1 \label{diraccond} \end{equation} between the minimal coupling strengths. A parallel for this exists also for $su(N)$ theories which for our present normalization convention reads as \cite{dualcomm}: \begin{equation} g \tilde{g} = 1, \label{Diraccond} \end{equation} and can be similarly derived.\footnote{The couplings here which follow the original Dirac convention are the so-called unrationalized couplings. For the rationalized couplings now more in common use, the conditions should read respectively $e \tilde{e} = 2 \pi$ and $g \tilde{g} = 4 \pi$.} The definition given above for the monopole charge is unfortunately a little abstract. In order to be useful, the monopole charge has to be expressed explicitly in terms of whatever field variables one may choose to adopt. In terms of the standard variables $A_\mu(x)$, monopole charges are always a little hard to handle. This can be seen already in the abelian theory. If $A_\mu(x)$ exists and is single-valued, then it follows that: \begin{equation} \partial_\mu F_{\nu\rho} + \partial_\nu F_{\rho\mu} + \partial_\rho F_{\mu\nu} = 0, \label{bianchi} \end{equation} or that for: \begin{equation} {}^*\!F_{\mu\nu} = -\frac{1}{2} \epsilon_{\mu\nu\rho\sigma} F^{\rho\sigma}, \label{fstar} \end{equation} we have: \begin{equation} \partial_\mu {}^*\!F^{\mu\nu} = 0. \label{bianchia} \end{equation} However, in the presence of a monopole, $\partial_\mu {}^*\!F^{\mu\nu}$ cannot vanish. Hence, $A_\mu(x)$ must be singular somewhere. This is the reason for the Dirac string \cite{Dirac}. The way out, of course, is to consider $A_\mu(x)$ as a patched quantity \cite{Wuyang1}. One covers, for example, the sphere by two patches ($N$ and $S$): \begin{eqnarray} (N): \ \ \ & 0 \leq \theta < \pi, \ \ \ & 0 \leq \phi < 2\pi, \nonumber \\ (S): \ \ \ & 0 < \theta \leq \pi, \ \ \ & 0 \leq \phi < 2\pi, \label{patchesNS} \end{eqnarray} and define $A_\mu^{(N)}$ and $A_\mu^{(S)}$ separately in $N$ and $S$, with the two potentials related to each other in the overlap region by a gauge transformation parametrized by the (patching) function: \begin{equation} S = \exp ie\alpha, \label{patchingf} \end{equation} where for $S$ to be well-defined the phase $\alpha$ has to change by an integral multiple of $2\pi$ for $\phi = 0 \rightarrow 2 \pi$, giving thus the Dirac quantization condition (\ref{diraccond}). Similarly for nonabelian Yang-Mills theory, one can introduce patched potentials to accommodate monopoles, the procedure being then a little more complicated. In either case, however, the patching depends on the locations of the monopoles, and the number of patches required increases exponentially with the number of monopoles introduced. It thus appears that a description of monopoles in terms of $A_\mu(x)$ is going to be very complicated, so much so that even the intrinsic difficulty of loop space formulations may now be worth facing by comparison. The definition above of the monopole charge being given in terms of loop quantities in the first place, it is not surprising that it has a simpler representation in the loop space formulation, especially for the nonabelian theory \cite{loop1,Book}. For illustration, it is sufficient to exhibit this explicitly only for the simplest example with gauge group $SO(3)$. Recall that $F_\mu[\xi|s]$ is by definition the logarithmic derivative of $\Phi[\xi]$. Hence, we may write: \begin{equation} \exp ig dt \int_0^{2\pi} ds\, F_\mu[\xi_t|s] (\partial \xi_t^\mu(s)/\partial t) \sim \Phi^{-1}[\xi_{t+dt}] \Phi[\xi_t]. \label{increPhi} \end{equation} The loop space holonomy $\Theta_\Sigma$ is the product ordered in $t$ of such factors and is thus the total change in $\Phi[\xi_t]$ as $t =0 \rightarrow 2\pi$. Now both $\Phi$ and $\Theta$ may be interpreted as an element of either the gauge group $SO(3)$ or its double cover $SU(2)$. However, if we wish to exhibit the monopole charge as an element of ${\bf Z}_2$ considered as a subgroup of $SU(2)$, then we should work in the latter. Remembering that the corresponding curve $\Gamma_\Sigma$ is a closed curve in $SO(3)$, we see that $\Theta_\Sigma$ must wind around $SU(2)$ an odd number of `half-times' if $\Sigma$ contains a monopole charge $-$, but an even number of `half-times' if $\Sigma$ contains no monopole. Hence we conclude: \begin{equation} \Theta_\Sigma = \zeta_\Sigma I, \label{zetasigma} \end{equation} where $\zeta_\Sigma$ is the monopole charge enclosed inside $\Sigma$. This formula (\ref{zetasigma}) actually holds for any theory with gauge group $SU(N)/{\bf Z}_N$. It is instructive to examine how this result arises in detail in terms of the (patched) gauge potential (Figure \ref{holoncurv}). \begin{figure} \centering \input{holoncurv.pstex_t} \caption{A curve representing an $SO(3)$ monopole} \label{holoncurv} \end{figure} Without loss of generality, we shall choose the reference point $P_0 = \xi_0^\mu$ to be in the overlap region, say on the equator which corresponds to the loop $\xi_{t_e}$. Starting at $t=0$, where $\Phi^{(N)}[\xi_0]$ is the identity, the phase factor $\Phi^{(N)}[\xi_t]$ traces out a continuous curve in $SU(2)$ (which has the topology of a 3-sphere) until it reaches $t = t_e$. At $t = t_e$ one makes a patching transformation and goes over to $\Phi^{(S)}[\xi_t]$. From $t = t_e$ onwards, the phase factor $\Phi^{(S)}[\xi_t]$ again traces out a continuous curve until $t$ reaches $2 \pi$, where it becomes again the identity and joins up with $\Phi^{(N)}[\xi_0]$. In order that the curve $\Gamma_\Sigma$ so traced out winds only half-way round $SU(2)$ while being a closed curved in $SO(3)$, as it should if $\Sigma$ contains a monopole, we must have \begin{equation} \Phi^{(N)}[\xi_{t_e}] = - \Phi^{(S)}[\xi_{t_e}], \label{PhiNSrel} \end{equation} which means for the holonomy \begin{equation} \Theta_\Sigma = (\Phi^{(S)}[\xi_{t_e}])^{-1}\,\Phi^{(N)}[\xi_{t_e}] =-I \end{equation} as required in (\ref{zetasigma}). The formula (\ref{zetasigma}) for the monopole charge in terms of the loop space holonomy $\Theta$, though already explicit, is not as convenient for our discussion as the differential formula in terms of the loop space curvature $G_{\mu\nu}$. As noted already in Section 1, curvature is just an infinitesimal version of the holonomy. Hence, in the absence of monopoles $G$ vanishes, but if the loop $\xi$ passes through a monopole at the point labelled by $s$, then $G_{\mu\nu}[\xi|s]$ must take on a value equal to the logarithm of the monopole charge $\zeta$ at that point. Hence, we can write, for a classical monopole moving along the world line $Y^\mu(\tau)$ \cite{monact1}: \begin{equation} G_{\mu\nu}[\xi|s] = - 4 \pi \tilde{g} \int d\tau \kappa[\xi|s] \epsilon_{\mu\nu\rho\sigma} \dot{\xi}^\rho(s) \frac{dY^\sigma(\tau)}{d\tau} \delta^4(\xi(s) - Y(\tau)), \label{Gausslaw} \end{equation} where $\kappa[\xi|s]$ satisfies: \begin{equation} \exp i \pi \kappa = \zeta. \label{condkappa} \end{equation} This formula has to be interpreted with some care since given $\zeta$ the solution for $\kappa$ in (\ref{condkappa}) in the Lie algebra is not unique. But in any case, for $\zeta \neq 1$, $\kappa \neq 0$ which means that monopoles can be regarded as sources of curvature in loop space, as already anticipated. For the abelian theory, the equation (\ref{Gausslaw}) reduces to the familiar formula for a classical point magnetic charge: \begin{equation} \partial_\nu {}^*\!F^{\mu\nu}(x) = - 4 \pi \tilde{e} \int d\tau \frac{dY^\mu(\tau)} {d\tau} \delta(x - Y(\tau)). \label{gausslaw} \end{equation} \setcounter{equation}{0} \section{Wu-Yang Criterion and Poincar\'e Lemma} An attractive feature of monopoles considered as topological obstructions in gauge fields is that their topology defines their own dynamics. This was first pointed out in a beautiful paper by Wu and Yang in 1976 \cite{Wuyangcr}. That this is so is intuitively clear. The assertion that there is a monopole at a certain point $x$ in space-time, as discussed in the last section, means that the gauge field surrounding $x$ has to have a certain topological structure, and if the monopole is displaced to another point, then the gauge field will have to rearrange itself so as to maintain the same topological structure around the new point. There is thus naturally a coupling between the gauge field and the position of the monopole, or in physical language a topologically induced interaction between the field and the monopole. To deduce the explicit form of this interaction, one can proceed as follows. One writes down first the free action of the gauge field together with that of a particle. For example, for electromagnetism and a classical particle of mass $m$, one has: \begin{equation} {\cal A}^0 = {\cal A}^0_F + {\cal A}^0_M, \label{calA0} \end{equation} where ${\cal A}^0_F$ is the usual Maxwell action, and \begin{equation} {\cal A}^0_M = - \int d\tau, \label{calA0M} \end{equation} the integral being taken along the world-line of the particle with $\tau$ being the proper time along the world-line. Extremizing this action with respect to the dynamical variables of the problem, namely $A_\mu(x)$ for the field and co-ordinates $Y^\mu(\tau)$ for the particle, one obtains the free equations of the field and of the particle. Suppose now, however, one stipulates that the particle carries a (magnetic) monopole charge and imposes on to the system the appropriate topological condition that this should be so. This condition couples the field and the particle, as already explained, so that if one extremizes again the action (\ref{calA0M}) under the imposed constraint, the equations of motion will no longer be free but coupled equations involving an `interaction' between the particle and the field. What are the coupled equations so obtained? Knowing as one does that classical electromagnetism is dual symmetric, one is not surprised that they turn out to be just the dual to the Maxwell and Lorentz equations for the motion of an electric charge in an electromagnetic field. Indeed, this was the way that Wu and Yang deduced the equations in their original paper \cite{Wuyangcr}. However, a direct attack on the problem so posed is not as easy as it might seem at first sight. The reason is that $A_\mu(x)$, which is the dynamical variable for the field, is a patched quantity in the presence of a monopole, as explained in the last section, where the patching depends on the other dynamical variable $Y^\mu(\tau)$ for the particle. Extremizing ${\cal A}^0$ with respect to $A_\mu(x)$ is thus not a simple matter, although (according to Wu in private communication) possible. There is, however, a very simple and elegant method for solving this problem \cite{monact1,Book}. The trick is to adopt as field variables not the gauge potential $A_\mu(x)$ but the field strength $F_{\mu\nu}(x)$. Being gauge invariant, $F_{\mu\nu}(x)$ is not patch-dependent even in the presence of a monopole. Further, the topological condition defining a (magnetic) monopole charge at $Y^\mu(\tau)$ can be expressed simply in terms of $F_{\mu\nu}(x)$ as (\ref{gausslaw}). Incorporating (\ref{gausslaw}) as a constraint on the action (\ref{calA0}) by means of a Lagrange multplier $\lambda_\mu(x)$ and extremizing with respect to $F_{\mu\nu}(x)$ and $Y^\mu(\tau)$, one obtains then easily the equations: \begin{equation} F^{\mu\nu}(x) = 4 \pi [\frac{1}{2} \epsilon^{\mu\nu\rho\sigma} (\partial_\sigma \lambda_\rho(x) - \partial_\rho \lambda_\sigma(x))], \label{ELeq1} \end{equation} and \begin{equation} m \frac{d^2 Y^\mu(\tau)}{d \tau^2} = - 4 \pi \tilde{e} [\partial_\nu \lambda_\mu(Y(\tau)) - \partial_\mu \lambda_\nu(Y(\tau))] \frac{d Y^\nu(\tau)}{d \tau}. \label{ELeq2} \end{equation} The equation (\ref{ELeq1}) says that the dual Maxwell field ${}^*\!F_{\mu\nu}(x)$ is also a gauge field: \begin{equation} {}^*\!F_{\mu\nu}(x) = \partial_\nu \tilde{A}_\mu(x) - \partial_\mu \tilde{A}_\nu(x), \label{dfmunu} \end{equation} with the (dual) potential: \begin{equation} \tilde{A}_\mu(x) = 4 \pi \lambda_\mu(x). \label{atilde} \end{equation} Then using this, one can rewrite the other equation (\ref{ELeq2}) as: \begin{equation} m \frac{d^2 Y^\mu(\tau)}{d \tau^2} = - \tilde{e}\, {}^*\!F^{\mu\nu}(Y(\tau)) \frac{d Y_\nu(\tau)}{d \tau}. \label{dlorentzeq} \end{equation} As expected, these equations, together with the constraint (\ref{gausslaw}), are exactly the dual of the equations of motion for an electric charge in an electromagnetic field. There is an important detail in the above derivation which at first sight looks like a flaw but which when understood has far reaching consequences for the development which follows. The variables $F_{\mu\nu}(x)$ adopted to solve the variational problem are more numerous than the original variables $A_\mu(x)$ (there being 6 components to $F_{\mu\nu}$ compared with 4 to $A_\mu$) and must therefore be regarded as `redundant' in the sense the term was used in Section 1 while discussing loop variables. In other words, given a set of values for $F_{\mu\nu}(x)$, there is no guarantee that they can be derived from a potential $A_\mu(x)$ unless their values are appropriately constrained. Now, the beauty of the above derivation is that the dynamical constraint (\ref{gausslaw}) imposed, representing the topological definition of the monopole charge, already ensures the existence of the potential, thereby removing automatically the redundancy in the new variables. Indeed, one notices from (\ref{gausslaw}) that except on the world-line $Y(\tau)$ of the monopole, one has: \begin{equation} \partial_\nu {}^*\!F^{\mu\nu}(x) = 0, \label{divfstar} \end{equation} which is exactly the condition which implies that $F_{\mu\nu}(x)$ is derivable from a potential. This well-known mathematical fact, to which we shall have ample occasions to recall, we shall refer to as the Poincar\'e Lemma, although it is only a very special case of that important theorem in differential geometry. The only place where the condition (\ref{gausslaw}) does not guarantee the existence of the potential is on the monopole world-line, but that is no problem since at the monopole position $A_\mu(x)$ should not exist in any case. The same derivation can be applied also to a quantum particle carrying a monopole charge. For example, for a Dirac particle, we replace the action (\ref{calA0M}) above by \cite{monact2}: \begin{equation} {\cal A}^0_M = \int d^4x\,\bar{\tilde{\psi}}(x)\, (i \partial_\mu \gamma^\mu - m)\,\tilde{\psi}(x), \label{calA0Mq} \end{equation} and the current on the right-hand side of (\ref{gausslaw}) by its quantum equivalent: \begin{equation} \partial_\nu {}^*\!F^{\mu\nu}(x) = - 4\pi \tilde{e} \bar{\tilde{\psi}}(x) \gamma^\mu \tilde{\psi}(x). \label{gausslawq} \end{equation} Extremizing then the action ${\cal A}^0$ with respect to $F_{\mu\nu}(x)$ and $\tilde{\psi}(x)$ under the constraint (\ref{gausslawq}) yields the equation (\ref{ELeq1}) together with the equation \cite{monact2}: \begin{equation} (i \partial_\mu \gamma^\mu - m) \tilde{\psi}(x) = - \tilde{e} \tilde{A}_\mu(x) \gamma^\mu \tilde{\psi}(x). \label{ELeq2q} \end{equation} Again, the equations obtained are exactly the dual for that of an electric charge as expected. Our next objective is now to generalize to nonabelian gauge theory to derive the equations of motion for monopoles. This is not possible using again as variables the field strength $F_{\mu\nu}(x)$ since, in contrast to the abelian theory, these are not gauge-invariant and has therefore also to be patched in the presence of a monopole. This difference with the abelian theory is of course very deep, and no simple modifications are likely to overcome this difficulty. For this reason, we turn to the loop formulation treated in Section 1 where the variables are constructed to be gauge invariant.\footnote{Actually, as defined, the variables $\Phi[\xi]$ and $F_\mu[\xi|s]$ depend on the gauge transformation at the reference point $P_0 = \xi^\mu_0$, but such a transformation is harmless as far as patching is concerned.} We write then the free field action ${\cal A}^0_F$ in terms of the loop variables $F_\mu[\xi|s]$ in the form given in (\ref{CalA0F}), and, by the Wu-Yang criterion, impose as a dynamical constraint the condition that the particle in ${\cal A}^0_M$ should carry a monopole charge. Now, according to (\ref{Gausslaw}), this condition can be written in terms of the loop curvature $G_{\mu\nu}[\xi|s]$ as: \begin{equation} G_{\mu\nu}[\xi|s] = -4\pi J_{\mu\nu}[\xi|s], \label{Gausslawj} \end{equation} where $J_{\mu\nu}[\xi|s]$ represents the monopole current. For a classical particle, this takes the form \cite{monact1}: \begin{equation} J_{\mu\nu}[\xi|s] = \tilde{g}\int d\tau\kappa[\xi|s] \epsilon_{\mu\nu\rho\sigma} \dot{\xi}^\rho(s) \frac{d Y^\sigma(\tau)}{d\tau} \delta^4(\xi(s)-Y(\tau)), \label{Jmunu} \end{equation} and for a Dirac particle, it takes the form \cite{monact2}: \begin{equation} J_{\mu\nu}[\xi|s] = \tilde{g} \epsilon_{\mu\nu\rho\sigma} [\bar{\tilde{\psi}}(\xi(s)) \Omega_\xi(s,0) \gamma^\rho \tau_i \Omega_\xi^{-1}(s,0) \tilde{\psi}(\xi(s))] \tau^i \dot{\xi}^\sigma(s). \label{Jmunuq} \end{equation} We shall return later to explain the meaning of the operator appearing in (\ref{Jmunuq}): \begin{equation} \Omega_\xi(s,0) =\omega(\xi(s)) \Phi_\xi(s,0). \label{Omega} \end{equation} which will be assigned a rather significant role in the applications of nonabelian duality to phenomenology. For the moment, it suffices only to note that it represents a frame rotation in internal symmetry space. As it stands, the variational problem posed is straightforward though somewhat complicated. Thus, incorporating the constraint (\ref{gausslawq}) into the action by means of the Lagrange multipliers $L_{\mu\nu}[\xi|s]$: \begin{equation} {\cal A}' = {\cal A}^0 \int \delta \xi ds\, {\rm Tr} \{ L^{\mu\nu}[\xi|s] \{G_{\mu\nu}[\xi|s] + 4\pi J_{\mu\nu}[\xi|s] \}\}, \label{calAp} \end{equation} and extremizing with respect to the variables $F_{\mu\nu}[\xi|s]$ and $Y^\mu(\tau)$ or $\tilde{\psi}(x)$, one obtains the equations of motion for the nonabelian monopole. We give here only the equations for the more interesting Dirac particle \cite{monact2}, namely: \begin{equation} \delta_\mu(s) F^\mu[\xi|s] = 0, \label{Polyakoveq} \end{equation} which, according to Polyakov \cite{Polyakov}, is the loop equivalent of the Yang-Mills field equation, and: \begin{equation} (i \partial_\mu \gamma^\mu - m) \tilde{\psi}(x) = - \tilde{g} \tilde{A}_\mu(x) \gamma^\mu \tilde{\psi}(x), \label{DiracYMeq} \end{equation} where the quantity $\tilde{A}_\mu(x)$ which couples to the Dirac particle like a dual potential can again be expressed in terms of the Lagrange multiplier $L_{\mu\nu}[\xi|s]$ but now as a rather complicated functional integral. The equations derived for nonabelian monopoles from the Wu-Yang criterion as outlined above are new since we know as yet of no nonabelian generalization to the abelian electric-magnetic duality by means of which we were able to infer in the abelian case the equations for the magnetic charge from those of the electric charge. However, we shall not examine the details of these new equations for the present, for we shall be able later to achieve a much better appreciation of them. What is of greater interest at the moment is a missing link in the derivation akin to that already encountered in the parallel derivation of the abelian equations, namely the question of the redundancy of the loop variables $F_\mu[\xi|s]$. What are the necessary constraints on $F_\mu[\xi|s]$ to ensure that one can recover from them the original $A_\mu(x)$, and have these constraints been satsified in the above treatment? To answer this question, one has to find a generalization to the nonabelian theory of the Poincar\'e Lemma which was used to answer a similar question in the abelian theory. Given the large number of loop variables, this at first sight looks very difficult. Fortunately, one is able to guess an answer by following a physical intuition based on what one has learned in the abelian case \cite{loop1}. We recall that by the Poincar\'e Lemma, what guaranteed the recovery of the potential at the point $x$ from $F_{\mu\nu}(x)$ was the condition (\ref{divfstar}), which means physically that at $x$ there is no magnetic charge. Could it not be then that even for the nonabelian theory, the absence of monopole charge at any point $x$ would guarantee the local existence of the gauge potential? If that is true, then our derivation is complete, for the dynamical constraint imposed (\ref{Gausslaw}) does imply that at all points outside the world-line of the monopole, the loop curvature $G_{\mu\nu}$ representing the monopole charge vanishes. This conjecture is found to be correct. One is indeed able to derive an Extended Poincar\'e Lemma \cite{loop1}, which states that given $F_\mu[\xi|s]$ depending on $\xi$ only up to the point $s$ as the notation $[\xi|s]$ in the argument implies, and satisfying the transversality condition: \begin{equation} F_\mu[\xi|s] \dot{\xi}^\mu(s) = 0, \label{transvers} \end{equation} then a gauge potential $A_\mu(x)$ can be locally constructed in regions of space-time where the loop space curvature $G_{\mu\nu}[\xi|s]$ vanishes. That there is a transversality condition (\ref{transvers}) on $F_\mu[\xi|s]$, is not surprising. As seen in (\ref{Fmuxis}), the longitudinal component of $F_\mu[\xi|s]$ represents the logarithmic variation of $\Phi[\xi]$ along the loop, just as that induced by a reparametrization of the loop. The condition (\ref{transvers}) thus corresponds to the parametrization independence of the phase factor $\Phi[\xi]$ and is the price one has to pay for working in {\it parametrized loop space} for removing the redundancy thus introduced, as was explained in Section 1. In terms of local field variables, as seen in (\ref{Fmuinx}), it corresponds to the statement that $F_{\mu\nu}(x)$ is antisymmetric in its indices $\mu$ and $\nu$. This additional constraint is relatively easy to take care of, affecting little the following arguments, and for this reason will largely be ignored. Apart then from a `proof' of the Extended Poincar\'e Lemma \cite{loop1,Book}, which we briefly outline below, the loop space formulation of nonabelian Yang-Mills theory introduced in Section 1, and the extension of the Wu-Yang criterion to the nonabelian Yang-Mills case, are now both complete. We shall indicate how one can arrive at an Extended Poincar\'e Lemma only for the case when there are no monopole charges anywhere in space. This will be sufficient to illustrate the idea. For the more general case with a number of isolated monopole charges, which is of interest to the above application of the Wu-Yang criterion, some more technical sophistication is required \cite{loop1} but the idea remains similar. First, we note that if all $G_{\mu\nu}[\xi|s]$ vanish, then $\Theta_\Sigma$ as defined in (\ref{Thetasigma}) equals the identity for all parametrized surfaces $\Sigma$. Recalling that $\Sigma$ is a loop in loop space, this implies that the following integral over an open path in loop space: \begin{equation} \Theta_\Sigma(t,0) = P_{t'} \exp ig \int_0^t dt' \int_0^{2\pi} ds F_\mu[\xi_{t'}|s] \frac{\partial \xi_{t'}^\mu(s)}{\partial t'}, \label{Theta0t} \end{equation} is path-independent and is a function only of the `end-point' $\xi_t$. We define then the phase factor as: \begin{equation} \Phi[\xi_t] = \Theta_\Sigma^{-1}(t,0), \label{Phiintheta} \end{equation} which by definition gives $F_\mu[\xi|s]$ as its logarithmic loop derivative. Second, by construction, one shows, from the fact that $F_\mu[\xi|s]$ is transverse and depends on $\xi$ only up to the point $s$, that $\Phi[\xi]$ obeys the composition law, namely that: \begin{equation} \Phi(C^2*C^1) = \Phi(C^2) \Phi(C^1), \label{complaw} \end{equation} where $C$ denotes the geometric loop in space-time corresponding to the parametrized loop $\xi$, and $C^2*C^1$ represents the loop obtained by first going around $C^1$ then $C^2$, as illustrated in Figure \ref{complawf}. \begin{figure} \centering \includegraphics{complaw.eps} \caption{Illustration for the composition law for loops} \label{complawf} \end{figure} Third, to every point $x$ in space-time, draw a straight line $\gamma_x$ joining the reference point $P_0 = \xi^\mu_0$ to $x$ and construct the phase factor $\Phi$ for the triangle formed by $\gamma_{x'}^{-1}* \gamma_{x'x}*\gamma_x$: \begin{equation} h(x',x) = \Phi(\gamma_{x'}^{-1}*\gamma_{x'x}*\gamma_x), \label{Phitriang} \end{equation} where $\gamma_{x'x}$ is the straight line joining $x$ to a neighbouring point $x'$. Define then the gauge potential $A_\mu(x)$ as: \begin{equation} A_\mu(x) = -\frac{i}{g} \lim_{\Delta \rightarrow 0} \{h(x',x) - 1\} \label{Amudef} \end{equation} for \begin{equation} x'^\nu = x^\nu + \Delta \delta_\mu^\nu. \label{xprime} \end{equation} Using the composition law as illustrated in Figure \ref{fanshape}, one shows that any $\Phi(C)$ can indeed be written in terms of this $A_\mu(x)$ in the usual manner. The recovery of the gauge potential from the loop variables $F_\mu[\xi|s]$ is then complete. \begin{figure} \centering \input{potconst.pstex_t} \caption{Construction of $\Phi(C)$ from $A_\mu(x)$} \label{fanshape} \end{figure} The above construction of $A_\mu(x)$ depends on the choice of the straight line $\gamma_x$ for each space time point $x$. This choice is not unique and can be replaced by any other choice of an open path for $\gamma_x$ linking the reference point to $x$ so long as neighbouring points are assigned neighbouring paths. A different choice for the paths $\gamma_x$, however, is seen to correspond just to a gauge transformation on $A_\mu(x)$ \cite{loop1,Book}. \setcounter{equation}{0} \section{Electric-Magnetic Duality and its Non\-abel\-ian Generalization} In vacuo, the field tensor $F_{\mu\nu}$ in Maxwell theory satisfies the equations: \begin{equation} \partial_\nu F^{\mu\nu}(x) = 0, \label{maxwell} \end{equation} and: \begin{equation} \partial_\nu {}^*\!F^{\mu\nu}(x) = 0, \label{maxwellstar} \end{equation} which are symmetric under the *-operation interchanging electricity with magnetism. This property of the theory is what is loosely known as electric-magnetic duality. In particular, just as the equation (\ref{maxwellstar}) guarantees that the field $F_{\mu\nu}$ is derivable from a potential $A_\mu(x)$, \begin{equation} F_{\mu\nu}(x) = \partial_\nu A_\mu(x) - \partial_\mu A_\nu(x), \label{fina} \end{equation} so also the equation (\ref{maxwell}) implies that there is a local `dual' potential $\tilde{A}_\mu(x)$ such that: \begin{equation} {}^*\!F_{\mu\nu}(x) = \partial_\nu \tilde{A}_\mu(x) - \partial_\mu \tilde{A}_\nu(x). \label{fstarinat} \end{equation} And just as $F_{\mu\nu}$ is invariant under the gauge transformation: \begin{equation} A_\mu(x) \longrightarrow A_\mu(x) + \partial_\mu \alpha(x), \label{atransf} \end{equation} for an arbitrary $\alpha(x)$, so is also ${}^*\!F_{\mu\nu}$ invariant under the gauge transformation: \begin{equation} \tilde{A}_\mu(x) \longrightarrow \tilde{A}_\mu(x) + \partial_\mu \tilde{\alpha}(x), \label{attransf} \end{equation} for another arbitrary $\tilde{\alpha}(x)$ which need have nothing to do with the $\alpha(x)$ in (\ref{atransf}). Hence, it follows that the theory is itself invariant under a doubled $U(1) \times \tilde{U}(1)$ gauge transformation where, apart from having the opposite parity (because of the *), the `magnetic' $\tilde{U}(1)$ as a group is the same as the `electric' $U(1)$. It is important to note, however, that although there are two independent gauge degrees of freedom as represented by (\ref{atransf}) and (\ref{attransf}), the physical degrees of freedom remain that given by either $F_{\mu\nu}$ or ${}^*\!F_{\mu\nu}$, but not both, since these two quantities are always related by the algebraic relation (\ref{fstar}) defining the Hodge star *. The logical steps in the above arguments are summarized in Chart \ref{chart1}. What happens when there are charges around? The Maxwell theory as usually formulated is then not dual symmetric because there are only electric but no magnetic charges. This is, however, merely a whim of nature; the theory itself still keeps the dual symmetry for there is in prinicple nothing (apart from experiment) to stop one introducing magnetic charges into the theory and write for the equations of motion: \begin{equation} \partial^\nu F_{\mu\nu}(x) = j_\mu(x), \label{maxwelle} \end{equation} and: \begin{equation} \partial^\nu {}^*\!F_{\mu\nu}(x) = \tilde{\jmath}_\mu{x}, \label{maxwellm} \end{equation} with $j_\mu$ and $\tilde{\jmath}_\mu$ as respectively the electric and magnetic currents. In the standard description in terms of the Maxwell field $F_{\mu\nu}$ and $A_\mu$, electric charges appear as sources of the field as per {(\ref{maxwelle}), while magnetic charges, as Dirac has taught us and as we have explained in Section 2, will appear as monopoles or topological obstructions in $A_\mu$. However, if one chooses to describe Maxwell theory in terms of the dual fields ${}^*\!F_{\mu\nu}$ and $\tilde{A}_\mu$, magnetic charges instead will appear as sources as per (\ref{maxwellm}) while electric charges will appear as monopoles. We can thus apply the Wu-Yang criterion to either electric or magnetic charges to derive their equations of motion. In particular, one sees that the standard Lorentz and Dirac equations for electric charges can be deduced as consequences of their topology when regarded as monopoles. The logical structure of duality for Maxwell theory with charges is summarized in Chart \ref{chart2}I. The question now is whether the above derivation is generalizable to nonabelian Yang-Mills theory. It is easy to see that a naive extension with the same *-operation as dual transform would fail. Although the field tensor satisfies in vacuo: \begin{equation} D^\nu F_{\mu\nu}(x) = 0, \label{Yangm} \end{equation} and, if $F_{\mu\nu}$ is derivable from a potential $A_\mu$, also the Bianchi identity: \begin{equation} D^\nu {}^*\!F_{\mu\nu}(x) = 0, \label{Yangmstar} \end{equation} these two equations, despite appearances, are not dual symmetric. The covariant derivative $D_\mu$ occurring in both (\ref{Yangm}) and (\ref{Yangmstar}) involves the potential $A_\mu$ of $F_{\mu\nu}$, whereas for the second equation to be dual to the first, it ought instead to involve a potential, say $\tilde{A}_\mu$, bearing the same relation to the dual field ${}^*\!F_{\mu\nu}$ as $A_\mu$ to $F_{\mu\nu}$, namely: \begin{equation} {}^*\!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{FstarinAt} \end{equation} and this $\tilde{A}(x)$ has no reason to be the same as $A_\mu$. Indeed, one does not even know whether such an $\tilde{A}_\mu$ exists. In contrast to the abelian theory where the equation of motion, namely the Maxwell equation (\ref{maxwell}), implies by the Poincar\'e Lemma that a potential exists for ${}^*\!F_{\mu\nu}$, the same is not true for the the nonabelian case. The equation of motion, which is here the Yang-Mills equation (\ref{Yangm}), does not imply a potential for the dual field ${}^*\!F_{\mu\nu}$. Worse in fact, since Gu and Yang \cite{Guyang} have exhibited many explicit counter-examples of solutions to the equation (\ref{Yangm}) for which no potential for ${}^*\!F$ exists. Does that mean then that electric-magnetic duality is not generalizable to Yang-Mills theory? Not necessarily, for it can be that if one defines the dual transform in a different way from *, then duality is retrieved. Supposing this to be true, let us first try to imagine what sort of properties such a generalized dual transform will need to possess. We would want, of course, the new transform to reduce to the Hodge star for the abelian theory as a special case, and to be reversible apart for a sign, like the Hodge star: ${}^*\!({}^*\!F) = - F$, so as to qualify as a 'duality'. But the most difficult property to satisfy is that discussed in the preceding paragraph of the existence of a dual potential $\tilde{A}_\mu$ and in order to recover that, we turn back to the abelian theory for inspiration. The reason why abelian duality works with the Hodge star lies in the fact that the source-free Maxwell equation (\ref{maxwell}) implies that there are no monoples in the dual field ${}^*\!F$, which is in physical terms exactly the condition required by the Poincar\'e Lemma for the existence of a gauge potential. Thus it appears that a crucial property of the new transform we seek is that the dual field should be so defined as to make the source-free Yang-Mills equation (\ref{Yangm}) equivalent to the statement that there are no monopole charges in the dual field. Given that the occurrence or otherwise of monopole charges in a nonabelian field was shown in our previous sections to be best expressed in loop space language, it is indicated that the same language be adopted also for constructing the generalized dual transform. These, then, are the leads which set us off to look for a generalized dual transform in loop space \cite{dualsymm}. To avoid obscuring the basic arguments with details, we shall begin with an outline of the construction and only return later to consider the `proof' of its validity as given in \cite{dualsymm}. First, in place of the $F_\mu[\xi|s]$ that we have employed up to now as variables in loop space, we introduce a new set: \begin{equation} E_\mu[\xi|s] = \Phi_\xi(s,0) F_\mu[\xi|s] \Phi_\xi^{-1}(s,0), \label{Emuxis} \end{equation} with $\Phi_\xi(s,0)$ defined as in (\ref{Phis2s1}). These $E_\mu[\xi|s]$ are not gauge invariant like $F_\mu[\xi|s]$ and may not be as useful in general but seem more convenient for dealing with duality. In particular, the condition that there be no monopole charge anywhere in space which, according to Section 3, is expressible in terms of $F_\mu[\xi|s]$ as the vanishing of the loop curvature $G_{\mu\nu}[\xi|s]$, is given in terms of $E_\mu[\xi|s]$ simply as: \begin{equation} \delta_\nu(s) E_\mu[\xi|s] - \delta_\mu(s) E_\nu[\xi|s] = 0. \label{curlE} \end{equation} On the other hand, the source-free condition (\ref{Yangm}) for Yang-Mills fields which was expressed by Polyakov \cite{Polyakov} as the vanishing of the loop divergence of $F_\mu[\xi|s]$ (\ref{Polyakoveq}) remains simply: \begin{equation} \delta^\mu(s) E_\mu[\xi|s] = 0 \label{divE} \end{equation} in terms of $E_\mu[\xi|s]$. Using these new variables $E_\mu[\xi|s]$ for the field, we now define their `dual' $\tilde{E}_\mu[\eta|t]$ as: \begin{eqnarray} & & \omega^{-1}(\eta(t)) \tilde{E}_\mu[\eta|t] \omega(\eta(t)) \\ &=& -\frac{2}{\bar{N}} \epsilon_{\mu\nu\rho\sigma} \dot{\eta}^\nu(t) \int\!\delta \xi ds\, E^\rho[\xi|s] \dot{\xi}^\sigma(s) \dot{\xi}^{-2}(s) \delta(\xi(s) -\eta(t)), \label{Dualtransf} \end{eqnarray} where $\omega(x)$ is a (local) rotation matrix tranforming from the frame in which the orientation in internal symmetry space of the fields $E_\mu[\xi|s]$ are measured to the frame in which the dual fields $\tilde{E}_\nu[\eta|t]$ are measured. The dual transform \ $\tilde{\ }$\ is by construction reversible apart from a sign, meaning that $\tilde{\tilde{E}} = - E$, and also reduces to the Hodge star for the abelian theory, as required. Further, it can be shown by differentiating (\ref{Dualtransf}) that: \begin{eqnarray} \lefteqn{ \omega^{-1}(\eta(t)) \{\delta_\nu(t) \tilde{E}_\mu[\eta|t] - \delta_\mu(t) \tilde{E}_\nu[\eta|t] \} \omega(\eta(t)) =} \nonumber\\ &&\!\!\!\!\!\!\!\!\! - \frac{1}{\bar{N}}\!\!\int\!\! \delta \xi ds \epsilon_{\mu\nu\rho\sigma} \{ \dot{\eta}^\beta(\!t\!) \dot{\xi}^\alpha(\!s\!)\!-\!\dot{\eta}^\alpha(\!t\!) \dot{\xi}^\beta(\!s\!) \} \delta_\rho(\!s\!) E^\rho[\xi|s] \dot{\xi}^{-2}(\!s\!) \delta(\xi(\!s\!)\!-\!\eta(\!t\!)). \label{curlEtdivE} \end{eqnarray} Now equation (\ref{curlEtdivE}) means that so long as the Yang-Mills field is source-free and therefore (\ref{divE}) is satisfied, then \begin{equation} \delta_\nu(t) \tilde{E}_\mu[\eta|t] - \delta_\mu(t) \tilde{E}_\nu[\eta|t] = 0, \label{curlEt} \end{equation} or that there are no monopoles in the dual field. Hence, by the Extended Poincar\'e Lemma discussed in Section 3, we deduce that a dual potential $\tilde{A}_\mu(x)$ will exist in that case. With this generalized dual transform in hand, we can now analyse the dual structure of nonabelian Yang-Mills theory. We shall begin with the pure theory with no source either in $E$ or $\tilde{E}$, for which the structure is summarized in Chart \ref{chart6}. Consider first the left half. The field action is ${\cal A}^0$ which can be expressed either in terms of the gauge potential $A_\mu(x)$ via the field tensor $F_{\mu\nu}(x)$, or in terms of the loop variable $E_\mu[\xi|s]$. Adopting $A_\mu(x)$ as variables, the theory can be developed along conventional lines. Extremizing ${\cal A}^0$ with respect to $A_\mu(x)$, one obtains the standard Yang-Mills equation (\ref{Yangm}), which can be written also in terms of $E_\mu[\xi|s]$ as (\ref{divE}). By (\ref{curlEtdivE}), however, this is equivalent to (\ref{curlEt}) and implies the existence of a dual gauge potential $\tilde{A}_\mu(x)$. This chain of arguments is summarized in the left-most column of Chart \ref{chart6}. Alternatively, adopting $E_\mu[\xi|s]$ as variables which are redundant, the constraint (\ref{curlE}) has to be imposed, which is incorporated via Lagrange multipliers $W_{\mu\nu}[\xi|s]$ into the action ${\cal A}$. Extremizing now ${\cal A}$ with respect to $E_\mu[\xi|s]$, one obtains an equation relating $E_\mu[\xi|s]$ to the Lagrange multipliers, from which relation the dual gauge potential $\tilde{A}_\mu(x)$ can then be constructed. This chain of arguments is summarized in the second column of Chart \ref{chart6}. Furthermore, the existence of the dual potential under gauge transformation of which the theory is invariant, together with the original invariance under gauge transformations of $A_\mu(x)$, implies that the theory has overall a doubled gauge symmetry $SU(N) \times \widetilde{SU}(N)$. The dual tranform (\ref{Dualtransf}) allows one to express the field action ${\cal A}^0$ also in terms of the dual loop variables $\tilde{E}_\mu[\xi|s]$. The fact now that the dual transform is reversible means that in going over into the dual representation, the steps outlined in the preceding paragraph can all be repeated with only occasional changes in signs as shown on the right half of Chart \ref{chart6}. Comparing this Chart with Chart \ref{chart1} shows a very close analogy with the abelian theory. Next, we turn to Yang-Mills theory with charges. In this case, one works exclusively with loop variables which admit an easier implementation of the Wu-Yang criterion. Consider first again the left-half of Chart \ref{chart7}. The free field-particle system as represented by the free action ${\cal A}^0$ is subjected to the constraint: \begin{equation} \delta_\nu(s) E_\mu[\xi|s] - \delta_\mu(s) E_\nu[\xi|s] = - 4\pi J_{\mu\nu}[\xi|s], \label{Jmunuconst} \end{equation} which is both the constraint imposed by the Wu-Yang criterion for deriving the dynamics of the monopoles and that required by the Extended Poincar\'e Lemma of Section 3 for removing the redundancy in the loop variables. Whether one is dealing with a classical or a Dirac particle is distinguished by the choice of the free particle action and the form of the current $J_{\mu\nu}[\xi|s]$, as shown in the separate columns of Chart \ref{chart7}. Incorporating the constraint (\ref{Jmunuconst}) into the action ${\cal A}$, and extremizing with respect to the field variables $E_\mu[\xi|s]$ and particle variables $Y^\mu(\tau)$ or $\psi(x)$, one obtains the equations of motion. One notes in particular the equation for the Dirac monopole, which couples the monopole to the dual potential $\tilde{A}(x)$ and takes a form exactly dual of that for a `colour' charge. That the monopole is so coupled to the field confirms that the dual potential constructed does play the expected role of a connection giving parallel phase transport for the monopole wave function. It also implies that the theory has an $\widetilde{SU}(N)$ symmetry corresponding to the phase of monopole wave functions in addition to the original $SU(N)$ gauge symmetry corresponding to the phase of the wave functions of `colour' charges, giving the theory thus in all an $SU(N) \times \widetilde{SU}(N)$ gauge symmetry. The equation for the classical monopole is perhaps less familiar but is also exactly dual to the so-called Wong equation \cite{Wong} for a classical `colour' charge. Again, given that the dual transform is reversible, one obtains a near symmetry apart from signs between the left and right halves of the Chart. This then completes our outline of nonabelian duality barring the derivations of some formulae that we have used, which, unfortunately, will involve some rather delicate operations in loop space, and are as yet far from rigorous. Even more than before, the lack of a general loop calculus, already mentioned in Section 1, is here strongly felt. We shall not go through all the details which the reader can find in our original publications \cite{dualsymm,dualsym}, but shall just pick up a few representative points for illustration. First, let us return to the variables $E_\mu[\xi|s]$ defined in (\ref{Emuxis}). Recalling the definition (\ref{Fmuxis}) of the Polyakov variable $F_\mu[\xi|s]$ one sees that $E_\mu[\xi|s]$ can be pictured as the bold curve in Figure \ref{Emuxisfig} where the phase factors $\Phi_\xi(s,0)$ in (\ref{Emuxis}) have cancelled parts of the faint curve representing $F_\mu[\xi|s]$. In contrast to $F_\mu[\xi|s]$, therefore, $E_\mu[\xi|s]$ depends really only on a ``segment'' of the loop $\xi$ from $s_-$ to $s_+$. Notice that the \begin{figure} \centering \input{segment.pstex_t} \caption{Illustration for $E_\mu[\xi|s]$} \label{Emuxisfig} \end{figure} $\delta$-function $\delta(s-s')$ inherent in our definition (\ref{loopderiv}) and (\ref{xiprime}) of the loop derivative $\delta_\mu(s)$ is represented in the figure as a bump function centred at $s$ with width $\epsilon = s_+ - s_-$. The reason for doing so is that, as usual in most functional formulations, our treatment here involves operations with the $\delta$-function which need to be ``regularized'' to be given a meaning. Our procedure is to take first the $\delta$-function as a bump function with finite width $\epsilon$ and height $h$, and afterwards take the zero width limit. In (\ref{loopderiv}) and (\ref{xiprime}), the limit $\epsilon \rightarrow 0$ with $\Delta = \epsilon h$ held fixed is taken first, to be followed by the limit $h \rightarrow 0$. For example, suppose we wish to take the loop derivative $\delta_\nu(s)$ of the quantity $E_\mu[\xi|s]$ at the same value of $s$. Clearly, a loop derivative has a meaning only if there is a segment of the loop on which it can operate. Therefore, to define this derivative, we shall first regard $E_\mu[\xi|s]$ as a segmental quantity dependent on the segment of the loop $\xi$ from $s - \epsilon/2$ to $s + \epsilon/2$. We then define the loop derivative $\delta_\nu(s)$ using the normal procedure on this segment, and afterwards take the limit $\epsilon \rightarrow 0$. Following this procedure, we have then by (\ref{Emuxis}) and (\ref{Fmuxis}): \begin{equation} \delta_\nu(s') E_\mu[\xi|s] = \Phi_\xi(s,0) \{ \delta_\nu(s') F_\mu[\xi|s] + ig \theta(s-s') [F_\nu[\xi|s'], F_\mu[\xi|s]] \} \Phi_\xi^{-1}(s,0), \label{difEmuxis} \end{equation} where $\theta(s)$ is the Heaviside $\theta$-function, so that: \begin{equation} G_{\mu\nu}[\xi|s] = \Phi_\xi^{-1}(s,0) \{\delta_\nu(s) E_\mu[\xi|s] - \delta_\mu(s) E_\nu[\xi|s]\} \Phi_\xi(s,0). \label{GmunuinE} \end{equation} This shows that the absence of monopoles, which was given in terms of the variables $F_\mu[\xi|s]$ by the vanishing of the loop curvature $G_{\mu\nu}[\xi|s]$ is here translated in terms of $E_\mu[\xi|s]$ to read as the vanishing of the `curl' of $E_\mu[\xi|s]$ as was claimed in (\ref{curlE}) above. Next, as another example, let us check that our generalized dual transform (\ref{Dualtransf}) does reduce to the Hodge star when the theory is abelian. To see this, we let the segmental width of $\tilde{E}_\mu[\eta|t]$ in (\ref{Dualtransf}) go to zero so that we can use the formula given in (\ref{Fmuinx}) for $F_\mu[\xi|s]$ to write the left-hand side in terms of local quantities: \begin{equation} \omega^{-1}(x) \tilde{F}_{\mu\nu}(x) \omega(x) = -\frac{2}{\bar{N}} \epsilon_{\mu\nu\rho\sigma} \int\! \delta\xi ds \,E^\rho[\xi|s] \dot{\xi}^\sigma(s) \dot{\xi}^{-2}(s) \delta(x - \xi(s)). \label{reducedual1} \end{equation} To evaluate the right-hand side, we recall that our procedure is to do the integral before taking the width of the segment in $E_\mu[\xi|s]$ to zero. In other words, within the integral, the loop $\xi$ can still vary by a $\delta$-functional bump as illustrated in Figure \ref{reducedualfig} (a). \begin{figure} \centering \input{abdual.pstex_t} \caption{Illustration for the Integrand in Dual Transform} \label{reducedualfig} \end{figure} For such a $\xi$, $E_\mu[\xi|s]$, which is obtained by making a $\delta$-functional variation along the direction $\mu$, will take on the shape depicted in Figure \ref{reducedualfig} (b). This last figure can be expressed as the product of three factors, namely Figures \ref{reducedualfig}(c),(d),(e) in the order indicated. In the abelian theory, the ordering of the factors is unimportant so that the factors of Figures (c) and (e) cancel in the limit when the segmental width $\epsilon \rightarrow 0$, leaving only the factor of Figure (d), which can as before be expressed as $F_{\mu\alpha}(\xi(s)) \dot{\xi}^\alpha(s)$, giving: \begin{eqnarray} \tilde{F}_{\mu\nu}(x) & = & -\frac{2}{\bar{N}} \epsilon_{\mu\nu\rho\sigma} \int\! \delta\xi ds \,F^{\rho\alpha}(\xi(s)) \dot{\xi}_\alpha(s) \dot{\xi}^\sigma(s) \dot{\xi}^{-2}(s) \delta(x-\xi(s)) \nonumber \\ & = & - \mbox{\small $\frac{1}{2}$} \epsilon_{\mu\nu\rho\sigma} F^{\rho\sigma}(x). \label{reducedual} \end{eqnarray} This is just the Hodge star relation if we identify $\tilde{F}_{\mu\nu}(x)$ with $\mbox{\mbox{}$^*\!$} F_{\mu\nu}(x)$. On the other hand, for a nonabelian theory, the factors of Figures \ref{reducedualfig} (c) and (e) cannot be commuted through the factor of Figure (d) so that the above reduction to the Hodge star relation will not go through. This last statement is important for otherwise our claim of nonabelian duality under the generalized dual transform (\ref{Dualtransf}) would be in contradiction with the result of Gu and Yang quoted above \cite{Guyang}. These two examples, we hope, should give a taste of the sort of arguments one had to go through to justify the results claimed above. They serve to illustrate the level of rigour one is able at present to achieve, which is unfortunately not as high as one could wish. Except for this reservation, one could claim that a nonabelian generalization of electric-magnetic duality is now established. \setcounter{equation}{0} \section{'t~Hooft's Order-Disorder Parameters} As in electromagnetism, the dual symmetry in nonabelian Yang-Mills theory is established only for classical fields. In contrast to electromagnetism, however, where the classical theory has already a wide range of applications, Yang-Mills theory applied to physics involves almost always the quantum theory. The construction of a quantized version of the above results looks difficult and only the most tentative of beginnings of an attempt have so far been made \cite{rudiments}. However, one very useful result for the quantum theory has already been derived which is the subject of this section. In his famous study of the confinement problem in nonabelian gauge theories, 't~Hooft \cite{thooft} introduced 2 loop-dependent operators $A(C)$ and $B(C)$ which satisfy the following commutation relation: \begin{equation} A(C) B(C') = B(C') A(C) \exp(2\pi i l/N) \label{ABcomm} \end{equation} for $su(N)$ symmetry and any 2 spatial loops $C$ and $C'$ with linking number $l$ between them. $A(C)$ is given explicitly as: \begin{equation} A(C) = {\rm Tr} \left[P \exp ig \oint_C A_i(x) dx^i \right], \label{AC} \end{equation} and in the words of 't~Hooft measures the magnetic flux through $C$ while creating electric flux along $C$. On the other hand, $B(C)$ measures the electric flux through $C$ while creating magnetic flux along $C$, and plays thus an exactly dual role to $A(C)$. For lack of a dual potential, however, $B(C)$ was not given a similar explicit expression to (\ref{AC}). Now, in the preceding section, one claims that there does indeed exist a dual potential $\tilde{A}_\mu$ for nonabelian gauge fields, although by duality is now meant no longer the Hodge star but a more complicated transform which reduces to the Hodge star only for the abelian theory. That being the case, one ought to have: \begin{equation} B(C) = {\rm Tr} \left[ P \exp i \tilde{g} \oint_C \tilde{A}_i(x) dx^i \right] \label{BC} \end{equation} as the explicit expression for $B(C)$, with the dual coupling $\tilde{g}$ related to $g$ by a Dirac quantization condition. In other words, starting from the formulae (\ref{AC}) and (\ref{BC}) with $A_\mu$ nd $\tilde{A}_\mu$ related in the manner detailed in the preceding section, one ought to be able to deduce the commutation relation (\ref{ABcomm}) required by `t~Hooft. The success in doing so would be an important check on the consistency of the proposed framework for duality. It also means that the duality as defined above accords with what 't~Hooft called duality so that the important results he derived are applicable to the present case. That (\ref{ABcomm}) is indeed satisfied for (\ref{AC}) and (\ref{BC}) is shown in \cite{dualcomm}. To appreciate easier how this obtains, recall first how the parallel assertion to (\ref{ABcomm}) can be deduced in the abelian case: \begin{equation} \left[ ie \oint_C A_i(x) dx^i,\; i \tilde{e} \oint_{C'} \tilde{A}_i(x) dx^i \right] = 2 \pi l. \label{abcomm} \end{equation} Using Stokes' theorem, the second integral over $C'$ in (\ref{abcomm}) can be written as a surface integral, thus: \begin{equation} - i \tilde{e} \int\!\int_{\Sigma_{C'}} {}^*\!F_{ij}\, d\sigma^{ij}, \label{intfstar} \end{equation} or, by the definition of the Hodge star, in terms of the electric field strength ${\cal E}_i = F_{0i}$ as: \begin{equation} i \tilde{e} \int\!\int_{\Sigma_{C'}} {\cal E}_i\, d\sigma^i, \label{intcalE} \end{equation} where $\Sigma_{C'}$ is some surface both spanning and bounded by $C'$. Consider first the simple case for linking number 1 between $C$ and $C'$. The loop $C$ in that case will intersect the surface $\Sigma_{C'}$ at some point $x_0$. ($C$ may of course intersect $\Sigma_{C'}$ more than once, but the extra intersections occurring pairwise with opposite orientations, their contributions to the commutator will all cancel, leaving in effect just one intersection.) Except at this point $x_0$, all points on $C$ are spatially separated from points on $\Sigma_{C'}$ so that, using the canonical commutation relation between $A_i(x)$ and ${\cal E}_i(x)$: \begin{equation} [{\cal E}_i(x), A_j(x')] = i\, \delta_{ij}\, \delta(x-x') \label{eacomm} \end{equation} we have: \begin{equation} \left[ ie \oint_C A_i(x) dx^i, i \tilde{e} \int \!\! \int_{\Sigma_{C'}} {\cal E}_j d \sigma^j \right] =i e \tilde{e}, \label{expcomm} \end{equation} which by the Dirac quantization condition (\ref{diraccond})(see footnote for the rationalized couplings adopted here) gives the answer (\ref{abcomm}) for $l = 1$ as required. In case $C'$ winds around $C$ more than once, say $l$ times, then $C$ will intersect $\Sigma_{C'}$ at effectively $l$ points for each of which the above applies, so that (\ref{abcomm}) still remains valid. What happens when we generalize to the nonabelian case? Then $A(C)$ and $B(C)$ are each a trace of an ordered product of noncommuting factors, {\it not}, in spite of appearances (due to the somewhat misleading standard notation), an exponential of a line integral for which Stokes' Theorem applies, so that the above arguments no longer work. Nevertheless, one finds that one may still associate with each a surface in an analogous fashion \cite{dualcomm}. Take $B(C')$, for example. The phase factor: \begin{equation} \tilde{\Phi}(C') = P \exp i \tilde{g} \oint_{C'} \tilde{A}_i dx^i \label{PhiCp} \end{equation} of which $B(C')$ is the trace, can be written, according to \cite{dualsymm}, as: \begin{equation} \tilde{\Phi}(C') = \prod_{t = 0 \rightarrow 2 \pi} (1 - i \tilde{g} \tilde{W}[\eta|t]) \sim \prod_{t = 0 \rightarrow 2 \pi} \exp (- i \tilde{g} \tilde{W}[\eta|t])\,, \label{PhiCpp} \end{equation} where $\eta$, for $t = 0 \rightarrow 2 \pi$, is a parametrization of $C'$, and $\tilde{W}$ is a segmental quantity the `loop gradient' of which, $\delta_\mu \tilde{W}$, as indicated in Chart \ref{chart6}, is the field variable $\tilde{E}_\mu$. One can thus write symbolically: \begin{equation} \tilde{W}[\eta|t] = \int_{\eta_0}^{\eta(t)} \delta \eta^{' \nu}(t) \,\tilde{E}_\nu[\eta'|t], \label{Wtildeint} \end{equation} where the integral in (\ref{Wtildeint}) denotes a `segmental' integral along some path from a reference point $\eta_0$ to the point $\eta(t)$, so that in ordinary space, this path appears as a ribbon. Piecing such ribbons together as $\eta(t)$ moves along $C'$ in (\ref{PhiCpp}), one obtains a surface $\Sigma_{C'}$ spanning over and bounded by $C'$ as suggested. In as much as the reference point $\eta_0$ and the path joining it to $\eta(t)$ are both arbitrary for (\ref{Wtildeint}) to hold, one can choose $\Sigma_{C'}$ to be completely space-like. This surface will again intersect the loop $C$ of $A(C)$ at some point $x_0$. (Previous remarks in the abelian case about multiple intersections and higher linking numbers between $C$ and $C'$ will still apply and need not be repeated.) The remaining arguments for deriving (\ref{ABcomm}) \cite{dualcomm} then follow along much the same lines as for (\ref{abcomm}) in the abelian case although they are naturally more complicated because of the noncommutative quantities involved. Thus, starting from the formula (\ref{Wtildeint}), we write $\tilde{E}[\eta|t]$ via the dual transform (\ref{Dualtransf}) given in the last section in terms of its dual $E[\xi|s]$, which is then related to the usual Yang-Mills local field tensor $F_{\mu\nu}(x)$. The commutation with the operator $A(C)$ using the canonical commutation relation between $A_i^\alpha(x)$ and the `electric' field strengths ${\cal E}_i^\beta(x) = F_{0i}^\beta$ then picks out the one point of intersection $x_0$ between the loop $C$ and the surface associated to $B(C')$ as described in the last paragraph. Again, the Dirac quantization condition comes in in relating the dual charge $\tilde{g}$ to $g$ giving rise to the factor $2\pi$ in the exponent on the right-hand side of (\ref{ABcomm}). For details, the reader is referred to the original reference \cite{dualcomm}. Although the derivation of (\ref{ABcomm}) represents but a small step on the road to quantizing the classical formalism developed in the previous sections, it is of much practical importance in allowing one to apply 't~Hooft's far-reaching results \cite{thooft}, which have been used to good effect in the Dualized Standard Model \cite{DSMrph98}. Furthermore, as a by-product, one has obtained in (\ref{BC}) an explicit formula for the operator $B(C)$ which may be useful in the problem of confinement. Although other, and presumably equivalent, explicit formulae for $B(C)$ have previously been suggested \cite{Zeni}, they are, in contrast to (\ref{BC}), usually given as dependent not only on the loop $C$ but also on the particular surface spanning it. \setcounter{equation}{0} \section{Concluding Remarks} Starting from a loop space formulation of Yang-Mills theory which helps to clarify the role that monopoles play in defining the dynamics and in guaranteeing the existence of gauge potentials, a generalization of electric-magnetic duality to the nonabelian theory is derived. The result implies in particular that the gauge symmetry is doubled, say from $SU(N)$ to $SU(N) \times \widetilde{SU}(N)$, where the second factor has opposite parity to the first. However, the physical degrees of freedom, of course, remain the same, and the theory can be described in terms of either the usual Yang-Mills potential $A_\mu(x)$ or a dual potential $\tilde{A}_\mu(x)$. The theory then admits both nonabelian `electric' and nonabelian `magnetic' charges, where the former appear as sources of $A_\mu$ but as monopoles of $\tilde{A}_\mu$, while the latter appear as monopoles of $A_\mu$ but sources of $\tilde{A}_\mu$. Although these results have been derived only for classical fields, one result is known for the quantum theory, namely that the Dirac phase factors (or Wilson loops) constructed out of $A_\mu$ and $\tilde{A}_\mu$ satisfy the 't~Hooft commutation relations, so that his results apply. Hence one concludes, in particular, that if $SU(N)$ is in the confined phase then its dual $\widetilde{SU}(N)$ is in the Higgs phase, and vice versa. When applied to the Standard Model with symmetry $su(3) \times su(2) \times u(1)$, one concludes first that each symmetry has its dual and particles can in principle carry an `electric' as well as a `magnetic' charge of each symmetry. Secondly, using the corollary stated above to `t~Hooft's result, one concludes that since colour $su(3)$ is confined, dual colour $\widetilde{su}(3)$ is broken, and since electroweak $su(2)$ is in the Higgs phase, dual electroweak $\widetilde{su}(2)$ is in the confined phase. If we assume that these effects have correspondence in Nature, then their manifestations should lead to very interesting consequences. A possible scenario for the physical realization of these effects is the subject of our companion paper \cite{DSMrph98}. \vspace{1cm} \noindent {\large \bf Acknowledgement} \vspace{.3cm} It is a pleasure to recall our most enjoyable collaborations at different periods with respectively Peter Scharbach and Jacqueline Faridani from which most of the material reviewed in this paper is drawn. The present version is based on some lectures given to the lattice gauge theory group at Pisa at the invitation of Adriano Di Giacomo. We are much indebted to him and to Angela Buonocorso and Kenichi Konishi for their kind hospitality during this visit. \clearpage \begin{chart} \vspace*{-2cm} \centerline{ \resizebox{!}{23cm} {\includegraphics{chart1.ps}} } \vspace*{-1.5cm} \caption[]{Pure Electromagnetism.} \label{chart1} \end{chart} \clearpage \begin{chart} \vspace*{-2cm} \centerline{ \resizebox{!}{23cm} {\includegraphics{chart2.ps}} } \vspace*{-1.5cm} \caption[]{Electromagnetism with charges.} \label{chart2} \end{chart} \clearpage \begin{chart} \vspace*{-2cm} \centerline{ \resizebox{!}{23cm} {\includegraphics{chart6.ps}} } \vspace*{-1.5cm} \caption[]{Pure Yang--Mills Theory.} \label{chart6} \end{chart} \clearpage \begin{chart} \vspace*{-2cm} \centerline{ \resizebox{!}{23cm} {\includegraphics{chart7.ps}} } \vspace*{-1.5cm} \caption[]{Yang--Mills with charges.} \label{chart7} \end{chart} \clearpage
\section{Introduction} Observations of spatial fluctuations in the cosmic microwave background (CMB) radiation are fundamental to our understanding of structure formation in the universe as they mark the earliest observable imprints of massive gravitational structures (see e.g. review by White, Scott \& Silk 1994). The distribution and amplitude of anisotropies in the CMB sky over scales from degrees to arcminutes can be used to discriminate between competing cosmological theories. On scales of 0.2$^{\circ}$ - 2$^{\circ}$, inflationary models predict that increased power should be seen in the CMB sky due to scattering of photons during acoustic oscillations of the photon-baryon fluid at recombination. Detection and study of these acoustic or `Doppler' peaks in the power spectrum is one of the primary goals of CMB astronomy. Furthermore, the amplitudes and angular scales of the acoustic peaks provide powerful constraints on basic cosmological parameters including $H_0$ and $\Omega$. The first clear indication of a downturn in the power spectrum on sub-degree scales was provided by the detection of CMB power by the Cambridge Cosmic Anisotropy Telescope (CAT) on scales of about half a degree (Scott et al. 1996; Paper I). The CAT is a three-element interferometer operating at frequencies between 13 and 17~GHz (Robson et al. 1993). It is sensitive to structure on angular scales of about $10'$ to $30'$ over a field of view covering 2$^{\circ}$$\times 2$$^{\circ}$\ (primary beam FWHM). This paper describes observations of a second field observed with CAT and the detection of CMB anisotropies within it, at levels consistent with measurements in the first field. \section{Observations and Data Reduction} Observations have been made with the CAT of a blank field centred at the position 17~00~00 $+64$~30~00 (B1950), which we call `CAT2'. The field was chosen to be relatively free from strong radio sources at frequencies up to 5~GHz (Condon, Broderick \& Seielstad 1989), lie at high Galactic latitude ($b>30$$^{\circ}$) and away from known Galactic features (e.g. the North Polar spur). In periods during the interval 1995 March -- 1997 June, CAT observed the CAT2 field at three frequencies, 13.5, 15.5 and 16.5 GHz. The three baselines of the array were scaled with frequency to achieve the same resolution at each frequency. The resulting synthesised beams measured $20'$ FWHM in right ascension and $24'$ in declination. The primary beam of the telescope, due to the symmetric nearly-Gaussian envelope beams of the three horn-reflector antennas, has a FWHM of 1.96$^{\circ}$\ at 15.5~GHz, scaling inversely with frequency. The telescope observes in two orthogonal linear polarisations (which rotate on the sky as it tracks), and has a system noise temperature of 50~K. An observing bandwidth of 500~MHz was used. Amplitude and phase calibrations were carried out daily with observations of Cas~A (using the flux scale of Baars et al. 1977), and cross checked periodically by observations of other 15-GHz calibrators from the VLA list (Perley 1982). Typical uncertainties in flux scaling are less than 10\%. Observations were generally carried out at night (about 80\% of the data) or pointing $>90^{\circ}$ away from the sun to avoid possible solar interference; no extra emission from the moon was detected at this declination. The CAT2 data were reduced using the same method as for the first field, CAT1, as described in detail by O'Sullivan et al. (1995). Phase rotation and flux calibration were applied first, and the data were then edited and analysed using standard tasks in AIPS. Excessively noisy data (reflecting periods of poor weather) were excised --- the visibility amplitudes were all checked by eye for periods where they regularly exceeded the mean value by more than 3$\sigma$ and then these data ($\pm 1$ hour) were removed from all baselines. Across the remaining dataset, individual visibilities with amplitudes exceeding the 3$\sigma$ threshold were also excluded. In total, about 40\% of the data were excluded by this process leaving 370, 310 and 1340 hours of good data at 13.5, 15.5 and 16.5~GHz respectively. Since the atmospheric coherence time is very short (10s) compared with the total integration time, any remaining atmospheric signals will be distributed uniformly across the synthesised map as noise, unlike true sky signals which will be modulated by the envelope beam pattern. The efficacy of atmospheric filtering for the CAT interferometer has already been demonstrated (Robson et al. 1994). CAT's sidelobe response and lack of crosstalk and correlator offsets are discussed in O'Sullivan et al. (1995). No radio interference was seen. \section{Source Subtraction} Radio sources contributing significantly to the CAT2 image were identified and monitored at 15~GHz using the Ryle Telescope (RT). Five RT antennas were used in a configuration giving a synthesised beam of $30''$ FWHM. The RT has an instantaneous field of view of $6'$ FWHM, but for these observations is used in a rastering mode which covers $30' \times 30'$ in 12 hours to a typical flux sensitivity of 1.5~mJy/beam. To detect sources within the the CAT2 field, the central 2$^{\circ}$$\times 2$$^{\circ}$\ area was scanned with the RT in raster mode in sixteen days. A source list was then compiled, including sources listed in the Green Bank 4.85-GHz survey (Condon et al. 1989). Pointed observations with the RT were then made, and repeated regularly to check for variability over the whole period of the CAT observations. In all, twenty-nine sources were detected, the faintest having a flux density of 4.5~mJy at 15~GHz. Flux densities at 13.5 and 16.5~GHz were extrapolated using spectral information obtained from lower frequency surveys (i.e. 4.85~GHz, Condon et al. 1989, and 1.4~GHz, Condon et al. 1998) where available. A flat spectrum between 13.5 and 16.5~GHz was assumed for three sources without other data. After correcting for CAT primary beam attenuation, the corresponding flux densities were subtracted as point sources from the visibility data. Totals of 33\,mJy at 13.5, 27\,mJy at 15.5 and 20\,mJy at 16.5 GHz were subtracted. The robustness of the source-subtraction procedure is illustrated in O'Sullivan et al. The strongest source in the field, at position 16~45~32~+63~35~29 (B1950), was highly variable (by factors of up to two over periods of a few days) and so was monitored regularly with the RT; no residuals at the source position remain after subtraction of the variable source. The successful subtraction of this source, 1$^{\circ}$ away from the pointing centre and clearly visible at the same position in all three maps, shows that CAT phase calibration and pointing are accurate. \section{The Images} The resulting source-subtracted images each show excess signal in the central 2$^{\circ}$$\times 2$$^{\circ}$\, falling away at larger radii as expected from the antenna envelope beam. These are displayed in Figure \ref{cat2ss}. The instrumental noise levels (measured directly from the visibilities) for each source-subtracted image are 6.1, 6.5 and 3.5 mJy/beam rms for 13.5, 15.5 and 16.5 GHz respectively. At the three frequencies, excess powers above the intrinsic noise level in the central 2$^{\circ}$$\times 2$$^{\circ}$\ area of $7.8 \pm 1.0$, $8.2 \pm 1.0$ and $8.3 \pm 0.5$ mJy/beam rms were found for the source-subtracted 13.5, 15.5 and 16.5~GHz images respectively. Checks were made by splitting the data in time (consistent excesses were seen), polarisation (no excess power was visible on polarisation difference maps) and cross-correlating the source-subtracted image with a reconstructed map of the radio sources as in O'Sullivan et al. (no correlation was found); most importantly, maps were made by correlating orthogonal polarisations and these gave the {\em same} noise levels as those above. To attempt to remove the instrumental response and thereby illustrate the distribution of features on the sky, we have co-added the data from the three frequencies weighted as $\nu^{2}$ and CLEANed the final image. This is shown in Figure \ref{cat2cln} and shows the presence of significant features in the central region as well as the diminution of the telescope sensitivity in accordance with the envelope beam. The strongest feature in the CAT2 images is a negative one centred at position 17~05~29~+64~47~37 (B1950) reaching $-39$ mJy at 16.5 GHz (i.e. 5$\sigma$, relative to the rms excess power in the sky). For comparison, the strongest positive feature in the 16.5-GHz dirty map has a peak flux density of 26 mJy at position 16~55~55 +63~49~47 (B1950). The negative feature can be seen at all three frequencies (most clearly in the 13.5 and 16.5-GHz images in Figure \ref{cat2ss}), both before and after source subtraction and in different time cuts, and its spectrum is consistent with that of CMB radiation. For example, even before source subtraction the hole reaches $-30$ mJy in the 16.5-GHz image, and only three weak sources ($S_{16.5} < 10$ mJy) lie within $40'$ of it. No obvious Galactic structures at the position of the negative feature were visible in IRAS 100\,$\mu$m (Wheelock et al. 1994) or H\,{\sc i\/} 21-cm sky survey images (Lockman \& Dickey 1995) or the $100\mu$m map of Schlegel, Finkbeiner \& Davis 1998. Indeed, there is no resemblance at all between any of these images of the CAT2 region and the CAT2 image itself. On the scales sampled by CAT, the negative feature is unlikely to have been caused by a Sunyaev-Zel'dovich (S--Z) effect (Sunyaev \& Zel'dovich 1972; Rephaeli 1995) towards a single massive cluster. Any cluster which might produce such an effect would have to be nearby (subtending a large angle on the sky, filling the CAT beam) and/or very massive to produce such a strong signature. The strongest S--Z decrements measured towards nearby clusters at 15~GHz with the RT (Grainge {\it et~al.} \ 1996; Jones {\it et~al.} \ 1993) are about $-0.5$\,mJy on arcminute scales. Observing a similar cluster at $z \sim 0.2$ (e.g. gas mass $10^{14} M_{\odot}$, with a King profile of core radius $\sim 300$ kpc, truncated at about ten core radii) with the larger CAT beam of $30'$ would produce a similar decrement amplitude, due to the balance between the effects of beam dilution and sampling cluster gas out to a larger radius. In order to maximise the S--Z signal in a $30'$ beam with the minimum gas mass, a nearby cluster ($z \approx 0.05$) with $10^{15} M_{\odot}$ of gas would be required, which is patently not observed. A system at any higher redshift would require an even higher mass because of beam dilution. A central portion of the CAT2 field has been observed serendipitously with ROSAT. The direction of the negative CMB feature lies close to the edge of the PSPC field ($45'$ off axis) and is partially shadowed by the support structure of the PSPC detector. An X-ray source (heavily distorted by the PSPC point spread function) is found nearby (at the position 17~05~49 +64~45~50, B1950) but its X-ray spectrum does not fit a thermal model for any reasonable temperature, and is fit much better by a power law, implying it is not a cluster (but perhaps an AGN or Galactic source). Four clusters (Abell 2246 and two others at $z = 0.25$ plus one at $z = 0.44$) --- and an optically luminous QSO (HS~1700+6416) --- have been detected by ROSAT and lie within the central $10'$ of the CAT2 field (Reimers et al. 1997; Vikhlinin et al. 1998). All four clusters are fainter than $10^{44}$ erg\,s$^{-1}$ at X-ray energies 0.4--2.0 keV and none is apparent in the CAT images. Finally, we emphasise that the presence of a 5$\sigma$ feature in a single CAT field should not be construed as evidence for non-Gaussianity of the CMB fluctuations: the sidelobe structure of the synthesised beam of the three-element CAT is significant, and analysis should be carried out in the aperture plane --- see Section \ref{analysis}. As a check, we have simulated CAT images given standard CDM-based realisations of CMB structure (using the actual CMB and Galactic mean power measured by CAT), and find that features which appear as strong as $5\sigma$ occur in about 10\% of cases. We return to this point in Section \ref{analysis}. \section{Determination of the CMB Component} \label{analysis} Due to the limited range of baselines and resulting sparse sampling of the $u-v$ plane by CAT, a statistical likelihood analysis was employed to estimate the relative contributions of CMB and Galactic components given the three-frequency CAT data. A Bayesian likelihood method was used, as in Paper I, as described by Hobson, Lasenby \& Jones (1995). This method uses the complex visibility data directly in the calculation of the likelihood function to avoid the problem of the long-range correlations present between different resolution elements in the image plane. As described in Paper I, power was estimated in two independent annular bins centred on spherical harmonic multipole values of $\ell =410$ and $\ell =590$ and with widths equivalent to the diameter of the antenna function, thus together spanning the range $\ell\approx 330-680$. The noise-weighted centroid positions for data in each bin are $\ell =422$ and $\ell =615$. CMB and Galactic signals were modelled as independent Gaussian distributions with a power-law spectrum ($S \propto \nu^{\alpha}$), the CMB with a fixed spectral index of $+2$ in flux density and the Galactic spectral index variable between 0 and $-1$ (as expected for Galactic free-free and synchrotron emission). After marginalising over the Galactic parameters, this analysis confirmed that the bulk of the power in the 16.5~GHz map (Fig. \ref{cat2ss}) arises from CMB fluctuations. The CMB component was clearly distinguished from possible Galactic contamination in the $\ell =422$ bin; $\Delta T/T = 2.1^{+0.4}_{-0.5} \times 10^{-5}$ was estimated for the CMB signal, compared with only $\Delta T/T = 0.8^{+0.5}_{-0.8} \times 10^{-5}$ for any Galactic component. The uncertainties quoted are $1\sigma$ values. The CMB--Galaxy separation was less certain in the $\ell =615$ bin --- a Galactic contribution of $\Delta T/T = 1.0^{+0.7}_{-1.0} \times 10^{-5}$ was given by the likelihood analysis, which is equivalent to an upper limit for the marginalised CMB power of $\Delta T/T < 2.0 \times 10^{-5}$ ($1\sigma$). These values are plotted in Figure \ref{deltat}. The average values of $\Delta T/T$ in the CAT2 field agree with the CAT1 result (Paper I) within $1\sigma$; for comparison the measured CMB powers in the CAT1 field were $\Delta T/T =1.9^{+0.5}_{-0.5} \times 10^{-5}$ ($\ell = 420$) and $\Delta T/T =1.8^{+0.7}_{-0.5} \times 10^{-5}$ ($\ell = 590$). We have also investigated how much of the CMB power is not associated with the strong negative feature discussed in Section 4. We removed the dip as a point source from the visibilities and repeated the above analysis. Significant power remained, at around half of the level with the negative feature included. \section{Estimation of Cosmological Parameters using new CAT results} The CAT2 points, taken in conjunction with the results from the Saskatoon experiment \cite{sask}, provide further evidence for a downturn in the CMB power spectrum for $\ell \mbox{$\stackrel{>}{_{\sim}}$} 300$. To assess the implications for cosmological parameters using current CMB data sets, we have extended the analysis presented previously by Hancock {\it et~al.} \ (1998). This analysis used a statistically independent subset of the current data and carried out $\chi^2$ fitting for a range of cosmological models and parameters. The extensions carried out for the present work were (a) the inclusion of the CAT2 point at $\ell = 422$; (b) the inclusion of new points from experiments Python (Python III, Platt {\it et~al.} \ 1997), MSAM (the 2nd and 3rd flights, Cheng {\it et~al.} \ 1996, 1997, Ratra {\it et~al.} \ 1997), ARGO (Aries+Taurus region, Masi {\it et~al.} \ 1996), FIRS \cite{firs} and BAM \cite{bam}, and with the latest calibration correction to the Saskatoon data (i.e. increased by 5\%, Leitch, private communication); (c) consideration of a wider class of cosmological models, all treated using exact power spectra rather than the generic forms assumed in Hancock {\it et~al.} \ (see also Rocha 1997; Rocha {\it et~al.} \ in preparation). The formalism and approach are otherwise the same as in Hancock {\it et~al.} \ (1998), to which the reader is referred. The cosmological models considered were: \begin{enumerate} \item Flat models with $\Lambda=0$, and with a range of spectral tilts (i.e. $n \neq 1$); \item Flat models with $\Lambda \neq 0$, and a range of spectral tilts; \item Open models with $\Lambda=0$ and open-bubble inflation spectrum \cite{sugi,ratra2,kami}. \end{enumerate} In cases (i) and (ii), nucleosynthesis constraints, with $0.009 \leq \Omega_{b} h^{2} \leq 0.02$ \cite{copi} were assumed. Theoretical power spectra from Seljak \& Zaldarriaga (1996) were used in cases (i) and (ii), and kindly provided by N. Sugiyama for case (iii). The parameters fitted for (unless fixed) were the Hubble constant $h$ (in units of $100 \,\hbox{km}\,\hbox{s}^{-1}\,\hbox{Mpc}^{-1}$ ), the spectral tilt $n$, the cosmological constant $\Omega_{\Lambda}$ and the matter density $\Omega_m$ in units of the critical density. The flat models are defined by $\Omega_m + \Omega_{\Lambda} = 1$. Note the ranges of $h$ and $\Omega_{\Lambda}$ considered were $0.3 - 0.8$ and $0.3 - 0.7$ respectively. The results are displayed in Table~\ref{tabum}. In each case the best fit values of the parameters are shown, together with marginalised error ranges. These marginalised errors are $\pm 1\sigma$ confidence limits formed by integrating over all the other parameters with a uniform prior. \begin{table} \begin{center} \begin{tabular}{llccc} \multicolumn{1}{l}{Model}&\multicolumn{1}{l}{}&\multicolumn{1}{l}{Best Fit}&\multicolumn{2}{l}{Marginalised ranges}\\ & &value & ($1\sigma$) & ($2\sigma$) \\ Flat models &$h$ & 0.3 &0.3 -- 0.55 &0.3 -- 0.8 \\ &$n$ & 0.88 &0.8 -- 1.15 &0.65 -- 1.3 \\ && & & \\ Lambda models &$\Omega_{\Lambda}$ &0.3&0.4 -- 0.7 &0.3 -- 0.7 \\ &$h$ &0.35 &0.3 -- 0.55 &0.3 -- 0.8 \\ &$n$ &0.9 &0.8 -- 1.1 &0.7 -- 1.2 \\ && & & \\ Open models &$\Omega_m$ &0.7 &0.5 -- 1.0 &0.3 -- 1.0 \\ &$h$ &0.6 &0.3 -- 0.7 &0.3 -- 0.8 \\ && & & \\ \end{tabular} \caption{Results of the fitting analysis. Non-stated parameters were marginalised over to estimate the $1\sigma$ and $2\sigma$ ranges. } \label{tabum} \end{center} \end{table} In common with other recent work on fitting to current CMB data (e.g. Lineweaver \& Barbosa 1998) a tendency to low $H_0$ values (except in the case of open models) is found, although it is clear from the marginalised ranges that the statistical significance of this is not yet high. For comparison with the results of Webster {\it et~al.} \ (1998), who worked jointly with recent CMB and IRAS large-scale structure results, we note that a flat $\Lambda$ model with normalization fixed to the COBE results and $n=1$ yielded a best fit of $\Omega_{\Lambda}=0.7$ and $h=0.6$. In conjunction with the CAT points, new results forthcoming from the Python, Viper and QMAP\footnote{Results from QMAP published after this paper was submitted show a rise in the power spectrum from $\ell\sim 40$ to $\ell \sim 200$ consistent with the Saskatoon results (de Oliviera Costa et al. 1998).} experiments may soon be very significant in ruling out open models and in further delimiting the Doppler peak, sharpening up these parameter estimates. Recent OVRO results at $\ell \sim 590$ (Leitch, private communication) agree well with the values found by CAT. The joint CAT and OVRO results will clearly be very significant in constraining the latest cosmic string power spectrum predictions, which now include a cosmological constant (Battye {\it et~al.} \ 1998). These predictions succeed in recovering a significant `Doppler peak' in the power spectrum, but have peak power for $\ell$ in the range 500 to 600, at variance with the trend of the current experimental results. \section{Conclusions} \begin{itemize} \item[(1)] We have imaged a 2$^{\circ}$ $\times 2$$^{\circ}$\ patch of sky at 13-17 GHz with the CAT; this is the second field observed by CAT. \item[(2)] Significant CMB anisotropy is detected in the CAT2 field with an average power of $\Delta T/T = 2.1^{+0.4}_{-0.5} \times 10^{-5}$ for the $\ell =422$ bin, and with an upper limit of $\Delta T/T < 2.0 \times 10^{-5}$ for $\ell =615$ (due to Galactic contamination). \item[(3)] This new result is consistent with the first detection made by CAT in a different area of sky. \item[(4)] Together with other CMB data over a range of angular scales, the inclusion of the new CAT2 detection restricts the likely values of cosmological parameters. \end{itemize} \section*{Acknowledgments} Staff at the Cavendish Laboratory and MRAO, Lords Bridge are thanked for their ongoing support in the running of CAT. We thank an anonymous referee for comments. We are pleased to acknowledge major PPARC support of CAT. G. Rocha wishes to acknowledge a NSF grant EPS-9550487 with matching support from the State of Kansas and from a K$\ast$STAR First award. We also thank N. Sugiyama, and U. Seljak and M. Zaldarriaga for access to their theoretical power spectra.
\section{Deffinitions} \newcommand{i.e.\ }{i.e.\ } \newcommand{\mean}[1]{\left\langle #1 \right\rangle} \newcommand{\abs}[1]{\left| #1 \right|} \newcommand{\commute}[2]{\left[ #1,#2 \right]} \newcommand{\anticommute}[2]{\left\{ #1,#2 \right\}} \newcommand{\flist}[2]{#1_1,#1_2,\ldots,#1_{#2}} \newcommand{\plist}[2]{(\flist{#1}{#2})} \newcommand{\infty}{\infty} \newcommand{\bra}[1]{\left\langle #1 \right\rvert} \newcommand{\ket}[1]{\left\lvert #1 \right\rangle} \newcommand{\tfrac{1}{2}}{\tfrac{1}{2}} \newcommand{\tfrac{i}{2}}{\tfrac{i}{2}} \newcommand{\partialwith}[1]{\frac{\partial}{\partial#1}} \newcommand{\mathop{\rm Tr}\nolimits}{\mathop{\rm Tr}\nolimits} \newcommand{\mathop{\rm tr}\nolimits}{\mathop{\rm tr}\nolimits} \newcommand{\diag}[1]{\mathop{\rm diag}\nolimits\{#1\}} \newcommand{\partial_{t}}{\partial_{t}} \newcommand{D_{t}}{D_{t}} \newcommand{\mathcal{L}}{\mathcal{L}} \newcommand{\mathcal{S}}{\mathcal{S}} \newcommand{\bar{y}}{\bar{y}} \newcommand{\cc}[1]{\bar{#1} \newcommand{\hc}[1]{{{#1}^\dag}} \renewcommand{\bar}[1]{\overline{#1}} \newcommand{\order}[1]{\mathcal{O}(#1)} \newcommand{_{\text{eff}}}{_{\text{eff}}} \newcommand{\mathbf{X}}{\mathbf{X}} \newcommand{\mathbf{A}}{\mathbf{A}} \newcommand{\mathbf{F}}{\mathbf{F}} \newcommand{\boldsymbol{\O}}{\boldsymbol{\O}} \newcommand{\mathbf{c}}{\mathbf{c}} \newcommand{\mathbf{c}}{\mathbf{c}} \newcommand{\boldsymbol{\Psi}}{\boldsymbol{\Psi}} \newcommand{\boldsymbol{\psi}}{\boldsymbol{\psi}} \newcommand{\boldsymbol{\sigma}}{\boldsymbol{\sigma}} \newcommand{_{\text{cl}}}{_{\text{cl}}} \newcommand{\mathbf{X}\cl}{\mathbf{X}_{\text{cl}}} \newcommand{\mathbf{A}\cl}{\mathbf{A}_{\text{cl}}} \newcommand{D\cl}{D_{\text{cl}}} \newcommand{_{\text{q}}}{_{\text{q}}} \newcommand{g_{\text{s}}}{g_{\text{s}}} \newcommand{g_{\text{A}}}{g_{\text{A}}} \newcommand{g_{\text{B}}}{g_{\text{B}}} \newcommand{l_{\text{p}}}{l_{\text{p}}} \newcommand{l_{\text{s}}}{l_{\text{s}}} \newcommand{\prop}[3]{\Delta_{\text{#1}}(#2|#3)} \newcommand{\Bprop}[1]{\prop{B}{\tau_1,\tau_2}{r^{2}#1}} \newcommand{\Fprop}[1]{\prop{F}{\tau_1,\tau_2}{#1}} \newcommand{r\negmedspace\negmedspace/}{r\negmedspace\negmedspace/} \newcommand{v\negmedspace\negthickspace/}{v\negmedspace\negthickspace/} \newcommand{s\negmedspace\negthickspace/}{s\negmedspace\negthickspace/} \newcommand{\partial{\mskip -9.5mu}/\,}{\partial{\mskip -9.5mu}/\,} \newcommand{_{\vphantom{\beta}}}{_{\vphantom{\beta}}} \newcommand{^{\vphantom{i}}}{^{\vphantom{i}}} \newcommand{\Gt}{\widetilde\Gamma \newcommand{{\tilde\phi}}{{\tilde\phi}} \newcommand{{\hat\phi}}{{\hat\phi}} \newcommand{{\tilde\eta}}{{\tilde\eta}} \newcommand{{\bar t}}{{\bar t}} \newcommand{{\bar\omega}}{{\bar\omega}} \newcommand{G_{N}}{G_{N}} \newcommand{R_{s}}{R_{s}} \newcommand{\greekdef}[3] {\def#1{\relax\ifmmode#2\else#3\fi}} \def\alpha{\alpha} \def\beta{\beta} \newcommand{\gamma}{\gamma} \def\delta{\delta} \newcommand{\epsilon}{\epsilon} \newcommand{\zeta}{\zeta} \newcommand{\eta}{\eta} \greekdef{\o}{\theta}{\char"1C} \greekdef{\i}{\iota}{\char"10} \def\kappa{\kappa} \greekdef{\l}{\lambda}{\char32l} \newcommand{\mu}{\mu} \newcommand{\nu}{\nu} \newcommand{\xi}{\xi} \newcommand{\pi}{\pi} \def\rho{\rho} \newcommand{\sigma}{\sigma} \def\tau{\tau} \def\upsilon{\upsilon} \newcommand{\phi}{\phi} \def\chi{\chi} \newcommand{\psi}{\psi} \newcommand{\omega}{\omega} \newcommand{\Gamma}{\Gamma} \newcommand{\Delta}{\Delta} \greekdef{\O}{\Theta}{\char"1F} \greekdef{\L}{\Lambda} {\leavevmode\setbox0\hbox{L}\hbox to\wd0{\hss\char32L}} \newcommand{\Xi}{\Xi} \greekdef{\P}{\Pi}{\mathhexbox27B} \greekdef{\S}{\Sigma}{\mathhexbox278} \newcommand{\Upsilon}{\Upsilon} \newcommand{\Phi}{\Phi} \newcommand{\Psi}{\Psi} \newcommand{\Omega}{\Omega} \newcommand{\varepsilon}{\varepsilon} \newcommand{\vartheta}{\vartheta} \newcommand{\varpi}{\varpi} \newcommand{\varrho}{\varrho} \newcommand{\varsigma}{\varsigma} \newcommand{\varphi}{\varphi} \begin{document} \setlength{\unitlength}{1mm} \begin{titlepage} \maketitle \thispagestyle{empty} \begin{table}[t!] \rightline{hep-th/9904128} \rightline{EFI-99-14} \end{table} \begin{abstract} Eigenvalue repulsion can explain the holographic growth of black holes in Matrix theory. The resulting picture is essentially the same as the Boltzman gas picture but avoids any assumption about the effective potential between the D0 branes. Further, eigenvalue repulsion extends the Boltzman gas picture past the BFKS point to $N\gg S$. The use of Boltzman statistics is natural in this picture. \end{abstract} \end{titlepage} \setcounter{footnote}{0} \section{Introduction} Since Matrix theory is conjectured to be the discrete light-cone quantization of M-theory, it is natural to use it as a way of investigating the quantum properties of black holes. Banks, Fischler, Klebanov and Susskind (BFKS) and Martinec and Li, have studied black holes whose entropy, $S$, roughly the same as their light-like momentum, $N$ \cite{Banks:1998hz,Klebanov:1997kv,Banks:1998tn}. They can be formed by reducing the energy of a highly excited cluster of D0 branes until the momentum of the individual D0 branes is the inverse of the cluster's transverse size (saturating the Heisenberg uncertainty bound). The transverse size is found by setting the potential equal to the kinetic energy (in accord with the virial theorem). The choice of potential is crucial. The obvious choice, $v^{4}\!/r^{7}$, gives the Schwarzschild radius, $R_{s}$. However, at this point the expansion parameter for the potential is 1, so the expansion is on the verge of breaking down \cite{Horowitz:1998fr}. Nonetheless, once the correct radius is found, Matrix theory makes many correct predictions about black holes: the right mass for a given radius \cite{Klebanov:1997kv,Horowitz:1998fr,Banks:1998tn} and the correct long range gravitational potential between equal mass black holes \cite{Banks:1998tn}. The constituents have the correct properties to be Hawking radiation \cite{Banks:1998tn}. If the particles are treated with Boltzman statistics, the entropy is even correct \cite{Volovich:1996gw,Horowitz:1998fr,Banks:1998tn}. The problem becomes more pronounced when $N \gg S$. Simply replacing the D0 branes with bound states and using the $v^{4}/r^{7}$ potential fails to give correct results. Li identified other terms in the matix theory effective potential that could give the correct radius, but it is not clear why those terms would dominate \cite{Li:1998iz}. Li and Martinec get the correct results using $v^{4}/r^{7}$ to approximate processes that exchange longitudinal momentum \cite{Li:1998ci}. Eigenvalue repulsion can be used to find the size of black holes without any knowledge of the potential. Using this size, all of the previous black hole results follow \cite{Horowitz:1998fr,Li:1998ci,Gao:1998fe}. Eigenvalue repulsion allows the results to be extended to $N \gg S$. Because eigenvalue repulsion involves the off-diagonal elements of the matrices, it will be natural to treat the system with Boltzman statistics. First, I describe eigenvalue repulsion in Matrix theory. Then I find the radius and energy at $R=R_{s}$ and discuss entropy in this case. Finally, I discuss $R \gg R_{s}$. All factors of order one are ignored throughout. \section{Low energies and eigenvalue repulsion} To study black holes we need to understand matrix theory in the region where the energy per D0 brane is very low (much less than the plank energy). This will mean two things: First, remote D0 branes are not connected by strings (i.e.\ the off-diagonal elements between well separated D0 branes are in the ground state). As the D0 branes get within $l_{\text{p}}$ of each other, the quantum fluctuations in the off diagonal elements connecting the D0 branes also become of order $l_{\text{p}}$, so the D0 branes are able to explore their full matrix degrees of freedom. This region is sometimes called the stadium. The second consequence of the low energy is that the wave-function must be almost constant across the width of the stadium since variations on this scale would require very large kinetic energies. The flatness of the wave function in the stadium means that we can look to the theory of random matrices for intuition about how the system will explore the full matrix degrees of freedom. It is known that the statistics of random matrices strongly favors matrices with well separated eigenvalues, pushing the D0 branes to the edge of the stadium. The role of eigenvalue repulsion can be made a bit more quantitative. In the stadium, the distance between two D0 branes is $r^{2} = -\mathop{\rm Tr}\nolimits X^{i}X^{i}$. For fixed $r$, this is the equation for a sphere in 27 dimensions (nine directions times the three independent generators of traceless, antihermitian, $2 \times 2$ matrices). Since the wave function is relatively flat in this region, the probability that the D0 branes are a distance $r$ apart falls off like $r^{26}$. That is 18 powers of $r$ faster than the $r^{8}$ probability that one would expect for the nine spatial directions. This means that the wave function is strongly dominated by configurations which have the D0 branes near the edge of the stadium. There is no actual potential pushing the D0-branes apart, but this statistical effect is the Matrix Theory manifestation of eigenvalue repulsion.% \footnote{ The gauge symmetry does not affect this result. The gauge field degrees of freedom are removed by going to $A=0$ gauge. The assumption that the wave function is flat satisfies the requirement that the state be annihilated by the gauge generators. } The situation is similar when $N$ D0 branes are present. When the separation between \emph{any two} D0 branes falls below $l_{\text{p}}$, new matrix degrees of freedom open up and eigenvalue repulsion turns on. This causes the wave function to strongly favor configurations with size at least $N^{1/9}l_{\text{p}}$, one D0 brane per plank volume. It has been shown that composite systems of D0 branes grow like $N^{1/3}$, much faster than the $N^{1/9}$ growth predicted by eigenvalue repulsion \cite{Susskind:1998vk}. However, this rapid growth is caused by fluctuations in very high energy, off-diagonal elements of the matrix connecting widely separated D0 branes. Only very high energy probes will be able to see these high frequency quantum fluctuations. Eigenvalue repulsion is a zero energy effect, so it should be seen by any probe. Similarly, a low energy probe will be captured by a black hole if its impact parameter is less than $R_{s}$, but a high energy probe (with energy much greater than the mass of the black hole) will be captured at a much greater distance. \section{Black holes at the BFKS point} A very highly excited clump of D0 branes has enough mass to create a black hole whose radius is larger than the light-like compactification radius. Lowering the system's energy will slow the constituent D0 branes until the momentum saturates the Heisenberg uncertainty bound, i.e.\ the momentum of individual D0 branes is one over the size of the clump. This is the BFKS point \cite{Banks:1998hz}. Using the size, $N^{1/9}l_{\text{p}}$, given by eigenvalue repulsion one can find the mean kinetic energy. \begin{align} E_{k} &= \frac{R}{\hbar} \mathop{\rm Tr}\nolimits \P^{i}\P^{i} \\ &\approx \frac{R}{\hbar} \sum_{a=1}^{N} \frac{\hbar^{2}}{R_{s}^{2}} \\ &\approx \frac{R R_{s}^{7}}{G_{N}} \end{align} We know from the virial theorem that the potential energy will be of the same order as the kinetic energy, so we do not need to know anything about its form. Other than the form of the potential, this is exactly the system studied by previous authors. As they have observed, this system has the right relationship between mass and radius \cite{Klebanov:1997kv,Horowitz:1998fr,Banks:1998tn}, \begin{equation} M^{2} = E\frac{N\hbar}{R} = \frac{R_{s}^{16}}{G_{N}^{2}}, \end{equation} and gives the correct long range gravitational interaction for equal mass black holes \cite{Banks:1998tn}. The constituent D0 branes have the right energy and momentum to be Hawking radiation, should they escape \cite{Banks:1998tn}. \section{Entropy} To get the correct entropy the D0 branes must be treated with Boltzman statistics \cite{Banks:1998tn}. This occurs because the wave function is dominated by states that break all of the gauge symmetry mixing remote D0 branes, even the permutation symmetry. The large separation between the majority of D0 branes breaks the continuous symmetry down to the permutation symmetry. If all the D0 branes were far apart one could gauge transform all of the matrices into diagonal form simultaneously. Since this diagonalization would only be unique up to permutations, the permutation symmetries would be unbroken. However, since neighboring D0 branes are near enough to explore the full matrix degrees of freedom, the off diagonal elements connecting them cannot be integrated out \cite{Li:1998ci}. The matrices cannot be simultaneously diagonalized, only put in band diagonal form.% \footnote{ Here I am using ``band diagonal'' in a very general sense. The D0 branes will be connected to their nearest neighbors in all nine dimensions. There will be no way to write these matrices as two dimensional arrays with zeros away from the diagonal. } This breaks the permutation symmetry because a gauge transformation that permutes remote D0 branes would take the matrices out of band diagonal form. For example, consider a situation where D0 branes $a,b$ and $c$ are close together and $e,f$ and $g$ are close together, but the two sets are far apart. This will be represented by matrices like this (All the 1's represent terms of order $l_{\text{p}}$, not literally 1): \begin{equation} X^i = \begin{bmatrix} \ddots & 1 & & & & & & & & &\\ 1 & x^i_a & 1 & & & & & & & &\\ & 1 & x^i_b & 1 & & & & & & &\\ & & 1 & x^i_c & 1 & & & & & &\\ & & & 1 & \ddots & \ddots & & & & &\\ & & & & \ddots & \ddots & 1 & & & &\\ & & & & & 1 & x^i_e & 1 & & &\\ & & & & & & 1 & x^i_f & 1 & &\\ & & & & & & & 1 & x^i_g & 1 &\\ & & & & & & & & 1 & \ddots &\\ \end{bmatrix} \end{equation} Permuting the $b$ and $f$ D0 branes takes the matrices out of band diagonal form: \begin{equation} \hc{U} X^i U = \begin{bmatrix} \ddots & 1 & & & & & & & & &\\ 1 & x^i_a & & & & & & 1 & & &\\ & & x^i_f & & & & 1 & & 1 & &\\ & & & x^i_c & 1 & & & 1 & & &\\ & & & 1 & \ddots & \ddots & & & & &\\ & & & & \ddots & \ddots & 1 & & & &\\ & & 1 & & & 1 & x^i_e & & & &\\ & 1 & & 1 & & & & x^i_b & & &\\ & & 1 & & & & & & x^i_g & 1 &\\ & & & & & & & & 1 & \ddots &\\ \end{bmatrix} \end{equation} where $U$ is the matrix permuting $b$ and $f$. The symmetries mixing neighboring D0 branes is not broken at all. However, the number of near neighbors is of order one, so in doing statistical machanics for large $N$ the D0 branes should be treated as distinguishable. This description is similar to the proposal of BFKS in which the D0 branes are tethered to a background configuration \cite{Banks:1998tn}. Here, however, there is no firm distinction between the background and the excitations. \section{Beyond the BFKS point} At the BFKS point the constituent D0 branes have already saturated their uncertainty bound, so the energy of the system cannot be lowered by further reducing their momentum. They will have to band together into bound states. These bound states will have a larger mass (in the Matrix theory, they are of course massless in M-theory) and will be able to live in a smaller volume with less energy \cite{Horowitz:1998fr}. The arguments of the previous section would appear to apply to the bound states of Matrix theory, causing them to be seen by low energy probes as large objects. However, even in the large $N$ limit, these objects should not have hard interactions for impact parameters greater than $l_{\text{p}}$. Some miracle of the bound state wave function and supersymmetry must cause cancellations that allow the objects to pass through each other without interaction. These miracles will not occur when the wave functions are excited. In eleven dimensions this corresponds to changing the supergravitons into black holes. Excitations will cause the graviton to break into a metastable collection of $n$ smaller gravitons. Each of these is unexcited, and is therefore still able to pass through the others without interaction. The distance at which the eigenvalue repulsion plays a role is still $l_{\text{p}}$. Consider a system of two well separated bound states, consisting of $m_{1}$ and $m_{1}$ D0 branes each. They will break the $U(m_{1}+m_{2})$ symmetry down to $U(m_{1}) \times U(m_{2})$. There are $m_{1}m_{2}$ strings connecting them, each with 8 transverse, complex polarizations. The remaining $U(m_{1}) \times U(m_{2})$ gauge symmetry and the $O(8)$ rotation symmetry about the axis of separation is enough to relate all of these off diagonal modes, so they must all have the same mass. The mass is given by the separation of the bound states exactly as in the case of two separated D0 branes. When the centers of the two bound states get within a distance $l_{\text{p}}$ of each other, all $16 m_{1}m_{2}$ of these degrees of freedom open up, causing tremendous eigenvalue repulsion. This bound state cluster picture closely mimics the description at the BFKS point, and has been studied by several previous authors (once the radius is found). Once the radius is found, many of the successes of $N = S$ can be carried over in a straightforward manner to $N \gg S$. In particular, the mass to radius relationship is correct, the constituent bound states have the right properties to become Hawking radiation, and Boltzman statistics give the correct entropy \cite{Li:1998ci}. The breaking of the statistics symmetries by off diagonal fluctuations should work for bound states exactly as it did for individual D0 branes, explaining the use of Boltzman statistics. \subsection{Newtonian Potential} As a demonstration, consider the potential between two static black holes of different masses.% \footnote{ This analysis is similar to the equal mass case studied by Banks et.~al.~\cite{Banks:1998tn}, but differs from Gao and Zhang's treatment \cite{Gao:1998fe}. } We will assume that their separation is greater than $R$, where $v^{4}/r^{7}$ dominates the two body interactions. It will be much more convenient to use the velocities of the bound states rather than their momenta. The mass of a bound state is $ \hbar m / R $. Since the bound states that make up a black hole have momenta $ \hbar / R_{s} $ the velocities are $ R / R_{s} m $. First, observe that the velocity of the constituent bound states is the boost parameter required to bring the black hole to the rest frame. \begin{align} \frac{E}{P_{-}} &= \frac{ER}{N\hbar} \\ &= \frac{R^{2}R_{s}^{7}}{G_{N} m N \hbar} \\ &= \frac{R^{2}}{m^{2} R_{s}^{2}} \\ &= v^{2} \end{align} Black holes that are at rest relative to each other will be made of bound states with the same velocity, though not of the same momentum. It will be useful to note: \begin{align} M^{2} &= E P_{-} = v^{2} P_{-}^{2} \\ M &= \frac{v N \hbar}{R} \end{align} Next, we find the energy shift due to the $v^{4}/r^{7}$ interaction. Since the velocities of the constituents are roughly the same in the two black holes, $(v_{1}-v_{2})^{4}=v^{4}$. \begin{align} \Delta E &= \frac{G_{N}\hbar^{2}}{R^{3}} \sum_{n_{1},n_{2}} m_{1}m_{2} \frac{(v_{1}-v_{2})^{4}}{r^{7}} \\ &= \frac{G_{N}\hbar^{2}}{R^{3}} N_{1}N_{2} \frac{v^{4}}{r^{7}} \\ &= G_{N} \frac{M_{1}M_{2} v^{2}}{R r^{7}} \end{align} Finally, this can be used to find the potential. The potential is much smaller than the masses of the black holes. \begin{align} E\frac{N\hbar}{R} &= M^{2} = (M_{1} + M_{2} + V)^{2} \\ V &= \frac{\Delta E (N_{1} + N_{2})\hbar}{R (M_{1} + M_{2})} \\ &= \frac{\Delta E}{v} \\ &= G_{N} \frac{M_{1}M_{2}}{R' r^{7}}, \end{align} where $R'=R/v$ is the compactification radius in the rest frame of the black holes. \section{Conclusions} Eigenvalue repulsion predicts the size of black holes without requiring knowledge of the effective potential between the constituent D0 branes, and suggests a mechanism for breaking the statistics symmetry of the constituents. However, this understanding is still rather primitive. Much work will have to be done even to turn the sketch presented here into a derivation of black hole thermodynamics. Little has been said about the fermion degrees of freedom in this theory. Might supersymmetry cause terms that cancel the eigenvalue repulsion? This does not appear to happen. Supersymmetry requires that fermionic ground state energy cancels bosonic ground state energy. However, eigenvalue repulsion is purely statistical and is not affected by the fermionic states. I would like to thank Jeff Harvey, Emil Martinec and Miao Li for helpful conversations about these matters. \providecommand{\href}[2]{#2}\begingroup\raggedright
\section{Dijet Cross Sections at Low $Q^2$ and Virtual Photon Structure} \subsection{Comparing Dijet Cross Sections with NLO QCD Calculations} H1 have measured the triple-differential cross-section, $\triple$ ($\ensuremath{\overline{E_T}^2} $ is the mean $E_T$ of the two highest $E_T$ jets) in the $\gamma^*p$ center-of-mass system (hadronic cms)~\cite{vanc2}. The photon virtuality spans the range $1.6 < Q^2 < 80 {\rm GeV}^2$ and $y=Pq/Pk$ is constrained to $0.1<y<0.7$, where $P$, $q$ and $k$ are the four-vectors of the proton, virtual photon, and electron. The momentum fraction of the parton from the photon entering the hard scattering, $\ensuremath{x^{jets}_{\gamma}} $, defined as \begin{equation} \label{xgam} \ensuremath{x^{jets}_{\gamma}} = \frac{\sum_{i=1,2} E^{jets}_{T_i}\exp (-\eta_i^{jets})}{W} \end{equation} is estimated from the two highest $E_T$ jets. The variable $W^2=(P+q)^2=2Pq -Q^2$ defines the hadronic cms energy.\footnote{One can theoretically define the variable $x_\gamma=p_0P/qP$, where $p_0=x_\gamma q$ is the four-vector of the incoming parton from the photon. This definition assumes a collinear emission of the parton from the photon, which is an approximation neglecting some $k_\perp$ contribution due to the finite $Q^2$. These contributions are, however, small for moderate $Q^2$. This variable differs from that defined in (\ref{xgam}). When comparing with QCD calculations at LO level, partons directly give jets and there are exactly two partons in the final state. Thus, $x_\gamma^{jets} = x_\gamma$. However, in NLO, $x_\gamma$ still gives the momentum fraction of the parton in the photon, but $x_\gamma^{jets} \ne x_\gamma$. At NLO, one also has contributions to $x_\gamma^{jets} <1$ from the direct contribution, whereas at LO $x_\gamma^{jets}=1$ for the direct process.} The triple differential cross-section is shown in figure~\ref{figure4} \begin{figure}[htb] \begin{center} \mbox{\epsfig{file=nlo-final.eps,width=11.2cm}} \end{center} \vspace{-0.5cm} \caption{The differential dijet cross-section \triple shown as a function of $\ensuremath{x^{jets}_{\gamma}} $ for different regions of $\ensuremath{\overline{E_T}^2} $ and $Q^2$ compared to a NLO QCD calculation which employed the SaS1M PDF's of the virtual photon. The direct component of this model is shown as the shaded histogram. The result including a resolved component in NLO is shown as the full line. \label{figure4}} \end{figure} as a function of $\ensuremath{x^{jets}_{\gamma}} $ in ranges of $Q^2$ and $\ensuremath{\overline{E_T}^2} $. The data (points) are corrected for detector effects and the error bar shows the quadratic sum of systematic and statistical errors. The cross section decreases rapidly with increasing $\ensuremath{\overline{E_T}^2} $ and with increasing $Q^2$. Hadronization corrections are not included in the data. These are believed to be small on average but will presumably change the shape of the $\ensuremath{x^{jets}_{\gamma}} $ distribution. QCD calculations of dijet cross-sections with virtual photons have recently been performed at next-to-leading order (NLO) \cite{1,2}. The calculations are implemented in the fixed order program {\tt JetViP} \cite{jv}. In NLO, in the direct component a logarithm $\ln E_T^2/Q^2$ occurs, which is proportional to the photon splitting function. This term is large for $E_T^2\gg Q^2$ and therefore subtracted and resummed in the virtual photon structure function. The condition $E_T^2\gg Q^2$ ensures that it is possible to resolve an internal structure of the virtual photon. The virtual photon PDF's are suppressed as $Q^2 \rightarrow E_T^2$ and various anz{\"a}tze have been used to interpolate between the regions of known leading-log behaviour~\cite{SAS,vpth2,vpth5,vpth6}. It should be noted that the logarithmic term is only subtracted for the transversely polarized photons, since it vanishes in the case of longitudinally polarized photons for $Q^2\to 0$. The hadronic content of longitudinal virtual photons should be very small and therefore negligible. The results of the NLO calculations are also shown in figure~\ref{figure4}. The direct component of this model is shown as the shaded histogram. The result including a resolved component in NLO is shown as the full line. The calculation including a resolved photon component compares better to the data and indicates a need for a resolved virtual photon component below $10$~GeV$^2$, especially in the forward rapidity region, corresponding to low $\ensuremath{x^{jets}_{\gamma}} $. The discrepancy of data and NLO calculation for $\ensuremath{x^{jets}_{\gamma}} >0.75$ will presumably be cured when hadronization effects are considered, which lower the NLO results at large $\ensuremath{x^{jets}_{\gamma}} $. \subsection{Effective Virtual Photon Parton Densities} H1 have used their studies to extract an effective parton denstity (EPDF) of the virtual photon. By using the Single Effective Subprocess Approximation~\cite{SES}, the cross-section for dijet production in LO can be written \[ \!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\! \frac{{\rm d}^5\sigma}{{\rm d}y{\rm d}x_{\gamma} {\rm d}x_{\rm p}{\rm d}\cos\theta^*{\rm d}Q^2} \sim \frac{f_{eff/\gamma}^k(x_\gamma,P_t^2,Q^2)}{x_\gamma} \frac{f_{eff/{\rm p}}(x_{\rm p},P_t^2)}{x_{\rm p}} |M_{SES}(\cos\theta^*)|^2 . \] Here the Effective Parton Densities (EPDF), $f_{eff/\gamma}^k$ and $f_{eff/{\rm p}}$, are defined as \begin{math} f_{eff/{\rm A}} \equiv (f_{q/A}+f_{\overline{q}/A}) + \frac{9}{4} f_{g/A}. \end{math} The EPDF is shown in figure~\ref{figure6}. \begin{figure}[htb] \begin{center} \mbox{\epsfig{file=fig6.eps,width=10.4cm}} \end{center} \caption{The leading order effective parton density of the photon $x(q + 9/4 g)$, divided by the fine structure constant $\alpha$, as a function of $x_\gamma$ for different values of $Q^2$ and the transverse momentum $P_t^2$. The data are displayed as points, with the inner error bar depicting the statistical error, and the total error bar the quadratic sum of statistical and systematic errors. Also shown are the prediction from the DG model using GRV-LO real photon parton densities and $\omega=0.1$ (solid line) and the SAS-1D (dashed line) and SAS-2D(dot-dashed line) parameterisations. \label{figure6}} \end{figure} We see that, independently of $Q^2$, the parton density tends to be flat or rising with $\ensuremath{x_{\gamma}} $ (not to be confused with $\ensuremath{x^{jets}_{\gamma}} $). This behaviour is maintained as the probing scale increases. These are features characteristic of photon structure. The data are compared with predictions from the SaS~\cite{SAS} and DG~\cite{vpth5,vpth6} models which are able to describe the data quite well except where $Q^2 \to P_t^2$ and various aspects of the model start to break down. The models tend to underestimate the data in these regions. The three parameterisations for the parton density all give a good description of the data both in the lowest $\ensuremath{x_{\gamma}} $ range and in the lowest two $Q^2$ bins but predict a more rapid suppression as $Q^2\to P_t^2$ than is seen in the data. \section{Three Jet Photoproduction} ZEUS has measured the high-mass three-jet cross section in photoproduction, $d\sigma/dM_{\mb{3J}}$, as shown in Figure~\ref{M3J}~\cite{ZEUS_3J}. \begin{figure}[htb] \begin{center} \mbox{\epsfig{file=M3J.eps,width=7.cm}} \end{center} \caption{The three-jet cross section $d\sigma/dM_{\mt{3J}}$. The dots show the data. The inner error bars show the statistical error. The outer error bars show the quadratic sum of the statistical and systematic uncertainties with the exception of the absolute energy scale uncertainty which is shown separately as a shaded band. $\mathcal{O}(\alpha \alpha_{s}^{2})$\ pQCD calculations by two groups of authors are shown by thick solid and dot-dashed lines. The thin solid and dashed histograms show the predictions from PYTHIA and HERWIG. \label{M3J}} \end{figure} The $\mathcal{O}(\alpha \alpha_{s}^{2})$\ pQCD calculations from two groups of authors~\cite{harris,klasen} provide a good description of the data, even though they are leading order for this process. Monte Carlo models also generate three-jet events through the parton shower mechanism and both PYTHIA~\cite{PYTHIA} and HERWIG~\cite{HERWIG} reproduce the shape of the $M_{\mb{3J}}$ distribution. For three-jet events there are two relevant scattering angles as illustrated in Figure~\ref{angles}(a). \begin{figure}[htb] \begin{center} \mbox{\epsfig{file=diagram.eps,width=4.6cm}}\mbox{\epsfig{file=angles.eps,width=9.cm}} \end{center} \caption{(a) Centre-of-mass frame diagram of the three-body scattering angles. The beam direction is indicated by $\vec{P}_{B}$. Distributions of $\cos \theta_3$ and $\psi_3$ are shown in (b) and (c). The dotted curve shows the distribution for a constant matrix element. Other details are as described for Figure~\ref{M3J}. \label{angles}} \end{figure} The distributions of $\cos \theta_3$ and $\psi_3$ are shown in Figure~\ref{angles}(b) and (c). The $\cos \theta_3$ distribution resembles that of $\cos \theta^*$ in dijet production~\cite{ZEUS_2J} and exhibits forward and backward peaks. It is well described in both $\mathcal{O}(\alpha \alpha_{s}^{2})$\ pQCD calculations and parton shower models. The $\psi_3$ distribution is peaked near 0 and $\pi$ indicating that the three-jet plane tends to lie near the plane containing the highest energy jet and the beam. This is particularly evident if one considers the $\psi_3$ distribution for three partons uniformly distributed in the available phase space. The phase space near $\psi_3 = 0$ and $\pi$ has been depleted by the $E_T^{\mb jet}$ cuts and by the jet-finding algorithm. The pQCD calculations describe perfectly the $\psi_3$ distribution. It is remarkable that the parton shower models PYTHIA and HERWIG are also able to reproduce the $\psi_3$ distribution. Within the parton-shower model it is possible to determine the contribution to three-jet production from initial-state radiation (ISR) and final-state radiation. It is also possible to switch the QCD phenomenon of colour coherence on and off. From a Monte Carlo study it has been determined that ISR is predominantly responsible for three-jet production. Also, it has been found that colour coherence can account for the suppression of large angle emissions which leads to the depletion of the $\psi_3$ distribution near $\psi_3 = \pi / 2$~\cite{ZEUS_3J}. \section{Summary} The dijet cross section at low $Q^2$ has been measured by the H1 collaboration and compared to NLO QCD calculations. The comparison shows a clear need for resolved virtual photon component. First measurements of a leading order effective virtual photon PDF have been made. Existing models for virtual photon PDF's are consistent with the measurements but systematic errors are still large. A first measurement of high transverse energy three-jet photoproduction has been performed. The distribution of the three jets is sensitive to colour coherence and is correctly predicted in $\mathcal{O}(\alpha \alpha_{s}^{2})$\ pQCD.
\section{\boldmath Models of substructure in $\ensuremath{e^{+}e^{-}}\to 4\pi$}\label{appa} In this section we discuss the details of the model which has been used for description of the experimental data. Since the initial hadron state (referred to as $\tilde\rho$) is created by a virtual photon ($e^+e^- \to \gamma^*\to\tilde\rho $) and then decays into four pions, it has the $\rho$-meson quantum numbers ($I^{G}J^{PC}=1^{+}1^{--}$). We use the notation ${\tilde e}_{\mu}$ for the polarization vector of this state. We assume that the main contribution to the amplitude of the process in the energy range under study is given by the intermediate state resonances having the masses close to the threshold of $\rho\pi$ production (\ref{eq1})-(\ref{eq7}). In the case of broad resonances ($a_1(1260)$,$a_2(1320)$,$\pi(1300)$ etc.), the form of their propagators $1/D(q)$ is very important for analysing the data. We learned that it is necessary to take into account the dependence of the imaginary part of the propagators (width) on virtuality while the corresponding corrections to the real part which can be expressed through the imaginary part by dispersion relations are not so important. We represent the function $D(q)$ in the form \begin{equation} D(q)=q^2-M^2 +iM\Gamma \frac{g(q^2)}{g(M^2)} \, , \end{equation} where $M$ and $\Gamma$ are the mass and width of the corresponding particle, and the function $g(s)$ describes the dependence of the width on the virtuality. If $q^2=M^2$ then $D=iM\Gamma$, in accordance with the usual definition of mass and width of the resonance. In the case of the $\rho$-meson we used the following representation for the function $g_{\rho}(q)$: \begin{equation} g_{\rho}(s)=s^{-1/2}(s-4m^2)^{3/2} \, . \end{equation} \subsection{The contribution of $a_1(1260)\pi$} The quantum numbers of the $a_1(1260)$ resonance are $I^{G}J^{PC}=1^{-}1^{++}$. Taking into account the quantum numbers of the pion, we can write the matrix elements corresponding to the processes $\tilde\rho(P)\to a_1(q)\pi(p)$ and $a_1(q)\to\rho(P^{\prime})\pi(p)$ as \begin{eqnarray}\label{M1} T(\tilde\rho\to a_1\pi)=F_{\tilde\rho a_1\pi}\mbox{$\varepsilon$}^{3ab} (P_{\mu}{\tilde e}_{\nu}-P_{\nu}{\tilde e}_{\mu}) q_{\mu}A_{\nu}^{a*}\phi^{b*}\, ,\nonumber\\ T(a_1\to \rho\pi)=F_{a_1\rho\pi}\mbox{$\varepsilon$}^{abc} q_{\mu}A_{\nu}^{a} (P_{\mu}^{\prime}e_{\nu}^{b*}-P_{\nu}^{\prime}e_{\mu}^{b*})\phi^{c*} \, , \end{eqnarray} where $a,b,c$ are isospin indices, $A_{\mu}^a$ and $e_{\mu}^a$ are the polarization vectors of $a_1$- and $\rho$-mesons, $\phi^b$ is the pion wave function, $F_{a_1\rho\pi}$ and $F_{\tilde\rho a_1\pi}$ are form factors depending on the virtuality of initial and final particles. The explicit form of these form factors in the energy range considered is not very essential (it contributes to the theoretical uncertainty of the model). The matrix element of the transition $T(\rho\to \pi\pi)$ reads \begin{equation}\label{Mrho} T(\rho\to \pi\pi)=F_{\rho\pi\pi}\mbox{$\varepsilon$}^{abc} e_{\mu}^{a}(p_{\mu}^{(b)}-p_{\mu}^{(c)})\phi^{b*}\phi^{c*}\, , \end{equation} where $p_{\mu}^{(b)}$ and $p_{\mu}^{(c)}$ are 4-momenta of the corresponding pions. Due to the helicity conservation only transverse space components (with respect to electron and positron momenta) of the 4-vector ${\tilde e}_{\mu}$ are not zero. We denote them as $\tilde{\bf e}_{\perp}$. For unpolarized electrons and positrons a square of matrix element for the process $\tilde\rho \to 4\pi$ is of the form \begin{equation} \mid T\mid^2 =\, \mid {\bf J}_{\perp}\mid^2 . \end{equation} Using \eq{M1} and \eq{Mrho} we obtain the following expression for the contribution of the $a_1(1260)$-meson to the current $ {\bf J}$ in the process \\ $\tilde\rho\to \pi^+{(p_1)}\pi^+{(p_2)}\pi^-{(p_3)}\pi^-{(p_4)}$ : \begin{eqnarray}\label{JAPPMM} {\bf J}_{a_1}^{++--}\!\!\!&=&\!\!\! G\left[{\bf t}_{a_1}(p_1,p_2,p_3,p_4)+{\bf t}_{a_1}(p_1,p_4,p_3,p_2) +{\bf t}_{a_1}(p_2,p_1,p_3,p_4) \right. \nonumber\\ &&+{\bf t}_{a_1}(p_2,p_4,p_3,p_1)+{\bf t}_{a_1}(p_1,p_2,p_4,p_3) +{\bf t}_{a_1}(p_1,p_3,p_4,p_2) \nonumber \\ &&+\left.{\bf t}_{a_1}(p_2,p_1,p_4,p_3) +{\bf t}_{a_1}(p_2,p_3,p_4,p_1)\right] \, , \end{eqnarray} where \begin{eqnarray}\label{j} &\displaystyle {\bf t}_{a_1}(p_1,p_2,p_3,p_4)=\frac{F_{a_1}^2(P-p_4)} {D_{a_1}(P-p_4)D_{\rho}(p_1+p_3)}\\ &\displaystyle \times\{(E-\mbox{$\varepsilon$}_4)[\mbox{${\bf p}$}_1(E\mbox{$\varepsilon$}_3-p_4p_3)-\mbox{${\bf p}$}_3(E\mbox{$\varepsilon$}_1-p_4p_1)] \nonumber \\ &\displaystyle - \mbox{${\bf p}$}_4[\mbox{$\varepsilon$}_1(p_4p_3)-\mbox{$\varepsilon$}_3(p_4p_1)]\}\, , \nonumber \end{eqnarray} $P$ is the initial 4-momentum ( $P^0=E$ , ${\bf P}=0$) , $p_i=(\mbox{$\varepsilon$}_i, \mbox{${\bf p}$}_i)$ , $1/D_A(q)$ and $1/D_{\rho}(q)$ are propagators of the $a_1$- and $\rho$- mesons, $F_{a_1}(q)$ is the form factor, $G$ is the function of the initial energy $E$ proportional to the amplitude of $\tilde\rho$ creation. Similarly to \eq{JAPPMM}, we obtain for the contribution of $a_1(1260)$ to the current $ {\bf J}$ in the process $\tilde\rho\to \pi^+{(p_1)}\pi^-{(p_4)}\pi^0{(p_2)}\pi^0{(p_3)}$: \begin{eqnarray}\label{JAPM00} {\bf J}_{a_1}^{+-00}&=&G\left[{\bf t}_{a_1}(p_1,p_2,p_3,p_4) -{\bf t}_{a_1}(p_4,p_2,p_3,p_1)\right. \\ &&+\left.{\bf t}_{a_1}(p_1,p_3,p_2,p_4)-{\bf t}_{a_1}(p_4,p_3,p_2,p_1)\right]\, . \nonumber \end{eqnarray} The function $g_{a_1}(s)$ in the propagator of $a_1$ reads: \begin{eqnarray}\label{g} g_{a_1}(s)&=& F_{a_1}^2(Q) \int \left | \frac{\mbox{$\varepsilon$}_2\mbox{${\bf p}$}_1-\mbox{$\varepsilon$}_1\mbox{${\bf p}$}_2}{D_{\rho}(p_1+p_2)}+ \frac{\mbox{$\varepsilon$}_2\mbox{${\bf p}$}_3-\mbox{$\varepsilon$}_3\mbox{${\bf p}$}_2}{D_{\rho}(p_2+p_3)}\right|^2 \nonumber \\ && \times \frac{d\mbox{${\bf p}$}_1\,d\mbox{${\bf p}$}_2\,d\mbox{${\bf p}$}_3\,\delta^{(4)}(p_1+p_2+p_3-Q)} {2\mbox{$\varepsilon$}_12\mbox{$\varepsilon$}_2 2\mbox{$\varepsilon$}_3(2\pi)^5}\, , \end{eqnarray} where $Q^0=\sqrt s$ and ${\bf Q}=0$. As a form factor, we used the function $F(q)=(1+m_{a_1}^2/\Lambda^2)/(1+q^2/\Lambda^2)$ with $\Lambda\sim$ 1~GeV. \subsection{The contribution of $\omega\pi$} The amplitude of the process $\tilde\rho({\cal P})\to \omega(q)\pi(p)$ has the form: \begin{equation}\label{om} T(\tilde\rho\to \omega\pi)=F_{\tilde\rho\omega\pi} \mbox{$\varepsilon$}_{\mu\nu\alpha\beta}{\cal P}_{\mu} q_{\nu}\tilde{e}_{\alpha}^a e_{\beta}^a\, , \end{equation} where $ e_{\beta}^a$ is the polarization vector of the $\omega$-meson. The matrix element of the transition $\omega\to\rho\pi$ can be written in the similar form. The $\omega$-meson contributes only to a channel $\tilde\rho\to \pi^+{(p_1)}\pi^-{(p_4)}\pi^0{(p_2)}\pi^0{(p_3)}$ . The corresponding current is equal to \begin{eqnarray}\label{JOMPM00} {\bf J}_{\omega}^{+-00}&=& G_{\omega}\left[{\bf t}_{\omega}(p_2,p_4,p_1,p_3)- {\bf t}_{\omega}(p_2,p_1,p_4,p_3)\right.\nonumber\\ &&-\left.{\bf t}_{\omega}(p_2,p_3,p_1,p_4) \right] + (p_2\leftrightarrow p_3) \, , \end{eqnarray} where \begin{eqnarray} &\displaystyle {\bf t}_{\omega}(p_1,p_2,p_3,p_4)= \frac{F_{\omega}^2(P-p_1)}{D_{\omega}(P-p_1)D_{\rho}(p_3+p_4)}\\ &\displaystyle \times \{(\mbox{$\varepsilon$}_4\mbox{${\bf p}$}_3-\mbox{$\varepsilon$}_3\mbox{${\bf p}$}_4)(\mbox{${\bf p}$}_1\mbox{${\bf p}$}_2)- \mbox{${\bf p}$}_2(\mbox{$\varepsilon$}_4\mbox{${\bf p}$}_1\mbox{${\bf p}$}_3-\mbox{$\varepsilon$}_3\mbox{${\bf p}$}_1\mbox{${\bf p}$}_4) \nonumber \\ &\displaystyle -\mbox{$\varepsilon$}_2[\mbox{${\bf p}$}_3(\mbox{${\bf p}$}_1\mbox{${\bf p}$}_4)-\mbox{${\bf p}$}_4(\mbox{${\bf p}$}_1\mbox{${\bf p}$}_3)]\} \, , \nonumber \end{eqnarray} $\mbox{$\varepsilon$}_i$ is the energy of the corresponding pion, $F_{\omega}(q)$ is the form factor. Since the width of the $\omega$ is small, we set $g_{\omega}(s)=1$ in the propagator $D_{\omega}(q)$. \subsection{The contribution of $h_1(1170)\pi$} The quantum numbers of the $h_1(1170)$ resonance are $I^{G}J^{PC}=0^{-}1^{+-}$. Similarly to the $\omega$-meson it gives the contribution to the mixed channel only. The matrix elements for the transitions $\tilde\rho(P)\to h_1(q)\pi^0(p)$ and $h_1(q)\to\rho(P)\pi(p)$ have the form \begin{eqnarray}\label{H1} T(\tilde\rho\to h_1\pi^0)=F_{\tilde\rho h_1\pi} (P_{\mu}{\tilde e}_{\nu}-P_{\nu}{\tilde e}_{\mu}) q_{\mu}H_{\nu}^*\phi^{3*}\, ,\nonumber\\ T(h_1\to \rho\pi)=F_{h_1\rho\pi} q_{\mu}H_{\nu} (P_{\mu}e_{\nu}^{a*}-P_{\nu}e_{\mu}^{a*})\phi^{a*} \, , \end{eqnarray} where $H_{\mu}$ is the polarization vector of $h_1$. Using \eq{H1} and \eq{Mrho} we obtain the following representation for the $h_1$-meson contribution to the current $ {\bf J}$ in the decay $\tilde\rho\to \pi^+{(p_1)}\pi^-{(p_4)}\pi^0{(p_2)}\pi^0{(p_3)}$ : \begin{eqnarray}\label{JHPM00} {\bf J}_{h_1}^{+-00}&=& G_{h_1}\left[{\bf t}_{h_1}(p_1,p_4,p_3,p_2)- {\bf t}_{h_1}(p_4,p_1,p_3,p_2)\right.\nonumber\\ &&-\left.{\bf t}_{h_1}(p_1,p_3,p_4,p_2)\right] + (p_2\leftrightarrow p_3) \, , \end{eqnarray} the current ${\bf t}_{h_1}$ is given by \eq{j} with the change of indices $a_1\to h_1$ , and the function $g_{h_1}(s)$ in the propagator of $h_1$ is: \begin{eqnarray}\label{gh} g_{h_1}(s)&=&F_{h_1}^2(Q) \int \left | \frac{\mbox{$\varepsilon$}_2\mbox{${\bf p}$}_1-\mbox{$\varepsilon$}_1\mbox{${\bf p}$}_2}{D_{\rho}(p_1+p_2)}- \frac{\mbox{$\varepsilon$}_2\mbox{${\bf p}$}_3-\mbox{$\varepsilon$}_3\mbox{${\bf p}$}_2}{D_{\rho}(p_2+p_3)}- \frac{\mbox{$\varepsilon$}_3\mbox{${\bf p}$}_1-\mbox{$\varepsilon$}_1\mbox{${\bf p}$}_3}{D_{\rho}(p_1+p_3)} \right|^2 \nonumber\\ && \times \frac{d\mbox{${\bf p}$}_1\,d\mbox{${\bf p}$}_2\,d\mbox{${\bf p}$}_3\,\delta^{(4)}(p_1+p_2+p_3-Q)} {2\mbox{$\varepsilon$}_12\mbox{$\varepsilon$}_2 2\mbox{$\varepsilon$}_3(2\pi)^5}\, . \end{eqnarray} \subsection {The contribution of $\rho^+\rho^-$} One more contribution to the mixed channel amplitude comes from the process $\tilde\rho\to \rho^+\rho^-\to 4\pi$. The matrix element corresponding to the transition $\tilde\rho(P)\to \rho^+(p)\rho^-(q)$ reads \begin{eqnarray}\label{rhorho} T(\tilde\rho\to \rho^+\rho^-)&=&F_{\tilde\rho \rho^+\rho^-} (P_{\mu}{\tilde e}_{\nu}-P_{\nu}{\tilde e}_{\mu}) \\ && \times [(p_{\mu}e_{\alpha}^{+*}-p_{\alpha}e_{\mu}^{+*}) (q_{\nu}e_{\alpha}^{-*}-q_{\alpha}e_{\nu})^{-*})-(\mu\leftrightarrow\nu)] \nonumber \end{eqnarray} where $e_{\mu}^+$ and $e_{\mu}^-$ are the polarization vectors of $\rho^+$ and $\rho^-$ respectively. Using \eq{rhorho}, we obtain the contribution of $\rho^+\rho^-$ to the current $ {\bf J}$ in the decay $\tilde\rho\to \pi^+{(p_1)}\pi^-{(p_4)}\pi^0{(p_2)}\pi^0{(p_3)}$ : \begin{eqnarray}\label{JRRPM00} {\bf J}_{\rho\rho}^{+-00}&=& \frac{G_{\rho\rho}F_{\rho}^2(p_1+p_2)F_{\rho}^2(p_3+p_4)} {D_{\rho}(p_1+p_2)D_{\rho}(p_3+p_4)} \\ && \times [\mbox{${\bf p}$}_1(\mbox{$\varepsilon$}_3\mbox{${\bf p}$}_2\mbox{${\bf p}$}_4-\mbox{$\varepsilon$}_4\mbox{${\bf p}$}_2\mbox{${\bf p}$}_3)- \mbox{${\bf p}$}_2(\mbox{$\varepsilon$}_3\mbox{${\bf p}$}_1\mbox{${\bf p}$}_4-\mbox{$\varepsilon$}_4\mbox{${\bf p}$}_1\mbox{${\bf p}$}_3)\nonumber \\ && -\mbox{${\bf p}$}_3(\mbox{$\varepsilon$}_1\mbox{${\bf p}$}_2\mbox{${\bf p}$}_4-\mbox{$\varepsilon$}_2\mbox{${\bf p}$}_1\mbox{${\bf p}$}_4)+ \mbox{${\bf p}$}_4(\mbox{$\varepsilon$}_1\mbox{${\bf p}$}_2\mbox{${\bf p}$}_3-\mbox{$\varepsilon$}_2\mbox{${\bf p}$}_1\mbox{${\bf p}$}_3)] + (2\leftrightarrow 3)\, . \nonumber \end{eqnarray} \subsection{The contribution of $\pi(1300)\pi$} The matrix elements for the transitions $\tilde\rho(P)\to \pi^{\prime}(q)\pi(p)$ and $\pi^{\prime}(q)\to\rho(P)\pi(p)$ are of the form (for brevity $\pi^{\prime}\equiv\pi(1300)$): \begin{eqnarray}\label{pp} T(\tilde\rho\to \pi^{\prime}\pi)=F_{\tilde\rho \pi^{\prime}\pi}\mbox{$\varepsilon$}^{3ab} (P_{\mu}{\tilde e}_{\nu}-P_{\nu}{\tilde e}_{\mu}) q_{\mu}p_{\nu}\phi^{\prime a*}\phi^{b*}\, ,\nonumber\\ T(\pi^{\prime}\to \rho\pi)=F_{\pi^{\prime}\rho\pi}\mbox{$\varepsilon$}^{abc} q_{\mu}p_{\nu} (P_{\mu}e_{\nu}^{b*}-P_{\nu}e_{\mu}^{b*})\phi^{\prime a}\phi^{c*} \, . \end{eqnarray} Then we get for the contribution of $\pi(1300)$ to the current $ {\bf J}$ in the decay $\tilde\rho\to \pi^+{(p_1)}\pi^+{(p_2)}\pi^-{(p_3)}\pi^-{(p_4)}$ : \begin{eqnarray}\label{JPiPPMM} {\bf J}_{\pi^{\prime}}^{++--}&=& G_{\pi^{\prime}}\left[{\bf t}_{\pi^{\prime}}(p_1,p_2,p_3,p_4)+ {\bf t}_{\pi^{\prime}}(p_2,p_1,p_3,p_4)+{\bf t}_{\pi^{\prime}}(p_1,p_2,p_4,p_3) \right. \nonumber\\ &&+{\bf t}_{\pi^{\prime}}(p_2,p_1,p_4,p_3)+{\bf t}_{\pi^{\prime}}(p_3,p_2,p_1,p_4) +{\bf t}_{\pi^{\prime}}(p_4,p_2,p_1,p_3)\nonumber\\ &&+\left.{\bf t}_{\pi^{\prime}}(p_3,p_1,p_2,p_4)+{\bf t}_{\pi^{\prime}} (p_4,p_1,p_2,p_3)\right]\, . \end{eqnarray} In the mixed channel $\tilde\rho\to\pi^+{(p_1)}\pi^-{(p_4)}\pi^0{(p_2)}\pi^0{(p_3)}$ the corresponding current reads: \begin{eqnarray}\label{JPiPP00} {\bf J}_{\pi^{\prime}}^{+-00}&=& G_{\pi^{\prime}}\left[ {\bf t}_{\pi^{\prime}}(p_1,p_2,p_3,p_4)+{\bf t}_{\pi^{\prime}}(p_1,p_3,p_2,p_4)\right. \nonumber\\ &&+\left.{\bf t}_{\pi^{\prime}}(p_4,p_1,p_3,p_2)+{\bf t}_{\pi^{\prime}}(p_4,p_1,p_2,p_3) \right] . \end{eqnarray} It \eq{JPiPPMM} and \eq{JPiPP00} we use the notation \begin{equation} {\bf t}_{\pi^{\prime}}(p_1,p_2,p_3,p_4)= \frac{F_{\pi^{\prime}}^2(P-p_1)}{D_{\pi^{\prime}}(P-p_1)D_{\rho}(p_2+p_4)} (p_2p_3-p_4p_3)(p_2+p_4)^2\,\mbox{${\bf p}$}_1 \, . \end{equation} For the function $g_{\pi^{\prime}}(s)$ in the $\pi^{\prime}$ propagator, one has: \begin{eqnarray}\label{gpi} g_{\pi^{\prime}}(s)&=&F_{\pi^{\prime}}^2(Q) \int \left | \frac{\mbox{$\varepsilon$}_1\mbox{${\bf p}$}_2\mbox{${\bf p}$}_3-\mbox{$\varepsilon$}_2\mbox{${\bf p}$}_1\mbox{${\bf p}$}_3}{D_{\rho}(p_1+p_2)}+ \frac{\mbox{$\varepsilon$}_3\mbox{${\bf p}$}_1\mbox{${\bf p}$}_2-\mbox{$\varepsilon$}_2\mbox{${\bf p}$}_1\mbox{${\bf p}$}_3}{D_{\rho}(p_2+p_3)} \right|^2 \nonumber\\ && \times \frac{d\mbox{${\bf p}$}_1\,d\mbox{${\bf p}$}_2\,d\mbox{${\bf p}$}_3\,\delta^{(4)}(p_1+p_2+p_3-Q)} {2\mbox{$\varepsilon$}_12\mbox{$\varepsilon$}_2 2\mbox{$\varepsilon$}_3(2\pi)^5}\, . \end{eqnarray} \subsection{The contribution of $\sigma\rho$} The quantum numbers of the $\sigma$ resonance are $I^{G}J^{PC}=0^{+}0^{++}$. The matrix element of the transition $\tilde\rho(P)\to \sigma(q)\rho^0(p)$ is of the form: \begin{equation}\label{sigrho} T(\tilde\rho\to \sigma\rho^0)=F_{\tilde\rho \sigma\rho} (P_{\mu}{\tilde e}_{\nu}-P_{\nu}{\tilde e}_{\mu}) q_{\mu}e_{\nu}^*\phi_{\sigma}^{*} . \end{equation} The corresponding contribution to the current $ {\bf J}$ in the decay \\ $\tilde\rho\to \pi^+{(p_1)}\pi^+{(p_2)}\pi^-{(p_3)}\pi^-{(p_4)}$ reads: \begin{eqnarray}\label{SigRhoPPMM} {\bf J}_{\sigma}^{++--}&=& G_{\sigma}\left[{\bf t}_{\sigma}(p_1,p_2,p_3,p_4)+ {\bf t}_{\sigma}(p_2,p_1,p_3,p_4)\right. \nonumber\\ &&+\left.{\bf t}_{\sigma}(p_1,p_2,p_4,p_3)+{\bf t}_{\sigma}(p_2,p_1,p_4,p_3)\right]\, . \end{eqnarray} In the mixed channel $\tilde\rho\to\pi^+{(p_1)}\pi^-{(p_4)}\pi^0{(p_2)}\pi^0{(p_3)}$ the current has the form \begin{equation}\label{SigRhoPP00} {\bf J}_{\sigma}^{+-00}=G_{\sigma}{\bf t}_{\sigma}(p_1,p_2,p_3,p_4)\, , \end{equation} where \begin{equation} {\bf t}_{\sigma}(p_1,p_2,p_3,p_4)= \frac{F_{\sigma}^2(p_2+p_3)}{D_{\sigma}(p_2+p_3)D_{\rho}(p_1+p_4)} (\mbox{$\varepsilon$}_4\mbox{${\bf p}$}_1-\mbox{$\varepsilon$}_1\mbox{${\bf p}$}_4) \, . \end{equation} The function $g_{\sigma}(s)$ in the propagator of $\sigma$ is equal to \begin{equation} g_{\sigma}(s)=(1-4m^2/s)^{1/2} \, . \end{equation} \subsection{The contribution of $a_2(1320)\pi$} The quantum numbers of the $a_2(1320)$ resonance are $I^{G}J^{PC}=1^{-}2^{++}$. The matrix elements for the transitions $\tilde\rho(P)\to a_2(q)\pi(p)$ and $a_2(q)\to\rho(P^{\prime})\pi(p)$ can be written in the form \begin{eqnarray}\label{a2} T(\tilde\rho\to a_2\pi)=F_{\tilde\rho a_2\pi}\mbox{$\varepsilon$}^{3ab} \mbox{$\varepsilon$}_{\mu,\nu\rho\lambda}P_{\mu}{\tilde e}_{\nu}p_{\rho} A_{\lambda\gamma}^{a*}p_{\gamma}q_{\mu}\phi^{b*}\, ,\nonumber\\ T(a_2\to \rho\pi)=F_{a_2\rho\pi}\mbox{$\varepsilon$}^{abc} \mbox{$\varepsilon$}_{\mu,\nu\rho\lambda}P_{\mu}^{\prime}{e}_{\nu}^{a*}p_{\rho} A_{\lambda\gamma}^{b}p_{\gamma}\phi^{c*}\, \end{eqnarray} where $A_{\mu\nu}^a$ is the polarization tensor of $a_2$ . The contributions of the $a_2$-meson to the current $ {\bf J}$ in the processes $\tilde\rho\to \pi^+{(p_1)}\pi^+{(p_2)}\pi^-{(p_3)}\pi^-{(p_4)}$ and $\tilde\rho\to \pi^+{(p_1)}\pi^-{(p_4)}\pi^0{(p_2)}\pi^0{(p_3)}$ are given by the formulae \eq{JAPPMM} and \eq{JAPM00} with the substitution ${\bf t}_{a_1}\to{\bf t}_{a2}$ , where \begin{eqnarray}\label{ja2} &\displaystyle {\bf t}_{a2}(p_1,p_2,p_3,p_4)=\frac{F_{a2}^2(P-p_4)} {D_A(P-p_4)D_{\rho}(p_1+p_3)}\\ &\displaystyle \times\{E(\mbox{${\bf p}$}_4\times\mbox{${\bf p}$}_2)[\mbox{${\bf p}$}_1\times\mbox{${\bf p}$}_3,\mbox{${\bf p}$}_2]+ [p_2p_4-(p_2q)(p_4q)/m_{a2}^2] \nonumber\\ &\displaystyle \times [\mbox{${\bf p}$}_1(\mbox{$\varepsilon$}_3\mbox{${\bf p}$}_2\mbox{${\bf p}$}_4-\mbox{$\varepsilon$}_2\mbox{${\bf p}$}_4\mbox{${\bf p}$}_3) -\mbox{${\bf p}$}_3(\mbox{$\varepsilon$}_1\mbox{${\bf p}$}_4\mbox{${\bf p}$}_2)-\mbox{$\varepsilon$}_2\mbox{${\bf p}$}_4\mbox{${\bf p}$}_1) \nonumber \\ &\displaystyle +\mbox{${\bf p}$}_2(\mbox{$\varepsilon$}_1\mbox{${\bf p}$}_3\mbox{${\bf p}$}_4-\mbox{$\varepsilon$}_3\mbox{${\bf p}$}_1\mbox{${\bf p}$}_4) ]\}\, . \nonumber \end{eqnarray} Here $q=P-p_4$, and $P$ is the initial 4-momentum. For the function $g_{a2}(s)$ in the propagator of $a_2$, we obtain: \begin{equation}\label{ga2} g_{a2}(s)=F_{a2}^2(Q)\, \int {\cal F}_{ij} {\cal F}_{ij}^* \frac{d\mbox{${\bf p}$}_1\,d\mbox{${\bf p}$}_2\,d\mbox{${\bf p}$}_3\,\delta^{(4)}(p_1+p_2+p_3-Q)} {2\mbox{$\varepsilon$}_12\mbox{$\varepsilon$}_2 2\mbox{$\varepsilon$}_3(2\pi)^5}\, , \end{equation} where \begin{equation}\label{gga2} {\cal F}_{ij}(s)=\frac{(\mbox{${\bf p}$}_2\times\mbox{${\bf p}$}_3)^i\mbox{${\bf p}$}_1^j}{D_{\rho}(p_2+p_3)}+ \frac{(\mbox{${\bf p}$}_1\times\mbox{${\bf p}$}_3)^i\mbox{${\bf p}$}_2^j}{D_{\rho}(p_1+p_3)}+(i\leftrightarrow j) \end{equation} \section{\bf Introduction. } Investigation of $\ensuremath{e^{+}e^{-}}$ annihilation into hadrons at low energies provides unique information about interaction of light quarks and spectroscopy of their bound states. At present the energy range below $J/\psi$ can not be satisfactorily described by QCD. Future progress in our understanding of the phenomena in this energy range is impossible without accumulation of experimental data vitally important to check the predictions of existing theoretical models. In addition, the total cross section of $\ensuremath{e^{+}e^{-}}$ annihilation into hadrons at low energies as well as the cross sections of exclusive channels are necessary for precise calculations of various effects. These include strong interaction contributions to vacuum polarization for (g-2)$_{\mu}$ and $\alpha(M_{Z}^{2})$ \cite{ej}, tests of standard model by the hypothesis of conserved vector current (CVC) relating $\ensuremath{e^{+}e^{-}} \to$ hadrons to hadronic $\tau$-lepton decays~\cite{CVC1,CVC2}, determination of the QCD parameters based on QCD sum rules~\cite{SVZ} etc. The energy behavior of the total cross section as well as that of the cross sections for exclusive channels is complicated and characteristic of various broad overlapping resonances (e.g. $\rho, \omega, \phi$ recurrencies) with numerous common decay channels having energy thresholds just in this energy range. Presence of the broad resonances in the intermediate state necessitates consideration of the quasistationary states and makes effects of their interference important. Until recently the investigation of $\ensuremath{e^{+}e^{-}}$ annihilation into hadrons was restricted by measurements of the cross sections only. Appearance of the new detectors with a large solid angle operating at high luminosity colliders and providing very large data samples opens qualitatively new possibilities for the investigation of the multihadronic production in $\ensuremath{e^{+}e^{-}}$ annihilation. Production of four pions is one of the dominant processes of $\ensuremath{e^{+}e^{-}}$ annihilation into hadrons in the energy range from 1.05 to 2.5 GeV. For the first time it was observed in Frascati~\cite{fr1} and Novosibirsk~\cite{nov1}. First experiments with limited data samples allowed one to qualitatively study the new phenomenon of multiple production of hadrons and estimate the magnitude of the corresponding cross sections. Subsequent measurements by different groups in Frascati, Orsay and Novosibirsk (see the references in~\cite{ej}) provided more detailed information on the energy dependence of the cross sections of the processes $\ensuremath{e^{+}e^{-}} \to \ensuremath{2\pi^{+}2\pi^{-}}$ and $\ensuremath{e^{+}e^{-}} \to \ensuremath{\pi^{+}\pi^{-}2\pi^{0}}$ in comparison with the previous measurements. One of the main difficulties in the experimental studies of four pion production was caused by the existence of different intermediate states via which the final state could be produced, such as \begin{eqnarray} \ensuremath{e^{+}e^{-}} & \to & \omega \pi \label{eq1}\\ \ensuremath{e^{+}e^{-}} & \to & \rho \sigma \\ \ensuremath{e^{+}e^{-}} & \to & a_{1}(1260) \pi \\ \ensuremath{e^{+}e^{-}} & \to & h_{1}(1170) \pi \\ \ensuremath{e^{+}e^{-}} & \to & \rho^{+} \rho^{-} \\ \ensuremath{e^{+}e^{-}} & \to & a_{2}(1320) \pi \\ \ensuremath{e^{+}e^{-}} & \to & \pi(1300) \pi \label{eq7} \end{eqnarray} The relative contributions of the above mentioned processes can't be obtained without the detailed analysis of the process dynamics. First attempts of this type were performed by MEA~\cite{mea} and DM1~\cite{dm1} in the energy range above 1.4 GeV and OLYA~\cite{olya2} and CMD~\cite{cmd} below 1.4 GeV and it was shown that the $\ensuremath{2\pi^{+}2\pi^{-}}$ final state is dominated by the $\rho^{0} \pi^+ \pi^-$ mechanism. ND~\cite{opi} measured the cross section of $\omega \pi$ production from 1.0 to 1.4 GeV with a magnitude which was confirmed by the subsequent $\tau$ decay studies at \nolinebreak{ARGUS~\cite{argus}} as well as by more recent results from CLEO~\cite{cleo} and ALEPH~\cite{aleph}. Later the DM2 group tried to perform partial wave analysis (PWA) of the mode with four charged pions \cite{dm2} in the energy range 1.35 to 2.40 GeV. Their analysis was based on the momentum distributions only while the angular dependence as well as interference between different waves were not taken into account. Although they obtained some evidence for the presence of $a_1(1260) \pi$ and $\rho \sigma$ states, a mechanism for a substantial part of the cross section was not determined. PWA for the mode $\ensuremath{\pi^{+}\pi^{-}2\pi^{0}}$ was not performed because of the insufficient number of completely reconstructed events. The abundance of various possible mechanisms and their complicated interference results in the necessity of simultaneous analysis of two possible final states ($\ensuremath{2\pi^{+}2\pi^{-}}$ and $\ensuremath{\pi^{+}\pi^{-}2\pi^{0}}$) which requires a general purpose detector capable of measuring energies and angles of both charged and neutral particles. The first detector of this type operating in the energy range below 1.4 GeV is the CMD-2 detector at VEPP-2M collider in Novosibirsk \cite{cmddec}. In this work, we present results from a model-dependent analysis of both possible channels in $\ensuremath{e^{+}e^{-}}$ annihilation into four pions based on data collected with the CMD-2 detector. To describe four pion production we used a simple model assuming quasitwoparticle intermediate states and taking into account the important effects of the identity of the final pions as well as the interference of all possible amplitudes. \section{\boldmath Selection of $\ensuremath{e^{+}e^{-}} \to \ensuremath{\pi^{+}\pi^{-}2\pi^{0}}$ events} \section{\boldmath Data sample and event selection} The analysis described here is based on 5.8 pb$^{-1}$ of $\ensuremath{e^{+}e^{-}}$ data collected at center-of-mass energies 2$E_{beam}$ from 1.05 up to 1.38 GeV. The data were recorded at the VEPP-2M $\ensuremath{e^{+}e^{-}}$ collider of the Budker Institute of Nuclear Physics in Novosibirsk, Russia with the CMD-2 detector in 1997. The energy range mentioned above was scanned twice with a step of 20 MeV: first by increasing energy from 1.05 to 1.37 GeV and then by decreasing energy from 1.38 to 1.06 GeV. The CMD-2 is a general purpose detector consisting of a drift chamber (DC) with about 250 $\mu$ resolution transverse to the beam and proportional Z-chamber used for trigger, both inside a thin (0.4 $X_0$) superconducting solenoid with a field of 1 T. Photons are detected in the barrel CsI calorimeter with 8-10 $\%$ energy resolution and the endcap BGO calorimeter with 6 $\%$ energy resolution. More details on the detector can be found elsewhere \cite{cmddec}. \subsection{Event selection} The present analysis is based on completely reconstructed $\ensuremath{\pi^{+}\pi^{-}2\pi^{0}}$ events. At the initial stage, events with one primary vertex with two opposite sign tracks and four or more reconstructed photons were selected. Both tracks should come from the interaction region: a distance from the track trajectories to the beam axis should be less than 0.3 cm and the vertex position along the beam axis should be inside $\pm$ 10 cm. To reject background from collinear events, the acollinearity angle between tracks in the $R-\phi$ plane should be greater than 0.1 radians. Both tracks are required in the DC fiducial volume: a $\theta$ angle should be inside $0.54\div\pi-0.54$ radians. Clusters of energy deposition that are not matched with charged track projection are paired to form $\pi^0$ candidates. These showers must have energies greater than 20 MeV, and invariant mass of the photon pair must lie within $3\sigma$ of the $\pi^0$ mass where $\sigma$ varies between $(5 \div 10)$ MeV. After that a kinematic fit was performed assuming the $\ensuremath{\pi^{+}\pi^{-}2\pi^{0}}$ hypothesis for all possible $\pi^0$ pairs. For further analysis the combination with $min(\chi^{2})$ was selected under the condition $\chi^{2}/ndf~<~2.5$. After such selection 22128 events remained in the energy range under study. \subsection{Final event sample} To understand the dynamics of the process we studied the distribution over the recoil mass for one of the $\pi^0$'s. Figure~\ref{omega} shows this distribution at the beam energy of 690 MeV. Each event gives two entries to the histogram corresponding to two $\pi^{0}$. A signal from the $\ensuremath{\omega\pi^{0}}$ final state is clearly seen. Points with errors are the data while the smooth line is our fit corresponding to the sum of a Breit-Wigner $\omega$ signal convoluted with detector resolution and a smooth combinatorial background. The detector contribution to the signal width is about 10 MeV. The number of events under the $\omega$ peak accounts for only $\sim 60\%$ of the observed events that indicates at the existence of additional intermediate states. \begin{figure}[htbp] \centering \epsfig{figure=omsignal.eps,width=0.7\linewidth} \caption{ Distribution over the $\pi^{0}$ recoil mass for $\ensuremath{\pi^{+}\pi^{-}2\pi^{0}}$ } \label{omega} \end{figure} For further analysis the data sample was subdivided into two classes: \begin{enumerate} \renewcommand{\theenumi}{\Roman{enumi}} \item $min(|M_{recoil}(\pi^{0})-M_{\omega}|)~<~70~ MeV$ \label{class1} \item $min(|M_{recoil}(\pi^{0})-M_{\omega}|)~>~70~ MeV$ \label{class2} \end{enumerate} where $M_{\omega}$ is the $\omega$ mass. The first class contains mostly $\ensuremath{\omega\pi^{0}}$ events while their admixture in the second class is relatively small, about $(1\div5)\%$ depending on the beam energy. Figs.~\ref{rhopm} and \ref{rhozero} show distributions over $M_{inv}(\pi^{\pm}\pi^{0})$ and $M_{inv}(\pi^{+}\pi^{-})$ for the events in the second class. In the spectrum of $M_{inv}(\pi^{\pm}\pi^{0})$ one can see a clear signal of $\rho^{\pm}$ while that for $M_{inv}(\pi^{+}\pi^{-})$ is relatively smooth and no signal from the $\rho^0$ is observed. The solid lines show our fit including smooth combinatorial background and gaussian $\rho$ ~signals. Presence of the $\rho^{\pm}$ signal and absence of the $\rho^{0}$ signal lead us to the natural assumption that $\rho$ mesons originate from an isospin 1 resonance. For the case of the resonance with $I=0$ one expects production of both neutral and charged $\rho$'s. \begin{figure}[htbp] \epsfig{figure=rchsignal.eps,width=0.5\textwidth} \hfill \epsfig{figure=rhzsignal.eps,width=0.5\textwidth} \\ \parbox[t]{0.47\textwidth} { \caption{ Distribution over $M_{inv}(\pi^{\pm}\pi^0)$ for $\ensuremath{\pi^{+}\pi^{-}2\pi^{0}}$ events from the class (\ref{class2}) } \label{rhopm} } \hfill \hfill \parbox[t]{0.47\textwidth} { \caption{ Distribution over $M_{inv}(\pi^+\pi^-)$ for $\ensuremath{\pi^{+}\pi^{-}2\pi^{0}}$ events from the class (\ref{class2}) } \label{rhozero} } \\ \end{figure} Possible candidates from the list of the processes (\ref{eq1})-(\ref{eq7}) are $a_{1}(1260)$, $a_{2}(1320)$ and $\pi(1300)$. These resonances have different spin and parity resulting in different angular distributions for a recoil pion (Table~\ref{reslist}). \begin{table}[hbtp] \caption{List of $I=1$ resonances with expected angular distributions} \label{reslist} \centering \renewcommand{\arraystretch}{1.1} \begin{tabular}[t]{|c|l|c|l|r|} \hline & & $J^{PC}$ & $d\sigma/d\cos(\theta)$ & $\chi^2/ndf$ \\ \hline \hline 1 & $a_{1}(1260)$ & $1^{++}$ & $const$ & $7.1/7$\\ \hline 2 & $\pi(1300)$ & $0^{-+}$ & $\sin^{2}(\theta)$ & $47.3/7$\\ \hline 3 & $a_{2}(1320)$ & $2^{++}$ & $1+\cos^{2}(\theta)$ & $37.0/7$\\ \hline \end{tabular} \end{table} To measure the angular distribution of the $\rho^{\pm}\pi^0$ system (or recoiled $\pi^{\mp}$) we fit the $\rho^{\pm}$ signal in eight ranges of the recoil $\pi^{\mp}$ angle with respect to the beam axis. The measured angular distribution is shown in Fig.~\ref{picosfit}. Three smooth curves show the fits corresponding to mentioned above hypotheses. The fit goodness is presented in the last column of Table~\ref{reslist}. One can see good agreement with the hypothesis of the $\ensuremath{a_{1}(1260)\pi} $ intermediate state. \begin{figure}[htbp] \centering \epsfig{figure=picosfit.eps,width=0.7\textwidth} \caption{ Angular distribution of the recoil $\pi^{\pm}$. Solid curves correspond to the following hypotheses: {\bf 1} -- $a_1(1260)\pi$; {\bf 2} -- $\pi(1300)\pi$; {\bf 3} -- $a_2(1320)\pi$ } \label{picosfit} \end{figure} Thus one can assume that our data can be explained by $\ensuremath{\omega\pi^{0}}$ and $\ensuremath{a_{1}(1260)\pi} $ intermediate states although small admixture of other states is not excluded. \subsection{Fitting method} For detailed comparison we perform the simulation of the processes (\ref{eq1})-(\ref{eq7}) taking into account their interference as well as the interference of the diagrams differing by permutations of identical final pions (see Appendix \ref{appa}). In order to extract a relative fraction of $\ensuremath{a_{1}(1260)\pi} $ and any other intermediate states, the unbinned maximum likelihood fit has been used \cite{kopylov}. Kinematics of each event is completely described by a set of measured four-momenta $p_{\pi}\equiv (\varepsilon , \vec{\mathbf{p}})$ of the final pions with the exception of smearing due to detector resolution and radiative effects. This set will be referred to as a vector of state $\vec{P}_{i}\equiv (p_{\pi^+},p_{\pi^-},p_{\pi^0},p_{\pi^0})_{i}$ where a subscript $i$ stands for the event number. The theoretical probability density function for $\vec{P}_{i}$ depends on process dynamics and can be expressed via the matrix element $|\mathfrak{M}(\vec{P}_{i},\vec{A})|^{2}$. Here $\vec{A}$ stands for a set of unknown model parameters like relative strength and phase of interference between the intermediate states. To obtain the optimal values of $\vec{A}$ we minimize \cite{minuit} the logarithmic likelihood function \begin{equation} \label{lfunc} L(\vec{A})=-\sum _{i=1}^{N_{2}}\ln w(\vec{P}_{i},\vec{A}) \, , \end{equation} where $w(\vec{P}_{i},\vec{A})$ is the probability to observe an event in state $\vec{P}_{i}$, and $N_{2}$ is the number of events in the second class. Since the detector resolution for the invariant mass of three pions is about 10 MeV, i.e. comparable to the $\omega$ width, smearing effects are significant for the events of the first class. Therefore we use the events of the second class only. To fix the fraction of events in the first class we add the following term to the function (\ref{lfunc}) \[ \frac{(r(\vec{A})-r_{1})^{2}}{2\sigma _{r_{1}}^{2}} \, , \] where $r_{1}=\frac{N_{1}}{N_{1}+N_{2}}$ is a measured fraction, $N_{1}$ is the number of events in the first class, and $r(\vec{A})$ is the expectation of $r_{1}$. The normalized probability density function is expressed via the matrix element \[ w(\vec{P}_{i},\vec{A}) =\frac{|\mathfrak{M}(\vec{P}_{i},\vec{A})|^{2}}{\sigma_{vis}(\vec{A})} \cdot\frac{1}{\prod_{j}2\varepsilon_{j}} \, , \] where $j$ is the particle number, and $\sigma_{vis}(\vec{A})$ is a model dependent visible cross section: \[ \sigma _{vis}(\vec{A})=\int|\mathfrak{M}(\vec{P}_{i},\vec{A})|^{2} \cdot\prod_{j=1}^{4}\frac{d^{3}\vec{\mathbf{p}}_{j}}{2\varepsilon_j} \, . \] The function $\sigma_{vis}(\vec{A})$ is calculated using Monte Carlo technique. For this purpose we use a set of events sampled according to the relativistic phase space distribution. In this case, $\sigma_{vis}(\vec{A})$ is calculated as \[ \sigma_{vis}(\vec{A})=\frac{1}{N_{MC}}\sum_{i=1}^{N_{MC}}|\mathfrak{M}(\vec{P}_{i},\vec{A})|^{2} \, , \] where $N_{MC}$ is the number of second class events in a generated set. The main goal of our fits was to find the minimal model compatible with our data. To this end, we write the matrix element as follows: \[ |\mathfrak{M}(\vec{P}_{i},\vec{A})|^{2} =|\jel{\ensuremath{\omega\pi^{0}} }+Z_{a_{1}}\cdot \jel{\ensuremath{a_{1}\pi} }+Z_{X}\cdot \jel{X}|^{2} \, , \] where $X$ stands for an admixture under study --- $\ensuremath{\rho \sigma} $, $\ensuremath{h_{1}(1170)\pi} $, $\ensuremath{\rho ^{+}\rho ^{-}}$, $\ensuremath{a_{2}(1320)\pi} $ or $\ensuremath{\pi(1300)\pi} $. In the above equation, expressions for $\jel{X}$ are taken from Appendix \ref{appa} while complex factors $Z$ are components of the vector $\vec{A}$. \subsection{\boldmath Comparison of $\ensuremath{\pi^{+}\pi^{-}2\pi^{0}}$ data with simulation} Since the matrix element for the channel $\ensuremath{\omega\pi^{0}}$ has a well-known structure, one can use the events from the first class to test the adequacy of the total MC simulation \cite{cmd2sim}. Figs.~\ref{omslide1},\ref{omslide3} show the distributions over $M_{inv}(\pi^{\pm}\pi^{0})$, $M_{recoil}(\pi^{\pm})$, $M_{inv}(\pi^{+}\pi^{-})$, $M_{inv}(\pi^{0}\pi^{0})$, $cos(\psi_{\pi^{\pm}\pi^{0}})$, $cos(\psi_{\pi^{+}\pi^{-}})$, $cos(\psi_{\pi^{0}\pi^{0}})$ and $M_{recoil}(\pi^{0})$ for the events of the first class at the beam energy of 690 MeV. Points with errors are the data, while the histograms correspond to the simulation of the processes $\ensuremath{\omega\pi^{0}}$ and $\ensuremath{a_{1}(1260)\pi} $. Good consistence of the data and MC makes us confident that MC simulation adequately reproduces both the kinematics of produced particles and the detector response to them. Similar distributions for the events of the second class are shown in Figs.~\ref{a1slide1},\ref{a1slide3}. One can see that the process $\ensuremath{\pi^{+}\pi^{-}2\pi^{0}}$ is satisfactorily described in the minimal model in which there are two intermediate states $\ensuremath{\omega\pi^{0}}$ and $\ensuremath{a_{1}(1260)\pi} $ only. Similar consistence is observed at other energies: we have also examined the energy points $2E_{beam}=(1.28\pm 0.01)$ and $(1.18\pm 0.03)$ GeV. To determine the admixture of other possible mechanisms we extend the minimal model above by adding each of the other states one by one and performing the fit. The results of these fits are shown in Table~\ref{admix}. From its third column one can see that the relative fractions $r_X$ of the additional intermediate state $X$ with respect to the $\ensuremath{a_{1}(1260)\pi} $ are small. This confirms our assumption of the $\ensuremath{a_{1}(1260)\pi} $ dominance. \begin{table}[htbp] \caption{The results of fit to different models with upper limits} \label{admix} \centering \renewcommand{\arraystretch}{1.4} \begin{tabular}{|c|c|c|c|} \hline Model& \( L_{min}/N_{ev} \)& $r_X$, $\%$ & Upper limit, \(\%\)\\ \hline \hline \( \ensuremath{\omega\pi^{0}} +\ensuremath{a_{1}\pi} \)& \( 1264/891 \)& ---& --- \\ \hline \( \ensuremath{\omega\pi^{0}} +\ensuremath{a_{1}\pi} +\ensuremath{\rho \sigma} \)& \( 1256/891 \)& \( 2.1_{-0.9}^{+1.2} \)& \( 4.3 \)\\ \hline \( \ensuremath{\omega\pi^{0}} +\ensuremath{a_{1}\pi} +\ensuremath{h_{1}\pi} \)& \( 1263/891 \)& \( 0.1_{-0.1}^{+0.2} \)& \( 0.4 \)\\ \hline \( \ensuremath{\omega\pi^{0}} +\ensuremath{a_{1}\pi} +\ensuremath{a_{2}\pi} \)& \( 1263/891 \)& \( 0.2_{-0.2}^{+0.4} \)& \( 0.8 \)\\ \hline \( \ensuremath{\omega\pi^{0}} +\ensuremath{a_{1}\pi} +\ensuremath{\pi^{\prime} \pi} \)& \( 1250/891 \)& \( 9.5_{-2.8}^{+3.2} \)& \( 15. \)\\ \hline \( \ensuremath{\omega\pi^{0}} +\ensuremath{a_{1}\pi} +\ensuremath{\rho ^{+}\rho ^{-}} \)& \( 1246/891 \)& \( 4.7_{-1.6}^{+2.0} \)& \( 7.7 \)\\ \hline \end{tabular} \end{table} Because of the theoretical uncertainty of the $\ensuremath{a_{1}(1260)\pi} $ matrix element, we do not consider the nonnegligible magnitude of the $\ensuremath{\rho ^{+}\rho ^{-}}$ and $\ensuremath{\pi(1300)\pi} $ contributions as significant. Instead, we prefer to set upper limits for the fraction of these intermediate states. The values of the likelihood function~(\ref{lfunc}) for optimal parameters ($L_{min}$) do not contradict to our expectation based on the simulation of $\ensuremath{\omega\pi^{0}}$ and $\ensuremath{a_{1}(1260)\pi} $. \begin{figure}[htbp] \epsfig{figure=slide1_omega_690.eps,width=1.0\textwidth} \caption{ Distributions over $M_{inv}(\pi^{\pm}\pi^{0})$, $M_{recoil}(\pi^{\pm})$, $M_{inv}(\pi^{+}\pi^{-})$, $M_{inv}(\pi^{0}\pi^{0})$ for $\ensuremath{\pi^{+}\pi^{-}2\pi^{0}}$ events from the class (\ref{class1}) } \label{omslide1} \end{figure} \begin{figure}[htbp] \epsfig{figure=slide3_omega_690.eps,width=1.0\textwidth} \caption{ Distributions over $cos(\psi_{\pi^{\pm}\pi^{0}})$, $cos(\psi_{\pi^{+}\pi^{-}})$, $cos(\psi_{\pi^{0}\pi^{0}})$ and $M_{recoil}(\pi^{0})$ for $\ensuremath{\pi^{+}\pi^{-}2\pi^{0}}$ events from the class (\ref{class1}) } \label{omslide3} \end{figure} \begin{figure}[htbp] \epsfig{figure=slide1_a1_690.eps,width=1.0\textwidth} \caption{ Distributions over $M_{inv}(\pi^{\pm}\pi^{0})$, $M_{recoil}(\pi^{\pm})$, $M_{inv}(\pi^{+}\pi^{-})$, $M_{inv}(\pi^{0}\pi^{0})$ for $\ensuremath{\pi^{+}\pi^{-}2\pi^{0}}$ events from the class (\ref{class2}) } \label{a1slide1} \end{figure} \begin{figure}[htbp] \epsfig{figure=slide3_a1_690.eps,width=1.0\textwidth} \caption{ Distributions over $cos(\psi_{\pi^{\pm}\pi^{0}})$, $cos(\psi_{\pi^{+}\pi^{-}})$, $cos(\psi_{\pi^{0}\pi^{0}})$ and $M_{recoil}(\pi^{0})$ for $\ensuremath{\pi^{+}\pi^{-}2\pi^{0}}$ events from the class (\ref{class2}) } \label{a1slide3} \end{figure} \subsection{\boldmath Comparison of $\ensuremath{2\pi^{+}2\pi^{-}}$ data with simulation} Consider now the $\ensuremath{2\pi^{+}2\pi^{-}}$ channel. In this case we don't have $\ensuremath{\omega\pi^{0}}$ in the intermediate state and therefore no additional complicacy to check the assumption of the $a_{1}(1260)\pi$ dominance arises. To this end four-track events were selected. All tracks should come from the interaction region: a distance from track trajectories to the beam axis should be less than 1~cm and the vertex position along the beam axis should be inside $\pm$ 15~cm. After that a kinematic fit was performed assuming the $\ensuremath{2\pi^{+}2\pi^{-}}$ hypothesis and events with $\chi^{2}/ndf~<~2.5$ were selected. Under these conditions 28552 events remain in the energy range under study. Figure~\ref{pi4slide} shows distributions over $M_{inv}(\pi^{+}\pi^{-})$, $M_{inv}(\pi^{\pm}\pi^{\pm})$, $M_{recoil}(\pi^{\pm})$ and $cos(\psi_{\pi^{+}\pi^{-}})$ for \ensuremath{2\pi^{+}2\pi^{-}} case. One can see that the hypothesis of the $a_{1}(1260)\pi$ dominance does not contradict to the data although one should note a slight deviation in the spectrum of the invariant masses of the likesign pions. This deviation can be possibly explained by the contribution of the $D-wave$ or some final state interaction. It can't be explained by the admixture of other possible processes because their fractions compared to $a_{1}(1260)\pi$ are small (see Table~\ref{admix}). The study of more complicated cases taking into account simultaneously a few intermediate states in addition to $\ensuremath{\omega\pi^{0}}$ and $\ensuremath{a_{1}(1260)\pi} $ is now in progress. \begin{figure}[htbp] \epsfig{figure=slide4_4pi_690.eps,width=1.0\textwidth} \caption{ Distributions over $M_{inv}(\pi^{+}\pi^{-})$, $M_{inv}(\pi^{\pm}\pi^{\pm})$, $M_{recoil}(\pi^{\pm})$ and $cos(\psi_{\pi^{+}\pi^{-}})$ for $4\pi^{\pm}$ events } \label{pi4slide} \end{figure} \section{Energy dependence of the cross sections} To obtain the $\ensuremath{a_{1}(1260)\pi} $ contribution to the total cross section, we subtracted from the latter the contribution of the $\ensuremath{\omega\pi^{0}}$ intermediate state. Strictly speaking, such a procedure loses its meaning when the interference between $\ensuremath{\omega\pi^{0}}$ and $\ensuremath{a_{1}(1260)\pi} $ is strong. In our case, however, the effects of the interference is numerically small ($\sim 5\%$) because of the small width of the $\omega$ meson. In order to diminish the influence of background from soft photons, the additional cut was applied: $cos(\psi)<0.7$ where $\psi$ is the angle between the photon direction in the $\pi^{0}$ rest frame and $\pi^0$ momentum. Cross sections were calculated from the formulae: \begin{eqnarray*} \sigma_{\omega\pi^0} & = & \frac{N_{\omega}} {B(\omega \to 3\pi)\cdot (1+\delta)\cdot L\cdot \epsilon} \; , \\ \sigma_{\ensuremath{\pi^{+}\pi^{-}2\pi^{0}}} & = & \frac{N_{a_{1}}} {(1+\delta)\cdot L\cdot \epsilon} \; , \\ \end{eqnarray*} where $L$ is the luminosity, $\epsilon$ is the detection efficiency from MC, and $\delta$ is a radiative correction calculated according to \cite{kurfad}. $N_{\omega}$ and $N_{a_{1}}$ are the numbers of events due to the corresponding reactions ($\ensuremath{e^{+}e^{-}} \to \ensuremath{\omega\pi^{0}}$ and $\ensuremath{e^{+}e^{-}} \to \ensuremath{a_{1}\pi} $ respectively) under the assumption that there is no interference. The values of $N_{\omega}$ and $N_{a_{1}}$ were determined from the equations: \begin{eqnarray*} N_1 & = & N_{\omega}\cdot \alpha + N_{a_1}\cdot(1-\beta) \, , \\ N_2 & = & N_{\omega}\cdot(1-\alpha) + N_{a_1}\cdot\beta \, , \\ \end{eqnarray*} where $N_{1}$ and $N_{2}$ are the numbers of events in the first and second classes respectively, $\alpha$ is the probability for the $\ensuremath{\omega\pi^{0}}$ event to enter the first class, and $\beta$ is the probability for the $\ensuremath{a_{1}(1260)\pi} $ event to enter the second class. The values of $\alpha$ and $\beta$ as a function of energy were taken from MC simulation. Figure~\ref{xsection} shows the cross sections obtained for $\ensuremath{e^{+}e^{-}} \to \ensuremath{\omega\pi^{0}}$ and $\ensuremath{e^{+}e^{-}} \to \ensuremath{\pi^{+}\pi^{-}2\pi^{0}}$ with the $\ensuremath{\omega\pi^{0}}$ contribution subtracted as explained above. Both cross sections rise with energy while the relative fraction of the $\ensuremath{a_{1}(1260)\pi} $ increases. \begin{figure}[htbp] \centering \epsfig{figure=xsection.eps,width=0.6\textwidth} \caption{ Energy dependence of $\sigma(\ensuremath{e^{+}e^{-}} \to \ensuremath{\omega\pi^{0}})$ and $\sigma(\ensuremath{e^{+}e^{-}} \to \ensuremath{\pi^{+}\pi^{-}2\pi^{0}})$ where the contribution of $\ensuremath{\omega\pi^{0}}$ is subtracted } \label{xsection} \end{figure} Figure~\ref{xs2pi} shows the total cross section vs energy. Only statistical errors are shown. The systematic uncertainty for this channel consists of the contributions from the uncertainties in the event reconstruction, radiative corrections and luminosity determination. The overall systematic uncertainty was estimated to be $\sim~15\%$. \begin{figure}[htbp] \centering \epsfig{figure=xs2pi.eps,width=\textwidth} \caption{ Energy dependence of the $\ensuremath{\pi^{+}\pi^{-}2\pi^{0}}$ cross section} \label{xs2pi} \end{figure} The cross section measured in our experiment is consistent with the previous measurement at OLYA \cite{olya1} and within systematic errors does not contradict to the recent result from SND \cite{snd98}. The value of the cross section from all three groups is significantly lower than that from ND \cite{nd,ndr}. Also shown above 1.4 GeV are results from Orsay \cite{dm2,m3n} and Frascati \cite{gg2} groups. To determine the cross section $\ensuremath{e^{+}e^{-}} \to \ensuremath{2\pi^{+}2\pi^{-}}$ the following criteria were added: \begin{itemize} \item $\chi^{2}/ndf~<~7$ \item $\theta_{min}~>~0.67$ \item $|\sum_{i} E_{i}~-~2\cdot E_{beam}|~<~0.2 \cdot E_{beam}$ \item $|\sum_{i} \overrightarrow{p_{i}}|~<~0.3 \cdot E_{beam}$ \end{itemize} where $E_{i}$, $\overrightarrow{p_{i}}$ are pion energies and momenta before the kinematic reconstruction, and $\theta_{min}$ is a minimal angle of pions with respect to the beam axis. Figure~\ref{xs4pi} shows the total cross section vs energy. Only statistical errors are shown. The systematic uncertainty for this channel consists of the contributions from the uncertainties in the event reconstruction, radiative corrections, selection criteria and luminosity determination. The overall systematic uncertainty was estimated to be $\sim~15\%$. \begin{figure}[htbp] \centering \epsfig{figure=xs4pi.eps,width=\textwidth} \caption{ Energy dependence of the $\ensuremath{2\pi^{+}2\pi^{-}}$ cross section} \label{xs4pi} \end{figure} The obtained cross section is consistent with the previous measurements at OLYA \cite{olya1}, ND \cite{nd,ndr} and CMD \cite{cmd} and within systematic errors does not contradict to the recent result from SND \cite{snd98}. Also shown above 1.4 GeV are results from Orsay \cite{m3n,dm1,dm2} and Frascati \cite{mea,gg2:xs} groups. Figure~\ref{xsratio} presents the ratio of the cross sections $\sigma(\ensuremath{e^{+}e^{-}} \to \ensuremath{2\pi^{+}2\pi^{-}})$ and $\sigma(\ensuremath{e^{+}e^{-}} \to \ensuremath{\pi^{+}\pi^{-}2\pi^{0}})$ where the contribution of $\omega\pi^{0}$ is subtracted. The solid curve shows the theoretical prediction based on the $a_{1}(1260)\pi$ dominance. \begin{figure}[htbp] \centering \epsfig{figure=xsratio.eps,width=0.7\textwidth} \caption{ Energy dependence of the ratio $\frac{\sigma(\ensuremath{e^{+}e^{-}}\to\ensuremath{2\pi^{+}2\pi^{-}})} {\sigma(\ensuremath{e^{+}e^{-}}\to\ensuremath{\pi^{+}\pi^{-}2\pi^{0}})}$. The contribution of $\omega\pi^{0}$ is subtracted. The solid curve shows the theoretical prediction based on the $a_{1}(1260)\pi$ dominance } \label{xsratio} \end{figure} In the energy range under consideration the experimentally measured ratio is about two. Calculations show that this ratio tends to four at very low energies close to the threshold of four pion production when the interference effects are maximal. The effects of interference decreases with energy since the produced resonances are moving at a high velocity. As a result, the ratio falls with energy and tends to unity. The importance of the interference between the different diagrams corresponding to permutations of identical pions was first noted in \cite{ei}. The neglection of interference can lead to wrong conclusions about the contributions of different intermediate states to the total cross section. \section{Conclusion} The detailed analysis of the process $\ensuremath{e^{+}e^{-}}\to\ensuremath{\pi^{+}\pi^{-}2\pi^{0}}$ unambiguously demonstrates that the dominant contribution to the cross section comes from $\ensuremath{\omega\pi^{0}}$ and $\rho^{\pm}\pi^{\mp}\pi^0$ intermediate states whereas the $\rho^0\pi^0\pi^0$ state is not observed. Moreover, the $\rho^{\pm}\pi^{\mp}\pi^0$ state is completely saturated by the $\ensuremath{a_{1}(1260)\pi} $ mechanism. The latter also dominates in the cross section of the process $\ensuremath{e^{+}e^{-}}\to\ensuremath{2\pi^{+}2\pi^{-}}$. This conclusion is based on the data sample corresponding to the integrated luminosity of 5.8 $pb^{-1}$ collected with the CMD-2 detector in the energy range 1.05 -- 1.38 GeV. Comparison of the experimental data with the calculations shows that account of the interference of different amplitudes with various intermediate resonances but the identical final state drastically changes predictions for the differential distributions and total cross sections. Through the hypothesis of CVC \cite{CVC1} experimentally tested to be valid within the 3-5$\%$ accuracy \cite{CVC2}, we can relate the values of the four pion cross sections measured in $\ensuremath{e^{+}e^{-}}$ to the hadronic spectra in corresponding four pion decays of the $\tau$-lepton. Moreover, our observation of the $a_{1}(1260) \pi$ dominance should be also applied to $\tau$-lepton decays. One should note that such a conclusion is in contradiction to that made by ARGUS collaboration \cite{argus}. However, their conclusion based on the distribution over the dipion mass is not compatible with the recent data from ALEPH \cite{aleph}. The model used by ARGUS didn't take into account the interference of the amplitudes corresponding to various intermediate states whereas our observation unambiguously points to significance of the interference effects. Our predictions for various two- and three-body invariant mass spectra will be addressed in more detail in the forthcoming paper which will compare them to the existing $\tau$-lepton data from ARGUS~\cite{argus}, CLEO~\cite{cleo} and ALEPH~\cite{aleph}. The detailed investigation of the energy behavior of exclusive channels of $\ensuremath{e^{+}e^{-}}$ annihilation into hadrons is crucial for understanding the structure of the initial hadronic state created by the virtual photon ($\rho(1450)$, $\rho(1700)$ etc.). Our work considers this problem in the energy range below 1.4 GeV. At higher energy, such analysis can't be performed because $\ensuremath{e^{+}e^{-}}$ annihilation data are not available now. One can hope to move forward by using data on semihadronic decays of the $\tau$-lepton. Our statement about the $\ensuremath{a_{1}(1260)\pi} $ dominance can be used for verification of theoretical models in which the problem of intermediate states has been discussed. In particular, recently there were numerous speculations about the possible existence of vector hybrids claiming that hybrids can be distinguished from the normal $q\bar{q}$ states by their decays, e.g. in the $\ensuremath{a_{1}(1260)\pi} $ \cite{hybrid}. \section*{Acknowledgements} The authors are grateful to the staff of VEPP-2M for excellent performance of the collider, to all engineers and technicians who participated in the design, commissioning and operation of CMD-2. Special thanks are due to N.N.~Achasov, M.~Benayoun, V.L.~Chernyak, V.F.~Dmitriev, V.S.~Fadin, V.B.~Golubev, I.B.~Khriplovich, L.~Montanet, S.I.~Serednyakov, A.I.~Vainshtein for numerous discussions and constant interest.
\section{Introduction} Restricted Trace Ensembles (RTEs) introduced a long time ago in \cite{met} are interesting for a couple of reasons. They possess compact support not only for infinite but also for finite $n$, where $n$ is the size of the matrix. In the canonical ensemble eq. (\ref{Pcan}) large values of matrix elements are only exponentially suppressed whereas in the RTEs a sharp cutoff is introduced. For this reason the latter can be regarded as the corresponding microcanonical ensemble. Much of the relevance of random matrix theory is related to universality properties of connected correlators in the large $n$ limit, that is their independence from the details of the probability density which defines the matrix ensemble. The most famous property is the limiting form of connected density-density correlator at ``short distances'', also called the ``sine law''. Very interesting is also a global universality property: it was found that smoothed connected correlators may be expressed by the same universal function. The original derivation made use of loop equations \cite{AJM}, it was later rediscovered by diagrammatical expansion \cite{BZ}. All derivations are valid for canonical ensembles with an arbitrary polynomial, therefore it was generally believed that this global universality property holds for all probability densities invariant under unitary transformations. In this note we investigate two major questions, namely whether the RTEs also possess universal global correlations, which are independent of the details of the distribution, and, second whether they are equivalent to the universality classes of the canonical ensemble. Notice that usual techniques as orthogonal polynomials fail for RTEs because of the additional constraint on the matrix-trace. In order to address to the above problems in a previous publication \cite{ACMV} we have introduced the following generalization of the RTEs \beqn {\cal P}_\phi(M)&\equiv&\frac{1}{{\cal Z}_\phi}\,\phi \left( A^{2}-\frac{1}{n} \mbox{Tr} V(M)\right) \ \ \ \ \ \ , \ V(M)=\sum_{l=1}^p g_{2l}M^{2l} \ \ , \nonumber\\ {\cal Z}_\phi &\equiv& \int {\cal D}M\, \phi \left(A^{2}-\frac{1}{n}\mbox{Tr} V(M)\right) \ \ , \label{PRTE} \eeqn where $\phi(x)=\delta(x)$ or $\theta(x)$, and compared them to the canonical ensemble \beqn {\cal P}(M)\equiv \frac{1}{{\cal Z}}\,\exp \left[-ng\mbox{Tr} V(M)\right] \ \ ,\ {\cal Z} \equiv \int {\cal D}M \, \exp \left[-ng\mbox{Tr} V(M)\right] \ \ . \label{Pcan} \eeqn We have calculated the spectral density $\rho(\lambda)=<1/n\mbox{Tr}\delta(\lambda-M)>$ of RTEs, which is equivalent to the one-point resolvent $G(z)=<1/n\mbox{Tr} (z-M)^{-1}>$. Comparing it to the canonical ensemble we have shown, that in the large-$n$ limit they agree provided that the scale factor $g$ takes a well defined value determined by the values of the couplings $g_{2l}$ in the potential $V(M)$ and by $A^2$. This holds despite the well known fact that the the spectral density itself is non-universal. From the factorization property of correlators at large-$n$ we then concluded that all finite moments of the three ensembles coincide. The question now is whether this equivalence holds also for the connected part of higher correlation functions and thus for higher orders in $1/n$. Therefore in this letter we investigate all $k$-point correlators. We start with $k=2$: \begin{equation} G_\phi (z,w) \equiv \left\langle\frac{1}{n}\mbox{Tr} \frac{1}{z-M} \frac{1}{n}\mbox{Tr}\frac{1}{w-M}\right\rangle_\phi \ - \ \left\langle\frac{1}{n}\mbox{Tr}\frac{1}{z-M}\right\rangle_\phi \left\langle\frac{1}{n}\mbox{Tr}\frac{1}{w-M}\right\rangle_\phi \ , \label{G2} \end{equation} where the subscript $\phi$ indicates the corresponding average. Here we have subtracted the factorized part. The 2-point correlator as well as all higher $k$-point correlators are known to be universal for the canonical ensemble \cite{AJM}. There, the subtraction corresponds to taking into account only connected diagrams of random surfaces. The corresponding connected density-density correlators can be obtained by taking the appropriate imaginary part, as given for example in \cite {AA} \footnote{In contrast to ref. \cite{AA} we define the $k$-point resolvents without a factor of $n^{2k-2}$.}. \setcounter{equation}{0} \section{Non-universality of $G_\phi(z,w)$} In the following we shall restrict ourselves to purely monomial potentials $V(M)=M^{2p}$. Since we want to show that the correlator $G_\phi(z,w)$ is {\it non-universal}, in principle only two examples of different potentials leading to a different correlator eq. (\ref{G2}) would be sufficient. In a first step we will show that all expectation values of products of matrices have a $1/n^2$-expansion for the RTEs. When relating the averages to the corresponding canonical ones we can explicitly extract their expansion coefficients, which enables us to calculate $G_\phi(z,w)$. It is useful to introduce the following representation for the $\delta$- and $\theta$-function \begin{equation} \phi (x)=\int_{-\infty}^\infty \frac{dy}{2\pi \frac{e^{(iy+\epsilon)x}}{(iy+\epsilon)^s} \ \ , \left\{ \begin{array}{ll} s=0 & \phi(x)=\delta(x)\\ s=1 & \phi(x)=\theta(x) \end{array} \right.\ , \label{phi} \end{equation} where $\epsilon=0^{+}$ is a small and harmless regulator which makes possible to interchange integrals. Next we calculate the matrix integral \begin{equation} I^{\{k\}}_\phi(n,A)\equiv\int {\cal{D}}M\,\phi \left( A^{2}-\frac{1}{n} \mbox{Tr} M^{2p}\right) \;M_{\alpha _{1}\beta _{1}}M_{\alpha _{2}\beta _{2}} \ldots M_{\alpha _{k}\beta _{k}} \ \ , \label{Idef} \end{equation} where the superscript $\{k\}$ summarizes the dependence on all the matrix indices. The volume element of the Hermitean $(n\times n)$-matrices $\prod_i dM_{ii} \prod_{i<j} d \mbox{Re} M_{ij} \prod_{i<j} d \mbox{Im} M_{ij}$ is the usual product of independent entries. In the particular case $k=0$ eq. (\ref{Idef}) is just the partition function ${\cal Z}_\phi$. By inserting the representation eq. (\ref{phi}) into the eq. (\ref{Idef}) and exchanging the order of integrations, we exhibit that (\ref{Idef}) is actually proportional to the analogous integral with canonical measure. Indeed \begin{equation} I_\phi^{\{k\}}(n,A) = \int_{-\infty }^\infty\frac{dy}{2\pi} e^{(iy+\epsilon)A^{2}} \frac{1}{(iy+\epsilon)^s} \int {\cal{D}}M\;e^{- [(iy+\epsilon)/n] \mbox{Tr} M^{2p}}\;M_{\alpha _{1}\beta_{1}} \ldots M_{\alpha _{k}\beta _{k}} \ . \label{2.1} \end{equation} The explicit dependence of the matrix integral \begin{equation} \int {\cal{D}}M\;e^{- a \, \mbox{Tr} M^{2p}}\;M_{\alpha _{1}\beta_{1}} \ldots M_{\alpha _{k}\beta _{k}} = a^{-\frac{n^2+k}{2p} } \int {\cal{D}}M\;e^{- \mbox{Tr} M^{2p}}\;M_{\alpha _{1}\beta_{1}} \ldots M_{\alpha _{k}\beta _{k}} \label{2.2} \end{equation} on the real positive parameter $a$, allows analytic continuation in the whole complex plane, cut along the negative part of the real axis. This provides a definition for (\ref{2.1}) and we obtain \beqn I_\phi^{\{k\}}(n,A) &=& \int_{-\infty }^\infty\frac{dy}{2\pi} e^{(iy+\epsilon)A^{2}} \frac{1}{(iy+\epsilon)^s} \left( \frac{gn^2}{iy+\epsilon}\right) ^{\frac{n^{2}+k}{2p}} \int {\cal{D}}M\;e^{-ng\mbox{Tr} M^{2p}}\;M_{\alpha _{1}\beta_{1}} \ldots M_{\alpha _{k}\beta _{k}} \nonumber\\ &=& \left(gn^2A^{2}\right)^{\frac{n^{2}+k}{2p}} \frac{\left( A^2\right)^{s-1}}{\Gamma \left( \frac{n^{2}+k}{2p}+s\right)}\; \int {\cal{D}}M\;e^{-ng\mbox{Tr} M^{2p}}\;M_{\alpha _{1}\beta_{1}} \ldots M_{\alpha _{k}\beta _{k}}\ \ . \label{I} \eeqn where $g>0$ and in the second step we have used Hankel's contour integral for the Gamma function \cite{Abramo}. As a consequence we obtain the RTE average expressed by the canonical average \begin{equation} \left\langle M_{\alpha _{1}\beta _{1}}\ldots M_{\alpha _{k}\beta _{k}}\right\rangle _\phi \ =\ \frac{I_\phi^{\{k\}}(n,A)}{I_\phi^{\{0\}}(n,A)} \ =\ s_{n,k}(g,A) \left\langle M_{\alpha _{1}\beta _{1}}\ldots M_{\alpha _{k}\beta _{k}}\right\rangle \ , \label{mom} \end{equation} where \begin{equation} s_{n,k}(g,A) \ = \ \left( gn^2A^2\right) ^\frac{k}{2p}\frac{\Gamma \left( \frac{n^{2}}{2p}+s\right) }{\Gamma \left( \frac{n^{2}+k}{2p}+s\right)} \ . \label{sn} \end{equation} On the r.h.s. of eq. (\ref{mom}) we average with the canonical measure eq. (\ref{Pcan}) for $V(M)=gM^{2p}$. The exact relation (\ref{mom}) may be exploited to relate the parameters of the RTEs to the parameters of the canonical model, so that at leading order in the large-$n$ limit all moments of the form (\ref{mom}) are identical. It is however impossible to relate the parameters to obtain that the scaling factor $s_{n,k}(g,A)$ is unity up to order $O(1/n^4)$. Indeed, if we assume \begin{equation} A^2 = \frac{1}{2 p g } + \frac{x}{n^2} + O\left(\frac{1}{n^4}\right) \label{2.3} \end{equation} and use the relation for ratios of Gamma functions \cite{Abramo}, we obtain \beqn s_{n,k}(g,A) &=& (2p\ g A^2)^\frac{k}{2p} \left[ 1 - \frac{k}{2 n^2}\left( \frac{k}{2p}+2s-1\right) + O\left(\frac{1}{n^4}\right)\right] \nonumber \\ &=& 1+\frac{k}{n^2}\left( g x - \frac{k}{4 p} -s +\frac{1}{2} \right) + \ O\left(\frac{1}{n^4}\right) \ \ . \label{sexp} \eeqn This shows that, with the general relation (\ref{2.3}), the $1/n^2$ expansion of the canonical measure translates into a $1/n^2$ expansion for the RTEs \footnote{Using Stirling's formula one can easily convince oneself that $z^{b-a}\frac{\Gamma(z+a)}{\Gamma(z+b)}$ has an expansion in $1/z$.} . The non-vanishing contribution at order $1/n^2$ in the scaling factor $s_{n,k}(g,A)$ with the $k^2$-dependence will be shown to lead to the non-universality of connected correlators. The relation between the coefficients $c_{\{k\}}^{(j)}$ of the topological $1/n^2$-expansion in the canonical ensemble are simply related to the corresponding coefficients $d_{\{k\}}^{(j)}$ for the RTEs \begin{equation} \left\langle M_{\alpha _{1}\beta _{1}}\ldots M_{\alpha _{k}\beta _{k}}\right\rangle \ =\ \sum_{j=0}^\infty c_{\{k\}}^{(j)} \frac{1}{n^{2j}} \ \ , \ \ \left\langle M_{\alpha _{1}\beta _{1}}\ldots M_{\alpha _{k}\beta _{k}}\right\rangle_\phi \ =\ \sum_{j=0}^\infty d_{\{k\}}^{(j)}\frac{1}{n^{2j}} \label{ddef} \end{equation} through eq. (\ref{sexp}) \beqn d_{\{k\}}^{(0)} &=& c_{\{k\}}^{(0)} \ \ ,\nonumber\\ d_{\{k\}}^{(1)} &=& c_{\{k\}}^{(1)} +k \left( g x -\frac{k}{4p}-s+\frac{1}{2}\right)c_{\{k\}}^{(0)} \ \ , \label{dc} \eeqn where we recall that we have $s=0,1$ for $\phi=\delta,\theta$ and the subscript $\{k\}$ summarizes all matrix indices. Eq. (\ref{dc}) immediately implies the identity of the one-point resolvents \cite{ACMV} \beqn G_\phi(z) &=& \frac{1}{n}\sum_{k=0}^\infty\frac{1}{z^{k+1}} \left\langle\mbox{Tr} M^k\right\rangle_\phi \ = \ \frac{1}{n}\sum_{k=0}^\infty\frac{1}{z^{k+1}}\left(d_{\{k\}}^{(0)} +O\left(\frac{1}{n^2}\right)\right) \nonumber\\ &\stackrel{n\to\infty}{\longrightarrow}& G(z) \ \ , \label{GG1} \eeqn which has been shown in \cite{ACMV} for a larger class of potentials. Note that $G_\phi(z)$ is of order 1 since $d_{\{k\}}^{(0)}$ contains a power of $n$ from the trace. Next we turn to the two-point resolvent $G_\phi(z,w)$. Inserting eq. (\ref{dc}) into the definition (\ref{G2}) we obtain \beqn G_\phi(z,w) &=&\frac{1}{n^{2}}\sum_{k,l=0}^{\infty }\frac{1}{z^{k+1}w^{l+1}} \left(\left\langle \mbox{Tr} M^{k}\mbox{Tr} M^{l}\right\rangle _\phi -\left\langle \mbox{Tr} M^{k}\right\rangle _\phi\left\langle \mbox{Tr} M^{l}\right\rangle _\phi\right) \nonumber\\ &=&\frac{1}{n^{2}}\sum_{k,l=0}^{\infty }\frac{1}{z^{k+1}w^{l+1}} \left( d_{\{ k,l\} }^{(0)}+d_{\{ k,l\} }^{(1)}\frac{1}{n^{2}} - \left[ d_{\{ k\} }^{(0)}+d_{\{ k\} }^{(1)}\frac{1}{n^{2}}\right] \left[ d_{\{ l\}}^{(0)}+d_{\{ l\} }^{(1)}\frac{1}{n^{2}}\right] + O\left(\frac{1}{n^{4}}\right)\right) \nonumber\\ &=&\frac{1}{n^{4}}\sum_{k,l=0}^{\infty }\frac{1}{z^{k+1}w^{l+1}} \left( c_{\{ k,l\} }^{(1)} - c_{\{ k\}}^{(0)}c_{\{ l\} }^{(1)} - c_{\{ k\}}^{(1)}c_{\{ l\} }^{(0)} - \frac{kl}{2p}c_{\{ k\}}^{(0)}c_{\{ l\} }^{(0)} + O\left(\frac{1}{n^{2}}\right)\right)\ . \label{GG2a} \eeqn Here we have made use of the fact that \beqn d_{\{ k,l\} }^{(0)} &=& c_{\{ k,l\} }^{(0)} = c_{\{k\} }^{(0)}c_{\{ l\} }^{(0)} \ , \nonumber \\ d_{\{ k,l\} }^{(1)} &=& c_{\{ k,l\} }^{(1)} +(k+l)\left( g x -\frac{k+l}{4p} -s +\frac{1}{2} \right) c_{\{ k,l\} }^{(0)} \ \ . \label{dcc} \eeqn As a consequence the leading terms in $G_\phi(z,w)$ cancel as they should, leaving the remaining part of order $1/n^2$ when counting properly factors of $n$ from the traces. Remarkably the result (\ref{GG2a}) does not depend upon the values of $x$ and $s$. The first three terms in the last line of eq. (\ref{GG2a}) give precisely the universal two-point resolvent of the canonical ensemble. The last term is new and can be written as a product of derivatives of the one-point resolvent $G(z)$ $(=G_\phi(z))$. The final result reads \begin{equation} n^2G_\phi(z,w)\ \stackrel{n\to\infty}{\longrightarrow}\ n^2G(z,w) \ -\ \frac{1}{2p}\partial_z(zG(z))\partial_w(wG(w)) \ \ , \label{GG2} \end{equation} which holds for all monomial potentials $V(M)=gM^{2p}$ , both $\phi=\delta,\theta$ and $A^2$ given by eq.(\ref{2.3}). Note that all terms in eq. (\ref{GG2}) are of order 1. The first term is the well known universal two-point resolvent \cite{AJM} \beqn n^2G(z,w) &=& \frac{1}{2(z-w)^2}\left( \frac{zw \ -\ a^2}{\sqrt{(z^2-a^2)(w^2-a^2)}} \ -\ 1\right) \ , \label{Guniv} \eeqn where $a$ denotes the support of the eigenvalues $[-a,\ a]$. The notion of universality means that eq. (\ref{Guniv}) is the same for any given polynomial potential sharing the same support \cite{AJM}. We only have to assume that the couplings $g_{2l}$ are such that the support is one arc. This is true in particular for the monomial potentials with $g>0$. The second term in eq. (\ref{GG2}), however, is {\it non-universal} as the one-point resolvent itself is non-universal. Let us give two examples, the Gaussian and the purely quartic potential: \beqn \underline{V(M)=gM^2}:\ \ \ G(z)&=&g\left(z - \sqrt{z^2-a^2}\right) \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ , \ a^2 = \frac{2}{g} \ ,\nonumber\\ \underline{V(M)=gM^4}:\ \ \ G(z)&=&g \left(2z^3 - (2z^2+a^2)\sqrt{z^2-a^2}\right) \ \ , \ a^4 = \frac{4}{3g} \ . \label{Gexpl} \eeqn Although in this case we have a potential depending only on one parameter $g$, which is thus in one to one correspondence to the endpoint of support $a$, the two resolvents in eq. (\ref{Gexpl}) are different {\it functions} of $z$. Inserting them into the result for $G_\phi(z,w)$ eq. (\ref{GG2}) we obtain \beqn \underline{V(M)=gM^2}&:& \nonumber\\ n^2G_\phi(z,w)&=& n^2G(z,w) - g^2 \left(2z-\frac{2z^2-a^2}{\sqrt{z^2-a^2}}\right) \left(2w-\frac{2w^2-a^2}{\sqrt{w^2-a^2}}\right) \ , \nonumber\\ \underline{V(M)=gM^4}&:& \nonumber\\ n^2G_\phi(z,w)&=& n^2G(z,w) - g^2 \left(8z^3-\frac{8z^4-4a^2z^2-a^4}{\sqrt{z^2-a^2}}\right) \left(8w^3-\frac{8w^4-4a^2w^2-a^4}{\sqrt{w^2-a^2}}\right) \ , \nonumber\\ \label{G2non} \eeqn where $n^2G(z,w)$ is given in eq. (\ref{Guniv}) and $A^2$ in eq.(\ref{2.3}). These examples clearly demonstrate the non-universality of $G_\phi(z,w)$, which cannot be repaired by a suitable parameter redefinition. Let us finally mention that we have verified eq. (\ref{GG2}) for the quadratic potential following an entirely different approach. As it has been already emphasized in \cite{ACMV} the fixed trace ensemble $\phi=\delta$ can be obtained from the ``trace squared ensemble'' \beqn {\cal P}_l(M)&\equiv& \frac{1}{{\cal Z}_l}\,\exp \left[-l\left( -2nA^2\mbox{Tr} V(M) + (\mbox{Tr} V(M))^2\right)\right] \ \ , \ \nonumber\\ {\cal Z}_l &\equiv& \int {\cal D}M \, \exp \left[-l\left( -2nA^2\mbox{Tr} V(M) + (\mbox{Tr} V(M))^2\right)\right] \ \ , \label{Psq} \eeqn when taking the limit $l\to\infty$. The trace square terms add so-called ``touching'' interactions to the triangulated surface \cite{das}. This representation of the fixed trace ensemble does not only provide us with a different technical tool to check eq. (\ref{GG2}) for $p=1$, which we do not display here. It also gives us a diagrammatical interpretation in terms of Feynman graphs, which explains the existence of the $1/n^2$-expansion in eq. (\ref{ddef}) for the Gaussian fixed trace ensemble as a topological expansion. It seems remarkable that the second term in eq. (\ref{GG2}), in the case of Gaussian resolvent $G_\phi(z) \to G(z) =\frac{2}{a^2} (z- \sqrt{z^2-a^2})$ , has the same form of the analogous term which appears in the connected correlator for Wigner ensembles \cite{pastur}, \cite{danna}, written in different but equivalent forms since \beqn \partial_z\left( z G(z) \right)= \frac{-2 a^2 [G(z)]^3}{4-a^2 [G(z)]^2} =\frac{a^2}{4} \partial_z[G(z)]^2 \ \ . \nonumber \eeqn \setcounter{equation}{0} \section{Equivalence of all higher-point resolvents of the RTEs} The two-point resolvent of the fixed and bounded trace ensemble has turned out to be identical in the large-$n$ limit although non-universal. It is therefore natural to ask whether this equivalence holds also for all higher $k$-point resolvents. In the following we will show that this is indeed the case. Let us define the two generating functionals \beqn Z_\phi[J] &\equiv& \sum_{k=0}^\infty \frac{1}{k!}\int dz_1\ldots dz_k \left\langle\frac{1}{n}\mbox{Tr}\frac{1}{z_1-M}\ldots \frac{1}{n}\mbox{Tr}\frac{1}{z_k-M}\right\rangle_\phi J(z_1)\ldots J(z_k) \ \ , \label{ZJ}\\ W_\phi[J] &\equiv& \sum_{k=0}^\infty \frac{1}{k!}\int dz_1\ldots dz_k \ G_\phi(z_1,\ldots,z_k)\ J(z_1)\ldots J(z_k) \ \ , \label{WJ} \eeqn where \begin{equation} G_\phi(z_1,\ldots,z_k) \equiv \left\langle\frac{1}{n}\mbox{Tr}\frac{1}{z_1-M}\ldots \frac{1}{n}\mbox{Tr}\frac{1}{z_k-M}\right\rangle_\phi - \sum_{\sigma\in P} G_\phi(z_{\sigma(1)},..,z_{\sigma(l_1)}) \ldots G_\phi(z_{\sigma(l_{k-1}+1)},..,z_{\sigma(k)}) \label{Gk} \end{equation} is the $k$-point resolvent. The sum runs over all different partitions $P$ of the $k$ arguments and thus over all different combinations of 1- to $(k-1)$- point resolvents with in total $k$ indices. In the canonical ensemble this subtraction corresponds to taking only connected graphs into account. For this reason the $k$-point resolvent is there of the order $1/n^{2k-2}$. From field theory we know that the following relation between the generating functionals holds: \begin{equation} Z_\phi[J] \ =\ e^{W_\phi[J]} \ \ . \label{ZW} \end{equation} The correlators can be obtained in the usual way \begin{equation} \frac{\delta^k}{\delta J^k} \left\{ \begin{array}{l}Z_\phi[J]\\W_\phi[J]\end{array} \right|_{J=0} \ =\ \left\{ \begin{array}{l}\langle\frac{1}{n}\mbox{Tr}\frac{1}{z_1-M}\ldots \frac{1}{n}\mbox{Tr}\frac{1}{z_k-M}\rangle_\phi \\ G_\phi(z_1,\ldots,z_k) \end{array}\right. \ . \end{equation} In \cite{ACMV} it has been shown that the ensemble averages of the fixed and bounded trace ensemble can be related: \begin{equation} \left\langle\ {\cal O}(M)\ \right\rangle_\delta\ =\ (1+c_n\partial_{A^2})\left\langle\ {\cal O}(M)\ \right\rangle_\theta \ \ , \label{rel} \end{equation} where we have \begin{equation} c_n \ =\ 2pA^2\frac{1}{n^2} \ \ \mbox{for} \ \ V(M)=M^{2p} \ . \label{cn} \end{equation} Consequently the same relation holds for the generating functionals $Z_\phi[J]$ following their definition (\ref{ZJ}) \begin{equation} Z_\delta[J] \ =\ (1+c_n\partial_{A^2})Z_\theta[J] \ \ . \label{ZZ} \end{equation} Using the relation eq. (\ref{ZW}) we can translate this to the generating functional for the resolvent operators \begin{equation} e^{W_\delta[J]} \ =\ \left( 1+ c_n (\partial_{A^2}W_\theta[J])\right) e^{W_\theta[J]} \ \ , \end{equation} or equivalently \begin{equation} W_\delta[J]\ =\ W_\theta[J]+\sum_{l=1}^\infty (-)^{l+1}\frac{1}{l} (c_n\partial_{A^2}W_\theta[J])^l \ \ , \end{equation} where we have expanded the logarithm. Taking the functional derivative $\delta^k/\delta J^k$ and setting $J=0$ will truncate the infinite sum for the following reason. From the definition we have $W_\phi[J\!=\!0]=1$ and thus $\partial_{A^2}W_\phi[J\!=\!0]=0$. For this reason only terms will persist where at least one functional derivative $\delta/\delta J$ acts on $\partial_{A^2}W_\phi[J]$. We finally obtain \beqn G_\delta(z_1,\ldots,z_k) &=& (1+c_n\partial_{A^2})G_\theta(z_1,\ldots,z_k) \nonumber\\ &+& \sum_{\sigma\in P;\ l=2,..,k} (-)^{l+1}\frac{1}{l} \left(c_n\partial_{A^2}G_\theta(z_{\sigma(1)},..,z_{\sigma(l_1)})\right)\ldots \left(c_n\partial_{A^2}G_\theta(z_{\sigma(l_{k-1}+1)},..,z_{\sigma(k)})\right) .\nonumber\\ \label{GGk} \eeqn Here the sum runs again over all partitions $P$ of the $k$ arguments and $l$ counts the number of blocks or resolvents into which the arguments are divided. To prove the desired equivalence between the two $k$-point resolvents we need to know the order in $1/n^2$ of all terms on the r.h.s. Following the diagrammatic approach mentioned at the end of the last section, where the fixed trace ensemble is represented by the trace squared one, we obtain the same counting of powers as in the canonical ensemble already mentioned \begin{equation} G_\delta(z_1,\ldots,z_k)\ =\ O\left(\frac{1}{n^{2k-2}}\right) \ , \label{GOn} \end{equation} at least in the Gaussian case. In the following we will assume that same holds for the monomial potentials. We have checked this explicitly for the 2- and 3-point resolvent using the definition (\ref{Gk}) and the relation (\ref{dc}). It now follows easily by induction that eq. (\ref{GOn}) also holds for the bounded trace ensemble and that we have \begin{equation} n^{2k-2}G_\theta(z_1,\ldots,z_k)\ \stackrel{n\to\infty}{\longrightarrow}\ n^{2k-2}G_\delta(z_z,\ldots,z_k) \ , \label{equi} \end{equation} which generalizes eq. (\ref{GG2a}) for the two-point resolvents. Namely in eq. (\ref{GGk}) on the r.h.s. the second term in the first line is obviously subleading, due to $c_n\sim1/n^2$. Using induction in the sum each term is of the order $O(n^{-(2l_1+2(l_2-l_1)+..+2(k-l_{k-1})})$ = $O(n^{-2k})$ which is also subleading. In the above derivation no explicit use has been made of the $\delta$- or $\theta$-measure apart from the fact that $\theta^{\prime}(x)=\delta(x)$. Instead of this we could have used for example $\phi(x)=x\theta(x)$ and $\phi(x)=\theta(x)$ because of $(x\theta(x))^{\prime}=\theta(x)$. More generally we can extend the proof of relation (\ref{equi}) to an infinite class of RTEs with \begin{equation} \phi(x)=\left\{ \ \delta(x),\ \theta(x),\ \left(\frac{1}{j!}x^j\theta(x)\right)_{j\in \mbox{N}_+}\right\}\ \ , \label{newphi} \end{equation} showing that all their $k$-point resolvents are equivalent at large-$n$. We only need to show the starting point for $k=2$ since we have used induction. This can be shown as follows. When we calculate the matrix integral $I^{\{k\}}_\phi(n,A)$ in eq. (\ref{Idef}) we allow the parameter $s$ in the representation eq. (\ref{phi}) to take all non-negative integer values, which is then a representation for all the measures introduced in eq. (\ref{newphi}). The same derivation goes through up to the result for the two-point resolvent eq. (\ref{GG2}) as we have kept $s$ general and explicit everywhere. Let us conclude this section with a final remark. In ref. \cite{ACKM} a topological expansion was introduced and calculated for each resolvent \begin{equation} G(z_1,\ldots,z_k) \ =\ \sum_{h=0}^\infty \frac{1}{n^{2h}} G_h(z_1,\ldots,z_k) \ . \end{equation} If we introduce the same expansion here for the $G_\phi(z_1,\ldots,z_k)$, a short look at relation (\ref{GGk}) tells us that already for $h=1$ (``genus one'') the equivalence eq. (\ref{equi}) breaks down: \begin{equation} n^{2k-2+2h}G_{h\geq1,\theta}(z_1,\ldots,z_k) \ \neq\ n^{2k-2+2h}G_{h\geq1,\delta}(z_1,\ldots,z_k) \ . \end{equation} In this sense we have shown that only the ``planar'' $(h=0)$ $k$-point resolvents of the fixed and extended bounded trace ensembles agree. \setcounter{equation}{0} \section{Conclusions} We have proved the non-universality of the two-point resolvent $G_\phi(z,w)$ of the generalized fixed and bounded trace ensembles by comparing it to the universal two-point resolvent of the canonical ensemble. Apart from the general results for $G_\phi(z,w)$ for all monomial potentials $V(M)=M^{2p}$ we have explicitly displayed its non-universal parts in two examples, the quadratic and the pure quartic potential. Furthermore, we have extended the equivalence of the generalized fixed and generalized bounded ensemble in the large-$n$ limit from all finite moments \cite{ACMV} to all $k$-point resolvents, which probe higher orders in $1/n^2$. While we have shown that global universality fails for the generalized RTEs, the issue of universality of correlators at short distance, possibly matching with the canonical ensemble, is still open. We plan to come back to this interesting question in the future.
\section{Introduction} We have previously reported the detection of a Sunyaev-Zel'dovich (SZ) decrement \cite{sandz} towards the $z = 0.217$ cluster Abell~773 using the Ryle Telescope~(RT)~\cite{grainge93}. (The SZ effect in this cluster has also been mapped by the millimeter array of the Owens Valley Radio Observatory~\cite{carl96}.) The RT observations of Abell~773 form part of a continuing programme to observe an X-ray luminosity-limited sample of rich, intermediate-redshift clusters in order to measure $ H_0$ by combining SZ and X-ray observations~\cite{jones01}. Such programmes~(e.g. Reese et al., 2002; Mason, Myers and Readhead, 2001; see also Birkinshaw, 1999 for a review) are direct measurements of $H_0$ free from distance-ladder arguments. In Grainge et al.~1993 we did not calculate an estimate of $H_0$ because no suitable X-ray image of A773 and no estimate of its gas temperature existed. A {\sl ROSAT} HRI image and {\sl ASCA} spectroscopic data have since become available, and we have also made additional RT observations. These now enable us to make an estimate of the Hubble constant from this cluster, which, when combined with other clusters from the sample, will give an estimate of $H_0$ unbiased by the individual shapes and orientations of the clusters. \section{Ryle Telescope observations and source subtraction} The RT \cite{jones91} is an east--west synthesis telescope of 13-m antennas with a bandwidth of 350~MHz and an average system temperature for these observations of 65~K at an observing frequency of 15.4~GHz. We used five antennas in a compact configuration, giving two baselines of 18~m, three of 36~m, and five more out to 108~m. The short baselines alone are sensitive to the SZ signal; the longer ones are used to recognize and subtract the radio sources in the field that would otherwise mask the SZ decrement. We have made a total of 30 12-h observations of A773, each with the pointing centre $ \rm {RA} \ 09^h \ 17^m \ 51^s.91 ,\ \rm {Dec.} $ $ +51^{\circ} \ 43' \ 32'' $~(J2000). Phase calibration using 0859+470 and flux calibration using 3C~48 and 3C~286 were carried at as described in Grainge et al \shortcite{grainge93}. Similarly, we used the {\sc Postmortem} package \cite{titterington91} to flag the data for interference and antenna pointing errors, and to weight them in accord with the continuously monitored system temperature of each antenna. As a standard check, we used the {\sc Aips} package to make a map of each 12-h run and then combined the data. We removed radio sources from the data by a simultaneous maximum-likelihood fit to several point sources and the SZ effect using a technique described by Grainger et al \shortcite{A611}. We use a model for the SZ signal as a function of baseline that is based on the $\beta$-model fit to the X-ray image described below (Section \ref{xrays}). We simultaneously fit flux densities for trial sources whose initial positions are determinied both from a map made from just the long-baseline data ($>2 \, \rm k \lambda$), and from a VLA 1.4-GHz image of the cluster field (Figure \ref{vla}). This allows us to fit the optimum flux densities of sources whose existence we know of from the VLA image but which would not give a significant detection from the RT data alone. The postitions and fitted flux densities are given in Table \ref{table:sources}. The image made from the long ($>2\rm k \lambda$) source-subtracted baselines is consistent with noise (Figure \ref{source-sub}). To image the decrement, we removed the sources in Table~\ref{table:sources} from all the visibilities and made a short-baseline map from baselines shorter than 1 k$\lambda$, and CLEANed this. The resulting image is shown in Figure~\ref{sz+xray}. The decrement is of $-527 \, \mu \rm \, Jy\, beam^{-1}$ with a noise (1-$\sigma$) of $60 \, \mu \rm Jy \, beam^{-1}$; the beam is $152 \times 119$ arcsec FWHM. Also shown is the X-ray image of the cluster; it can be seen that the alignment with the X-ray image is very good. The extension of the SZ image to the north-east is of marginal significance. The magnitude of the decrement is consistent with that of $-590 \pm 116 \ \mu$Jy, in the same beam, reported in Grainge et al \shortcite{grainge93}. An alternative way of looking at the data is shown in Figure~\ref{visplot}, which shows the real part of the source-subtracted visibilities binned radially, along with the best-fitting model based on the X-ray data. These data have the advantage, unlike the image pixels, of having independent gaussian noise on each point; it is these that are used in the fitting for $H_0$. \section{X-ray observations and fitting}\label{xrays} We measure the gas temperature from {\sl ASCA} observations on 1994 April 29 of 46240 s (GIS) and 39904 s (SIS), using standard {\sc XSPEC} tools. Times of high background flux were excluded and both GIS and SIS data were used. We took the Galactic absorbing column density predicted by Dickey and Lockman \shortcite{dickey90} in the direction of A773 of $1.3 \times 10^{24}$ H atoms m$^{-2}$. Using a Raymond-Smith model, we find a temperature of $8.7 \pm 0.7$ keV ($90\%$-confidence error bounds) and a metallicity of 0.25 solar. The 2--10 keV flux from A773 is $6.7{\pm 1.0}$ x $10^{-13}$ W m$^{-2}$. Our temperature estimate is consistent with that of Allen and Fabian \shortcite{allen98} who find a temperature of $9.29^{+0.69}_{-0.60}$ keV ($90\%$-confidence error bounds). For the X-ray surface-brightness fitting we used a {\sl ROSAT} HRI image of A773 with an effective exposure of 16518~s obtained on 13--15 April 1994 and analysed using standard {\sc ASTERIX} routines. We calculate the {\sl ROSAT} HRI count rate, given our estimates of metallicity and Galactic column and with the K-correction appropriate to the redshift of A773, to be $1.53 (\pm 0.08) \times 10^{-69}$ counts s$^{-1}$ from a $1 \, \rm m^3$ cube of gas of electron density $1 \, \rm m^{-3}$ at the temperature of A773 and at a luminosity distance of 1 Mpc. We then fitted an ellipsoidal King profile to the X-ray image. Since the high spatial resolution of the HRI leads to a low count rate per pixel, we use Poisson rather than Gaussian statistics to fit for the measured count in each pixel. For $c_i$ counts measured at position $x_i$, and for a mean number $f(x_i|a)$ of counts predicted by the model given parameters $a$ (such as core radius), the probability of obtaining $c_i$ counts is \[ P(c_i|a)=\frac{\left(f(x_i|a)\right)^{c_i}}{c_i!}e^{-f(x_i|a)} \ , \] and the most likely value of $a$ can be obtained in a computationally efficient way by maximizing \[ \ln P(c|a)=\sum_i (c_i\ln f(x_i|a)-\ln c_i!-f(x_i|a)). \] We fitted an ellipsoidal King profile to the HRI data with $\theta_1$ and $\theta_2$ as the perpendicular angular sizes in the plane of the image, assuming that the length along the line of sight is the geometric mean of the other two. We find $\theta_1 = 60''$ and $\theta_2 = 44''$, with the major axis at position angle $16^{\circ}$, $\beta =0.64$, and central electron density $n_0 = 6.80 \times 10^3 \, h_{50}^{-1/2} \rm m^{-3}$ where $H_0 = 50\,h_{50}$ km s$^{-1}$ Mpc$^{-1}$. Fig \ref{residual} shows the HRI image, the model, and the residual image with the best model subtracted. To assess the goodness of fit, we made 50 realisations of the image with the appropriate Poisson noise added, and calculated the mean and standard deviation of their Poisson likelihoods. The likelihood of the observed HRI image is 0.32 standard deviations from the mean; we therefore conclude that the fit is good and the cluster is well represented by a $\beta$ model. There is a strong degeneracy in the fit between $\beta$ and $\theta_{1,2}$; however this has little effect on the comparison with the SZ data and the derived value of $H_0$. Figure \ref{margplot} shows the likelihood contours for the fit in the $\beta$--$\theta_1$ plane, marginalised over $n_0$ and using the best-fit value of the axial ratio (which is very well constrained). Overlaid are the contours of predicted mean observed SZ flux density on the shortest RT baseline. It can be seen that despite the degeneracy between $\beta$ and $\theta_1$, the range of SZ flux densities corresponding to the 1-$\sigma$ limits of the model fit is only $\pm 3 \%$. Since the SZ flux density varies as $H_0^{-2}$, this corresponds to a $6 \%$ error in $H_0$ due to the model fitting. This lack of sensitivity to the $\beta$--$\theta$ degeneracy is characteristic of observations that are sensitive to spatial frequencies around the cluster core size (see eg Reese et al \shortcite{reese}) and contrasts with the sensitivity to the model fitting of measurements that measure only lower spatial frequencies (eg Birkinshaw \& Hughes\shortcite{birk94}). \section{$H_0$ estimation} To measure $H_0$, we compared the real SZ data with a simulation of the SZ effect from the X-ray gas model. We use the expression of Challinor \& Lasenby \shortcite{challinor98} to provide a relativistic correction to the standard non-relativistic SZ expression; in the case of A773, the effect is to increase our estimate of the $y$-parameter by $2.4\%$. We then simulated RT observations of the SZ effect due to the model gas distribution and to compared these with the real source-subtracted RT visibilities on the same baselines, and adjusted $H_0$ to get the best fit. Using our temperature of $8.7{\pm 0.7}$ keV we find $H_0 = 77^{+13}_{-11}$ km s$^{-1}$ Mpc$^{-1}$, assuming an Einstein-de-Sitter universe. The $1$-${\sigma}$ error quoted is that due solely to noise in the SZ data. For the best fit $\beta, \ \theta_{1,2}$ model, the corresponding central density $n_0$ is $8.44 \times 10^3$ m$^{-3}$ and the central decrement $737 \pm 85 {\mu}$K. Grainge et al \shortcite{grainge02} consider at some length the contributions to error in the $H_0$ determination from A1413. The situation in A773 is very similar. The dominant contributions to the error in $H_0$ in A773 are $\pm 16 \%$ from noise in the SZ measurement, $\pm 12 \%$ from our estimation of the gas temperature and a likely error of $\pm 14 \%$ from the uncertain line-of-sight depth. This is obtained by considering the range of axial ratios of simulated clusters that is needed to reproduce the projected axial ratio distribution observed in clusters with redshift similat to that of A773 \cite{will_thesis}. Clearly this estimate is rather uncertain for a single object, but can be significantly reduced by averaging a sample of clusters with random orientations. Table \ref{errors} shows the complete error budget, and the final 1-$\sigma$ error limits of $H_0 = 77^{+19}_{-15}\, \rm km \, s^{-1}\, Mpc^{-1}$ if $(\Omega_{\rm m}, \Omega_{\Lambda}) = (1.0, 0.0)$ and $H_0 = 85^{+20}_{-17}\rm \, km \, s^{-1}\, Mpc^{-1}$ if $(\Omega_{\rm m}, \Omega_{\Lambda}) = (0.3, 0.7)$. \section{Conclusions} Using {\sl ASCA}, {\sl ROSAT} HRI, and RT observations of A773, we find: \begin{enumerate} \item there are eight radio sources detectable in the field of the cluster that we have removed from the data, which would otherwise contaminate the measurement of the SZ effect; \item the correlated fitting errors on the shape parameters $\beta$ and $\theta$ have negligable effect on the derived value of $H_0$, a feature characteristic of observations on the scale of the cluster core size; \item the estimated value of $H_0$ is $77^{+19}_{-15}\, \rm km \, s^{-1}\, Mpc^{-1}$ if $(\Omega_{\rm m}, \Omega_{\Lambda}) = (1.0, 0.0)$ or $85^{+20}_{-17}\, \rm km \, s^{-1}\, Mpc^{-1}$ if $(\Omega_{\rm m}, \Omega_{\Lambda}) = (0.3, 0.7)$ , where the 1-$\sigma$ error bars include estimates from the main sources of error---noise in the SZ data, X-ray temperature uncertainty, and uncertain line-of-sight depth. \end{enumerate} \subsection*{ACKNOWLEDGMENTS} We thank the staff of the Cavendish Astrophysics group who maintain and operate the Ryle Telescope, which is funded by PPARC. AE acknowledges support from the Royal Society; WFG acknowledges the support of a PPARC studentship; RK acknowledges support from an EU Marie Curie Fellowship.
\section{Introduction} The concept of Black Holes (BHs) is one of the most important plinths of modern physics and astrophysics. As is well known, the basic concept of BHs actually arose more than two hundred years ago in the cradle of Newtonian gravitation\cite{1}. In General Theory of Relativity (GTR), the gravitational mass is less than the baryonic mass ($M\le M_0$). Further, as the body contracts and emits radiation $M$ keeps on decreasing progressively alongwith $r$. Thus, given an initial gravitational mass $M_i$, one can not predict with certainty the value of $M_f$ when we would have $2M_f/r =1$ ($G=c=1$). Neither are the values of $M_i$, $M_f$ and $M_0$ related by any combination of fundamental constants though, it is generally assumed that $M_i\approx M_f$. Ideally, one should solve the Einstein equations analytically to fix the value of $M_f$ for a given initial values of $M_i$ and $M_0$ for a realistic equation of state (EOS) and energy transport properties. However even when one does away with the EOS by assuming the matter to behave like a dust, $p\equiv 0$, one does not obtain any unique solution if the dust is inhomogeneous. Depending on the various initial conditions and assumptions (like self-similarity) employed one may end up finding either a BH or a ``naked singularity''\cite{2}. By further assuming the dust to be {\em homogeneous} Oppenheimer and Snyder (OS)\cite{3} found {\em asymptotic} solution of the problem by approximating Eq.(36) of their paper. The region exterior to the event horizon ($r >r_g=2M$) can be described by the Schwarzschild coordinates $r$ and $t$\cite{4,5}: \begin{equation} ds^2 = g_{tt} dt^2 +g_{rr} dr^2 + g_{\theta \theta} d\theta^2 + g_{\phi \phi} d\phi^2 \end{equation} where $g_{tt} =(1- 2M/r)$, $ g_{rr} =-(1-2M/r)^{-1}$, $ g_{\theta \theta}=-r^2$, and $g_{\phi \phi} = -r^2 \sin^2 \theta$. Here, we are working with a spacetime signature of +1, -1, -1, -1 and $r$ has a distinct physical significance as the {\em invariant circumference radius}. For $r>r_g=2M$, the worldline of a freeling falling radial material particle is indeed timelike $ds^2 >0$ and the metric coefficients have the right signature, $g_{tt} >0$, $g_{rr} <0$, $g_{\theta \theta} <0$ and $g_{\phi \phi} <0$. But at $r=2M$, $g_{rr}$ blows up and as $r <2M$, the $g_{tt}$ and $g_{rr}$ suddenly exchange their signatures {\em though the signatures of $g_{\theta \theta}$ and $g_{\phi \phi}$ remain unchanged}. This is interpreted by saying that, inside the event horizon, $r$ becomes ``time like'' and $t$ becomes ``spacelike''\cite{4,5}. However, we see that actually $r$ continues to retain, atleast partially, its spacelike character by continuing to be ``invariant circumference radius''. Also, note that, if physically measurable quantities like the Rimennian curvature components behaved like $\sim M/r^3$ outside the EH, they continue to behave in a similar manner, {\em and not like $\sim M/t^3$} inside the EH. And it should be borne in mind here that by a fresh relabelling or by any other means, the curvature components can not be made to assume the form $\sim M/t^3$. One particular reason for this is that, we would see later that, inside the EH, we have $t=\infty$ while, of course, the value of $r$ remains finite. Thus it may not actually be justified to conclude that $r$ becomes the ``timelike coordinate'' inside the EH even though $g_{rr}$ changes its sign. So far, it has not been possible to resolve this enigma of the {\em duality} in the behaviour of $r$ for $r <2M$, and the present paper intends to attend to this problem. Since $ds$ is the proper time, we may also write \begin{equation} ds^2=dt^2 \left( 1- {2M\over r} \right) \end{equation} Therefore, the radial geodesic of a {\em material particle} in the Schwarzschild metric becomes, {\em unphysically} {\em null} ($ds^2=0$) and then {\em spacelike} ($ds^2 <0$) as one moves inside the event horizon (EH). In contrast, any physically meaningful coordinate system must be free of such anomalies. Although $g_{rr}$ blows up at $r=2M$, as mentioned before, the curvature components of the Rimennian tensor behave perfectly normally at $r=r_g$, $R^{ij}_{kl} \sim M/r^3$. Further, the determinant of the metric coefficients continues to be negative and finite $g = r^4 \sin^2 \theta~ g_{rr}~ g_{tt} = -r^4 \sin^2 \theta \le 0 $. Such realizations gave rise to the idea that the Schwarzschild coordinate system suffers from a ``coordinate singularity'' at the event horizon and must be replaced by some other well behaved coordinate system. It is known that a comoving coordinate system is naturally singularity free and Lemaitre suggested that the region inside $r\le r_g$ may be represented by such a coordinate system\cite{6} whereas the exterior region is still described by the old Schwarzschild coordinates. It is only in 1960 that Kruskal and Szekeres\cite{4,5,7} discovered a one-piece coordinate system which can describe both the interior and exterior regions of a BH. They achieved this by means of the following coordinate transformation for the exterior region (Sector I): \begin{equation} u=f_1(r) \cosh {t\over 4M}; \qquad v=f_1(r) \sinh {t\over 4M}; r\ge 2M \end{equation} where \begin{equation} f_1(r) = \left({r\over 2M} -1\right)^{1/2} e^{r/4M} \end{equation} It would be profitable to note that \begin{equation} {df_1\over dr} = {r\over 8M^2} \left({r\over 2M} -1\right)^{-1/2} e^{r/4M} \end{equation} And for the region interior to the horizon (Sector II), we have \begin{equation} u=f_2(r) \sinh {t\over 4M};\qquad v=f_2(r) \cosh {t\over 4M}; r\le 2M \end{equation} where \begin{equation} f_2(r) = \left(1- {r\over 2M}\right)^{1/2} e^{r/4M} \end{equation} and \begin{equation} {df_2\over dr} = {-r\over 8M^2} \left(1- {r\over 2M}\right)^{-1/2} e^{r/4M} \end{equation} Given our adopted signature of spacetime ($-2$), in terms of $u$ and $v$, the metric for the entire spacetime is \begin{equation} ds^2 = {32 M^3\over r} e^{-r/2M} (dv^2 -d u^2) - r^2 (d\theta^2 +d\phi^2 \sin^2 \theta) \end{equation} The metric coefficients are regular everywhere except at the intrinsic singularity $r=0$, as is expected. Note that, the angular part of the metric remains unchanged by such transformations and $r(u,v)$ continues to signal its intrinsic spacelike nature. In either region we have \begin{equation} u^2-v^2= \left({r\over 2M} -1\right) e^{r/2M} \end{equation} so that \begin{equation} u^2-v^2 > 1; \qquad u/v >\pm 1; \qquad r >2M, \end{equation} \begin{equation} u^2-v^2 \rightarrow 0; \qquad u =\pm v; \qquad r= 2M \end{equation} and \begin{equation} u^2-v^2 <0; \qquad u/v < \pm 1; \qquad r <2M \end{equation} So, each of these above three inequalities, and, in particular, the $r=0$ point corresponds to not one but two conditions! \begin{equation} v= \pm (1+u^2)^{1/2} \end{equation} Here, one point needs to be hardly overemphasized; astronomical observations and experiments actually conform to the idea that atleast far from massive bodies or probable BHs, the spacetime is well described by the $r,t$ coordinate system. In fact, although in the (normal) physical spacetime, in a spherically symmetric spatial geometry (as defined by the implications of $r$ as an ``invariant circumference radius''), the physical singularity corresponds to a mathematical point, in the Kruskal world view, this central singularity corresponds to a pair of hyperbolas in the ($u-v$) plane. While the ``+ve'' sign of equation corresponds to the central BH singularity, the ``-ve'' sign corresponds to the singularity inside a so-called {\em White Hole} which may spew out mass-energy spontaneously in ``our universe''\cite{4,5}. The white hole singularity belongs to ``other universe'' whose presence is suggested by the fact that the Kruskal metric remains unaffected by the following additional transformations: \begin{equation} u=-f_1(r) \cosh {t\over 4M};\qquad v=-f_1(r) \sinh {t\over 4M}; r\ge 2M \end{equation} defining Sector (III) and \begin{equation} u=-f_2(r) \sinh {t\over 4M}; \qquad v=-f_2(r) \cosh {t\over 4M}; r\le 2M \end{equation} defining Sector (IV). Thus not only does the region interior to the EH correspond to two different universes, (Sector II and IV) but the structure of the physical spacetime outside the EH, too, effectively corresponds to two universes (Sector I and III). If there exists $N$ number of BHs, the (normal) physical spacetime may be much more complex. The aim of this paper is to {\em explicitly verify} whether the (radial) geodesics of material particles are indeed {\em timelike} at the EH which they must be if this idea of a finite mass Schwarzschild BH is physically correct. First we focus attention on the region $r\ge 2M$ and differentiate Eq.(3) to see \begin{equation} {du\over dr} = {\partial u\over \partial r} + {\partial u\over \partial t} {dt\over dr}= {df\over dr} \cosh {t\over 4M} + {f\over 4M} \sinh {t\over 4M}{dt\over dr} \end{equation} Now by using Eq. (4-6) in the above equation, we find that \begin{equation} {du\over dr} = {ru\over 8M^2} (r/2M -1)^{-1} + {v\over 4M} {dt\over dr}; \qquad r\ge 2M \end{equation} and \begin{equation} {dv\over dr} = {rv\over 8M^2} (r/2M -1)^{-1} + {u\over 4M} {dt\over dr}; \qquad r\ge 2M \end{equation} By dividing equation (18) by (19), we obtain \begin{equation} {du\over dv} = {{ru\over 2M} + v {dt\over dr} (r/2M-1) \over {rv\over 2M} + u {dt\over dr} (r/2M -1)} \end{equation} Similarly, starting from Eq. (6), we end up obtaining a form of $du/dv$ for the region $r <2M$ which is exactly similar to the foregoing equation. Now, by using Eq.(12) ($u=\pm v$) in Eq. (20), we promptly find that \begin{equation} {du\over dv}\rightarrow {{\pm r\over 2M} + {dt\over dr} (r/2M-1) \over { r\over 2M} \pm {dt\over dr} (r/2M -1)} \rightarrow \pm 1;\qquad r\rightarrow 2M \end{equation} Thus, we are able to find the precise value of $du/dv$ at the EH in a most general manner {\em irrespective of the precise relationship} between $t$ and $r$. Armed with this value of $du/dv$, we are in a position now to complete our task by rewriting the {\em radial part} of the Kruskal metric ($d\theta =d\phi =0$) as \begin{equation} ds^2 = {32 M^3\over r} e^{-r/2M} dv^2 \left[1- \left({du\over dv}\right)^2\right] \end{equation} Or, \begin{equation} ds^2 = 16 M^2 e^{-1} dv^2 (1-1)=0; \qquad r=2M \end{equation} We have found that {\em for the Lemaitre coordinate too}, $ds^2 =0$ at $r=2M$. This implies that although the metric coefficients can be made to appear regular, the radial geodesic of a {\em material particle becomes null} at the event horizon of a finite mass BH in contravention of the basic premises of GTR! And since, now, we can not blame the coordinate system to be faulty for this occurrence, the only way we can explain this result is that {\em the Event Horizon itself corresponds to the physical singularity} or, in other words, the mass of the Schwarzschild BHS $M\equiv 0$. And then, the entire conundrum of ``Schwarzschild singularity'', ``swapping of spatial and temporal characters by $r$ and $t$ inside the event horizon ({\em when the angular part of all metrics suggest that $r$ has a spacelike character even within the horizon}), ``White Holes'' and ``Other Universes'' get resolved. Here we recall the conjecture of Rosen\cite{8} ``so that in this region $r$ is timelike and $t$ is spacelike. However, this is an impossible situation, for we have seen that $r$ defined in terms of the circumference of a circle so that $r$ is spacelike, and we are therefore faced with a contradiction. We must conclude that the portion of space corresponding to $r< 2M$ is non-physical. This is a situation which a coordinate transformation even one which removes a singularity can not change. What it means is that the surface $r=2M$ represents the boundary of physical space and should be regarded as an impenetrable barrier for particles and light rays.'' This idea of Rosen is also in accordance with the idea of Einstein that the Schwarzschild type singularity is unphysical and can not occur for realistic cases\cite{9}. And this paper indeed shows that {\em in order that the radial worldlines of free falling material particles do not become null at a mere coordinate singularity}, Nature (GTR) refuses to have any spacetime within the EH. Although, having made our basic point, {\em we could have ended this paper at this point}, for the sake of further insight, we shall study the behaviour of $ds^2$ for the entire spacetime by, again assuming, for a moment, the existence of a finite mass BH. It can be found that in the region $r>2M$, one would indeed have $ds^2 >0$ for $r >2M$. And to see the behaviour of $du/dv$ inside the EH, we recall the relationship between $t$ and $r$ (see pp. 824 of ref.[4] or pp. 343 of ref.[5]): \begin{equation} {t\over 2M} = \ln\mid{(r_\infty/2M-1)^{1/2} + \tan{(\eta/2)} \over (r_\infty /2M-1)^{1/2} - \tan{(\eta/2)}}\mid + 2M\left({r_\infty\over 2M}-1\right)^{1/2} \left[\eta + \left({r_\infty\over 4M}\right)(\eta +\sin \eta)\right] \end{equation} where the particle is released with zero velocity from $r=r_\infty$ at $t=0$ and the ``cyclic'' coordinate $\eta$ is defined by \begin{equation} r ={r_\infty\over 2} (1+\cos \eta) \end{equation} Since $\tan{(\eta/2)} = (r_\infty/r -1)$ we find from Eq. (24) that, as $r\rightarrow 2M$, the logarithmic term blows up and $t\rightarrow \infty$, which is a well known result. And since $t$ continues to increase as the particle enters the EH, we have the general result that $t =\infty$ for $r\le 2M$. In this limit, we have \begin{equation} \cosh{t\over 4M} \rightarrow \sinh{t\over 4M} \rightarrow {e^{t/4M}\over 2} =\infty \end{equation} Consequently, even though, $u^2 -v^2$ continues to be finite we obtain \begin{equation} {u\over v} =\pm 1; \qquad r\le 2M \end{equation} Hence we obtain a more general form of Eq. (21) \begin{equation} {du\over dv}\rightarrow \pm 1;\qquad r\le 2M \end{equation} irrespective of the {\em precise} form of $dt/dr$. Then from Eq. (22), we find that the metric would {\em continue to be null} for $r <2M$: \begin{equation} ds^2 =0; \qquad r \le 2M \end{equation} And this unphysical happening is of course avoided when we realize that $M=0$ and there is no additional spacetime between the EH and the central singularity. We may mention now that we have recently shown that the OS work too actually suggests that the mass of the resultant BH must be $M\equiv 0$\cite{10}. The basic reason for this assertion is extremely simple. The Eq.(36) of OS paper connects $t$ and $r$ through a relationship which, for large values of $t$ is \begin{equation} t \sim \ln{y^{1/2} +1\over y^{1/2} -1} \end{equation} where at the boundary of the fluid \begin{equation} y={r\over r_g} = {r\over 2M} \end{equation} Since the argument of a {\em logarithmic function can not be negative}, in order that $t$ is definable at all that we must have \begin{equation} y={r\over 2M} \ge 1; \qquad {2M\over r}\le 1 \end{equation} Thus atleast for the collapse of a homogeneous dust, ``trapped surfaces'' do not form and if the collapse continues to the point $r\rightarrow 0$ we must have $M_f\rightarrow 0$. This independent finding is in complete agreement with what we have shown in the present paper that Schwarzschild BHs must have $M=0$. Although, there is no modulus here in the argument of the logarithmic of Eq. (30) (unlike Eq. [24]), some readers may wish there were one. Even if one imagined the existence of such a modulus, one would run into contradiction in the following way. Of course we will have $t \rightarrow \infty$ as $r\rightarrow 2M$. But during the collapse if one would enter $r <2M$ (if $M >0$), $t$ {\bf would start decreasing}! However, unlike the case of Newtonian gravity, in GTR, $M=0$ state need not correspond to a configuration with zero baryonic mass. The $M=0$ state is simply one in which the negative gravitational energy exactly offsets the positive energy associated with $M_0$ and internal energy, and may indeed represent a physical singularity with infinite energy density and tidal acceleration. For instance, if the collapse process leads to the $y=1$ limit, then the curvature components $R^{ij}_{kl} \sim M/ r^3 \sim r^{-2} \rightarrow \infty$ as $r\rightarrow 0$. Note also that, the metric coefficients $g_{uu}$ and $g_{vv}$ for the zero-mass BH blow up in a similar fashion at the EH. It may be noted that the ``naked singularities'' too may be characterized by $M=0$\cite{11}. In the context of the dust collapse, we see that, for, $M=0$, the proper time for the formation of the BH would be infinite \begin{equation} \tau = \pi \left({r_\infty^3\over 8 M}\right)^{1/2} =\infty \end{equation} Further, we have shown elsewhere that the crucial condition (32), $y\ge 1$, is valid not only for the OS problem, but also for any generic spherical gravitational collapse\cite{12}. And similarly, $\tau\rightarrow \infty$ as $r\rightarrow 0$ not only for dust collapse, but also for the collapse of any physical fluid\cite{12}. Thus at any given finite proper time there would be no BH, and on the other hand there could be dynamically collapsing configurations with arbitrary high surface redshifts. In fact it can be found that the {\em proper length} of a radial geodesic becomes infinite too\cite{12}. And therefore, even if, such dynamically configurations with large surface red-shifts may be collapsing with relativistic velocities, the collapse process will never terminate in any finite amount of time. This happens because spacetime would get infinitely stretched by infinite curvature near $r=0$. This is a purely general relativistic effect, and is difficult to comprehend by ``common astronomical sense''. Observationally, such configurations may be identified as Black Holes. And if some of these configurations are collapsing with nearly free fall speed, accretion onto such configurations would emit little radiation if the accretion flow happens to be advection dominated. To conclude, irrespective of the observational consequences, we have {\em directly} shown that, if GTR is correct, Schwarzschild BHs must have $M\equiv 0$ in order that the radial geodesics of material particles remain timelike at a finite value of $r$.
\section{Introduction} Multiparticle production in high energy particle collisions is dominated by classical statistics. Bose-Einstein statistics of pions, nevertheless, proved to lead to quantum coherence effects which survive in the final multiparticle states \cite{GyuKauWil}. The corresponding mechanism can be handled within standard quantum statistics. Yet, some recent works \cite{Zaj,IMS} continued, mainly for historic reasons as Weiner \cite{Wei} points out, to use the "traditional" method of wave function with tedious explicit Bose-Einstein-symmetrization. We shall see how standard second quantization methods lead to the correct results in a shorter and clearer way. I shall discuss the dynamical conditions of Bose-Einsein condensation, and I outline a master equation suitable for Monte-Carlo simulations. For the concentrated study of effects of Bose-statistics, a simple scheme of 'independent' emission has been proposed \cite{IMS,BiaZal.PLB,BiaZal.APS,BiaZal.EJP}. Sec.2 recapitulates the features of these multiparticle states, sparing the burden of separate 'symmetrization'. Sec.3 introduces the generating functional of the measured counts \cite{GF} directly in second quantized formalism. Introducing chaotic classical currents, advocated e.g. by Ref.\cite{GyuKauWil} and used e.g. in \cite{AndPluWei,Ornetal}, I construct the simplest quantum dynamics reproducing the corresponding multiparticle quantum states (Sec.4). The measured multidetector counts turn out to be identical to the corresponding spectral intensities of the effective current (Sec.5). I show that the existence of Bose-Einstein condensate imposes explicit analytic constraints on the intensity and on the spectrum of the external effective current (Sec.6). Finally, I generalize the simple quantum dynamics and propose a quantum master equation suitable to efficient Monte Carlo simulation of the multiparticle density matrix itself (Sec.7). The Letter concludes with summary (Sec.8). \section{Independent multiboson states} When searching for a class of multiparticle density operators $\hat\rho$ representing independent bosons, consider first the Gibbs canonical state for non--interacting bosons at inverse temperature $\beta$. The bosons remain independent if, formally, we assign different instantaneous temperatures $\beta_{\bf k}}\def\kp{{\k^\prime}$ to each modes ${\bf k}}\def\kp{{\k^\prime}$, i.e. we assume $\hat\rho\sim\left(-\sum_{\bf k}}\def\kp{{\k^\prime} \beta_{\bf k}}\def\kp{{\k^\prime}{\hat a}_{\bf k}}\def\kp{{\k^\prime}^\dagger{\hat a}_{\bf k}}\def\kp{{\k^\prime}\right)$. Moreover, the bosons remain independent if we assign different temperatures to a generic (maybe non--stationary) set of orthogonal modes instead of the momentum eigenstates. Hence, we arrive at the following class of 'independent multiboson states' (IMS): \begin{equation} \hat\rho= det(1-e^{-\beta})\exp\left(-\sum_{{\bf k}}\def\kp{{\k^\prime},\kp} \beta_{\kp{\bf k}}\def\kp{{\k^\prime}}\akp^\dagger\a_\k}\def\akp{\a_\kp\right) \end{equation} where $\beta$ positive matrix. Let us define the 1-particle density matrix from the above state: \begin{equation} \rho_{\kp{\bf k}}\def\kp{{\k^\prime}}=\frac{\bra{\kp}\hat\rho\ket{{\bf k}}\def\kp{{\k^\prime}}}{\sum_{\bf k}}\def\kp{{\k^\prime}\bra{{\bf k}}\def\kp{{\k^\prime}}\hat\rho\ket{{\bf k}}\def\kp{{\k^\prime}}}~. \end{equation} Using Eq.~(1), we find the following matrix relation: \begin{equation} e^{-\beta}=\nu\rho~~~~(\nu=tre^{-\beta})~. \end{equation} The IMS (1) can be rewritten in terms of the 1-particle density matrix $\rho$ and the parameter $\nu$ (whose physical interpretation remains a bit involved): \begin{equation} \hat\rho= det(1-e^{-\nu\rho})\exp\left(\nu\sum_{{\bf k}}\def\kp{{\k^\prime},\kp} \rho_{\kp{\bf k}}\def\kp{{\k^\prime}}\akp^\dagger\otimes\a_\k}\def\akp{\a_\kp\right)\hat\rho_0~. \end{equation} This form might give an insight into the kinematics of the particle creation from the vacuum $\hat\rho_0$. We have to note that the IMS are non-stationary quantum states. Yet, the measured quantities $\hat n_{\bf k}}\def\kp{{\k^\prime}={\hat a}_{\bf k}}\def\kp{{\k^\prime}^\dagger{\hat a}_k$ are not sensitive to the time evolution of the IMS $\hat\rho$. This will be formulated in the next Section. \section{Generating functional} We introduce generating functionals for the multiparticle final state momentum distributions. A compact heuristic form of definition is the following: \begin{equation} G[u]=tr\left(\hat\rho\prod_{\bf k}}\def\kp{{\k^\prime}(u_{\bf k}}\def\kp{{\k^\prime})^{\hat n_{\bf k}}\def\kp{{\k^\prime}}\right)~, \end{equation} where $u_{\bf k}}\def\kp{{\k^\prime}$ are auxiliary variables and $\hat n_{\bf k}}\def\kp{{\k^\prime}=\a_\k}\def\akp{\a_\kp^\dagger\a_\k}\def\akp{\a_\kp$. If we introduce the diagonal matrix $u$ by $u_{\kp{\bf k}}\def\kp{{\k^\prime}}=\delta_{\kp{\bf k}}\def\kp{{\k^\prime}}u_{\bf k}}\def\kp{{\k^\prime}$ then, using the IMS density operator (4), the generating functional takes the following form: \begin{equation} G[u]=\frac{det(1-\nu\rho)}{det(1-\nu u\rho)}~. \end{equation} The logarithmic generating functional $g=\log G$ can be expressed through its Taylor--expansion in a transparent way \cite{BiaZal.PLB,BiaZal.EJP}: \begin{equation} g[u]=\sum_{r=1}^\infty \frac{\nu^r}{r}\left(tr(u\rho)^r-tr\rho^r\right)~. \end{equation} For $u_{\bf k}}\def\kp{{\k^\prime}\equiv u$, it yields the (logarithmic) multiplicity generating function \begin{equation} g(u)=\sum_{r=1}^\infty \frac{\nu^r}{r}tr\rho^r(u^r-1) \end{equation} whose Taylor--coefficients are the combinants (c.f. \cite{GyuKauHeg}). The derivatives of the generating functionals at $u=0$ yield the {\it exclusive} distribution/correlation functions. In experiments, we can easily measure the {\it inclusive} distributions instead, which are the derivatives at $u=1$ \cite{GF}. To make these derivations more convenient, let us substitute $\nu\rho$ in the generating functionals (6-8) by $\nu\rho=\alpha/(1+\alpha)$, where $\alpha$ will be the correlation matrix of 'currents' of Sec.4: \begin{equation} G[u]=\frac{1}{det[1-(u-1)\alpha)]}~, \end{equation} \begin{equation} g[u]=\sum_{r=1}^\infty \frac{\nu^r}{r} tr(u\alpha-\alpha)^r~, \end{equation} \begin{equation} g(u)=\sum_{r=1}^\infty \frac{\nu^r}{r}tr\rho^r(u-1)^r~. \end{equation} Comparing these expressions with the Eqs.~(6-8) we see that the inclusive distributions/correlations will depend on the current correlation matrix $\alpha$ exactly the same way as the exclusive distributions/correlations depend on ($\nu$-times) the 1-particle density matrix $\rho$. \section{Dynamics} Multiparticle production can not be derived from 'first principles'. I can certainly not overcome the well-known difficulties. Instead, I present the simplest quantum dynamics which produces exactly the class (1) of IMS. I postulate the following effective Hamiltonian: \begin{equation} \hat H=\delta(t)\sum_{\bf k}}\def\kp{{\k^\prime}\left(J_\k}\def\Jkp{J_\kp^\star\a_\k}\def\akp{\a_\kp+J_\k}\def\Jkp{J_\kp\a_\k}\def\akp{\a_\kp^\dagger\right)~, \end{equation} where $J_\k}\def\Jkp{J_\kp$ denote the Fourier--components of a certain effective external field $J$ exciting the boson--field. The 'current' $J(x)=J({\bf x})\delta(t)$ is non-zero in the collision area and we assume that the collision time can be taken infinite short. Let us calculate the unitary effect of the Hamiltonian (5) on the vacuum: \begin{equation} \ket{\psi_J}=\exp\left(-i\sum_{\bf k}}\def\kp{{\k^\prime}(J_\k}\def\Jkp{J_\kp^\star\a_\k}\def\akp{\a_\kp+J_\k}\def\Jkp{J_\kp\a_\k}\def\akp{\a_\kp^\dagger)\right)\ket{0} =\exp\left(-\frac{1}{2}\sum_{\bf k}}\def\kp{{\k^\prime}\vertJ_\k}\def\Jkp{J_\kp\vert^2 -i\sum_{\bf k}}\def\kp{{\k^\prime}J_\k}\def\Jkp{J_\kp\a_\k}\def\akp{\a_\kp^\dagger\right)\ket{0}~, \end{equation} which is otherwise a product coherent state $\prod_{\bf k}}\def\kp{{\k^\prime} \otimes\ket{-iJ_\k}\def\Jkp{J_\kp}$. These final states $\ket{\psi_J}$ are pure states whereas the IMS are mixed ones. Obviously, no unitary dynamics can create mixed states from pure ones. Therefore, I consider unitary dynamics in {\it random} external fields: I assume Gaussian distribution for the stochastic fluctuations of the current $J$. Let the mean values $M[J_\k}\def\Jkp{J_\kp]$ be always zero. Also we assume that $M[\JkpJ_\k}\def\Jkp{J_\kp]\equiv0$, which is equivalent to a random phase for all $J_\k}\def\Jkp{J_\kp$. We denote the only non--vanishing correlations by the non-negative Hermitian matrix $\alpha$: \begin{equation} M[\JkpJ_\k}\def\Jkp{J_\kp^\star]=\alpha_{\kp{\bf k}}\def\kp{{\k^\prime}}~. \end{equation} After these preparations, we can define the density operator $\hat\rho$ of the final state as the stochastic mean value of the pure coherent states (13): \begin{equation} \hat\rho=M\bigl[\ket{\psi_J}\bra{\psi_j}\bigr]~. \end{equation} Substituting the Eq.~(13) and taking the stochastic mean over $J$ of the Gaussian correlation (14) we are led directly to the form (4) of IMS density operators. The 1-particle density matrix $\rho$ and the parameter $\nu$ are related to the correlation matrix $\alpha$ of the current by easily invertible matrix relations: \begin{equation} \nu\rho=\frac{\alpha}{1+\alpha}~,~~~\alpha=\frac{\nu\rho}{1-\nu\rho}~. \end{equation} Measuring the 1-particle density matrix we could, up to the validity of the model, calculate the structure of the external current. Although such measurement is (so far) not completely possible we shall see in Sec.5 that the inclusive correlation function gives the modulus of $\alpha$ directly. It is also seen from Eqs.~(16) that a Gaussian shape, like \cite{BiaZal.EJP} \begin{equation} \rho_{\kp,{\bf k}}\def\kp{{\k^\prime}}\sim\exp \left(-\frac{1}{2\Delta^2}{\bf k}}\def\kp{{\k^\prime}_+^2-\frac{1}{2}R^2{\bf k}}\def\kp{{\k^\prime}_-^2\right), ~~{\bf k}}\def\kp{{\k^\prime}_+=\frac{{\bf k}}\def\kp{{\k^\prime}+\kp}{2},~~{\bf k}}\def\kp{{\k^\prime}_-={\bf k}}\def\kp{{\k^\prime}-\kp~, \end{equation} for the 1-particle density matrix is not compatible with a Gaussian shaped current correlation matrix $\alpha_{\kp,{\bf k}}\def\kp{{\k^\prime}}$ and {\it vice versa}. \section{Final state distribution {\it vs.} external current} The final state distributions in IMS can be directly related to the currents $J$. The generating functional (5) can conveniently be re-expressed as an averaged functional over the fluctuating external current $J$: \begin{equation} G[u]=M\left[\exp\left(\sum_{\bf k}}\def\kp{{\k^\prime} (u_{\bf k}}\def\kp{{\k^\prime}-1)\vertJ_\k}\def\Jkp{J_\kp\vert^2\right)\right]~, \end{equation} which is of course equivalent to Eqs. (6) or (9). The above equation has numerous useful consequences. The multiplicity distribution can be written in this form: \begin{equation} p_r=M \Bigl[\left(\sum\vert J\vert^2\right)^r\exp\left(-\sum\vert J\vert^2\right) \Bigr]~, \end{equation} while the factorial moments take the same form but without the exponential factor $exp(-\sum\vert J\vert^2)$, i.e.: \begin{equation} F_r=M \Bigl[\left(\sum\vert J\vert^2\right)^r\Bigr]~. \end{equation} This phenomenon also characterizes the differences between the expressions of the exclusive and the inclusive distribution functions, respectively: \begin{eqnarray} f(1,2,\dots,r)=M\Bigl[\vert J_1\vert^2,\vert J_2\vert^2,\dots,\vert J_r\vert^2 \cases{\exp(-\sum\vert J\vert^2)\Bigr]~~~&(exclusive)\cr \Bigr] &(inclusive) } \end{eqnarray} as well as of the correlation functions. In particular, the inclusive correlation functions take the following form: \begin{equation} C(1,2,\dots,r)= M\left[\vert J_1\vert^2 \vert J_2\vert^2 \dots \vert J_r\vert^2\right]_c~, \end{equation} the inclusive ones would contain the ominous exponential factor, too. The notation $M[\dots]_c$ means that in the 'expectation value' only the 'connected grafs' are to be taken into the account. In case of Eq.~(22) it yields $(r-1)!$ 'cycles', i.e. the 'cycle' $\alpha_{12}\alpha_{23}\dots\alpha_{r1}$ and its variants for permutations of $2,\dots,r$ \cite{BiaZal.PLB}. One can easily summarize the main result of this Section as follows. The counts $n_{\bf k}}\def\kp{{\k^\prime}$, measured simultaneously in a collision event, are {\it statistically identical} to the corresponding spectral intensities $\vertJ_\k}\def\Jkp{J_\kp\vert^2$. Like their distributions, their corresponding moments are identical as well: \begin{equation} \langle n_1 n_2 \dots n_r\rangle = M\left[\vert J_1\vert^2 \vert J_2\vert^2 \dots \vert J_r\vert^2\right]. \end{equation} \section{Bose-Einstein-condensation} The IMS class of density operators (1) has a particular asymptotics. The 'inverse temperature' matrix $\beta$ must be positive. If it were degenerate the state (1) would not exist at all. A degenerate $\beta$ can formally be interpreted as if the mode of zero eigenvalue became infinite hot. This mode is, in fact, becoming more and more populated but the infinite population is unattainable. Nonetheless, an IMS with almost degenerate $\beta$ would really be a Bose-Einstein condensate since this only requires a big finite number of bosons in a single quantum state. Speculations that the point of degeneracy, i.e. the point $\nu=1/\Vert\rho\Vert$, is the point of condensation (like in thermal Bose--systems) can not be verified for the IMS. Let us first recapitulate the kinematics of an IMS condensate. The condensate mode does not interfere with the other modes so we can discuss it separately. We assume that our IMS is dominated by the condensate mode. The 1-particle density matrix has the form $\rho_{\kp{\bf k}}\def\kp{{\k^\prime}}=\varphi_\kp\varphi_{\bf k}}\def\kp{{\k^\prime}^\dagger$ where $\varphi_{\bf k}}\def\kp{{\k^\prime}$ is the condensate mode's wave function. If we introduce the condensate absorption operator ${\hat a}_c=\sum_{\bf k}}\def\kp{{\k^\prime}\varphi_{\bf k}}\def\kp{{\k^\prime}{\hat a}_{\bf k}}\def\kp{{\k^\prime}$ then, using Eqs.(1-4), the condensate IMS can be written as a thermal equilibrium state at temperature $T=-1/\log\nu$: \begin{equation} \hat\rho_c=(1-e^{-1/T})\exp\left(-\frac{{\hat a}_c^\dagger{\hat a}_c}{T}\right)~. \end{equation} This state assumes a Hamiltonian ${\hat a}_c^\dagger{\hat a}_c$ which is not the real case, the condensate is not even stationary in general. Yet, the form (24) is completely proper to calculate characteristics of the state by a thermal analogy. For instance, Eq.~(5) yields directly the generating functional in the form: \begin{equation} G[u]=\frac{1-e^{-1/T}}{1-\exp(-1/T)\sum_{\bf k}}\def\kp{{\k^\prime} u_{\bf k}}\def\kp{{\k^\prime}\vert\varphi_{\bf k}}\def\kp{{\k^\prime}\vert^2}~, \end{equation} with the canonical thermal multiplicity distribution \begin{equation} p_n=(1-e^{-1/T})e^{-n/T} \end{equation} of mean multiplicity \begin{equation} \langle n\rangle = \frac{1}{e^{n/T}-1}~. \end{equation} Let us observe that approaching the 'condensation point' corresponds to $T\rightarrow\infty$ and the population of the 'Bose-condensate' increases to the infinity while it is remaining thermally distributed all the time. Now I turn to the dynamic conditions for the fluctuating external current $J$. In the special case of the condensate IMS, the second relation in Eq.~(16) becomes simply $\alpha_{\kp{\bf k}}\def\kp{{\k^\prime}}=\langle n\rangle\rho_{\kp{\bf k}}\def\kp{{\k^\prime}}=\langle n\rangle\varphi_\kp\varphi_{\bf k}}\def\kp{{\k^\prime}^\star$. Recall the definition (14) of $\alpha$ as the current's correlation matrix, which yields the following relation: \begin{equation} M[\JkpJ_\k}\def\Jkp{J_\kp^\star]=\langle n\rangle \varphi_\kp\varphi_{\bf k}}\def\kp{{\k^\prime}^\star~. \end{equation} Regarding that $M[\JkpJ_\k}\def\Jkp{J_\kp]$ should vanish by assumption (Sec.4), the Gaussian fluctations satisfying the above relation must take the form \begin{equation} J_\k}\def\Jkp{J_\kp=z\sqrt{\langle n\rangle}\varphi_k \end{equation} for all ${\bf k}}\def\kp{{\k^\prime}$, where $z$ is a random complex number of the standard Gauss--distribution $(1/\pi)\exp(-\vert z\vert^2)d^2z$. Taking the stochastic mean of the modulus square of both sides we obtain: \begin{equation} \vertJ_\k}\def\Jkp{J_\kp\vert^2=\langle n\rangle \vert\varphi_{\bf k}}\def\kp{{\k^\prime}\vert^2~, \end{equation} which also leads to \begin{equation} \langle n\rangle=\sum_{\bf k}}\def\kp{{\k^\prime} \vertJ_\k}\def\Jkp{J_\kp\vert^2~. \end{equation} The Eqs.~(29-31) show the simple way how the pulse of the effective current $J$ determines the condensate wave function and the mean population. Actually, the mean multiplicity is identical to the overall intensity of the current pulse (31). The pulse's normalized spectral density is equal to the modulus square of the condensate wave function (30). The Eq.~(29) seems, however, to be very restrictive since it imposes the same random phase and weight simultaneously for all current amplitudes $J_{\bf k}}\def\kp{{\k^\prime}$. \section{Outlook: master equations} I outline a possible generalization of the simple dynamics proposed in Sec.4. Let us replace the Hamiltonian (12) by \begin{equation} \hat H(t)=g(t)\sum_{\bf k}}\def\kp{{\k^\prime}\left(J_\k}\def\Jkp{J_\kp^\star(t)\a_\k}\def\akp{\a_\kp+J_\k}\def\Jkp{J_\kp(t)\a_\k}\def\akp{\a_\kp^\dagger\right)~, \e where $g(t)$ is a normalized function of characteristic width $\Delta t$, controlling the intensity of particle creation. The time-dependent currents $J_{\bf k}}\def\kp{{\k^\prime}(t)$ are random functions of zero mean; let their correlator be of non-stationary white-noise type: \b g(t)M[\Jkp(t^\prime)J_\k}\def\Jkp{J_\kp^\star(t)]=\delta(t^\prime-t)\alpha_{\kp{\bf k}}\def\kp{{\k^\prime}}~. \end{equation} In the limit $\Delta t\rightarrow0$, the random dynamics represented by Eqs.~(32,33) reduces to the simplistic dynamics (12,14) of Sect.~4. The Hamiltonian (32) with the white-noise currents (33) yield the following master equation \cite{GorKosSud} for the noise-averaged density operator (in interaction picture): \begin{equation} \frac{d\hat\rho}{dt}=g \sum_{{\bf k}}\def\kp{{\k^\prime},\kp}\alpha_{\kp{\bf k}}\def\kp{{\k^\prime}} \left(\akp^\dagger\hat\rho\a_\k}\def\akp{\a_\kp-\frac{1}{2}\{\a_\k}\def\akp{\a_\kp\akp^\dagger,\hat\rho\}\right)~. \end{equation} One solves the master equation with vacuum initial state. If $1/\Delta t$ is much greater than the typical pion energy then, via the relationships (16), the final state will tend to the IMS (1). For larger $\Delta t$ the simple relations (16) do not hold. Though analytic calculations are still possible one can turn to very powerful Monte-Carlo methods \cite{Dio88} developped for Markovian master equations. These MC algorithms will yield the density operators of the multiparticle final states without "struggling through" \cite{Zaj93} the usual Wigner-function formalism. \section{Summary} The aim of the Letter was partly pedagogical. To avoid Bose--symmetrization 'by hand', I used standard quantum mechanical considerations to construct and to analyze the 'independent multiboson states'. I showed how these states emerge from a simplistic version of chaotic current models and I derived the relationship between the IMS states and the correlator of the currents. I briefly recapitulated the generating functional representation of multiparticle counts. Beyond methodological matters, I found that the 'Bose-Einstein condensate' would be thermally populated and the condensation point corresponds to the infinite hot state. I restricted my analysis for the IMS of Sec.1 and for the simplest chaotic current mechanisms with trivial time-dependences. The simple choice allows for transparent relationships between the current and the final multiparticle state. (Other works, like e.g. Ref.\cite{Ornetal}, incorporate a more realistic space-time evolution of the multiparticle source.) There is, nonetheless, a particular advantage of any underlying dynamics whether realistic or not. Usually it allows economic simulation methods for the physical quantities of interest. To this end, I proposed a quantum master equation known to be suitable to efficient Monte-Carlo simulations. \section*{Acknowledgement} I thank Tam\'as Cs\"org\H o and S\'andor Hegyi for useful discussions.
\section{INTRODUCTION} Non-abelian gauge theories are endowed with local gauge invariance \cite{cheng}. Local gauge invariance leads to relations between Green's functions of gauge and or ghost fields collectively denoted by WT identities \cite{iz}. Formulation of gauge theories in covariant gauges necessities inclusion of unphysical degrees of freedom corresponding to the longitudinal and the time like gauge fields. Unitarity of S-Matrix (whenever defined ) requires that these modes do not contribute to the intermediate states in the cutting equations \cite{tht}. The contributions from such intermediate states are canceled by contributions from diagrams containing ghost intermediate states. This is demonstrated in gauge theories with the use of the ( on-shell) WT identities. Thus the cancellation of intermediate states coming from longitudinal /time like gauge degrees of freedom and the ghost degrees of freedom ( we denote this set by R) together is one of the essential consequence of WT identities. These, in turn, follow from the BRS symmetry ( or gauge invariance )\cite{iz}. This, in turn, suggests that there should be a formulation of BRS symmetry where the above set of R of degrees of freedom are explicitly linked together. There exist many attempts to link $(A, c, \bar{c})$ fields together. In view of both the commuting and anti-commuting degrees of freedom involved, this points to a `` Supersymmetric/Superfields " formulation. A number of superspace /superfield formulation have been written down which exhibit the BRS symmetry in terms of translations or rotations in superspace \cite{pr2,sdj}. For a brief summary of superspace /superfields formulations and their comparison see comments in Ref. \cite{pro} and references therein. The superspace formulation of Ref. \cite{ssp} constructed superfields $ A(x, \theta, \bar{ \theta }), c(x, \theta ,\bar{ \theta })$ and $ \bar{c}(x, \theta ,\bar{\theta })$ by hand by ascribing the values of the additional components $ ( A_{\theta }, A_{\bar{\theta }}\cdots $ etc) equal to the BRS/Anti-BRS variations \cite{btm} of these. They exhibited the BRS/Anti-BRS structure thereby. However as the structure of the superfields was restricted there one could not construct a full-fledged field theory of these superfields. The works of references \cite{sdj} and \cite{zpc} (and subsequent works ) attempted to constructed a field theory of superfields in superspace. Here the superfields were entirely unconstrained and the superrotations could be carried out in the formulation. In fact the BRS and Anti-BRS were identified with these superrotations and the corresponding WT identities understood as arising from these. \cite{zpc}. These constructions had a broken $ OSp(3,1|2) $ symmetry. While these superspace formulation exhibited the BRS/Anti-BRS structure \cite{sp,zpc} , and the renormalization properties \cite{D49} of gauge theories compactly and correctly. They treated the anti-ghost field asymmetrically ( and as far as we know it is necessary to do this, to exhibit the renormalization properties in linear gauges. Moreover, the underlying $ OSp(3,1|2) $ symmetry was broken one. Following the motivations outlined earlier, we have attempted, in this work, a formulation that (i) is a superspace field theories as in Ref. \cite{sdj} (ii) treats gauge, ghost and anti-ghost fields together in one single supermultiplet. (iii)has an underlying formal $ OSp(3,1|2) $ symmetry as the basis of construction as a limiting symmetry of the Lagrange density. (iv)has WT identities that formally imply that this symmetry becomes exact as gauge parameter $\eta \rightarrow \infty $ (v) corresponds to the Yang-Mills theory in one of its formulation. In fact we find that the superspace formulation presented here corresponds to the BRS/Anti-BRS invariant formulation of Baulieu and Thierry-Mieg \cite{btm} with $\beta =1$ ( $ c, \bar{c} $ symmetric case). We interpret heuristically the last property in the following manner. We note ( as done in sec. IIC) that as $ \eta \rightarrow \infty $ the gauge boson propagator is dominated by the longitudinal and time like modes. Thus in this limit, the multiplet $(A, c, \bar{c} )$ is dominated by just the set R of extra modes which enter the unitarity discussion via WT identities . It is precisely in this limit, the $ OSp(3,1|2) $ symmetry is becoming exact. We now briefly present the plan of the paper. In Sec II, we shall review the underlying superspace /superfield structure and the OSp group properties. We briefly discuss the BRS/Anti-BRS symmetric formulation of Reference \cite{btm} . We also include a brief discussion on the mode structure of propagator as $\eta\rightarrow \infty $. In section III, we present the superspace formulation and show its equivalence to the BRS/Anti-BRS symmetric formulation with $\beta =1$ \cite{btm}. In this section IV, we show that the generating functional $W[\bar{X}] $ is asymptotically $(\eta\rightarrow \infty )$ invariant under the $ OSp(3,1|2) $ group. In sec V, we elaborate on the physical meaning of the result so obtained. \section{PRELIMINARY} \subsection{BRS/Anti-BRS symmetric action} In this section, we shall review the known results on BRS and anti-BRS symmetries of effective action in gauge theories \cite{btm}. We consider the most general effective action in linear gauges given by Baulieu and Thierry - Mieg \cite{btm} that has BRS/anti-BRS invariance, when expressed entirely in terms of necessary fields $A,c, \bar{c}$ (and no auxiliary fields) \begin{equation} S_{eff}[A,c,\bar{c}]=\int \,d^4x \left[-\frac{1}{4} F_{\mu\nu}^{\alpha}F^{\alpha \mu\nu} - \sum_{\alpha}\frac{(\partial \cdot A^{\alpha} )^2}{2\eta }-{\cal L}_G \right] \label{o1} \end{equation} with \begin{eqnarray} {\cal L}_G & =& (1-\frac{1}{2}\beta)\partial ^\mu \bar{c} D_\mu c + \frac{\beta}{2}D^\mu\bar{c}\partial_\mu c -\frac{1}{2}\beta(1-\frac{1}{2}\beta) \frac{\eta }{2} g^2[f^{\alpha\beta\gamma}\bar{c}^\beta c^\gamma]^2\label{o2}\\ &=&\partial^\mu\bar{c}D_\mu c+\frac{\beta}{2}gf^{\alpha\beta\gamma}\partial\cdot A^\alpha\bar{c}^ \beta c^\gamma+\frac{1}{8}\beta(1-\frac{1}{2}\beta)\eta g^2f^{\alpha\beta\gamma}\bar{c}^\beta\bar{c}^\gamma f^{\alpha\eta\xi}c^\eta c^\xi \label{o3} \end{eqnarray} Here we are assuming a Yang- Mills theory with a simple gauge group and introducing the following notations:\\ \begin{eqnarray*} \mbox{Lie Algebra:}\;\; [T^\alpha,T^\beta] &=& i f^{\alpha\beta\gamma} T^\gamma \\ \mbox{Covariant derivative:}\;\;{(D_\mu c)}^\alpha &\equiv & D^{\alpha\beta}_ \mu c^\beta =(-\partial_\mu \delta^{\alpha\beta}+gf^{\alpha\beta\gamma} A^{\gamma}_\mu)c^\beta\\ \end{eqnarray*} $f^{\alpha\beta\gamma}$ are totally antisymmetric. Note here we have changed the convention for the covariant derivative just to bring it in line with notations of Ref.~\cite{sdj}. This action has the global symmetries under the following transformations\\ \begin{eqnarray} \mbox{BRS :}\nonumber \\ \delta A^{\alpha}_\mu & =&(D_\mu c)^\alpha \delta \Lambda \nonumber \\ \delta c^\alpha & =&-\frac{1}{2}gf^{\alpha\beta\gamma}c^\beta c^\gamma \delta\Lambda \nonumber\\ \delta \bar{c}^\alpha&=& \left(-\frac{\partial\cdot A^\alpha}{\eta }-\frac{1}{2}\beta g f^{\alpha\beta\gamma}\bar{c}^\beta c^\gamma \right)\delta\Lambda\label{o4} \\ \mbox{ and anti-BRS:}\nonumber \\ \delta A^{\alpha}_\mu &=& {(D_\mu \bar{c})}^\alpha \delta\Lambda \nonumber \\ \delta \bar{c}^\alpha &=&-\frac{1}{2}gf^{\alpha\beta\gamma}\bar{c}^\beta\bar{c}^ \gamma \delta \Lambda \nonumber\\ \delta c^\alpha &=&\left(-\frac{\partial\cdot A^\alpha}{\eta } -(1-\frac{1}{2}\beta)g f^{\alpha \beta\gamma}\bar{c}^\beta c^\gamma \right)\delta\Lambda\label{o5} \end {eqnarray} In the anti-BRS transformations the role of $ c$ and $\bar{c} $ are interchanged in addition to change in some coefficients. Note that $\beta=0$ case yields the usual Faddeev- Popov action and $\beta=1$ yields an action symmetric in $c$ and $\bar{c}$. \subsection{ Superspace, Superfields and Invariants} We shall work in superspace formulation of Yang-Mills theory given in Ref. [5], which we briefly review in this section. The superspace formulation uses an underlying six-dimensional superspace described by superspace coordinate $ \bar{x}^i \equiv (x^\mu, \lambda , \theta )$ with $\lambda , \theta $ are real Grassmannian variables. Superfields and supersources are function of superspace coordinates. The superspace is endowed with a metric $g_{ij}$ , with only non zero components \begin{equation} g_{00}= -g_{11}= -g_{22} =-g_{33}= -g_{45}= g_{54} = 1 \end{equation} The infinitesimal orthosymplectric coordinate transformations which leaves the norm of the supervector, $x^ig_{ji} x^j$ invariant are consists of (i) Six Lorentz transformations which leaves $g_{\mu\nu}x^{\mu}x^{\nu}$ invariant. (ii) Three simplectic transformations which leaves $\lambda \theta $ invariant and characterized by three infinitesimal parameter. (iii) and eight SUSY transformations given by \begin{eqnarray} x^{\prime \mu} &=& x^\mu + \epsilon ^\mu a \lambda +\delta ^\mu b \theta \nonumber \\ \lambda ^\prime &=& \lambda +\delta ^\mu x_\mu b \nonumber \\ \theta ^\prime &=& \theta -\epsilon ^\mu x_\mu a \label{susy} \end{eqnarray} [Where $ \epsilon _\mu , \delta _\mu$ are arbitrary four vectors and $a,b$ are real infinitesimal Grassmannians] generated by $S_{4\mu}$ and $S_{5\mu}$. $S_{4\mu} $ generates transformations with $ \delta =0$ and $ S_{5\mu}$ generates transformations with $ \epsilon =0$. $\frac{ \partial }{\partial x_i}$ are transforms as a covariant vector under the $ OSp(3,1|2) $ transformations and they are give by \begin{eqnarray} \frac{ \partial} { \partial x^{\prime \mu}} &=&\frac{ \partial }{\partial x^\mu} +\epsilon ^\mu a \frac{ \partial }{ \partial \theta } -\delta ^\mu b \frac{ \partial }{\partial \lambda } \nonumber \\ \frac{ \partial }{ \partial \lambda ^\prime} &=&\frac{ \partial }{\partial \lambda} +\epsilon ^\mu a \frac{ \partial }{\partial x_\mu} \nonumber \\ \frac{ \partial }{\partial \theta^\prime }&=& \frac{ \partial }{\partial \theta } + \delta ^\mu b \frac{ \partial }{\partial x_\mu} \end{eqnarray} and the vector superfields $ A_i(\bar{x})\equiv ( A_\mu(\bar{x}) ,c_4 (\bar{x}) , c_5 (\bar{x}) )$ also transform as covariants vectors under these $ OSp(3,1|2) $ transformations and given by \begin{eqnarray} A^{\prime}_\mu &=& A_\mu + \epsilon _\mu a c_5 -\delta _\mu b c_4 \nonumber \\ c_4 ^\prime &=& c_4 +\epsilon ^\mu a A_\mu \nonumber \\ c_5 ^\prime &=& c_5 +\delta ^\mu b A_\mu \label{29} \end{eqnarray} The transformations for the vector supersource $ \bar{X}^i (\bar{x}) $ are such that $\bar{X}^i (\bar{x}) \bar{A}_i (\bar{x}) $ remain invariant under $ OSp(3,1|2) $ We define the scalar product as \begin{equation} A\cdot B = A_i g^{ji}B_j \equiv A^jB_j \end{equation} And the tensor invariants are defined as \begin{equation} A\cdot B C\cdot D = C_iA_jB_kD_l g^{kj}g^{li} = T_{ij}T_{kl}g^{kj}g^{li} \end{equation} where $A^iB_i$ is a commuting quantity. Using the above definitions of scalar products we construct the following OSp invariant quantities $(i)\ F_{ij}F_{kl}g^{kj}g^{li} \ \ (ii)\ \partial ^i[A_i \partial ^jA_j ] \ \ (iii) \ \partial ^i[A^j \partial _jA_i] \ \ (iv)\ \partial ^i[ (\partial _i A^j)A_j]$. Where the superspace field strength tensor, $F_{ij}$ is defined as \begin{equation} F^\alpha _{ij} (\bar{x}) = \partial _i A_j^\alpha (\bar{x}) -A^\alpha _i \stackrel{\leftarrow}{\partial }_j +gf^{\alpha \beta \gamma }A^\beta _i (\bar{x}) A^\gamma _j (\bar{x}) \end{equation} \subsection{ Mode structure of gauge propagator} The propagator in the linear gauges is given by \begin{equation} i\Delta_{F\mu\nu} (k,\eta) = \frac{-i}{k^2+i \epsilon }\left [g_{\mu\nu}- \frac{ k_\mu k_\nu}{k^2+i \epsilon } (1-\eta) \right ] \end{equation} We imagine expanding the gauge field (in the momentum space ) in the basis consisting of the transverse, the longitudinal and the time like degrees of freedom, \begin{equation} A_\mu(k) = \sum_{i=1}^4 \epsilon ^i_\mu(k) a_{(i)} (k^2) \end{equation} with $\epsilon_\mu^{(1)} $ and $\epsilon ^{(2)}_\mu(k) $ are transverse degrees of freedom with \begin{equation} \epsilon _0^{(i)}(k) =0\ , \ \ \ \vec{k}\cdot\vec{\epsilon }^{(i)}(\vec{k}) =0 \ \ \ i=1,2 \end{equation} and \begin{equation} \epsilon ^{(3)}_\mu(k) = \left (0, \frac{ -\vec{k}}{|k|} \right ); \ \ \ \epsilon ^{(4)}_\mu = (1,0,0,0) \end{equation} We note the orthonormality properties, \begin{equation} \epsilon ^{*(i)}\cdot \epsilon ^{(j)} = -\delta ^{ij} + 2 \delta _{i0} \delta _{j0} \ \ \ \ \ (i,j= 1,2,3,4) \end{equation} Then the gauge boson propagator \begin{equation} \left < A_\mu(k) \ A_\nu(-k) \right > = \sum_{i,j=1}^4 \epsilon ^{(i)}_\mu (k) \epsilon _\nu ^{(j)}(-k) \left <a_{(i)}(k^2)a_{(j)}(k^2) \right > \end{equation} We recall \begin{equation} \epsilon ^{(i)}_\mu (k) \epsilon_\nu^{( i)}(-k)=-(g_{\mu\nu} -\delta _{\mu 0} \delta _{\nu 0})+ \frac{ k_\mu k_\nu (1-\delta _{\mu 0})(1-\delta _{\nu 0})}{\vec{k}^2} \end{equation} We then find by comparison, \begin{eqnarray} \left <a_{(i)}(k^2),\ a_{(j)}(k^2) \right > &=& \frac{ \delta _{ij}}{k^2+ i \epsilon } \ \ \ 1\le i,j\le 2 \nonumber \\ \left < a_j(k^2),\ a_i(k^2) \right >& = &\left <a_{(i)}(k^2),\ a_{(j)}(k^2) \right > = 0 \ \ \ 1\le i\le 2;\ 3\le j\le 4 \nonumber \\ \left <a_{(3)}(k^2),\ a_{(3)}(k^2)\right >&=& -(\eta-1)\frac{|\vec{k}|^2}{ (k^2+i \epsilon )^2}\nonumber \\ \left <a_{(4)}(k^2),\ a_{(4)}(k^2) \right >&=& -\left [1+\frac{(\eta-1) |\vec{k}|^2}{k^2+i \epsilon }\right ] \frac{ 1}{k^2+ i \epsilon } \nonumber \\ \left <a_{(3)}(k^2),\ a_{(4)}(k^2) \right > &=& \frac{ (\eta-1)}{2} \frac{ |\vec{k}|k_0}{(k^2+i \epsilon )^2} \end{eqnarray} Thus we see that as $ \eta \rightarrow \infty$, the correlation functions of modes containing $a_{(3)}$ ( the longitudinal ) or $a_{(4)}$ ( the time like ) go to $\infty, $ while those containing the transverse components remain unaltered. We now scale as, \begin{equation} a_{(3)} = \sqrt{ \frac{ \eta}{3}}\tilde{a_3};\ \ \ a_{(4)} = \sqrt{ \frac{ \eta}{3}}\tilde{a_4};\ \ \tilde{a_i} \equiv a_i \ \ i=1,2 \end{equation} ( The factor of $ \frac{ 1}{3} $ is for future convenience only.) Then all correlation functions $\left < \tilde{a_i}(k^2),\ \tilde{a_j}(k^2) \right >$ have $\eta$- independent limits. Then the expansion of the gauge field reads \begin{eqnarray} A_\mu(k)& =& \sum_{i=1}^2 \epsilon ^{(i)}_\mu(k)\tilde{a}_{(i)}(k^2)+ \sqrt{ \frac{ \eta}{3} } \epsilon ^{(3)}_\mu\tilde{a}_{(3)}(k^2) + \sqrt{ \frac{ \eta}{3}}\epsilon ^{(4)}_\mu \tilde{a}_{(4)}(k^2) \nonumber \\ &\equiv& A^T_\mu +\sqrt{ \frac{ \eta}{3}}A^L_\mu +\sqrt{ \frac{ \eta}{3}} A^t_\mu \end{eqnarray} The relation above exhibits explicitly the $\sqrt{\eta}$ factors that say that (after suitable normalization) the longitudinal and the time like components of a general gauge field become dominant as $\eta \rightarrow\infty$. This remark will find application in See. V in the context of the supermultiplet structure of fields introduced. \section{ CONSTRUCTION OF SUPERSPACE ACTION} In this section, we shall present the construction of the superspace action which is equivalent to the Yang-Mills theory in its BRS/Anti-BRS invariant formulation with $\beta =1$ [ See sec. IIA]. The building block of the superspace action is a covariant vector field $\bar{A}_i(\bar{x}) = \left ( A_\mu (\bar{x}) ,\ c_4 (\bar{x}) ,\ c_5 (\bar{x}) \right ) ( c_4$ and $ c_5 $ will turn out to be related to the antighost field $\bar{c}$ and the ghost field $c$). We shall also introduce commuting contravariant vector source $\bar{X}^i (\bar{x}) $ . Unlike Ref. \cite{pro} we don,t however need a scalar superfield and scalar supersource. As we shall see later, the Lagrange density turns out to have a graded structure as the gauge parameter $\eta\rightarrow \infty$ (i) ${\cal L}_0 $ is an OSp invariant action of $O(\eta^0)$ (ii) ${\cal L}_1 $ turns out to be also an OSp invariant, but of $O(\eta^{-1})$ and (iii) ${\cal L}_2 $ is an Sp(2) invariant symmetry breaking term of $O(\eta^{-2})$. Explicitly\footnote{The parameter $\beta$ in ${\cal L}_1$ is not to be confused with $\beta$ in the BRS /anti-BRS invariant action of Sec. IIA which will always be taken to be 1 in this work.}, \begin{eqnarray} {\cal L}_0 &=& \frac{ 1}{4} F_{ij}F_{kl}g^{kj}g^{li} \nonumber \\ {\cal L}_1 &=& \alpha \partial ^i \left [A_i \partial ^jA_j \right ] +\beta \partial ^i \left [A^j \partial _j A_i \right ] + \gamma \partial ^i \left [A^j \partial _i A_j \right ] \nonumber \\ {\cal L}_2 &=& \frac{\kappa}{2} c^a \partial ^i \partial _i c_a \ \ \ \ a =4,5 \end{eqnarray} To this we add the source terms \begin{equation} {\cal L}_s = \frac{ \partial }{\partial \theta } \frac{ \partial }{\partial \lambda }[\bar{X}^i\bar{A}_i ] \end{equation} Under $ OSp(3,1|2) $ transformations ${\cal L}_s$ changes at most by a total derivative. We then construct the generating functional \begin{equation} W[\bar{X}] = \int {\cal D}A \exp{i\int d^4x \left [ {\cal L}_0 + {\cal L}_1+{\cal L}_2 +{\cal L}_s \right ]} \end{equation} where \begin{equation} {\cal D}\bar{A}\equiv \prod_{i=0}^5{\cal D}A_i (\bar{x}) \equiv \prod_{i=0}^5 {\cal D}A_i(x) {\cal D} A_{i, \lambda }(x) {\cal D}A_{i,\theta }(x) \end{equation} In order to establish the equivalence of the above generating functional with that of the Yang-Mills theory, we carry out the integrations over the variables $A_{i, \lambda } , A_{i, \theta }$ explicitly as in Ref. \cite{sdj}. The procedure is very straightforward and hence we shall not present the details; but only the final result. Omitting the source terms for the present ( as these are not relevant to the equivalence ) we find \begin{equation} W[\bar{X}] = \int {\cal D} A_\mu (x) {\cal D} c_4(x) {\cal D} c_5(x)\exp {i\left [ S_0[A_\mu,c_4,c_5] +\mbox{Source terms} \right ]} \end{equation} with\footnote{ $A_{i, \lambda \theta }$ here are not dynamical fields \cite{sdj} and can be dropped (i.e. can be set to zero by hand ) in future.} omitting redundant terms in $A_{i, \lambda \theta }$. \begin{eqnarray} S_0[A,c_4,c_5] &=& - \frac{ 1}{4} F_{\mu\nu}F^{\mu\nu} - \frac{ (\alpha -\beta)^2}{2(2 \alpha -\gamma -\beta) }(\partial \cdot A)^2 - \frac{ \alpha -\beta}{1-2 \gamma } \left [ D_\mu c_4 \partial ^\mu c_5 + \partial _\mu c_4 D^\mu c_5 \right ] \nonumber \\ &+& \frac{ 2 \gamma }{1-2 \gamma } D_\mu c_4D^\mu c_5 - \frac{ 3(\gamma +\beta)}{2(\beta+\gamma +1)} (gfc_4c_5)^2-\left [ (\alpha -\beta)^2- \kappa \right ]\partial _\mu c_4 \partial ^\mu c_5 \label{s0} \end{eqnarray} Comparing with Eq. (\ref{o1}) , we see that $S_0$ of (\ref{s0}) is compatible with the action in Eq. (\ref{o1}) only if \begin{equation} \mbox{coefficient of } (fA_\mu c_4)(fA_\mu c_5) = 0 \ \ \Rightarrow \gamma =0 \end{equation} and \begin{equation} \eta = \frac{ 2 \alpha -\beta}{(\alpha -\beta)^2} \end{equation} Further, we use the freedom to define $c_4$ and $c_5$ to set \begin{equation} c_5= \frac{ 1}{\sqrt{2(\alpha -\beta)}} c;\ \ \ c_4 = \frac{ 1}{\sqrt{2(\alpha -\beta)}} \bar{c} \label{39} \end{equation} then the two actions coincide if, further, \begin{eqnarray} - \frac{ 3\beta}{2(\beta+1)} &=& \frac{ 2 \alpha -\beta}{2} \label{cc} \\ \mbox{ and } \kappa & =& (\alpha -\beta)^2 \label{cc1} \end{eqnarray} The quadratic equation of (\ref{cc}) has solutions \begin{equation} \beta = (\alpha +1) \left [ 1\pm \sqrt{1+ \frac{ 2 \alpha }{(\alpha +1)^2}} \right ] \label{root} \end{equation} Either values of $\beta$ would be acceptable for our purpose. We shall see in sec. IV that the solution in (\ref{root}) with -ve sign leads to a superspace Lagrange density that has asymptotic ( i.e. as $\eta\rightarrow \infty $) symmetry ; and hence we shall make this choice. Thus the equivalence of the two action with $\beta $ and $ \kappa $ given in terms of (\ref{cc}) and (\ref{cc1}) is established completely. We shall, however, be particularly interested in a special case. We further use the freedom we have in choosing the free parameter $\alpha $ to let $0<\alpha \ll 1$ then, \begin{equation} \beta = (\alpha +1) \left [ 1-\sqrt{1+ \frac{ 2 \alpha }{(\alpha +1)^2} }\right ] \simeq - \frac{ \alpha }{\alpha +1} \simeq - \alpha \end{equation} Then the gauge parameter becomes \begin{equation} \eta = \frac{ 2 \alpha -\beta}{(\alpha -\beta)^2} \simeq \frac{ 3 \alpha }{(2 \alpha )^2} = \frac{ 3}{4 \alpha } \end{equation} Thus, as $ \alpha \rightarrow 0^+ ,$ our superspace action represents the BRS/Anti-BRS action with the parameter $\beta$ in \ref{o2} set equal to 1 and $\eta\rightarrow \infty. $ Further, \begin{equation} \kappa = ( \alpha -\beta)^2 \simeq 4 \alpha ^2 \end{equation} expressing all parameters in terms of $\eta (\mbox{ as } \eta \rightarrow \infty )$, \begin{equation} -\beta \simeq \alpha \simeq (\frac{3}{4\eta}), \ \ \ \kappa \simeq 4.\frac{9}{16\eta^2} = \frac{9}{4\eta^2} \end{equation} and the scaling of (\ref{39}) are re-expressed as \begin{equation} c_5 = \sqrt{ \frac{ \eta}{3} } c , \ \ \ \ c_4 = \sqrt{ \frac{ \eta}{3}} \bar{c} \end{equation} To summarize, the superspace action with one free parameter $\eta$ \begin{equation} \int d^4x \left \{ {\cal L}_0 + \alpha \partial ^i \left [ \partial ^jA_j A_i- A^j \partial _jA_i \right ] + 4 \alpha ^2 c^a \partial ^i \partial _i c_a \right \} \end{equation} ( with $\alpha = \frac{3}{ 4\eta}$ and coefficients valid for $\eta $ large ) is equivalent in the superspace generating functional to the BRS/Anti-BRS symmetric action with gauge parameter $\eta (\rightarrow \infty)$ . In the next section we shall establish the asymptotic OSp invariance for $W[\bar{K}]$ in other words, the formal equation of the form \begin{equation} \left [W[\bar{X}^\prime ] - W[\bar{X}]\right ]|_{X^i_{,\theta}=0} = 0( \frac{ 1}{\eta}) \end{equation} \section{ $OSp(3,1|2)$ WT IDENTITIES} In this section, we shall consider the consequence of the $ OSp(3,1 2) $ transformations on the source $\bar{X}^i (\bar{x}) $ to obtain the WT identities for the broken $ OSp(3,1|2) $ symmetry. The result is summarized by the statement which in effect says that $\eta \rightarrow \infty \ \ W$ recovers $ OSp(3,1|2) $ invariance under the conditions clarified under. It is also shown how this WT identity embodies exact BRS/anti-BRS symmetry in the form of the statements \ref{w24a} and \ref{w24b}. We begin with the generating functional \begin{equation} W[\bar{X} (\bar{x}) ] = \int {\cal D} \bar{A}(\bar{x}) \exp \left \{i\int d^4x \left [{\cal L}_0[\bar{A}] +{\cal L}_1[ \bar{A}]+{\cal L}_2[\bar{A}]+{\cal L}_s \right ] \right \} \label{w} \end{equation} We perform an $ OSp(3,1|2) $ rotation on the sources $\bar{X}^i$ \begin{equation} \bar{X}^i (\bar{x}) \rightarrow \bar{X}^{i\prime}(\bar{x} ) \label{w2} \end{equation} with \begin{equation} \bar{X}^{\prime i} (\bar{x}) = \bar{X}^j(\Lambda^{-1}\bar{x})\Lambda_j^i \label{w3} \end{equation} Under this transformation, we have the invariance \begin{eqnarray} X^i (\bar{x}) A_i (\bar{x}) &=& X^{\prime i}( \Lambda \bar{x})A^\prime _i (\Lambda \bar{x}) \nonumber \\ &=& X^{\prime i} (\Lambda \bar{x})\tilde{\Lambda}_i^jA_j (\bar{x}) \label{w4} \end{eqnarray} [Here $\tilde{\Lambda}$ is defined in \ref{29}, in particular for $S_{4\mu}$ and $S_{5\mu}$ transformations ]. Then using (\ref{w4}), we have \begin{eqnarray} \int d^4x {\cal L}_s = \int \frac{ \partial }{\partial \theta } \frac{ \partial }{\partial \lambda } \left [ \bar{X}^i (\bar{x}) A_i (\bar{x}) \right ]\nonumber \\ &=& \int \frac{ \partial }{\partial \theta } \frac{ \partial }{\partial \lambda } \left [ \bar{X}^{i\prime } (\Lambda\bar{x}) \tilde{\Lambda}^j_i A_j (\bar{x}) \right] \label{w5} \end{eqnarray} In view of the $ SO(3,1) \times Sp(2)$ invariance of the entire $S$, we expect new informations to emerge from the transformations associated with additional supersymmetries $S_{4\mu}$ and $S_{5\mu}.$ Hence we now restrict ourselves to the $\tilde{\Lambda}$ of Eq. (\ref{29}) given in Sec. IIB. We note now that \begin{equation} \Lambda (\bar{x}) = \left ( x^\mu + \epsilon ^\mu a \lambda + \delta ^\mu b \theta , \lambda + \delta ^\mu b x_\mu , \theta - \epsilon ^\mu a x_\mu \right ) \label{w6} \end{equation} and express \begin{equation} \int d^4x {\cal L}_s =\int d^4x \frac{ \partial }{ \partial \theta } \frac{ \partial }{ \partial \lambda } \left [ \bar{X}^{i \prime }(\bar{x}) \tilde{\Lambda}^j_i A_j (\bar{x}) + \left \{ (\epsilon ^\mu a \lambda + \delta b \theta ) \partial _\mu \bar{X}^i (\bar{x}) + \delta \cdot x b \bar{X}^i_{, \lambda }-\epsilon \cdot x a \bar{X}^i_{, \theta } \right \} A_i (\bar{x}) \right ] \label{w7} \end{equation} In the last term, we have used the infinitesimal nature of $\epsilon ^\mu $ and $\delta ^\mu $ to replace $\bar{X}^\prime \rightarrow \bar{X} $ and $\tilde{\Lambda}\rightarrow 1$. Further, \begin{eqnarray} \int d^4x {\cal L}_s(\bar{X})&-& \int d^4x {\cal L}_s(\bar{X}^\prime )= \int d^4x \left [ +( \epsilon ^\mu a \lambda + \delta ^\mu b \theta ) \frac{ \partial }{ \partial \theta } \frac{ \partial }{ \partial \lambda } (\partial _\mu \bar{X}^i (\bar{x}) A_i (\bar{x}) )\right ]\nonumber \\ &+& \int d^4x \left [ \epsilon ^\mu a \frac{ \partial }{\partial \theta }(\partial _\mu \bar{X}^iA_i) + \delta ^\mu b \frac{ \partial }{ \partial \lambda }( \partial _\mu \bar{X}^iA_i) \right ]+ \int d^4x \frac{ \partial }{ \partial \theta } \frac{ \partial }{ \partial \lambda } \left [ \bar{X}^i_{, \lambda \theta } \epsilon \cdot x A_{i, \theta }-\bar{X}^i_{, \lambda \theta } \delta \cdot x A_{i, \lambda }\right ] \nonumber \\ &+& \int d^4 x \frac{ \partial }{\partial \theta } \frac{ \partial }{ \partial \lambda } \left [ \bar{X}^\mu( \epsilon _\mu a c_5 - \delta _\mu b c_4) + \bar{X}^4(\epsilon ^\mu a A_\mu +\bar{X}^5 \delta ^\mu b A_\mu \right ] \label{w8} \end{eqnarray} Using (\ref{w8}) we can write down change in $W[\bar{X}]$ under an infinitesimal OSp transformation (\ref{susy}) \begin{eqnarray} \delta W[\bar{X}] && = W[\bar{X}]-W[\bar{X}^\prime ] = << \int d^4x (\epsilon ^\mu a \lambda + \delta ^\mu b \theta ) \sum_S \partial _\mu S(\bar{x}) \frac{ \delta W}{\delta S} \nonumber \\ && +i\int d^4x \left [ \bar{X}^\mu _{,\lambda \theta }\left [ \epsilon \cdot x a A_{\mu, \theta }+ \epsilon _\mu a c_5 - \delta _\mu b c_4 - \delta \cdot x b A_{\mu, \lambda } \right ] + \bar{X}^4_{ ,\lambda \theta }\left [\epsilon \cdot x a c_{4, \theta}+\epsilon ^\mu a A_\mu - \delta \cdot x b c_{4, \lambda } \right ] \nonumber\right . \\ && \left . + \bar{X}^5_{, \lambda \theta } \left [ \epsilon \cdot x a c_{5, \theta }+ \delta ^\mu b A_\mu - \delta \cdot x b c_{5, \lambda } \right ] + \bar{X}^\mu _{, \lambda } \left [-\delta _\mu b c_{4, \theta }+ \epsilon _\mu a c_{5, \theta } \right ] - \bar{X}^4_{, \lambda } \epsilon ^\mu a A_{\mu , \theta} -\bar{X}^5_{, \lambda } \delta ^\mu b A_{\mu, \theta }\right . \nonumber \\ && \left . -\bar{X}^\mu_{, \theta } \left [ \epsilon _\mu a c_{5, \lambda }- \delta _\mu b c_{4, \lambda } \right ] + \bar{X}^4_{, \theta } \epsilon ^\mu a A_{\mu , \lambda }+\bar{X}^5_{, \theta } \delta ^\mu b A_{\mu , \lambda }+ \left [ \epsilon^\mu a \partial _\mu \bar{X}^i_{, \theta } + \delta ^\mu b \partial _\mu \bar{X}^i_{, \lambda } \right ] A_i \nonumber\right . \\ && \left . \epsilon^\mu a \partial _\mu \bar{X}^\nu A_{\nu , \theta } + \delta ^\mu b \partial _\mu \bar{X}^\nu A_{\nu , \lambda } - \epsilon^\mu a \partial _\mu (\bar{X}^4c_{4, \theta }+\bar{X}^5 c_{5, \theta }) - \delta ^\mu b \partial _\mu (\bar{X}^4c_{4, \lambda }+\bar{X}^5c_{5, \lambda }) \right ] >> \label{w9} \end{eqnarray} Here we have dropped terms proportional to $A_{, \lambda \theta }^i$ ( as these fields can be set to zero). The double bracket, $<< >> $ has been used to denote that the expression inside it is actually inside the path integral. We now evaluate $\delta W[\bar{X}]$ for the``supersymmetry transformations " $ S_{4\mu}$ only; i.e. set $ \delta =0$. We, further, note that the sources $-\bar{X}^\mu_{, \theta }$ are to generate the Green,s functions of the composite operator involved in the anti-BRS transformations . These are not required to evaluate the basic Green,s functions of the Yang-Mills theory, nor the BRS WT identities. Hence, we evaluate (\ref{w9}) at $\bar{X}^\mu _{, \theta } =0$. [ The spurious terms involving $\partial _\mu \bar{X}^i$ can also be set to zero as $\bar{X}^i$ are sources for $A_{, \lambda \theta }$ which are redundant field]. Now, the first term on the right hand side of (\ref{w9}) ( here $\sum_S$ goes over the sources $\bar{X}^i; \bar{X}^i_{, \lambda }; \bar{X}^i_{, \theta };\bar{X}^i_{,\lambda \theta}$) vanishes by the translational invariance of $W[\bar{X}]$ in $x^\mu$. We organize the rest of the terms in $ \delta W$ as \begin{eqnarray} \delta W[\bar{X}]|_{\bar{X}^i_{, \theta }=0=\bar{X}^i(x)} &=& <<i \int d^4x \left \{ \bar{X}^\mu_{, \lambda \theta }\left ( \epsilon\cdot x a A_{\mu, \theta }+ \epsilon _\mu a c_5 \right ) +\bar{X}^4_{, \lambda \theta } \left ( \epsilon\cdot x a c_{4, \theta }+ \epsilon^\mu a A_\mu \right ) \right . \nonumber \\ && \left . + \bar{X}^5_{ ,\lambda \theta }(\epsilon\cdot x a c_{5, \theta }) + \bar{X}^ \mu_{, \lambda }\epsilon _\mu a c_{5, \theta } - \bar{X}^4_{, \lambda } \epsilon ^\mu aA_{\mu , \theta } \right \} >> \label{w10} \end{eqnarray} We shall simplify the expression on the right hand side employing the 6-D gauge invariance of ${\cal L}_0$ \cite{bpm}. We consider the gauge transformations \begin{equation} \delta A_i = D_i( c_5 \epsilon\cdot x a) \label{w11} \end{equation} and the consequent transformations \begin{equation} \delta A_{i,\theta } = \frac{ \partial }{ \partial \theta } \left [ D_i( c_5 \epsilon\cdot x a) \right ]; \ \ \ \delta A_{i,\lambda } = \frac{ \partial }{ \partial \lambda } \left [ D_i( c_5 \epsilon\cdot x a) \right ] \label{w11a} \end{equation} Under (\ref{w11}) and (\ref{w11a}) ; ${\cal L}_0$ is gauge invariant \begin{equation} \int d^4x \left [ \delta A_i \frac{ \delta S_0}{\delta A_i} + \delta A_{i, \theta } \frac{ \delta S_0}{ \delta A_{i, \theta} } + \delta A_{i, \lambda } \frac{ \delta S_0}{\delta A_{i, \lambda }} \right ]=0 \label{w12} \end{equation} We now invoke the equations of motion \begin{eqnarray} \frac{ \delta S_0}{\delta A_\mu}&=& -\bar{X}^\mu_{, \lambda \theta }+ (\alpha -\beta) \partial _\mu(c_{4, \theta }-c_{5, \lambda }) \nonumber \\ \frac{ \delta S_0}{\delta c_4}&=& \bar{X}^4_{, \lambda \theta }+ (\alpha -\beta) \partial ^\mu(A_{\mu, \theta })+\kappa \partial ^2 c_5 \nonumber \\ \frac{ \delta S_0}{\delta c_5}&=& \bar{X}^5_{, \lambda \theta }- (\alpha -\beta) \partial ^\mu A_{\mu, \lambda }-\kappa \partial ^2 c_4\nonumber \\ \frac{ \delta S_0}{\delta A_{\mu , \lambda } }&=& \bar{X}^\mu_{, \theta }- (\alpha -\beta) \partial _\mu c_5 \nonumber \\ \frac{ \delta S_0}{\delta c_{4, \lambda }}&=& \bar{X}^4_{, \theta }- 2 \beta c_{5, \theta } \nonumber \\ \frac{ \delta S_0}{\delta c_{5, \lambda }}&=& \bar{X}^5_{, \theta }+ (\alpha -\beta) \partial\cdot A-( \alpha -\beta) c_{5, \lambda }+ 2 \alpha c_{4, \theta } \nonumber \\ \frac{ \delta S_0}{\delta A_{\mu , \theta }}&=& -\bar{X}^\mu_{, \lambda}+ (\alpha -\beta) \partial _\mu c_4 \nonumber \\ \frac{ \delta S_0}{\delta c_{4, \theta }}&=&- \bar{X}^4_{, \lambda }+ (\alpha -\beta) \partial\cdot A-( \alpha -\beta) c_{4, \theta }+ 2 \alpha c_{5,\lambda } \nonumber \\ \frac{ \delta S_0}{\delta c_{5 , \theta }}&=& -\bar{X}^5_{, \lambda}+ 2\beta c_{4, \lambda } \label{w13} \end{eqnarray} [It is understood that these Eqs are in double brackets.] Using (\ref{w13}) in (\ref{w12}), we obtain \begin{eqnarray} << -i\int d^4x && \left [ D_\mu(\epsilon \cdot x a c_5) \bar{X}^\mu_{, \lambda \theta } -D_4(\epsilon\cdot x a c_5)\bar{X}^4_{, \lambda \theta } -D_5(\epsilon\cdot x ac_5)\bar{X}^5_{, \lambda \theta } + \frac{ \partial }{\partial \theta }[D_\mu(\epsilon\cdot x ac_5)]\bar{X}^\mu_{, \lambda }\nonumber \right .\\ && \left . \frac{ \partial }{\partial \theta }[D_4 \epsilon \cdot x a c_5 ] \bar{X}^4_{, \lambda }+ \frac{ \partial }{\partial \theta }[D_5(\epsilon\cdot x a c_5)]\bar{X}^5_{, \lambda } \right ] \nonumber \\ && = O( \alpha , \bar{X}^i_{, \theta }) \label{w14} \end{eqnarray} We now subtract (\ref{w14}) from (\ref{w10}) to obtain \begin{eqnarray} \delta W[\bar{X}] |_{\bar{X}^i_{, \theta} =0=\bar{X}^i}= << i\int d^4x \epsilon\cdot x a \left \{ (A_{\mu, \theta }+D_\mu c_5) \bar{X}^\mu_{, \lambda \theta}-(c_{4, \theta }-D_4c_5)\bar{X}^4_{, \lambda \theta }-(c_{5, \theta }-D_5c_5)\bar{X}^5_{, \lambda \theta } \nonumber \right .\\ \left . -\bar{X}^4_{, \lambda \theta } \epsilon ^\mu a A_\mu - \frac{ \partial }{ \partial \theta }(D_\mu c_5)\bar{X}^\mu_{, \lambda }+ \frac{ \partial }{\partial \theta }(D_4c_5) \bar{X}^4_{, \lambda }+ \frac{ \partial }{\partial \theta }(D_5c_5)\bar{X}^5_{, \lambda } -\bar{X}^4_{, \lambda } \epsilon ^\mu a A_{\mu , \theta }\right \}>> \label{w15} \end{eqnarray} Now we recall the equation of motion \begin{eqnarray} << A_{\mu , \theta }+D_\mu c_5 + (\alpha -\beta)\partial _\mu c_5 - \bar{X}_{, \theta }^\mu>> =0 \label{w16a}\\ << c_{5, \theta } - D_5c_5 + 2\beta c_{5, \theta } -\bar{X}^4_{, \theta} >> =0 \label{w16b} \end{eqnarray} and \begin{equation} \frac{ \partial }{\partial \theta }(D_5c_5) = -2f^{\alpha \beta\gamma}c^\beta_{5, \theta }c^\gamma_5 =0 \label{w17} \end{equation} Which can be obtained by using (\ref{w16b}) at $\bar{X}^4_{, \theta }=0$. We further have the equation of motion of $c_{4, \theta }$ and $c_{5, \lambda }$. \begin{eqnarray} <<c_{4, \theta } + c_{5, \lambda } + g fc_4c_5 +\beta(c_{4, \theta } + c_{5, \lambda }) - \frac{ \bar{X}^4_{, \lambda }-\bar{X}^5_{, \theta }}{2}>> =0 \label{w16c} \\ <<(2 \alpha -\beta)(c_{4, \theta }-c_{5, \lambda }) + (\alpha -\beta) \partial \cdot A- \frac{ \bar{X}^5_{, \theta }-\bar{X}^4_{ ,\lambda}}{2}>>=0 \label{w16d} \end{eqnarray} Subtracting (\ref{w16d}) from (\ref{w16c}) and setting $\bar{X}^5_{, \theta }$ we obtain \begin{eqnarray} << c_{4, \theta }&-&D_4c_5>>|_{\bar{X}^5_{, \theta }=0}\nonumber \\ && = -\beta<<c_{4, \theta }+c_{5, \lambda }>>+ (2 \alpha -\beta)<<c_{4, \theta} - c_{5, \lambda }>>+ (\alpha -\beta) << \partial \cdot A>> = O( \alpha ) \label{w18} \end{eqnarray} Further, using (\ref{w16a}) and (\ref{w16b}), we obtain ( at $ \bar{X}^i_{, \theta } =0$ ) \begin{equation} \frac{ \partial }{\partial \theta }(D_\mu c_5) = O( \alpha ) \label{w19} \end{equation} [ We recognize in (\ref{w17}) and in (\ref{w19})the usual BRS invariance statement of $\frac{ 1}{2} fcc$ and $ D_\mu c.$ ] We further recall the equation of motion of $c_4$ \begin{eqnarray} <<-D_\mu^{\beta \alpha } (A_{\mu , \theta } +D_\mu c_5)^\alpha + f^{\alpha \beta \gamma }c^\gamma _4(2c_{5, \theta }^\alpha +gf^{\alpha \eta \delta }c_5^\eta c_5^\delta )-gf^{\alpha \beta \gamma }c_5^\gamma (c_{4, \theta }^\alpha +c_{5, \lambda }^\alpha +gf^{\alpha \eta \delta }c_4^\eta c_5^\delta ) \nonumber \\ -\bar{X}^4_{, \lambda \theta } -( \alpha -\beta) \partial ^\mu A_{\mu , \theta } +\kappa \partial ^2 c_5 >> =0 \label{w20} \end{eqnarray} On account of (\ref{w16a}),(\ref{w16b}) and (\ref{w18}) used successively in the left hand side of (\ref{w20}) these terms vanish at $\bar{X}^i_{, \theta } =0$. We thus conclude, \begin{equation} << \bar{X}^4_{, \lambda \theta } \epsilon^\mu a A_\mu >> = <<O( \alpha )>> \label{w21} \end{equation} Using (\ref{w16a}),(\ref{w18}),(\ref{w16b}),(\ref{w21}),(\ref{w19}),(\ref{w16d}),(\ref{w17}) and (\ref{w16d}) in the successive terms on the right hand side of (\ref{w15}) we obtain, \begin{equation} \delta W[\bar{X}]|_{\bar{X}^i_{, \theta } =0=\bar{X}^i} = <<O(\frac{ 1}{\eta} )>> \label{w22} \end{equation} We could have alternatively considered the symmetry associated with $S_{5\mu}$ transformations $ ( \delta \neq 0, \epsilon =0 )$ in (\ref{w9}). In view of the overall $Sp(2)$ symmetry of the formulation, we will obtain the analogous relation \begin{equation} \delta W[\bar{X}]|_{X^i_{, \lambda }=0=X^i} = O( \frac{ 1}{\eta}) \label{w23} \end{equation} The relations (\ref{w22}) and (\ref{w23}) are statements of formal $ OSp(3,1| 2) $ symmetry as $\eta\rightarrow \infty $. These contain in them the consequences of BRS and anti-BRS invariance. These consequences can be obtained in a manner analogous to the argument following Eq. (19) of the Ref. \cite{sp} ( See also \cite{bpm} for alternative procedure for the entire derivation). They result in equations \begin{eqnarray} \frac{ \partial W}{\partial \theta }|_{X^i_{, \theta }=0=X^i} &=& O( \frac{ 1}{\eta}) \label{w24a} \\ \frac{ \partial W}{\partial \lambda }|_{X^i_{, \lambda }=0=X_i} &=& O( \frac{ 1}{\eta}) \label{w24b} \end{eqnarray} for BRS and anti-BRS symmetry respectively. These equations can also be alternatively verified evaluating $W[X]$ along the lines of Ref \cite{sdj} and evaluating $\frac{ \partial W}{\partial \theta }$ and $ \frac{ \partial W}{\partial \lambda }$ along the line of Ref \cite{zpc,sdj2} using BRS /anti-BRS symmetry of the resultant $W$ \section{ PHYSICAL MEANING OF $OSp(3,1|2)$ INVARIANCE} We expand the multiplet $\bar{A}_i (\bar{x}) $ explicitly as \begin{eqnarray} \bar{A}_i (\bar{x}) &=& \left (\begin{array}{l} A_\mu (\bar{x}) \\ c_4 (\bar{x}) \\ c_5 (\bar{x}) \end{array} \right ) \equiv \left ( \begin{array}{l} A_\mu^T (\bar{x}) +\sqrt{ \frac{ \eta}{3}}A_\mu^L (\bar{x}) +\sqrt{ \frac{ \eta}{3}}A_\mu^t (\bar{x}) \\ \sqrt{ \frac{ \eta}{3}}\bar{c} (\bar{x}) \\ \sqrt{ \frac{ \eta}{3}}c (\bar{x}) \end{array} \right ) \nonumber \\ &=& \sqrt{ \frac{ \eta}{3}} \left ( \begin{array}{l} A_\mu^L (\bar{x}) +A_\mu^t (\bar{x}) \\ \bar{c} (\bar{x}) \\ c (\bar{x}) \end{array} \right ) + \left ( \begin{array}{l} A_\mu^T (\bar{x}) \\0\\ \end{array} \right ) \nonumber \\ &\equiv& \sqrt{\frac{\eta}{3}} \left [\bar{A}_\mu^R (\bar{x}) +\sqrt{ \frac{ 3}{\eta}} A_\mu^T (\bar{x}) \right ] \end{eqnarray} Here $\bar{A}_\mu^R (\bar{x}) $, in particular, contains the fields corresponding to the set $R$ We further expand the transformation laws for fields under $ OSp(3,1|2) $ viz. \begin{eqnarray} A_\mu ^\prime (\bar{x}) &=& A_\mu (\bar{x}) -\delta _\mu b c_4 (\bar{x}) + \epsilon _\mu a c_5 (\bar{x}) -( \epsilon a \lambda +\delta b \theta )^\nu \partial _\nu A_\mu (\bar{x}) \nonumber \\ &-& \delta \cdot x b A_{\mu, \lambda } (\bar{x}) + \epsilon \cdot x a A_{\mu, \theta } (\bar{x}) \nonumber \\ c^ \prime _4 (\bar{x}) &=& c_4 (\bar{x}) + \epsilon ^\mu a A_\mu (\bar{x}) -( \epsilon a \lambda + \delta b \theta )^\nu \partial _\nu c_4 \nonumber \\ &-& \delta \cdot x b c_{4, \lambda } (\bar{x}) + \epsilon \cdot x c_{4, \theta } (\bar{x}) \nonumber \\ c_5^ \prime (\bar{x}) &=& c_5 (\bar{x}) + \delta ^\mu b A_\mu (\bar{x}) -( \epsilon a \lambda +\delta b \theta )^\nu \partial _\nu c_5 \nonumber \\ &-& \delta \cdot x b c_{5, \lambda }(\bar{x}) + \epsilon \cdot x a c_{5, \theta } (\bar{x}) \end{eqnarray} in powers of $\eta $. We find that these read \begin{eqnarray} A_\mu^{R \prime } (\bar{x}) &=& A_\mu^R (\bar{x}) +P^R_{\mu\nu}[- \delta _\mu b \bar{c} (\bar{x}) +\epsilon _\mu a c (\bar{x}) -( \epsilon a \lambda + \delta b \theta )^\nu \partial _\nu A^R_\mu (\bar{x}) \nonumber \\ &-& \delta \cdot x b A_{\mu, \lambda }^R (\bar{x}) + \epsilon \cdot x a A^R_{\mu, \theta} (\bar{x})] + 0( \frac{ 1}{\sqrt{\eta}}) \nonumber \\ \bar{c}^\prime (\bar{x}) &=& \bar{c} (\bar{x}) + \epsilon a A_\mu ^R (\bar{x}) -( \epsilon a \lambda + \delta b \theta )^\nu \partial _\nu \bar{c} (\bar{x}) \nonumber \\ &-& \delta \cdot x b \bar{c}^R_{, \lambda }(\bar{x}) + \epsilon \cdot x a \bar{c}_{, \theta } (\bar{x}) + 0( \frac{ 1}{\sqrt{\eta}}) \nonumber \\ c^\prime (\bar{x}) &=& c (\bar{x}) + \delta ^\mu b A_\mu^R (\bar{x}) -( \epsilon a \lambda + \delta b \theta )^\nu \partial _\nu c (\bar{x}) \nonumber \\ &-& \delta \cdot x b c_{, \lambda } (\bar{x}) + \epsilon \cdot x a c_{, \theta } (\bar{x}) + 0( \frac{ 1}{\sqrt{\eta}}) \label{53} \end{eqnarray} and \begin{equation} A^{\prime T}_\mu (\bar{x}) = A_\mu ^T (\bar{x}) + 0(\sqrt{\eta}) \label{53a} \end{equation} [$P^R_{\mu\nu}$ is the projection operator that projects away the transverse part]. We note that as $\eta\rightarrow\infty$, (\ref{53}) refers to the transformations within the set $A^R$ only. Thus, in the limit $\eta \rightarrow \infty $, the $ OSp(3,1|2) $ transformations, in particular, contain a set of symmetry transformations among the members of the redundant set R. The WT identities are a particular consequence of these symmetries. A special consequence of the WT identities is the cancellation of the contributions from the set R in the intermediate states in the unitarity relations using the Cutkowsky rules \cite{tht} . In the present superspace formulation, we have an explicit construction of a set of symmetry transformations amongst this set R; originating from the original $ OSp(3,1|2) $ transformation which, as we have shown, lead to WT identities in particular. Thus, this formulation can be looked upon as an explicit realization of that relationship that is expected to exist with in the fields of the set R that is ultimately known to lead to mutual cancellations in the cutting equations. {\bf Acknowledgment:} SDJ acknowledges the hospitality provided by the theory group, Saha Institute of Nuclear Physics, Calcutta, India, where part of the work was done. \newpage
\section{Introduction} Many models for new physics include particles which may live long enough to traverse our detector. Such a particle may be long-lived due to a new conserved quantum number (such as R-parity in SUSY), or its decays may be suppressed. If these particles are charged, they could be detected directly~\cite{theories,phenom,penet}. The most stringest limits from direct searches for such particles come from LEPII and CDF. LEPII has set mass limits of about 82 GeV/c$^{2}$ on stable sleptons within the framework of Gauge-Mediated Supersymmetry Breaking (GMSB)~\cite{lep2}. CDF set a limit of M $>$ 139 GeV/c$^{2}$ on stable quarks from data taken in 88/89~\cite{cdf_run1}. Other searches have also been performed at LEP and other accelerators~\cite{other_searches}. We have searched for Charged Massive Particles (CHAMPS) in 90 pb$^{-1}$ of Run1B data at CDF. We have aimed to be as model independent as possible, but the search naturally divides into two separate searches, one for strongly produced particles and one for those produced via the weak interaction. The strongly produced particles would have a larger production cross section, so the region of interest is at high mass where the background is expected to be low. We expect that these particles would fragment into an integer charged meson within a jet. For the strong search, we use a 4$^{th}$ generation quark as a reference model. Weakly interacting CHAMPS would have a lower production cross-section so the region of sensitivity is at lower mass where the background is higher. These weakly produced particles are expected to be isolated, which allows us to reduce the background significantly. We use Drell-Yan production of a slepton within a GMSB scenario as a reference model. \section{Signature} Massive particles will be produced with relatively low $\beta\gamma$, so the signature is a highly ionizing track. We use the Central Tracking Chamber (CTC) and the Silicon Vertex Detector (SVX) to independently measure dE/dx. For both searches, we make a cut of $\beta\gamma <$ 0.85, and for the strong search we additionally look at $\beta\gamma <$ 0.70 for added sensitivity at low mass. We searched track by track from events which came in on three different triggers -- the muon trigger, $E_{T}\hspace{-0.17in}/\hspace{0.1in}> $ 35 GeV trigger, and the electron trigger. The muon trigger is directly sensitive to a CHAMP since massive particles would be penetrating and appear as muons, while all three triggers are sensitive to production of sleptons in cascade decays of other sparticles which produce in addition a neutrino, electron or muon. \section{Kinematic Cuts} All considered tracks are required to pass quality cuts which reduce backgrounds from misreconstructed tracks. Due to timing considerations, particles moving slower than $\beta\gamma$ = 0.4 are not reconstructed. For a particle of mass 90 GeV/c$^{2}$, this corresponds to a minimum reconstructable momentum of 35 GeV. We therefore require that each track has a momentum greater than 35 GeV/c, which eliminates much of our low momentum background. We also cut on the mass $M_{SVX}$ calculated from the dE/dx measurement in the SVX and momentum of the track. For the strong search, we perform a search in bins of mass M, and make a cut on $M_{SVX} > 0.6 \times M$. Our background falls off with momentum, and this $M_{SVX}$ cut forces a stiffer dE/dx cut at low momentum where the background is higher. For the weak search, the mass cut is simply $M_{SVX} > 60 GeV/c^{2}$. These tracks must additionally pass an isolation cut, namely we require less than 4 GeV of calorimeter energy or total track pT within a cone of 0.4 = $\sqrt{\Delta\eta^{2}+\Delta\phi^{2}}$ of the track. \section{Background} The only background expected is from fakes, where a track has a dE/dx which fluctuated high, or where overlapping tracks reconstruct as a single track. In order to estimate the background, we use a control sample at low momentum (15 $< |\vec{p}| <$ 35 GeV) to calculate the expected fake rate, then multiply the fake rate by the number of tracks which enter our sample in the signal region $|\vec{p}| >$ 35 GeV to get the expected background. The fake rate is calculated in the control sample by taking the ratio of the number of tracks which pass the dE/dx cut in both the CTC and the SVX to the total number of tracks. Since the fake rates are extrapolated from low momentum to high momentum, we plot the fake rate as a function of momentum and check that it is flat. We observe that the fake rate falls off below 20 GeV/c, especially in the CTC. This is because our control sample is contaminated with kaons, which are still along the relativistic rise on their dE/dx vs. p curve below 20 GeV/c and pull down the fake rate in this region in momentum. For the muon triggered data sample, we take the control sample to be tracks which lie in the momentum region 20 $< |\vec{p}| <$ 35 GeV and for the $E_{T}\hspace{-0.17in}/\hspace{0.1in}$ sample and the electron sample our control region is 25 $< |\vec{p}| <$ 35 GeV. \section{Results} The results of each search are tabulated below. \begin{table} \caption{Strong Search} \begin{tabular}{lrr} Sample & Bkgd. & Obs. \\ \tableline Muon Trigger& 12 $\pm$ 2 & 12 \\ $E_{T}\hspace{-0.17in}/\hspace{0.1in}$ Trigger & 63 $\pm$ 9 & 45 \\ Muon Trigger ($\beta\gamma < $0.70)& 2.5 $\pm$ 0.8 & 2 \\ \end{tabular} \end{table} \begin{table} \caption{Weak Search (Includes Isolation Cut} \begin{tabular}{lrr} Sample &Bkgd.&Obs.\\ \tableline Muon Trigger& 0.85 $\pm$ 0.25 & 0\\ $E_{T}\hspace{-0.17in}/\hspace{0.1in}$ Trigger &4.0 $\pm$ 2.8 &1\\ Electron Trigger & 0.72 $\pm$ 0.54 & 0\\ \end{tabular} \end{table} We calculate the expected mass distributions from the control sample. To do this, we take the momentum of each track and calculate what its dE/dx would be if it came from a particle with mass M, and then use the control sample to find the probability that the track fakes that dE/dx. We do this for the entire range in M and sum over all tracks. The data and expected background distributions are shown in Figures~\ref{Mass_dist_utrig} and ~\ref{Mass_dist_mettrig}. No excess over background is observed. \begin{figure}[ht] \centerline{\epsfxsize 2.3 truein \epsfbox{connolly713fig1.ps}} \vskip -.2 cm \caption[]{ \label{Mass_dist_utrig} \small Observed M$_{SVX}$ distribution for tracks passing all the cuts for the strong search in the muon triggered data sample. The solid curves are the expected background distributions.} \end{figure} \begin{figure}[ht] \centerline{\epsfxsize 1.8 truein \epsfbox{connolly713fig2.ps}} \vskip -.2 cm \caption[]{ \label{Mass_dist_mettrig} \small Observed M$_{SVX}$ distribution for tracks passing all the cuts for the strong search in the $E_{T}\hspace{-0.17in}/\hspace{0.1in}$ triggered data sample. The solid curve is the distribution of the 63 $\pm$ 9 expected background events.} \end{figure} \section{Efficiencies and Systematics} For the strong search, the efficiency depends on the quark charge due to fragmentation effects. It increases from 1.5 - 3\% for q=2/3 and 0.8-1.5\% for q=1/3 in the mass range 100-240 GeV/c$^{2}$. The largest systematic uncertainties come from modeling interactions in the calorimeter. This gives an uncertainty of 20\% for q=1/3 and 13\% for q=2/3. Systematic uncertainties from luminosity (7.5\%) and from the choice of PDF (7\%) are also significant. The total systematic uncertainty is 23\% and 17\% for q=1/3 and q=2/3 respectively. For the weak search, the total efficiency is 3\% for Drell-Yan slepton production, to which only the muon trigger is sensitive. Once slepton production from cascade decays are included, the efficiency for finding sleptons using only the muon triggered data sample increases to 6\%. The total efficiency after including cascades and all three triggers is 8\%. The largest systematics on this efficiencies come from the luminosity (7.5\%). Other systematics include track quality cuts (4.9\%) and choice of PDF (5.5\%). Once cascades are included, the $E_{T}\hspace{-0.17in}/\hspace{0.1in}$ trigger and initial/final state radiation (ISR/FSR) become significant systematics. \section{Limits} Figure~\ref{hsq_limits} shows the cross-section limits we derive for heavy stable quarks from the results of our strong search. From comparison with the theoretical prediction, we conclude M $>$ 195 GeV/c$^{2}$ for q=1/3 and M $>$ 220 GeV/c$^{2}$ for q=2/3. These limits are the most stringent limits from a direct search to date. Figure~\ref{slepton_limits} shows the cross-section limits derived for Drell-Yan production of long-lived sleptons in a GMSB scenario with the stau as the Next-to-Lightest Susy Particle (NLSP). We are over an order of magnitude away from being sensitive to the theoretical prediction. When all sparticle production modes with cascade decays to a slepton are included, the efficiency goes up, bringing our limit down and the theoretically predicted slepton production cross section increases. This limit was derived for one model point with three slepton co-NLSP's (N$_{5}$=3, M/$\Lambda$=3, tan$\beta$=3, $\mu>$ 0) and found to be $\sigma_{95\%}\approx 550 fb$, still nearly an order of magnitude away from the predicted 80 fb. \begin{figure}[ht] \centerline{\epsfxsize 2.3 truein \epsfbox{connolly713fig3.ps}} \vskip -.2 cm \caption[]{ \label{hsq_limits} \small Limits on heavy stable quarks derived from results of the strong search.} \end{figure} \begin{figure}[ht] \centerline{\epsfxsize 2.3 truein \epsfbox{connolly713fig4.ps}} \vskip -.2 cm \caption[]{ \label{slepton_limits} \small Limits on long-lived Drell-Yan sleptons derived from results of the strong search.} \end{figure} \section{Prospects for RunII} RunII at the Tevatron will have an increased factor of 20 in luminosity and CDF will have a number of detector improvements. The upgraded detector will have double-sided silicon with stereo strips for z measurement, 7 silicon layers instead of 4, and smaller cell size in the new tracking chamber. This smaller cell size will cut down on the background from overlapping tracks. Additionally, the now approved Time of Flight system with a resolution of approximately 0.1ns will greatly increase our region of sensitivity in momentum from $\approx$ 85 GeV/c to $\approx$ 175 GeV/c for a 100 GeV/c$^{2}$ particle. \section{Conclusions} We have searched for heavy stable particles using 90$pb^{-1}$ of Run1B data at CDF. No excess of events over background was observed. We set preliminary limits on heavy stable quarks at M $>$ 195 GeV/c$^{2}$ and M $>$ 220 GeV/c$^{2}$ for q=1/3 and q=2/3 respectively. Using GMSB with three slepton co-NLSP's as a reference model, we set a limit on slepton production at $\sigma_{95\%} < 550 fb$. This limit is a factor of 7 away from the theoretical prediction. With increased luminosity and detector upgrades in RunII we expect to have sensitivity to slepton production in GMSB models. \section{Acknowledgements} I am grateful to D. Stuart for introducing me to this interesting topic, and for his guidance throughout this analysis. I would also like to thank M. Shapiro for her advice and support.
\section*{Introduction} Let $(X,{\cal O}_X(1)={\cal O}_X(H))$ be a smooth polarized variety defined over an algebraic closed field of arbitrary characteristic. We assume ${\cal O}_X(1)$ to be very ample. Additionally, let $E$ be a $\mu$-semistable vector bundle of rank two on $X$. We want to show that there exists an integer $m$ only depending on the characteristic numbers $H^{\dim(X)}$ and $(c_1(E)^2-4c_2(E)).H^{\dim(X)-2}$ such that the restriction of $E$ to a general element of $|mH|$ is semistable. Such effective bounds have been known only for the case that the characteristic is zero. In this case the restriction theorem of Flenner (see \cite{Fle}) gives effective bounds on $m$ for semistable bundles of arbitrary rank. On the other hand there are results of Mehta and Ramanathan which say that the restriction of $E$ to a divisor in $|mH|$ is semistable (or stable, for $E$ a stable vector bundle) if $m \gg 0$ (cf. \cite{MR1}, and \cite{MR2}). A detailled overview on restriction theorems is given in \S7 of the book \cite{HL} of Huybrechts and Lehn. First we discuss the case of rank two bundles on a surface $X$. Theorem \ref{res1} shows that for a semistable $X$-vector bundle $E$ of rank two there exists an integer $m$ such that the restriction of $E$ to a general curve in the linear system $|mH|$ is semistable. Using this result we provide a boundedness result for semistable rank two bundles (proposition \ref{bound}). For surfaces defined over ${\mathbb C}$ a semistable bundle $E$ cannot have positive discriminant $\Delta(E)$ (Bogomolov's theorem, cf. \cite{Bog}). In positive characteristic this does not hold. No more than the Kodaira vanishing holds for positive characteristic (see \cite{Ray}). It is remarkable that semistable bundles which contradict Bogomolov's theorem behave well with respect to restrictions. Applying our restriction result, we obtain a weak form of Bogomolov's theorem (corollary \ref{res3}, cf.~also \cite{Meg} for vector bundles of arbitrary rank), and a weak form of Kodaira's vanishing theorem in arbitrary characteristic (corollary \ref{res4}). The reader familiar with the vector bundle techniques presented in Lazarsfeld's lectures, \cite{Laz} will deduce Reider type theorems for surfaces in arbitrary characteristic. If the $X$-bundle $E$ is semistable but not stable, then it is easy to see that the restriction of $E$ to a general curve in $|H|$ is semistable but not stable. Conversely, we may ask whether stable bundles do restrict to stable objects. Theorem \ref{res5} gives an affirmative answer to this question. The proof follows an idea of Bogomolov (see \cite{Bog1} and \cite{HL} theorem 7.3.5) using the weak Bogomolov inequality deduced before. Finally, we present with theorem \ref{res6} the higher dimensional version of theorem \ref{res1}. It turns out that its proof is easier than the proof in the surface case. The main reason for this simplification is the fact that two general hyperplanes in a linear system intersect in a irreducible subscheme. All these results should generalize to vector bundles of arbitrary rank. To prove the corresponding results it seems necessary to consider the complete Harder-Narasimhan filtration. {\vspace{0.5em} } {\bf Acknowledgement:} The author would like to thank D.~Huybrechts for many helpful remarks. \section{Preliminaries} Let $X,H$ be a polarized projective variety. We will identify line bundles on $X$ and their corresponding Cartier divisor classes. Moreover, to any class in the Chow group ${\rm CH}^{\dim(X)}(X)$ of codimension $\dim(X)$ cycles is assigned via evaluation on the fundamental class $[X]$ of $X$ its characteristic number. This allows us to interpret $c_i(E).H^{\dim(X)-i}$ as integers. For a coherent $X$-sheaf $E$, we write $E(n)$ instead of $E \otimes {\cal O}_X(H)^{\otimes n}$. The Hilbert polynomial $\chi_E:n \mapsto \chi(E(n))$ can be written in the following form $$\chi_E(n) = a_0(E)\bino{n +\dim X}{\dim X}+ a_1(E)\bino{n +\dim X -1}{\dim X -1} + \ldots \,.$$ If $H$ is sufficiently general in the linear system $|H|$ (i.e., ${\rm Tor}^{{\cal O}_X}_1(E,{\cal O}_H)=0$), we have $a_i(E)=a_i(E|_H)$, for all integers $i < \dim X$. We define the $H$-slope $\mu_H(E)$ of $E$ to be the quotient $a_1(E)/a_0(E)$. A coherent sheaf $E$ is called Mumford semistable (resp. stable) with respect to $H$, if $E$ is torsion free, and for all proper subsheaves $F \subset E$ the inequality $\mu_H(F) \leq \mu_H(E)$ (resp. $\mu_H(F) < \mu_H(E)$) holds true. This kind of stability is also named slope stability, weak stability, or $\mu$-stability. For brevity we simply write stability because we only use this stability concept. We will frequently use the following facts on stable and semistable coherent sheaves: \begin{enumerate} \item If $E$ and $F$ are semistable with $\mu_H(E)>\mu_H(F)$, then the group ${\rm Hom}(E,F)$ vanishes. \item For a stable bundle $E$ on a variety defined over an algebraically closed field the endomorphism group ${\rm End}(E)$ consists of the scalar multiples of the identity. \item If a rank two vector bundle $E$ is not semistable, then there exists a unique maximal subsheaf $E_1 \subset E$ of rank one which is the maximal destabilizing subsheaf. Or equivalently, there exists a unique destabilizing quotient $E \to Q$. The flag $0 \subset E_1 \subset E$ is the Harder-Narasimhan filtration of $E$. \end{enumerate} See, for example, the article \cite{Sha} of Shatz. An important invariant of a vector bundle is its discriminant. Let $E$ be a vector bundle of rank $r$ with Chern roots $\{\alpha_i \}_{i=0,\ldots ,r}$. As the name discriminant suggests we define the discriminant $\Delta(E)$ of the vector bundle $E$ by $\Delta(E)= \sum_{i<j}(\alpha_i-\alpha_j)^2$. Obviously $\Delta(E)$ can be expressed in terms of the Chern classes of $E$, namely $\Delta(E)=(r-1)c_1(E)^2-2rc_2(E)$. (Unfortunately, there are different definitions of $\Delta(E)$ in literature, differing by a sign or a constant.) In particular, we have $\Delta(E)=c_1(E)^2-4c_2(E)$, for a rank two vector bundle. A rank two vector bundle $E$ on a surface $X$ is named Bogomolov unstable if there exists an injection $A \rarpa{\iota} E$ of coherent sheaves where $A$ is an $X$-line bundle, the cokernel of $\iota$ is torsion free, and the inequalities $(2A-c_1(E))^2>0$, and $(2A-c_1(E)).H>0$ are satisfied for a polarization $H$ of $X$. For a rational number $q$, let $\lceil q \rceil$ be the least integer not smaller than $q$, $\lfloor q \rfloor$ the largest integer smaller or equal to $q$, and $[q]_+$ the maximum of $q$ and $0$. \section{Rank two bundles on surfaces} \subsection{The semistable restriction theorem} \begin{theorem}\label{res1} Let $X$ be a smooth surface over an algebraically closed field with a very ample line bundle ${\cal O}_X(1)={\cal O}_X(H)$. For an $X$-vector bundle $E$ of rank two which is semistable with respect to ${\cal O}_X(1)$ the following holds: \begin{enumerate} \item If $\Delta(E) \geq 0$, then the restriction of $E$ to a general curve of the linear system $|H|$ is semistable; \item For $\Delta(E)<0$ and any integer $l$ with $l \geq \log_2 \left(\sqrt{\frac{-\Delta(E)}{H^2}}+1 \right)$ the restriction of $E$ to a general curve in $|2^lH|$ is semistable. \end{enumerate} \end{theorem} {\bf Proof: } We divide the proof in several steps. First we outline its strategy: \begin{itemize} \item We define the objects which are needed. In particular, we define the non negative integer $A(m)$ which measures the instability of the restriction of $E$ to a general curve of the linear system $|mH|$ (step 1-3); \item We next (step 4-7) compute an upper bound for $A(1)$. This bound depends only on the Chern number $\Delta(E)$ and $H^2$; \item After that, we give an upper bound for $A(2m)$ in terms of $A(m)$ (step 8-12); \item Finally, we combine both estimates to conclude the theorem (step 13). \end{itemize} \step{1} Let $m$ be a positive integer. For the linear system $|mH|$ we denote by ${\cal C}_{|mH|}$ the universal curve over $|mH|$. We have the morphisms $$\xymatrix{|mH| & {\cal C}_{|mH|} \ar[r]^q \ar[l]_p & X} .$$ The space $|mH|$ is isomorphic to ${\mathbb P}^{h^0(m)-1}$ where $h^0(m)$ denotes the dimension of $H^0(X,{\cal O}_X(m))$. Since $|mH|$ is base point free, $q$ is a ${\mathbb P}^{h^0(n)-2}$-bundle. We denote by $g_m$ the genus of a smooth curve of $|mH|$. A curve $C\subset X$ rationally equivalent to $mH$ corresponds to a geometric point in $|mH|$ which we denote by $[C]$. \step{2} For all integers $a$ with $2a<c_1(E).(mH)$ we consider the Quot scheme $${\rm Quot}_{m,a}:={\rm Quot}^{P_a}_{q^*E / {\cal C}_{|mH|} / |mH|}$$ of $p$-flat quotients of $q^*E$ with Hilbert polynomial $P_a(k)=(mH^2)k+a+1-g_m$ with respect to the very ample line bundle $q^*{\cal O}_X(1)$ see \cite{Gro}. For $a<g-1-h^1(E)-h^2(E(-mH))$ the scheme ${\rm Quot}_{m,a}$ is the empty scheme. To see this, we remark that for any curve $[C] \in |mH|$ we have the inequality $h^1(E|_C) \leq h^1(E)+h^2(E(-mH))$. Hence, any quotient of the restriction $E|_C$ has at least Euler characteristic $-(h^1(E)+h^2(E(-mH)))$. From that bound, using the Riemann-Roch theorem for curves, we obtain the above bound for the degree of quotients of $E|_C$. Thus, we are considering only a finite number of Quot schemes. \step{3} Since the schemes ${\rm Quot}_{m,a}$ are projective over $|mH|$, their images dominate $|mH|$ if and only if at least one ${\rm Quot}_{m,a}$ is surjective over $|mH|$. If they do not cover $|mH|$, then the restriction of $q^*E$ to the general fiber of $p$ is semistable. In this case we define the number $A(m)$ to be zero. Otherwise we define $A(m)$ by $$A(m):= \max \left\{c_1(E).(mH)-2a \left| \begin{array}{ll} 2a<c_1(E).(mH) &\mbox{ and} \\ {\rm Quot}_{m,a} \to |mH| &\mbox{ is surjective.} \end{array} \right. \right\} \,.$$ By definition $A(m)$ measures how far the restriction of $E$ to the general curve of $|mH|$ is from being semistable. The projectivity of the Quot schemes implies that if $A(m)>0$, then the restriction of $E$ to any curve $[C]\in |mH|$ has a quotient $Q$ with Hilbert polynomial $$\chi(Q(k))=(mH^2)k+\frac{1}{2}(c_1(E).(mH)-A(m))+1-g_m \, .$$ We will apply this specialization property (see also \cite{Sha}) in the sequel to reducible curves $[C] \in |2mH|$ with $C=C' \cup C''$ where $C'$ and $C''$ are smooth curves in $|mH|$, to bound $A(2m)$ in terms of $A(m)$. \step{4} From now on we assume that $A(m)$ is positive. We set $b(m):=\frac{1}{2}(c_1(E).(mH)-A(m))$. By definition of $A(m)$ the subset $$Y := \bigcup_{a<b(m)} {\rm im}( {\rm Quot}_{m,a} \to |mH|) $$ is a proper closed subset of $|mH|$. For all points of the open subset $U'_m= |mH| \setminus Y$ the restriction of $E$ the corresponding hyperplane has a minimal destabilizing quotient of degree $b(m)$. The minimality of the destabilizing quotient implies its uniqueness. Therefore the restriction ${\rm Quot}_{m,b(m)} \times_{|mH|} U'_m$ of the Quot scheme ${\rm Quot}_{m,b(m)}$ to $U'_m$ gives a bijection of geometric points of ${\rm Quot}_{U'_m}:={\rm Quot}_{m,b(m)} \times_{|mH|} U'_m$ and $U'_m$. Thus, $p_{U'_m}:{\rm Quot}_{U'_m} \to U'_m$ is an isomorphism or completely inseparable. If $[ \xymatrix{E|_C \ar@{->>}[r]^\alpha & F}]$ is a geometric point of ${\rm Quot}_{U'_m}$, then we have ${\rm Hom}(\ker (\alpha), F)=0$. Therefore the relative tangent bundle of $p_{U'_m}$ vanishes. Eventually, we conclude that $p_{U'_m}$ is an isomorphism. \step{5} If $[C]$ is a smooth curve in $U'_m$, then the minimal quotient of degree $b(m)$ has to be a quotient line bundle. Therefore, by considering the open subset $U_m$ of $U'_m$ parametrizing smooth curves, we obtain the following situation: $$\xymatrix{ U_m & {\cal C}_{U_m} \ar[r]^q \ar[l]_p & X}$$ and a destabilizing quotient line bundle $L$ of $q^*E$ which is $p$-flat. Furthermore the degree of $L$ on all fibers of $p$ is $b(m)$. The surjection $q^*E \to L$ defines the following diagram: $$\xymatrix{{\cal C}_{U_m} \ar[d]_p \ar[r]^\xi \ar[dr]^q & {\mathbb P}(E) \ar[d]\\ U_m & X}$$ For a curve $C \subset X$ which is parametrized by $U_m$ we call $\xi(p^{-1}[C])$ its canonical $m$-lifting. \step{6} Now we take two smooth curves $H_1$ and $H_2$ in $X$ which meet transversally and which are contained in $U_1 \subset |H|$. The pencil spanned by these curves defines a rational map $\xymatrix{{\mathbb P}^1_{\mbox{ }} \ar@{-->}[r] & U_1 \ar[r]^-{\sim} & {\mbox{ }}} \!\!{\rm Quot}_{1,b(1)} \times U_1$. Since ${\rm Quot}_{1,b(1)}$ is projective, we obtain a morphism ${\mathbb P}^1 \to {\rm Quot}_{1,b(1)}$. This corresponds to a flat family of degree $b(1)$ quotients for all restrictions of $E$ to curves of the pencil. To be precise we have the following situation: ${\mathbb P}^1 \larpa{p} \tilde X \rarpa{q} X$ where $\tilde X$ denotes the blow up of $X$ in the points of $H_1 \cap H_2$, and a destabilizing $p$-flat quotient $q^*E \to Q$ which for all $p$-fibers is of degree $b(1)$. Over ${\mathbb P}^1 \cap U_1$ the quotient $Q$ is a line bundle (see step 5). $Q$ is flat and its restriction to most fibers is torsion free. Hence, $Q$ itself is torsion free of projective dimension at most one. The kernel $K$ of $q^*E \to Q$ is a line bundle on $\tilde X$ which is isomorphic to $q^*L+\sum_{i=1}^{H^2} a_iE_i$ where the $\{ E_i \}_{i=1,\ldots,H^2}$ are the exceptional fibers of the blow up $q$. Restricting the exact sequence $0 \to K \to q^*E \to Q \to 0$ to $E_i$ we see that the integer $a_i$ is at least zero. \step{7} It results that $Q$ is of the form ${\cal O}(q^*c_1(E)-q^*L-\sum a_i E_i) \otimes {\cal J}_Z$ where ${\cal J}_Z$ denotes the ideal sheaf of a closed subscheme $Z$ of $\tilde X$ of finite length. Semistability of $E$ implies: \begin{eqnarray} H.D& \leq& 0 \qquad \end{eqnarray} where $D=(2L-c_1(E)) $. Chern class computation gives: $$\begin{array}{rcl} c_2(E) & = & c_1(K)c_1(Q) +{\rm length}(Z) \\ & = & L.c_1(E)-L^2+\sum a_i^2 +{\rm length}(Z) \\ \end{array}$$ It follows \begin{eqnarray} \sum a_i^2 &\leq & c_2(E)+L^2-L.c_1(E) = c_2(E)-\frac{c_1(E)^2}{4} +\frac{D^2}{4} \end{eqnarray} The discrepancy to semistability is the number $A(1)$ \begin{eqnarray} A(1) & = & 2c_1(K).\left( H -\sum E_i\right) - c_1(E).H = D.H + 2\sum a_i \end{eqnarray} Now we use the inequality $$\sum_{i=1}^{H^2} a_i \leq \sqrt{H^2\sum_{i=1}^{H^2} a_i^2}$$ and inequalities (2) and (3) to deduce: $$\begin{array}{rcl} A(1) & \leq & D.H +2 \sqrt{H^2\cdot(c_2-\frac{c_1(E)^2}{4})+H^2 \cdot \frac{D^2}{4}} \\ & \leq & D.H +\sqrt{ -H^2\Delta(E) + H^2 \cdot D^2} \end{array}$$ By the Hodge index theorem $H^2 \cdot D^2 \leq (D.H)^2$. Thus, we eventually obtain: $$A(1) \leq D.H+\sqrt{(D.H)^2-H^2\Delta(E) } \, .$$ The basic properties of the function $x \mapsto x+\sqrt{x^2-H^2\Delta(E)}$ together with (1) give a bound for $A(1)$: If $\Delta(E) \geq 0$, then $A(1)=0$. For $\Delta(E) < 0$ we have the upper bound $A(1) < \sqrt{-H^2\Delta(E)}$. \step{8} Take a reducible curve $C=C' \cup C''$ where $C',C'' \in |mH|$ are smooth curves which intersect transversally. The singular divisor of $C$ consisting of $m^2 \cdot H^2$ nodes we denote by $D$. Let $E|_C \to Q$ be a torsion free quotient of $E$ with Hilbert polynomial $$\chi_Q(k) = \chi( Q \otimes {\cal O}_X(k))=2m \cdot H^2\cdot k+b+1-g_{2m}$$ where $g_{2m}$ denotes the arithmetic genus of $C$. Torsion free means: $Q$ does not contain a subsheaf of dimension zero. By the Mayer-Vietoris exact sequence $$0 \to {{\cal O}_C}\to {{\cal O}_{C'} \oplus {\cal O}_{C''}} \to {{\cal O}_D} \to 0$$ we see that $g_{2m}=2g_m+ m^2 \cdot H^2-1$. Furthermore, we obtain from this exact sequence the following diagram with exact rows and surjective columns $$\xymatrix{0 \ar[r] & E|_C \ar[d] \ar[r] & E|_{C'} \oplus E|_{C''} \ar[d] \ar[r] & E|_D \ar[d] \ar[r] & 0 \\ {\rm Tor}_1^{{\cal O}_C}(Q,{\cal O}_D) \ar[r] & Q \ar[r] & Q|_{C'} \oplus Q|_{C''} \ar[r] & Q|_D \ar[r] & 0}$$ Since ${\rm Tor}_1^{{\cal O}_C}(Q,{\cal O}_D)$ is concentrated in $D$, and $Q$ was assumed to be torsion free, the image of ${\rm Tor}_1^{{\cal O}_C}(Q,{\cal O}_D)$ in $Q$ is zero. Therefore, the equality \begin{eqnarray}\label{chiq} \chi(Q)+{\rm length}(Q|_D)& =&\chi( Q|_{C'}) +\chi( Q|_{C''}) \end{eqnarray} holds true. There are three cases for the ranks of $Q|_{C'}$ and $Q|_{C''}$. The pair $({\rm rk}(Q|_{C'}),{\rm rk}(Q|_{C''}))$ has to be $(1,1)$, $(2,0)$, or $(0,2)$. \step{9} We next show that if the ranks of $Q|_{C'}$ and $Q|_{C''}$ do not coincide, then the quotient $Q$ is not destabilizing. Assume that $Q|_{C'}$ has rank two and $Q|_{C''}$ is torsion. It results that $Q|_{C'}$ is isomorphic to $E|_{C'}$, and $Q|_{C''}$ is isomorphic to $E|_D$. Hence, $Q$ is isomorphic to $E|_{C'}$. Therefore we find $$\chi(Q(k))=\chi(E(k))-\chi(E(k-m))=(2mH^2)k+\chi(E)-\chi(E(-m)) \,.$$ Analogously we compute the Euler characteristic of $E|_C$ to be $$\chi(E|_C(k))=(4m \cdot H^2)k+\chi(E)-\chi(E(-2m)) \,.$$ In order to prove that $Q$ is not destabilizing we must show that the inequality $$\frac{\chi(E)-\chi(E(-m))}{2mH^2} > \frac{\chi(E)-\chi(E(-2m))}{4mH^2}$$ holds. This inequality is equivalent to $$2(\chi(E)-\chi(E(-m))) > \chi(E)-\chi(E(-2m)) \,.$$ The last inequality holds because the function $k \mapsto \chi(E(k))$ is strictly convex by the Riemann-Roch theorem for surfaces. \step{10} {\bf (General intersection lemma) \/}{\em Let $C''$ be an irreducible curve in $X$ with a lifting $\tilde C''$ to ${\mathbb P}(E)$. For a general curve $[C']$ in $U_m$ its canonical $m$-lifting $\tilde {C'}$ in ${\mathbb P}(E)$ intersects $\tilde C''$ in zero or $C'.C''$ points.} {\bf Proof: } For brevity, we write $U$ instead of $U_m$. We consider the following situation: $$\xymatrix{& {\cal C}_U \times_X C'' \ar[r] \ar[d] \ar[ddr]^(0.3)\psi & C'' \ar[d]^\iota \\ {\cal C}_U \times_{{\mathbb P}(E)} \tilde C'' \ar[ur]^\varphi \ar[r] \ar[d] & {\cal C}_U \ar[r]|\hole \ar[d] \ar[dr]|\hole & X \\ \tilde C'' \ar[r]^{\tilde \iota} & {\mathbb P}(E) \ar[ur]|(0.65)\hole & U}$$ Since $C'' \to X$ and ${\cal C}_U \to U$ are projective morphisms, so is $\psi$. The same way, we see that the composition morphism $\psi \circ \varphi$ is projective. For a geometric point $[C'] \in U$ the fiber of $\psi$ over $[C']$ is the intersection $C' \cap C''$. Thus, $\psi$ is of relative dimension zero. Analogously we identify the fiber of $\psi \circ \varphi$ with the intersection of the liftings. By construction ${\cal C}_U$ is an open subset in a ${\mathbb P}^n$-bundle over $X$. We conclude the irreducibility of ${\cal C}_U \times_X C''$. Thus we see, that if $\varphi$ is a dominant morphism, then the canonical lifting $\tilde C'$ of a general curve $C'$ intersects $\tilde C''$ in $C'.C''$ points. If the morphism $\varphi$ is not dominant, then the canonical lifting of a general curve $C'$ is disjoint from $\tilde C''$. \hfill{$\Box$ (of step 10)} \step{11} Let $[C'']$ be a point in $U_m$, and $\tilde C''$ be its canonical lifting to ${\mathbb P}(E)$. Let us assume that the lifting $\tilde C'$ of a general curve $[C'] \in U_m$ intersects $\tilde C''$ in $C'.C''$ points. If we consider the pencil spanned by $C''$ and $C'$, then we obtain (see step 6) a family $Q$ over this pencil with all $a_i$ equal to zero. Indeed, if one $a_i$ is positive, then the $m$-lifting of general curve contained in the pencil spanned by $C'$ and $C''$ does not intersect $\tilde C''$ in the point $P_i$. This would imply (see (1) and (3) of step 7) that $A(m)=0$. By the above lemma we can assume that $[C'],[C''] \in U_m$ are two smooth curves whose canonical liftings are disjoint. We consider now for $C=C' \cup C''$ a minimal quotient $Q$ of $E|_C$ having rank one on $C'$ and $C''$. We call a point $P \in D=C' \cap C''$ a point of discord if the dimension of $Q \otimes k(P)$ is two. Let $M$ be the number of points of discord. It is obvious that the maximal torsion subsheaf of $Q|_{C'}$ is concentrated in the points of discord and has length $M$. The quotient of $Q_{C'}$ modulo its torsion is denoted by $Q'$, and analogous we have the $C''$-line bundle $Q''$. We obtain from (\ref{chiq}) that \begin{eqnarray}\label{chiq2} \chi(Q) & = & \chi(Q') + \chi(Q'') + M -m^2H^2 \end{eqnarray} \step{12} {\em The inequality $A(2m) \leq [2A(m)-2m^2H^2]_+$ holds.} {\bf Proof: } We consider the unique destabilizing quotient $L'$ of $E|_{C'}$. The kernel of $ E|_{C'} \to L$ we denote by $F'$. We now consider the composition $\beta':F' \to E|_{C'} \to Q'$. \case{1} The morphism $\beta'$ (or $\beta''$) is not trivial. If $\beta'$ is not trivial, then it follows that $\deg(Q') > \deg(F')$ and $\chi(Q'(k)) \geq \chi(F'(k)) = \chi(F''(k))$. We obtain from (\ref{chiq2}) that $$\begin{array}{rcl} \chi(Q(k)) & \geq & \chi(F''(k))+ \chi(Q''(k))-m^2H^2 \\ & \geq & \chi(F''(k))+ \chi(L''(k))-m^2H^2 \\ & = & \chi(E|_{C''}(k))-m^2H^2 \\ & = & \frac{1}{2}(\chi(E|_C(k))) \,. \end{array}$$ Thus, in this case the quotient $Q$ is not destabilizing. \case{2} If $\beta'$ is trivial, then we obtain $F' \subset \ker(E_{C'} \to Q')$. However $F'$ is the rank one subbundle of $E_{C'}$ of maximal degree. Hence we have $Q' \cong L'$. Since the canonical liftings of $C'$ and $C''$ do not intersect, we must have $M=m^2H^2$. The equality (\ref{chiq2}) consequently yields: $$\begin{array}{rcl} \chi(Q(k)) & = & 2\chi(L'(k))\\ &=&\chi(E|_C'(k))-A(m) \\ &=& \frac{1}{2}\chi(E|_C(k))+(m^2H^2-A(m)) \end{array}$$ This gives the asserted inequality for $A(2m)$. To complete the proof of step 12 we just remark that a similar computation shows that $A(m)=0$ implies $A(2m)=0$. \hfill{$\Box$ (of step 12)} \step{13} Using induction on $l$, we obtain from the inequality of step 12 $$A(2^l) \leq [ 2^lA(1)-2^l(2^l-1)H^2]_+ \,.$$ Combining this with the upper bound for $A(1)$ computed in step 7 the theorem follows. { \hfill $\Box$} {\vspace{0.5em} } {\bf Remark 1:} The theorem still holds true for a semistable coherent $X$-sheaf $E$ of rank two. Indeed, consider the embedding $E \rightarrow E^{\lor \lor}$ of $E$ into its double dual. In this way, we obtain the semistable vector bundle $E^{\lor \lor}$ which obviously satisfies $\Delta(E^{\lor \lor}) \geq \Delta(E)$. Hence, the theorem applies. {\vspace{0.5em} } {\bf Remark 2:} We can extend the theorem to projective surfaces $X$ with isolated singularities. Reviewing the proof, we see that it is enough to have smooth curves in the linear systems $|mH|$. By the same argument, we see that the theorem holds true if we require ${\cal O}_X(1)$ to be base point free. {\vspace{0.5em} } {\bf Remark 3:} If we know the ideal $\{ L.H \}_{L \in {\rm Pic}(X)} \subset {\mathbb Z}$ of intersections with $H$, then we can sharpen the inequality of step 7. To illustrate this, let us assume that the Picard group of $X$ is generated by ${\cal O}_X(1)$. Furthermore, suppose that $E$ is a semistable $X$-vector bundle. We have $\det(E)=nH$. If $\Delta(E)<0$, then we can improve the bound for $A(1)$ of step 7 by $$A(1) \leq \left\{ \begin{array}{ll} -2H^2+\sqrt{4(H^2)^2-H^2\Delta(E)} & \mbox{ for } n \mbox{ even;} \\ -H^2+\sqrt{(H^2)^2-H^2\Delta(E)} & \mbox{ for } n \mbox{ odd.} \\ \end{array} \right.$$ \subsection{A boundedness result} Let $X$, ${\cal O}_X(1)$ be as before. Furthermore, let $E$ be a semistable $X$-vector bundle of rank two. We next give a bound $M$ depending only on the characteristic numbers $c_2(E)$ and $c_1(E).H$ such $E(M)$ becomes globally generated. We use Mumford's concept of $m$-regularity: A coherent sheaf $E$ on a polarized variety $X$ with very ample line bundle ${\cal O}_X(1)$ is called $m$-regular, if $h^i(E(m-i))=0$, for all $i>0$. The following lemma (cf. \S 14 in \cite{Mum}) resumes properties of $m$-regular sheaves. \begin{lemma}\label{bound1} Let $X$ be a projective variety with a very ample line bundle ${\cal O}_X(1)={\cal O}_X(H)$, and $E$ be a coherent $X$-sheaf. If $E$ is $m$-regular, then it is globally generated, and $E(m+k)$-regular for all $k \geq 0$. Let $D \in |H|$ be a divisor such that the sequence $0 \to E(-1) \to E \to E|_D \to 0$ is exact. If $E|_D$ is $m$-regular, then $E$ is $(m+h^1(E(m-1)))$-regular. { \hfill $\Box$} \end{lemma} This lemma outlines our strategy. We first show that for a suitable curve $C \in |H|$ and an integer $m_1$ the restriction $E_C=E|_C$ is $(m_1+1)$-regular. In order to obtain the boundedness result, we then compute an upper bound for $h^1(E(m_1))$. \begin{lemma}\label{bound2} Let $E_C$ be a rank two vector bundle on a smooth curve $C$ of genus $g$ defined over an algebraically closed field. We define the number $A$ to be zero if $E_C$ is semistable. Otherwise we set $$A = \max \{ \deg(E_C)-2\deg(Q) \, | \, E_C \to Q \mbox{ is surjective, and }{\rm rk}(Q)=1 \} \, .$$ \begin{enumerate} \item[(1)] If $L$ is a $C$-line bundle with $\deg(L)>\frac{A-\deg(E_C)}{2}+2g-2$,\\ then $H^1(C,E_C \otimes L)=0$; \item[(2)] For any $C$-line bundle $L$ the inequality\\ $h^0(E_C\otimes L) \leq 2 \left[ 1+\deg(L)-\frac{\deg(E_C)-A}{2} \right]_+$ holds true. \end{enumerate} \end{lemma} {\bf Proof: } (1) If $h^1(E_C \otimes L)>0$, then there exists, by Serre duality, a non trivial homomorphism $\varphi:E \to \omega_C \otimes L^{-1}$. (Here $\omega_C$ denotes the dualizing sheaf of $C$.) Thus, the image of $\varphi$ is a rank one quotient of degree at most $2g-2-\deg(L)$. By the very definition of the number $A$ we obtain $\deg(L) \leq \frac{A-\deg(E_C)}{2}+2g-2$. (2) Analogously, we see that for $\deg(L)<\frac{\deg(E_C)-A}{2}$ there are no global sections of $E_C \otimes L$. Thus, the assertion holds for all line bundles $L$ of degree less than $\frac{\deg(E_C)-A}{2}$. Let $P \in C$ be a geometric point of $C$. Then from the exact sequence $0 \to E_C \otimes L(-P) \to E_C \otimes L \to E_C|_P \to 0$ we obtain $h^0(E_C \otimes L) \leq 2+ h^0(E_C \otimes L(-P))$ which proves the second statement. { \hfill $\Box$} {\vspace{0.5em} } We now take a smooth curve $C$ of genus $g$ in the linear system $|H|$ such that for the restriction $E_C$ the number $A$ of the above lemma is at most $\sqrt{ [-H^2 \cdot \Delta(E)]_+}$. We have seen in step 7 of the proof of theorem \ref{res1} that this is possible. The adjunction formula gives $2g-2=H.(H+K_X)$. Obviously, the degree of the $C$-line bundle ${\cal O}_X(mH)|_C$ is $mH^2$. Thus, setting $$m_1:= \left\lfloor \frac{1}{H^2} \left( \frac{\sqrt{ [-H^2 \cdot \Delta(E)]_+}-c_1(E).H}{2}+H.(H+K_X) \right) \right\rfloor +1$$ we obtain by lemma \ref{bound2} that $E_C$ is $(m_1+1)$-regular. The semistability of $E$ implies that $h^0(E(m_2-1))=0$, for $m_2:=\left\lceil \frac{-H.c_1(E)}{2H^2} \right\rceil$. Applying the inequality $h^0(E(m))<h^0(E(m-1))+h^0(E_C(m))$ obtained from the long exact cohomology sequence yields $$h^0(E(m_1)) \leq m_3 := 2\sum_{m=m_2}^{m_1} \left[ 1+ mH^2-\frac{c_1(E).H-\sqrt{ [-H^2 \cdot \Delta(E)]_+}}{2} \right]_+ \, .$$ Since $h^2(E(m_1))=0$ we deduce that $h^1(E(m_1)) \leq m_3-\chi(E(m_1))$. Setting $m_4:=m_1+m_3-\chi(E(m_1))$, we obtain by lemma \ref{bound1}: \begin{proposition}\label{bound} Let $X$ be a smooth projective surface over an algebraically closed field, and ${\cal O}_X(1)$ a very ample line bundle on $X$. Furthermore, let $E$ be a rank two $X$-bundle which is semistable with respect to ${\cal O}_X(1)$. Then for $m \geq m_4$ we have that $E(m)$ is globally generated. The number $m_4$ defined above depends only on the characteristic numbers of $E$. { \hfill $\Box$} \end{proposition} It follows that any semistable sheaf $E$ of rank two with given $c_1(E).H$, $c_1(E).K_X$, and $c_2(E)$ is a quotient of ${\cal O}_X(-m_4)^{\oplus \chi_E(m_4)}$. Considering the Quot scheme ${\rm Quot}^{\chi_E}_{{\cal O}_X(-m_4)^{\oplus \chi_E(m_4)}/X}$ together with its natural ${\rm SL}_{\chi_E(m_4)}$-action we obtain (see \cite{GIT}) the coarse moduli space of semistable coherent sheaves on $X$ with Hilbert polynomial $\chi_E$. This proves the next corollary. \begin{corollary} There exists a projective coarse moduli space for semistable coherent sheaves of rank two with fixed characteristic numbers on a smooth projective surface. \end{corollary} \subsection{Further applications} \begin{proposition}\label{res2} Let $X$ be a smooth projective surface over an algebraically closed field with a very ample line bundle ${\cal O}_X(1)={\cal O}_X(H)$. If $E$ is a $X$-bundle, which is stable with respect to ${\cal O}_X(1)$ and of rank 2, then the inequality $$\Delta(E) \leq \left\{ \begin{array}{ll} 1-4\chi({\cal O}_X) & \mbox{if } K_X.H<0 \,; \\ 2-4\chi({\cal O}_X) & \mbox{if } K_X.H=0 \,; \\ \left[ 6-4\chi({\cal O}_X) +4\cdot \left\lceil \frac{K_X.H}{H^2} \right\rceil K_X.H \right]_+ & \mbox{if } K_X.H>0 \,. \\ \end{array} \right.$$ holds. \end{proposition} {\bf Proof: } We compute, using the Riemann-Roch theorem for surfaces that $$\chi(E \otimes E^\lor) = \Delta(E)+4\chi({\cal O}_X)\,. $$ The stability of $E$ implies that $H^0(E \otimes E^\lor) = {\rm Hom}(E,E)$ is of dimension one. Now we want to bound $h^2:=h^2(E \otimes E^\lor)$. By Serre duality, $h^2$ equals the dimension of ${\rm Hom}(E,E(K_X))$ where $K_X$ denotes the canonical class on $X$. Thus, we obtain for $K_X.H \leq 0$ $$h^2(E \otimes E^\lor) \leq \left\{ \begin{array}{ll} 0 & \mbox{if } K_X.H < 0 \, ;\\ 1 & \mbox{if } K_X.H = 0 \, .\\ \end{array} \right.$$ If $K_X.H >0$, we set $m=\lceil \frac{K_X.H}{H^2}\rceil$ and consider a smooth curve $C$ in the linear system $|mH|$. If $\Delta(E) \geq 0$, then by theorem \ref{res1}, we may assume that the restriction $E|_C$ is semistable. Thus, ${\rm Hom}(E|_C,E|_C)$ is at most of dimension 4. By induction we see that for a $C$-line bundle $L$ of degree $d$ we can bound the dimension of ${\rm Hom}(E|_C,E|_C \otimes L)$, by $4+4\cdot d$. By definition of $m$, we have $(K_X-mH).H \leq 0$. Thus, we can bound the dimension of ${\rm Hom}(E,E(K_X-C))$ by one. From the exact sequence $$0 \to {\rm Hom}(E,E(K_X-C)) \to {\rm Hom}(E,E(K_X)) \to {\rm Hom}(E,E(K_X)|_C) $$ we obtain the estimate $$h^2(E \otimes E^\lor) \leq 5+4 \cdot \left\lceil \frac{K_X.H}{H^2}\right\rceil K_X.H \, .$$ Applying the obvious inequality $\chi(E \otimes E^\lor) \leq h^0(E \otimes E^\lor) + h^2(E \otimes E^\lor)$ we obtain the estimation of the proposition. { \hfill $\Box$} \begin{corollary}\label{res3} {\bf (Weak Bogomolov inequality)} Let $X,H$ be a very ample polarized smooth surface over an algebraically closed field. Let $E$ be a rank 2 vector bundle on $X$ satisfying $$\Delta(E) > \left\{ \begin{array}{ll} [1-4\chi({\cal O}_X)]_+ & \mbox{if } K_X.H<0 \,; \\ \left[ 2-4\chi({\cal O}_X)\right]_+ & \mbox{if } K_X.H=0 \,; \\ \left[ 6-4\chi({\cal O}_X) +4\cdot \left\lceil \frac{K_X.H}{H^2} \right\rceil K_X.H \right]_+ & \mbox{if } K_X.H>0 \,. \\ \end{array} \right.$$ Then $E$ is Bogomolov unstable. \end{corollary} \setcounter{equation}{0} {\bf Proof: } By proposition \ref{res2} $E$ cannot be stable with respect to the given polarization $H$. Thus, we have a short exact sequence $$0 \to A \to E \to {\cal J}_Z(c_1(E)-A) \to 0 \, ,$$ where $Z\subset X$ is a closed subscheme of codimension 2. Since $A$ is destabilizing we have $(c_1(E)-2A).H \leq 0$. Using the exact sequence to compute $c_2(E)$ yields $$(c_1(E)-2A)^2 = \Delta(E)+4\cdot {\rm length}(Z) > 0 \,.$$ Thus, the Hodge index theorem implies $(c_1(E)-2A).H<0$. Consequently, $E$ is not semistable. Now the half c\^one in the N\'eron-Severi group ${\rm NS}(X)$ defined by positive self intersection and negative intersection with an ample class $H$ does not depend on $H$. { \hfill $\Box$} {\vspace{0.5em} } {\bf Remark:} G. Megyesi proves that for a vector bundle $E$ of arbitrary rank on a smooth surface defined over a field of characteristic $p>0$ with $\Delta(E)>0$ the pullback $(F^n)^*E$ of $E$ by a large power $n$ of the absolute Frobenius $F$ is Bogomolov unstable (see \cite{Meg}). Corollary \ref{res3} gives an effective bound for $n$, for a vector bundle $E$ of rank two. \begin{corollary}\label{res4} {\bf (Weak Kodaira vanishing)} Let $X$ be a smooth projective surface defined over an algebraically closed field with very ample line bundle ${\cal O}_X(H)$. Let $L$ be a nef $X$-line bundle such that $$L^2> \left\{ \begin{array}{ll} \left[ 1-4\chi({\cal O}_X) \right]_+ & \mbox{if } K_X.H<0 \,; \\ \left[ 2-4\chi({\cal O}_X)\right]_+ & \mbox{if } K_X.H=0 \,; \\ \left[ 6-4\chi({\cal O}_X) +4\cdot \left\lceil \frac{K_X.H}{H^2} \right\rceil K_X.H \right]_+ & \mbox{if } K_X.H>0 \,. \\ \end{array} \right.$$ Then the first cohomology group $H^1(X,L^{-1})$ vanishes. \end{corollary} {\bf Proof: } \setcounter{equation}{0} We take an extension $E$ of ${\cal O}_X$ by $L^{-1}$. Since $H^1(X,L^{-1}) = {\rm Ext}^1({\cal O}_X, L^{-1})$, we have to show that the short exact sequence $$0 \to L^{-1} \to E \to {\cal O}_X \to 0$$ splits. We compute $c_1(E)=-L$, $c_2(E)=0$, and $\Delta(E)=L^2$. Consequently, by corollary \ref{res3}, $E$ has a destabilizing subsheaf $A$ of rank one with $(2A+L)^2>0$, and $(2A+L).H > 0$, for all ample classes $H$. Since nef bundles are limits of ample classes, we obtain \begin{eqnarray}\label{r41} (2A+L).L & \geq &0 \,. \end{eqnarray} By the same reason, the Hodge index theorem applies \begin{eqnarray}\label{r42} A^2 & \leq & \frac{(A.L)^2}{L^2} \,. \end{eqnarray} The subsheaf $A$ of $E$ cannot be contained in $L^{-1}$ because $A$ destabilizes $E$ whereas $L^{-1}$ does not. Thus, $A$ is contained in ${\cal O}_X$. We conclude \begin{eqnarray}\label{r43} A.L & \leq & 0 \,. \end{eqnarray} Computing the second Chern class of $E$ in terms of $A$ and $E/A$ we obtain \begin{eqnarray}\label{r44} A.L+A^2 \geq & 0 \end{eqnarray} Combining (\ref{r42}) and (\ref{r44}) yields $A.L \leq -L^2$ or $A.L \geq 0$. In view of (\ref{r41}) and (\ref{r43}) we deduce that $A.L=0$. This equality, and the inequalities (\ref{r42}) and (\ref{r44}) imply $A^2=0$. Now we claim that $A.H=0$. Suppose this were not the case. Then $(L+a\cdot A).H$ would be zero, for a rational number $a$. Applying once again the Hodge index theorem yields $(L+a\cdot A)^2=L^2 \leq 0$ which contradicts our assumptions on $L$. Since $A$ is contained in ${\cal O}_X$ we conclude that $A={\cal O}_X$. Thus the injection $A \to E$ splits the exact sequence. { \hfill $\Box$} \subsection{The stable restriction theorem} \begin{theorem}\label{res5} Let $X$ be a smooth projective surface over an algebraically closed field with very ample line bundle ${\cal O}_X(H)$. Let $E$ be an $X$-vector bundle of rank 2 which is stable with respect to the polarization $H$. Assume furthermore, that the positive integer $m$ satisfies \begin{itemize} \item $m>\frac{a}{2H^2}-\frac{\Delta(E)}{2a}$ where $a>0$ is an integer not larger than the positive generator of the ideal $\{ A.H \, | \, A \in {\rm Pic}(X) \}$; \item $m^2H^2+\Delta(E)>\left\{ \begin{array}{ll} \left[ 1-4\chi({\cal O}_X) \right]_+ & \mbox{if } K_X.H<0 \,; \\ \left[ 2-4\chi({\cal O}_X) \right]_+ & \mbox{if } K_X.H=0 \,; \\ \left[ 6-4\chi({\cal O}_X) +4\cdot \left\lceil \frac{K_X.H}{H^2} \right\rceil K_X.H \right]_+ & \mbox{if } K_X.H>0 \,. \\ \end{array} \right.$ \end{itemize} Then the restriction of $E$ to any smooth curve of the linear system $|mH|$ is stable. \end{theorem} {\bf Proof: } \setcounter{equation}{0} Let $C$ be a smooth curve rationally equivalent to $mH$. Suppose that $E|_C$ has a quotient line bundle $L$ of degree $d$ where \begin{eqnarray}\label{r31} 2d & \leq & c_1(E).(mH) \, . \end{eqnarray} We denote the kernel of the surjection $E \to L$ by $E'$. Then: \begin{eqnarray}\label{r32} c_1(E')=c_1(E)-mH & & c_2(E')=c_2(E)+d-c_1(E).(mH) \end{eqnarray} Hence, we obtain from (\ref{r31}) $$ \Delta(E') = \Delta(E)+m^2H^2+2(c_1(E).(mH)-2d) \geq \Delta(E)+m^2H^2 \, .$$ From proposition \ref{res2} and the assumptions on $m$, we conclude that $E'$ cannot be stable with respect to $H$. Thus there is a line bundle $A$ destabilizing $E'$. We may assume that $E'/A$ is torsion free. We consider the following short exact sequence. $$0 \to A \to E' \to {\cal J}_Z(c_1(E)-mH-A) \to 0 $$ where $Z\subset X$ denotes a closed subscheme of codimension 2. Since $A$ is destabilizing we obtain $H.(2A+mH-c_1(E)) \geq 0$. We want to show that $A$ destabilizes $E$. It is convenient to introduce the divisor $B:=2A-c_1(E)$. Therefore we have to show that $B.H \geq 0$. Using this notation the last inequality reads \begin{eqnarray}\label{r33} B.H & \geq & -mH^2 \,. \end{eqnarray} Computing $c_2(E)$ via the above exact sequence yields $$c_2(E')=A.(c_1(E)-mH-A)+{\rm length}(Z) \,.$$ Using ${\rm length}(Z) \geq 0$, (\ref{r31}), and (\ref{r32}) we obtain $$A.(c_1(E)-mH-A) \leq c_2(E)-\frac{1}{2}c_1(E).(mH) \,.$$ This is equivalent to $B^2+2B.(mH)-\Delta(E) \geq 0$. Combined with the Hodge index theorem $\frac{(B.H)^2}{H^2}\geq B^2$ this yields \begin{eqnarray}\label{r34} (B.H)^2+2mH^2\cdot (B.H)-\Delta(E)\cdot H^2& \geq &0\,. \end{eqnarray} Our second assumption on $m$ implies that the quadratic equation $x^2+2mH^2\cdot x-\Delta(E)\cdot H^2 = 0$ for the indeterminant $x$ has a positive discriminant. Therefore, it results from (\ref{r33}) and (\ref{r34}) that $$B.H \geq -mH^2+\sqrt{(mH^2)^2+\Delta(E) \cdot H^2} \,.$$ By the assumption $m>\frac{a}{2H^2}-\frac{\Delta(E)}{2a}$ we eventually obtain from the last inequality $B.H>-a$. By the very definition of $a$ this shows that $B.H$ is non negative. { \hfill $\Box$} {\vspace{0.5em} } {\bf Remark:} Of course the number $a$ in theorem \ref{res5} can always be set to one. However, the larger $a$ the sharper becomes the estimation. In particular, if $H$ itself is the $k$th multiple of a divisor class, then we can set $a=k$. \section{The higher dimensional case} Before we generalize theorem \ref{res1} to varieties of dimension at least three, we present a lemma which is needed in the proof of \ref{res6}. First let us fix notations. Let $X$ be smooth variety of dimension $n>2$ defined over an algebraically closed field. Furthermore, let ${\cal O}_X(H)$ be a very ample line bundle. A torsion free $X$-sheaf $F$ of rank one is a line bundle outside a set of codimension two. Thus, the first Chern class $c_1(L)$ is well defined. \begin{lemma}\label{h0bound} Let $L$ be a torsion free $X$-sheaf of rank one. If $c_1(L).H^{n-1} < d_1$, then there is an integer $d_2$ only depending on $X$, $H$, and $d_1$ such that $h^0(L)<d_2$. \end{lemma} {\bf Proof: } We show this by induction on the dimension of $X$. The case $n=0$ being trivial. Suppose now that the result holds for schemes of dimension $n-1$. Since $L(-d_1H)$ has by assumption no global sections. Take a smooth divisor $D$ in the linear system $|d_1H|$ such that the restriction $L|_D$ is torsion free, and the sequence $0 \to L(-d_1H) \to L \to L|_D \to 0$ is exact. Then we have $h^0(L) \leq h^0(L|_D)$. Thus, we have reduced the case to dimension $n-1$. { \hfill $\Box$} \begin{theorem}\label{res6} Let $E$ be a rank two $X$-vector bundle which is semistable with respect to $H$. Let $l$ be an integer satisfying $l \geq \log_2\left( \sqrt{ \left[ \frac{-\Delta(E).H^{n-2}}{H^n} \right]_+ } +1 \right)$. Then the restriction of $E$ to a general divisor of the linear system $|2^lH|$ is semistable. \end{theorem} {\bf Proof: } We follow step by step the proof of theorem \ref{res1}. There is no problem in generalizing most steps. By definition of stability, we need only the first terms of Hilbert polynomials. Thus, for computations with Chern classes and Hilbert polynomials we are allowed to restrict to surfaces $S \subset X$ where $S$ is the intersection of $n-2$ divisors of the linear system $|H|$. In step 2 we replace $P_a$ by polynomials $P(k)=\frac{1}{2}a_0(E) {k+n-1 \choose n-1}+a{k+n-2 \choose n-2}+ \ldots$ with $a< \frac{a_1(E)}{2}$. The first Chern class of these destabilizing quotients is bounded above. Furthermore, for $m\geq m_0 \gg 0$ we have $H^q(E(m))=0$, for all $q>0$. Since a destabilizing quotient $Q$ on a divisor is a quotient sheaf of $E$, it results that $H^q(Q(m))=0$, for $q>0$ and $m \geq m_0$. It follows from lemma \ref{h0bound}, that there are upper bounds for $h^0(Q(m))$. Thus there exists only a finite number of possible Hilbert polynomials, for destabilizing quotients. Now taking the minimal polynomial $P$ such that the associated Quot scheme dominates $|mH|$ we can copy the above proof. { \hfill $\Box$}
\section*{Acknowledgements} I would like to thank Prof.\ A. Bia{\l}as for reading the manuscript and many suggestions and comments and Dr.\ R.\ Janik for discussions. This work was supported in part by Polish Government grant Project (KBN) 2P03B04214.
\section*{Acknowledgements} I would like to thank W. Bernreuther, M. Flesch, M. Spira, and P. Uwer for many enlightening discussions. \newpage
\section{Introduction} The idea of using extra dimensions in describing physical phenomena is a fairly mature one and dates back to the early decades of the twentieth century. During that time, attempts at unifying the theories of electromagnetism and gravitation were made by assuming the existence of an extra spatial dimension \cite{NKK}. A new application of extra dimensional theories has recently been proposed in Refs.\cite{ADD1, ADD2}, where it was suggested that the fundamental scale of gravity $M_F$ could be as low as the weak scale $\Lambda_w \sim 1$ TeV, assuming that there were $n$ large compactified extra dimensions. It was shown in Refs.\cite{ADD1, ADD2} that gravitational data allow $n \geq 2$. This proposal has significant phenomenological implications for collider experiments at the scale $\Lambda_w$, where Weak Scale Quantum Gravity (WSQG) effects are assumed to become strong. Lately, a great deal of effort has been made to constrain the proposal for WSQG \cite{Recent, SNGAM, KC}. In the case of $n = 2$, the most stringent constraints come from astrophysical and cosmological observations\cite{ADD2}, and it is argued that $M_F \,\,\rlap{\raise 3pt\hbox{$>$}}{\lower 3pt\hbox{$\sim$}}\,\, 100$ TeV \cite{SNGAM}. However, terrestrial experimental data have constrained WSQG to have $M_F \,\,\rlap{\raise 3pt\hbox{$>$}}{\lower 3pt\hbox{$\sim$}}\,\, 1$ TeV, and in the case of $n \geq 3$, there is no evidence of a more severe constraint. In this paper, we show that the process $\gamma \gamma \to \gamma \gamma$ at energies of order 1 TeV can be used to constrain WSQG over a large range in the TeV region. This process can be studied at a proposed Next Linear Collider (NLC) \cite{NLC}, where high energy Compton backward scattered photon beams with energies of order 1 TeV and luminosities of order 100 fb$^{-1}$ per year can be produced. The process $\gamma \gamma \to \gamma \gamma$ has the advantage that it receives contributions from the Standard Model (SM) only at the loop level and, therefore, could in principle be sensitive to new physics at the tree level. We will show that this process provides a good test of WSQG in the TeV regime. The rest of this paper will be organized as follows. In section 2, we present the basic ideas of WSQG in theories with large extra dimensions. An estimate of the expected size of the gravity contribution to the photon scattering process at weak scale energies will also be given in this section. We conclude that the effect is strong, but the SM contribution could be comparable and must be included. Section 3 contains the SM and the gravity amplitudes used in our calculations. This section also includes our discussion of the approximations that have been made in writing down the various amplitudes, and the conditions of their validity. In section 4, we discuss the method used in computing the predicted cross sections for $\gamma \gamma \to \gamma \gamma$ at the NLC, in the photon collider mode. The results of our computations for cross sections and the NLC reach for the effective scale of WSQG are presented in section 5. Finally, section 6 contains our concluding remarks. \section{WSQG and Its Contribution to $\gamma \gamma \to \gamma \gamma$} In this work, we assume the fundamental scale of gravity $M_F \,\,\rlap{\raise 3pt\hbox{$>$}}{\lower 3pt\hbox{$\sim$}}\,\, 1$ TeV and that there are $n \geq 2$ compact extra dimensions of size $R$, even though there are astrophysical and cosmological considerations that suggest $M_F \,\,\rlap{\raise 3pt\hbox{$>$}}{\lower 3pt\hbox{$\sim$}}\,\, 100$ TeV for $n = 2$\cite{SNGAM}. With these assumptions, Gauss' law yields the relation\cite{ADD1, ADD2} \begin{equation} M_P^2 \sim M_F^{n + 2} R^n, \label{MP} \end{equation} where $M_P \sim 10^{19}$ GeV is the Planck mass. The exact relation among $M_P$, $M_F$, and $R$, as presented in the appendix, depends on the convention and the compactification manifold used. However, for order of magnitude estimates, relation (\ref{MP}) suffices. Given the above assumptions, we expect gravitationally mediated processes at TeV energies to be important. To estimate the size of the WSQG effect in $\gamma \gamma \to \gamma \gamma$, we take $M_F \sim 1$ TeV, and the center of mass energy $\sqrt{s_{\gamma \gamma}} \sim 1$ TeV. The gravity contribution $\sigma_G$ to the total cross section is then given by \begin{equation} \sigma_G \sim \frac{2 \pi}{(16 \pi)^2} \left(\frac{E_\gamma}{M_F}\right)^6 \left(\frac{1}{M_F}\right)^2, \label{sigG} \end{equation} and we obtain \begin{equation} \sigma_G \sim 10 \, {\rm fb}, \label{sigGfb} \end{equation} where $E_\gamma = \sqrt{s_{\gamma \gamma}}/2$. We note that, in the TeV regime, the SM total cross section $\sigma_{SM} \sim 10$ fb is measurable at the NLC \cite{NLC, Jikia}, and Eq. (\ref{sigGfb}) suggests that the signal for WSQG in $\gamma \gamma \to \gamma \gamma$ can be large and measurable, as well. Our estimate also suggests that, although the effects of gravity can be large, the SM contribution is comparable and has to be included in our analysis. In the next section, we present the SM and gravity amplitudes used in our calculations. \section{The Amplitudes} We consider the process $\gamma (k_1) \gamma (k_2) \to \gamma (p_1) \gamma (p_2)$, where $k_1$ and $k_2$ are the initial and $p_1$ and $p_2$ are the final 4-momenta of the photons. We define $s \equiv (k_1 + k_2)^2, t \equiv (k_1 - p_1)^2$, and $u \equiv (k_1 - p_2)^2$. Each photon can have either $+$ or $-$ helicity. In what follows, we denote a helicity amplitude by $M_{i j k l}$, where $i, j, k, l = \pm$, and $(i, j)$ are the helicities of the $(k_1, k_2)$ photons, and $(k, l)$ are the helicities of the $(p_1, p_2)$ photons. The 1-loop helicity amplitudes of the SM are in general complicated. However, in the limit $s, |t|, |u| \gg m^2$, where $m$ is the mass of a $W$ boson, a quark, or a charged lepton, these amplitudes can be approximated by those parts of them that receive logarithmic enhancements\cite{Gounaris}. Except for the contribution of the top quark loop which does not affect our results significantly\cite{Gounaris}, these leading amplitudes provide a good approximation at the energies of interest to us, namely those of the NLC. We will discuss the necessary cuts and the regime of validity of these amplitudes, in the next section. The process $\gamma \gamma \to \gamma \gamma$ has many symmetries that reduce the number of independent helicity amplitudes. It can be shown \cite{Gounaris} that only three helicity amplitudes, out of 16, are independent, and they are $M_{++++}(s, t, u), M_{+++-}(s, t, u)$, and $M_{++--}(s, t, u)$. In the high energy limit that we are considering, of these three amplitudes, the only logarithmically enhanced one is $M_{++++}(s, t, u)$, for both the fermion and the $W$ loops\cite{Gounaris}. For the $W$ loop amplitude we have\cite{Gounaris} \[ \frac{M_{++++}^{(W)}(s, t, u)} {\alpha^2} \approx 12 + 12 \left(\frac{u - t}{s}\right) \left[\ln \left(\frac{-u - i\varepsilon}{m_W^2} \right) - \ln \left(\frac{-t - i\varepsilon}{m_W^2}\right) \right] \] \[ + 16 \left(1 - \frac{3 t u}{4 s^2}\right)\left(\left[\ln \left(\frac{-u - i\varepsilon}{m_W^2}\right) - \ln \left(\frac{-t - i\varepsilon} {m_W^2}\right)\right]^2 + \pi^2 \right) \] \[ +16 s^2 \left[\frac{1}{s t} \ln \left(\frac{-s - i\varepsilon}{m_W^2}\right) \ln \left(\frac{-t - i\varepsilon}{m_W^2}\right) + \frac{1}{s u} \ln \left(\frac{-s - i\varepsilon}{m_W^2}\right) \ln \left(\frac{-u - i\varepsilon}{m_W^2}\right) \right. \] \begin{equation} + \left.\frac{1}{t u} \ln \left(\frac{-t - i\varepsilon}{m_W^2}\right) \ln \left(\frac{-u - i\varepsilon}{m_W^2}\right) \right], \label{Wamp} \end{equation} where $\alpha \approx 1/137$ and $m_W$ is the mass of the $W$ boson; $m_W = 80$ GeV. This value of $\alpha$ corresponds to that appropriate for real initial and final state photons\footnote{We thank I. Ginzburg for bringing this point to our attention.}. \newpage For the fermion loops, we have \[ \frac{M_{++++}^{(f)}(s, t, u)}{\alpha^2 Q_f^4} \approx -8 - 8 \left(\frac{u - t}{s}\right) \left[\ln \left(\frac{-u - i\varepsilon}{m_f^2} \right) - \ln \left(\frac{-t - i\varepsilon}{m_f^2}\right) \right] \] \begin{equation} - 4 \left(\frac{t^2 + u^2}{s^2}\right)\left(\left[\ln \left(\frac{-u - i\varepsilon}{m_f^2}\right) - \ln \left(\frac{-t - i\varepsilon} {m_f^2}\right)\right]^2 + \pi^2 \right), \label{famp} \end{equation} where $Q_f$ is the fermion charge in units of the positron charge, and $m_f$ is the mass of the fermion in the loop. In our approximation, there are only two more leading helicity amplitudes that will enter our computations. These are \begin{equation} M_{+-+-}(s, t, u) = M_{++++}(u, t, s) \label{M+-+-} \end{equation} and \begin{equation} M_{+--+}(s, t, u) = M_{+-+-}(s, u, t). \label{M+--+} \end{equation} The gravity amplitudes are all at the tree level, and they are \[ M^{(G, s)} = (2 \pi) \, \varepsilon^\rho(k_1) \, \varepsilon^\sigma(k_2) \, \varepsilon^{* \gamma}(p_1) \, \varepsilon^{* \delta}(p_2) \, B^{\mu \nu, \alpha \beta}(k_1 + k_2) \, D(s) \] \begin{equation} \times [(k_1 \cdot k_2) C_{\mu \nu, \rho \sigma} + D_{\mu \nu, \rho \sigma}(k_1, k_2)] \, [(p_1 \cdot p_2) C_{\alpha \beta, \gamma \delta} + D_{\alpha \beta, \gamma \delta}(p_1, p_2)], \label{MGs} \end{equation} \[ M^{(G, t)} = (2 \pi) \, \varepsilon^\rho(k_1) \, \varepsilon^\sigma(k_2) \, \varepsilon^{* \gamma}(p_1) \, \varepsilon^{* \delta}(p_2) \, B^{\mu \nu, \alpha \beta}(k_1 - p_1) \, D(t) \] \begin{equation} \times [(k_1 \cdot p_1) C_{\mu \nu, \rho \gamma} + D_{\mu \nu, \rho \gamma}(k_1, p_1)] \, [(k_2 \cdot p_2) C_{\alpha \beta, \sigma \delta} + D_{\alpha \beta, \sigma \delta}(k_2, p_2)], \label{MGt} \end{equation} and \[ M^{(G, u)} = (2 \pi) \, \varepsilon^\rho(k_1) \, \varepsilon^\sigma(k_2) \, \varepsilon^{* \gamma}(p_1) \, \varepsilon^{* \delta}(p_2) \, B^{\mu \nu, \alpha \beta}(k_1 - p_2) \, D(u) \] \begin{equation} \times [(k_1 \cdot p_2) C_{\mu \nu, \rho \delta} + D_{\mu \nu, \rho \delta}(k_1, p_2)] \, [(k_2 \cdot p_1) C_{\alpha \beta, \sigma \gamma} + D_{\alpha \beta, \sigma \gamma}(k_2, p_1)], \label{MGu} \end{equation} where $\varepsilon^\mu (p)$ denotes the polarization vector of a photon with 4-momentum $p$, and the function $D(x)$ is given by\cite{Han} \[ D(x) \approx M_S^{-4} \ln \left(\frac{M_S^2}{|x|}\right) \, \, \, \, {\rm for} \, \, \, \, n = 2 \] and \begin{equation} D(x) \approx M_S^{-4} \left(\frac{2}{n - 2}\right) \, \, \, \, {\rm for} \, \, \, \, n > 2; \label{D(x)} \end{equation} the expressions for $B_{\mu \nu, \lambda \sigma}(k)$, $C_{\mu \nu, \lambda \sigma}$, and $D_{\mu \nu, \lambda \sigma}(k, p)$ are given in the appendix. Here, we would like to make a few comments regarding the amplitudes (\ref{MGs}), (\ref{MGt}), and (\ref{MGu}). First, we note that the expressions for $D(x)$ depend on the cutoff scale $M_S \gg s, |t|, |u|$, introduced to regulate the divergent sum over the infinite tower of Kaluza-Klein states. This dependence is a result of our implicit assumption that $M_S = M_F$. However, if $M_S$ is much smaller than $M_F$ then \begin{equation} D(x) \to \left(\frac{M_S}{M_F}\right)^{(n + 2)} D(x) \, \, \, \, {\rm for} \, \, \, \, n \geq 2, \label{DMF} \end{equation} resulting in a suppression\cite{Giudice}. Secondly, it should be kept in mind that the amplitudes (\ref{MGs}), (\ref{MGt}), and (\ref{MGu}) are derived from an effective Lagrangian\cite{Han} with the lowest dimension operators that describe the coupling of the Kaluza-Klein gravitons to various fields, in our case, the photon field. In this effective description of quantum gravity, we need to introduce a cutoff $M_S \,\,\rlap{\raise 3pt\hbox{$<$}}{\lower 3pt\hbox{$\sim$}}\,\, M_F$, in order to get finite results. As with any effective Lagrangian, there are terms of higher dimension that should in principle be included in the Lagrangian. The term $\lambda (F_{\mu \nu} F^{\mu \nu})^2/M_S^4$, where $\lambda$ is an unknown coefficient, is one such term that contributes at the same order in powers of $1/M_S$ to our calculations. The coefficient $\lambda$ cannot be calculated, unless the fundamental theory of gravity at scale $M_F$ is known. In principle, the size of the contribution from this term can be larger than the one calculated in this paper, and may even have the opposite sign. However, since $\lambda$ is unknown, we have chosen to consider only the lowest dimension local terms in the Lagrangian, and simply add the contributions from Eqs. (\ref{MGs}), (\ref{MGt}), and (\ref{MGu}) to those obtained from Eqs. (\ref{Wamp}) and (\ref{famp}). This is a reasonable choice, as long as one is only interested in an order of magnitude estimate of the effects. \section{The NLC as a Photon Collider} We mentioned before that high energy and luminosity $\gamma$ beams can be achieved at the NLC. The basic proposed mechanism uses backward Compton scattering of laser photons from the high energy $e^+ e^-$ beams at the NLC\cite{Ginzburg}. The $\gamma$ beams that are obtained in this way have distributions in energy and helicity that are functions of the $\gamma$ energy and the initial polarizations of the electron beams and the laser beams. Laser beam polarization $P_l$ can be achieved close to $100\%$, however, electron beam polarization $P_e$ is at the $90\%$ level. We take $|P_l| = 1$ and $|P_e| = 0.9$ for our calculations. Let $E_e$ be the electron beam energy, and $E_\gamma$ be the scattered $\gamma$ energy in the laboratory frame. The fraction of the beam energy taken away by the photon is then \begin{equation} x = \frac{E_\gamma}{E_e}. \label{x} \end{equation} We take the laser photons to have energy $E_l$. Then, the maximum value of $x$ is given by \begin{equation} x_{max} = \frac{z}{1 + z}, \label{xmax} \end{equation} where $z = 4 E_e E_l/m_e^2$, and $m_e$ is the electron mass. One cannot increase $x_{max}$ simply by increasing $E_l$, since this makes the process less efficient because of $e^+ e^-$ pair production through the interactions of the laser photons and the backward scattered $\gamma$ beam. The optimal value for $z$ is given by \begin{equation} z_{_{OPT}} = 2 \left(1 + {\sqrt 2}\right). \label{zOPT} \end{equation} The photon number density $f(x, P_e, P_l)$ and average helicity $\xi_2 (x, P_e, P_l)$ are functions of $x$, $P_e$, $P_l$, and $z$, however, we always set $z = z_{_{OPT}}$ in our calculations. We give the expressions for these two functions in the appendix. Let $M_{i j k l}$ be a helicity amplitude for $\gamma \gamma \to \gamma \gamma$. We define \begin{equation} |M_{++}|^2 \equiv \sum_{k, l} |M_{++ k l}|^2 \label{M++2} \end{equation} and \begin{equation} |M_{+-}|^2 \equiv \sum_{k, l} |M_{+- k l}|^2, \label{M+-2} \end{equation} where the summation is over the final state helicities of the photons. Then, for various choices of the pairs $(P_{e_1}, P_{l_1})$ and $(P_{e_2}, P_{l_2})$ of the the two beams, the differential cross section $d \sigma/d \Omega$ is given by \[ \frac{d \sigma}{d \Omega} = \frac{1}{128 \, \pi^2 \, s_{ee}} \int \int d x_1 d x_2 \left[\frac{f(x_1) \, f(x_2)}{x_1 \, x_2}\right] \] \begin{equation} \times \left[\left(\frac{1 + \xi_2(x_1) \, \xi_2(x_2)}{2}\right)|M_{++}|^2 + \left(\frac{1 - \xi_2(x_1) \, \xi_2(x_2)}{2}\right)|M_{+-}|^2\right], \label{diffcs} \end{equation} where $x_1$ and $x_2$ are the energy fractions for the two beams, given by Eq. (\ref{x}), and $s_{ee} = 4 E_e^2$. Different choices of $(P_{e_1}, P_{l_1})$ and $(P_{e_2}, P_{l_2})$ in $(f(x_1), \xi_2(x_1))$ and $(f(x_2), \xi_2(x_2))$, respectively, yield different polarization cross sections. We note that the expressions for $|M_{++}|^2$ and $|M_{+-}|^2$ are actually functions of the $\gamma \gamma$ center of mass energy squared ${\hat s} = x_1 x_2 s$, and the center of mass scattering angle $\theta_{cm}$. We also have ${\hat t} = x_1 x_2 t$ and ${\hat u} = x_1 x_2 u$. In the previous section, we introduced the logarithmically enhanced SM amplitudes, valid when $s, |t|, |u| \gg m_W^2$. However, we see that to have a good approximation, we must demand ${\hat s}, |{\hat t}|, |{\hat u}| \gg m_W^2$. To avoid restricting the phase space too much, and in order to have a good approximation to the SM amplitudes, we will make the following cuts \[ \theta_{cm} \in [\pi/6, 5 \pi/6], \] \[ x_1 \in [\sqrt{0.4}, x_{1 max}], \] and \begin{equation} x_2 \in [\sqrt{0.4}, x_{2 max}], \label{cuts} \end{equation} where $x_{1 max}$ and $x_{2 max}$ are given by Eq. (\ref{xmax}); $x_{1 max} = x_{2 max}$. These cuts ensure that the integrations are always performed in a region where ${\hat s}, |{\hat t}|, |{\hat u}| > m_W^2$. \section{Results} In this section, we present our numerical results for the expected size of the WSQG effects at TeV energies. However, here, we would like to make a few remarks regarding our calculations. First of all, in obtaining our results, we have assumed $M_S = M_F$. The effects of departure from this assumption are given in Eq. (\ref{DMF}). Secondly, the only dependence on the number of extra dimensions $n$ in our computations comes from Eq. (\ref{D(x)}). We only distinguish between the cases with $n = 2$ and $n > 2$. In the case with $n = 2$, in the limit $M_S^2 \gg s$, the WSQG amplitude is enhanced logarithmically compared to the case with $n > 2$. In our computations, for $n > 2$, we have $\ln (M_S^2/{\hat s}) > 2/(n - 2)$ over most of the parameter space considered. We choose $n = 6$ as a representative value for $n > 2$; other choices result in a rescaling of the effective value of $M_S$. In Fig. (\ref{sgamma}), we present the $\gamma \gamma \to \gamma \gamma$ cross sections for SM + WSQG and SM, assuming a mono-energetic beam of photons. For the gravity contribution, we have chosen $M_S = 3$ TeV, and $n = 6$. The curves in Fig. (\ref{sgamma}) are obtained for photons with $++$ and $+-$ initial helicities. At the NLC, the $\gamma$ beams will have a distribution in photon energies and helicities, and these cross sections will not be observed. However, the cross sections presented in Fig. (\ref{sgamma}) show the relative size of the contribution of each initial helicity state to the predicted cross section at the NLC, as obtained from Eq. (\ref{diffcs}). The six SM + WSQG cross sections, for $M_S = 3$ TeV and $n = 6$, in Fig. (\ref{sixpols}), correspond to six independent choices for the polarizations $(P_{e_1}, P_{l_1}, P_{e_2}, P_{l_2})$ of the electron and the laser beams at the NLC, in the photon collider mode. These cross sections are plotted versus the center of mass energy of the beam, $\sqrt{s_{ee}}$. The curves in this figure show a sensitive dependence on the choices of the polarizations for $\sqrt{s_{ee}} \,\,\rlap{\raise 3pt\hbox{$>$}}{\lower 3pt\hbox{$\sim$}}\,\, 1$ TeV, with the $(+, -, +, -)$ polarization giving the largest cross section at high energies. In Fig. (\ref{26sm}), choosing $M_S = 3$ TeV and $n =2, 6$, we compare the SM + WSQG cross sections with that of the SM in the typical proposed NLC center of mass energy range $\sqrt{s_{ee}} \in [500, 1500]$ GeV. We have chosen the $(+, -, +, -)$ polarization for all three curves, since this choice yields the largest gravity cross section, as shown in Fig. (\ref{sixpols}). The SM and SM + WSQG differential cross sections $d \sigma/d (\cos \theta_{cm})$, at $\sqrt{s_{ee}} = 500$ GeV, are compared in Fig. (\ref{dcs}). Again, we have chosen $M_S = 3$ TeV and the $(+, -, +, -)$ polarization. The endpoint behavior of the differential cross sections are caused by our choice for the cuts, given in Eqs. (\ref{cuts}). At this center of mass energy, and given our cuts, the $n = 6$ result does not offer a distinctive signal for WSQG. For $n = 2$, we see that the differential cross section for SM + WSQG, in the region where $\cos \theta_{cm} \approx 0$, is larger than that for SM by about a factor of 2. Next, we present our results for the typical $M_S$ reach of the NLC, at various stages. The stages at which $\sqrt{s_{ee}} = 500$ GeV, $\sqrt{s_{ee}} = 1000$ GeV, and $\sqrt{s_{ee}} = 1500$ GeV are denoted by NLC0.5, NLC1.0, and NLC1.5, respectively. Throughout, we assume that the luminosity $L = 100$ fb$^{-1}$ per year. We use the $\chi^2 (M_S)$ variable, given by \begin{equation} \chi^2 (M_S) = \left(\frac{L}{\sigma_{_{SM}}}\right)\left[\sigma_{_{SM}} - \sigma (M_S)\right]^2, \label{chi2} \end{equation} where $\sigma_{_{SM}}$ and $\sigma (M_S)$ refer to the cross sections for the SM and SM + WSQG, respectively. We have chosen the $(+, -, +, -)$ polarization for computing $\sigma (M_S)$, since this choice gives the largest high energy SM + WSQG cross sections. To establish the reach in each case, we require a one-sided $95\%$ confidence level, corresponding to $\chi^2 (M_S) \geq 2.706$. The plots in Figs. (\ref{reach0.5}), (\ref{reach1.0}), and (\ref{reach1.5}) show the $95\%$ confidence level experimental reach for $M_S$ at NLC0.5, NLC1.0, and NLC1.5, respectively. As one can see from these figures, the largest reach at each stage is for $n = 2$. This is because of the logarithmic enhancement of the WSQG amplitude for $n = 2$, as given by Eqs. (\ref{D(x)}). The lowest reach in $M_S$ is about 2 TeV for $n = 6$ at NLC0.5 and the largest $M_S$ reach is about 9 TeV for $n = 2$ at NLC1.5. Note that these values are obtained for $L = 100$ fb$^{-1}$ per year, and by increasing $L$, the reach in $M_S$ will be improved. \section{Concluding Remarks} The process $\gamma \gamma \to \gamma \gamma$ at TeV energies is an important test channel for WSQG theories, since the tree level gravity contribution to the process in these theories is expected to be significant, whereas the SM contributes only at the loop level. In this paper, we have used the high energy limit SM helicity amplitudes and the gravity amplitudes from the lowest dimension WSQG effective Lagrangian to compute scattering cross sections. The SM + WSQG cross sections can significantly differ from those of the SM alone. We have shown that the NLC in the photon collider mode can be effectively used to constrain theories of quantum gravity at the weak scale. The size of the expected effect shows a strong dependence on the choice of initial electron and laser polarizations. Our computations suggest that studying $\gamma \gamma \to \gamma \gamma$ at the NLC, operating at $\sqrt{s_{ee}} \in [500, 1500]$ GeV and $L = 100$ fb$^{-1}$ per year, can constrain the scale $M_S$ at which quantum gravity becomes important, over the range 1 TeV $\,\,\rlap{\raise 3pt\hbox{$<$}}{\lower 3pt\hbox{$\sim$}}\,\, M_S \,\,\rlap{\raise 3pt\hbox{$<$}}{\lower 3pt\hbox{$\sim$}}\,\,$ 10 TeV. \section*{Acknowledgements} It is a pleasure to thank T. Rizzo for many helpful discussions. The author would also like to thank N. Arkani-Hamed, S. Brodsky, L. Dixon, I. Ginzburg, J.L. Hewett, M. Peskin, J. Rathsman, and M. Schmaltz for various comments and conversations. While this work was being completed, we received a paper\cite{KC} by K. Cheung whose contents have some overlap with those of this work.
\section{INTRODUCTION} One of the puzzles in heavy hadron decays is the explanation of the low lifetime ratio $\tau(\Lambda_b)/\tau(B_d)=0.79\pm 0.06$ \cite{ad}. Inclusive heavy hadron decays involve both nonperturbative and perturbative corrections to the tree-level $b$ quark decay amplitudes. In the framework of heavy quark effective theory (HQET) \cite{aa}, nonperturbative corrections are expanded in powers of $1/M_b$, $M_b$ being the $b$ quark mass, with the coefficient of each power proportional to a hadronic matrix element of local operators, and perturbative corrections are evaluated order by order at the quark level. The HQET prediction for the ratio $\tau(\Lambda_b)/\tau(B_d)$ to $O(1/M_b^2)$ is about 0.99 \cite{NS}. When including the $O(1/M_b^3)$ corrections, the ratio depends on six unknown parameters, and reduces to around 0.97 for various model estimates, which is still far beyond the experimental data. On the other hand, a phenomenological ansatz was proposed, in which the overall $M_b^5$ factor in front of nonleptonic decay widths is replaced by the corresponding hadron mass $M_H^5$ \cite{ac}. This ansatz provides a solution to the puzzle, since $(M_B/M_{\Lambda_b})^5=0.73$, $M_B$ and $M_{\Lambda_b}$ being the $B$ meson mass and the $\Lambda_b$ baryon mass, respectively, is close to the observed ratio. As pointed out in \cite{ae}, the same replacement also explains the absolute $B$ meson decay rate, while the HQET prediction using the expansion parameter $M_b$ accounts for only 80\% of the decay rate. An alternative approach to inclusive heavy hadron decays is the perturbative QCD (PQCD) factorization theorem \cite{ab}. In this formalism radiative corrections to $b$ quark decays are absorbed into different factors in factorization formulas for decay widths according to their characteristic scales. The soft corrections characterized by the hadronic scale of order $\Lambda_{\rm QCD}$ are factorized into a universal heavy hadron distribution function. The corrections characterized by the heavy hadron mass and by the $W$ boson mass are factorized into a hard $b$ quark decay amplitude and a ``harder" function, respectively. The heavy hadron distribution function provides a summation over the initial states of the $b$ quark, such that the virtual and real soft gluon corrections cancel exactly. The introduction of a distribution function indicates that a residual momentum of the $b$ quark, whcih arises as an effect from the light degrees of freedom in the heavy hadron, is allowed. The inclusion of the light degrees of freedom then demands the use of the heavy hadron kinematics, whose difference from the $b$ quark kinematics sets a constraint on the magnitude of the residual momentum. The soft gluons are collected into the distribution function by employing an eikonal approximation, under which a $b$ quark propagator is simplified into the propagator associated with the large component of the rescaled $b$ quark field in HQET. This observation implies the identity of the heavy hadron distribution function to the one constructed in the HQET framework \cite{N}. Therefore, nonperturbative corrections also start from $O(1/M_b^2)$ and the quark-hadron duality is respected in the PQCD formalism, the same as in HQET. However, as transforming the $b$ quark kinematics to the heavy hadron kinematics through $M_b=M_H-{\bar\Lambda}$, the $O({\bar\Lambda}/M_H)$ correction occurs \cite{N}. It has been explicitly demonstrated, taking the $B\to X_u l{\bar\nu}$ as an example \cite{ab}, that the $O(1/M_B)$ correction from the $B$ meson distribution function cancels that from the phase space factor $M_B^5$. As to perturbative corrections, various large logarithms contained in $b$ quark decays are summed to all orders in the PQCD approach. The results are the renormalization-group (RG) evolutions among the three characteristic scales, such that the factorization formulas are independent of the renormalization scale $\mu$. Double logarithms associated with light energetic final-state quarks are organized by the Collins-Soper resummation technique \cite{CS}, leading to a Sudakov factor which smears end-point singularities and improves the applicability of PQCD. These perturbative factors, depending on kinematics of the initial heavy hadrons and of final states, differ among various decay modes of a hadron and among various hadrons. This is natural, because different decay modes should involve different energy releases. In the HQET approach perturbative corrections are evaluated to finite orders explicitly at the quark level with the renormalization scale $\mu$ set to a common value for all decay modes. It is then not a surprise that HQET predicts the lifetime ratio $\tau(\Lambda_b)/\tau(B_d)\approx 0.99$, since nonperturbative corrections, supposed to distinguish the decays of different $b$-hadrons, are suppressed by $1/M_b^2$, and perturbative corrections to $b$ quark decays are the same. The $B$ meson distribution function has been extracted from the photon energy spectrum of the $B\to X_s\gamma$ decay \cite{ag}, which provides the information of the $b$ quark mass $M_b$ and of the HQET parameter $\lambda_1$ for the $B_d$ meson. Based on this information, we propose reasonable $B_s$ meson and $\Lambda_b$ baryon distribution functions by assuming that they correspond to the same $\lambda_1$. Convoluting these distribution functions with the perturbative factors, we predict the correct lifetimes of the $b$-hadrons: $\tau(B_d)=1.56$ ps, $\tau(B_s)=1.46$ ps and $\tau(\Lambda_b)=1.22$ ps, and thus the correct ratios $\tau(B_s)/\tau(B_d)=0.94$ and $\tau(\Lambda_b)/\tau(B_d)=0.78$. By varying the $B$ meson distribution function slightly, we obtain the $B_u$ meson lifetime $\tau(B_u)=1.62$ ps, and the ratio $\tau(B_u)/\tau(B_d)=1.04$. It will be shown that perturbative contributions play an essential role in the explanation of the $b$-hadron lifetimes, and that our predictions are insensitive to the variation of the distribution functions, as expected from the quark-hadron duality. Our results of the semileptonic branching ratio $B_{\rm SL}=B(B\to Xl{\bar\nu})=10.16\%$ and of the average charm yield $\langle n_c\rangle=1.17$ per $B$ decay are also consistent with experimental data. We present the semileptonic branching ratios and the charm yields in $B_s$ meson and $\Lambda_b$ baryon decays as well, which can be tested in future experiments. In Sec.~II the factorization theorem for semileptonic $b$-hadron decays is constructed. The equivalence of the $B$ meson distribution function to that derived in the HQET framework is demonstrated in Sec.~III. The three-scale factorization theorem for nonleptonic decays is developed in Sec.~IV. Various logarithmic corrections are summed to all orders using the resummation technique and RG equations in Sec.~V. In Sec.~VI we present the factorization formulas for the $b$-hadron decay widths and their numerical analysis. Section VII is the conclusion. The appendix contains the detailed derivation of the allowed phase space for heavy hadron decays. \section{FACTORIZATION OF SEMILEPTONIC DECAYS} We start with the factorization of the simplest case, the semileptonic decays $B\to X_cl{\bar \nu}$, which correspond to the $b\to cl{\bar\nu}$ decays at the quark level. These decays involve only the $B$ meson distribution function and the hard $b$ quark decay amplitudes due to the lack of the characteristic scale of the $W$ boson mass. Nonperturbative dynamics is reflected by infrared poles in radiative corrections to quark-level amplitudes in perturbation theory. According to PQCD factorization theorems, these poles are absorbed into a hadron distribution function, which must be derived by nonperturbative methods or extracted from experimental data. A distribution function, being universal, is determined once for all, and then employed to make predictions for other processes containing the same hadron. In this section we demonstrate how to isolate infrared poles from radiative corrections and factorize them into the $B$ meson distribution function. Consider the one-loop corrections to the $b\to cl{\bar\nu}$ decays shown in Fig.~1. Because both the $b$ and $c$ quarks are massive, there are no collinear (mass) divergences, and we concentrate only on soft divergences from vanishing loop momenta. The self-energy correction to the $b$ quark in Fig.~1(a) is written as \begin{eqnarray} \Sigma^{(a)} u_b=-ig^2C_F\mu^{2\epsilon} \int\frac{d^{4-2\epsilon}l}{(2\pi)^{4-2\epsilon}} \frac{1}{l^2}\frac{{\not p}_b+M_b}{p_b^2-M_b^2}\gamma_\mu \frac{{\not p}_b-\not l+M_b}{(p_b-l)^2-M_b^2}\gamma^\mu u_b \delta(p_c^2-M_c^2)\;, \label{a1} \end{eqnarray} with $C_F=4/3$ a color factor, $p_b$ and $u_b$ the $b$ quark momentum and spinor, respectively, and $p_c=p_b-q$ and $M_c$ the $c$ quark momentum and mass, respectively, $q$ being the lepton pair momentum. The $\delta$-function from the final-state cut on the outgoing $c$ quark is included for the discussion of the soft cancellation between virtual and real corrections below. The $b$ quark propagator $({\not p}_b+M_b)/(p_b^2-M_b^2)$ after the self-energy correction helps the extraction of the soft pole in Eq.~(\ref{a1}). The soft divergence of $\Sigma^{(a)}$ is isolated by the eikonal ($l\to 0$) approximation, \begin{eqnarray} \Sigma^{(a)}_{\rm soft} u_b=-ig^2C_F\mu^{2\epsilon} \int\frac{d^{4-2\epsilon}l}{(2\pi)^{4-2\epsilon}} \frac{1}{l^2}\frac{{\not p}_b+M_b}{p_b^2-M_b^2} \gamma_\mu\frac{2p_b^\mu}{-2p_b\cdot l+p_b^2-M_b^2} u_b \delta(p_c^2-M_c^2)\;, \label{1a1} \end{eqnarray} where the term $\not l$ in the numerator and $l^2$ in the denominator of the $b$ quark propagator have been neglected, and the equation of motion $({\not p}_b-M_b)u_b=0$ has been applied to obtain the factor $2p_b^\mu$. We make the following expansion because of the on-shell condition $p_b^2-M_b^2\to 0$: \begin{equation} \frac{1}{-2p_b\cdot l+p_b^2-M_b^2} =-\frac{1}{2p_b\cdot l}-\frac{p_b^2-M_b^2}{4(p_b\cdot l)^2}\;, \end{equation} where the first term on the right-hand side does not contribute, since it leads to an integrand with an odd power in $l$. The numerator of the second term removes the on-shell pole of the $b$ quark propagator after the self-energy correction. Employing the equation of motion again, we obtain \begin{eqnarray} \Sigma^{(a)}_{\rm soft}=ig^2C_F\mu^{2\epsilon}\int\frac{d^{4-2\epsilon}l} {(2\pi)^{4-2\epsilon}}\frac{v^2}{l^2(v\cdot l)^2}\delta(p_c^2-M_c^2)\;, \label{eik} \end{eqnarray} where $v\equiv p_b/M_b$ is the $b$ quark velocity. Obviously, Eq.~(\ref{eik}) does not depend on the spin and mass of the $b$ quark. The loop integral for Fig.~1(b) with a real gluon attaching the $b$ quarks before and after the final-state cut is written as \begin{eqnarray} u_b\Sigma^{(b)}u_b&=&-g^2C_F\mu^{2\epsilon} \int\frac{d^{4-2\epsilon}l}{(2\pi)^{4-2\epsilon}} u_b\gamma_\mu\frac{{\not p}_b-\not l+M_b}{(p_b-l)^2-M_b^2} \Gamma\frac{{\not p}_b-\not l+M_b}{(p_b-l)^2-M_b^2}\gamma^\mu \nonumber\\ & &\times 2\pi\delta(l^2)\delta((p_c-l)^2-M_c^2)\;, \label{b1} \end{eqnarray} where $\Gamma$ represents the other irrelevant vertices and propagators. The $\delta$-function is associated with the final-state $c$ quark with the momentum $p_c-l=p_b-q-l$ in the case of real gluon emission. The soft divergence in Eq.~(\ref{b1}) is also extracted by applying the eikonal approximation and the equation of motion to the two $b$ quark propagators, leading to \begin{eqnarray} \Sigma^{(b)}_{\rm soft}=-g^2C_F\mu^{2\epsilon} \int\frac{d^{4-2\epsilon}l}{(2\pi)^{4-2\epsilon}} \frac{v^2}{(v\cdot l)^2}2\pi\delta(l^2)\delta((p_c-l)^2-M_c^2)\;. \label{rel} \end{eqnarray} The above expression indicates that under the eikonal approximation, the $b$ quark propagator is replaced by the eikonal propagator $1/(v\cdot l)$ and the quark-gluon vertex $\gamma^\mu$ is replaced the eikonal vertex $v^\mu$. These Feynman rules for an eikonal line are exactly the same as those associated with the large component $h_v(x)$ of the rescaled $b$ quark field $b_v(x)$, \begin{equation} h_v(x)=\frac{1+\not v}{2}b_v(x)\;,\;\;\;\; b_v(x)=\exp(iM_bv\cdot x)b(x)\;, \label{hq} \end{equation} that appears in HQET. This observation will become essential, as we demonstrate the identity of the $B$ meson distribution functions constructed in the PQCD factorization theorems and in HQET. A straightforward calculation gives \begin{eqnarray} \Sigma^{(a)}_{\rm soft}&=&\frac{\alpha_s}{2\pi}C_F\frac{1}{-\epsilon} \delta(p_c^2-M_c^2)\;, \label{1as}\\ \Sigma^{(b)}_{\rm soft}&=&-\frac{\alpha_s}{\pi}C_F \frac{(\mu^2 p_c^2)^\epsilon}{(p_c^2-M_c^2)^{1+2\epsilon}}\;. \label{1bs} \end{eqnarray} The ultraviolet pole in $\Sigma^{(a)}_{\rm soft}$ has been subtracted, $1/(-\epsilon)$ with $\epsilon < 0$ denotes the soft pole, and the other finite terms irrelevant to our discussion have been dropped. The real gluon correction $\Sigma^{(b)}_{\rm soft}$ in fact contains a soft pole as $p_c^2\to M_c^2$. Following the standard factorization of deep inelastic scattering, this pole is extracted through the convolution with a $B$ meson structure function. The introduction of such a structure function implies that the $b$ quark momentum $p_b$ is allowed to vary. A variable $p_b$ is natural, since the light degrees of freedom in the $B$ meson share various amount of meson momentum, such that the $b$ quark is not always at rest. Define the residual momentum of the $b$ quark as $k=p_b-M_bv$, $v=(v^+,v^-,{\bf v}_\perp)=(1,1,{\bf 0})/\sqrt{2}$ in the light-cone notation, for which $k=(k^+,0,{\bf 0})$ is a convenient parametrization. The $B$ meson structure function $f(k^+)$ then describes the probability of finding a $b$ quark with the residual momentum $k^+$ inside the $B$ meson. Consequently, the one-loop corrections in Figs.~1(a) and 1(b) are modified into the convolutions of Eqs.~(\ref{a1}) and (\ref{b1}) with $f(k^+)$, respectively, where the $c$ quark momentum is replaced by $p_c=M_bv+k-q$. It is legitimate to reexpress the factor $1/(p_c^2-M_c^2)^{1+2\epsilon}$ in Eq.~(\ref{1bs}) as \begin{eqnarray} \frac{1}{(p_c^2-M_c^2)^{1+2\epsilon}} =\frac{1}{-2\epsilon}\delta(p_c^2-M_c^2) +\frac{1}{(p_c^2-M_c^2)_+}\;. \label{ind0} \end{eqnarray} The second term in the above expression is defined via the convolution with an arbitrary function $F(k^+)$, \begin{eqnarray} \int dk^+\frac{F(k^+)}{(p_c^2-M_c^2)_+} =\int dk^+\frac{F(k^+)-F({\bar k})}{p_c^2-M_c^2}\;, \label{ind} \end{eqnarray} in which \begin{eqnarray} {\bar k}=-\frac{M_b^2-2M_bq^0+q^2-M_c^2}{2(M_bv^--q^-)}\;, \end{eqnarray} is the value of $k^+$ determined by the on-shell condition $p_c^2=M_c^2$. Apparently, Eq.~(\ref{ind}) is infrared finite as $k^+\to {\bar k}$ ($p_c^2\to M_c^2$), and thus the second term in Eq.~(\ref{ind0}) is negligible. The first term in Eq.~(\ref{ind0}) leads to \begin{eqnarray} \Sigma^{(b)}_{\rm soft}=-\frac{\alpha_s}{2\pi}C_F \frac{1}{-\epsilon}\delta(p_c^2-M_c^2)\;, \label{sib} \end{eqnarray} which cancels $\Sigma^{(a)}_{\rm soft}$ in Eq.~(\ref{1as}). A similar cancellation occurs between the self-energy correction to the $c$ quark in Fig.~1(c) and its corresponding real gluon correction in Fig.~1(d), when they are convoluted with the $B$ meson structure function. The loop integrals are the same as those in Eqs.~(\ref{eik}) and (\ref{rel}) but with the $b$ quark velocity $v$ replaced by the $c$ quark velocity $v_c\equiv p_c/M_c$. Using the eikonal approximation, the soft structure of the vertex correction in Fig.~1(e) with the virtual gluon attaching the $b$ and $c$ quarks is given by \begin{eqnarray} \Sigma^{(e)}_{\rm soft}=-ig^2C_F\mu^{2\epsilon} \int\frac{d^{4-2\epsilon}l}{(2\pi)^{4-2\epsilon}} \frac{v\cdot v_c}{l^2 v\cdot l v_c\cdot l}\delta(p_c^2-M_c^2)\;. \end{eqnarray} The soft structure extracted from the corresponding real correction in Fig.~1(f) is \begin{eqnarray} \Sigma^{(f)}_{\rm soft}=g^2C_F\mu^{2\epsilon} \int\frac{d^{4-2\epsilon}l}{(2\pi)^{4-2\epsilon}} \frac{v\cdot v_c}{v\cdot l v_c\cdot l}2\pi\delta(l^2) \delta((p_c-l)^2-M_c^2)\;. \end{eqnarray} Following the similar reasoning, we show that $\Sigma^{(e)}_{\rm soft}$ and $\Sigma^{(f)}_{\rm soft}$ cancel exactly. The above soft cancellation can be generalized to all orders straightforwardly using the eikonal approximation, under which soft gluons detach from the quarks. In conclusion, the exact soft cancellation between virtual and real corrections to $b$ quark decays must be implemented by introducing the $B$ meson structure function. This is consistent with the Kinoshita-Lee-Nauenberg theorem, which states that infrared divergences in radiative corrections to a QCD process cancel when both final and initial states are summed. The $B$ meson structure function $f(k^+)$ simply provides a weighting factor for the summation over the initial $b$ quark states. The one-loop contributions to the hard $b$ quark decay amplitude $H$ for the semileptonic decays are defiined as the difference of the full radiative corrections and their eikonalized versions, that have been absorbed into the $B$ meson structure function. As a demonstration, we perform the factorization of the self-energy correction to the $b$ quark: \begin{eqnarray} & &({\rm tree}\;\;{\rm diagram})+{\rm Fig.~1(a)} \nonumber\\ &=&({\rm tree}\;\;{\rm diagram})+{\rm Fig.~1(a)} -({\rm tree}\;\;{\rm diagram})\times \Sigma^{(a)}_{\rm soft} +({\rm tree}\;\;{\rm diagram})\times \Sigma^{(a)}_{\rm soft}\;, \nonumber\\ &=&[({\rm tree}\;\;{\rm diagram})+{\rm Fig.~1(a)} -({\rm tree}\;\;{\rm diagram})\times \Sigma^{(a)}_{\rm soft}]\times (1+\Sigma^{(a)}_{\rm soft})+O(\alpha_s^2)\;. \label{fas} \end{eqnarray} The first and second factors in the last line of the above expression belong to the first-order $H$ and $f(k^+)$, respectively. Extending the above procedures to all orders, we derive the factorization formula for the semileptonic decays, which is a convolution of the hard amplitude with the $B$ meson structure function. Because of the soft subtraction $({\rm tree}\;\;{\rm diagram})\times \Sigma^{(a)}_{\rm soft}$, $H$ is infrared finite and calculable at the quark level in perturbation theory. \section{MOMENTS OF THE DISTRIBUTION FUNCTION} A formal definition of the $B$ meson structure function $f(k^+)$ should reflect the soft dynamics associated with the $B$ meson. Hence, we replace the $b$ quark line by an eikonal line in the direction $v$, which collects infinite many soft gluons radiated by the $b$ quark. From Eq.~(\ref{hq}), the dependence on the $b$ quark mass $M_b$ is removed by employing the rescaled $b$ quark field $b_v$. Specifying the large component $h_v$ of $b_v$, the spin of the $b$ quark remains fixed as radiating gluons, and its dependence also decouples. Therefore, the propagator and the gluon vertex associated with $h_v$ in HQET are exactly the same as the eikonal propagator and vertex, implying that $f(k^+)$ is defined in terms of $h_v$. The initial and final $b$ quarks have a separation $y^-$ in the minus coordinate, since the $b$ quark momentum varies in the plus direction. At last, to render the definition gauge invariant, we insert a path-ordered exponential $P\exp[i\int_0^{y^-}dtn\cdot A(tn)]$ in between the initial and final $b$ quark fields, with $n=(0,1,{\bf 0})$ a vector on the light cone, forming \begin{equation} f(k^+)=\int\frac{dy^-}{2\pi}e^{ik^+y^-} \langle B(v)|{\bar h}_v(0)P\exp\left[-i\int_0^{y^-}dtn\cdot A(tn)\right] h_v(y^-)|B(v)\rangle\;, \label{deb} \end{equation} where $|B(v)\rangle$ is the initial state of the $B$ meson. The Feynman rules for the path-ordered exponential are an eikonal line in the direction $n$ with the propagator $1/(n\cdot l)$ and the vertex $n^\mu$. The effects of the variation of the $b$ quark momentum from the mass shell, determined by $f(k^+)$, are nonperturbative. Because the variation is of order $\Lambda_{\rm QCD}$, it is appropriate to expand the nonperturbative corrections contained in $f(k^+)$ in terms of its moments \cite{N}: \begin{eqnarray} & &\int dk^+ f(k^+)=\langle B(v)|{\bar h}_v(0)h_v(0)|B(v)\rangle=1\;, \label{mo0}\\ & &\int dk^+ k^+ f(k^+)= \langle B(v)|{\bar h}_v(0)iD^+ h_v(0)|B(v)\rangle=0\;, \label{mo1}\\ & &\int dk^+ k^{+2} f(k^+)= \langle B_v|{\bar h}_v(0)(iD^+)^2h_v(0)|B(v)\rangle\equiv -\lambda_1/6\;, \label{mo2} \end{eqnarray} where Eq.~(\ref{mo0}) is the normalization of $f(k^+)$, Eq.~(\ref{mo1}) is the consequence of the equation of motion for $h_v$, and Eq.~(\ref{mo2}) defines the HQET parameter $\lambda_1$. Equation (\ref{mo1}) implies that the $O(1/M_b)$ nonperturbative correction to the semileptonic decays does not exist, as observed in HQET. That is, the quark-hadron duality, which states the equality of heavy hadron decay widths to heavy quark decay widths up to small difference of $O(1/M_b^2)$ from $\lambda_1$, is respected in the PQCD formalism. As argued before, the $B$ meson structure function allows a variable $b$ quark momentum, whose source is the effect of the light degrees of freedom in the $B$ meson. This effect demands the consideration of $B$ meson, instead of $b$ quark, decays, and thus the $B$ meson kinematics. Therefore, we transform the structure function $f(k^+)$ into the usual $B$ meson distribution function $f_B(z)$, $z=p_b^+/P_B^+$, which describes the probability to find a $b$ quark with the plus momentum $zP_B^+$ \cite{N}. The relation between $f(k^+)$ and $f_B(z)$ cab be extracted from their moments by employing the variable transformation, \begin{equation} k^+=p_b^+-M_bv^+=\frac{zM_B-M_b}{\sqrt{2}}\;. \label{var} \end{equation} Equation (\ref{mo0}) leads to the normalization of $f_B(z)$, \begin{equation} \int_0^1 dz f_B(z)=1\;,\;\;\;\; f_B(z)\equiv \frac{M_B}{\sqrt{2}}f\left(\frac{zM_B-M_b} {\sqrt{2}}\right)\;. \label{nor} \end{equation} Equation (\ref{mo1}) gives \begin{eqnarray} \int dz (1-z) f_B(z)=\left(1-\frac{M_b}{M_B}\right)\int dz f_B(z) =\frac{{\bar\Lambda}}{M_B} +O\left(\frac{\Lambda^2_{\rm QCD}}{M_B^2}\right)\;, \label{ke} \end{eqnarray} into which Eq.~(\ref{nor}) has been inserted. Applying a similar manipulation to Eq.~(\ref{mo2}), we arrive at \begin{equation} \int dz (1-z)^2 f_B(z)=\frac{1}{M_B^2}\left({\bar\Lambda}^2- \frac{\lambda_1}{3}\right) +O\left(\frac{\Lambda^3_{\rm QCD}}{M_B^3}\right)\;. \label{ke1} \end{equation} The HQET parameters ${\bar\Lambda}$ and $\lambda_1$ satisfy the relation \begin{eqnarray} {\bar\Lambda}=M_B-M_b+\frac{\lambda_1}{2M_b}\;. \label{mass} \end{eqnarray} Therefore, the heavy hadron kinematics is introduced into the PQCD formalism rigorously via Eqs.~(\ref{deb})-(\ref{ke1}). It is then clear that the PQCD approach is equivalent to HQET in the treatment of nonperturbative corrections, and the moment ${\bar\Lambda}/M_B$ is attributed to the replacement of the $b$ quark mass by the $B$ meson mass. An advantage of the $B$ meson kinematics is that it provides the correct kinematic bounds. Take the lepton energy spectrum as an example. Using the $b$ quark kinematics, the maximal lepton energy $M_b/2$ is smaller than the correct value $M_B/2$. The $B$ meson distribution function must be determined by means outside the PQCD regime. For its functional form, we propose \cite{ab} \begin{equation} f_B(z)=N\frac{z(1-z)^2}{[(z-a)^2+\epsilon z]^2}\;, \label{eq: aub} \end{equation} where the three parameters $N$, $a$ and $\epsilon$ are obtained from the best fit to experimental data. We have extracted $f_B(z)$ from the photon energy spectrum of the radiative decay $B\to X_s\gamma$, given by \cite{ag} \begin{equation} f_B(z)= \frac{0.02647 z(1-z)^2}{[( z - 0.95 )^2 + 0.0034 z]^2}\;. \label{bdf} \end{equation} The maximum of $f_B$ occurs at $z\sim 1$ as expected, since the $b$ quark carries most of the $B$ meson momentum. Substituting the above expression into Eqs.~(\ref{nor}), (\ref{ke}) and (\ref{ke1}), which relate the three parameters $N$, $a$ and $\epsilon$ to the HQET parameters, we obtain ${\bar\Lambda}=0.65$ GeV and $\lambda_1=-0.71$ GeV$^2$. As to the distribution functions for other $b$-hadrons, we assume the parametrizations of the $B_s$ meson and $\Lambda_b$ baryon distribution functions, $f_{B_s}(z)$ and $f_{\Lambda_b}(z)$, respectively, the same as Eq.~(\ref{eq: aub}). The moments of $f_{B_s(\Lambda_b)}$ need to be treated as free parameters, since they have not been determined yet. The relations of the parameters $N$, $a$ and $\epsilon$ to the first three moments of $f_{B_s(\Lambda_b)}$ are similar to Eqs.~(\ref{nor}), (\ref{ke}) and (\ref{ke1}): \begin{eqnarray} & &\int_0^1 f_{B_s(\Lambda_b)}(z) dz=1\;, \nonumber\\ & &\int_0^1 (1-z)f_{B_s(\Lambda_b)}(z) dz= \frac{{\bar\Lambda}_{B_s(\Lambda_b)}} {M_{B_s(\Lambda_b)}}+O(\Lambda^2_{\rm QCD}/M_{B_s(\Lambda_b)}^2)\;, \nonumber \\ & &\int_0^1 (1-z)^2f_{B_s(\Lambda_b)}(z) dz=\frac{1}{M^2_{B_s(\Lambda_b)}} \left({\bar\Lambda}^2_{B_s(\Lambda_b)} -\frac{1}{3}\lambda_{1B_s(\Lambda_b)}\right) +O(\Lambda^3_{\rm QCD}/{M_{B_s(\Lambda_b)}^3}), \label{eq: ap} \end{eqnarray} with \begin{eqnarray} & &\lambda_{1B_s(\Lambda_b)}=-6 \langle B_s(\Lambda_b)|{\bar b}(iD_\perp)^2b|B_s(\Lambda_b)\rangle \nonumber\\ & &M_{B_s(\Lambda_b)}=M_b+{\bar\Lambda}_{B_s(\Lambda_b)}- \frac{\lambda_{1B_s(\Lambda_b)}}{2M_b}\;. \label{eq: ar} \end{eqnarray} Note that only $\lambda_{1B_s}$ and $\lambda_{1\Lambda_b}$ are free parameters. Because of the parameters ${\bar\Lambda}=0.65$ GeV and $\lambda_1=-0.71$ GeV$^2$ and $M_B=5.279$ GeV, the $b$ quark mass $M_b=4.551$ GeV is fixed by Eq.~(\ref{mass}). Substituting this $M_b$ into Eq.~(\ref{eq: ar}) with $M_{B_s}=5.369$ GeV and $M_{\Lambda_b}=5.621$ GeV, ${\bar\Lambda}_{B_s}$ and ${\bar\Lambda}_{\Lambda_b}$ will be derived, if $\lambda_{1B_s}$ and $\lambda_{1\Lambda_b}$ are chosen. Using the values of ${\bar\Lambda}_{B_s(\Lambda_b)}$ and $\lambda_{1B_s(\Lambda_b)}$, combined with the normalization of $f_{B_s(\Lambda_b)}$, we obtain the corresponding parameters $N$, $a$ and $\epsilon$. In the numerical analysis we shall assume that the HQET parameters $\lambda_1$ are the same for all $b$-hadrons, {\it i.e.}, $\lambda_{1B_s}=\lambda_{1\Lambda_b}=\lambda_1=-0.71$ GeV$^2$, which correspond to ${\bar\Lambda}_{B_s}=0.74$ GeV and ${\bar\Lambda}_{\Lambda_b}=0.99$ GeV. The $B_s$ meson and $\Lambda_b$ baryon distribution functions are then given by \begin{eqnarray} f_{B_s}(z)&=& \frac{0.02279 z(1-z)^2}{[( z - 0.93 )^2 + 0.0043 z]^2}\;, \label{bds}\\ f_{\Lambda_b}(z)&=& \frac{0.02095 z(1-z)^2}{[( z - 0.89 )^2 + 0.0068 z]^2}\;. \label{bdb} \end{eqnarray} For the $B_u$ meson distribution function, the reasoning is a bit different. In fact, we can not distinguish the $B_u$ and $B_d$ meson distribution functions, if $f_B$ is extracted from the photon energy spectrum of the $B\to X_s \gamma$ decay, which acquires contributions from both mesons \cite{C}. To explain the $B_u$ meson lifetime in Sec.~VI, we shall vary $f_B$ slightly: \begin{eqnarray} f_{B_u}(z)= \frac{0.03358 z(1-z)^2}{[( z - 0.96 )^2 + 0.0034 z]^2}\;, \label{bdu} \end{eqnarray} which corresponds to ${\bar\Lambda}_{B_u}=0.65$ GeV and $\lambda_{1B_u}=-0.81$ GeV$^2$, about 10\% deviation from $\lambda_1$ for the $B_d$ meson. \section{FACTORIZATION OF NONLEPTONIC DECAYS} The factorization theorem for nonleptonic $B$ meson decays is more complicated than the semileptonic case. Nonleptonic decays involve three scales: the $W$ boson mass $M_W$, at which the matching condition of the effective Hamiltonian to the full Hamiltonian is defined, the characteristic scale $t$, which reflects the specific dynamics of a decay mode, and the $b$ quark transverse momentum $p_\perp$, which serves as a factorization scale. Below the factorization scale, dynamics is regarded as being purely nonperturbative, and absorbed into the $B$ meson structure function, if it is soft, or into jet functions associated with light energetic final-state quarks, if it is collinear. Above the factorization scale, PQCD is reliable, and radiative corrections are absorbed into a hard $b$ quark decay amplitude characterized by $t$, or into a harder function characterized by $M_W$. In this section we shall demonstrate how to construct the three-scale factorization theorem for nonleptonic $B$ meson decays \cite{ah}. Consider a nonleptonic $b$ quark decay $b\to cq{\bar q}'$ through a $W$ boson emission up to $O(\alpha_{s})$ virtual gluon corrections. The sum of the full diagrams does not possess ultraviolet divergences because of the current conservation and the presence of the $W$ boson propagator. Reexpress a full diagram into two terms, where the first term is obtained by shrinking the $W$ boson line into a point, and the second term is the difference of the full diagram and the first term. The first term corresponds to a local four-fermion operator $({\bar q} q')({\bar c} b)$ appearing in the full or effective Hamiltonian with $({\bar q} q')={\bar q} \gamma_\mu(1-\gamma_5)q'$ the $V-A$ current, and is absorbed into a decay amplitude $H'(p_b,p_j,\mu)$, $p_j$ being the outgoing quark momenta. $H'$ is characterized by momenta smaller than the $W$ boson mass $M_W$, since gluons in $H'$ do not ``see" the $W$ boson. Note that the factorization in $H'$, which still contains the contributions characterized by the hadronic scale, is not complete yet. The second term, characterized by momenta of order $M_W$ due to the subtraction term, is absorbed into the harder function $H_r(M_W,\mu)$. Following the steps that lead to Eq.~(\ref{fas}), we derive the factorization formula for the $b$ quark decay width \cite{ah}, \begin{equation} \Gamma=H_r(M_W,\mu)\times H'(p_b,p_j,\mu)\;, \label{fa1} \end{equation} Though the full diagrams are ultraviolet finite, the factorization introduces the dependence on the renormalization scale $\mu$ into $H_r$ and $H'$, because the diagrams with the $W$ boson line shrunk are divergent. We then investigate infrared divergences from the diagrams contained in $H'$. These diagrams are basically similar to those in Fig.~1, where the four-fermion vertices have been adopted, and the final-state quark may represent the $c$, $q$, or ${\bar q}'$ quark. At this stage real gluon corrections are taken into account. If the final-state quark is $c$, the discussion the same as in Sec.~II leads to the $B$ meson structure function defined by Eq.~(\ref{deb}), which is thus universal. If the final-state quark is $q$, the loop integrals for Figs.~1(c) and 1(d) are written as \begin{eqnarray} ({\not p}_j+m)\Sigma^{'(c)}&=&-ig^2C_F\mu^{2\epsilon} \int\frac{d^{4-2\epsilon}l}{(2\pi)^{4-2\epsilon}}({\not p}_j+m) \gamma_\mu\frac{{\not p}_j-\not l+m}{(p_j-l)^2-m^2}\gamma^\mu \frac{{\not p}_j+m}{p_j^2-m^2} \frac{1}{l^2} \nonumber\\ & &\times \delta(p_j^2-m^2)\;, \label{3a}\\ ({\not p}_j+m)\Sigma^{'(d)}&=&-g^2C_F\mu^{2\epsilon} \int\frac{d^{4-2\epsilon}l}{(2\pi)^{4-2\epsilon}} \frac{{\not p}_j+m}{p_j^2-m^2} \gamma_\mu({\not p}_j-\not l+m)\gamma^\mu\frac{{\not p}_j+m}{p_j^2-m^2} \nonumber\\ & &\times 2\pi\delta(l^2)\delta((p_j-l)^2-m^2)\;, \label{3b} \end{eqnarray} with $p_j$ and $m$ the final-state quark momentum and mass, respectively. Performing the loop integrations directly, we obtain \begin{eqnarray} \Sigma^{'(c)}&=&\frac{\alpha_s}{2\pi}C_F\left(\frac{1}{-\epsilon} -\frac{3}{2}\ln\frac{\mu^2}{m^2}\right)\delta(p_j^2-m^2)\;, \label{3a3}\\ \Sigma^{'(d)}&=&-\frac{\alpha_s}{\pi}C_F \frac{(\mu^2p_j^2)^\epsilon}{(p_j^2-m^2)^{1+2\epsilon}}\;. \label{3b3} \end{eqnarray} The terms finite as $m\to 0$ have been dropped and the ultraviolet pole has been subtracted in Eq.~(\ref{3a3}). Similarly, Eq.~(\ref{3b3}) is reexpressed, following Eq.~(\ref{ind0}), as \begin{eqnarray} \Sigma^{'(d)}=-\frac{\alpha_s}{2\pi}C_F\frac{1}{-\epsilon} \delta(p_j^2-m^2)\;. \label{3b2} \end{eqnarray} Obviously, the soft poles $1/(-\epsilon)$ disappear in the sum of Eqs.~(\ref{3a3}) and (\ref{3b2}) as expected. For $m\not= 0$, there are no other infrared divergences, the same as in the $c$ quark case. However, an additional infrared divergence, {\it i.e.}, the collinear divergence, appears when the invariant mass $p_j^2$ of the final-state quark is equal to $m^2$ and vanishes, which remains even in the sum of Eqs.~(\ref{3a3}) and (\ref{3b2}). This divergence comes from the region with the loop momentum $l$ being parallel to $p_j$ and not small, for which both $(p_j-l)^2$ and $l^2$ approach zero. Note that $p_j^2$ vanishes only in the kinematic end-point region. To absorb this end-point singularity, we introduce a jet function associated with the energetic outgoing light quark. Figures 1(c) and 1(d) are then trivially factorized into the jet function. For Figs.~1(e) and 1(f) with a gluon attaching the $b$ quark and the $q$ quark, the discussion is the same as for the $c$ quark case, if the $q$ quark is massive: there are no collinear divergences, and soft divergences cancel, as the two diagrams are convoluted with the $B$ meson structure function. If the $q$ quark is massless, the loop integrals for Figs.~1(e) and 1(f) are written as \begin{eqnarray} {\not p}_j\Sigma^{'(e)}&=&-ig^2C_F\mu^{2\epsilon} \int\frac{d^{4-2\epsilon}l}{(2\pi)^{4-2\epsilon}}{\not p}_j \gamma_\mu\frac{{\not p}_j-\not l}{(p_j-l)^2} \Gamma_4\frac{{\not p}_b-\not l+M_b}{(p_b-l)^2-M_b^2}\gamma^\mu \frac{1}{l^2}\delta(p_j^2)\;, \label{3c}\\ {\not p}_j\Sigma^{'(f)}&=&-g^2C_F\mu^{2\epsilon} \int\frac{d^{4-2\epsilon}l}{(2\pi)^{4-2\epsilon}} \frac{{\not p}_j}{p_j^2}\gamma_\mu({\not p}_j-\not l) \Gamma_4\frac{{\not p}_b-\not l+M_b}{(p_b-l)^2-M_b^2}\gamma^\mu 2\pi\delta(l^2)\delta((p_j-l)^2)\;, \label{3d} \end{eqnarray} where $\Gamma_4$ represents the four-fermion vertex. Similarly, collinear divergences from $l$ parallel to $p_j$ with $p_j^2\to 0$ exist and remain after summing $\Sigma^{'(e)}$ and $\Sigma^{'(f)}$. These collinear divergences should be absorbed into the jet function associated with the $q$ quark. The factorization of Figs.~1(e) and 1(f) in the collinear region requires an eikonal approximation. For convenience, we assume that $p_j$ is in the plus direction at the kinematic end point. When $l$ is parallel to $p_j$, only $\gamma_-$ in the gamma matrices $\gamma_\mu$ contributes, indicating that only $\gamma^-$ in $\gamma^\mu$ contributes, and that $\not l$ in the numerator of the $b$ quark propagator is negligible. Using the equation of motion for the $b$ quark spinor, we have the eikonal approximation, \begin{equation} \frac{{\not p}_b+M_b}{(p_b-l)^2-M_b^2}\gamma^\mu =\frac{2p_b^\mu}{l^2-2p_b\cdot l} \approx\frac{p_b^-\delta^{-\mu}}{-p_b\cdot l} =-\frac{n^\mu}{n\cdot l}\;. \label{jfa} \end{equation} That is, the collinear gluon decouples from the $b$ quark, and attaches the eikonal line in the direction $n$ defined in Sec.~II. Using Eq.~(\ref{jfa}), Eqs.~(\ref{3c}) and (\ref{3d}) reduce to \begin{eqnarray} \Sigma^{'(e)}_{\rm coll}&=&ig^2C_F\mu^{2\epsilon} \int\frac{d^{4-2\epsilon}l}{(2\pi)^{4-2\epsilon}} \gamma_\mu\frac{{\not p}_j-\not l}{(p_j-l)^2} \frac{n^\mu}{n\cdot l}\frac{1}{l^2}\delta(p_j^2)\;, \label{3c1}\\ \Sigma^{'(f)}_{\rm coll}&=&g^2C_F\mu^{2\epsilon} \int\frac{d^{4-2\epsilon}l}{(2\pi)^{4-2\epsilon}} \frac{1}{p_j^2}\gamma_\mu({\not p}_j-\not l) \frac{n^\mu}{n\cdot l}2\pi\delta(l^2)\delta((p_j-l)^2)\;. \label{3d1} \end{eqnarray} Since the eikonal approximation in Eq.~(\ref{jfa}) also holds for a massless quark propagator with $M_b=0$, the factorization of the other set of diagrams with a gluon attaching the $q$ and ${\bar q}'$ quarks in the collinear region is similar: the collinear gluon decouples from the ${\bar q}'$ quark and attaches the eikonal line in the direction $n$. Extending the above analysis to all orders, the jet function is defined as the collection of infinite many gluon exchanges among the $q$ quark and the eikonal lines in the direction $n$ on both sides of the final-state cut. Note that the self-energy corrections to the eikonal line are excluded. The above conclusion applies to the other final-state ${\bar q}'$ quark: if the ${\bar q}'$ quark is massless, there exist collinear divergences at the kinematic end point, and another jet function needs to be introduced. With the jet functions and the $B$ meson structure function defined in Eq.~(\ref{deb}), all the infrared divergences in nonleptonic $B$ meson decays are extracted and factorized appropriately. The hard $b$ quark decay amplitude $H(t,\mu)$ is then defined as the difference of $H'$ and those diagrams which have been absorbed into the jet functions and the $B$ meson structure funciton, and thus calculable in perturbation theory. Following Eq.~(\ref{fas}), we obtain the factorization of $H'$, \begin{equation} H'(p_b,p_j,\mu)=H(t,\mu)\times \prod_jJ_j(p_j,\mu) \times f(k^+,\mu)\;, \label{fa2} \end{equation} where the index $j$ runs over the light final-state quarks. Note that a trace of $H$ and $J_j$ is necessary, since the jet function carries the spin structure of the corresponding final-state quark. Combining Eqs.~(\ref{fa1}) and (\ref{fa2}), the three-scale factorization formula for nonleptonic $B$ meson decay widths is written as \begin{equation} \Gamma=H_r(M_W,\mu)\times H(t,\mu)\times \prod_jJ_j(p_j,\mu) \times f_B(z,\mu)\;, \label{fat} \end{equation} where the structure function $f(k^+,\mu)$ has been transformed into the distribution function $f_B(z,\mu)$ \cite{N}. The $\mu$ dependence will disappear after performing a RG analysis, since a decay width is scale-independent. Note that the $\mu$ dependence of the $B$ meson distribution function for the semileptonic decays was not considered. We shall show in the next section that such a $\mu$ dependence is not necessary in our formalism. \section{LOGARITHMIC SUMMATIONS} To regularize the collinear divergences, we associate extra transverse momentum $p_\perp$, originating from the initial $b$ quark, with the final-state quarks, which renders them off-shell by $p_\perp$. This transverse momentum can be regarded as the factorization scale, above which RG evolutions are reliable, and below which dynamics is absorbed into the $B$ meson distribution function and the jet functions. We work in the impact parameter $b$ space, which is the Fourier conjugate variable of $p_{\perp}$. Radiative corrections to nonleptonic decays then produce two types of large logarithms, $\ln(M_W/t)$ and $\ln(tb)$. Especially, the double logarithms $\ln^2({\bar p}_jb)$, ${\bar p}_j$ being the large longitudinal component of the final-state quark momentum $p_j$, appear in the jet functions at the kinematic end points. These logarithmic corrections should be summed to all orders using RG equations and the resummation technique \cite{CS} in order to improve perturbative expansions. The summation of the logarithms $\ln(M_W/t)$ is identified as the Wilson coefficients in the effective Hamiltonian, which describes the evolution from the characteristic scale $M_W$ of the harder function to the characteristic scale $t$ of the hard amplitude. The summation of the logarithms $\ln(tb)$ leads to the evolution from $t$ to the factorization scale $1/b$. The resummation of the double logarithms $\ln^2({\bar p}_jb)$ for a jet with a small invariant mass is written, in the $b$ space, as \cite{ab,af} \begin{equation} J_j({\bar p}_j,b,\mu)=J_j(b,\mu)\exp[-2s({\bar p}_j,b)]\;. \label{sdc} \end{equation} The Sudakov exponent $s$ is given by \begin{equation} s({\bar p}_j,b)=\int_{1/b}^{{\bar p}_j}\frac{d \mu}{\mu} \left[\ln\left(\frac{{\bar p}_j}{\mu} \right)A(\alpha_s(\mu))+B(\alpha_s(\mu))\right]\;, \label{fsl} \end{equation} where the factors $A$ and $B$ are expanded as \begin{eqnarray} A(\alpha_s)&=&A^{(1)}\frac{\alpha_s}{\pi}+A^{(2)} \left(\frac{\alpha_s}{\pi}\right)^2\;, \nonumber \\ B(\alpha_s)&=&\frac{2}{3}\frac{\alpha_s}{\pi}\ln\left(\frac{e^{2\gamma-1}} {2}\right)\;, \end{eqnarray} in order to take into account the next-to-leading-logarithm summation. We adopt the one-loop running coupling constant, \begin{equation} \frac{\alpha_s(\mu)}{\pi}=\frac{1}{\beta_1\ln(\mu^2/\Lambda^2)}\;. \end{equation} The above coefficients $A^{(1)}$, $A^{(2)}$ and $\beta_1$ are \begin{eqnarray} \beta_{1}=\frac{33-2n_{f}}{12}\;,\;\;\; A^{(1)}=\frac{4}{3}\;, \;\;\; A^{(2)}=\frac{67}{9}-\frac{\pi^{2}}{3}-\frac{10}{27}n_f +\frac{8}{3}\beta_{1}\ln\left(\frac{e^{\gamma}}{2}\right)\; , \label{12} \end{eqnarray} where $n_f=5$ is the number of quark flavors, and $\gamma$ the Euler constant. The Sudakov factor $\exp(-2s)$ exhibits strong suppression at large $b$, and approaches unity as ${\bar p}_j< 1/b$. In this region the final-state quarks are regarded as being highly off-shell, and absorbed into the hard amplitude. Hence, double logarithms do not exist, {\it i.e.}, $\exp(-2s)\to 1$, and it is not necessary to introduce a jet function. For the detailed derivation of Eq.~(\ref{sdc}), refer to \cite{af,L1}. The large scale ${\bar p}_j$ will be chosen as the sum of the longitudinal components of $p_j$, {\it i.e.}, ${\bar p}_j=p_j^++p_j^-$ in the factorization formulas presented in the next section. The initial condition $J_j(b,\mu)$ of the resummation still contains single logarithms $\ln(b\mu)$, which are summed by RG equations. The RG solution of the jet function is given by \begin{eqnarray} J_j(b,\mu)=J_j^{(0)}\exp\left[-\int_{1/b}^\mu\frac{d{\bar\mu}} {\bar\mu}\gamma_j(\alpha_s({\bar\mu}))\right]\;, \label{jet} \end{eqnarray} with the anomalous dimension $\gamma_j$. It is more convenient to compute $\gamma_j$ in the axial gauge $n\cdot A=0$, since the eikonal lines in the direction $n$ disappear. The $n$ dependence goes into the gluon propagator $(-i/l^2)N^{\mu\nu}(l)$, with \begin{equation} N^{\mu\nu}=g^{\mu\nu}-\frac{n^\mu l^\nu+n^\nu l^\mu}{n\cdot l}\;. \label{gp} \end{equation} The lowest-order self-energy correction to the final-state quark is written as \begin{eqnarray} \Sigma_q=-ig^2C_F\mu^{2\epsilon}\int\frac{d^{4-2\epsilon}l} {(2\pi)^{4-2\epsilon}}\gamma_\mu \frac{{\not p}_j-{\not l}}{(p_j-l)^2}\gamma_\nu \frac{N^{\mu\nu}}{l^2}\frac{{\not p}_j}{p_j^2}\;, \label{anoj} \end{eqnarray} which is obtained by replacing the gluon propagator $-ig^{\mu\nu}/l^2$ in Eq.~(\ref{3a}) by $-iN^{\mu\nu}/l^2$. Extracting the ultraviolet poles from the above loop integral, we find that the $g^{\mu\nu}$ term and the second term in Eq.~(\ref{gp}) give $-\alpha_s C_F/(4\pi)\times (1/\epsilon)$ and $\alpha_s C_F/\pi\times (1/\epsilon)$, respectively. Their sum, $\Sigma_q=\alpha_s/\pi\times (1/\epsilon)$, leads to the quark anomalous dimension $\gamma_q=-\alpha_s/\pi$ in the axial gauge. $\gamma_j$ is twice of $\gamma_q$, {\it i.e.}, $\gamma_j=2\gamma_q$, because the self-energy corrections occur before and after the final-state cut. The initial condition $J_j^{(0)}$, with the large logarithms collected by the Sudakov factor and by the RG evolution, can be approximated by its tree-level expression, {\it i.e.}, the final-state cut. The RG solution of the $B$ meson distribution function is \begin{eqnarray} f_B(z,\mu)&=&f_B(z)\exp\left[-\int_{1/b}^\mu\frac{d{\bar\mu}} {\bar\mu}\gamma_S(\alpha_s({\bar\mu}))\right]\;. \label{sof} \end{eqnarray} where the initial condition $f_B(z)$ absorbs the nonperturbative dynamics below the scale $1/b$. The anomalous dimension $\gamma_S$ is also computed in the axial gauge, under which the path-ordered exponential in Eq.~(\ref{deb}) is equal to unity. Hence, we need to consider only the self-energy correction to the $h_v$ field in Fig.~1(a). The loop integral is written as \begin{eqnarray} \Sigma_v=-ig^2C_F\mu^{2\epsilon}\int\frac{d^{4-2\epsilon}l} {(2\pi)^{4-2\epsilon}} \frac{v_\mu}{v\cdot k}\frac{v_\nu}{v\cdot (k-l)}\frac{N^{\mu\nu}}{l^2}\;, \label{ano} \end{eqnarray} where the residual momentum $k$ approaches zero. Using the expansion \begin{equation} \frac{1}{v\cdot (k-l)}=-\frac{1}{v\cdot l}-\frac{v\cdot k}{(v\cdot l)^2}\;, \end{equation} Eq.~(\ref{ano}) reduces to \begin{eqnarray} \Sigma_v=ig^2C_F\mu^{2\epsilon}\int\frac{d^{4-2\epsilon}l} {(2\pi)^{4-2\epsilon}} \frac{v_\mu v_\nu}{(v\cdot l)^2}\frac{N^{\mu\nu}}{l^2}\;, \label{ano1} \end{eqnarray} which is the same as $\Sigma^{(a)}_{\rm soft}$ in Eq.~(\ref{eik}) but with the gluon propagator $-ig^{\mu\nu}/l^2$ replaced by $-iN^{\mu\nu}/l^2$. Therefore, Eq.~(\ref{deb}) indeed generates the factor $1+\Sigma^{(a)}_{\rm soft}$ in Eq.~(\ref{fas}) at first order. It is easy to find that the contribution from the second term in Eq.~(\ref{gp}) vanishes, and that the $g^{\mu\nu}$ term gives $\Sigma_v=\alpha_s C_F/(2\pi)\times (1/\epsilon)$. Adding the self-energy corrections on both sides of the final-state cut, we derive the anomalous dimension $\gamma_S=-\alpha_s C_F/\pi$ of the $B$ meson distribution function. The RG solution of $H_r$ is \begin{equation} H_r(M_W,\mu)=H_r(M_W,M_W) \exp\left[\int_{\mu}^{M_W}\frac{d{\bar\mu}} {\bar\mu}\gamma_{H_r}(\alpha_s({\bar\mu}))\right]\;, \label{hrh} \end{equation} with $\gamma_{H_r}$ the anomalous dimension of $H_r$. The initial condition $H_r(M_W,M_W)$ can be safely approximated by its lowest-order expression $H^{(0)}_r=1$, since the large logarithms $\ln(M_W/\mu)$ have been organized into the exponential, which is identified as the Wilson coefficient $c(\mu)$, \begin{eqnarray} c(\mu)\equiv \exp\left[\int_\mu^{M_W}\frac{d{\bar\mu}}{\bar\mu} \gamma_{H_r}(\alpha_s({\bar\mu}))\right]\;. \label{wil} \end{eqnarray} For the explicit expressions of the Wilson coefficients, refer to \cite{aj}. At last, the anomalous dimension of the hard amplitude $H$ is given by $\gamma_H=-\gamma_{H_r}-\sum_j\gamma_j-\gamma_S$ because of the scale invariance of a decay width in the full theory. Applying the RG analysis, we obtain \begin{eqnarray} H(t,\mu)=H(t,t)\exp\Bigg\{-\int_{\mu}^t\frac{d{\bar\mu}}{\bar\mu} \left[\gamma_{H_r}(\alpha_s({\bar\mu}))+\sum_j\gamma_j(\alpha_s({\bar\mu}))+ \gamma_S(\alpha_s({\bar\mu}))\right]\Bigg\}\;. \label{h} \end{eqnarray} The scale $t$ will be chosen as the maximal relevant scales, \begin{equation} t = \max \left( {\bar p}_j, \frac{1}{b} \right)\;. \label{tsc} \end{equation} Since the large $b$ region is Sudakov suppressed by $\exp(-2s)$ as stated before, that is, $t$ remains as a hard scale, and the large logarithms $\ln(t/\mu)$ in $H$ have been grouped into the exponential, the initial condition $H(t,t)$ is calculable in perturbation theory. Hence, the Sudakov factor, though important only in the end-point region, improves the applicability of PQCD to inclusive nonleptonic heavy hadron decays. Substituting Eqs.~(\ref{sdc}), (\ref{jet}), (\ref{sof}), (\ref{hrh}) and (\ref{h}) into Eq.~(\ref{fat}), we derive the RG improved factorization formula for the nonleptonic $B$ meson decay widths, \begin{equation} \Gamma=c(t)H(t,t)f_B(z)\exp\Bigg\{-\sum_j\left[ 2s({\bar p}_j,b)+\int_{1/b}^t\frac{d{\bar\mu}}{\bar\mu} \gamma_j(\alpha_s({\bar\mu}))\right] -\int_{1/b}^t\frac{d{\bar\mu}}{\bar\mu} \gamma_S(\alpha_s({\bar\mu}))\Bigg\}\;, \label{main} \end{equation} where the cancellation of the $\mu$ dependences among the the convolution factors is explicit, and the two-stage evolutions from $1/b$ to $t$ and from $t$ to $M_W$ have been established. Note that Eq.~(\ref{main}) in fact denotes the integrand appearing in the factorization formula, and that the variables $t$ and $b$ will be integrated out. The initial conditions $J_j^{(0)}$ in Eq.~(\ref{jet}) have been incorporated into the evaluation of $H(t,t)$. We emphasize that the Wilson coefficient appears as a convolution factor in the three-scale factorization theorem, instead of a constant (once its argument $\mu$ is set to a common value for all decay modes) in the HQET approach. Since the perturbative factors depend on the hadron kinematics, they vary in different decay modes of a hadron and in different hadrons. It will be shown that these perturbative effects play an essential role in the explanation of the $b$-hadron lifetimes. In the semileptonic case there are not the Wilson coefficient $c(t)$ and the Sudakov factor $\exp(-2s)$ because of the absence of the double logarithms. Therefore, the large scales ${\bar p}_j$ do not exist, and $t$ is equal to $1/b$, for which the RG evolution factors governed by the anomalous dimensions $\gamma_j$ and $\gamma_S$ disappear. This is the reason we did not perform the RG analysis for the semileptonic decays in Sec.~II. \section{FACTORIZATION FORMULAS AND NUMERICAL ANALYSIS} After developing the factorization theorems, we present the factorization formulas for the $B$ meson semileptonic and nonleptonic decay widths. Define the momentum of the light degrees of freedom as $p=(p^+,0,{\bf p}_\perp)$. The $B$ meson is at rest with the mometum $P_B=M_B/\sqrt{2}(1,1,{\bf 0})$. The $b$ quark momentum is then written as $p_b=P_B-p=(zM_B/\sqrt{2},M_B/\sqrt{2},-{\bf p}_\perp)$, where the momentum fraction $z\equiv p_b^+/P_B^+=1-\sqrt{2}p^+/M_B$ is the same as that defined in Sec. III. The lepton and neutrino momenta involved in the semileptonic decays $B(P_{B}) \rightarrow X_{c}+l(p_l)+\bar{\nu}(p_{\nu})$ are expressed, in terms of light-cone coordinates, as \begin{equation} p_l=(p^+_l,p_l^-,0_{\perp})\;,\mbox{\ \ } p_{\nu}=(p_{\nu}^+,p_{\nu}^-,{\bf p}_{\nu \perp})\;, \end{equation} where the minus component $p_l^-$ vanishes for a massless lepton. For convenience, we adopt the scaling variables, \begin{equation} x=\frac{2E_l}{M_B}\;,\mbox{\ \ } y=\frac{q^2}{M_B^2}\;,\mbox{\ \ } y_0=\frac{2q_0}{M_B}\;, \label{dl} \end{equation} with the kinematic ranges, \begin{eqnarray} 2\sqrt{\alpha}\leq&x& \leq 1+\alpha-\beta\;, \nonumber\\ \alpha\leq &y& \leq \alpha+(1+\alpha-\beta-x) \frac{x+\sqrt{x^2-4\alpha}}{2-x-\sqrt{x^2-4\alpha}}\;, \nonumber\\ x+\frac{2(y-\alpha)}{x+\sqrt{x^2-4\alpha}}\leq &y_0&\leq 1+y-\beta\;, \label{kcs} \end{eqnarray} where $E_l$ is the lepton energy and $q \equiv p_l+p_{\nu}$ the lepton pair momentum. The constants $\alpha$ and $\beta$ are \begin{equation} \alpha \equiv \frac{M^2_l}{M^2_B}\;,\;\;\;\; \beta \equiv \frac{M^2_D}{M^2_B}\;, \end{equation} $M_l$ and $M_D$ being the lepton mass and the $D$ meson mass, respectively. $M_D$ arises as the minimal invariant mass of the decay product $X_c$. For the derivation of Eq.~(\ref{kcs}), refer to the Appendix. The factorization formula for the semileptonic decay width is given, in the $b$ space, by \cite{ai} \begin{equation} \frac{\Gamma_{\rm SL}}{\Gamma_0}=\frac{M_B^2}{2\pi}\int dxdydy_0 \int^1_{z_{\rm min}}{dz} \int_0^\infty db b f_B(z){\tilde J}_c(x,y,y_0,z,b)H(x,y,y_0,z)\;, \label{asb} \end{equation} with $\Gamma_0 \equiv (G_F^2/16\pi^3)|V_{cb}|^2M^5_B$. The momentum fraction $z$ approaches 1 as the $b$ quark carries the whole $B$ meson momentum in the plus direction. The minimum of $z$, determined by the condition $p_c^2 > M_c^2$, is \begin{equation} z_{\rm min} = \frac{\displaystyle\frac{y_0}{2}-y+\frac{M_c^2}{M_B^2}- \frac{x}{\sqrt{x^2-4\alpha}} \left[-\frac{y_0}{2}+\frac{y}{x}+ \frac{\alpha}{x}\right] } {\displaystyle 1-\frac{y_0}{2}-\frac{x}{\sqrt{x^2-4\alpha}} \left[-\frac{y_0}{2}+\frac{y}{x}+\frac{\alpha}{x} \right] }\;. \end{equation} The function ${\tilde J}_c$ denotes the Fourier transformation of the final-state cut on the $c$ quark line, which is in fact part of the hard amplitude. $J_c$ and the lowet-order $H$ in momentum space are \cite{ai} \begin{eqnarray} J_c&=&\delta(p_c^2-M_c^2)\;, \nonumber\\ &=& \delta\left( M_B^2\left\{z-(1+z)\frac{y_0}{2}+y+ \frac{x(1-z)}{\sqrt{x^2-4\alpha}} \left[-\frac{y_0}{2}+\frac{y}{x}+\frac{\alpha}{x}\right] \right \}-M_c^2-{\bf p}^2_{\perp}\right), \nonumber \\ & &\label{jc}\\ H&=&(p_b\cdot p_\nu)(p_l\cdot p_c)\;, \nonumber\\ &=&\left((y_0-x)\left\{1-\frac{(1-z)}{2} \left(1-\frac{x}{\sqrt{x^2-4\alpha}}\right)\right\} -\frac{(1-z)}{\sqrt{x^2-4\alpha}} (y-\alpha)\right)\nonumber\\ &&\times \left(\frac{x}{2}\left\{1+z+(1-z) \frac{\sqrt{x^2-4\alpha}}{x}\right\} -y-\alpha\right). \label{hi} \end{eqnarray} The universal $B$ meson distribution function $f_B$, determined from the $B \rightarrow X_s \gamma$ decay, has been given in Eq.~(\ref{bdf}), which minimizes the model dependence of our predictions. For the nonleptonic decays, we consider the modes $b\to c{\bar c}s$ and $b\to c{\bar u}d$. Ignoring the penguin operators, the effective Hamiltonian for the $b\to c{\bar c}s$ decay is \begin{equation} H_{\rm eff} = \frac{4G_{F}}{\sqrt{2}} V_{cb} V^{\ast}_{cs} [ \, c_{1}(\mu) O_{1}(\mu) + c_{2}(\mu) O_{2}(\mu) \, ]\;, \label{eff1} \end{equation} with $G_F$ the Fermi coupling constant, $V$'s the Cabibbo-Kabayashi-Maskawa (CKM) matrix elements, the four-quark operators $O_{1}=(\bar{s}b)(\bar{c}c)$ and $O_{2}=(\bar{c} b)(\bar{s}c)$, and the initial conditions $c_1(M_W)=1$ and $c_2(M_W)=0$. For the $b\to c{\bar u}d$ decay, $V_{cs}$ and the ${\bar c}$ and $s$ quark fields are replaced by $V_{ud}$ and the ${\bar u}$ and $d$ quark fileds, respectively. It is simpler to work with the operators $O_{\pm}=\frac{1}{2} (O_{2} \pm O_{1})$ and their corresponding coefficients $c_{\pm}(\mu) = c_{2}(\mu) \pm c_{1}(\mu)$, since they are multiplicatively renormalized. In the leading logarithmic approximation $c_{\pm}$ are given by \cite{aj} \begin{equation} c_{\pm}(\mu) = \left[ \frac{\alpha_{s}(M_{W})}{\alpha_{s}(\mu)} \right] ^{\frac{-6\gamma_{\pm}}{33-2n_{f}}} \;, \end{equation} with the constants $2\gamma_+=-\gamma_-=-2$ and $n_f=5$. To simplify the analysis, we route the transverse momentum $p_\perp$ of the $b$ quark through the outgoing $c$ quark as in the semileptonic case, such that the $c$ quark momentum is the same as before. For kinematics, we make the correspondence with the ${\bar c}$ (${\bar u}$) quark carrying the momentum of the massive (light) lepton $\tau$ ($e$ and $\mu$) and with the $s$ and $d$ quarks carrying the momentum of ${\bar\nu}$. The scaling variables are then defined exactly by Eq.~(\ref{dl}). The factorization formula for the nonleptonic decay widths is written, according to the three-scale factorization theorem in Sec. IV, as \cite{ai} \begin{eqnarray} \frac{\Gamma_{\rm NL}}{\Gamma_0} &=&\frac{M_B^2}{2\pi}\int dxdydy_0\int_{z_{\rm min}}^1 dz \int_0^\infty bdb \left[ \frac{N_c\!+\!1}{2} c_{+}^2(t) + \frac{N_c\!-\!1}{2} c_{-}^2(t) \right] \nonumber \\ & & \times f_B(z){\tilde J}_c(x,y,y_0,z,b)H(x,y,y_0,z) S({\bar p}_j,t,b)\;. \label{non} \end{eqnarray} The $B$ meson distribution function $f_B$, being universal, is the same as that for the semileptonic decays. The factor $S$ is the result of the Sudakov resummation and the RG evolutions, given by \begin{equation} S({\bar p}_j,t,b)=\exp\Bigg\{-\sum_j\left[ 2s({\bar p}_j,b)+\int_{1/b}^t\frac{d{\bar\mu}}{\bar\mu} \gamma_j(\alpha_s({\bar\mu}))\right] -\int_{1/b}^t\frac{d{\bar\mu}}{\bar\mu} \gamma_S(\alpha_s({\bar\mu}))\Bigg\}\;, \end{equation} with $j=s$ for the $b \rightarrow c\bar{c}s$ mode and $j={\bar u}$, $d$ for the $b \rightarrow c\bar{u}d$ mode. The explicit expressions of ${\bar p}_j$ are \begin{eqnarray} {\bar p}_s={\bar p}_d=\frac{M_B}{\sqrt{2}}(y_0-x)\;, \;\;\;\; {\bar p}_u=\frac{xM_B}{\sqrt{2}}\;. \end{eqnarray} The anomalous dimensions $\gamma_j=-2\alpha_s/\pi$ and $\gamma_S=-C_F\alpha_s/\pi$ have been computed in Sec. V. It has been found that the single-logarithm evolutions governed by $\gamma_j$ and $\gamma_S$ enhance the nonleptonic branching ratios and thus lower the semileptonic branching ratios \cite{ai}, consistent with the observation in \cite{bagan}. The above formalism can be generalized to the $B_s$ meson and $\Lambda_b$ baryon decays straightforwardly, for which the functional forms of the hard amplitudes $H$, the final-state cut $J_c$, the Wilson coefficients $c_\pm$, and the Sudakov evolution factor $S$ remain the same, since they are evaluated at the quark level. The only differences arise from the replacement of the $B_d$ meson mass $M_B$ by the $B_s$ meson mass $M_{B_s}$ and by the $\Lambda_b$ baryon mass $M_{\Lambda_b}$, and from the heavy hadron distribution functions given in Eqs.~(\ref{bds}) and (\ref{bdb}). We then proceed with the numerical analysis of the factorization formulas in Eqs.~(\ref{asb}) and (\ref{non}) for the semileptonic and nonleptonic $b$-hadron decays, respectively, choosing the CKM matrix elements $|V_{cs}|=|V_{ud}|=1.0$ and $|V_{cb}|=0.044$, and the masses $M_c=1.6$ GeV, $M_D=1.869$ GeV and $M_\tau=1.7771$ GeV. Our predictions are not sensitive to the change of $M_c$: the difference of the results for $M_c=1.5$ GeV from those for $M_c=1.6$ GeV is less than 5\%. We obtain the lifetimes $\tau(B_d)=1.56$ ps, $\tau(B_s)=1.46$ ps and $\tau(\Lambda_b)=1.22$ ps, or the ratios $\tau(B_s)/\tau(B_d)=0.94$ and $\tau(\Lambda_b)$/$\tau(B_d)=0.78$. The experimental data of the $b$-hadron lifetimes are summarized below. The lifetimes of the $B_d$ and $B_s$ mesons and of the $\Lambda_b$ baryon are $\tau(B_d)=(1.55\pm 0.03)$ ps, $\tau(B_s)=(1.47\pm 0.06)$ ps and $\tau(\Lambda_b)=(1.23\pm 0.05)$ ps, respectively, from the CERN $e^+e^-$ collider LEP measurements \cite{ad}. Recent CDF results yield $\tau(\Lambda_b)=(1.32\pm 0.16)$ ps \cite{al}. The lifetime ratios are then $\tau(B_s)/\tau(B_d)=0.95\pm 0.06$ and $\tau(\Lambda_b)$/$\tau(B_d)=0.79\pm 0.06$ from LEP, and $\tau(\Lambda_b)$/$\tau(B_d)=0.85\pm 0.11$ from CDF. Obviously, our predictions are almost the same as the central values of the LEP data. Employing the $B_u$ meson distribution function in Eq.~(\ref{bdu}), we obtain $\tau(B_u)=1.62$ ps or the ratio $\tau(B_u)/\tau(B_d)=1.04$, which is in agreement with the experimental data $\tau(B_u)=1.65\pm 0.04$ ps in \cite{C}. To test the sensitivity of our predictions to the variation of $\lambda_{1B_s}$ and $\lambda_{1\Lambda_b}$, we adopt the HQET parameters $\lambda_1^{\rm meson}\sim\lambda_1^{\rm baryon}=-0.4$ GeV$^2$ from QCD sum rules for $\lambda_{1B_s}$ and $\lambda_{1\Lambda_b}$ (with the correspondiong ${\bar\Lambda}_{B_s}=0.77$ GeV and ${\bar\Lambda}_{\Lambda_b}=1.03$ GeV). The lifetimes $\tau(B_s)=1.49$ ps and $\tau(\Lambda_b)=1.26$, or the ratios $\tau(B_s)/\tau(B_d)=0.96$ and $\tau(\Lambda_b)$/$\tau(B_d)=0.81$, are derived, which are still much smaller than those from the HQET approach. This test indicates that our results are insensitive to the variation of the distribution functions (less than 4\% under 40\% variation of the HQET parameters), and that the explanation of the experimental data is mainly due to the PQCD effects. To confirm this observation, we set all the $b$-hadron masses in the phase space factors $\Gamma_0$ and in the perturbative factors to the $b$ quark mass $M_b$ in Eqs.~(\ref{asb}) and (\ref{non}). That is, the difference among the factorization formulas for $b$-hadron decays resides only in the distribution functions. The lifetime ratios $\tau(B_s)/\tau(B_d)=1.01$ and $\tau(\Lambda_b)/\tau(B_d)=1.09$, which are even greater than unity, are obtained. Therefore, the perturbative contributions are indeed responsible for the low $b$-hadron lifetime ratios. The above results are aummarized in Table I. In Table I we also present the branching ratio of each mode in the $B_d$ and $B_s$ meson decays and in the $\Lambda_b$ baryon decays in terms of the quantities $r_{\tau\nu}=B(b\to c\tau{\bar \nu})/B(b\to cl{\bar \nu})$, $r_{ud}=B(b\to c{\bar u}d)/B(b\to cl{\bar \nu})$, $r_{cs}=BR(b\to c{\bar c}s)/B(b\to cl{\bar \nu})$, the semileptonic branching ratio $B_{\rm SL}$ and the average charm yield $\langle n_c\rangle$ per $b$-hadron decay. It is observed that our predictions of $B_{\rm SL}=10.16\%$ and $\langle n_c\rangle=1.17$ for the $B$ meson are consistent with the experimental data: $B_{\rm SL}=(10.19\pm 0.37)$ and $\langle n_c\rangle=(1.12\pm 0.05)$ from the CLEO group\cite{ak}, and $B_{\rm SL}=(11.12\pm 0.20)$ and $\langle n_c\rangle=(1.20\pm 0.07)$ from the LEP measurements \cite{LEP}. It is also observed that $B_{\rm SL}$, $\langle n_c\rangle$ and all $r$'s except $r_{ud}$ increase a bit with the $b$-hadron masses. It is interesting to examine these tendencies in future experiments. \section{CONCLUSION} In this paper we have developed the PQCD factorization theorems for inclusive $b$-hadron decays by carefully analyzing the infrared divergences in radiative corrections to $b$ quark decays. Radiative corrections characterized by the hadronic scale are absorbed into the heavy hadron distribution function or into the jet functions. Above the hadronic scale, perturbative contributions characterized by the $b$-hadron mass and the $W$ boson mass are absorbed into the hard $b$ quark decay amplitude and the harder function, respectively. Various large logarithmic corrections have been summed to all orders using the resummation technique and the RG equations, leading to the evolution factors among the three characteristic scales, such that the factorization formulas are $\mu$-independent. We have shown that the soft cancellation between virtual and real corrections demands the introduction of the heavy hadron distribution function, and thus the $b$-hadron kinematics into our formalism. Using the PQCD factorization theorems, we have been able to explain the lifetimes of the $b$-hadrons $B_d$, $B_u$, $B_s$ and $\Lambda_b$. It has been confirmed that our predictions are insensitive to the variation of the distribution functions as expected from the quark-hadron duality, and that the low $b$-hadron lifetime ratios are mainly attributed to the perturbative contributions. \vskip 0.5cm I thank Profs. H.Y. Cheng and H.L. Yu for useful discussions. This work was supported by the National Science Council of R.O.C. under Grant No. NSC 88-2112-M-006-013. \vskip 1.0cm \centerline{\bf APPENDIX} \vskip 0.5cm In this appendix we derive the phase space constraints for the decay $b\to cl{\bar\nu}$. The constraints for the decay $b\to c{\bar c}s$ are the same with the ${\bar c}$ and $s$ quarks corresponding to the $l$ and ${\bar\nu}$ leptons, respectively. The constraints for the decay $b\to c{\bar u}d$ is the massless lepton case of the results presented below. We derive the constraint of $x=2E_l/M_H$ from the energy and momentum conservations associated with the decay $H\to X_cl{\bar\nu}$, \begin{eqnarray} M_H&=&E_X+E_l+E_\nu\;, \nonumber\\ 0&=&{\bf p}_X+{\bf p}_l+{\bf p}_\nu\;. \label{con} \end{eqnarray} Because of $E_l^2=|{\bf p}_l|^2+M_l^2$, the minimum of $E_l$ occurs at the minimum of $|{\bf p}_l|$, $|{\bf p}_l|=0$, and thus the minimum of $x$ is given by \begin{equation} x_{\min}=\frac{2M_l}{M_H}\;. \label{xmin} \end{equation} To obtain the maximum of $x$, we find the maximum of $|{\bf p}_l|$ under Eq.~(\ref{con}), \begin{equation} M_H=\sqrt{({\bf p}_l+{\bf p}_\nu)^2+M_X^2}+ \sqrt{|{\bf p}_l|^2+M_l^2}+|{\bf p}_\nu|^2\;, \end{equation} into which ${\bf p}_X=-({\bf p}_\tau+{\bf p}_\nu)$ has been inserted. It is easy to observe that the maximum of $|{\bf p}_l|$ corresponds to ${\bf p}_\nu=0$, leading to \begin{equation} |{\bf p}_{l\max}|^2=\frac{(M_H^2+M_l^2-M_X^2)^2-4M_H^2M_l^2}{4M_H^2}\;, \end{equation} and \begin{equation} x_{\max}=1+\frac{M_l^2}{M_H^2}-\frac{M_X^2}{M_H^2}\;. \label{xmam} \end{equation} Combining Eqs.~(\ref{xmin}) and (\ref{xmam}), we obtain \begin{equation} \frac{2M_l}{M_H}\le x\le 1+\frac{M_l^2}{M_H^2}-\frac{M_D^2}{M_H^2}\;, \end{equation} where $M_X$ has been replaced by the $D$ meson mass $M_D$, the mass of the lightest charmed hadron. We then derive the constraint of the kinematic variable $y=q^2/M_H^2$ with \begin{equation} q^2=(p_l+p_\nu)^2=M_l^2+2(E_l-|{\bf p}_l|\cos\theta)E_\nu\;, \label{q2} \end{equation} where $\theta$ is the angle between the vectors ${\bf p}_l$ and ${\bf p}_\nu$. The minimum of $q^2$ occurs, as $\cos\theta=1$ and $E_\nu$ takes its minimal value under the constraint from Eq.~(\ref{con}), \begin{equation} M_H-E_l=\sqrt{|{\bf p}_l|^2+E_\nu^2+2|{\bf p}_l| E_\nu\cos\theta+M_X^2}+E_\nu\;. \label{cons2} \end{equation} Obviously, the minimum of $E_\nu$ is zero, which can be achieved by increasing $M_X$. We then have \begin{equation} (q^2)_{\min}=M_l^2\;. \label{ymin} \end{equation} On the other hand, the maximum of $q^2$ occurs, as $\cos\theta=-1$ and $E_\nu$ takes its maximal value under Eq.~(\ref{cons2}), which corresponds to the minimum of $M_X$. Setting $M_X=M_D$, we obtain \begin{equation} E_{\nu\max}=\frac{M_H^2+M_l^2-M_D^2-2M_HE_l} {2(M_H-E_l-\sqrt{E_l^2-M_l^2})}\;, \end{equation} and \begin{equation} (q^2)_{\max}=M_l^2+(M_H^2+M_l^2-M_D^2-2M_HE_l) \frac{E_l+\sqrt{E_l^2-M_l^2}} {M_H-E_l-\sqrt{E_l^2-M_l^2}}\;. \label{ymax} \end{equation} Combining Eqs.~(\ref{ymin}) and (\ref{ymax}), we have \begin{equation} \frac{M_l^2}{M_H^2}\le y\le \frac{M_l^2}{M_H^2} \left(1+\frac{M_l^2}{M_H^2}-\frac{M_D^2}{M_H^2}-x\right) \left(\frac{2}{2-x-\sqrt{x^2-4M_l^2/M_H^2}}-1\right)\;. \end{equation} At last, we derive the range of the variable $y_0=2q_0/M_H$ with \begin{equation} y_0=E_l+E_\nu\;, \end{equation} or equivalently, the range of $E_\nu$ under the constraints of Eqs.~(\ref{q2}) and (\ref{cons2}). Since $E_\nu$ decreases with $\cos\theta$ in order that $q^2$ maintains constant, the minimum of $E_\nu$ occurs at $\cos\theta=-1$. Equation (\ref{q2}) then gives \begin{equation} E_{\nu\min}=\frac{q^2-M_l^2}{2(E_l+\sqrt{E_l^2-M_l^2})}\;, \label{evm} \end{equation} and \begin{equation} y_{0\min}=x+\frac{2(y-M_l^2/M_H^2)}{x+\sqrt{x^2-4M_l^2/M_H^2}}\;. \label{ymi} \end{equation} It is easy to observet that $E_\nu$ increases as $M_X$ decreases. Setting $M_X=M_D$, and substituting the expression of $\cos\theta$ from Eq.~(\ref{q2}) into (\ref{cons2}), we obtain \begin{equation} E_{\nu\max}=\frac{M_H^2-2M_HE_l-M_D^2+q^2}{2M_H}\;, \label{evmx} \end{equation} and \begin{equation} y_{0\max}=1+y-\frac{M_D^2}{M_H^2}\;. \label{yma} \end{equation} Combining Eqs.~(\ref{ymi}) and (\ref{yma}), we derive \begin{equation} x+\frac{2(y-M_l^2/M_H^2)}{x+\sqrt{x^2-4M_l^2/M_H^2}} \le y_0\le 1+y-\frac{M_D^2}{M_H^2}\;. \end{equation} \newpage
\section{Introduction} The differences between the BCS scenario of superconductivity and superconductivity in high-$T_c$ materials are well accepted as experimental facts, although there is no theoretical consensus about their origin. One of the most convincing manifestations is the pseudogap, or a depletion of the single particle spectral weight around the Fermi level (see for example, \cite{Timusk}). Another transparent manifestation is the temperature and carrier density dependencies of the superfluid density in high-$T_c$ superconductors (HTSC) \cite{Corson,Panagopoulos,Mesot} which do not fit the canonical BCS behaviour. In particular the value of the zero temperature superfluid density is substantially less than the total density of doped carriers \cite{Emery.conference}. Currently there are many possible explanations for the unusual properties of HTSC. One of these is based on the nearly antiferromagnetic Fermi liquid model \cite{Pines}. Another explanation, proposed by Anderson, relies on the separation of spin and charge degrees of freedom. One more approach, which we will follow in this paper, relates the observed anomalies to precursor superconducting fluctuations. Different authors argue that alternative types of superconducting fluctuations are responsible for the pseudogap, e.g. \cite{Levin}, and the scenario based on the fluctuations of the order parameter phase suggested by Emery and Kivelson \cite{Emery}. The latter scenario we believe to be more relevant due to low superfluid density and practically 2D character of the conductivity in HTSC mentioned above. A microscopic 2D model which elaborates the abovementioned scenario \cite{Emery} has been studied in the papers \cite{Gusynin.JETP,Gusynin.new}. The results obtained show that the condensate phase fluctuations indeed lead to features which are experimentally observed in HTSC both in the normal and superconducting states \footnote{Of course, the contribution from the phase fluctuations need not be the only or even the major contribution.}. It is obvious, however, that the present treatment of the phase fluctuations is incomplete due both to the oversimplified character of the model and the absence of an explanation for the more recent advanced experiments \cite{Corson,Panagopoulos,Mesot} on the temperature and doping dependencies of the superfluid density. It is well known, however, that the theoretical study of HTSC faces a lot of computational difficulties due to, for example, a non-conventional order parameter symmetry, complex frequency-momentum dependence of the effective quasiparticle attraction, general form of the quasiparticle dispersion law, etc. Therefore, in order to obtain analytical results, we have to date only considered nonretarded $s$-wave pairing in the absence of impurities. (The attempts to consider the retardation effects were done in \cite{Turkowski}.) Nevertheless, a discussion of the effect of impurities seems to be crucial for a realistic model of the HTSC. Indeed, it is known, that the itinerant holes in HTSC are created by doping which in turn introduces a considerable disorder into the system, for instance, from the random Coulomb fields of chaotically distributed charged impurities (doped ions) \cite{Pogorelov}. Thus one the purposes of the present paper is to study the model \cite{Gusynin.JETP,Gusynin.new} but in the presence of nonmagnetic impurities. In the theory of ``common metals'' the Fermi energy $\epsilon_{F}$ and the mean transport quasiparticle time $\tau_{tr}$ are independent quantities which are always assumed to satisfy the criterion $\epsilon_{F} \tau_{tr} \gg 1$. In HTSC which are ``bad metals'' \cite{Emery} both $\epsilon_{F}$ and $\tau_{tr}$ are dependent on the doping and the abovementioned criterion may fail \cite{Pogorelov}. As an illustration we refer to the remarkable linear dependence of the normal state resistivity \cite{Timusk} which implies indeed that $\epsilon_{F} \tau_{tr}$ may be $\sim 1$. It has been shown \cite{Pogorelov} that for strongly disordered metallic system superconductivity is absent if the scattering-to-pairing ratio exceeds a critical value and the existence of superconductivity in a finite range of doping if this ratio is not exceeded. We shall not study this case but rather consider here the more usual (and in some sense simple) situation, originally studied in the papers of Anderson \cite{Anderson} and Abrikosov-Gor'kov (AG) \cite{Abrikosov} (see also \cite{Abrikosov.book}), when the superconducting order is preexisting and the criterion $\epsilon_{F} \tau_{tr} \gg 1$ is satisfied. The Anderson theorem \cite{Anderson} states that in 3D the BCS critical temperature is unchanged in the presence of nonmagnetic impurities. However, as discussed in \cite{Gusynin.JETP}, the BCS critical temperature in 2D is the temperature $T^{\ast}$ at which the pseudogap opens, while the superconducting transition temperature transition is the temperature $T_{\rm BKT}$ of the Berezinskii-Kosterlitz-Thouless (BKT) transition. In contrast to the former, the latter is defined by a bare superfluid density (given by the delocalized carriers) which is dependent on (see below) the concentration of impurities. Thus in 2D case the superconducting transition temperature, $T_{\rm BKT}$ decreases with increasing impurity concentration. Thus, in the model under consideration, the relative size of the pseudogap phase, $(T^{\ast} - T_{\rm BKT})/T^{\ast}$ is larger in the presence of impurities than in the clean limit \cite{Gusynin.JETP}. Therefore it can be observed over a wider range of densities. The second result obtained is that the value of the zero temperature superfluid density is less than the total density of carriers (dopants), so that the presence of impurities may contribute into this diminishing and in its turn explain the experimental results \cite{Emery.conference}. Finally we attempt to interpret qualitatively the recent experiments on the temperature dependence of the superfluid density \cite{Corson,Panagopoulos} within our scenario. A brief overview of the paper follows: In Sec.~\ref{sec:model} we present the model and derive the main equations. In Sec.~\ref{sec:comparison} we compare the results obtained for clean and dirty limits. In particular we compare the values of $T_{\rm BKT}$, the relative sizes of the pseudogap region and the values of the bare superfluid density at $T = 0$ and $T$ close to $T_{\rho}$. In Sec.~\ref{sec:superfluid} an attempt is made to give an explanation for the experimental results \cite{Corson,Panagopoulos}. \section{Model and main equations} \label{sec:model} Our starting point is a continuum version of the two-dimensional attractive Hubbard model defined by the Hamiltonian \cite{Gusynin.JETP,Gusynin.new}: \begin{eqnarray} H = && \int d^2 r \left[ \psi_{\sigma}^{\dagger}(x) \left(- \frac{\nabla^{2}}{2 m} - \mu \right) \psi_{\sigma}(x) - V \psi_{\uparrow}^{\dagger}(x) \psi_{\downarrow}^{\dagger}(x) \psi_{\downarrow}(x) \psi_{\uparrow}(x) \right. \nonumber \\ && \left. + U_{imp} (\mbox{\bf r}) \psi_{\sigma}^{\dagger}(x) \psi_{\sigma}(x) \right], \label{Hamilton} \end{eqnarray} where $x= \mbox{\bf r}, \tau$ denotes the space and imaginary time variables, $\psi_{\sigma}(x)$ is a fermion field with spin $\sigma =\uparrow,\downarrow$, $m$ is the effective fermion mass, $\mu$ is the chemical potential, $V$ is an effective local attraction constant, and $U_{imp}(\mbox{\bf r})$ is the static potential of randomly distributed impurities; we take $\hbar = k_{B} = 1$. The model with the Hamiltonian (\ref{Hamilton}) is equivalent to a model with an auxiliary BCS-like pairing field which is given in terms of the Nambu variables as \begin{eqnarray} H = && \int d^2 r \left\{ \Psi^{\dagger}(x) \left[\tau_3 \left(- \frac{\nabla^{2}}{2 m} - \mu \right) \right. \right. \nonumber \\ && - \left. \left. \tau_{+}\Phi(x) - \tau_{-}\Phi^\ast(x) + \tau_{3} U_{imp}(\mbox{\bf r}) \right] \Psi(x) +\frac{|\Phi(x)|^2}{V} \right\}, \end{eqnarray} where $\tau_{\pm} = (\tau_{1} \pm i \tau_{2})/2$, $\tau_{3}$ are the Pauli matrices, and $\Phi(x)= V \Psi^\dagger(x)\tau_{-}\Psi(x)= V \psi_\downarrow(x) \psi_\uparrow (x)$ is the complex order field. Then the partition function can be presented as a functional integral over Fermi fields (Nambu spinors) and the auxiliary fields $\Phi$, $\Phi^{\ast}$. However, in contrast to the usual method, the modulus-phase parametrization $\Phi(x) = \rho(x) \exp[i \theta(x)]$ is necessary for the 2D model at finite temperatures (see \cite{Gusynin.JETP,Gusynin.new,Ramakrishnan} and references therein). To be consistent with this replacement one should also introduce the spin-charge variables for the Nambu spinors \begin{equation} \Psi(x) =\exp[i\tau_3\theta(x)/2]\Upsilon(x) \label{Nambu.phase} \end{equation} with $\Upsilon(x)$ the field operator for neutral fermions. From the Hamiltonian (\ref{Hamilton}), following \cite{Gusynin.JETP}, one can derive an effective one which is the Hamiltonian of the classical XY-model \begin{equation} H_{XY} = \frac{1}{2} J(\mu, T, \rho) \int d^2 r [\nabla \theta (x)]^2 \label{XY.Hamilton} \end{equation} where \begin{eqnarray} && J(\mu, T, \rho) = \frac{T}{16 m \pi^2} \sum_{n = -\infty}^{\infty} \int d^2 k \mbox{tr} [\tau_3 \langle \mathcal{ G}(i \omega_{n}, \mbox{\bf k}) \rangle] \nonumber \\ && + \frac{T}{32 m^2 \pi^2} \sum_{n = -\infty}^{\infty} \int d^2 k k^2 \mbox{tr} [ \langle \mathcal{ G}(i \omega_{n}, \mbox{\bf k}) \rangle \langle \mathcal{ G}(i \omega_{n}, \mbox{\bf k}) \rangle ] \label{J} \end{eqnarray} is the bare (i.e. unrenormalized by the phase fluctuations, but including pair breaking thermal fluctuations) superfluid stiffness. Here \begin{equation} \langle \mathcal{ G}(i \omega_{n}, \mbox{\bf k}) \rangle = - \frac{ (i \omega_{n} \hat{I} - \tau_{1} \rho ) \eta_{n} + \tau_{3} \xi(\mbox{\bf k})} {(\omega_{n}^{2} + \rho^{2}) \eta_{n}^{2} + \xi^{2}(\mbox{\bf k})} \label{Green.impurity} \end{equation} with \begin{equation} \eta_{n} = 1 + \frac{1}{2 \tau_{tr} \sqrt{\omega_{n}^{2} + \rho^{2}}}, \qquad \xi(\mbox{\bf k}) = \frac{\mbox{\bf k}^2}{2m} -\mu, \qquad \omega_{n} = \pi (2n+1)T \label{eta} \end{equation} is the AG \cite{Abrikosov.book} Green's function of neutral fermions averaged over a random distribution of impurities and written in the Nambu representation \cite{Evans,Rickayzen.book}. In writing (\ref{J}) we assumed that $ \langle \mathcal{ G}(i \omega_{n}, \mbox{\bf k}) \mathcal{ G}(i \omega_{n}, \mbox{\bf k}) \rangle \simeq \langle \mathcal{ G}(i \omega_{n}, \mbox{\bf k}) \rangle \langle \mathcal{ G}(i \omega_{n}, \mbox{\bf k}) \rangle$. This approximation, as shown by AG \cite{Abrikosov.book}, does not change the final result for $J$. Note also that the Green's function (\ref{Green.impurity}) is valid only when $\epsilon_{F} \tau_{tr} \gg 1$ which demands the presence of a well developed Fermi surface, which in turn implies that $\mu \simeq \epsilon_{F}$. Thus one cannot use the expression (\ref{Green.impurity}) in the so called Bose limit with $\mu < 0$ \cite{Gusynin.JETP}. On the other hand, Fermi surface can be formed even in the bad metals when Ioffe-Regel-Mott criterion proves to be fulfilled \cite{Pogorelov}. Substituting (\ref{Green.impurity}) into (\ref{J}), and using the inequalities $\mu \gg T, \rho$ to extend the limits of integration to infinity, one arrives at \begin{equation} J = \frac{\mu}{4 \pi} + \frac{T \mu }{4 \pi} \sum_{n=-\infty}^{\infty} \int_{-\infty}^{\infty} dx \left( \frac{1}{x^2 + (\omega_{n}^{2} + \rho^{2}) \eta_{n}^{2}} - \frac{2 \omega_{n}^{2} \eta_{n}^{2}} {[x^2 + (\omega_{n}^{2} + \rho^{2}) \eta_{n}^{2}]^2} \right) . \label{J.intermediate} \end{equation} Eq.(\ref{J.intermediate}) is formally divergent and demands special care due to the fact that one has to perform the integration over $x$ before the summation \cite{Abrikosov.book}. Finally one can formally cancel the divergence \cite{Abrikosov.book} to obtain \begin{equation} J = \frac{\mu \rho^{2} T}{4} \sum_{n = -\infty}^{\infty} \frac{1}{(\omega_{n}^{2} + \rho^{2}) \left[ \sqrt{\omega_{n}^{2} + \rho^{2}} + \frac{\displaystyle 1}{\displaystyle 2 \tau_{tr}} \right]}. \label{J.final} \end{equation} The temperature of the BKT transition for the XY-model Hamiltonian (\ref{XY.Hamilton}) is determined by the equation \begin{equation} T_{\rm BKT} = \frac{\pi}{2} J (\mu, T_{\rm BKT}, \rho(\mu, T_{\rm BKT})) . \label{BKT.temperature} \end{equation} The self-consistent calculation of $T_{\rm BKT}$ as a function of the carrier density $n_f = m \epsilon_{F}/\pi$ requires additional equations for $\rho$ and $\mu$, which together with (\ref{BKT.temperature}) form a complete set \cite{Gusynin.JETP}. When the modulus of the order parameter $\rho(x)$ is treated in the mean field approximation, the equation for $\rho$ takes the following form \cite{Gusynin.JETP} \begin{equation} \frac{2 \rho}{V} = \sum_{n = -\infty}^{\infty} \int \frac{d^2 k}{(2 \pi)^{2}} \mbox{tr} [\tau_1 \langle \mathcal{ G}(i \omega_{n}, \mbox{\bf k}) \rangle] , \label{rho.gap} \end{equation} which formally coincides with the gap equation of the BCS theory. This coincidence allows one to use the Anderson theorem \cite{Anderson} which states the dependence of $\rho(T)$ is the same as that for the clean superconductor and not affected by the presence of nonmagnetic impurities. It is important to recall that this theorem is, of course, valid only for the $s$-wave pairing and small disorder. There is, however, both physical and mathematical differences between the gap in the BCS theory and $\rho$ \cite{Gusynin.JETP,Gusynin.new}. In particular the temperature $T_{\rho}$ which is estimated by the condition $\rho = 0$ is not related to the temperature of the superconducting transition, but is interpreted as the temperature of pseudogap opening $T^{\ast}$ (see details in \cite{Gusynin.JETP}). The main point, which we would like only to stress here, is that due to the Anderson theorem \cite{Anderson} the value of $T_{\rho}$ does not depend on the presence of impurities, while the temperature $T_{\rm BKT}$, as we will show, is lowered. The chemical potential $\mu$ is defined by the number equation \begin{equation} \sum_{n = -\infty}^{\infty} \int \frac{d^2 k}{(2 \pi)^{2}} \mbox{tr} [\tau_3 \langle \mathcal{ G}(i \omega_{n}, \mbox{\bf k}) \rangle] = n_f. \label{number} \end{equation} Since we are interested in the high carrier density region the solution of (\ref{number}) is $\mu \simeq \epsilon_{F}$, so that in Eqs.(\ref{J.final}) - (\ref{rho.gap}) one can replace $\mu$ by $\epsilon_{F}$. Having the temperatures $T_{\rho}$ and $T_{\rm BKT}$ as functions of the carrier density one can build the phase diagram of the model \cite{Gusynin.JETP} which consists of three regions. The first one is the superconducting (here BKT) phase with $\rho \ne 0$ at $T < T_{\rm BKT}$. In this region there is algebraic order, or a power law decay of the $\langle\Phi^{\ast}\Phi\rangle$ correlations. The second region corresponds to the pseudogap phase ($T_{\rm BKT} < T < T_{\rho} $). In this phase $\rho$ is still non-zero but the correlations mentioned above decay exponentially. The third is the normal (Fermi-liquid) phase at $T > T_{\rho}$ where $\rho= 0$. Note that $\langle \Phi(x) \rangle= 0$ everywhere. While the given phase diagram was derived for the idealized 2D model, there are indications that even for as complicated layered systems as HTSC the value of the critical temperature for them may be well estimated using $T_{\rm BKT}$ \cite{Abrikosov.1997,LQSh.1999} even though the transition undoubtedly belongs to the 3D XY class. It was also pointed out in \cite{Abrikosov.1997} that a nonzero gap in the one-particle excitation spectrum can persists even without long-range order. \section{The comparison of clean and dirty limits} \label{sec:comparison} \subsection{Clean limit} The transport time $\tau_{tr}$ is infinite in the clean limit, so that \begin{equation} J(\epsilon_{F}, T, \rho(\epsilon_{F},T)) = \frac{\epsilon_{F} \rho^{2} T}{4} \sum_{n = -\infty}^{\infty} \frac{1}{(\omega_{n}^{2} + \rho^{2})^{3/2}} . \label{J.clean} \end{equation} Near $T_{\rho}$ one can obtain from (\ref{J.clean}) \begin{equation} J(\epsilon_{F}, T \to T_{\rho}^{-}, \rho \to 0) = \frac{7 \zeta(3)}{16 \pi^3} \frac{\rho^{2}}{T_{\rho}^{2}} \epsilon_{F}, \label{J.clean.expand} \end{equation} where $\zeta(x)$ is the zeta function. This expression must coincide with the result from \cite{Gusynin.JETP} which was derived using the opposite order for the summation and integration. Inserting the well-known dependence of $\rho(T)$ (see, for example \cite{Schrieffer}) \begin{equation} \rho^{2}(T \to T_{\rho}^{-}) = \frac{8 \pi^{2}}{7 \zeta(3)} T_{\rho}^{2} \left(1 - \frac{T}{T_{\rho}} \right) \label{rho.dependence} \end{equation} and then substituting (\ref{J.clean.expand}) into (\ref{BKT.temperature}) one obtains the following asymptotic expression for the BKT temperature in the clean limit for high carrier densities \cite{Halperin,Gusynin.JETP,Babaev} \begin{equation} T_{\rm BKT} = T_{\rho} \left(1 - \frac{4T_{\rho}}{\epsilon_{F}} \right), \qquad T_{\rm BKT} \lesssim T_{\rho}. \label{BKT.clean} \end{equation} In the high density limit one can also use the equation \begin{equation} T_{\rho} = \frac{\gamma}{\pi} \sqrt{2 |\varepsilon_{b}| \epsilon_{F}}, \label{rho.temperature} \end{equation} where $\gamma \simeq 1.781$ and $\varepsilon_{b}$ is the energy of the two-particle bound state in vacuum which is a more convenient parameter than the four-fermion constant $V$ \cite{Miyake,Gusynin.JETP}. It is obvious from (\ref{BKT.clean}) and (\ref{rho.temperature}) that the pseudogap region shrinks rapidly for high carrier densities \footnote{ In 2D for $s$-wave pairing the high density limit is in fact equivalent the weak coupling BCS limit.} and one may ask (see, for example, \cite{Randeria}) whether this scenario can explain the pseudogap anomalies which are observed over a wide range of temperatures and carrier densities, since in the clean limit the relative size of the pseudogap region $(T_{\rho} - T_{\rm BKT})/T_{\rho}$ is, for instance, less than $1/2$ when the dimensionless ratio $\epsilon_{F}/|\varepsilon_{b}| \lesssim 128 \gamma^{2}/\pi^2 \simeq 41$. A crude estimate for the dimensionless ratio for optimally doped cuprates gives $\epsilon_{F}/|\varepsilon_{b}| \sim 3 \cdot 10^2$ - $10^3$ \cite{Carter} which indicates that in the clean superconductor the pseudogap region produced by the phase fluctuations is too small. Of course, all these estimations are qualitative due to the simplicity of the model. The value of the bare superfluid density, $n_{s}(T)$ is straightforwardly expressed via the bare phase stiffness, $n_s(T) = 4m J(T)$. In particular, it follows from (\ref{J.clean}) that $n_s(T=0) = n_f$. This is not surprising since $n_s(T=0)$ must be equal to the full density $n_f$ for any superfluid ground state in a translationary invariant system \cite{Leggett} and the clean system is translationary invariant. We note, however, as stated above that in HTSC $n_{s}(T=0) \ll n_f$ \cite{Emery.conference}. Substituting (\ref{rho.dependence}) into (\ref{J.clean}) one obtains for $T$ close to $T_{\rho}$ the bare superfluid density as $n_{s}(T \to T_{\rho}^{-}) = 2 n_{f} (1 - T/T_{\rho})$. This behaviour of the bare superfluid density is formally the same as the behaviour of the full superfluid density in the BCS theory. Nevertheless it is important to remember that the full superfluid density in the present model undergoes the Nelson-Kosterlitz jump at $T_{\rm BKT}$ and is zero for $T > T_{\rm BKT}$. We note that one can probe experimentally both the bare superfluid density in high-frequency measurements \cite{Corson} and the full superfluid density in low-frequency measurements \cite{Panagopoulos}. \subsection{Dirty limit} In the dirty limit the quasiparticle transport time $\tau_{tr}$ is small ($\tau_{tr} \ll \rho^{-1}(T=0)$) so that one can neglect the radical in the bracket of (\ref{J.final}) \cite{Abrikosov.book}. The remaining series is easily summed and one obtains for the bare superfluid stiffness \begin{equation} J(\epsilon_{F}, T, \rho(\epsilon_{F},T), \tau_{tr}) = \frac{\epsilon_{F} \tau_{tr} \rho}{4} \tanh \frac{\rho}{2T} . \label{J.dirty} \end{equation} As explained above, due to the Anderson theorem, the expressions (\ref{rho.dependence}) for $\rho$ and (\ref{rho.temperature}) for $T_{\rho}$ remain unchanged in the presence of impurities. Again substituting (\ref{rho.dependence}) into (\ref{J.dirty}) one obtains \begin{equation} T_{\rm BKT} = T_{\rho} \left( 1 - \frac{14 \zeta(3)}{\pi^3} \frac{1}{\epsilon_{F} \tau_{tr}} \right), \qquad T_{\rm BKT} \lesssim T_{\rho}. \label{BKT.dirty} \end{equation} One can see that the size of the pseudogap region is now controlled by the new phenomenological parameter $\tau_{tr}$ which is an unknown function of $\epsilon_{F}$ for HTSC. The experimental data \cite{Timusk} suggest that $\tau_{tr}$ is almost independent on doping level in the underdoped region. It is difficult to obtain more than a qualitative estimate using Eq.(\ref{BKT.dirty}) since in its derivation we have assumed that $\epsilon_{F} \tau_{tr} \gg 1$. In HTSC however, as discussed above (see also \cite{Pogorelov}), this assumption is not always justified. Bearing in mind that the dirty limit implies that the condition $\tau_{tr}^{-1} \gg \rho(T=0) \sim T_{\rho}$ is satisfied, one can easily see that the value of $T_{\rm BKT}$ for this case is less than that given by (\ref{BKT.clean}) for the clean superconductor. Since impurities are inevitably present in HTSC, phase fluctuations can in fact give rise to a pseudogap region that is of comparable size to that experimentally observed. We note that our arguments are in fact quite similar to that given in \cite{Halperin} for the best observing conditions for the BKT physics in superconducting films. However, in contrast to this paper, the gap opening below $T_{\rho}$ is particularly emphasized here. While Eq.(\ref{BKT.dirty}) was derived under assumption $T_{\rm BKT} \lesssim T_{\rho}$ in general case when $T_{\rm BKT}$ can be substantially less than $T_{\rho}$ the self-consistent equation (\ref{BKT.temperature}) with $J(\epsilon_{F}, T_{\rm BKT}, \rho(\epsilon_{F},T_{\rm BKT}))$ given by (\ref{J.dirty}) must be solved. Recall, however, that to make any quantitative estimates, the more realistic $d$-wave model has to be considered and the inequality $\epsilon_{F} \tau_{tr} \gg 1$ should not be assumed \cite{Pogorelov}. The value of the zero temperature superfluid density is now given by $n_{s}(T=0) = \pi n_{f} \tau_{tr} \rho \ll n_{f}$ since $\tau_{tr} \rho \ll 1$. This does not contradict the results of \cite{Leggett} because the system is not translationary invariant in the presence of impurities \cite{Randeria.1998}. Furthermore as one can see, the low value of the superfluid density in HTSC \cite{Emery.conference} may be related to the impurities which are inevitably present in HTSC. Another reason which leads to lowering of the superfluid density is the presence of lattice which also destroys a continuous translational invariance. We note that as was pointed \cite{Chakraverty} quantum fluctuations also lead to a decrease in the superfluid density. \section{The temperature dependence of the bare superfluid density} \label{sec:superfluid} In this section we try to correlate the temperature dependence of the observed in-plane resistivity $\rho_{ab}(T)$ with the recently measured temperature dependence of the bare superfluid density \cite{Corson}. For $T > T_{\rm BKT}$ the expression for the bare superfluid density in the dirty limit (\ref{J.dirty}) can be rewritten in terms of the in-plane conductivity, $\sigma = e^2 n_f \tau_{tr} /m$, where $e$ is the charge of electron: \begin{equation} J(\sigma(\epsilon_{F},T), \rho(\epsilon_{F},T)) = \frac{\pi}{4} \frac{\sigma \rho}{e^{2}} \tanh \frac{\rho}{2T} . \label{J.conductivity} \end{equation} The in-plane resistivity, $\rho_{ab} \sim \sigma^{-1}$ in cuprates has been extensively studied \cite{Timusk} and its temperature and concentration dependencies must reflect the pseudogap properties observed in other experiments. One can say that $\rho_{ab}(T)$ is linear above $T^\ast \simeq T_\rho$ and roughly linear between $T_{\rm BKT}$ and $T_{\rho}$ but with a lower slope. Thus in the interval $T_{\rm BKT} < T < T_{\rho}$ the resistance can be approximately written as $\rho_{ab}(T) = aT + b$, where $a$ and $b$ are functions of $\epsilon_{F}$ but not of temperature. Now, substituting $\sigma \sim \rho_{ab}^{-1}(T)$ into Eq.(\ref{J.conductivity}), one obtains \begin{equation} n_{s} (T) \sim \frac{\rho}{aT+b} \tanh \frac{\rho}{2T} . \label{sup.phenomenological} \end{equation} Our estimations based on Eq.(\ref{sup.phenomenological}) are shown in Fig.~\ref{fig:1}. One can see that, in contrast to the almost linear BCS dependence of $n_s(T)$, we have convex behaviour and the superfluid density becomes zero at $T_{\rho}$. We stress that the curvature of $n_{s}(T)$ is the result of both the temperature dependence of $\rho(T)$ and $\sigma(T)$ for $T_{\rm BKT} < T < T_{\rho}$. More importantly the slope of the curve $n_s(T)$ at $T_{\rho}$ for the dirty metal is substantially less than for the clean one. The experiment \cite{Corson} shows the same curvature for $n_{s}(T)$ but indicates that the bare superfluid density disappears at a lower temperature, $T_{s} < T^{\ast}$. Since the slope $dn_{s}(T)/dT$ at $T_{\rho}$ is very small, as predicted by Eq.(\ref{sup.phenomenological}) and observed experimentally, the non-zero value of $n_s(T)$ between $T_s$ and $T_\rho$ may however simply be too small to be experimentally observed. A definitive answer to this question demands further experiments and theoretical studies. In particular, $\rho$-fluctuations should be taken into account \cite{Gusynin.JETP,Gusynin.new}. One can also comment on the experimentally observed change in the curvature of the full superfluid density, $\mathcal{ N}_{s}(T)$, with changing carrier density \cite{Panagopoulos} even though $\mathcal{ N}_{s}(T)$ cannot be directly linked to the bare superfluid density $n_{s}(T)$ discussed here. Although the full superfluid density disappears above $T_{\rm BKT}$, the curvature present in the bare superfluid density $n_s(T)$ seems to be retained as a curvature in the full superfluid density, $\mathcal{ N}_{s}(T)$, below $T_{\rm BKT}$ \cite{Corson,Panagopoulos}. For low carrier densities (the underdoped region) the pseudogap region, $T_{\rm BKT} < T < T_{\rho}$, is larger and therefore the curvature in $n_s(T)$ is more pronounced. This behaviour seems to be reflected in the full superfluid density, $\mathcal{ N}_s(T)$ below $T_{\rm BKT}$ \cite{Panagopoulos}. It is important however to study experimentally and theoretically the concentration dependence of the bare superfluid density, $n_{s}(T)$ to make a full comparison with the results from \cite{Panagopoulos} for $\mathcal{ N}_{s}(T)$. The experimental data of \cite{Panagopoulos} also show that $\mathcal{ N}_{s}(T)$ does not display the Nelson-Kosterlitz jump. This is probably related to the influence of the interlayer coupling, see Refs. in \cite{Gusynin.new}. \section{Conclusion} Since in HTSC the pairing scale $T^\ast$ is different from the superconducting transition temperature the role of non-magnetic impurities is not traditional and they in fact define the superconducting properties of a ``bad metal''. In particular the presence of nonmagnetic impurities strongly increases the size of the pseudogap phase originating from the fluctuations of the phase of the order parameter. In addition the behaviour of the superfluid density in the presence of the impurities is closer to that experimentally observed. Our results are only qualitative since we have considered the model with non-retarded $s$-wave attraction and an isotropic fermion spectrum. However, it is likely that for $d$-wave pairing, the properties obtained will persist. There is, of course, a problem why a strong disorder does not destroy $d$-wave superconductivity when non-magnetic impurities are pair-breaking. As was suggested by Sadovskii \cite{Sadovskii} even $d$-wave pairing may persist if coupling is strong enough. Further studies are necessary, for example, it is important to explain the concentration dependence of the superfluid slope, $d n_{s} (T)/dT$ at $T=0$ \cite{Panagopoulos,Mesot}. Our results also indicate that it would be interesting to study the BCS-Bose crossover problem in the presence of impurities, especially in $d$-wave case \cite{Sadovskii}. \noindent We gratefully acknowledge V.P.~Gusynin and Yu.G.~Pogorelov for many useful discussions. One of us (S.G.Sh) is grateful to the members of the Department of Physics of the University of Pretoria for hospitality. R.M.Q and S.G.Sh acknowledge the financial support of the Foundation for Research Development, Pretoria.
\section{Introduction} \label{intro} Through the hard work of many excellent experimenters, the detection and characterization of temperature fluctuations in the cosmic microwave background is now on sure footing. Following the initial detection by the DMR instrument aboard COBE \cite{Smoetal90}, numerous ground and balloon-based detections have been made, and the first reasonably large temperature maps at angular scales of a degree have been constructed \cite{Devetal98}. The upcoming satellites MAP and Planck promise full sky temperature maps of unprecedented resolution and sensitivity, as detailed elsewhere in these lectures. Theoretically, much of this intensive effort has been motivated by the realization that the microwave background temperature fluctuations contain a wealth of fundamental cosmological information \cite{Junetal96,ZalSelSpe97,EisHuTeg99}. Polarization of the microwave background is a different story. Polarization is expected in every cosmological model, for the simple reason that the Thomson scattering which thermalizes the radiation has a polarization-dependent cross section. But the polarization signal is generically expected to be a factor of 10 to 50 smaller than the temperature fluctuations, presenting that much greater of an experimental challenge. Only upper limits on polarization of around a part in $10^5$ now exist, but a new generation of experiments optimized for polarization are currently being constructed, which potentially have both the raw sensitivity and the control over systematic errors necessary to make the first detection. In many ways, the experimental study of polarization today is at about the same stage that temperature was ten years ago. This contribution aims to explain how polarization is physically characterized, how it is generated in the microwave background, the mathematical description of the associated power spectra, and the physical effects which might be probed via polarization measurements. Another elementary reference about microwave background polarization from a somewhat different perspective is \citeasnoun{WhiHu97}. \section{Review of Stokes Parameters} \label{stokes} Polarized light is conventionally described in terms of the Stokes parameters, which are presented in any optics text. Consider a nearly monochromatic plane electromagnetic wave propogating in the $z$-direction; nearly monochromatic here means that its frequency components are closely distributed around its mean frequency $\omega_0$. The components of the wave's electric field vector at a given point in space can be written as \begin{equation} E_x=a_x(t)\cos\left[\omega_0 t - \theta_x(t)\right],\quad E_y=a_y(t)\cos\left[\omega_0 t - \theta_y(t)\right]. \label{efield} \end{equation} The requirement that the wave is nearly monochromatic guarantees that the amplitudes $a_x$ and $a_y$ and the phase angles $\theta_x$ and $\theta_y$ will vary slowly relative to the inverse frequency of the wave. If some correlation exists between the two components in Eq.~(\ref{efield}), then the wave is polarized. The Stokes parameters are defined as the following time averages: \label{stokesdef} \begin{eqnarray} I\,&&\equiv \langle a_x^2\rangle + \langle a_y^2\rangle;\\ Q\,&&\equiv \langle a_x^2\rangle - \langle a_y^2\rangle;\\ U\,&&\equiv \langle 2a_xa_y\cos(\theta_x -\theta_y)\rangle;\\ V\,&&\equiv \langle 2a_xa_y\sin(\theta_x -\theta_y)\rangle. \end{eqnarray} The averages are over times long compared to the inverse frequency of the wave. The parameter $I$ gives the intensity of the radiation which is always positive and is equivalent to the temperature for blackbody radiation. The other three parameters define the polarization state of the wave and can have either sign. Unpolarized radiation, or ``natural light,'' is described by $Q=U=V=0$. The parameters $I$ and $V$ are physical observables independent of the coordinate system, but $Q$ and $U$ depend on the orientation of the $x$ and $y$ axes. If a given wave is described by the parameters $Q$ and $U$ for a certain orientation of the coordinate system, then after a rotation of the $x-y$ plane through an angle $\phi$, it is straightforward to verify that the same wave is now described by the parameters \begin{eqnarray} Q'&&=Q\cos(2\phi) + U\sin(2\phi),\nonumber\\ U'&&=-Q\sin(2\phi) + U\cos(2\phi). \label{uqtransf} \end{eqnarray} From this transformation it is easy to see that the quantity $Q^2+U^2$ is invariant under rotation of the axes, and the angle \begin{equation} \alpha\equiv{1\over 2}\tan^{-1}{U\over Q} \label{alpha} \end{equation} transforms to $\alpha -\phi$ under a rotation by $\phi$ and thus defines a constant orientation, which physically is parallel to the electric field of the wave. The Stokes parameters are a useful description of polarization because they are {\it additive} for incoherent superposition of radiation; note this is not true for the magnitude or orientation of polarization. While polarization has a magnitude and an orientation, it is not a vector quantity because the orientation does not have a direction, describing only the plane in which the electric field of the wave oscillates. Mathematically, the Stokes parameters are identical for an axis rotation through an angle of $\pi$, whereas for a vector, such a rotation would lead to an inverted vector and a full rotation through $2\pi$ is required to return to the same situation. The transformation law in Eq.~(\ref{uqtransf}) is characteristic of the second-rank {\it tensor} \begin{equation} \rho={1\over 2} \pmatrix{\vphantom\int I+Q & U-iV\cr \vphantom\int U+iV& I-Q}, \label{densmatrix} \end{equation} which also corresponds to the quantum mechanical density matrix for an ensemble of photons \cite{Kos96} (the matrix is 2 by 2 because the photon has two helicity states). \section{Polarization from Thomson Scattering} \label{polgeneration} Polarization in the microwave background is generated through the polarization-dependent cross-section for Thomson scattering. Consider Thomson scattering of an incoming unpolarized beam of electromagnetic radiation by an electron; this discussion closely follows those in \citeasnoun{Kos96} and \citeasnoun{Kos98}. The total scattering cross-section, defined as the radiated intensity per unit solid angle divided by the incoming intensity per unit area, is given by \begin{equation} {d\sigma\over d\Omega} = {3\sigma_T\over 8\pi} \left|{\hat\varepsilon}'\cdot{\hat\varepsilon}\right|^2 \label{thomson} \end{equation} where $\sigma_T$ is the total Thomson cross section and the vectors $\hat\varepsilon$ and ${\hat\varepsilon}'$ are unit vectors in the planes perpendicular to the propogation directions which are aligned with the outgoing and incoming polarization, respectively. Consider first a nearly monochromatic, unpolarized incident plane wave of intensity $I'$ and cross-sectional area $\sigma_B$ which is scattered into the $z$-axis direction. Defining the $y$-axes of the incoming and outgoing coordinate systems to be in the scattering plane, the Stokes parameters of the outgoing beam, defined with respect to the $x$-axis, follow from Eq.~(\ref{thomson}) as \begin{eqnarray} I &=& {3\sigma_T\over 8\pi\sigma_B}I'(1+\cos^2\theta),\\ Q &=& {3\sigma_T\over 8\pi\sigma_B}I'\sin^2\theta,\\ U &=& 0, \end{eqnarray} \label{stokesout} where $\theta$ is the angle between the incoming and outgoing beams. By symmetry, Thomson scattering can generate no circular polarization, so $V=0$ always and will not be considered further. (Note that Eqs.~(\ref{stokesout}) give the well-known result that sunlight from the horizon at midday is linearly polarized parallel to the horizon). The net polarization produced by the scattering of an incoming, unpolarized ratiation field of intensity $I'(\theta,\phi)$ is determined by integrating Eqs.~(\ref{stokesout}) over all incoming directions. Note that the coordinate system for each incoming direction must be rotated about the $z$-axis so that the outgoing Stokes parameters are all defined with respect to a common coordinate system, using the transformation of $Q$ and $U$ under rotations. The result is \begin{eqnarray} I({\bf\hat z}) &=& {3\sigma_T\over 16\pi\sigma_B}\int d\Omega (1+\cos^2\theta) I'(\theta,\phi),\\ Q({\bf\hat z}) - iU({\bf\hat z}) &=& {3\sigma_T\over 16\pi\sigma_B}\int d\Omega \sin^2\theta e^{2i\phi} I'(\theta,\phi). \end{eqnarray} \label{stokesint} Expanding the incident radiation field in spherical harmonics, \begin{equation} I'(\theta,\phi) = \sum_{lm}a_{lm}Y_{lm}(\theta,\phi), \end{equation} leads to the following expressions for the outgoing Stokes parameters: \begin{eqnarray} I({\bf\hat z})&=&{3\sigma_T\over 16\pi\sigma_B} \left[{8\over 3}\sqrt{\pi}\, a_{00} + {4\over 3}\sqrt{\pi\over 5} a_{20} \right],\\ Q({\bf\hat z}) - iU({\bf\hat z}) &=& {3\sigma_T\over 4\pi\sigma_B} \sqrt{2\pi\over 15} a_{22}. \end{eqnarray} \label{a22} Thus polarization is generated along the outgoing $z$-axis provided that the $a_{22}$ quadrupole moment of the incoming radiation is non-zero. To determine the outgoing polarization in a direction making an angle $\beta$ with the $z$-axis, the same physical incoming field must be multipole expanded in a coordinate system rotated through the Euler angle $\beta$; the rotated multipole coefficients are \begin{eqnarray} {\tilde a}_{lm} &=& \int d\Omega Y^*_{lm}(R\Omega) I'(\Omega)\nonumber\\ &=&\sum_{m'=-m}^m D^{l\,*}_{m'm}(R)\int d\Omega Y^*_{lm'}(\Omega) I'(\Omega), \label{almtilde} \end{eqnarray} where $R$ is the rotation operator and $D^l_{m'm}$ is the Wigner D-symbol. (For a wonderfully complete reference on representations of the rotation group, see \citeasnoun{varshalovich}). In the rotated coordinate system, the multipole coefficient generating polarization is ${\tilde a}_{22}$ by Eqs.~(\ref{a22}). The unrotated multipole components which contribute to polarization will all have $l=2$ by the orthogonality of the spherical harmonics. If the incoming radiation field is independent of $\phi$, as it will be for individual Fourier components of a density perturbation, then \begin{equation} {\tilde a}_{22} = a_{20} d^{2\, *}_{02}(\beta) ={\sqrt{6}\over 4}a_{20}\sin^2\beta, \label{a20} \end{equation} which has used an explicit expression for the reduced D-symbol $d^l_{m'm}$. The outgoing Stokes parameters are finally \begin{equation} Q({\bf\hat n})-iU({\bf\hat n}) ={3\sigma_T\over 8\pi\sigma_B} \sqrt{\pi\over 5}\,a_{20}\sin^2\beta. \label{a20} \end{equation} In other words, an azimuthally-symmetric radiation field will generate a polarized scattered field if it has a non-zero $a_{20}$ multipole component, and the magnitude of the scattered polarization will be proportional to $\sin^2\beta$. Since the incoming field is real, $a_{20}$ will be real, $U=0$, and the polarization orientation will be in the plane of the $z$-axis and the scattering direction. Similar relationships can be derived for radiation fields which are not azimuthally symmetric, which occur in the cases of vector and tensor metric perturbations. In short, what this section shows is that unpolarized quadrupolar radiation fields get Thomson scattered into polarized radiation fields. This is the key fact which must be appreciated to understand why the microwave background should be polarized and what the magnitude of the polarization is expected to be. \section{Generation of Microwave Background Polarization} \label{generation} At times significantly before decoupling, the universe is hot enough that protons and electrons exist freely in a plasma. During this epoch, the rate for photons to Thomson scatter off of free electrons is large compared to the expansion rate of the universe. Thus, the photons and electrons stay in thermal equilibrium at a common temperature and are said to be tightly coupled. As the universe drops below a temperature of around 0.1 eV at a redshift of around 1300, the electrons and protons begin to ``recombine'' into neutral hydrogen. Within a short time, almost all the free electrons are converted to neutral hydrogen, the rapid Thomson scattering ceases for lack of scatterers, and the radiation is said to decouple. At this point, the radiation will propogate freely until the universe reionizes at some redshift greater than 5. During the tight coupling epoch, the photons must have a distribution which mirrors that of the electrons. An immediate consequence is that the angular dependence of the radiation field at a given point can only possess a monopole (corresponding to the temperature) and a dipole (corresponding to a Doppler shift from a peculiar velocity) component, and that the radiation field is unpolarized. Any higher multipole moment will rapidly damp away as the electrons scatter off the free electrons, and no net polarization can be produced through scattering. A quadrupole is subsequently produced at decoupling as free streaming of the photons begins. A single Fourier mode of the radiation field can be described by the temperature distribution function $\Theta(k,\mu,\eta)$ where $k$ is the wavenumber, $\mu = {\bf\hat k}\cdot{\bf\hat n}$ is the angle between the vector $\bf k$ and the propagation direction $\bf\hat n$, and $\eta$ is conformal time. (For mathematical simplicity only a flat universe is considered here, although the non-flat cases are no more complicated conceptually.) Ignoring gravitational potential contributions, free streaming of the photons is described by the Liouville equation $\dot\Theta + ik\mu\Theta = 0$. If the free streaming begins at time $\eta_*$, then the solution at a later time is simply $\Theta(k,\mu,\eta) = \Theta(k,\mu,\eta_*)\exp(-ik\mu(\eta - \eta_*))$. We can reexpress the $\mu$ dependence as a multipole expansion \begin{equation} \Theta(k,\mu,\eta) = \sum_{l=0}^\infty (-i)^l\Theta_l(k,\eta)P_l(\mu); \end{equation} using the identity \begin{equation} e^{iz\cos\phi}=\sum_{n=0}^\infty (2n+1)i^n j_n(z)P_n(\cos\phi) \end{equation} the free streaming becomes \begin{equation} \Theta_l(k,\eta) = (2l+1)[ \Theta_0(k,\eta_*)j_l(k\eta - k\eta_*) + \Theta_1(k,\eta_*)j'_l(k\eta - k\eta_*)], \label{freestream} \end{equation} where $j_l$ is the usual spherical Bessel function. We are interested in the behavior of the free streaming at times near decoupling; at later times, the number density of free electrons which can Thomson scatter has dropped to negligible levels and no further polarization can be produced. The physical length scales of interest for microwave background fluctuations will be larger than the thickness of the last scattering surface, so $k(\eta-\eta_*)$ will be small compared to unity. For small arguments $x\ll 1$, $j_l(x)/j_l'(x)\sim x/l$, which implies that if the monopole and dipole radiation components are initially of comparable size, free streaming through the region of polarization generation with thickness $\Delta$ will generate a quadrupole component from the dipole which is a factor of $2/(k\Delta)$ larger than the quadrupole component from the monopole. In other words, on length scales large compared to the thickness of the surface of last scattering, the quadrupole moment and thus the polarization couples much more strongly to the velocity of the baryon-photon fluid than to the density. Note that on smaller scales with $k\Delta\gtrsim 1$, the polarization can couple more strongly to either the velocity or the density, depending on the scale, but for standard recombination these scales are always small enough that the microwave background fluctuations are strongly diffusion damped. \section{Polarization and Sound Waves} \label{acoustic} Inflation produces acoustic oscillations in the early universe which are {\it coherent}: all Fourier modes of a given wavelength have the same phase. Such acoustic oscillations have a very specific relationship between velocity and density perturbations, which shows up in the relative angular scales of features in the temperature and polarization power spectra. As emphasized in \citeasnoun{HuSug96}, the photon-baryon density perturbation in the tight-coupling regime obeys the differential equation for a forced, damped harmonic oscillator with the damping coming from the expansion of the universe and the forcing from gravitational potential perturbations. The solution is of the form \begin{equation} \Theta_0(k,\eta) = A_1(\eta)\cos(kr_s) + A_2(\eta)\sin(kr_s) \label{theta0} \end{equation} where the amplitudes vary slowly in time and $r_s\simeq \eta/\sqrt{3}$ is the sound horizon. The velocity perturbation follows from the photon continuity equation $\dot\Theta_0 = -k\Theta_1/3$, again neglecting gravitational potential perturbations. A detailed consideration of boundary conditions reveals that initial isentropic density perturbations couple to the cosine harmonic in the small-scale limit, and this approximation is good even for the largest-wavelength acoustic oscillations \cite{HuWhi96}. Thus in an inflationary model, at the surface of last scattering, the photon monopole has a $k$-dependence of approximately $\cos(k\eta_*/\sqrt{3})$, while the dipole, which is the main contributor to the polarization, has a $k$-dependence of approximately $\sin(k\eta_*/\sqrt{3})$. For initial isocurvature perturbations, the density perturbations couple instead to the sine harmonic, but the photon monopole and dipole are still $\pi/2$ out of phase. Squaring these amplitudes gives the rough behavior of the CMB power spectra. Acoustic peaks in the temperature power spectrum occur at scales where $\cos^2(k\eta_*/\sqrt{3})$ has its maxima. The amplitude of the velocity perturbations are suppressed by a factor of $c_s$ with respect to the density perturbations, so the temperature peaks reflect only the density perturbations. The polarization couples to the temperature dipole on scales larger than the thickness of the last scattering surface, and acoustic peaks in the polarization power spectrum will be present at scales where $\sin^2(k\eta_*/\sqrt{3})$ has its maxima. In other words, the temperature peaks represent density extrema, the polarization peaks represent velocity extrema, and for coherent oscillations these two sets of maxima are at {\it interleaved angular scales} (see Fig.~\ref{fig:interleave}). This is a generic signature of coherent acoustic oscillations and is likely the most easily measurable physics signal in microwave background polarization. If two peaks are detected in the temperature power spectrum, the angular scale between the two makes a tempting target for polarization measurements. \begin{figure}[htbp] \centerline{\psfig{file=acoustfig.eps,width=5in}} \caption{The power spectra for temperature fluctuations (top), polarization (center) and temperature-polarization cross-correlation (bottom) for a typical inflationary model. The oscillations remain in phase up to $l=3000$. } \label{fig:interleave} \end{figure} The cross-correlation between the temperature and polarization will have extrema as $-\cos(k\eta_*/\sqrt{3})\sin(k\eta_*/\sqrt{3})$ which fall between the temperature and polarization peaks. (The correlation between the polarization and the velocity contribution to the temperature averages to zero because of their different angular dependences.) The sign of the cross-correlation peaks can be used to deduce whether a temperature peak represents a compression or a rarefaction, which can be checked against the alternating peak-height signature if the universe has a large enough baryon fraction \cite{HuSug96}. A combination of isentropic and isocurvature fluctuations shifts all acoustic phases by the same amount if the ratio of their amplitudes is independent of scale, thus leaving the acoustic signature intact. If the amplitude ratio depends on scale, the coherent acoustic oscillations could be modified, but fine tuning would be required to wash them out completely. Multi-field inflation models generically produce both isocurvature and isentropic perturbations \cite{KofLin87,MukSte98} but the resulting microwave background power spectra are just beginning to be studied in detail \cite{Kanetal98}. \section{The Tensor Harmonic Expansion} \label{tensor} The last two sections have pulled a fast one. We began by discussing polarization as a two component tensor quantity, but then started discussing the production of polarization as if only its amplitude were relevant. A more complete formalism for describing the polarization field has been worked out and will be presented in this section (see \citeasnoun{KamKosSte97} for a more extensive discussion). An equivalent formalism employing spin-weighted spherical harmonics has been used extensively by \citeasnoun{ZalSel97}. Note that the normalizations employed by Seljak and Zaldarriaga are slightly different than those adopted here and by \citeasnoun{KamKosSte97}. The microwave background temperature pattern on the sky $T({\bf \hat n})$ is conventionally expanded in a complete set of orthonormal basis functions, the spherical harmonics: \begin{equation} {T({\bf \hat n}) \over T_0}=1+\sum_{l=1}^\infty\sum_{m=-l}^l a^{\rm T}_{(lm)}\,Y_{(lm)}({\bf \hat n}) \label{Texpansion} \end{equation} where \begin{equation} a^{\rm T}_{(lm)}={1\over T_0}\int d{\bf \hat n}\,T({\bf \hat n}) Y_{(lm)}^*({\bf \hat n}) \label{temperaturemoments} \end{equation} are the temperature multipole coefficients and $T_0$ is the mean CMB temperature. Similarly, we can expand the polarization tensor for linear polarization, \begin{equation} P_{ab}({\bf \hat n})={1\over 2} \left( \begin{array}{cc} Q({\bf \hat n}) & -U({\bf \hat n}) \sin\theta \\ \noalign{\vskip6pt} - U({\bf \hat n})\sin\theta & -Q({\bf \hat n})\sin^2\theta \\ \end{array} \right) \label{whatPis} \end{equation} (compare with Eq.~\ref{densmatrix}; the extra factors are convenient because the usual spherical coordinate basis is orthogonal but not orthonormal) in terms of {\it tensor spherical harmonics}, a complete set of orthonormal basis functions for symmetric trace-free $2\times2$ tensors on the sky, \begin{equation} {P_{ab}({\bf \hat n})\over T_0} = \sum_{l=2}^\infty\sum_{m=-l}^l \left[ a_{(lm)}^{\rm G} Y_{(lm)ab}^{\rm G}({\bf \hat n}) + a_{(lm)}^{\rm C} Y_{(lm)ab}^{\rm C} ({\bf \hat n}) \right], \label{Pexpansion} \end{equation} where the expansion coefficients are given by \begin{eqnarray} a_{(lm)}^{\rm G}&=&{1\over T_0}\int \, d{\bf \hat n} P_{ab}({\bf \hat n}) Y_{(lm)}^{{\rm G} \,ab\, *}({\bf \hat n}), \\ a_{(lm)}^{\rm C}&=&{1\over T_0}\int d{\bf \hat n}\, P_{ab}({\bf \hat n}) Y_{(lm)}^{{\rm C} \, ab\, *}({\bf \hat n}), \label{defmoments} \end{eqnarray} which follow from the orthonormality properties \begin{equation} \int d{\bf \hat n}\,Y_{(lm)ab}^{{\rm G}\,*}({\bf \hat n})\, Y_{(l'm')}^{{\rm G}\,\,ab}({\bf \hat n}) =\int d{\bf \hat n}\,Y_{(lm)ab}^{{\rm C}\,*}({\bf \hat n})\, Y_{(l'm')}^{{\rm C}\,\,ab}({\bf \hat n}) =\delta_{ll'} \delta_{mm'},\nonumber \end{equation} \begin{equation} \int d{\bf \hat n}\,Y_{(lm)ab}^{{\rm G}\, *}({\bf \hat n})\, Y_{(l'm')}^{{\rm C}\,\,ab}({\bf \hat n}) =0. \label{norms} \end{equation} These tensor spherical harmonics have been used primarily in the literature of gravitational radiation, where the metric perturbation can be expanded in these tensors. Explicit forms can be derived via various algebraic and group theoretic methods; see \citeasnoun{Thorne} for a complete discussion. A particularly elegant and useful derivation of the tensor spherical harmonics (along with the vector spherical harmonics as well) is provided by differential geometry \cite{Ste96}. Given a scalar function on a manifold, the only related vector quantity at a given point of the manifold is the covariant derivative of the scalar function. The tensor basis functions can be derived by taking the scalar basis functions $Y_{lm}$ and applying to them two covariant derivative operators on the manifold of the two-sphere (the sky): \begin{equation} Y_{(lm)ab}^{\rm G} = N_l \left( Y_{(lm):ab} - {1\over2} g_{ab} Y_{(lm):c}{}^c \right), \label{Yplusdefn} \end{equation} and \begin{equation} Y_{(lm)ab}^{\rm C} = { N_l \over 2} \left(\vphantom{1\over 2} Y_{(lm):ac} \epsilon^c{}_b +Y_{(lm):bc} \epsilon^c{}_a \right), \label{Ytimesdefn} \end{equation} where $\epsilon_{ab}$ is the completely antisymmetric tensor, the ``:'' denotes covariant differentiation on the 2-sphere, and \begin{equation} N_l \equiv \sqrt{ {2 (l-2)! \over (l+2)!}} \label{Nleqn} \end{equation} is a normalization factor. Note that the somewhat more familiar vector spherical harmonics used to describe electromagnetic multipole radiation can likewise be derived as a single covariant derivative of the scalar spherical harmonics. While the formalism of differential geometry may look imposing at first glance, the expansion of the polarization field has been cast into exactly the same form as for the familiar temperature case, with only the extra complication of evaluating covariant derivatives. Explicit forms for the tensor harmonics are given in \citeasnoun{KamKosSte97}. Note that the underlying manifold, the two-sphere, is the simplest non-trivial manifold, with a constant Ricci curvature $R=2$, so the differential geometry is easy. One particularly useful property for doing calculations is that the covariant derivatives are subject to integration by parts: \begin{equation} \int d{\bf\hat n}AB_{:a} = -\int d{\bf\hat n}A_{:a}B \end{equation} with no surface term if the integral is over the entire sky. Also, the scalar spherical harmonics are eigenvalues of the Laplacian operator: \begin{equation} Y_{(lm):a}{}^{:a}\equiv \nabla^2 Y_{(lm)} = -l(l+1)Y_{(lm)}. \end{equation} The existence of two sets of basis functions, labeled here by ``G'' and ``C'', is due to the fact that the symmetric traceless $2\times2$ tensor describing linear polarization is specified by two independent parameters. In two dimensions, any symmetric traceless tensor can be uniquely decomposed into a part of the form $A_{:ab}-(1/2)g_{ab} A_{:c}{}^c$ and another part of the form $B_{:ac}\epsilon^c{}_b+B_{:bc}\epsilon^c{}_a$ where $A$ and $B$ are two scalar functions. This decomposition is quite similar to the decomposition of a vector field into a part which is the gradient of a scalar field and a part which is the curl of a vector field; hence we use the notation G for ``gradient'' and C for ``curl''. In fact, this correspondence is more than just cosmetic: if a linear polarization field is visualized in the usual way with headless ``vectors'' representing the amplitude and orientation of the polarization, then the G harmonics describe the portion of the polarization field which has no handedness associated with it, while the C harmonics describe the other portion of the field which does have a handedness (just as with the gradient and curl of a vector field). This geometric interpretation leads to an important physical conclusion. Consider a universe containing only scalar perturbations, and imagine a single Fourier mode of the perturbations. The mode has only one direction associated with it, defined by the Fourier vector $\bf k$; since the perturbation is scalar, it must be rotationally symmetric around this axis. (If it were not, the gradient of the perturbation would define an independent physical direction, which would violate the assumption of a scalar perturbation.) Such a mode can have no physical handedness associated with it, and as a result, the polarization pattern it induces in the microwave background couples only to the G harmonics. Another way of stating this conclusion is that primordial density perturbations produce {\it no} C-type polarization as long as the perturbations evolve linearly. This property is very useful for constraining or measuring other physical effects, several of which are considered below. Finally, just as temperature fluctuations are commonly characterized by their power spectrum $C_l$, polarization fluctuations possess analogous power spectra. We now have three sets of multipole moments, $a_{(lm)}^{\rm T}$, $a_{(lm)}^{\rm G}$, and $a_{(lm)}^{\rm C}$, which fully describe the temperature/polarization map of the sky. Statistical isotropy implies that \begin{eqnarray} \left\langle a_{(lm)}^{{\rm T}\,*}a_{(l'm')}^{\rm T}\right\rangle = C^{\rm T}_l \delta_{ll'}\delta_{mm'},&\qquad\qquad& \left\langle a_{(lm)}^{{\rm G}\,*}a_{(l'm')}^{\rm G}\right\rangle = C_l^{\rm G} \delta_{ll'}\delta_{mm'}, \cr \left\langle a_{(lm)}^{{\rm C}\,*}a_{(l'm')}^{\rm C}\right\rangle = C_l^{\rm C} \delta_{ll'}\delta_{mm'},&\qquad\qquad& \left\langle a_{(lm)}^{{\rm T}\,*}a_{(l'm')}^{\rm G}\right\rangle = C_l^{\rm TG}\delta_{ll'}\delta_{mm'}, \cr \left\langle a_{(lm)}^{{\rm T}\,*}a_{(l'm')}^{\rm C}\right\rangle = C_l^{\rm TC}\delta_{ll'}\delta_{mm'},&\qquad\qquad& \left\langle a_{(lm)}^{{\rm G}\,*}a_{(l'm')}^{\rm C}\right\rangle = C_l^{\rm GC}\delta_{ll'}\delta_{mm'}, \label{cldefs} \end{eqnarray} where the angle brackets are an average over all realizations of the probability distribution for the cosmological initial conditions. Simple statistical estimators of the various $C_l$'s can be constructed from maps of the microwave background temperature and polarization. For Gaussian theories, the statistical properties of a temperature/polarization map are specified fully by these six sets of multipole moments. In addition, the scalar spherical harmonics $Y_{(lm)}$ and the G tensor harmonics $Y_{(lm)ab}^{\rm G}$ have parity $(-1)^l$, but the C harmonics $Y_{(lm)ab}^{\rm C}$ have parity $(-1)^{l+1}$. If the large-scale perturbations in the early universe were invariant under parity inversion, then $C_l^{\rm TC}=C_l^{\rm GC}=0$. The arguments in the previous paragraph about handedness further imply that for {\it scalar} perturbations, $C_l^{\rm C}=0$. A question of substantial theoretical and experimental interest is what kinds of physics produce measurable nonzero $C_l^{\rm C}$, $C_l^{\rm TC}$, and $C_l^{\rm GC}$. This question is addressed in the following section. The power spectra can be computed for a given cosmological model through well-known numerical techniques. A set of power spectra for scalar and tensor perturbations in a typical inflation-like cosmological model, generated with the CMBFAST code \cite{SelZal96} are displayed in Fig.~\ref{fig:clsplot}. \begin{figure}[htbp] \centerline{\psfig{file=clt0.ps,width=7in}} \bigskip \caption{ Theoretical predictions for the four nonzero CMB temperature-polarization spectra as a function of multipole moment $l$. The solid curves are the predictions for a COBE-normalized scalar perturbations, while the dotted curves are COBE-normalized tensor perturbations. Note that the panel for $C_l^{\rm C}$ contains no dotted curve since scalar perturbations produce no ``C'' polarization component; instead, the dashed line in the lower right panel shows a reionized model with optical depth $\tau=0.1$ to the surface of last scatter. } \label{fig:clsplot} \end{figure} \section{Polarization and Physical Effects} \label{physics} What is microwave background polarization good for? One basic and model-independent answer to this question was outlined above: polarization can provide a clean demonstration of the existence of acoustic oscillations in the early universe. The fact that three of the six polarization-temperature power spectra are zero for linear scalar perturbations gives several other interesting and model-independent probes of physics. The most important is that the ``curl'' polarization power spectrum directly reflects the existence of any vector (vorticity) or tensor (gravitational wave) metric perturbations. Inflation models generically predict a nearly-scale invariant spectrum of tensor perturbations, while defects or other active sources produce significant amounts of both vector and tensor perturbations. If the measured temperature power spectrum of the microwave background turns out to look different than what is expected in the broad class of inflation-like cosmological models, polarization will tell what part of the temperature anisotropies arise from vector and tensor perturbations. More intriguingly, in inflation models, the amplitude of the tensor perturbations is directly proportional to the energy scale at which inflation occurred, so characterizing the gravitational wave background becomes a probe of GUT-scale physics at $10^{16}$ GeV! Inflation also predicts potentially measurable relationships between the amplitudes and power law indices of the primordial density and gravitational wave perturbations (see \cite{Lidetal97} a comprehensive overview), and measuring a $C_l^{\rm C}$ power spectrum appears to be the only way to obtain precise enough measurements of the tensor perturbations to test these predictions. A microwave background map with forseeable sensitivity could measure gravitational wave perturbations with amplitudes smaller than $10^{-3}$ times the amplitude of density perturbations \cite{KamKos98}, thanks to the fact that the density perturbations don't contribute to $C_l^{\rm C}$. The tensor perturbations generally contribute significantly to the temperature perturbations at angular scales larger than two degrees ($l\lesssim 100$) in a flat universe but have a much broader range of scales in polarization ($50 \lesssim l \lesssim 500$). For tensor and vector perturbations, the amplitude of the C-polarization is generally about the same as that of the G-polarization; if the perturbations inducing the COBE temperature anisotropies are 10\% tensors, then we expect the peak of $l^2C_l^{\rm C}\simeq 10^{-15}$ at angular scales around $l=80$. An experimental challenge not for the faint of heart! A second source of C-type polarization is gravitational lensing. The mass distribution in the universe between us and the surface of last scatter will bend the geodesics of the microwave background photons. This lensing can be described by an effective displacement field, in which the temperature and polarization at each point of the sky in an unlensed universe is mapped to a nearby but different point on the sky when lensing is accounted for. The displacement alters the shape of temperature contours in the microwave background, and likewise distorts the polarization pattern, inducing some curl component to the polarization field. Detailed calculations of this effect and the induced $C_l^{\rm C}$ have been made by \citeasnoun{ZalSel98}. The amplitude of this effect is expected to be around $l^2C_l^{\rm C}\simeq 10^{-14}$ on a broad range of subdegree angular scales ($200\lesssim l \lesssim 3000$) with the power spectrum peaking around $l=1000$ in a flat universe. This lensing polarization signal is just at the limit of detectability for the upcoming Planck satellite; future polarization satellites with better sensitivity could make detailed lensing maps based on the curl component of microwave background polarization. It is interesting to note that tensor perturbations and gravitational lensing are substantially distinguishable by their different angular scales. Note that the most recent version of the publicly available CMBFAST code by Seljak and Zaldarriaga \cite{SelZal96} computes polarization from both tensor modes and from gravitational lensing. A third source of C-type polarization is a primordial magnetic field. If a magnetic field was present at recombination, the linear polarization of electromagnetic radiation would undergo a Faraday rotation as it propagated through the surface of last scatter while significant numbers of free electrons were still present. (Such rotation could also occur after reionization, but both the electron density and the field strength would be much smaller and the resulting rotation is small compared to the primordial signal). This effect rotates an initial G-type polarization field into a C-type polarization field. A detailed estimate of the magnitude of this effect \cite{KosLoe97} shows that a primordial field with present strength $10^{-9}$ gauss induces a measurable one-degree rotation in the polarization at a frequency of 30 GHz. Faraday rotation depends quadratically on wavelength of the radiation, so down at 3 GHz, the rotation would be a huge 100 degrees (although the polarized emission from synchrotron radiation would also be correspondingly larger). Such a rotation will induce $l^2 C^{\rm C}$ at a level of between $10^{-15}$ for one degree of rotation and $10^{-11}$ for large rotations. Additionally, it has been pointed out that Faraday rotation will contribute also to the $C_l^{\rm TC}$ cross-correlation at corresponding levels \cite{ScaFer97} as well as to $C_l^{\rm GC}$. Investigation of the angular dependence and detectability of such a signal is ongoing \cite{MacKos99}. The best current constraints on a homogeneous component of a primordial magnetic field come from COBE constraints on anisotropic Bianchi spacetimes \cite{BarFerSil97}, because a universe which contains a homogeneous magnetic field cannot be statistically isotropic. Detection of a significant primordial magnetic field would both provide the seed field needed to generate current galactic and subgalactic-scale magnetic fields via the dynamo mechanism, and also provide a very interesting constraint on fundamental particle physics, particularly if a field on large scales is detected (see, e.g., \citeasnoun{TurWid88} or \citeasnoun{GasGioVen95}). Faraday rotation from magnetic fields is a special case of cosmological birefringence: rotation of polarization by differing amounts depending on direction of observation. Such rotation could arise from interactions between photons and other unknown fields. Constraints on the C-polarization of the microwave background could strongly constrain new pseudoscalar particles (see, e.g., \citeasnoun{CarFie97}). More generally, non-zero cosmological contributions to the $C_l^{\rm TC}$ and $C_l^{\rm GC}$ cross correlations, which must be zero if parity is a valid symmetry of the cosmological perturbations, would indicate some intrinsic parity to either the primordial perturbations \cite{LueWanKam98} or to some interaction of the microwave background photons \cite{Car98}. These types of effects are generally independent of photon frequency, so they can be distinguished from Faraday rotation through microwave background frequency dependence. The above signals are all model-independent probes of new physics using microwave background polarization. An additional less daring but initially more useful and important use of polarization is in determining and constraining the basic background cosmology of the universe. It has been appreciated for several years now that the microwave background offers the cleanest and most powerful constraint on the gross features of the universe \cite{Junetal96}. If the universe is described by an inflation-type model, with nearly scale-invariant initial adiabatic perturbations which evolve via gravitational instability, then the power spectrum of microwave background temperature fluctuations can strongly constrain nearly all cosmological parameters describing the universe: densities of various matter and energy components, amplitudes and power laws of initial density and gravitational wave perturbations, the Hubble parameter, and the redshift of reionization. More recent work \cite{ZalSelSpe97,EisHuTeg99} has shown that the addition of polarization information can help tighten these constraints considerably, mainly because the new information now gives four theoretical power spectra to match instead of just one. Polarization particularly helps constrain the reionization redshift and the baryon density \cite{ZalHar95}. Polarization will also be important for deciphering the universe if measurements of the temperature anisotropies reveal that the universe is {\it not} described by the simple class of inflation-like cosmological models: it is a strong discriminator between vector and tensor perturbations and scalar perturbations \cite{KamKosSte97}. Finally, no discussion of this sort would be completely honest without mentioning the thorny issue of foreground emission. We are gradually concluding that foregrounds have some non-negligible effect on temperature anisotropies, but that the amplitudes of various foregrounds are small enough that they will not substantially hinder our ability to draw cosmological conclusions from microwave background temperature maps (see, e.g., \citeasnoun{Teg98} for a recent estimate). Whether the same will prove true for polarization is unknown at present. Free-free emission is likely to have only negligible polarization, but synchrotron emission will be strongly polarized, and the polarization of dust emission is difficult to estimate reliably \cite{DraLaz98}. Polarized emission from radio point sources is another potential problem. No present measurements have had sufficient sensitivity to detect polarized emission from any of these foreground sources, so it is difficult to predict the foreground impact. My own guess is that the G-polarization component, from which acoustic oscillations can be confirmed and from which parameter estimation can be significantly improved, will face foreground contamination comparable to the temperature anisotropies. If so, and if the polarization foregrounds are divided evenly between C and G polarization components, then control of foregrounds will become crucial for the very interesting physics probed by the cosmological C polarization. But I fully expect that through a combination of techniques, including carefully tailored sky cuts, measurements at many frequencies, improved theoretical understanding, foreground nongaussianity, and foreground template matching, we will separate out the small cosmological polarization signals from whatever polarized foregrounds are out there. The next five years will bring us microwave background temperature maps of vastly improved sensitivity and resolution, and almost certainly the first detection of microwave background polarization. These observations will provide us with very tight constraints on our cosmological model, or else will reveal some new and unexpected aspect of our universe. Either way, the microwave background will be the cornerstone of a mature cosmology. What is left to do after Planck? One good answer to this question, I believe, is very high sensitivity measurements of microwave background polarization. Such observations hold the promise of probing the potential driving inflation, detecting primordial magnetic fields, mapping the matter distribution in the universe, and likely a variety of other interesting physics yet to be explored. This work has been supported by the NASA Theory Program. Portions of this work were done at the Institute for Advanced Study.
\section{Introduction} The scheduled beginning of experiments at the Relativistic Heavy Ion Collider this year will start an exciting new era of nuclear and particle physics. The estimated high energy density in central heavy ion collisions at RHIC will lead to the formation of a large region of deconfined matter of quarks and gluons, the Quark Gluon Plasma (QGP). This gives us an opportunity to study the properties of QGP and its transition to the hadronic matter, which would then shed light on the underlying fundamental theory of strong interactions, the Quantum Chromodynamics. To study the interactions of many strongly interacting particles and to relate the experimental results to the underlying theories, Monte Carlo event generators are needed. In this talk, we report a recently developed transport model that starts from initial conditions that are motivated by the perturbative QCD and incorporates the subsequent partonic and hadronic space-time evolution. In particular, we will describe the HIJING model \cite{hijing1} that is used to generate the initial phase space distribution of partons and the ZPC model \cite{zpc1} that models the parton cascade. We will also give a brief discussion of the modifications made to the default HIJING fragmentation scheme and ART \cite{art1} hadron evolution. This is followed by some results and a summary. \section{Elements of the new transport model} \subsection{HIJING initial conditions} HIJING is a Monte-Carlo event generator for hadron-hadron, hadron-nucleus, and nucleus-nucleus collisions. The main assumption made in simulating nucleus-nucleus collisions is the binary approximation in which a nucleus-nucleus collision is decomposed into binary nucleon-nucleon collisions. For each pair of nucleons, the impact parameter is calculated using nucleon transverse positions generated according to a Wood-Saxon nuclear geometry. The eikonal formalism is then used to determine the probability for a collision to occur. For a given collision, one further determines if it is an elastic or an inelastic collision, a soft or hard inelastic interaction, and the number of jets produced in the hard interaction. To take into account the nuclear effect in hard scatterings, an impact parameter dependent parton distribution function based on the Mueller-Qiu parameterization of the nuclear shadowing is used. Afterwards, the PYTHIA routines are called to describe the hard interactions, while the soft interactions are treated according to the Lund soft momentum transfer model. After the momentum information of each produced parton is known, HIJING uses a simple model, in which gluons lose energy to the wounded nucleons close to their straight line trajectories, to quench the produced jets (minijets). In the present model, we replace the quenching in HIJING by the cascade of produced partons. \subsection{ZPC parton evolution} The parton cascade in our calculation is carried out with the ZPC model. At present, this model includes only gluon-gluon elastic scatterings. To generate the initial phase space distribution, the formation time for each parton is determined according to a Lorentzian distribution with a half width $t_f=E/m_T^2$. Positions of formed partons are calculated by straight line propagation of the parton from its parent nucleon position. During the time of formation, partons are considered to be part of the coherent cloud of their parents and thus do not suffer rescatterings. The gluon-gluon scattering cross section is taken to be the leading divergent cross section regulated by a medium generated screening mass, which is related to the phase space density of produced partons. In the present study, the simple expression, $\mu^2\approx 3\pi\alpha_S/R_A^2\times(N_G/2Y)$, is used to estimate the screening mass. \subsection{Hadronization} Once partons stop interacting, they are converted into hadrons using the HIJING fragmentation scheme after an additional proper time of approximate $1.2$ fm. In the default HIJING fragmentation scheme, a diquark is treated as a single entity, and this leads to an average rapidity shift of about one unit. We modify this fragmentation scheme to allow the formation of diquark-antidiquark pairs. In addition, the $BM\bar{B}$ probability is $80\%$ among the produced diquark-antidiquark pairs, while the rest are $B\bar{B}$'s. \subsection{ART hadron transport model} For hadron evolution, we use the ART model, which is a very successful transport model for heavy ion collisions at AGS energies. To extend the model for heavy ion collisions at RHIC, we have included anti-nucleon annihilation channels, the inelastic interactions of kaons and anti-kaons, and neutral kaon production. In the ART model, multiparticle production is modeled through the formation of resonances. Since the inverse double resonance channels have smaller cross sections than those calculated directly from the detailed balance, we thus adjust the double resonance absorption cross sections to improve the model. \section{Results} \subsection{comparison with SPS data} \begin{figure}[ht] \psfig{file=fig1.ps,height=2.3in,angle=-90} \caption{ Comparisons between our model calculation and the NA49 data for (a) net baryon rapidity distribution and (b) negative hadron rapidity distribution. \label{fig:sps}} \end{figure} \begin{figure}[ht] \psfig{file=fig2.ps,height=2.3in,angle=-90} \caption{ Predictions of (a) baryon and (b) meson rapidity distributions for RHIC Au+Au central (b=0) collisions. \label{fig:rhic}} \end{figure} Fig.\ref{fig:sps}(a) shows a comparison of the net baryon rapidity distribution between our model results and the NA49 data \cite{na49a} for $5\%$ central Pb(158GeV)+Pb collision. The dotted curve gives the result using the default HIJING fragmentation scheme with afterburner. It shows a visible smaller rapidity shift than that obtained using the modified HIJING fragmentation scheme with afterburner, shown by the solid curve. The latter is seen to give a satisfactory description of the data. Fig.\ref{fig:sps}(b) shows the negative hadron distribution. We see that the modification to HIJING fragmentation scheme increases the rapidity width due to the exchange of meson and baryon positions and that final-state interactions included in the default ART model lowers the central rapidity density. When one turns off the double resonance cross sections, the central rapidity density increases to a reasonable value comparing to the NA49 data \cite{na49a}. An enhancement of strange particles is also observed but the magnitude is very sensitive to their formation time. \subsection{predictions for RHIC} We show our predictions for RHIC Au+Au central (b=0) collisions in Fig.\ref{fig:rhic}. The net baryon peak position is similar to the prediction \cite{brahms1} of Fritiof1.7, but the peak height at 80 is smaller than the 100 predicted by Fritiof1.7. On the other hand, the central rapidity height is around 15 which is similar to the Venus4.02 prediction but is much higher than the Fritiof1.7 prediction. We see that many anti-protons still survive after absorption in the hadronic matter, leading to a value of about 10 at central rapidities. The final meson central rapidity height is much large than the one at the initial time when many rho mesons exist. Although the central rapidity height is similar to Venus, our calculation gives a much wider rapidity distribution. Results using the default HIJING show a similar distribution except that the central rapidity density is higher. Including the afterburner but without double resonances induced pion absorption also gives a similar meson distribution. Also shown in the figure is the distribution of kaons produced from both string fragmentation and hadronic production. The latter is seen to enhance kaon production significantly. \section{Summary} The present study shows that because of the production of diquark-antidiquark pairs, there is a relatively large rapidity shift of net baryons comparing to the default HIJING fragmentation scheme. Many anti-protons survive the final-state interactions and are expected to be observed at RHIC. Also, our model gives a wider rapidity plateau at central rapidity than the predictions from default HIJING model. Furthermore, kaon production is appreciably enhanced due to production from hadronic interactions. However, many new elements need to be incorporated into the code, e.g., parton inelastic scatterings, dynamical screening, improved hadronization, etc. Predictions based on such an improved model will be reported in the future. \section*{Acknowledgments} This work is supported by the NSF Grant PHY-9870038, the Welch Foundation Grant A-1358, and the Texas Advanced Research Project FY97 010366-068. \section*{References}
\section{Introduction} \singlespace After hydrogen and helium, oxygen is the most abundant element in the Universe. It is well-known that knowledge of [O/Fe]-ratios in old, metal-poor stars is required if one is to test theories of Galactic chemical evolution. In addition, the oxygen abundance also affects both the energy generation and opacity of stars near the main sequence turnoff region of the color-magnitude diagram, and thus affects the determination of ages of globular clusters and the oldest stars (\cite{rc85}, \cite{v85}). The sensitivity of age to [O/Fe] is not negligible: according to \cite{v92}, the gradient is $\Delta T/\Delta$ [O/Fe] = -4 Gyr/ 1 dex. Until recently, most determinations of [O/Fe] among old, metal-poor stars, have been confined to the analysis of the forbidden [O I] doublet at $\lambda\lambda$ 6300, 6364 \AA. These transitions arise from the ground state and appear in absorption in the sun and also in metal-poor giants. All such studies agree (cf. reviews by \cite{sun93}, \cite{k94}, \cite{b94}, \cite{m97}) that [O/Fe] increases from $\sim0.0$ at solar metallicity to a value near +0.4 at [Fe/H] $\sim$ -1.0, after which [O/Fe] levels off near this last-named value and is essentially constant (with some scatter) to metal-poor values as low as [Fe/H] $\sim$ -3. The behavior of oxygen therefore mimics that of the other so-called alpha elements, e.g., Mg, Si and Ca. The only exception seems to be the behavior of [O/Fe] among globular cluster giants where, presumably as a result of deep mixing of the stellar envelope through the CNO hydrogen-burning shell, oxygen is significantly depleted, sometimes well below the value [O/Fe] $\sim$ 0.0, and transmuted to nitrogen (e.g., \cite{k97}). Apparently oxygen depletion among field halo giants, similar to that experienced by globular cluster giants, does not take place (\cite{h98}). However, since main sequence stars do not mix their surface layers into regions where oxygen depletion could occur, one naturally assumes that oxygen abundances determined from subdwarfs would be more secure than those determined from their giant star analogues. Unfortunately, it is impractical to use the [O I] doublet to determine [O/Fe] among metal-poor subdwarfs. Owing to that fact that these stars are generally hotter, have higher atmospheric density and have much decreased ratio of line to continuous opacity in comparison with giants, the $\lambda\lambda$ 6300, 6364 \AA$\;$ lines virtually disappear from their spectra. The solar EW of $\lambda$ 6300 \AA$\;$, the stronger of the two forbidden lines, is 6 m\AA. In a giant of metallicity [Fe/H] = -2.2, for example, the EW(6300) $\sim$ 30 m\AA$\;$ when [O/Fe] = +0.4 (e.g., \cite{s97}). We calculated model synthetic spectra of the $\lambda$ 6300 \AA$\;$ region in a subdwarf having solar values of $T_{eff}$ and log g and also having [Fe/H] = -2.2 and found that EW(6300) = 1 m\AA, even for an [O/Fe]-ratio as large as +1.0. High resolution, high S/N analysis of the two well-known subdwarfs HD 74000 and HD 25329 (\cite{bs94}) confirms this surmise. Clearly the discrimination of [O/Fe]-values among very metal-poor subdwarfs in the [O/Fe]-range 0.0 to +1.0 is beyond present observational capabilities, if one employs only the forbidden [O I] doublet. The IR permitted O I triplet lines centered near $\lambda$ 7774 \AA$\;$ provide an alternative means of estimating [O/Fe], their EWs being in the 10-20 m\AA$\;$ range even for very metal-poor subdwarfs having $T_{eff}$ and log g similar to that of the sun (\cite{ar89}). Analysis of these lines has in general led to values of [O/Fe] several tenths dex higher than those derived from the [O I] lines in metal-poor giants (\cite{ar89}, \cite{t92}, \cite{c97}). The discrepancy increases with decreasing metallicity and amounts to more than 0.5 dex as [Fe/H] approaches -3. Since the triplet lines arise from a level more than 9 eV above the ground state of the oxygen atom, \cite{k93} suggested that an increase in the adopted $T_{eff}$ scale (Carney 1983) of subdwarfs by 150-200K would resolve the problem. This solution was criticized by \cite{bc96}, who, in analyzing the infrared OH and CO bands of the well-known, intermediately metal-poor ([Fe/H] = -1.22) subdwarf HD 103095, determined the low oxygen abundance of [O/Fe] = +0.29, consistent with the originally adopted \cite{c83} $T_{eff}$ scale. Other approaches to the problem involving radiative transfer (\cite{kn95}) and chromospheric heating (\cite{t95}) effects have also been invoked, but a generally agreed upon solution has not be forthcoming. Analysis of the low-excitation bands of OH, which lie in the (optical) UV just longward of the atmospheric cutoff, provides a third means of obtaining [O/Fe]-ratios (\cite{bhc84}), but the procedure is rendered difficult because of flux limitations and severe line blending even in metal-poor stars. \cite{bsr91} and \cite{n94} derived [O/Fe]-values for a few metal-poor stars, and concluded that these were in agreement with values derived from analysis of the [O I] lines. Recently, however, an analysis of 23 stars, mostly low-metallicity subdwarfs, based on the UV OH features, has led to the conclusion (\cite{i98}, hereafter I98) that oxygen is strongly overabundant, the [O/Fe]-ratio increasing from $\sim$+0.6 to $\sim$+1.0 as [Fe/H] decreased from -1.5 to -3.0. These results are in close agreement with the results obtained earlier from the IR permitted oxygen triplet. A similar conclusion has been reached from a study of 24 (mostly) unevolved low-metallicity stars (\cite{b99}, hereafter B99), again based on the UV OH bands. If these higher [O/Fe]-values are indeed correct, three conclusions emerge, if at the same time, the smaller [O/Fe]- values derived for giants from analysis of the [O I] lines are also correct. First, all metal-poor giants must indeed have undergone deep mixing which has permitted the ashes of O to N processing to reach the surface. This would, of necessity, be true for halo giants, even those not members of globular clusters. Moreover, even those globular cluster giants having [O/Fe] near +0.4 must themselves also have undergone deep mixing. Second, oxygen does not share the same degree of modest ``overabundance" as do other alpha elements such as Mg, Si and Ca, all of which are predicted to arise from the same nucleosynthetic processes in Type II supernovae (e.g. \cite{ar95}, Fig 1). Finally, the ages of globular clusters are presumably 1 to 2 Gyr younger than one would have inferred from the [O I]-based oxygen abundances (\cite{v85}, \cite{ch98}). It is also possible, of course, that the disagreement between the oxygen abundances arises not because of some real physical effect, but rather from some inadequacy of the analysis or modelling procedure. We have already noted that, among low-metallicity subdwarfs, the [O I] lines become too weak to provide useful discriminatory input when [O/Fe] $<$ +1.0. However, two stars in the list of I98, BD +23 3130 and BD +37 1458, are actually subgiants: their spectra, as we shall show, contain measureable lines of the [O I] doublet. These two stars are also the most overabundant in oxygen of all stars in the I98 list, with [O/Fe] values, derived from the UV OH bands, of +1.17 and +0.97, respectively. The star BD +37 1458 is also to be found in the list by B99, for which they derive a similarly large oxygen abundance of [O/Fe] = +0.83 (King 1993 $T_{eff}$ scale). The remainder of this paper concerns the analysis of [O/Fe] for these two subgiants, making use of observations of the $\lambda$ 6300 [O I] line derived from high resolution, high S/N spectra. We conduct a full LTE analysis of a large sample of Fe I and Fe II lines, from which we determine $T_{eff}$ and log g directly from Kurucz model atmospheres, without appealing to temperature scales derived from colors. This procedure leads to oxygen abundances that do not agree with those derived either from the IR triplet permitted lines or the UV OH bands. Instead, we derive [O/Fe]-values that are consistent with those derived from analysis of [O I]-doublet in bright, low-metallicity field giants. \section{Observations and Reductions} The two stars selected for observation, BD +23 3130 and BD +37 1458, were taken from the lists of metal-poor stars analyzed by I98 and B99 because we inferred that their surface gravities corresponded more closely to subgiants than subdwarfs. As such, we suspected that they would have measureable [O I] lines at $\lambda$ 6300 \AA. Both stars were observed on two different nights using the Hamilton Echelle Spectrograph (\cite{v87}), operated at the coude focus of the Shane 3-m Telescope of Lick Observatory. The spectrograph was set at a (two pixel) spectral resolution of R = 50,000. The CCD detector was a thinned, cosmetically excellent, LL2-LA6-0 chip with dimensions of 2048 x 2048 pixels. Observational details are listed in Table 1. The frames were reduced using standard IRAF routines. First the overscan pedestal was removed, followed by the bias. A normalized flat field was created from several quartz exposures, and used to correct the pixel-to- pixel variations. Occasional bad columns were corrected by interpolation. After apertures were defined and traced, the frames were corrected for scattered light and the spectra extracted. Finally, a wavelength solution, created from Th/A lamp exposures, was attached to the extracted spectra. During both nights of observation, extremely high S/N spectra were taken of hot, rapidly rotating stars. These were used to remove the atmospheric $O_2$ lines present in the $\lambda$ 6300 \AA$\;$ spectral region. Following the procedures outlined in detail elsewhere (e.g., \cite{matt96}), the hot star spectra were adjusted in strength so that the $O_2$ lines matched those of the program star; the adjusted hot star spectrum was then divided from the program star. Fortunately the [O I] line did not lie near an $O_2$ feature in either star, but the $\lambda\lambda$ 6297 \AA$\;$ and 6301 \AA$\;$ Fe I lines of BD +23 3130 were blended with atmospheric $O_2$ lines, as was the $\lambda$ 6301 \AA$\;$ Fe I line of BD +37 1458. Thus great care was taken in matching the $O_2$ features of the hot star with those of the program objects. None of these Fe I lines were used to determine the atmospheric parameters of either program star. The corrected spectrum of this region is, however, shown in the syntheses discussed later (see Section 5), in which the flat fielding was accomplished using the hot stars themselves, without using the quartz exposures. Experience shows (e.g., \cite{s91}) that hot-star flatfields are superior to quartz flatfields as a basis for testing synthetic spectra, especially of weak lines, possibly because the hot star beam fills the collimator of the spectrograph in the same manner as does the beam of the program star. Finally, we estimate that the resulting program star spectra have a S/N $\sim$ 250 in the vicinity of the $\lambda$ 6300 feature. \section{Atomic Data and Equivalent Width Measurements} The atomic data for the Fe I lines used here were taken from the work of the Oxford group (\cite{bl86}) and from \cite{o91}. Only lines with gf-values certain to 10\% or better were adopted from the latter source. In cases where a line was listed in both of the sources, we found that the agreement was generally good, but we adopted the Oxford values in the end. The atomic data for the Fe II lines come from \cite{b91} and \cite{kk87} (see also the discussion in \cite{s91}). The atomic data for the [O I] doublet and the O I permitted triplet lines are well known (cf. \cite{l78}). The gf-values for other elements (see Section 6) follow the recommendations given in \cite{s94}, Shetrone (1996) and Kraft et al. (1997), to which the reader if referred for details. The EWs of the lines were measured using basic IRAF routines. In most cases, we fitted a Gaussian profile to each line, but in the few instances in which a Gaussian fit proved to be poor, we measured the line by direct integration under the continuum. The two methods gave consistent results when tested on lines for which the Gaussian fit was satisfactory. Lines that were badly blended or located on (occasional) bad columns of the CCD chip were not measured. In the spectrum of BD +37 1458, we measured 69 lines of Fe I and 9 lines of Fe II, ranging in EW from 4 to 100 m\AA. In the spectrum of the more metal-poor BD +23 3130, we were able to measure 56 Fe I and 7 Fe II lines, over the same range of EW. The S/N proved high enough that measurements down to the 2-3 m\AA$\;$ level can be considered reliable. Table 2 lists the adopted atomic parameters and EWs of all the lines used in this study. \section{Stellar Parameters and Fe Abundances} Analysis of the EWs was done iteratively using the LTE code MOOG (\cite{s73}) together with \cite{k92} model atmospheres. The latter were also used by both I98 and B99. Trial values for the stellar parameters ($T_{eff}$, log g and $v_t$) were those of I98. The method of analysis closely followed that of the ``Lick/Texas" group, and the reader is referred to the last two of a series of papers by that group (\cite{k97}, \cite{s97}) for a more lengthy discussion of the procedures. \footnote{We adopt a solar Fe-abundance of log epsilon(Fe) = 7.52. Otherwise all solar abundances adopted here are those of \cite{ag89}.} The ab initio stellar parameters were adjusted so that (1) the stronger Fe I lines gave the same abundance on the average as the weaker lines, (2) the Fe I lines gave an Fe abundance that was on the average independent of excitation potential of the Fe I lines, and (3) the average abundance of Fe based on Fe II lines was the same as that of the Fe I lines (to within 0.02 dex). Because the ratios [Mg/Fe] and [Si/Fe] among metal poor stars are generally positive (and near +0.4 to +0.5), and the solar abundances of Fe, Mg and Si are comparable, the input model should allow for the increased electron density coming from the ionization of Mg and Si. In prior investigations of the Lick/Texas group, that increase was simulated in the input models by adopting a value of [Fe/H] higher than the final iterated value. Thus, for example in the present case, in which the iterated final value of [Fe/H] for BD +37 1458 proved to be -2.31, we adopted an input model abundance of [Fe/H] = -2.0. Since, as proved finally to be the case, [Fe/H] = -2.3 whereas [Mg/Fe] and [Si/Fe] $\sim$ +0.5, a compromise input metallicity of -2.0 should allow for an ``average" increase in the free electron supply from Mg, Si and Fe by a factor of about 2 (i.e. factors of 3, 3, and 1 for Mg, Si and Fe, respectively). The multiple iteration of the basic parameters following the prescriptions of (1), (2) and (3) above, led to the final adopted values listed in Table 3. Plots of Fe abundance vs Fe I excitation potential (sensitive to the adopted $T_{eff}$) and Fe abundance vs Fe I line- strength (sensitive to microturbulent velocity $v_t$) are shown in Figures 1a for BD +21 3130 and 1b for BD +37 1458 for the parameters adopted here. As a check on the reliability of our procedure in which we simulate the effect on the models of an ``overabundance" of Mg and Si relative to Fe by substituting an input model with a ``compromise" increase in all three elements, we considered several LTE Kurucz models kindly supplied by Bruce Carney in which actual overabundances of Mg and Si by 0.5 dex relative to Fe were directly incorporated. For example, we compared the effect of the ``simulated" ($T_{eff} = 5100$ K, log g = 2.9, [Fe/H] = -2.0, [Mg/Fe] = [Si/Fe] = 0.0) and ``real" ($T_{eff} = 5100$ K, log g = 2.9, [Fe/H] = -2.31, [Mg/Fe] = [Si/Fe] = +0.5) model on the final iterated Fe and O abundances of BD +37 1458, in which the similated values led to [Fe/H] = -2.31. We find that the changes are extremely minor, amounting at most to 0.01 dex in [Fe/H] and 0.05 dex in [O/Fe]. The reason these changes are minor is no doubt a result of the small modifications induced in the free electron supply, since at each reference depth in the continuous opacity $\tau_{5000}$, the change in $T_{eff}$ and $N_e$ amounts to no more than 20 K and 7 percent at the level of the continuum, respectively. This in turn reflects the fact that for the overabundances of Fe, Mg and Si dicussed above, 60 percent or more of the free electron supply is derived from the ionization of H at every optical depth in the atmosphere, and thus is virtually independent of the precise value of the input abundances of these elements. In Table 3, we list our final values of the stellar parameters along with those adopted by I98 and B99. We obtain values of $T_{eff}$, derived from our LTE analysis of roughly 60 Fe I lines, that are considerably lower (by 150 to 350K) than those of the other investigations. For comparison purposes, we plot Fe abundance vs Fe I excitation potential and vs Fe I line strength, using our measured Fe I EWs, for the values of $T_{eff}$ and $v_t$ adopted by I98 and B99 (Figures 2a-d). For the B99 paper, the plot corresponds to the adoption of the \cite{k93} $T_{eff}$ scale. Inspection of these Figures reveals that adoption of the I98 and B99 values of $T_{eff}$ for these two stars leads to unacceptably large slopes and scatter in both the excitation and turbulence plots. Compared with the other two investigations, our values of $T_{eff}$ and log g are lower, but the differences in metallicity, i.e., [Fe/H], are fairly small ($\leq$ 0.2 dex). As noted above, we obtain log g using the requirement that the Fe abundance derived from the Fe II lines shall closely equal that derived from the Fe I lines. In Table 3, we list the values of log $\epsilon$(Fe I) and log $\epsilon$(Fe II) from our model parameters for the two stars, and it will be seen that this condition is met to within 0.01 dex. We derive also (Table 3) the mean abundances of Fe I and Fe II for the I98 and B99 model parameters, making use of our EW measurements. These lead to discrepancies in the mean abundance of Fe from Fe I and Fe II of 0.1 to 0.2 dex. We note in the next section that the value of [O/Fe] is somewhat sensitive to the choice of log g. Thus, as an experiment, we examined the Fe ionization equilibrium for an arbitrary 0.3 dex change in log g, having fixed our value of $T_{eff}$ in accordance with the Fe I excitation plot exhibited in Figures 1a and 1b. The changes in log $\epsilon$(Fe) from Fe I and Fe II lines are listed in Table 3, and lead to discrepancies in Fe abundance somewhat more than 0.1 dex. Generally it has been found (cf. \cite{k97}, \cite{s97}, \cite{gw98}) that, among globular cluster giants, estimates of surface gravity based on the ionization equilibrium of Fe I and Fe II are in excellent agreement with estimates based on values of the effective temperature, mass and luminosity expected for such stars. Thus, having settled on values of $T_{eff}$ for BD +23 3130 and BD +37 1458, we can check whether our corresponding log g's are consistent with expectation based on stellar evolution, using the color-magnitude arrays of globular clusters. BD +37 1458, for example, has a metallicity ([Fe/H] = -2.3) which is close to that of M92 (\cite{zw84}, \cite{s91}, \cite{matt96}). Theoretical models of a cluster with this metallicity require log g = 2.7 if $T_{eff}$ = 5100K and the age is 14 Gyr (\cite{sc91}); the models of \cite{c82} for M92 yield log g = 2.8 for this $T_{eff}$. Our value of log g = 2.9 seems about right. If this star were actually located in M92 it would have $M_{Vo}$ = +1.9 for an assumed distance modulus of $(m-M)_o$ = 14.49 (\cite{d93}) and $(B-V)_o$ = +0.62 (\cite{sh88}). The observed $(B-V)$ = +0.60 (\cite{clla}). Interstellar reddening in this direction appears to be zero or quite small even though b = 9.8 deg: for five stars selected from the Bright Star Catalog (\cite{bsc}) within 5 deg of BD +37 1458 and having $b \leq 10$ deg, $<E(B-V)>$ = +0.02 $\pm$ 0.03 ($\sigma$). Thus the observed and ``predicted" \footnote{From the same source, one finds $<E(B-V)>$ = +0.01 $\pm$ 0.13 for the five stars nearest to BD +37 1458, and having D $>$ 100 pc. Stars fainter than those in the Bright Star Catalog, and close to BD +37 1458, are found in the \cite{hip}. There are 3 stars (a K0 giant, a K0 supergiant, and a main-sequence A2 star) between $l = 171^o$ and $175^o$, having distances greater than 150 pc and $b < 8^o$. For these $<E(B-V)> = 0.01 \pm 0.03$. Unfortunately the spectral types are those assigned by the Draper Catalog; MK classifications are unknown. In the direction of BD +37 1458, the dust reddening maps (\cite{sfd98}) lead to E(B-V) = 0.28. Evidently some dust clouds are located in a region beyond BD +37 1458.} M92 colors are quite close. The distance to BD +37 1458 thus becomes 256 pc, which is only slightly beyond the upper limit of the Hipparcos distance plus its estimated error: 173 (+60, -32) pc. The I98 model is also consistent with ``membership" in M92, the increase in $T_{eff}$ (from 5100K to 5260K) being compensated by the corresponding increase in log g (from 2.8 to 3.0), and the $M_{Vo}$ derived leads to a distance lying within the error bars of the Hipparcos-based value. The values of $T_{eff}$ and log g assigned to this star by B99 are still larger than the I98 values and again are consistent with the color- magnitude array of M92. Identification of BD +37 1458 as a star consistent with membership in M92 therefore does not help to discriminate which of the models discussed here is correct. But all models do place the star above the main sequence in the subgiant domain. There are no galactic globular clusters with metallicities as low as [Fe/H] = -2.8, so a direct comparison of BD +23 3130 with globular cluster stars cannot be made. There appear also to be no available theoretical isochrones corresponding to clusters with metallicities below that of M92. Examination of the Straniero and Chieffi isochrones with respect to [Fe/H] suggests that subgiant and giant branches crowd more closely together as metallicity decreases; thus extrapolation toward a metallicity as low as [Fe/H] = -2.8 is a dangerous procedure. Nevertheless, we estimate that, relative to BD +37 1458, BD +23 3130 should be $\sim$0.05 mag bluer in (B-V) because of its reduced metallicity and $\sim$0.10 mag redder because of its reduced $T_{eff}$, and thus if $(B-V)_o$ = 0.60 for the former star, $(B-V)_o$ = 0.65 for the latter. The observed (B-V) = 0.64 (\cite{clla}). Once again we find $<E(B-V)>$ = 0.01 $\pm$ 0.07 from the five nearest stars in the Bright Star Catalog, so interstellar reddening is quite small or vanishing. However, because in color-magnitude diagrams, cluster giant branches are extremely steep, we suggest that using estimates of color to derive values of $M_V$ is an unwise procedure. The best we can say is that we expect BD +23 3130 to have a lower log g than BD +37 1458, as we see from our spectroscopic determination. In Figures 3a and 3b, we synthesize three Fe I lines of similar intermediate excitation (2.4 to 2.6 eV) and differing strengths in the spectral region $\lambda$ 6134 \AA$\;$ to $\lambda$ 6140 \AA, using our parameters and those of I98 and B99. The Hamilton spectra, binned by two pixels (= effective spectral resolution) are shown as dots. The parameters adopted by I98 and B99 produce synthesized lines that are too weak compared with the observed spectra. Obviously the discrepancies could be reduced simply by increasing the Fe abundance, but such a change would introduce a serious discordance between models and observations for lines of higher and lower excitation, as seems clear from inspection of Figures 2a-d. \section{Oxygen Abundances} Using the same techniques we employed to create Figures 3a and 3b, we show in Figures 4a,b and 5a-d the spectral region surrounding the $\lambda$ 6300 [O I] line in BD +23 3130 and BD +37 1458, respectively. Again our observed Lick spectra (binned by 2) are shown as dots. In each figure, the Fe abundance was held constant while the O abundance was changed to create plots (thin lines) for values of [O/Fe] = 0.0, +0.5, +1.0 and +1.5. The model parameters and Fe abundances are those corresponding to each individual investigation discussed in this paper. Turning first to the spectrum of BD +23 3130, we see that the present synthesis gives a satisfactory representation of the Fe I lines at $\lambda \lambda$ 6297.8 \AA, 6301.5 \AA$\;$ and 6302.5 \AA, and leads to an estimated oxygen abundance of [O/Fe] = +0.35 $\pm$ 0.2 (est. error, based on the S/N). However, the [O I] line is so weak (we estimate EW $\sim$ 1 m\AA) that the quoted [O/Fe]-value may be an upper limit, in which case the true oxygen abundance is even smaller than [O/Fe] = +0.35. The I98 model produces Fe I lines that are much too weak, but corresponds to an estimated oxygen abundance of [O/Fe] = +0.7 $\pm$ 0.2. Neither model comes close, however, to the value [O/Fe] = +1.17 obtained by I98 from synthesis of the OH features. Thus for this star, there is a serious discrepancy between the oxygen abundances derived from [O I] and OH, even if we adopt the parameters $T_{eff}$, log g, $v_t$ and [Fe/H] suggested by I98. The situation with respect to BD +37 1458 is less decisive. The synthesis using the present parameters leads to a satisfactory representation of the three Fe I lines listed in the preceding paragraph, and the corresponding oxygen abundance is [O/Fe] = +0.5 $\pm$ 0.2. The I98 model produces Fe I lines at $\lambda\lambda$ 6297.8 \AA$\;$ and 6301.5 \AA$\;$ that are much too weak, although the modelling of the remaining Fe I line is satisfactory. The corresponding oxygen abundance is [O/Fe] = +0.75 $\pm$ 0.2, a value reasonably close to the value [O/Fe] = +0.97 obtained by I98 from the OH features. This star was also considered by B99, who found [O/Fe] = 0.875 and 0.83, based on modelling the OH features, using the \cite{c83} and \cite{k93} $T_{eff}$ scales, respectively. These values of [O/Fe] are statistically indistiguish- able from the value of [O/Fe] we derive from the [O I] line, using their choices of the basic parameters. However, as we have seen, these models give a poor representation of the Fe I lines (Figures 3a,3b), and do not satisfy the requirements of the Fe I excitation and turbulence plots (Figures 2a-d). We may also explore the changes in oxygen abundance derived from OH that would be induced if I98 and B99 had adopted the values of $T_{eff}$ preferred by our analysis. According to B99, an increase in $T_{eff}$ by 100 K increases the derived O from OH by +0.16 dex. But it also increases Fe by $\sim$ 0.07 dex. So the net increase in [O/Fe] is about 0.09 dex. If I98 had adopted our $T_eff$ for BD +37 1458 their [O/Fe] would have gone down by $\sim$ 0.15 dex. In the case of BD +23 3130, the reduction would be $\sim$ 0.25 dex. \section{Other Elements} In Table 4 and 5 we list, along with the determinations of [O/Fe] cited in Section 5, abundances [el/Fe] for Na, the alpha elements Mg, Si, Ca and Ti, and the Fe-peak element Ni for the two stars studied here. In both Tables, the [el/Fe]-ratios are based on the EWs determined in this study. The Tables differ in that the adopted Fe abundances of Table 4 are taken as the Fe I abundances we derive using these EWs and the model parameters ($T_{eff}$, log g) used by I98 and B99, whereas the Fe abundances of Table 5 are those actually adopted by I98 and B99. We list also in these Tables the mean [el/Fe]-ratios found by \cite{m95} for a large sample of metal-poor stars having [Fe/H] $<$ -2.0. At any given [Fe/H], the scatter in [el/Fe]-ratios in the \cite{m95} sample is typically $\sim$ 0.2 dex for stars more metal-rich than [Fe/H] =-3.2. Thus we see from inspection of Tables 4 and 5 that almost all models give [el/Fe]-ratios for BD +37 1458 and BD +23 3130 that are in reasonable agreement with the mean [el/Fe]-ratios of the \cite{m95} sample. The only significant departures seem to be those of the I98 models tabulated in Table 5. For these the [el/Fe]-ratios appear to be rather large, and the values for Na and Ni are particularly uncomfortable. However, this departure is much reduced if we use the (mean) Fe-abundances we derive from the present set of EWs and the I98 model parameters, rather than the Fe abundances actually adopted by I98. Except for these two models of Table 5, the [el/Fe] ratios for the ``$\alpha$ elements", Mg, Si and Ca are close to their ``normal" expected values of $\sim$ +0.4 to +0.6 [cf., e.g., [$\alpha$/Fe]-ratios in M15 giants (Sneden et al. 1997).]. \section{Conclusions} In an analysis of the OH bands ($\lambda$ 3100 \AA) of 23 presumed subdwarfs, I98 found that the oxygen abundances were generally higher than the traditionally accepted values that had been determined from the analysis of the [O I] doublet in the spectra of low-metallicity giants. Similar conclusions have also been reached by B99, again from analysis of the OH bands in subdwarfs. The differences increase with decreasing metallicity, and reach $\sim$+0.6 dex as [Fe/H] approaches -3.0. In particular, the two highest [O/Fe]-ratios in the I98 sample were those found in BD +37 1458 and BD + 23 3130, for which [O/Fe] = +0.97 and +1.17 for [Fe/H] = -2.40 and -2.90, respectively. In their analysis of the former star, B99 found a similar high value of [O/Fe] = 0.87 or 0.83, depending on the choice of temperature scale. Atmospheric modelling of very metal-poor ([Fe/H $<$ -2) subdwarfs shows that the [O I]-doublet, even for [O/Fe]-ratios as high as +1.0, practically ``disappears": the EW of the stronger member of the pair ($\lambda$ 6300A) approaches 1.0 m\AA$\;$ or less. However, examination of the I98 sample indicates that the two stars discussed above have surface gravities more like subgiants than subdwarfs. These two stars thus presented the possibility that the EWs of the $\lambda$ 6300 \AA$\;$ line of [O I] might be measureable in spectra having sufficiently high S/N, and thus provide a check on the oxygen abundances derived from the OH feature. Thus, on the basis of high resolution (R = 50000), high S/N ($\sim$250) echelle spectra of BD +23 3130 and BD +37 1458, we have determined abundances [Fe/H], and [el/Fe]-ratios for Na, Ni, and the traditional alpha elements Mg, Si, Ca and Ti. We conducted a full-scale abundance analysis based on the EWs of about 60 Fe I and 7-9 Fe II lines, from which we determined from LTE models the appropriate values of $T_{eff}$, log g, [Fe/H] and $v_t$, independent of considerations based on the calibration of color scales. We determined oxygen abundances from synthetic spectra of the region encompassing the [O I] line at $\lambda$ 6300 \AA. Our conclusions are these: (1) Based on the ionization equilibrium of Fe, we find that these two stars are indeed subgiants, with surface gravities somewhat lower than those adopted by I98 and B99. (2) Our values of $T_{eff}$ are also lower than those obtained by I98 and B99, by amounts between $\sim$150 and 300K. Our [Fe/H] values are nevertheless within $\sim$0.2 dex of those determined by I98 and B99. The metallicity of BD +37 1458 is essentially the same as that of M92; all models are in reasonable agreement with the assumption that this star is a surrogate for an M92 star lying on the steep subgiant branch of that cluster. The metallicity of +23 3130 is lower than that of any known galactic globular cluster. (3) Except for two models listed in Table 5, all investigations lead to plausible [el/Fe]-ratios for alpha elements, Na and Ni, when compared with those found for metal-poor stars by \cite{m95}. (4) Our synthetic spectra of the region of the [O I] line at $\lambda$ 6300 \AA$\;$ lead to the traditional oxygen abundances of [O/Fe] = +0.35 $\pm$ 0.2 for BD +23 3130 and +0.50 $\pm$ 0.2 for BD +37 1458. The [O/Fe]-value quoted for BD +23 3130 may be an upper limit. (5) Use of our Fe I EWs leads to a fairly low value of [O/Fe] = +0.7 in the case of BD +23 3130, even if we adopt the parameters of I98. For the other star, an [O/Fe]-ratio closer to the value derived from the OH feature is found if the parameters adopted by I98 and B99 are adopted. These parameters, however, yield a poor representation of the Fe I features in this star, as do the parameters adopted by I98 for BD +23 3130. (6) A discrepancy therefore remains among these two stars between the [O/Fe]-ratio derived from [O I] and the OH feature. This is of some significance especially when one realizes that these two stars have the highest [O/Fe]-ratios among the stars considered by I98. The origin of these differences remains unclear. (7) The procedures adopted here are identical with those previously employed (e.g., \cite{s97}, \cite{k97}) in the analysis of the spectra of globular cluster and halo field giants and subgiants. In these studies, analysis of the Fe I and Fe II line spectra is a basic requirement in the determination of atmospheric parameters. (8) The results obtained here suggest that it is too early to conclude that the oxygen abundances of old, metal-poor stars need to be revised drastically upward. We suggest that input models for such stars need to be examined in the light of a full LTE analysis of the Fe I and Fe II spectrum, before values of $T_{eff}$, log g, $v_t$ and [Fe/H] are assigned. \acknowledgments We are indebted to C. Sneden for use of his MOOG code, and have benefitted from useful conversations with R. Peterson. We are especially grateful to Bruce Carney for supplying us with several $\alpha$-enhanced Kurucz model atmospheres and for reading and commenting on a preliminary version of this manuscript. This research has been supported by NSF Grant AST 96-18351, which we greatly appreciate. \clearpage \begin{deluxetable}{lcc} \footnotesize \tablenum{1} \tablecaption{Log of Observations.} \tablewidth{0pt} \tablehead{ & \colhead{BD +23 3130} & \colhead{BD +37 1458}} \startdata UT Date & 8 Aug 1998 & 31 Oct 1998 \nl Exposure & 3600 s & 3600 s \nl V & 8.95 & 8.92 \nl B-V & 0.64 & 0.60 \nl \enddata \end{deluxetable} \clearpage \begin{deluxetable}{ r r r r r r} \footnotesize \tablenum{2} \tablecaption{Line Measurements.} \tablewidth{0pt} \tablehead{ \colhead{Wavelength (\AA)} & \colhead{E.P. (eV)} & \colhead{log(gf)} & \colhead{EW (m\AA)} & \colhead{EW (m\AA)} & \colhead{Ref} \nl & & & \colhead{BD +37 1458} & \colhead{BD +23 3130} &} \startdata \multicolumn{1}{l}{Fe I} & & & & & \nl 4531.15 & 1.49 & -2.155 & 58.2 & 44.5 & 1d\nl 4592.66 & 1.56 & -2.449 & 45.4 & 29.4 & 1d\nl 4602.01 & 1.61 & -3.154 & 12.1 & 7.6 & 1d\nl 4602.94 & 1.49 & -2.208 & 54.8 & 45.3 & 1d\nl 4643.46 & 3.64 & -1.147 & 7.5 & -- & 2\nl 4647.43 & 2.94 & -1.351 & 23.6 & 13.4 & 2\nl 4733.60 & 1.49 & -2.987 & 24.3 & 14.0 & 1d\nl 4736.77 & 3.20 & -0.752 & 36.7 & -- & 2\nl 4786.81 & 3.00 & -1.606 & 12.4 & 7.6 & 2\nl 4789.65 & 3.53 & -0.957 & 15.9 & -- & 2\nl 4871.32 & 2.85 & -0.362 & 69.6 & 54.8 & 2\nl 4872.14 & 2.87 & -0.567 & 60.6 & 45.3 & 2\nl 4890.75 & 2.86 & -0.394 & 72.6 & 54.8 & 2\nl 4891.49 & 2.84 & -0.111 & 84.5 & -- & 2\nl 4918.99 & 2.85 & -0.342 & 74.5 & 59.5 & 2\nl 4938.81 & 2.86 & -1.077 & 36.4 & 24.6 & 2\nl 4939.69 & 0.86 & -3.340 & 38.2 & 31.2 & 1b\nl 4994.13 & 0.91 & -3.080 & 48.8 & 38.5 & 1b\nl 5006.12 & 2.82 & -0.615 & 62.3 & 45.5 & 2\nl 5028.13 & 3.56 & -1.122 & 9.0 & -- & 2\nl 5041.07 & 0.95 & -3.086 & 49.2 & 37.3 & 2\nl 5044.21 & 2.85 & -2.017 & 9.0 & -- & 2\nl 5048.44 & 3.94 & -1.029 & 5.1 & -- & 2\nl 5049.82 & 2.27 & -1.355 & 56.7 & 44.1 & 2\nl 5051.64 & 0.91 & -2.795 & 63.9 & 54.1 & 1b\nl 5068.77 & 2.93 & -1.041 & -- & 22.2 & 2\nl 5079.23 & 2.20 & -2.067 & 30.6 & 20.1 & 1b\nl 5079.74 & 0.99 & -3.220 & 38.1 & 28.1 & 1b\nl 5083.34 & 0.96 & -2.958 & 52.6 & 42.4 & 1b\nl 5107.45 & 0.99 & -3.087 & 43.8 & 34.0 & 1b\nl 5107.65 & 1.56 & -2.418 & 44.6 & 33.8 & 1d\nl 5123.72 & 1.01 & -3.068 & 44.4 & 35.2 & 1b\nl 5127.36 & 0.91 & -3.307 & 37.3 & 30.6 & 1b\tablebreak 5151.92 & 1.01 & -3.322 & 32.2 & 22.1 & 1b\nl 5191.45 & 3.03 & -0.551 & 54.9 & 37.6 & 2\nl 5192.34 & 2.99 & -0.421 & 60.8 & 45.8 & 2\nl 5198.71 & 2.22 & -2.135 & 25.2 & 16.4 & 1c\nl 5216.28 & 1.61 & -2.150 & 56.9 & 43.8 & 1d\nl 5217.39 & 3.20 & -1.162 & 21.9 & 13.3 & 2\nl 5232.94 & 2.94 & -0.057 & 82.7 & 67.8 & 2\nl 5242.49 & 3.62 & -0.967 & -- & 6.0 & 2\nl 5247.05 & 0.09 & -4.946 & 11.2 & -- & 1a\nl 5288.53 & 3.68 & -1.508 & 3.9 & -- & 2\nl 5307.37 & 1.61 & -2.987 & 18.6 & 11.4 & 1d\nl 5367.47 & 4.42 & 0.443 & -- & 16.7 & 2\nl 5383.37 & 4.31 & 0.645 & -- & 28.1 & 2\nl 5397.13 & 0.91 & -1.993 & 96.8 & -- & 1b\nl 5405.78 & 0.99 & -1.844 & 98.3 & -- & 1b\nl 5434.53 & 1.01 & -2.122 & 88.9 & -- & 1b\nl 5501.46 & 0.95 & -3.046 & 51.2 & 40.1 & 2\nl 5506.78 & 0.99 & -2.797 & 60.9 & 51.0 & 1b\nl 5701.55 & 2.56 & -2.216 & 11.1 & 6.1 & 1f\nl 6065.49 & 2.61 & -1.530 & 35.5 & 22.9 & 1f\nl 6136.62 & 2.45 & -1.400 & 51.2 & 34.2 & 1e\nl 6137.70 & 2.59 & -1.403 & 45.2 & 30.0 & 1f\nl 6173.34 & 2.22 & -2.880 & 8.0 & -- & 1e\nl 6219.29 & 2.20 & -2.433 & 16.5 & 9.9 & 1c\nl 6230.73 & 2.56 & -1.281 & 54.5 & 37.4 & 1f\nl 6246.32 & 3.59 & -0.877 & 19.8 & 9.9 & 2\nl 6252.56 & 2.40 & -1.687 & 38.9 & 28.7 & 1c\nl 6265.14 & 2.18 & -2.550 & 15.9 & 9.9 & 1e\nl 6297.80 & 2.22 & -2.740 & 11.0 & -- & 1c\tablenotemark{a}\nl 6322.69 & 2.59 & -2.426 & 7.9 & -- & 1f\nl 6358.69 & 0.86 & -4.468 & 7.6 & -- & 1b\nl 6411.65 & 3.64 & -0.717 & -- & 13.1 & 2\nl 6421.36 & 2.28 & -2.027 & 33.6 & 18.6 & 1e\nl 6430.85 & 2.18 & -2.006 & 39.4 & 25.2 & 1c\nl 6494.99 & 2.40 & -1.273 & 61.2 & 48.9 & 1e\tablebreak 6593.88 & 2.43 & -2.422 & 11.9 & 6.1 & 1c\nl 6677.99 & 2.68 & -1.418 & 42.3 & 27.5 & 2\nl 6945.21 & 2.42 & -2.482 & -- & 6.7 & 1c\nl 6750.15 & 2.42 & -2.621 & 8.0 & -- & 1e\nl 6945.21 & 2.42 & -2.482 & 11.4 & -- & 1c\nl 6978.86 & 2.48 & -2.500 & 10.9 & -- & 1c\nl 7511.02 & 4.16 & 0.099 & 32.9 & 19.7 & 2\nl \multicolumn{1}{l}{Fe II} & & & & & \nl 4555.88 & 2.83 & -2.290 & 33.2 & 28.7 & 3\nl 4576.34 & 2.84 & -2.822 & 13.4 & 10.6 & 4\nl 4582.83 & 2.84 & -3.094 & 10.9 & 6.2 & 4\nl 4620.52 & 2.83 & -3.079 & -- & 5.7 & 4\nl 4923.92 & 2.89 & -1.240 & 77.0 & -- & 3\nl 5197.58 & 3.23 & -2.233 & 25.6 & 19.6 & 4\nl 5234.63 & 3.22 & -2.151 & 28.5 & 23.0 & 4\nl 5316.62 & 3.15 & -1.850 & 44.9 & 35.7 & 3\nl 6247.56 & 3.89 & -2.329 & 6.2 & -- & 4\nl 6456.39 & 3.90 & -2.075 & 10.1 & -- & 4\nl \multicolumn{1}{l}{Na I} & & & & & \nl 5688.21 & 2.10 & -0.370 & 8.7 & -- & 5\nl 5889.96 & 0.00 & 0.110 & 165.1 & 128.7 & 6\nl 5895.94 & 0.00 & -0.190 & 152.0 & 108.8 & 6\nl \multicolumn{1}{l}{Mg I} & & & & & \nl 4730.03 & 4.33 & -2.310 & 3.6 & -- & 7\nl 5172.70 & 2.71 & -0.320 & 292.8 & 188.7 & 8\tablenotemark{a}\nl 5183.42 & 2.72 & -0.080 & 353.2 & 203.7 & 8\tablenotemark{a}\nl 5528.42 & 4.35 & -0.360 & 85.5 & 56.8 & 7\nl 5711.09 & 4.33 & -1.730 & 18.3 & 8.3 & 7\tablebreak \multicolumn{1}{l}{Si I} & & & & & \nl 5701.12 & 4.93 & -2.050 & 2.9 & -- & 9\nl 5708.41 & 4.93 & -1.470 & -- & 3.4 & 9\nl 5948.55 & 5.08 & -1.230 & 9.2 & 6.5 & 9\nl 7415.96 & 5.61 & -0.710 & 11.5 & 5.3 & 9\nl 7423.52 & 5.61 & -0.780 & 13.4 & 7.7 & 9\nl \multicolumn{1}{l}{Ca I} & & & & & \nl 4578.56 & 2.52 & -0.700 & 13.5 & 6.6 & 10\nl 5261.71 & 2.52 & -0.580 & 18.8 & 9.9 & 10\nl 5349.47 & 2.71 & -0.310 & 20.2 & 11.2 & 10\nl 5581.97 & 2.52 & -0.560 & 25.0 & 9.5 & 10\nl 5588.76 & 2.52 & 0.360 & 62.8 & 37.4 & 10\nl 5590.12 & 2.52 & -0.570 & 18.7 & 9.7 & 10\nl 5857.46 & 2.93 & 0.240 & 35.3 & 21.8 & 10\nl 6122.23 & 1.89 & -0.320 & 72.0 & 46.8 & 11\nl 6161.30 & 2.52 & -1.270 & 4.2 & -- & 10\nl 6162.18 & 1.90 & -0.090 & 87.5 & 61.0 & 11\nl 6163.75 & 2.52 & -1.290 & 7.5 & -- & 10\nl 6166.44 & 2.52 & -1.200 & 10.6 & -- & 10\nl 6169.04 & 2.52 & -0.800 & 15.6 & 8.9 & 10\nl 6169.56 & 2.52 & -0.480 & 21.5 & 12.3 & 10\nl 6439.08 & 2.52 & 0.390 & 66.7 & 42.8 & 10\nl 6455.60 & 2.52 & -1.290 & 6.1 & -- & 10\nl 6499.65 & 2.52 & -0.820 & 15.0 & -- & 10\nl \multicolumn{1}{l}{Ti I} & & & & & \nl 4555.49 & 0.85 & -0.490 & 8.0 & 5.5 & 12\tablenotemark{a}\nl 4617.28 & 1.75 & 0.390 & 7.7 & 5.6 & 12\nl 4623.10 & 1.74 & 0.110 & 4.7 & -- & 12\nl 4681.92 & 0.05 & -1.070 & 18.7 & 10.5 & 12\nl 4840.88 & 0.90 & -0.510 & 10.2 & 5.6 & 12\nl 4885.09 & 1.89 & 0.360 & 5.0 & -- & 12\nl 4991.07 & 0.84 & 0.380 & 43.6 & 32.5 & 12\nl 4999.51 & 0.83 & 0.250 & 38.2 & 24.8 & 12\tablebreak 5020.03 & 0.84 & -0.410 & 13.7 & 8.5 & 12\nl 5024.85 & 0.82 & -0.600 & 9.5 & 5.6 & 12\nl 5064.66 & 0.05 & -0.990 & 22.8 & 15.5 & 12\nl 5173.75 & 0.00 & -1.120 & 19.0 & 11.4 & 12\tablenotemark{a}\nl 5210.39 & 0.05 & -0.880 & 27.4 & 17.5 & 12\nl 6258.71 & 1.46 & -0.270 & 5.0 & -- & 12\nl 6261.11 & 1.43 & -0.480 & 4.0 & -- & 12\nl \multicolumn{1}{l}{Ti II} & & & & & \nl 4708.67 & 1.24 & -2.210 & 13.5 & 9.5 & 12\nl 5154.08 & 1.57 & -1.920 & 22.1 & 16.5 & 12\nl 5185.91 & 1.89 & -1.350 & 21.2 & 18.6 & 12\nl 5226.54 & 1.57 & -1.300 & 47.0 & 33.3 & 12\nl 5336.79 & 1.58 & -1.700 & 26.9 & 19.6 & 12\nl \multicolumn{1}{l}{Ni I} & & & & & \nl 4756.52 & 3.48 & -0.340 & -- & 5.3 & 13\nl 4829.03 & 3.54 & -0.330 & 7.1 & -- & 13\nl 4831.18 & 3.61 & -0.420 & 6.2 & -- & 13\nl 4904.42 & 3.54 & -0.170 & -- & 5.3 & 13\nl 5017.58 & 3.54 & -0.080 & 10.2 & 6.7 & 13\nl 5081.12 & 3.85 & 0.300 & 14.6 & 7.5 & 13\nl 5476.92 & 1.83 & -0.890 & 60.3 & 46.9 & 13\nl 5754.67 & 1.93 & -2.330 & 6.4 & -- & 13\nl 6643.65 & 1.68 & -2.300 & 10.6 & 6.7 & 13\nl 6767.78 & 1.83 & -2.170 & 12.1 & 6.8 & 13\nl \tablenotetext{a}{Line not used in abundance analysis.} \tablerefs{ (1) ``Oxford" group: (a) Blackwell et al. 1979a, (b) Blackwell et al. 1979b, (c) Blackwell et al. 1982a, (d) Blackwell et al. 1980, (e) Blackwell et al. 1982b, (f) Blackwell et al. 1982c; (2) O'Brian et al. 1991; (3) Kroll \& Kock 1987; (4) Biemont et al. 1991; (5) Lambert \& Warner 1968; (6) Weise et al. 1969; (7) Fuhrmann et al. 1995; (8) Thevenin 1989; (9) Garz 1973; (10) Smith \& Raggett 1981; (11) Weise \& Martin 1980; (12) Martin et al. 1988; (13) Fuhr et al. 1988. } \enddata \end{deluxetable} \clearpage \begin{deluxetable}{lcccccccc} \footnotesize \tablenum{3} \tablecaption{Atmospheric Parameters.} \tablewidth{0pt} \tablehead{ & \colhead{$T_{eff}$} & \colhead{$\log g$} & \colhead{[Fe/H]} & \colhead{$v_t$} & \colhead{$\log \epsilon$(Fe I)} & \colhead{$\sigma$(Fe I)} & \colhead{$\log \epsilon$(Fe II)} & \colhead{$\sigma$(Fe II)}} \startdata BD +23 3130 \nl This Paper &4850 K&2.00& -2.84 & 1.35 & 4.69 & 0.04 & 4.68 & 0.08\nl This Paper, log g down &4850 K&1.70& -2.84 & 1.35 & 4.70 & 0.04 & 4.58 & 0.08\nl This Paper, log g up &4850 K&2.30& -2.84 & 1.35 & 4.67 & 0.04 & 4.78 & 0.08\nl I98 &5130 K&2.50& -2.90 & 1.00 & 5.02 & 0.09 & 4.89 & 0.08\nl \hline BD +37 1458 \nl \nl This Paper &5100 K&2.90& -2.31 & 1.40 & 5.21 & 0.04 & 5.21 & 0.08\nl This Paper, log g down &5100 K&2.60& -2.31 & 1.40 & 5.22 & 0.04 & 5.11 & 0.07\nl This Paper, log g up &5100 K&3.20& -2.31 & 1.40 & 5.20 & 0.04 & 5.32 & 0.08\nl I98 &5260 K&3.00& -2.40 & 1.00 & 5.47 & 0.10 & 5.30 & 0.08\nl B99, K93 scale &5554 K&3.62& -2.06 & 1.50 & 5.64 & 0.09 & 5.48 & 0.07\nl B99, C83 scale &5408 K&3.41& -2.14 & 1.50 & 5.50 & 0.07 & 5.40 & 0.08\nl \enddata \end{deluxetable} \clearpage \begin{deluxetable}{l|cc|cccc|c} \footnotesize \tablenum{4} \tablecaption{Derived Abundances.} \tablewidth{0pt} \tablehead{ \colhead{} & \multicolumn{2}{|c|}{BD +23 3130} & \multicolumn{4}{c|}{BD +37 1458} & \colhead{McWilliam} \nl \colhead{} & \multicolumn{1}{|c}{This Paper} & \multicolumn{1}{c|}{$\;\;$I98$\;\;$} & \multicolumn{1}{c}{This Paper} & \multicolumn{1}{c}{$\;\;$I98$\;\;$} & \multicolumn{1}{c}{B99-K93} & \multicolumn{1}{c|}{B99-C83} & \colhead{et al. 1995} } \startdata log $\epsilon$(Fe I) & 4.69 & 5.02 & 5.21 & 5.47 & 5.64 & 5.50 & \\ $[Fe/H]_{Fe I}$ & -2.83$\;$ & -2.50$\;$ & -2.31$\;$ & -2.05$\;$ & -1.88$\;$ & -2.02$\;$ &\\ $\sigma$ & 0.04 & 0.09 & 0.04 & 0.10 & 0.09 & 0.07 & \\ \hline log $\epsilon$(Fe II)& 4.68 & 4.89 & 5.21 & 5.30 & 5.48 & 5.40 & \\ $[Fe/H]_{Fe II}$ & -2.84$\;$ & -2.63$\;$ & -2.31$\;$ & -2.22$\;$ & -2.04$\;$ & -2.12$\;$ & \\ $\sigma$ & 0.08 & 0.08 & 0.08 & 0.08 & 0.07 & 0.08 & \\ \hline log $\epsilon$([O I])& 6.45 & 6.73 & 7.12 & 7.28 & 7.67 & 7.49 & \\ $[O/Fe]$ & 0.35 & 0.30 & 0.50 & 0.40 & 0.62 & 0.58 & \\ $\sigma$ & 0.20 & 0.20 & 0.20 & 0.20 & 0.20 & 0.20 & \\ \hline log $\epsilon$(Na I) & 3.54 & 3.86 & 4.07 & 4.27 & 4.32 & 4.24 & \\ $[Na/Fe]$ & 0.05 & 0.04 & 0.06 & 0.00 & -0.12$\;$ & -0.06$\;$ &$\sim 0.1$ \\ $\sigma$ & 0.06 & 0.09 & 0.10 & 0.16 & 0.13 & 0.11 & \\ \hline log $\epsilon$(Mg I) & 5.34 & 5.54 & 5.83 & 5.99 & 6.04 & 5.98 & \\ $[Mg/Fe]$ & 0.60 & 0.47 & 0.57 & 0.47 & 0.35 & 0.43 & 0.42 \\ $\sigma$ & 0.06 & 0.18 & 0.13 & 0.19 & 0.16 & 0.15 & \\ \hline log $\epsilon$(Si I) & 5.43 & 5.55 & 5.87 & 5.94 & 6.05 & 6.00 & \\ $[Si/Fe]$ & 0.72 & 0.51 & 0.64 & 0.45 & 0.39 & 0.48 & 0.50 \\ $\sigma$ & 0.17 & 0.19 & 0.12 & 0.14 & 0.13 & 0.12 & \\ \hline log $\epsilon$(Ca I) & 3.98 & 4.17 & 4.52 & 4.67 & 4.77 & 4.70 & \\ $[Ca/Fe]$ & 0.52 & 0.38 & 0.54 & 0.43 & 0.36 & 0.43 & 0.43\\ $\sigma$ & 0.09 & 0.12 & 0.12 & 0.16 & 0.14 & 0.13 & \\ \hline log $\epsilon$(Ti) & 2.34 & 2.65 & 2.87 & 3.05 & 3.32 & 3.18 & \\ $[Ti/Fe]_{Ti}$ & 0.19 & 0.17 & 0.20 & 0.12 & 0.22 & 0.22 & 0.32\\ $\sigma$ & 0.10 & 0.12 & 0.09 & 0.13 & 0.12 & 0.11 & \\ \hline log $\epsilon$(Ni I) & 3.46 & 3.72 & 3.96 & 4.13 & 4.34 & 4.22 & \\ $[Ni/Fe]$ & 0.05 & -0.02$\;$ & 0.03 & -0.06$\;$ & -0.02$\;$ & 0.00 & $\sim$0.0\\ $\sigma$ & 0.06 & 0.11 & 0.08 & 0.15 & 0.15 & 0.12 & \\ \enddata \end{deluxetable} \clearpage \begin{deluxetable}{l|cc|cccc|c} \footnotesize \tablenum{5} \tablecaption{Derived Abundances.} \tablewidth{0pt} \tablehead{ \colhead{} & \multicolumn{2}{|c|}{BD +23 3130} & \multicolumn{4}{c|}{BD +37 1458} & \colhead{McWilliam} \nl \colhead{} & \multicolumn{1}{|c}{This Paper} & \multicolumn{1}{c|}{$\;\;$I98$\;\;$} & \multicolumn{1}{c}{This Paper} & \multicolumn{1}{c}{$\;\;$I98$\;\;$} & \multicolumn{1}{c}{B99-K93} & \multicolumn{1}{c|}{B99-C83} & \colhead{et al. 1995} } \startdata log $\epsilon$(Fe I) & 4.69 & 4.62 & 5.21 & 5.12 & 5.46 & 5.38 &\\ $[Fe/H]_{Fe I}$ & -2.83$\;$ & -2.90$\;$ & -2.31$\;$ & -2.40$\;$ & -2.06$\;$ & -2.14$\;$ &\\ \hline log $\epsilon$(Fe II)& 4.68 & 4.62 & 5.21 & 5.12 & 5.46 & 5.38 &\\ $[Fe/H]_{Fe II}$ & -2.84$\;$ & -2.90$\;$ & -2.31$\;$ & -2.40$\;$ & -2.06$\;$ & -2.14$\;$ &\\ \hline log $\epsilon$([O I])& 6.45 & 6.73 & 7.12 & 7.28 & 7.67 & 7.49 &\\ $[O/Fe]$ & 0.35 & 0.70 & 0.50 & 0.75 & 0.80 & 0.70 &\\ $\sigma$ & 0.20 & 0.20 & 0.20 & 0.20 & 0.20 & 0.20 &\\ \hline log $\epsilon$(Na I) & 3.54 & 3.86 & 4.07 & 4.27 & 4.32 & 4.24 &\\ $[Na/Fe]$ & 0.05 & 0.44 & 0.06 & 0.35 & 0.06 & 0.06 & $\sim$0.1\\ $\sigma$ & 0.05 & 0.02 & 0.09 & 0.12 & 0.09 & 0.09 &\\ \hline log $\epsilon$(Mg I) & 5.34 & 5.54 & 5.83 & 5.99 & 5.98 & 6.04 &\\ $[Mg/Fe]$ & 0.60 & 0.87 & 0.57 & 0.82 & 0.53 & 0.55 & 0.42\\ $\sigma$ & 0.05 & 0.16 & 0.12 & 0.16 & 0.13 & 0.13 &\\ \hline log $\epsilon$(Si I) & 5.43 & 5.55 & 5.87 & 5.94 & 6.05 & 6.00 &\\ $[Si/Fe]$ & 0.72 & 0.91 & 0.64 & 0.80 & 0.57 & 0.60 & 0.50\\ $\sigma$ & 0.17 & 0.17 & 0.11 & 0.10 & 0.10 & 0.10 &\\ \hline log $\epsilon$(Ca I) & 3.98 & 4.17 & 4.52 & 4.67 & 4.77 & 4.70 &\\ $[Ca/Fe]$ & 0.52 & 0.78 & 0.54 & 0.78 & 0.54 & 0.55 & 0.43\\ $\sigma$ & 0.08 & 0.08 & 0.11 & 0.13 & 0.11 & 0.11 &\\ \hline log $\epsilon$(Ti) & 2.34 & 2.65 & 2.87 & 3.05 & 3.32 & 3.18 &\\ $[Ti/Fe]_{Ti}$ & 0.19 & 0.57 & 0.20 & 0.47 & 0.40 & 0.34 & 0.32\\ $\sigma$ & 0.09 & 0.07 & 0.08 & 0.09 & 0.09 & 0.08 &\\ \hline log $\epsilon$(Ni I) & 3.46 & 3.72 & 3.96 & 4.13 & 4.34 & 4.22 &\\ $[Ni/Fe]$ & 0.05 & 0.38 & 0.03 & 0.29 & 0.16 & 0.12 & $\sim$0.0\\ $\sigma$ & 0.05 & 0.07 & 0.07 & 0.11 & 0.12 & 0.10 &\\ \enddata \end{deluxetable} \clearpage \noindent {\bf Figure 1a}. Plots of abundance vs. excitation potential and line strength for the individual Fe I lines of BD +23 3130 using the stellar parameters adopted by this paper. \noindent {\bf Figure 1b}. Same as Figure 1a, expect for BD +37 1458. \noindent {\bf Figure 2a}. Same as Figure 1a, except using the parameters of I98. \noindent {\bf Figure 2b}. Same as Figure 1b, except using the parameters of I98. \noindent {\bf Figure 2c}. Same as Figure 1b, except using the parameters of B99 (King 1993 scale). \noindent {\bf Figure 2d}. Same as Figure 1b, except using the parameters of B99 (Carney 1983 scale). \noindent {\bf Figure 3a}. Observed and synthetic spectra of Fe I lines in the $\lambda$ 6137 \AA$\;$ region of BD +23 3130. The dots are the observed data points (binned 2X), and the lines are the synthetic spectra using the stellar parameters adopted by this paper (solid) and by I98 (dotted). \noindent {\bf Figure 3b}. Same as Figure 3a, except for BD +37 1458. The lines are the synthetic spectra for the stellar parameters adopted by this paper (solid), I98 (dotted), B99 using the King 1993 scale (short dash), and B99 using the Carney 1983 scale (long dash). \noindent {\bf Figure 4a}. Observed and synthetic spectra for the $\lambda$ 6300 \AA$\;$ region of BD +23 3130. The dots are the observed data points (binned 2x) and the lines are the synthetic spectra for the stellar parameters adopted by this paper for [O/Fe] = +0.0, +0.5, +1.0 and +1.5. \noindent {\bf Figure 4b}. Same as Figure 4a, except for BD +37 1458. \noindent {\bf Figure 5a}. Same as Figure 4a, except using the parameters of I98. \noindent {\bf Figure 5b}. Same as Figure 4b, except using the parameters of I98. \noindent {\bf Figure 5c}. Same as Figure 4b, except using the parameters of B99 (King 1993 scale). \noindent {\bf Figure 5d}. Same as Figure 4b, except using the parameters of B99 (Carney 1983 scale). \plotone{Fulbright.fig1a.ps} \plotone{Fulbright.fig1b.ps} \plotone{Fulbright.fig2a.ps} \plotone{Fulbright.fig2b.ps} \plotone{Fulbright.fig2c.ps} \plotone{Fulbright.fig2d.ps} \plotone{Fulbright.fig3a.ps} \plotone{Fulbright.fig3b.ps} \plotone{Fulbright.fig4a.ps} \plotone{Fulbright.fig4b.ps} \plotone{Fulbright.fig5a.ps} \plotone{Fulbright.fig5b.ps} \plotone{Fulbright.fig5c.ps} \plotone{Fulbright.fig5d.ps}
\section{Introduction } \label{sec-introd} \andy{intro} The evolution law in quantum mechanics is governed by unitary operators \cite{Dirac}. This entails, by virtue of very general mathematical properties, that the decay of an unstable quantum system cannot be purely exponential. In general, a rigorous analysis based on the Schr\"odinger equation shows that the decay law is quadratic for very short times \cite{shortt} and governed by a power law for very long times \cite{longtt}. These features of the quantum evolution are well known and discussed in textbooks of quantum mechanics \cite{Sakurai} and quantum field theory \cite{Brown}. The temporal behavior of quantum systems is reviewed in Ref.\ \cite{temprevi}. Although the domain of validity of the exponential law is limited, the Fermi ``Golden Rule" \cite{Fermigold} works very well and no deviations from the exponential behavior have ever been observed for truly unstable systems \cite{expttlaw}. The quantum mechanical derivation of this law is based on the sensible idea that the temporal evolution of a quantum system is dominated by a pole near the real axis of the complex energy plane (Weisskopf-Wigner approximation \cite{seminal}). This yields an irreversible evolution, characterized by a master equation and exponential decay \cite{VanKampen}. An important contribution to this issue was given in the 50's by Van Hove \cite{vanHove}, who rigorously showed that it is possible to obtain a master equation (leading to exponential behavior) for a quantum mechanical system endowed with many (infinite) degrees of freedom, by making use of the so-called ``$\lambda^2 t$" limit. The crucial idea is to consider the limit \andy{vHlimit} \begin{equation} \lambda\to0 \quad\mbox{keeping}\quad \tilde t = \lambda^2t\quad\mbox{finite ($\lambda$-independent constant)}, \label{eq:vHlimit} \end{equation} where $\lambda$ is the coupling constant and $t$ time. One then looks at the evolution of the quantum system as a function of the rescaled time $\tilde t$. There has recently been a renewed interest in the physical literature for this time-scale transformation and its subtle mathematical features: see \cite{Accardi}. The purpose of this paper is to consider the effects that arise when the coupling constant is small but nonvanishing. This will enable us to give general estimates for deviations from exponential behavior. The paper is organized as follows. We shall first look, in Section 2, at a simple system: we summarize some recent results on a characteristic transition of the hydrogen atom in the two-level approximation. In Section 3 we consider the action of the Van Hove limiting procedure on a generic two-level atom in the rotating-wave approximation and then generalize our result when the other discrete levels and the counter-rotating terms are taken into account. We look in particular at the scaling procedure from the perspective of the complex energy plane, rather than in terms of the time variable. This enables us to pin down the different sources of non-exponential behavior. In Section 4 our analysis is extended to a general field-theoretical framework: general estimates are given of all deviations from the exponential law (both at short and long times) at leading orders in the coupling constant. \setcounter{equation}{0} \section{Hydrogen atom in the two-level approximation } \label{sec-2levels} \andy{2levels} We start our considerations from a simple field-theoretical model. Consider the Hamiltonian ($\hbar=c=1$) \andy{tothaml1,2,3} \begin{eqnarray} H & = & H_0 + \lambda V , \label{eq:tothaml1} \\ H_0 & = & H_{\rm atom} + H_{\rm EM} \nonumber \\ & \equiv & \omega_0 b^\dagger_2 b_2 + \sum_\beta \int_0^\infty d\omega \, \omega a^\dagger_{\omega\beta} a_{\omega\beta} , \label{eq:tothaml2} \\ V & = & \sum_\beta \int_0^\infty d\omega \left[ \varphi_\beta(\omega) b^\dagger_1 b_2 a^\dagger_{\omega \beta} + \varphi^*_\beta(\omega) b^\dagger_2 b_1 a_{\omega \beta} \right], \label{eq:tothaml3} \end{eqnarray} where $H_{\rm atom}$ is the free Hamiltonian of a two-level atom ($\omega_0$ being the energy gap between the two atomic levels), $b_j, b_j^\dagger$ are the annihilation and creation operators of the atomic level $j$, obeying anticommutation relations \andy{fermicomm} \begin{equation} \{b_k, b^\dagger_\ell \} = \delta_{k\ell} \qquad (k,\ell=1,2), \label{eq:fermicomm} \end{equation} $H_{\rm EM}$ is the Hamiltonian of the free EM field, $\lambda$ the coupling constant and $V$ the interaction Hamiltonian. We are working in the rotating-wave approximation and with the energy-angular momentum basis for photons \cite{Heitler}, with $\sum_\beta = \sum_{j = 1}^\infty\sum_{m=-j}^j\sum_{\epsilon = 0}^1 $, where $\epsilon$ defines the photon parity $P=(-1)^{j+1+\epsilon}$, $j$ is the total angular momentum (orbital+spin) of the photon, $m$ its magnetic quantum number and \andy{boscomm2} \begin{equation} [a_{\omega j m \epsilon}, a^\dagger_{\omega' j' m'\epsilon'}] = \delta(\omega-\omega')\delta_{j j'}\delta_{m m'}\delta_{\epsilon\epsilon'}. \label{eq:boscomm2} \end{equation} We shall focus our attention on the 2P-1S transition of hydrogen, so that $\omega_0 = \frac{3}{8} \alpha^2 m_e \simeq 1.550 \cdot 10^{16}$ rad/s ($\alpha$ is the fine structure constant and $m_e$ the electron mass) and the matrix elements $\varphi_{\beta}(\omega)$ of the interaction are known exactly \cite{Moses,Seke} \andy{phidef1} \begin{eqnarray} \varphi_\beta (\omega) &=& \langle 1; 1_{\omega\beta}|V|2;0\rangle = \varphi_{\bar{\beta}}(\omega)\delta_{\beta\bar{\beta}}\nonumber\\ &=& i(\Lambda)^\frac{1}{2}\frac{\left( \frac{\omega}{\Lambda}\right)^\frac{1}{2}}{\left[1 + \left( \frac{\omega}{\Lambda}\right)^2 \right]^2} \delta_{j1}\delta_{mm_2}\delta_{\epsilon 1}, \label{eq:phidef1} \end{eqnarray} where \andy{stati} \begin{equation} |1; 1_{\omega\beta}\rangle \equiv |1\rangle \otimes |\omega, j, m, \epsilon \rangle, \quad |2; 0\rangle \equiv |2\rangle \otimes |0\rangle \label{eq:stati} \end{equation} (the first ket refers to the atom and the second to the photon) and the selection rule, due to angular momentum and parity conservation, entails that the only nonvanishing term in (\ref{eq:tothaml3}) and (\ref{eq:phidef1}) is $\bar{\beta} = (1, m_2, 1)$. We emphasize that the so-called "retardation effects" are taken into account in (\ref{eq:phidef1}). The normalization reads \andy{normaliz} \begin{equation} \langle 1; 1_{\omega\beta} |1; 1_{\omega'\beta'}\rangle = \delta (\omega-\omega') \delta_{\beta\beta'}, \quad \langle 2;0 |2; 0\rangle =1 \label{eq:normaliz} \end{equation} and the quantities \andy{lambdachi} \begin{eqnarray} \Lambda &=& \frac{3}{2} \alpha m_e = \frac{3}{2a_0} \simeq 8.498\cdot 10^{18} \mbox{rad/s}, \nonumber \\ \lambda &=& \left( \frac{2}{\pi} \right)^{1/2} \left(\frac{2}{3}\right)^{9/2} \alpha^{3/2} \simeq .802 \cdot 10^{-4}, \label{eq:lambdachi} \end{eqnarray} are the natural cutoff of the atomic form factor, expressed in terms of the Bohr radius $a_0$, and the coupling constant, respectively. Observe that there are no free parameters in (\ref{eq:tothaml1})-(\ref{eq:tothaml3}). The above model was analyzed in a previous paper \cite{FP1}, where we mainly concentrated our attention on the deviations from exponential, both at short and long times. There is interesting related work on this subject \cite{KnightMilonni'76}. Let us summarize the main results, by concentrating our attention on the role played by the coupling constant $\lambda$. Assume one can prepare (at time $t=0$) the atom in the initial state $|2;0\rangle$. This is an eigenstate of the unperturbed Hamiltonian $H_0$, whose eigenvalue is $\omega_0$. The evolution is governed by the unitary operator $U(t) = \exp (-iHt)$ and the ``survival" or nondecay amplitude and probability at time $t$ are defined as (interaction picture) \andy{survampl,ndq1} \begin{eqnarray} {\cal A}(t) & = & \langle 2;0|e^{i H_0 t}U(t)|2;0\rangle , \label{eq:survampl} \\ P(t) & = & |\langle 2;0 |e^{i H_0 t}U(t) |2;0 \rangle |^2 . \label{eq:ndq1} \end{eqnarray} The survival probability at short times reads \andy{ndq, naiedef} \begin{equation} P(t) = 1 - t^2/\tau_{\rm Z}^2 + \cdots, \qquad \tau_{\rm Z} \equiv (\lambda^2\langle 2;0|V^2|2;0\rangle)^{-1/2} . \label{eq:naiedef} \end{equation} The quantity $\tau_{\rm Z}$ is the so-called ``Zeno time" and yields a quantitative estimate of the deviation from exponential at very short times. Strictly speaking, $\tau_{\rm Z}$ is the convexity of $P(t)$ in the origin. One finds \cite{FP1} \andy{taudet} \begin{equation} \tau_{\rm Z} = \frac{\sqrt{6}}{\lambda\Lambda} = (3 \pi)^{\frac{1}{2}} \left(\frac{3}{2}\right)^{\frac{7}{2}} \frac{1}{\alpha^{\frac{5}{2}} m_e} \simeq 3.593 \cdot 10^{-15} \mbox{s}. \label{eq:taudet} \end{equation} It is possible to obtain a closed expression for ${\cal A}(t)$, valid at all times, as an inverse Laplace transform: \andy{survamp1,qs3} \begin{eqnarray} {\cal A}(t) &=& \frac{e^{i\omega_0 t}}{2 \pi i} \int_{\rm B} ds \frac{e^{s \Lambda t}}{s + i \frac{\omega_0}{\Lambda} + \lambda^2 Q(s)}, \label{eq:survamp1} \\ & & \quad Q(s) \equiv -i \int_0^\infty dx \frac{x}{(1+x^2)^4} \frac{1}{x-is} , \label{eq:qs3} \end{eqnarray} where B is the Bromwich path, i.e.\ a vertical line at the right of all the singularities of the integrand. $Q$ is a self-energy contribution and can be computed exactly: \andy{qs4} \begin{eqnarray} Q(s) &=& \frac{-15\pi i -(88-48\pi i)s - 45\pi is^2 + 144s^3} {96(s^2 - 1)^4}\nonumber\\ & &+\frac{15\pi is^4- 72s^5 - 3\pi is^6 + 16s^7-96 s\log s}{96(s^2 - 1)^4} . \label{eq:qs4} \end{eqnarray} At short and long times one gets \andy{shortt,largetimes1} \begin{eqnarray} \!\! P(t) \! &\sim& \! 1 - \frac{t^2}{\tau_{\rm Z}^2} \qquad (t\ll \tau_{\rm Z}), \label{eq:shortt} \\ \!\! P(t) \! &\sim& \! {\cal Z}^2 e^{-\gamma t}+ \lambda^4\frac{{\cal C}^2}{(\omega_0 t)^4} - 2 \lambda^2\frac{{\cal C}{\cal Z}}{(\omega_0 t)^2} e^{-\frac{\gamma}{2} t} \cos\left[(\omega_0-\Delta E) t -\zeta \right] \label{eq:largetimes1} \\ & & \qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad(t\gg \Lambda^{-1}), \nonumber \end{eqnarray} where \andy{fgr, shift,Zz,Cchi} \begin{eqnarray} \gamma &=& 2\pi \lambda^2|\varphi_{\bar{\beta}}(\omega_0)|^2 + {\rm O}(\lambda^4) = 2\pi\lambda^2\omega_0 +{\rm O}(\lambda^4) \simeq 6.268\cdot10^8\mbox{s}^{-1}, \label{eq:fgr}\nonumber\\ \\ \Delta E &=&\lambda^2 {\cal P}\! \!\int_0^\infty d\omega |\varphi_{\bar{\beta}}(\omega)|^2 \frac{1}{\omega -\omega_0}+{\rm O}(\lambda^4)\simeq 0.5 \lambda^2 \Lambda, \label{eq:shift} \\ {\cal Z} e^{i\zeta} &\simeq& (1-4.38\lambda^2)e^{-i1.00\pi\lambda^2}=1 + {\rm O}(\lambda^2), \label{eq:Zz} \\ {\cal C} & \simeq& 1+5.38\lambda^2 = 1 + {\rm O}(\lambda^2). \label{eq:Cchi} \end{eqnarray} The first two formulae give the Fermi ``Golden Rule," yielding the lifetime \andy{fgrule} \begin{equation} \tau_{\rm E}=\gamma^{-1} \simeq 1.595\cdot 10^{-9}\mbox{s}, \label{eq:fgrule} \end{equation} and the second order correction to the energy level $\omega_0$. The exact expressions for the quantities (\ref{eq:fgr})-(\ref{eq:Cchi}) are given in Ref.\ \cite{FP1}. \setcounter{equation}{0} \section{Van Hove's limit } \label{sec-VHL} \andy{VHL} Let us look at Van Hove's $``\lambda^2 t"$ limiting procedure applied to the model of the previous section. Before proceeding to a detailed analysis, it is worth putting forward a few preliminary remarks: we shall scrutinize (in terms of the coupling constant) the mechanisms that make the nonexponential contributions in (\ref{eq:shortt})-(\ref{eq:largetimes1}) vanish. To this end, observe first that as $\lambda \to 0$ the Zeno time (\ref{eq:taudet}) diverges, while the {\em rescaled} Zeno time vanishes \andy{rescz} \begin{equation} \tilde\tau_{\rm Z} \equiv \lambda^2\tau_{\rm Z}=\lambda\frac{\sqrt{6}}{\Lambda} ={\rm O}(\lambda). \label{eq:rescz} \end{equation} On the other hand, the rescaled lifetime (\ref{eq:fgrule}) remains constant [see (\ref{eq:fgr})]: \andy{resce} \begin{equation} \tilde\tau_{\rm E} \equiv \lambda^2\tau_{\rm E}=\frac{1}{2\pi\omega_0}={\rm O}(1). \label{eq:resce} \end{equation} Moreover, the transition to a power law occurs when the first two terms in the right hand side of (\ref{eq:largetimes1}) are comparable, so that \andy{nuovapot} \begin{equation} (\omega_0 t)^2e^{-\frac{\gamma}{2}t}\simeq\lambda^2, \label{eq:nuovapot} \end{equation} because both ${\cal C}$ and ${\cal Z}$ are $\simeq 1$. In the limit of small $\lambda$, (\ref{eq:nuovapot}) yields $t = \tau_{\rm pow}$, with \andy{interr1} \begin{equation} 2 \log (\omega_0 \tau_{\rm pow}) - \frac{\gamma}{2}\tau_{\rm pow} \simeq 2 \log \lambda, \label{eq:interr1} \end{equation} namely, by (\ref{eq:fgr}), \andy{interr2} \begin{equation} \frac{\tau_{\rm pow}}{\tau_{\rm E}} \simeq 4\log \frac{1}{\lambda} + 4\log \frac{\tau_{\rm pow}}{2\pi \lambda^2 \tau_{\rm E}} = 12\log \frac{1}{\lambda} + 4\log \frac{\tau_{\rm pow}}{\tau_{\rm E}} + 4 \log \frac{1}{2\pi}. \label{eq:interr2} \end{equation} Therefore, when time is rescaled, \andy{rescpl} \begin{equation} \tilde\tau_{\rm pow} \equiv \lambda^2\tau_{\rm pow}=12\tilde\tau_{\rm E} \log\frac{1}{\lambda}+{\rm O}\left(\log\log\frac{1}{\lambda} \right)= {\rm O}\left(\log\frac{1}{\lambda}\right) . \label{eq:rescpl} \end{equation} Finally, the power contributions are $\sim {\rm O}(\lambda^{3\alpha})\tilde t^{-\alpha} \,(\alpha=2,4)$, the period of the oscillations [last term in (\ref{eq:largetimes1})] behaves like $\lambda^2/\omega_0$ and the quantities (\ref{eq:Zz})-(\ref{eq:Cchi}) become both unity. In conclusion, only the exponential law survives in the limit (\ref{eq:vHlimit}), with the correct normalization factor (${\cal Z}=1$), and one is able to derive a purely exponential behavior (Markovian dynamics) from the quantum mechanical Schr\"{o}dinger equation (unitary dynamics). It is important to notice that, in order to obtain the exponential law, a normalizable state (such as a wave packet) must be taken as initial state. Our initial state $|2;0\rangle$ is indeed normalizable: see (\ref{eq:normaliz}). \subsection{Two-level atom in the rotating-wave approximation} \label{sec-VH2lev} \andy{VH2lev} Let us now proceed to a more formal analysis. Write the evolution operator as \andy{survFT} \begin{equation} \label{eq:survFT} U(t)=\frac{i}{2\pi}\int_{\rm C}dE\frac{e^{-iEt}}{E-H}, \end{equation} where the path C is a straight horizontal line just above the real axis (this is the equivalent of the Bromwich path in the Laplace plane). By defining the resolvents ($\Im E>0$) \begin{equation} S(E)\equiv\langle 2;0|\frac{1}{E-H_0}|2;0\rangle=\frac{1}{E-\omega_0}, \qquad S'(E)\equiv\langle 2;0|\frac{1}{E-H}|2;0\rangle, \end{equation} Dyson's resummation reads \begin{equation} S'(E)=S(E)+\lambda^2S(E)\Sigma^{(2)}(E)S(E)+\lambda^4S(E)\Sigma^{(2)}(E)S(E) \Sigma^{(2)}(E)S(E)+\dots \end{equation} where $\Sigma^{(2)}(E) = \langle 2;0|V(E-H_0)^{-1}V|2;0\rangle$ is the 1-particle irreducible self-energy function. In the rotating-wave approximation $\Sigma^{(2)}(E)$ consists only of a second order diagram and can be evaluated exactly: \begin{equation} \Sigma^{(2)}(E)\equiv\int_0^\infty d\omega\frac{|\varphi(\omega)|^2}{E-\omega} =i\Lambda Q(-i E/\Lambda), \end{equation} where the matrix element $\varphi=\varphi_{\bar{\beta}}$ in (\ref{eq:phidef1}) and $Q$ is the function in (\ref{eq:qs4}). In the complex $E$-plane $\Sigma^{(2)}(E)$ has a branch cut running from 0 to $\infty$, a branching point in the origin and no singularity on the first Riemann sheet. Summing the series \andy{sumser} \begin{equation} S'(E)=\frac{1}{S(E)^{-1}-\lambda^2\Sigma^{(2)}(E)}= \frac{1}{E-\omega_0-\lambda^2\Sigma^{(2)}(E)}, \label{eq:sumser} \end{equation} we obtain for the survival amplitude \andy{survE} \begin{eqnarray} {\cal A}(t)&\equiv&\langle2;0|e^{i H_0 t} U(t)|2;0\rangle= \frac{i}{2\pi}\int_{\rm C}dE e^{-iEt} S'(E+\omega_0)\nonumber\\ &=& \frac{i}{2\pi}\int_{\rm C}dE\frac{e^{-iEt}} {E-\lambda^2\Sigma^{(2)}(E+\omega_0)} . \label{eq:survE} \end{eqnarray} In Van Hove's limit one looks at the evolution of the system over time intervals of order $t=\tilde t/\lambda^2$ ($\tilde t$ independent of $\lambda$), in the limit of small $\lambda$. Our purpose is to see how this limit works in the complex-energy plane, i.e.\ what is the limiting form of the propagator. To this end, by rescaling time $\tilde t\equiv \lambda^2 t$, we can write \andy{survEsc} \begin{equation} {\cal A}\left(\frac{\tilde{t}}{\lambda^2}\right)= \frac{i}{2\pi}\int_{\rm C} d\tilde E\frac{e^{-i\tilde{E}\tilde t}} {\tilde E-\Sigma^{(2)}(\lambda^2\tilde E+\omega_0)}, \label{eq:survEsc} \end{equation} where we are naturally led to introduce the rescaled energy $\widetilde E\equiv E/\lambda^2$. Taking the Van Hove limit we get \andy{survEsclim} \begin{equation} \widetilde {\cal A}(\tilde{t})\equiv \lim_{\lambda\to0}{\cal A}\left(\frac{\tilde{t}}{\lambda^2}\right) =\frac{i}{2\pi}\int_{\rm C} d\tilde E e^{-i\tilde{E}\tilde t}\widetilde S'(\tilde E). \label{eq:survEsclim} \end{equation} \begin{figure}[t] \epsfig{file=cmplxe.eps, width=13.5cm} \caption{Singularities of the propagator (\ref{eq:survE}) in the complex-$E$ plane. The first Riemann sheet ({\rm I}) is singularity free. The logarithmic cut is due to $\Sigma^{(2)}(E)$ and the pole is located on the second Riemann sheet ({\rm II}). In the complex-$\tilde E$ plane, the pole has coordinates (\ref{eq:polecoord1})-(\ref{eq:polecoord2}). } \label{fig:cmplxe} \andy{sigma4} \end{figure} where the propagator in the rescaled energy reads \andy{rescS} \begin{equation} \widetilde S'(\tilde E)\equiv\lim_{\lambda\to0} \frac{1}{\tilde E-\Sigma^{(2)}(\lambda^2\tilde E+\omega_0)} =\frac{1}{\tilde E-\Sigma^{(2)}(\omega_0+i0^+)}, \label{eq:rescS} \end{equation} the term $+i0^+$ being due to the fact that $\Im\tilde E>0$. The self-energy function in the $\lambda \to 0$ limit becomes \andy{sigmaomega0} \begin{equation} \label{eq:sigmaomega0} \Sigma^{(2)}(\omega_0+i0^+)=-\int_0^\infty d\omega\frac{|\varphi(\omega)|^2} {\omega-\omega_0-i0^+}=\Delta(\omega_0)-\frac{i}{2}\Gamma(\omega_0) \end{equation} where \andy{polecoord1,2} \begin{eqnarray} \Delta(\omega_0)&\equiv& {\cal P}\int_0^\infty d\omega\frac{|\varphi(\omega)|^2} {\omega_0-\omega}, \label{eq:polecoord1} \\ \Gamma(\omega_0)&\equiv& 2\pi|\varphi(\omega_0)|^2, \label{eq:polecoord2} \end{eqnarray} which yields a purely exponential decay (Weisskopf-Wigner approximation and Fermi Golden Rule). In Figure \ref{fig:cmplxe} we endeavoured to clarify the role played by the time-energy rescaling in the complex-$E$ plane. One can get a more detailed understanding of the mechanisms that underpin the limiting procedure by looking at higher order terms in the coupling constant. The pole of the original propagator (\ref{eq:sumser}) satisfies the equation \begin{equation} E_{\rm pole}-\lambda^2\Sigma^{(2)}(E_{\rm pole}+\omega_0)=0 , \end{equation} which can be solved by expanding the self-energy function around $E=\omega_0$ in power series \begin{equation} \Sigma^{(2)}(E+\omega_0)=\Sigma^{(2)}(\omega_0)+ E \Sigma^{(2)'}(\omega_0) +\frac{E^2}{2} \Sigma^{(2)''}(\omega_0) + \dots, \end{equation} whose radius of convergence is $\omega_0$, due to the branching point of $\Sigma^{(2)}$ in the origin. We get (iteratively) \begin{equation} E_{\rm pole}=\lambda^2 \Sigma^{(2)}(\omega_0)+\lambda^4 \Sigma^{(2)'} (\omega_0)\Sigma^{(2)}(\omega_0)+{\rm O}(\lambda^6), \end{equation} which, due to (\ref{eq:sigmaomega0}), becomes \andy{poledef} \begin{equation} \label{eq:poledef} E_{\rm pole}\equiv \Delta E-\frac{i}{2}\gamma=\lambda^2\Delta(\omega_0)- i\frac{\lambda^2}{2}\Gamma(\omega_0)+ {\rm O}(\lambda^4). \end{equation} In the rescaled energy (\ref{eq:poledef}) reads \begin{equation} \tilde E_{\rm pole}=\frac{E_{\rm pole}}{\lambda^2}=\Delta(\omega_0)- \frac{i}{2}\Gamma(\omega_0)+ {\rm O}(\lambda^2)\stackrel{\lambda\rightarrow 0} {\longrightarrow}\Delta(\omega_0)-\frac{i}{2}\Gamma(\omega_0), \end{equation} which is the same as (\ref{eq:sigmaomega0}). This is again the Fermi Golden Rule. \subsection{$N$-level atom with counter-rotating terms } \label{sec-VHNlev} \andy{VHNlev} Before proceeding to a general analysis it is interesting to see how the above model is modified by the presence of the other atomic levels and the inclusion of counter-rotating terms in the interaction Hamiltonian. This will enable us to pin down other salient features of the $\lambda^2t$ limit. The Hamiltonian is \andy{newHV} \begin{equation} H=H_0^\prime+\lambda V^\prime, \label{eq:newHV} \end{equation} where \andy{newH0,V} \begin{eqnarray} H_0^\prime & \equiv & \sum_\nu \omega_\nu b^\dagger_\nu b_\nu + \sum_\beta \int_0^\infty d\omega \, \omega a^\dagger_{\omega\beta} a_{\omega\beta} , \label{eq:newH0} \\ V^\prime & = & \sum_{\mu,\nu} \sum_\beta \int_0^\infty d\omega \left[ \varphi^{\mu\nu}_\beta(\omega) b^\dagger_\mu b_{\nu} a^\dagger_{\omega \beta} + \varphi^{\mu\nu*}_\beta(\omega) b^\dagger_{\nu} b_\mu a_{\omega \beta} \right], \label{eq:newV} \end{eqnarray} where $\nu$ runs over all the atomic states and $b^\dagger_\nu, b_\nu$ and $a^\dagger_{\omega\beta}, a_{\omega\beta}$ satisfy anticommutation and commutation relations, respectively. [The Hamiltonian (\ref{eq:tothaml1})-(\ref{eq:tothaml3}) is recovered if we set $\omega_2=\omega_0, \; \omega_1=0$ and neglect the counter-rotating terms.] Starting from the initial state $|\mu;0\rangle$, Dyson's resummation yields \andy{NBorn} \begin{equation}\label{eq:NBorn} S'(E)=\frac{1}{S(E)^{-1}-\lambda^2\Sigma(E)}= \frac{1}{E-\omega_\mu-\lambda^2\Sigma(E)} \end{equation} and the 1-particle irreducible self-energy function takes the form \andy{SEF} \begin{equation} \Sigma(E)=\Sigma^{(2)}(E)+\lambda^2\Sigma^{(4)}(E)+\dots, \label{eq:SEF} \end{equation} with \andy{gensigma2} \begin{equation}\label{eq:gensigma2} \Sigma^{(2)}(E)\equiv\sum_{\nu,\beta}\int_0^\infty d\omega \frac{|\varphi^{\nu\mu}_\beta(\omega)|^2}{E-\omega_\nu-\omega} . \end{equation} Both $\Sigma^{(2)}$ and $\Sigma^{(4)}$ are shown as Feynman diagrams in Figure \ref{fig:sigma4}. \begin{figure \epsfig{file=auto.eps, width=13.5cm} \caption{Graphic representation of (\ref{eq:SEF}): $\Sigma^{(2)}$ and $\Sigma^{(4)}$ are in the first and second line, respectively.} \label{fig:sigma4} \andy{sigma4} \end{figure} In the Van Hove limit one obtains \andy{sigma2lim} \begin{equation} \Sigma(\lambda^2\tilde E+\omega_\mu)\stackrel{\lambda\rightarrow 0} {\longrightarrow}\Sigma^{(2)}(\lambda^2\tilde E+\omega_\mu)\Big|_ {\lambda=0}=\Sigma^{(2)}(\omega_\mu+i0^+). \label{eq:sigma2lim} \end{equation} The propagator in the rescaled energy takes now the form \begin{equation} \widetilde S'(\tilde E)=\lim_{\lambda\to0} \frac{1}{\tilde E-\Sigma^{(2)} (\lambda^2\tilde E+\omega_\mu) +{\rm O}(\lambda^2)} =\frac{1}{\tilde E-\Sigma^{(2)}(\omega_\mu+i0^+)}, \end{equation} where \begin{equation} \Sigma^{(2)}(\omega_\mu+i0^+)=\sum_{\nu,\beta}\int_0^\infty d\omega \frac{|\varphi^{\nu\mu}_\beta(\omega)|^2}{\omega_\mu-\omega_\nu-\omega+i0^+}. \end{equation} The last two equations correspond to (\ref{eq:rescS})-(\ref{eq:sigmaomega0}): the propagator reduces to that of a generalized rotating-wave approximation. We see that the Van Hove limit works by following two logical steps. First, it constrains the evolution in a Tamm-Dancoff sector: the system can only ``explore" those states that are directly related to the initial state $\mu$ by the interaction $V'$. In other words, in this limit, the ``excitation number" ${\cal N}_\mu \equiv b^\dagger_\mu b_\mu + \sum_{\beta, \omega} a^\dagger_{\omega\beta} a_{\omega\beta}$ becomes a conserved quantity (even though the original Hamiltonian contains counter-rotating terms) and, as a consequence, the self-energy function consists only of a second order contribution that can be evaluated exactly. Second, it reduces this second order contribution, which depends on energy as in (\ref{eq:gensigma2}), to a constant (its value in the energy $\omega_\mu$ of the initial state), like in (\ref{eq:sigma2lim}). Hence the analytical properties of the propagator, which had branch-cut singularities, reduce to those of a single complex pole, whose imaginary part (responsible for exponential decay) yields the Fermi Golden rule, evaluated at second order of perturbation theory. Notice that it is the latter step (and not the former one) which is strictly necessary to obtain a dissipative behavior: Indeed, substitution of the pole value in the total self-energy function yields exponential decay, including, as is well known, higher-order corrections to the Fermi Golden Rule. On the other hand, the first step is very important when one is interested in computing the leading order corrections to the exponential behavior. To this purpose one can solve the problem in a restricted Tamm-Duncoff sector of the total Hilbert space (i.e., in an eigenspace of ${\cal N}_\mu$ --- in our case, ${\cal N}_\mu=1$) and exactly evaluate the evolution of the system with its deviations from exponential law. Let us add a final remark. As is well known, a nondispersive propagator yields a Markovian evolution. Let us briefly sketch how this occurs in the present model. From (\ref{eq:NBorn}), antitransforming, \begin{equation} \frac{i}{2\pi}\int_C dE e^{-iEt} \left(E S'(E+\omega_\mu)-1\right)= \frac{i}{2\pi}\int_C dE e^{-iEt} \lambda^2\Sigma (E+\omega_\mu) S'(E+\omega_\mu), \end{equation} we obtain (for $t>0$) \andy{nonlocal} \begin{equation} i\dot {\cal A}(t)=\lambda^2\int_0^t d\tau \sigma(t-\tau){\cal A}(\tau) , \label{eq:nonlocal} \end{equation} where ${\cal A}(t)$ is the survival amplitude (\ref{eq:survampl}) and \begin{equation} \sigma(t)\equiv\frac{1}{2\pi}\int_C dE e^{-iEt}\Sigma(E+\omega_\mu) =\frac{e^{i\omega_\mu t}}{2\pi}\int_C dE e^{-iEt}\Sigma(E). \end{equation} Equation (\ref{eq:nonlocal}) is clearly nonlocal in time and all memory effects are contained in $\sigma(t)$, which is the antitransform of the self-energy function. If such a self-energy function is a complex constant (energy independent), $\Sigma(E)=C$, then $\sigma(t)=C\delta(t)$ and equation (\ref{eq:nonlocal}) becomes \begin{equation} i\dot {\cal A}(t)=\lambda^2 C {\cal A}(t), \end{equation} describing a Markovian behavior, without memory effects \cite{VanKampen}. In particular, the Van Hove limit is equivalent to set $C=\Sigma^{(2)}(\omega_\mu+i0^+)$ and the Weisskopf-Wigner approximation is $C=\Sigma^{(2)}(\omega_\mu+i0^+)+{\rm O}(\lambda^2)$. In conclusion, in the Van Hove limit, the evolution of our system, which was nonlocal in time due to the dispersive character of the propagator (the self-energy function depended on $E$) becomes local and Markovian (only the value of the self-energy function in $\omega_\mu$ determines the evolution). \setcounter{equation}{0} \section{General framework} \label{sec-general} \andy{general} We can now further generalize our analysis: consider the Hamiltonian \begin{equation} H=H_0+\lambda V \end{equation} and suppose that the initial state $|a\rangle$ has the following properties \andy{aprop} \begin{eqnarray} & & H_0|a\rangle=E_a|a\rangle,\qquad \langle a|V|a\rangle=0, \nonumber\\ & & \langle a|a\rangle=1. \label{eq:aprop} \end{eqnarray} The survival amplitude of state $|a\rangle$ reads \andy{survEgen} \begin{eqnarray} {\cal A}(t)&\equiv&\langle a|e^{i H_0 t} U(t)|a\rangle= \frac{i}{2\pi}\int_{\rm C}dE e^{-iEt} S'(E+E_a)\nonumber\\ &=& \frac{i}{2\pi}\int_{\rm C}dE\frac{e^{-iEt}} {E-\lambda^2\Sigma(E+E_a)}, \label{eq:survEgen} \end{eqnarray} where $S'(E)\equiv\langle a|(E-H)^{-1}|a\rangle$ and $\Sigma(E)$ is the 1-particle irreducible self-energy function, that can be expressed by a perturbation expansion \andy{sigmagen} \begin{equation} \lambda^2 \Sigma(E) = \lambda^2 \Sigma^{(2)}(E) + \lambda^4 \Sigma^{(4)}(E) +\cdots . \label{eq:sigmagen} \end{equation} The second order contribution has the general form \andy{sigma2gen} \begin{eqnarray} \Sigma^{(2)}(E)&\equiv&\langle a|V P_d \frac{1}{E-H_0} P_d V|a\rangle =\sum_{n\neq a} \left| \langle a|V|n\rangle \right|^2\frac{1}{E-E_n} \nonumber\\ &=& \int_0^\infty \frac{dE'}{2\pi} \frac{\Gamma(E')}{E-E'}, \label{eq:sigma2gen} \end{eqnarray} where $P_d=1-|a\rangle\langle a|$ is the projector over the decayed states, $\{|n\rangle\}$ is a complete set of eingenstates of $H_0$ ($H_0|n\rangle=E_n|n\rangle$ and we set $E_0=0$) and \andy{genGamma} \begin{equation} \Gamma(E)\equiv2\pi\sum_{n\neq a}\left| \langle a|V|n\rangle \right|^2\delta(E-E_n). \label{eq:genGamma} \end{equation} Notice that $\Gamma(E)\geq0$ for $E>0$ and is zero otherwise. In the Van Hove limit we get \andy{gensurvEsclim} \begin{equation} \widetilde {\cal A}(\tilde{t})\equiv \lim_{\lambda\to0}{\cal A}\left(\frac{\tilde{t}}{\lambda^2}\right) =\frac{i}{2\pi}\int_{\rm C} d\tilde E e^{-i\tilde{E}\tilde t}\widetilde S'(\tilde E), \label{eq:gensurvEsclim} \end{equation} where the resulting propagator in the rescaled energy $\tilde E=E/\lambda^2$ reads \andy{genrescS} \begin{equation} \widetilde S'(\tilde E) =\frac{1}{\tilde E-\Sigma^{(2)}(E_a+i0^+)}. \label{eq:genrescS} \end{equation} To obtain this result we used \andy{gensigma2lim} \begin{equation} \Sigma(\lambda^2\tilde E+E_a)\stackrel{\lambda\rightarrow 0} {\longrightarrow}\Sigma^{(2)}(\lambda^2\tilde E+E_a)\Big|_ {\lambda=0}=\Sigma^{(2)}(E_a+i0^+) \label{eq:gensigma2lim} \end{equation} (Weisskopf-Wigner approximation and Fermi Golden Rule). Just above the positive real axis we can write \andy{sigmaEa} \begin{equation} \label{eq:sigmaEa} \Sigma^{(2)}(E+i0^+)=\Delta(E)-\frac{i}{2}\Gamma(E), \end{equation} where \andy{genpolecoord1} \begin{equation} \Delta(E) = {\cal P}\int_0^\infty \frac{dE'}{2\pi}\frac{\Gamma(E')}{E-E'}. \label{eq:genpolecoord1} \\ \end{equation} Let $\Gamma(E)$ be sommable in $(0,+\infty)$. Then \andy{Gammalim} \begin{equation} \Gamma(E)\propto E^{\eta-1}\quad\mbox{for}\quad E\to0, \label{eq:Gammalim} \end{equation} for some $\eta>0$, and one gets the following asymptotic behavior at short and long times: \andy{genshortt,genlargetimes1} \begin{eqnarray} \!\! P(t) \! &\sim& \! 1 - \frac{t^2}{\tau_{\rm Z}^2} \qquad (t\ll \tau_{\rm Z}), \label{eq:genshortt} \\ \!\! P(t) \! &\sim& \! |Z|^2 e^{-t/\tau_{\rm E}}+ \lambda^4\frac{|C|^2}{(E_a t)^{2\eta}} + 2 \lambda^2\frac{|CZ|}{(E_a t)^\eta} e^{-t/2\tau_{\rm E}} \cos\left[(E_a+\Delta E) t -\zeta \right] \label{eq:genlargetimes1} \nonumber\\ & & \qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad(t\gg \tau_{\rm Z}), \end{eqnarray} where \andy{gentauZ,genfgr,genshift,genzeta,genZz,genCchi} \begin{eqnarray} \tau_{\rm Z} &=& \frac{1}{\lambda} \left[\int_0^\infty \frac{dE}{2\pi}\Gamma(E)\right]^{-1/2}, \label{eq:gentauZ}\\ \tau_{\rm E} &=& \frac{1}{\lambda^2 \Gamma(E_a)}, \label{eq:genfgr}\\ \Delta E &=&\lambda^2 \Delta(E_a), \label{eq:genshift} \\ \zeta&=& {\rm Arg} Z - {\rm Arg} C, \label{eq:genzeta}\\ Z &=& 1 + {\rm O}(\lambda^2), \label{eq:genZz} \\ C &=& 1 + {\rm O}(\lambda^2). \label{eq:genCchi} \end{eqnarray} The transition to a power law occurs when the first two terms in the r.h.s. of (\ref{eq:genlargetimes1}) are comparable, namely for $t=\tau_{\rm pow}$, where $\tau_{\rm pow}$ is solution of the equation \andy{tpoweq} \begin{equation} \frac{\tau_{\rm pow}}{\tau_{\rm E}}=4(\eta+1)\log\frac{1}{\lambda} +2\eta\log\frac{E_a}{\Gamma(E_a)}+\log\left|\frac{Z}{C}\right| +\eta\log\frac{\tau_{\rm pow}}{\tau_{\rm E}}, \label{eq:tpoweq} \end{equation} i.e., for $\lambda\to 0$ \andy{gentaupow} \begin{equation} \tau_{\rm pow} = 4 \tau_{\rm E} (\eta+1) \log\lambda^{-1} +{\rm O}\left(\log\log \lambda^{-1}\right). \label{eq:gentaupow} \end{equation} Let us now look at the temporal behavior for a small but {\em finite} value of $\lambda$, using Van Hove's technique. In the rescaled time, $\tilde t=\lambda^2 t$, the Zeno region vanishes \andy{resctauZ} \begin{equation} \tilde\tau_{\rm Z}\equiv\lambda^2\tau_{\rm Z}=\lambda\left[\int_0^\infty \frac{dE}{2\pi}\Gamma(E)\right]^{-1/2}={\rm O}(\lambda) \label{eq:resctauZ} \end{equation} and Eq.\ (\ref{eq:genlargetimes1}) becomes valid at {\em shorter and shorter} (rescaled) times and reads \andy{genlargetimes2} \begin{eqnarray} & & P(\tilde t) \sim |Z|^2 e^{-\tilde t/\tilde\tau_{\rm E}} + \lambda^{4(\eta+1)}\frac{|C|^2}{(E_a \tilde t)^{2\eta}} \nonumber\\ & &\quad\quad\quad\quad\quad + 2 \lambda^{2(\eta+1)}\frac{|CZ|}{(E_a t)^\eta} e^{-\tilde t/2\tilde\tau_{\rm E}} \cos\left(\frac{E_a+\Delta E}{\lambda^2} \tilde t -\zeta \right), \label{eq:genlargetimes2} \end{eqnarray} where \andy{resctauE, resctaupow} \begin{eqnarray} \tilde\tau_{\rm E} &\equiv& \lambda^2\tau_{\rm E}= \frac{1}{\Gamma(E_a)}={\rm O}(1), \label{eq:resctauE}\\ \tilde\tau_{\rm pow}&\equiv& \lambda^2\tau_{\rm pow}\simeq 4 \tilde\tau_{\rm E} (\eta+1) \log\frac{1}{\lambda} ={\rm O}\left( \log\frac{1}{\lambda} \right). \label{eq:resctaupow} \end{eqnarray} Figure \ref{fig:summa} displays the main features of the temporal behavior of the survival probability. \begin{figure}[t] \epsfig{file=zeno2.eps, width=13.5cm} \caption{Essential features (not in scale!) of the survival probability as a function of the rescaled time $\tilde t$. The Zeno time is ${\rm O}(\lambda)$, the lifetime ${\rm O}(1)$, during the whole evolution there are oscillations of amplitude ${\rm O}(\lambda^{2\eta+2})$ and the transition to a power law occurs after a time ${\rm O}(\log (1/\lambda))$ [see (\ref{eq:resctauZ})-(\ref{eq:resctaupow})]. From (\ref{eq:genZz}), the normalization factor becomes unity like $1-{\rm O}(\lambda^2)$. The dashed line is the exponential and the dotted line the power law.} \label{fig:summa} \andy{summa} \end{figure} The typical values of the physical constants [see for instance (\ref{eq:lambdachi})] yield very small deviations from the exponential law. For this reason, we displayed in Figure \ref{fig:summa} the survival probability by greatly exaggerating its most salient features. The Van Hove limit performs several actions at once: It makes the initial quadratic (quantum Zeno) region vanish, it ``squeezes" out the oscillations and it ``pushes" the power law to infinity, leaving only a clean exponential law at all times, with the right normalization. All this is not surprising, being implied by the Weisskopf-Wigner approximation. However, the concomitance of these features is so remarkable that one cannot but wonder at the effectiveness of this limiting procedure. In atomic and molecular physical systems the smallness of the coupling constant and other physical parameters makes the experimental observation of deviations from exponential a very difficult task (see for example the simple model investigated in Section 2). The eventuality that alternative physical systems might exhibit experimentally observable non-exponential decays, as well as the possibility of modifying the lifetimes of unstable systems by means of intense laser beams \cite{MPS} are at present under investigation.
\section{Introduction} \par Recently there has been progress in the theory of Anderson localization for two dimensional continuous models of an electron moving in a random potential and a uniform magnetic field (\cite{DMP2}, \cite{DMP3}, \cite{CH}, \cite{W}). In these works it is established that the states at the edges of the Landau bands are exponentially localized and the corresponding energies form a pure point spectrum. However, the nature of the generalized eigenfunctions of the Schr\"odinger operator for energies near the centre of the Landau bands has not been established. A first step towards the resolution of this problem was made in \cite{DMP1} for a Hamiltonian restricted to the first Landau band with a random potential consisting of point impurities with random strength and located on the sites of a square lattice. There it was shown that, for a sufficiently strong magnetic field, all the eigenstates are localized except for a single energy at the centre of the band. This energy is an infinitely degenerate eigenvalue with probability one. \par In the present paper we extend the results of \cite{DMP1} to a similar model where the restriction to the lowest Landau band is removed. The technique used here is different and yields much stronger results. Formally the Hamiltonian of the electron is given by \begin{equation} H = H_0 + \sum_{\bf n} v_{\bf n} \delta({\bf r - n}), \label{a.1} \end{equation} where \begin{equation} H_0 = {1\over 2} \(- i\nabla -{\bf A}({\bf r} )\)^2, \label{a.2} \end{equation} ${\bf A}({\bf r})={1\over 2} ({\bf r}\times {\bf B})$ and $v_{\bf n}$, the strengths of the impurities which are located on the sites ${\bf n}$ of a two-dimensional square lattice, are i.i.d. random variables. It is well known that the definition of Hamiltonians with point scatterer in more than one dimension is delicate and requires a renormalization procedure. This is the subject of Section 2. \par The main results of this paper are the following. Let $E_n = (n + {1\over 2})B$, $n = 0,1,2, \ldots $, be the Landau levels corresponding to the kinetic part, $H_0$, of the Hamiltonian. Given an integer $N$, there exists $B_0 (N)$ such that for $B> B_0 (N)$, the spectrum is completely characterized for energies $E < E_N$. We show that for $n = 0, 1,2, \ldots N - 1$, the Landau levels $E_n$ are infinitely degenerate eigenvalues of $H$ with probability one. All other energies in this part of the spectrum correspond to exponentially localized eigenfunctions with a localization length which is uniformly bounded as a function of the energy. Thus the localization length does not diverge at the centres of the bands when the magnetic field is strong enough, at least for the lower bands. Our analysis breaks down for energies greater than $E_N$ and in fact we expect a different behaviour for high energies. \par There is an extensive literature on the problem of point scatterers with a magnetic field, but it appears that little is known on the rigorous level for the two-dimensional random case considered here. For the periodic case, that is, when all the $v_{\bf n}$'s are identical, we refer the reader to the review \cite{G} and the references therein. The case when the potential is periodic in the $x$-direction and random in the $y$-direction has been discussed recently in \cite{GZAA}. Finally the density of states for models similar to ours with a restriction to the first Landau level has been computed analytically in \cite{BGI} (see also \cite{E} which deals with the existence of Lifshitz tails). The infinite degeneracy of the Landau levels had already been noticed in various ways in the past (\cite{H}, \cite{BGI}, \cite{BF}). For example in \cite{BGI} it appears as a delta function in the density of states of the first level. The result suggests that it is in fact {\sl macroscopic}, in other words, there is a positive density per unit volume. Our results characterize completely the rest of the spectrum and also give information about the localization length. \par Let us say a few words about the method used to arrive at these results. The scatterers in (\ref{a.1}) are similar to rank one perturbations of the kinetic energy so that by using the resolvent identity one can express the Green's function corresponding to $H$ in terms of the Green's function of the kinetic energy and a matrix which contains all the randomness. Thus the problem is reduced completely to the study of this random matrix which has random elements on the diagonal and rapidly decaying non-random off-diagonal elements. It turns out that the method invented by Aizenman and Molchanov \cite{AM} is very well suited to study the decay of eigenvectors of this matrix. These eigenvectors are related by an explicit formula to the eigenfunctions of $H$ in such a way that exponential decay of the former implies exponential decay of the latter. In fact it follows from the structure of the random matrix that, in the strong magnetic field regime, the off-diagonal elements are much smaller than the diagonal elements, and this is true even for energies near the band centres. Therefore our problem is analagous to the high disorder regime in the usual Anderson model and this is the reason why we have access to the whole spectrum. \par It is instructive to discuss the physical implications of our results in the context of the quantum Hall effect. A basic ingredient used to explain the occurrence of plateaux in the Hall conductivity is the localization of electrons due to the random potential. This has been established in a mathematically precise way in \cite{Ku} (see also \cite{Be}), by assuming the existence of localized states. Usually it is difficult to obtain quantitative results on the localization length. The Network Model of Chalker and Coddington \cite{CC} and numerical simulations \cite{H} suggest that it is finite except at the band centres where it diverges like $\vert E - E_n \vert^{ - \nu}$ with $\nu \approx 2\cdot 35$ for the first few $n$'s. In the Network Model one must work with smooth equipotential lines of the random potential so that it is difficult to compare to our situation. The model in this paper has been treated numerically only in a regime where the magnetic length, which is of the order of $B^{-1/2}$, is much greater than the average spacing between impurities. The regime covered by our analysis is such that the magnetic length is smaller than the average spacing between impurities, and we prove that there is no divergence in the localization length at least for the first few bands. One might think that this means that there is no quantum Hall effect in this regime. However this is not the case because the energy at the band centre is an infinitely degenerate eigenvalue. One can compute explicitly the eigenprojector associated to each degenerate eigenvalue and check that the corresponding Chern number is equal to unity \cite{DMP4}. From this result and the equivalence between Hall conductivity and Chern number, when the Fermi level lies in the region of localized states or in a spectral gap, we conclude that the Hall conductivity takes a non-zero quantized value equal to the number of Landau levels below the Fermi energy. This has made mathematically precise in \cite{Ku} (see also \cite{Be} and \cite{AG}). \par The picture which emerges out of the combination of our analytical results with those of simulations is that in the present model one has to distinguish at least two regimes. In the first one, the magnetic length is much greater than the spacing between impurities: the localization length diverges and there is no degenerate eigenvalue at the band centres. In the second the magnetic length is much smaller than the spacing between impurities: the localization length does not diverge and there is a degenerate eigenvalue at the band centres. Whether there exists one or more intermediate regimes or not is an open question. It is instructive to note that in the model studied in \cite{BGI}, it turns out that, at the level of the density of states, one must also distinguish between various regimes, more than two in fact. Finally, we wish to stress that the quantized Hall plateaux exist in both regimes and that an interesting open question is whether the different behaviour of the localization length is reflected in the transition between two successive Hall plateaux. \par The paper is organized as follows. In Section 2 we give the precise definition of the model and the Hamiltonian and also collect useful Green's function identities. Our main theorem (Theorem 2.2) is stated at the end of this section. The infinite degeneracy of the first $N(B)$ Landau levels is proved in Section 3 and the spectrum is characterized as a set. The connection between generalized eigenfunctions of $H$ and eigenvectors of the random matrix is established in Section 4. Finally, the Aizenman-Molchanov method is applied in Section 5 where the proof of our main theorem is completed. The appendices contain more technical material. \section{Definition of the Hamiltonian} In this section we define our Hamiltonian. It is well known that Hamiltonians with $\delta$-function potentials in dimensions greater than one require renormalization. This was first done rigorously in \cite{BF}. The magnetic field case was developed in \cite{G}. We refer the reader also to \cite{AGHKH} though this does not deal explicitly with the case of a magnetic field. \par Let $\omega_n$, $n \in \hbox{\BB Z}[i]\equiv \{n_1+in_2: (n_1,n_2)\in \hbox{\BB Z}^2\}$, the Gaussian integers, be i.i.d. random variables. We shall assume that their distribution is given by an absolutely continuous probability measure $\mu_0$ whose support is an interval $X=[-a,a]$ with $0<a <\infty $. We require that $\mu_0$ is symmetric about the origin and that its density $\rho_0$ is differentiable on $(-a,a)$ and satisfies the following condition \begin{equation} \sup_{\zeta\in (0,a)} \frac{\rho_0'(\zeta)} {\rho_0(\zeta)}< \infty. \end{equation} These conditions on $\mu_0$ can be weakened, but we have chosen the above because they allow us to check the regularity of the distribution of $1/\omega_n $, in the sense of \cite{AM} very simply. We let $\Omega=X^{\hbox{\BB Z}[i]}$ and $\hbox{\BB P}=\prod_{n\in \hbox{\BB Z}[i]}\mu_0$. For $m\in \hbox{\BB Z}[i]$ let $\tau_m$ be the measure preserving automorphism of $\Omega$ defined by \begin{equation}\label{b.1} \(\tau_m \omega\)_n =\omega_{n-m}. \end{equation} The group $\{\tau_m\ :\ m\in\hbox{\BB Z}[i]\}$ is ergodic for the probability measure $\hbox{\BB P}$. \par Let ${\cal H}=L^2(\hbox{\BB C})$ and let $H_0$ be the operator on ${\cal H}$ defined by \begin{equation}\label{b.2} H_0=(1/8\kappa)(-i\nabla-A(z))^2 -1/2 \end{equation} where $A(z)=(-2\kappa {\cal I}z, 2\kappa {\cal R}z)$. Here $\kappa=B/4$ and $H_0$ is the same as the Hamiltonian in (\ref{a.2}) apart from the multiplicative constant $1/8\kappa$ and the shift by $1/2$ which are inserted for convenience so that the Landau levels coincide with the set of non-negative integers, $\hbox{\BB N}_0$. Let ${\cal H}_m$ be the eigenspace corresponding to the $m$th Landau level of the Hamiltonian $H_0$ defined in (\ref{b.2}) and let $P_m$ be the orthogonal projection onto ${\cal H}_m$. The projection $P_m$ is an integral operator with kernel \begin{equation}\label{b.4} P_m \(z,z'\) = L_m(2\kappa | z-z'|^2)P_0 \(z,z'\), \end{equation} where $L_m$ is the Laguerre polynomial of order $m$ and \begin{equation}\label{b.5} P_0 \(z,z'\) = {{2\kappa} \over { \pi}} \exp [- {\kappa} | z - z' |^2 - 2i \kappa z \wedge z'], \end{equation} with $ z \wedge z'={\cal R}z{\cal I}z'-{\cal I}z{\cal R}z'$, ${\cal R}z$ and ${\cal I}z$ being the real and imaginary parts of $z$ respectively. \par For $\lambda \in \hbox{\BB C}\setminus \hbox{\BB N}_0$, let $G^\lambda_0=(H_0-\lambda)^{-1}$, the resolvent of $H_0$ at $\lambda$. $G^\lambda_0$ has kernel (cf \cite{G}) \begin{equation} G^\lambda_0 (z,z') = \Gamma (-\lambda) P_0 (z,z') U (- \lambda, 1, 2 \kappa | z - z' |^2 ), \label{b.0} \end{equation} where \begin{equation} U(a, 1, \rho)= - {1\over {\Gamma (a) }}\[M(a, 1, \rho) \ln \rho + \sum^\infty_{r=0} {{(a)_r} \over {r!}} \rho^r \{\psi(a + r) - 2 \psi (1 + r)\}\] \end{equation} is the logarithmic solution of Kummer's equation (\cite{AS} Chap. 13): \begin{equation} \rho{{d^2U}\over{d\rho^2}}+(1-\rho) {{dU}\over{d\rho}}-a\rho=0 \end{equation} Here $\Gamma$ is the Gamma function, $\psi(a)=\Gamma'(a)/\Gamma(a)$ is the Digamma function, \begin{equation} (a)_r=a(a+1)(a+2)\ldots (a+r-1),\ \ \ \ \ \ \ \ (a)_0=1, \end{equation} and \begin{equation} M(a, 1, \rho)= \sum^\infty_{r=0} {{(a)_r} \over {r!}} \rho^r \end{equation} is Kummer's function. \par Let ${\cal M}=l^2 ([\hbox{\BB Z} [i]) $ and for $\lambda \in \hbox{\BB C}\setminus \hbox{\BB N}_0$, define $U_\lambda:{\cal H}\to {\cal M}$ by \begin{equation} \langle n | U_\lambda \phi\rangle = (G^\lambda_0 \phi)(n). \end{equation} From the bounds in Propositions 6.1 and 6.2 in Appendix A one can see that $U_\lambda$ is a bounded operator. Its adjoint $U_\lambda^*:{\cal M}\to {\cal H}$ is given by \begin{equation} (U_\lambda^* \xi)(z) = \sum_{n\in \hbox{\BB Z}[i]}G^{\bar \lambda}_0(z,n) \langle n | \xi\rangle. \end{equation} For $\lambda \in \hbox{\BB C} \setminus \hbox{\BB N}_0$ let \begin{equation} c^\lambda_n = {{2 \kappa}\over\pi} \(\psi (- \lambda) -{{2 \pi} \over {\omega_n}}\) \end{equation} and define the operators $D^\lambda$, $A^\lambda$ and $M^\lambda$ on ${\cal M}$ as follows. $D^\lambda$ is diagonal and \begin{equation} \langle n | D^\lambda | n \rangle = c^\lambda_n, \end{equation} \begin{equation} \langle n | A^\lambda | n' \rangle = \cases{ 0 & if $ n = n'$ \cr G^\lambda_0 (n,n') & if $ n \not= n'$, \cr} \end{equation} and \begin{equation} M^\lambda = D^\lambda -A^\lambda. \label{b.3} \end{equation} Note that $D^\lambda$ is a closed operator on the domain \begin{equation} D(D^\lambda)=\{\xi\in {\cal M}:\ \ \sum_{n\in\hbox{\BB Z}[i]}| c^\lambda_n |^2 |\langle n| \xi\rangle|^2 <\infty\}, \end{equation} and $A^\lambda$ is bounded, therefore $M^\lambda$ is closed on $ D(M^\lambda)=D(D^\lambda)$. Note also that $(M^\lambda)^*=M^{\bar \lambda}$ and that for $\lambda \in \hbox{\BB R}$, $M^\lambda$ is self-adjoint. For $\lambda \in \hbox{\BB C}\setminus \hbox{\BB N}_0$ such that $0\notin \sigma(M^\lambda)$ let \begin{equation} \Gamma^\lambda = (M^\lambda)^{-1}. \end{equation} To define our Hamiltonian $H$ we use the following lemma. : \par\noindent {\bf Lemma 2.1}\ {\sl For each $\kappa>0$, there exists $\lambda_\kappa\in \hbox{\BB C}\setminus \hbox{\BB R}$ such that $0\notin \sigma(M^{\lambda_\kappa})$ and \begin{equation} | \langle n | \Gamma^{\lambda_\kappa} | n' \rangle | \leq K(\kappa) e^{-\kappa | n - n' |^{1\over 2}}. \end{equation} } \par\noindent {\bf Proof:}\ Let $\lambda=-r(1+i)$, with $r>0$. By Proposition 6.1 in Appendix A, we have for $n, \ n' \in \hbox{\BB Z}[i]$, $n\neq n'$, \begin{equation} | G_0^\lambda (n,n') | \leq C_{r,\kappa} e^{- \kappa | n - n' |^2}, \end{equation} where \begin{equation} C_{r,\kappa}=C\kappa\left \{ {1\over r}+ e^{-(2\kappa r)^{1\over 2} }(1+|\ln (2\kappa)|)\right \}, \end{equation} $C<\infty$ being a constant. Therefore $|| A^\lambda|| \leq C_{r,\kappa}|| S||$ where $S$ is the operator with matrix \begin{equation} \langle n| S| n'\rangle = e^{-\kappa | n - n' |^2}. \end{equation} Let ${\tilde \Gamma}^\lambda=(D^\lambda)^{-1}$, then \begin{equation} || A^\lambda{\tilde \Gamma}^\lambda|| \leq {\pi\over{2\kappa}} {{ C_{r,\kappa}}\over {| {\cal I}\psi(-\lambda)|}} || S|| <1/2, \end{equation} if $r$ is large enough. Note that by (6.3.18) in \cite{AS} \begin{equation} \lim_{r\to\infty}{\cal I}\psi(-\lambda)=\pi/4. \end{equation} Then $ \sum^\infty_{k=1} (A^\lambda {\tilde \Gamma}^\lambda)^k$ converges and consequently $M^\lambda$ is invertible, \begin{equation}\label{b.6} \Gamma^\lambda = {\tilde \Gamma}^\lambda (I + \sum^\infty_{k=1} (A^\lambda {\tilde \Gamma}^\lambda)^k) \end{equation} and $|| I + \sum^\infty_{k=1} (A^\lambda {\tilde \Gamma}^\lambda)^k || \leq 2$. Clearly \begin{equation} \langle n | A^\lambda {\tilde \Gamma}^\lambda | n' \rangle = \cases{0 & for $n = n'$ \cr {1 \over {c_{n'}^\lambda}} \ \ G_0^\lambda (n,n') & if $n \not= n'$ .\cr} \end{equation} Thus \begin{equation} | \langle n | A^\lambda {\tilde \Gamma}^\lambda | n' \rangle | \leq B_{r,\kappa}e^{- \kappa | n - n' |^2} \leq B_{r,\kappa}e^{- \kappa | n - n' |^{1\over 2}} \label{b.11} \end{equation} where $ B_{r,\kappa} = {\pi\over{2\kappa}} {{ C_{r,\kappa} } \over {| {\cal I}\psi(- \lambda)|}} $. Now, there exists a constant $ c_0<\infty $ such that for $\kappa >1$ (see Lemma 3.3 in \cite{DMP1}), \begin{equation} \sum_{n'' \in \hbox{\BB Z}[i] } e^{- \kappa | n - n'' |^{1\over 2}} e^{- \kappa | n'' - n' |^{1\over 2}}\leq c_0 e^{- \kappa | n - n' |^{1\over 2}}. \end{equation} This bound, together with (\ref{b.11}), gives \begin{equation} | \langle n | (A^\lambda {\tilde \Gamma}^\lambda)^k | n'\rangle | \leq c_0^{k-1} B_{r,\kappa}^k e^{- \kappa | n - n' |^{1\over 2}} \end{equation} and thus from (\ref{b.6}) \begin{equation} | \langle n | \Gamma^\lambda | n' \rangle | \leq K e^{-\kappa | n - n' |^{1\over 2}} \end{equation} if $c_0 B_{r,\kappa} < {1\over 2}$. \par\noindent\rightline{$\square$} \par For $\lambda \in \hbox{\BB C}\setminus\hbox{\BB N}_0$, we have the formula (\cite{AS} 6.3.16) \begin{equation} \psi(- \lambda) = - \gamma - \sum^\infty_{m=0} {{(\lambda +1)} \over {(m+1)(m- \lambda)}} \end{equation} where $\gamma$ is Euler's constant. Thus if $\lambda_1, \lambda_2 \in \hbox{\BB C}\setminus\hbox{\BB N}_0 $ \begin{equation} \psi(- \lambda_1) - \psi(- \lambda_2) = (\lambda_2 - \lambda_1) \sum^\infty_{m=0} {1 \over {(m - \lambda_1)(m - \lambda_2)}}. \end{equation} On the other hand we have \begin{equation} G^{\lambda_1}_0 G^{\lambda_2}_0 = \sum^\infty_{m=0} {{P_m} \over {(m - \lambda_1) (m - \lambda_2)}} \end{equation} and thus \begin{equation} (G^{\lambda_1}_0 G^{\lambda_2}_0) (n, n) = \sum^\infty_{m=0} {{P_m (n,n)} \over {(m - \lambda_1)(m - \lambda_2)}} = {{2 \kappa} \over \pi} \sum^\infty_{m=0} {1 \over {(m - \lambda_1)(m - \lambda_2)}}. \label{b.10} \end{equation} Therefore \begin{equation} \langle n | M^{\lambda_1} - M^{\lambda_2} | n \rangle = {{2 \kappa} \over \pi} \{\psi(- \lambda_1) - \psi(-\lambda_2)\} = (\lambda_2 - \lambda_1)(G_0^{\lambda_1} G_0^{\lambda_2})(n,n). \label{b.7} \end{equation} On the other hand, for $n \not= n'$, using the resolvent identity, we get \begin{equation}\label{b.8} \langle n | M^{\lambda_1} - M^{\lambda_2} | n' \rangle = G_0^{\lambda_2} (n,n') - G^{\lambda_1}_0 (n,n') = (\lambda_2 - \lambda_1)(G_0^{\lambda_1} G_0^{\lambda_2})(n,n'). \end{equation} Therefore combining the two identities (\ref{b.7}) and (\ref{b.8}) we obtain \begin{equation}\label{b.9} M^{\lambda_1} - M^{\lambda_2} = (\lambda_2 - \lambda_1) U_{\lambda_2}U_{{\bar\lambda}_1}^*. \end{equation} It is clear from this equation that $U_{\lambda_2}U_{\bar{\lambda}_1}^* = U_{\lambda_1}U_{{\bar\lambda}_2}^*$. \par Note that $H_0$ is essentially self-adjoint on ${\cal S}(\hbox{\BB C})$ (\cite{RS2} Theorem X.34). Define $V_\kappa:{\cal S}(\hbox{\BB C})\to {\cal H}$ by $V_\kappa=U^*_{{{\bar \lambda}_\kappa}}\Gamma^{\lambda_\kappa}T$ where $\langle n | T\psi \rangle=\psi(n)$. Let \begin{equation} D(H)=\{\phi=\psi +V_\kappa\psi\ : \ \psi \in {\cal S}(\hbox{\BB C})\}, \end{equation} and for $\phi \in D(H)$ \begin{equation} H\phi=H_0\psi +{\lambda_\kappa}V_\kappa\psi. \end{equation} This definition implies that $(H-{\lambda_\kappa})\phi=(H_0-{\lambda_\kappa})\psi$ and therefore since $H_0$ is essentially self-adjoint on ${\cal S}(\hbox{\BB C})$, ${\rm Ran} (H-{\lambda_\kappa})$ is dense in ${\cal H}$. Let $\psi'\in {\cal S}(\hbox{\BB C})$ and let $\psi=\psi'+({\bar\lambda}_\kappa-\lambda_\kappa) G_0^{\lambda_\kappa}U^*_{\lambda_\kappa} \Gamma^{{{\bar \lambda}_\kappa}}T\psi'$. Then $\psi \in {\cal S}(\hbox{\BB C})$ and $T\psi=M^{\lambda_\kappa}\Gamma^{{{\bar \lambda}_\kappa}}T\psi'$. Note that $0\notin \sigma(M^{{\bar \lambda}_\kappa})$ and $ | \langle n | \Gamma^{{\bar \lambda}_\kappa} | n' \rangle | =| \langle n' | \Gamma^{\lambda_\kappa} | n \rangle | \leq K(\kappa) e^{-\kappa | n - n' |^{1\over 2}} $. Let $\phi=\psi+V_\kappa\psi$. Then \begin{eqnarray} (H-{{\bar \lambda}_\kappa})\phi &=& (H_0-{{\bar \lambda}_\kappa})\psi +(\lambda_\kappa-{\bar\lambda}_\kappa) V_\kappa\psi\nonumber \\ &=& (H_0-{{\bar \lambda}_\kappa})\psi' -(\lambda_\kappa-{\bar\lambda}_\kappa) G_0^{\lambda_\kappa}(H_0-{{\bar \lambda}_\kappa})U^*_{\lambda_\kappa} \Gamma^{{{\bar \lambda}_\kappa}}T\psi'\nonumber \\ & &\ \ \ \ \ \ +(\lambda_\kappa-{\bar\lambda}_\kappa) V_\kappa\psi' -(\lambda_\kappa-{\bar\lambda}_\kappa)^2 V_\kappa G_0^{\lambda_\kappa}U^*_{\lambda_\kappa} \Gamma^{{{\bar \lambda}_\kappa}}T\psi'\nonumber \\ &=& (H_0-{{\bar \lambda}_\kappa})\psi' -(\lambda_\kappa-{\bar\lambda}_\kappa)U^*_{{{\bar\lambda}_\kappa}} \Gamma^{{{\bar \lambda}_\kappa}}T\psi' +(\lambda_\kappa-{\bar\lambda}_\kappa) V_\kappa\psi' \nonumber \\ & &\ \ \ \ \ \ -(\lambda_\kappa-{\bar\lambda}_\kappa)^2 U^*_{{{\bar \lambda}_\kappa}}\Gamma^{\lambda_\kappa} U_{\lambda_\kappa}U^*_{\lambda_\kappa} \Gamma^{{{\bar \lambda}_\kappa}}T\psi'\nonumber \\ &=& (H_0-{{\bar \lambda}_\kappa})\psi' -(\lambda_\kappa-{\bar\lambda}_\kappa)U^*_{{{\bar\lambda}_\kappa}} \Gamma^{{{\bar \lambda}_\kappa}}T\psi' +(\lambda_\kappa-{\bar\lambda}_\kappa) V_\kappa\psi' \nonumber \\ & &\ \ \ \ \ \ -(\lambda_\kappa-{\bar\lambda}_\kappa) U^*_{{{\bar \lambda}_\kappa}}\Gamma^{\lambda_\kappa} (M^{{{\bar \lambda}_\kappa}}-M^{\lambda_\kappa}) \Gamma^{{{\bar \lambda}_\kappa}}T\psi' \nonumber \\ &= &(H_0-{{\bar\lambda}_\kappa})\psi'. \nonumber \end{eqnarray} Therefore ${\rm Ran} (H-{{\bar \lambda}_\kappa})$ is dense in ${\cal H}$ and $H$ is essentially self-adjoint on $D(H)$. \par For $\lambda \in \hbox{\BB C}\setminus \hbox{\BB N}_0$ such that $0\notin \sigma(M^\lambda)$, define \begin{equation} G^\lambda\equiv G^\lambda_0 + U_{\bar \lambda}^* \Gamma^\lambda U_\lambda. \end{equation} One can check using the resolvent identity and identity (\ref{b.9}) that (\cite{G}, see also \cite{AGHKH}) \begin{equation} G^\lambda(H-\lambda)\phi =\phi, \end{equation} so that \begin{equation} G^\lambda=(H-\lambda)^{-1}. \end{equation} We now state the main theorem of this paper. (a) is proved in Lemma 3.2, (c) in Lemma 3.1 and (b) and (d) in Theorem 5.8. \par\noindent {\bf Theorem 2.2}\ {\sl \newline (a)\ The spectrum of $H$ contains bands around the Landau levels $\hbox{\BB N}_0$ and an interval extending from $-\infty$ to a finite negative point. \newline For each $N\in\hbox{\BB N}$ there exists $\kappa_0 > 0$ such that for $\kappa >\kappa_0$, with probability one, \newline (b)\ $\sigma_{{\rm cont}}(H)\cap (-\infty, N)=\emptyset $, \newline (c)\ if $m \in \hbox{\BB N}_0\cap (-\infty,N)$, then $m$ is an eigenvalue of $H$ with infinite multiplicity \newline (d)\ if $\lambda \in \sigma(H)\cap (-\infty,N)\setminus \hbox{\BB N}_0$, is an eigenvalue of $H$ and the corresponding eigenfunction is $\phi_\lambda $, then for any compact subset $B$ of $\hbox{\BB C}$, $\int_B|\phi_\lambda(z-z')|^2dz'$ decays exponentially in $z$ with exponential length less than or equal to $ 2/\kappa$.} \section{The Spectrum} In this section we study the spectrum of the Hamiltonian. We first show that the Landau levels are still infinitely degenerate eigenvalues. We then prove that the spectrum contains bands around the Landau levels and an infinite interval in the negative half-line. \par Let $\{U_z: \ z \in \hbox{\BB C}\}$ be the family of unitary operators on ${\cal H}$ corresponding to the magnetic translations: \begin{equation} \(U_z f\)\(z'\) = e^{2i\kappa z \wedge z'} f \(z+z'\). \label{c.1} \end{equation} These satisfy $U_{z_1}U_{z_2} = e^{2i\kappa z_2 \wedge z_1}U_{z_1+z_2}$. For $n\in \hbox{\BB Z}[i]$ \begin{equation} U_n G^\lambda\(\omega\) U_n^{-1} = G^\lambda\(\tau_n\omega\). \label{c.2} \end{equation} The ergodicity of $\{\tau_m\ :\ m\in\hbox{\BB Z}[i]\}$ and equation (\ref{c.2}) together imply that the spectrum of $H(\omega)$ and its components are non random (see for example \cite{CL}, Theorem V.2.4). We shall first prove that almost surely the lower Landau levels are infinitely degenerate eigenvalues for large $\kappa$. This lemma is a generalization of similar results in \cite{DMP1} and \cite{ARB}. The main idea of the proof is to construct states in ${\cal H}_m$ which vanish at all the impurity sites, so that they are also eigenfunctions of $H$. These states involve the entire function in (\ref{c.3}) which vanishes at all the points of $\hbox{\BB Z}[i]$ and consequently grows like $e^{A\vert z \vert^2}$ for large $|z|$. The condition that the states are square integrable then requires that the magnetic field be sufficiently large in order to compensate this growth by the factor $e^{-\kappa \vert z \vert^2}$. \vskip .25 cm \noindent {\bf Lemma 3.1} \ {\sl For each $N\in\hbox{\BB N}$, there exists $\kappa_0(N) > 0$, such that for $\kappa >\kappa_0$, with probability one, each Landau level $m$, with $m\leq N$, is infinitely degenerate.} \par \noindent {\bf Proof:} \ The elements of the space ${\cal H}_0$ are of the form \begin{equation} \phi(z)=\psi(z)e^{-\kappa \vert z \vert^2}, \end{equation} where $\psi$ is an entire function and, of course, $\phi\in L^2(\hbox{\BB C})$. Let \begin{equation} \psi_0 (z) = z\!\!\!\!\!\!\prod_{n\in\hbox{\BB Z}[i]\setminus\{0\}}\!\!\!\!\! (1 - {z \over n})e^{{z\over n} +{{z^2}\over {2n^2}}}. \label{c.3} \end{equation} Then $\psi_0$ is an entire function with zeros at all the points of $\hbox{\BB Z}[i]$. It follows from the theory of entire functions (see \cite{B} 2.10.1) that there exists $A>0$ such that $\vert \psi_0(z)\vert \leq e^{A\vert z \vert^2}$. For $k \in \hbox{\BB N}_0$, let \begin{equation} \phi_{0,k} (z) = z^k\psi_0 (z) e^{-\kappa \vert z \vert^2}, \end{equation} then, if $\kappa >A$, $\phi_{0,k} \in {\cal H}_0$ and since $V_\kappa\phi_{0,k}=0$, $H\phi_{0,k}=0$. Also if for $M\in \hbox{\BB N}_0$, $\sum^M_{k=0} b_k \phi_{0,k} = 0$, then $\sum^M_{k=0} b_k z^k = 0$ for $ z \notin \hbox{\BB Z}[i]$. Therefore $\sum^M_{k=0} b_k z^k \equiv 0$ and thus the $ b_k$'s are zero implying that the $\phi_{0,k}$'s are linearly independent. So the $\phi_{0,k}$'s form an infinite linearly independent set of eigenfunctions of $H$ with eigenvalue $0$. For the higher levels we modify this argument with the use of the creation and anihilation operators for the Hamiltonian $H_0$, $a^*$ and $a$, defined by $\displaystyle{ a^*=(1/\sqrt {2\kappa})\(-{{\partial\phantom z}\over{\partial z}}+\kappa \bar z\)}$ and $\displaystyle {a=(1/\sqrt{2\kappa})\({{\partial\phantom z}\over{\partial \bar z}}+\kappa z\)}$. \hfill\break These operators satisfy the commutation relation $[a,a^*]=1$. Also if $\phi \in {\cal H}_m$ then $a^*\phi \in {\cal H}_{m+1}$ and $a\phi \in {\cal H}_{m-1}$ except when $m=0$, in which case $a\phi=0$. For $m\leq N$ and $k \in \hbox{\BB N}_0$, let \begin{equation} {\tilde\phi}_{m,k} (z) = z^k\(\psi_0 (z)\)^{m+1} e^{-\kappa \vert z \vert^2}, \end{equation} then, if $\kappa >A(N+1)$, ${\tilde \phi}_{m,k} \in {\cal H}_0$. Now let $\phi_{m,k}=(a^*)^m{\tilde \phi}_{m,k}$. Then $\phi_{m,k} \in {\cal H}_m$ and $\phi_{m,k} (n)= 0$ for all $n\in\hbox{\BB Z}[i] $ since ${\tilde \phi}_{m,k}$ has a zero of order greater then $m$ at each point of $\hbox{\BB Z}[i]$. Therefore since $V_\kappa\phi_{m,k}=0$, $H\phi_{m,k}=m\phi_{m,k}$. Moreover since $[a,a^*]=1$ and $a{\tilde \phi}_{m,k}=0$, $a^m\phi_{m,k}=m!{\tilde \phi}_{m,k}$. So, if for $M\in \hbox{\BB N}_0$, $\sum^M_{k=0} b_k \phi_{m,k} = 0$, then \begin{equation} \sum^M_{k=0} b_k {\tilde \phi}_{m, k} = (m!)^{-1}a^m\left(\sum^M_{k=0} b_k \phi_{m,k}\right) = 0. \end{equation} This means that $\sum^M_{k=0} b_k z^k = 0$ for $ z \notin \hbox{\BB Z}[i]$ and as for $m=0$ it follows that the $\phi_{m,k}$'s form an infinite linearly independent set of eigenfunctions of $H$ with eigenvalue $m$. \par\noindent\rightline{$\square$} \par In the case of one impurity of strength $\omega$ at the origin, the Green's function is given by \begin{equation} G^\lambda= G^\lambda_0 +{1\over {c^\lambda}} G^\lambda_0(\cdot,0) G^\lambda_0(0,\cdot), \end{equation} where \begin{equation} c^\lambda={{2\kappa}\over \pi}\(\psi(-\lambda)-{{2\pi}\over \omega}\). \end{equation} It is clear that in this case the spectrum consists of Landau levels and the values of $\lambda$ for which $c^\lambda=0$. For small $\omega$ the latter correspond to points close to the Landau levels and in the case of $\omega>0$, there is another point which is negative and of the order of $\exp(2\pi/|\omega|)$. \begin{figure}[hbt] \begin{center}\includegraphics[width=10cm]{digamma.eps} \end{center} \caption{$\lambda \mapsto \psi(-\lambda)$} \label{ } \end{figure} In the next lemma we shall show that in our case these points are also in the spectrum in the sense that the spectrum of our Hamiltonian contains bands around the Landau levels and an interval extending from $-\infty$ to a finite negative point. \par Let $Y=\{2\pi/ x \ :\ x\in X\setminus \{ 0\}\}$. \vskip .25 cm \noindent {\bf Lemma 3.2} \ { \sl With probability one} \begin{equation} -\psi^{-1}\(Y\) \subset \sigma (H (\omega)). \end{equation} \par \noindent {\bf Proof:} \ It is sufficient to prove that for each $\lambda \in -\psi^{-1}\(Y\)$ and for all $\epsilon> 0$, there exists $\Omega'$ with $\hbox{\BB P} (\Omega') > 0$ and $\psi \in {\cal H}$ with $\Vert \psi \Vert = 1$ such that for all $\omega \in \Omega'$, $\Vert \(G^{\lambda_\kappa}(\omega) - \(\lambda-{\lambda_\kappa}\)^{-1}\) \psi \Vert < \epsilon$. Let $\langle v \vert n \rangle =\delta_{n 0}$ and let $\psi=C U^*_{\bar \lambda}v$, where $C^{-2}= (2\kappa/\pi)\sum_{m=0}^\infty (m-\lambda)^{-2}$. Note that $\psi(z)=C G^\lambda_0(z,0)$ and $\Vert \psi \Vert=1$ by (\ref{b.10}). Then \begin{eqnarray} \bigl (G^{\lambda_\kappa} - \!\!\!\!\!\!\!&&\!\!\!\!\!\!\! \(\lambda-{\lambda_\kappa}\)^{-1}\bigr ) \psi \nonumber \\ &=& \(G^{\lambda_\kappa}_0 + U^*_{\bar{\lambda}_\kappa}\Gamma^{\lambda_\kappa}U_{\lambda_\kappa} + \({\lambda_\kappa}-\lambda\)^{-1}\) \psi \nonumber \\ &=& \({\lambda_\kappa}-\lambda\)^{-1} C\(\({\lambda_\kappa}-\lambda\) G^{\lambda_\kappa}_0 U^*_{\bar \lambda} - U^*_{\bar{\lambda}_\kappa}\Gamma^{\lambda_\kappa} (M^{\lambda_\kappa}-M^\lambda) +U^*_{\bar \lambda}\) v \nonumber \\ &=& \({\lambda_\kappa}-\lambda\)^{-1} C\(U^*_{\bar{\lambda}_\kappa}-U^*_{\bar \lambda} - U^*_{\bar{\lambda}_\kappa} + U^*_{\bar{\lambda}_\kappa}\Gamma^{\lambda_\kappa} M^\lambda+U^*_{\bar \lambda}\) v \nonumber \\ &=&\({\lambda_\kappa}-\lambda\)^{-1} CU^*_{\bar{\lambda}_\kappa}\Gamma^{\lambda_\kappa}M^\lambda v.\nonumber \end{eqnarray} By using (\ref{b.9}) we get \begin{eqnarray} \Vert \(G^{\lambda_\kappa}(\omega) - \(\lambda-{\lambda_\kappa}\)^{-1}\) \psi \Vert^2 &=& C^2\vert \lambda-{\lambda_\kappa}\vert^{-2} \({\cal I}\lambda_\kappa\)^{-1} {\cal I}\langle M^\lambda v, \Gamma^{\lambda_\kappa}M^\lambda v\rangle\nonumber \\ &\leq& 2C^2 \vert \lambda-{\lambda_\kappa}\vert^{-2} \vert {\cal I}\lambda_\kappa\vert^{-1} \Vert M^\lambda v\Vert\ \Vert {\tilde\Gamma}^{{\bar\lambda}_\kappa}M^\lambda v\Vert \nonumber \end{eqnarray} by (\ref{b.6}). Choose $R$ such that $\sum_{\vert n \vert > R} \vert G^{\lambda}_0(n,0)\vert^2 < \delta$ and let \begin{equation} \Omega' = \{\omega: \vert c^\lambda_0\vert < \delta, \ \ \min_{\vert n\vert \leq R, n \neq 0} \vert c^{\lambda_\kappa}_n \vert > 1/\delta \}. \end{equation} Since $\psi(-\lambda)\in Y$ and $0$ is in the support of $\mu$, $\hbox{\BB P} (\Omega') > 0$. We have \begin{equation} \langle n\vert M^\lambda v\rangle = \cases { c^\lambda_0, & if $n=0$ \cr -G^\lambda_0(n,0), & if $n\neq 0$.\cr} \end{equation} Therefore \begin{equation} \Vert M^\lambda v\Vert^2 \leq \delta^2 + \sum_{n \neq 0} \vert G^{\lambda}_0(n,0)\vert^2 \end{equation} and \begin{eqnarray} \Vert {\tilde\Gamma}^{{\bar\lambda}_\kappa} M^\lambda v\Vert^2 &=& \vert c_0^\lambda\vert^2\vert c_0^{\lambda_\kappa}\vert^{-2} + \sum_{n \neq 0} \vert c_n^{\lambda_\kappa}\vert^{-2} \vert G^{\lambda}_0(n,0)\vert^2 \nonumber \\ &\leq& \delta^2(\pi/2\kappa)^2 \vert {\cal I}\psi(-\lambda_\kappa)\vert^{-2} + \delta^2 \sum_{\vert n \vert \leq R}\vert G^{\lambda}_0(n,0)\vert^2 \nonumber \\ &&\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ + (\pi/2\kappa)^2 \vert {\cal I}\psi(-\lambda_\kappa)\vert^{-2} \sum_{\vert n \vert >R}\vert G^{\lambda}_0(n,0)\vert^2 \nonumber \\ &\leq &\delta^2(\pi/2\kappa)^2 \vert {\cal I}\psi(-\lambda_\kappa)\vert^{-2} + \delta^2 \sum_{n \neq 0}\vert G^{\lambda}_0(n,0)\vert^2 \nonumber \\ &&\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ + \delta(\pi/2\kappa)^2 \vert {\cal I}\psi(-\lambda_\kappa)\vert^{-2}. \nonumber \end{eqnarray} Thus $\Vert \(G^{\lambda_\kappa}(\omega) - \(\lambda-{\lambda_\kappa}\)^{-1}\) \psi \Vert<\epsilon $ if $\delta$ is small enough. \par\noindent\rightline{$\square$} \par In the next section we relate the generalized eigenvectors of $H$ with those of $M^\lambda$. \section{ Generalized eigenfunctions of $H$} In this section we show that a generalized eigenfunction of $H$ with eigenvalue $\lambda$, say, which is not a Landau level, is related in a simple way to an eigenvector $v$ of $M^{\lambda}$ with eigenvalue zero. Furthermore if $v$ decays then so does the corresponding eigenfunction. Since this reduces the problem to a lattice problem, it makes it possible for us to use the Aizenman-Molchanov method. \par\noindent {\bf Proposition 4.1}\ \ {\sl If $\phi$ is a generalized eigenfunction of $H$ with eigenvalue $\lambda\notin \hbox{\BB N}_0$, then $v=\Gamma^{\lambda_\kappa} U_{\lambda_\kappa}\phi$ is a generalized eigenvector of $M^{\lambda}$ with eigenvalue zero and $\phi=(\lambda-{\lambda_\kappa})U^*_{\lambda}v$. Moreover if $v$ decays exponentially, then for any compact subset $B$ of $\hbox{\BB C}$, $\int_B\vert \phi(z-z')\vert^2 dz'$ decays exponentially in $z$.} \par\noindent {\bf Proof:}\ Suppose $\phi$ is a generalized eigenvector of $H$ with eigenvalue $\lambda$. Then \begin{equation} G^{\lambda_\kappa} \phi = (\lambda - {\lambda_\kappa})^{-1} \phi \label{d.1} \end{equation} or \begin{equation} G_0^{\lambda_\kappa} \phi + U_{\bar {\lambda_\kappa}}^* \Gamma^{\lambda_\kappa} U_{\lambda_\kappa}\phi = (\lambda - {\lambda_\kappa})^{-1} \phi. \label{d.2} \end{equation} Thus \begin{equation} U_{\lambda} G_0^{\lambda_\kappa} \phi + U_{\lambda} U_{\bar {\lambda_\kappa}}^* \Gamma^{\lambda_\kappa} U_{\lambda_\kappa}\phi = (\lambda - {\lambda_\kappa})^{-1} U_{\lambda}\phi. \label{d.3} \end{equation} Using $U_{\lambda} G_0^{\lambda_\kappa} = (\lambda - {\lambda_\kappa})^{-1} (U_{\lambda} - U_{\lambda_\kappa})$, we get \begin{equation} U_{\lambda} U_{\bar{\lambda_\kappa}}^* \Gamma^{\lambda_\kappa} U_{\lambda_\kappa}\phi = (\lambda - {\lambda_\kappa})^{-1} U_{\lambda_\kappa}\phi \label{d.4} \end{equation} which by (\ref{b.9}) can be written in the form \begin{equation} M^{\lambda} \Gamma^{\lambda_\kappa} U_{\lambda_\kappa}\phi = 0. \end{equation} Therefore if $v=\Gamma^{\lambda_\kappa} U_{\lambda_\kappa}\phi$, \begin{equation} M^{\lambda} v = 0. \end{equation} From (\ref{d.2}) we get \begin{equation} (\lambda - {\lambda_\kappa})G_0^{\lambda_\kappa} \phi + (\lambda - {\lambda_\kappa})U_{\bar {\lambda_\kappa}}^* v = \phi. \label{d.5} \end{equation} Thus \begin{equation} (\lambda - {\lambda_\kappa})G_0^{\lambda}G_0^{\lambda_\kappa} \phi + (\lambda - {\lambda_\kappa})G_0^{\lambda}U_{\bar {\lambda_\kappa}}^* v = G_0^{\lambda}\phi. \label{d.6} \end{equation} By using the resolvent identity we can write this as \begin{equation} U_{\lambda}^* v =G_0^{\lambda_\kappa} \phi + U_{\bar {\lambda_\kappa}}^* v \label{d.7} \end{equation} and therefore $\phi=(\lambda-{\lambda_\kappa})U^*_{\lambda}v $ by (\ref{d.5}). From Propositions 6.1 and 6.2 in Appendix A we get for $\lambda\notin \hbox{\BB N}_0$ \begin{equation} \vert G^{\lambda}_0 (z,z') \vert < C e^{- {\kappa \over 2} \vert z - z' \vert^2} (1 + {\bf 1}_{B(0, 1/\sqrt{2 \kappa})} (\vert z - z' \vert) \vert \ln \vert z - z' \vert \vert) \end{equation} where $C$ depends on $\lambda$ and $\kappa$. From the equation \begin{equation} \phi (z) = (\lambda - {\lambda_\kappa}) \sum_{n \in \hbox{\BB Z} [i]} G_0^{\lambda} (z,n) \langle n \vert v\rangle, \end{equation} we get, assuming $\langle n \vert v\rangle \leq C' e^{- \alpha \vert n \vert}$, that \begin{eqnarray} \vert \phi (z) \vert & \leq &\vert \lambda - {\lambda_\kappa} \vert C' \sum_{n \in \hbox{\BB Z} [i]} \vert G_0^{\lambda} (z, n) \vert e^{- \alpha \vert n \vert} \nonumber \\ & \leq & C'' \sum_{n \in \hbox{\BB Z} [i]} e^{- {\kappa \over 2} \vert z - n \vert^2} e^{- \alpha \vert n \vert} \nonumber \\ & + & C'' \sum_{n \in \hbox{\BB Z} [i]} e^{- {\kappa \over 2} \vert z - n \vert^2} e^{- \alpha \vert n \vert} {\bf 1}_{B(0, 1/\sqrt{2 \kappa})} (\vert z - n \vert) \vert \ln \vert z - n \vert \vert \nonumber \\ & = & S_1 + S_2. \nonumber \end{eqnarray} Now \begin{eqnarray} S_1 & \leq & C'' \sum_{\vert n - z \vert \geq 1} e^{- {\kappa \over 2} \vert z - n \vert} e^{- \alpha \vert n \vert} + C'' e^\alpha e^{- \alpha \vert z \vert} \nonumber \\ & \leq & C'' e^{- \beta \vert z \vert} \sum_{n \in \hbox{\BB Z} [i]} e^{- \beta \vert n \vert} + C'' e^\alpha e^{- \alpha \vert z \vert} \nonumber \end{eqnarray} where $\beta = {1 \over 4}\min(\kappa, 2 \alpha)$. Thus, $S_1 \leq C''' e^{- \beta \vert z \vert}$. Similarly \begin{equation} S_2 \leq C''' e^{- \beta \vert z \vert} \sum_{n \in \hbox{\BB Z} [i]} e^{- \beta \vert n \vert}{\bf 1}_{B(0,{1 \over {\sqrt{2 \kappa}}})} (\vert z - n \vert)\ln \vert z - n \vert \vert. \end{equation} Therefore \begin{equation} \vert \phi (z) \vert^2 \leq C e^{- 2 \beta \vert z \vert} (1 + 3 \sum_{n \in \hbox{\BB Z} [i]} e^{- \beta \vert n \vert} {\bf 1}_{B(0, 1/\sqrt{2 \kappa})} (\vert z - n \vert) \vert \ln \vert z - n \vert \vert^2). \end{equation} Let $B \subset \hbox{\BB C}$ be compact and let $R = \sup \{\vert z \vert: z \in B\}$. Then for $z' \in B$ \begin{equation} \vert \phi (z - z') \vert^2 \leq C e^{2 \beta R} e^{- 2 \beta \vert z \vert} (1 + 3 \sum_{n \in \hbox{\BB Z} [i]} e^{- \beta \vert n \vert} {\bf 1}_{B (0, {1 \over {\sqrt{2 \kappa}}})} (\vert z - z' - n \vert) \vert \ln \vert z - z' - n \vert \vert^2). \end{equation} Therefore \begin{equation} \int_B\vert \phi (z - z') \vert^2 dz' \leq C e^{2 \beta R} e^{- 2 \beta \vert z \vert} (\vert B \vert + 3 \sum_{n \in \hbox{\BB Z} [i]} e^{- \beta \vert n \vert} \int_{\vert z' \vert < {1\over {\sqrt{2 \kappa}}}} \vert \ln \vert z' \vert \vert^2). \end{equation} \par\noindent\rightline{$\square$} \par We do not dwell on the existence of the generalized eigenfunctions. It suffices to say that the arguments of Theorem II.4.5 in \cite{CL} can be used with $e^{-tH}$ replaced by $G^{\lambda_\kappa}$ since from the bound in Lemma 2.1 and the bounds in Appendix A for $|G^\lambda_0(z,z')|$ it follows that \begin{equation} \sup_z \int_{\hbox{\BB C}}|G^{\lambda_\kappa}_0(z,z')|^2 dz'<\infty. \end{equation} The same bounds guarantee also that $v$ is a generalized eigenvector of $M^\lambda$. In the next section we apply the Aizenman-Molchanov method to the lattice operator $M^\lambda$. \section{An Application of the Aizenman-Molchanov Method} In this section we apply the Aizenman-Molchanov method to $M^\lambda$, where $\lambda$ is not a Landau level. The main ingredient in this method is the Decoupling Principle for $\tau$-regular measures. We start by stating this principle, not in its full generality but in the form in which it will be used here. \par\noindent {\bf Definition} {\sl A measure $\mu$ on $\hbox{\BB R}$ is said to be $\tau$-regular, with $\tau\in (0,1]$, if there exists $\nu >0$ and $C<\infty$ such that \begin{equation} \mu([x - \delta, \ x + \delta]) \leq C \delta^\tau \mu([x - \nu, x+\nu]) \end{equation} for all $x \in \hbox{\BB R}$ and $0 < \delta < 1$.} \par\noindent {\bf Lemma 5.1 (A Decoupling Principle)} {\sl Let $\mu$ be a $\tau$-regular measure and let $\int |u |^\epsilon \mu (du) < \infty$ for some $\epsilon >0$. Then for all $0<s < \min(\tau, \epsilon)$ there exists $\xi_s $, a positive, increasing function on $\hbox{\BB R}_+$ with $\xi_s(0)>0 $ satisfying \begin{equation} \lim_{x\to\infty} {{\xi_s (x)} \over x} = 1, \end{equation} such that for all $\eta$, $a$ and $b\in\hbox{\BB C}$, \begin{equation} \int |u - \eta |^s |au+b|^{-s} \mu (du) \geq (\xi_s ( |\eta |))^s \int |au+b |^{-s} \mu (du). \label{e.1} \end{equation} } \noindent Let $\mu (A)=\mu_0(\{\omega\ :\ 1/\omega \in A\})$. In Appendix B we shall show that $\mu$ is 1-regular and $\int |u |^\epsilon \mu (du) < \infty$ for all $\epsilon <1$. Thus the inequality ({\ref{e.1}) is valid for all $s\in (0, 1)$. As in \cite{AM} we use this lemma to obtain an exponential bound on $\langle n | \Gamma^\lambda (z) |0\rangle $ where $\Gamma^\lambda (z)= (M^\lambda - z)^{-1}$. This bound then allows us to apply the results of \cite{SW} to deduce that the spectrum of $M^\lambda$ in a neighbourhood of the origin consists of eigenvalues and that the corresponding eigenvectors decay exponentially. We then combine this result with Proposition 4.1 to translate it into a statement about the properties of the spectrum of $H$. \par It is convenient here to introduce a notation for the intervals between the Landau levels. We let $I_0=(-\infty, 0)$ and $I_N =(N-1, N)$ for $N\in \hbox{\BB N}$. \par\noindent {\bf Lemma 5.2}\ {\sl For all $N\in \hbox{\BB N}_0$, for all $s\in ({1\over 2}, 1)$ and for all $\gamma < s$ there exists $\kappa_0 (N,s) < \infty$ such that for all $\kappa > \kappa_0 (N,s)$, for all $\lambda \in (-\infty, N)\setminus \hbox{\BB N}_0$ and for all $z\in \hbox{\BB C}$ with ${\cal I}z\neq 0$ and $|{\cal R}z |\leq 1$, \begin{equation} \hbox{\BB E} \{\sum_{n \in \hbox{\BB Z} [i]} |\langle n |\Gamma^\lambda (z) |0 \rangle |^s e^{\gamma\kappa|n |} \} \leq 1/\{2\kappa(\xi_s(0))^s\}. \label{e.2} \end{equation} } {\bf Proof:}\ The starting point is the following equation: For $z\notin \hbox{\BB R}$, \begin{equation} \sum_{n' \in \hbox{\BB Z} [i]} \langle n |M^\lambda - z |n' \rangle \langle n' |\Gamma^\lambda (z) |n'' \rangle = \delta_{nn''}.\end{equation} This becomes using (\ref{b.3}) \begin{equation} (c^\lambda_n - z) \langle n |\Gamma^\lambda (z) |n'' \rangle - \sum_{n' \not= n} G_0^\lambda (n, n') \langle n' |\Gamma^\lambda (z) |n'' \rangle = \delta_{nn''}. \end{equation} Now we take $n \not= n''$ and $0 < s < 1$ to get \begin{eqnarray} |c_n^\lambda - z |^s |\langle n | \Gamma^\lambda (z) |n'' \rangle |^s & = &|\sum_{n' \not= n} G_0^\lambda (n,n') \langle n' |\Gamma^\lambda (z) |n'' \rangle |^s \nonumber \\ & \leq &\sum_{n'\not= n} |G_0^\lambda (n, n') |^s |\langle n' |\Gamma^\lambda (z) |n'' \rangle |^s. \nonumber \end{eqnarray} Thus \begin{equation}\hbox{\BB E} \{ |c^\lambda_n - z |^s |\langle n |\Gamma^\lambda (z) |n'' \rangle |^s \} \leq \sum_{n' \not= n} |G_0^\lambda (n,n') |^s \hbox{\BB E} \{ |\langle n' |\Gamma^\lambda (z) |n'' \rangle |^s\}. \label{e.2b} \end{equation} Now \begin{equation} \hbox{\BB E} \{ |c_n^\lambda - z |^s |\langle n |\Gamma^\lambda (z) |n'' \rangle |^s \} = {\tilde {\hbox{\BB E}}}_n \hbox{\BB E}_n \{|c_n^\lambda - z |^s |\langle n |\Gamma^\lambda (z) |n'' \rangle |^s\} \end{equation} where $\hbox{\BB E}_n$ is the expectation with respect to $\omega_n$ and ${\tilde {\hbox{\BB E}}}_n$ is with respect to all other $\omega_{n'}$'s. Let \begin{equation} \langle n' |M_n^\lambda |n'' \rangle = \langle n' |M^\lambda |n'' \rangle - (4\kappa/\omega_n) \delta_{nn'} \delta_{nn''}. \end{equation} Then $M^\lambda_n$ is independent of $\omega_n$ and using the resolvent identity \begin{equation} \langle n |\Gamma^\lambda (z) |0 \rangle = {A \over {1 + (4\kappa/\omega_n) B}} \end{equation} where $A= \langle n|(M_n^\lambda - z)^{-1} |0 \rangle$ and $B= \langle n |(M_n^\lambda - z)^{-1} |n \rangle$. Then \begin{eqnarray} \hbox{\BB E}_n \{|c_n^\lambda - z |^s \!\!\!\!\!\! &| &\!\!\!\!\!\!\langle n | \Gamma^\lambda (z) |0 \rangle |^s \} \nonumber \\ & = & \hbox{\BB E}_n \left\{{{|c_n^\lambda - z |^s} \over {|1 + (4\kappa/\omega_n)B |^s}} \right\} |A |^s \nonumber \\ & \geq &(4\kappa)^s \int {{|u -\eta|^s} \over {|1+4\kappa u B|^s}} \mu (du) |A |^s \nonumber \end{eqnarray} where $u=1/\omega_n$ and $2\pi \eta= \psi(- \lambda)-{{\pi} \over {2 \kappa}} E $, $E$ being the real part of $z$. Thus by Lemma 5.1 \begin{eqnarray} \hbox{\BB E}_n \{|c_n^\lambda - z |^s |\langle n | \Gamma^\lambda (z) |0 \rangle |^s \} & \geq &(4\kappa)^s (\xi_s (|\eta |))^s \hbox{\BB E}_n {{|A |^s} \over {|1 + c_n^\lambda B|^s}} \nonumber \\ & = &(4\kappa)^s (\xi_s (|\eta|))^s \hbox{\BB E}_n \{|\langle n | \Gamma^\lambda (z) |0 \rangle |^s\} \nonumber \end{eqnarray} Using (\ref{e.2b}) this gives \begin{equation} (4\kappa)^s(\xi_s (|\eta |))^s \hbox{\BB E} \{ |\langle n | \Gamma^\lambda (z) |0 \rangle |^s \} \leq \sum_{n' \not= n} |G_0^\lambda (n,n') |^s \hbox{\BB E} \{|\langle n' |\Gamma^\lambda (z) |0 \rangle |^s \}. \end{equation} or \begin{equation} \hbox{\BB E} \{ |\langle n |\Gamma^\lambda (z) |0 \rangle |^s \} \leq(1/4\kappa)^s(\xi_s (|\eta|))^{-s} \sum_{n' \not= n} |G_0^\lambda (n, n') |^s \hbox{\BB E} \{ |\langle n' |\Gamma^\lambda (z) |0 \rangle |^s \}. \end{equation} Let $\gamma>0$ and define \begin{equation} \Xi (s) = \hbox{\BB E}\{\sum_{n \in \hbox{\BB Z} [i]} e^{{\gamma\kappa} |n |} |\langle n |\Gamma^\lambda (z) |0 \rangle |^s \}. \end{equation} Then \begin{eqnarray} \Xi (s) &=& \hbox{\BB E} \{|\langle 0 |\Gamma^\lambda (z) |0 \rangle |^s \} + \sum_{n \not= 0} e^{{\gamma\kappa} |n |} \hbox{\BB E}\{|\langle n |\Gamma^\lambda (z) |0 \rangle |^s \}\nonumber \\ &\leq &\hbox{\BB E} \{|\langle 0 |\Gamma^\lambda (z) |0 \rangle |^s \}\nonumber \\ & & \ \ \ \ + (1/4\kappa)^s(\xi_s (|\eta|))^{-s} \sum_{n \not= 0} \sum_{n' \not= n} e^{{\gamma\kappa} |n |} |G_0^\lambda (n,n') |^s \hbox{\BB E}\{|\langle n' | \Gamma^\lambda (z) |0 \rangle |^s\}.\nonumber \end{eqnarray} Thus \begin{eqnarray} \Xi (s) &\leq &\hbox{\BB E}\{|\langle 0 |\Gamma^\lambda (z) |0 \rangle |^s\}\nonumber \\ & & + (1/4\kappa)^s (\xi_s (|\eta |))^{-s} \sum_{n'} \sum_{n\neq n'} \!\!\!\! e^{{\gamma\kappa} |n - n' |} |G_0^\lambda (n,n') |^s e^{\gamma\kappa |n' |} \hbox{\BB E} \{|\langle n'|\Gamma^\lambda (z) |0 \rangle |^s\}\nonumber \end{eqnarray} so that \begin{equation} \Xi (s) \leq \hbox{\BB E} \{ |\langle 0 |\Gamma^\lambda (z) |0 \rangle |^s\} + (1/4\kappa)^s \sup_{n'} {{\sum_{n \neq n'}e^{{\gamma\kappa} |n - n' |} |G_0^\lambda (n,n') |^s} \over {(\xi_s |\eta |))^s}} \Xi (s).\end{equation} Let \begin{equation} F(s,\lambda) = (1/4\kappa)^s \sup_{n'} {{\sum_{n \not= n'} e^{{\gamma\kappa} |n - n' |} |G_0^\lambda (n,n') |^s} \over {\left(\xi_s (|\eta |)\right)^s}}. \end{equation} If $F (s,\lambda) < 1/2$ then \begin{equation} \Xi (s) \leq {{\hbox{\BB E}\{|\langle 0 |\Gamma^\lambda (z) |0 \rangle |^s\}} \over {1 - F (s, \lambda)}} \leq 2\hbox{\BB E}\{|\langle 0 |\Gamma^\lambda (z) |0 \rangle |^s\}. \end{equation} But \begin{equation} \langle 0 |\Gamma^\lambda (z) |0 \rangle = {{\langle 0 |(M_0^\lambda - z)^{-1} |0 \rangle} \over {1 + c_0^\lambda \langle 0 |(M_0^\lambda - z)^{-1} |0 \rangle}} \end{equation} so that \begin{equation} |\langle 0 |\Gamma^\lambda (z) |0 \rangle | = {1 \over {4\kappa|b+1/\omega_0 |}} \end{equation} where $b$ is independent of $\omega_0$. Using this and Lemma 5.1 with $a=1$ and $\eta=-b$, we get \begin{equation} \hbox{\BB E}_0\{|\langle 0 |\Gamma^\lambda (z) |0 \rangle |^s\}\leq 1/(4\kappa \xi_s(0))^s \end{equation} and therefore \begin{equation} \hbox{\BB E}\{| \langle 0 |\Gamma^\lambda (z) |0 \rangle |^s\} \leq 1/(4\kappa \xi_s(0))^s. \end{equation} This proves (\ref{e.2}). To prove that $F (s,\lambda) < 1/2$, First assume that $ \lambda \in I_N$ with $N\in \hbox{\BB N}$. By Proposition 6.2, we have \begin{equation} |G_0^\lambda (n,n') |\leq (C \kappa)^{N +1}N^N |\Gamma (- \lambda) ||n - n' |^{2N} e^{- \kappa |n - n'|^2} \end{equation} for $n \not= n'$, and $ \lambda \in I_N$, $N\in \hbox{\BB N}$. Therefore \begin{equation} \sum_{n \not= n'} e^{{\gamma\kappa} |n - n' |} |G_0^\lambda (n, n') |^s \leq (C \kappa)^{(N+1)s} N^{Ns} |\Gamma (- \lambda) |^s \sum_{n \not= 0} e^{{\gamma\kappa} |n |} |n |^{2Ns} e^{- \kappa s |n |^2}. \end{equation} Let $\gamma< \alpha <s $. Using the bounds $ e^{{\gamma\kappa}x} e^{- \alpha \kappa x^2}\leq e^{-(\alpha - \gamma)\kappa}$ for $x\geq 1$, \break $ x^{2Ns} e^{- \kappa (s -\alpha)x^2/2} < (2sN/(s-\alpha)\kappa )^{Ns}$ for $x\geq 1$ and \begin{equation} \sum_{n \in \hbox{\BB Z}[i]} e^{- t |n |^2} \leq K(t) \end{equation} where $K(t)=\(1+e^{-t/4}+(\pi/t)^{1/2}\)^2 $ for $t>0$ (see Lemma 2.1 in \cite{DMP1}), we get \begin{eqnarray} \sum_{n \not= n'} e^{{\gamma\kappa} |n - n' |}\!\!\!\!\!\!\! &&\!\!\!\!\!\!\! |G_0^\lambda (n, n') |^s \nonumber \\ &\leq & \(K(\kappa (s-\alpha)/2) -1\) (C^{N+1}\kappa)^s (2sN^2/(s-\alpha))^{Ns} |\Gamma (- \lambda) |^s e^{-(\alpha -\gamma)\kappa}.\nonumber \end{eqnarray} Thus \begin{equation} F(s, \lambda) \leq \(K\({{\kappa (s-\alpha)}\over 2}\) -1\) (C^{N+1}/4)^s (2sN^2/(s-\alpha))^{Ns} e^{-(\alpha -\gamma)\kappa} \left |{{ \Gamma(- \lambda)} \over {\xi_s (|\eta |) }}\right|^s. \end{equation} Now since for $N\in \hbox{\BB N}$, the limits $\lim_{\lambda \to N}|\Gamma(-\lambda)/\psi(-\lambda) | $ and \hfill \break $\lim_{\lambda \to N-1}|\Gamma(-\lambda)/\psi(-\lambda) | $ are finite, we have for $\lambda\in I_N$, \begin{equation} |\Gamma(-\lambda)| \leq C_N (1 + |\psi(-\lambda) |) \leq C_N(1 + (\pi/2\kappa)+2\pi|\eta|). \end{equation} Therefore \begin{equation} \left |{{ \Gamma(- \lambda)} \over {\xi_s (|\eta |) }}\right| \leq C_N\(\[\{1+(\pi/2\kappa)\}/\xi_s (0)\] +2\pi\sup_{x\in \hbox{\BB R}_0}\{x/\xi_s (x)\}\). \end{equation} Thus there exists $\kappa (N_1,s) < \infty$ such that for all $\kappa > \kappa_0 (N_1,s)$, $F(s,\lambda)< 1/2$ for all $\lambda\in (0, N_1)\setminus \hbox{\BB N}$. \par For $\lambda \in I_0$ we have by Proposition 6.1 \begin{equation} |G_0^\lambda (n,n') |\leq (C \kappa) { 1\over {|\lambda|}} e^{- \kappa |n - n' |^2} \end{equation} and therefore \begin{equation} \sum_{n \not= n'} e^{{\gamma\kappa} |n - n' |} |G_0^\lambda (n, n') |^s \leq \(K(\kappa (s-\alpha)) -1\) (C\kappa)^s \left ( { 1\over {|\lambda|}} \right )^s e^{-(\alpha -\gamma)\kappa} \end{equation} and thus \begin{equation} F(s, \lambda) \leq \(K(\kappa (s-\alpha)) -1\)(C/4)^s e^{-(\alpha -\gamma)\kappa} s \left ( {1\over {|\lambda|\xi_s (|\eta|)}} \right )^s. \end{equation} Therefore by the same argument $F(s,\lambda)< 1/2$ for all $\lambda\in I_0$ if $\kappa$ is large enough. \par\noindent\rightline{$\square$} \par We have from Theorems 8 and 9 in \cite{SW} that if for all $E\in(-1,1)$ and a.e. $\omega$, \begin{equation} \lim_{\epsilon\downarrow 0} \sum_{n \in\hbox{\BB Z}[i]}|\langle n|\Gamma^\lambda(E+i\epsilon)|0\rangle|^2 <\infty, \label{e.3a} \end{equation} then $\sigma_{{\rm cont}}(M^\lambda)\cap(-1,1)=\emptyset $ for a.e. $\omega$. If, furthermore, for a.e. pair $(\omega, E)$, $\omega\in\Omega$ and $E\in (-1,1)$, \begin{equation} \lim_{\epsilon\downarrow 0} |\langle n |\Gamma^\lambda(E+i\epsilon)|0\rangle|<C_{\omega,E}e^{-m(E)|n|}, \label{e.3b}\end{equation} then with probability one, the eigenvectors $v^\lambda_E$ of $M^\lambda$ with eigenvalue $E\in (-1,1)$ obey \begin{equation} |\langle v^\lambda_E |n \rangle|<D_{\omega,E}e^{-m(E)|n|}. \label{e.4} \end{equation} We shall use the results of \cite{SW}, Lemma 5.2 and Proposition 4.1 to prove the following lemma. \par\noindent {\bf Lemma 5.3}\ {\sl For each $N\in\hbox{\BB N}$ there exists $\kappa_0 > 0$ such that for $\kappa >\kappa_0$, for each $\lambda\in (-\infty,N)\setminus \hbox{\BB N}_0 $ with probability one, if $\lambda $ is a generalized eigenvalue of $H$ with corresponding generalized eigenfunction $\phi_\lambda$, then for any compact subset $B$ of $\hbox{\BB C}$, $\int_B|\phi_\lambda(z-z')|^2dz'$ decays exponentially in $z$ with exponential length less than or equal to $ 2/\kappa $.} \newline {\bf Proof:}\ \ From Lemma 5.2 we have for all $\lambda \in(-\infty,N)\setminus \hbox{\BB N}_0$, $z\in \hbox{\BB C}$ with ${\cal I}z\neq 0$ and $|{\cal R}z |\leq 1$, \begin{eqnarray} \hbox{\BB E} \Bigl \{\Bigl [\sum_{n \in \hbox{\BB Z} [i]} |\langle n |\Gamma^\lambda (z) |0 \rangle |^2 e^{2\gamma\kappa|n |/s}\Bigr ]^{s/2} \Bigr \} &\leq & \hbox{\BB E} \{\sum_{n \in \hbox{\BB Z} [i]} |\langle n |\Gamma^\lambda (z) |0 \rangle |^s e^{\gamma\kappa|n |} \} \nonumber \\ &\leq & 1/\{(2\kappa)(\xi_s(0))^s\}. \label{e.5} \end{eqnarray} Now for a.e. pair $(\omega, E)$, $\omega\in\Omega$ and $E\in (-1,1)$, $ \lim_{\epsilon\downarrow 0} \langle n |\Gamma^\lambda (E+i\epsilon) |0 \rangle $ exists. Therefore by Fatou's Lemma, \begin{eqnarray} \hbox{\BB E} \Bigl\{\Bigl [\sum_{n \in \hbox{\BB Z} [i]} \lim_{\epsilon\downarrow 0} |\langle n |\Gamma^\lambda (E \! \! \! \! \! &+&\! \! \! \! \! i\epsilon) |0 \rangle |^2 e^{2\gamma\kappa|n |/s} \Bigr ]^{s/2}\Bigr \} \nonumber \\ &\leq & \hbox{\BB E} \Bigl \{\liminf_{\epsilon\downarrow 0} \Bigl [\sum_{n \in \hbox{\BB Z} [i]} |\langle n |\Gamma^\lambda (E+i\epsilon) |0 \rangle |^2 e^{2\gamma\kappa|n |/s} \Bigr]^{s/2}\Bigr \} \nonumber \\ &\leq & \liminf_{\epsilon\downarrow 0}\hbox{\BB E} \Bigl \{ \Bigl [ \sum_{n \in \hbox{\BB Z} [i]} |\langle n |\Gamma^\lambda (E+i\epsilon) |0 \rangle |^2 e^{2\gamma\kappa|n |/s} \Bigr ]^{s/2} \Bigr \} \nonumber \\ &\leq & 1/\{(2\kappa)(\xi_s(0))^s\}. \label{e.6} \end{eqnarray} Thus (\ref{e.3a}) and (\ref{e.3b}) are satisfied. Therefore if $\lambda$ is a generalized eigenvalue of $H$ with corresponding generalized eigenfunction $\phi_\lambda$, then by Proposition 4.1, $v_\lambda= \Gamma^{\lambda_\kappa}U_{\lambda_\kappa}\phi_\lambda$ is a generalized eigenvector of $M^\lambda$ with eigenvalue $0$ and must satisfy \begin{equation} |\langle v_\lambda |n \rangle|<D_{\omega }e^{-2\gamma\kappa|n |/s}. \label{e.7} \end{equation} Then again by Proposition 4.1, for any compact subset $B$ of $\hbox{\BB C}$, $\int_B|\phi_\lambda(z-z')|^2dz'$ decays exponentially in $z$ with exponential length less than or equal to \hfill\break $\max(s/(2\gamma\kappa), 2/\kappa)$. If we choose $\gamma=s/2$, then $\max (s/(2\gamma\kappa), 2/\kappa)= 2/\kappa $.\par \par\noindent\rightline{$\square$} \par By Fubini's Theorem, we can deduce from Lemma 3.5 the result about the decay of eigenfunctions with probability one and a.e. $\lambda$ with respect to Lebesgue measure and therefore with probability one $\sigma_{{\rm ac}}(H)\cap (-\infty, N)=\emptyset $. However to be able to make a statement about $\sigma_{{\rm cont}}(H)$ we have to replace {\sl a.e. $\lambda$ with respect to Lebesgue measure} with {\sl a.e. $\lambda$ with respect to the spectral measure of $H(\omega)$}. We do this in the Lemma 5.7 by using the ideas of \cite{DLS} and the following four lemmas. \par\noindent We state the first lemma without proof. \vskip .2 cm \noindent {\bf Lemma 5.4} \ {\sl Let $\{f_n\}$ be a total countable subset of normalized vectors of a Hilbert space ${\cal H}$ and $H$ a self-adjoint operator on ${\cal H}$ with spectral projections ${\bf E}(\ \cdot \ )$. Let $c_n > 0$, $\sum_n c_n < \infty$ and $\nu = \sum_n c_n \mu_n$, where $\mu_n = (f_n, {\bf E} (\ \cdot \ ) f_n)$. Then $\nu(A) = 0$ implies that ${\bf E} (A) = 0$.} \par\noindent {\bf Lemma 5.5} \ {\sl For each $N \in \hbox{\BB N}_0$, there exists an open set $J_N\subset \hbox{\BB C}$, containing $I_N$, such that for $\kappa$ sufficiently large with probability one, $M^\lambda$ is invertible for all $\lambda \in J_N \setminus I_N$.} \par \noindent {\bf Proof }\ Let $\lambda \in I_N$ and $| \epsilon | < 1$, $\epsilon \neq 0$. Let \begin{equation} \langle n | X | n' \rangle = \cases{c_n^{\lambda + i \epsilon} & if $n = n'$ \cr - G_0^\lambda (n,n') & if $n \neq n'$ \cr} \end{equation} Then $|| X \xi || \geq {{2 \kappa} \over \pi} |{\cal I} \psi (- (\lambda + i \epsilon))| ||\xi ||$. Therefore $X$ is invertible and \begin{equation} || X^{-1} || \leq {\pi \over {2 \kappa}} {1 \over {|{\cal I}(\psi (- (\lambda + i \epsilon))|}} \end{equation} Let \begin{equation} \langle n | Y | n' \rangle = \cases{0 & if $n = n'$ \cr - i \epsilon (G_0^\lambda G_0^{\lambda + i \epsilon})(n, n') & if $n \neq n'$ \cr} \end{equation} so that \begin{equation} M^\lambda = X + Y = X (1 + X^{-1} Y). \end{equation} From Proposition 6.2 in Appendix A we have for $\lambda $ with ${\cal R} \lambda \in I_N$, $N\in \hbox{\BB N}$, and $|{\cal I} \lambda | \leq 1$, \begin{equation} | G_0^\lambda (z, z') | \leq C^N \kappa N^N | \Gamma (- {\cal R}\lambda )| (1 + \ln (2 \kappa | z - z' |^2) \end{equation} for $2 \kappa | z - z' |^2 < 1$ and \begin{equation} | G_0^\lambda (z, z') | \leq C^N \kappa N^{2N} | \Gamma(- {\cal R} \lambda) | e^{- {\kappa \over 2} | z - z' |^2} \end{equation} for $2 \kappa | z - z' |^2 \geq 1$. \par \noindent Therefore if $\lambda \in I_N$, $N\in \hbox{\BB N}$, $| \epsilon | <1$, $\kappa > 2$ and $ n, n' \in \hbox{\BB Z} [i]$ with $n\neq n'$, \begin{eqnarray} | (G_0^\lambda G_0^{\lambda + i \epsilon}) (n, n') | &\leq & C^{2N}\kappa^2 | \Gamma (-\lambda)|^2 N^{3N}\nonumber \\ & & \ \ \ \ \ \ \times \Bigl \{\int_{2 \kappa | z - n |^2 < 1} dz (1 + \ln (2\kappa | z - n |^2) e^{- {\kappa \over 2} | z - n' |^2} \nonumber \\ & + & \ \ \ \ \ \ \ \int_{2 \kappa | z - n' |^2 < 1} dz (1 + \ln (2 \kappa | z - n' |^2) e^{- {\kappa \over 2} | z - n |^2} \nonumber \\ & + & N^N \int dz e^{- {\kappa \over 2} | z - n|^2} e^{- {\kappa \over 2} | z - n' |^2} \Bigr \} \nonumber \\ & \leq & 2\pi C^{2N} \kappa | \Gamma (- \lambda) |^2 N^{3N} e^{-{\kappa \over 8} | n - n' |^2}\!\! \left\{ \int^1_0 \!\!(1 + \ln r^2) rdr + N^N e^{- {\kappa \over 8}}\right\} \nonumber \\ & \leq & 2 C^{2N} \kappa^2 | \Gamma (- \lambda) |^2 N^{4N} e^{- {\kappa \over 8} | n - n' |^2} \nonumber \\ & \leq & e^{- {\kappa \over {32}}} C^{2N} | \Gamma (- \lambda) |^2 N^{4N} e^{- {1 \over 8} | n - n' |^2} \end{eqnarray} if $\kappa$ is large enough. Therefore \begin{equation} || Y || \leq \epsilon e^{- {\kappa \over {32}}} C^{2N} | \Gamma (- \lambda) |^2 || T || \end{equation} where $T$ is the operator with matrix $\langle n | T | n' \rangle = e^{- {1/8} | n - n' |^2}$. Now take $\lambda \in (N - 1, N - {1\over 2}]$ and $| \epsilon | < \lambda - N + 1$. In this interval \begin{equation} | \Gamma (- \lambda) | \leq {{a_N} \over {(\lambda - N + 1)}} \end{equation} On the other hand by \cite{AS} 6.3.16 \begin{equation} {\cal I} \psi (- (\lambda + i \epsilon)) = - \epsilon \sum^\infty_{k = 0} {1 \over {(\lambda - k)^2 + \epsilon^2}} \end{equation} Therefore \begin{eqnarray} | {\cal I}\psi (- (\lambda + i \epsilon))| & = & | \epsilon | \sum^\infty_{k = 0} {1 \over {(\lambda - k)^2 + \epsilon^2}} >{{| \epsilon |} \over {(\lambda - N + 1)^2 + \epsilon^2}} \nonumber \\ & > & {{| \epsilon |} \over {2(\lambda - N + 1)^2}}. \end{eqnarray} \par \noindent Thus \begin{equation} || X^{-1} Y || \leq || X^{-1} || || Y || \leq {\pi \over {4 \kappa}} a^2_N C^{2N} e^{- {\kappa \over {32}}} || T || < 1 \end{equation} if $\kappa$ is sufficiently large. Thus $M^{\lambda + i \epsilon }$ is invertible. We can use the same argument if $\lambda \in [N - {1\over 2}, N)$ and $ | \epsilon | < N - \lambda $. \par Using the bounds in Proposition 6.1, a similar calculation to the above gives for $\lambda \in I_0$, \begin{equation} || Y || \leq \epsilon e^{- {\kappa \over {32}}} C^{2N} | \lambda |^{-2} || T ||. \end{equation} Then using the inequality \begin{equation} | {\cal I}\psi (- (\lambda + i \epsilon))| >{{| \epsilon |} \over {|\lambda|^2 + \epsilon^2}}, \end{equation} we can show that $M^{\lambda + i \epsilon }$ is invertible if $ | \epsilon | < |\lambda| $. \par\noindent\rightline{$\square$} \par \noindent {\bf Lemma 5.6} \ {\sl For $n \in \hbox{\BB Z} [i]$ and $\lambda \in J_N \setminus I_N$, let $\phi^\lambda_n = c_{\lambda, n} U_\lambda^* \Gamma^\lambda | n \rangle$ where $c_{\lambda, n} = || U_\lambda^* \Gamma^\lambda | n \rangle ||^{-1}$ so that $|| \phi_n^\lambda || = 1$. Then if $[a,b]\subset I_N$, the set $\{\phi^\lambda_n: n \in \hbox{\BB Z} [i], \ \lambda \in (J_N \setminus I_N )\cap \hbox{\BB Q} [i]\}$ is total in ${\bf E} ([a,b]) {\cal H}$. } \par \noindent {\bf Proof:}\ For $n \in \hbox{\BB Z} [i]$ and $\lambda \in J_N$ let ${\tilde \phi}_n^\lambda = U^*_\lambda | n \rangle $. Then if $\lambda \in J_N \setminus I_N$, \begin{equation} {\tilde \phi}^\lambda_n = \sum_{n' \in \hbox{\BB Z} [i]} c^{-1}_{\lambda, n'} \langle n' | M^\lambda | n \rangle \phi^\lambda_{n'}. \end{equation} Also if $\lambda \to \lambda'$ then $ {\tilde \phi}^\lambda_n \to {\tilde \phi}_n^{\lambda'}$. Therefore it is sufficient to prove that the set $\{{\tilde \phi}^\lambda_n: n \in \hbox{\BB Z} [i], \lambda \in I_n\}$ is total. We do this by showing that the orthogonal complement of this set is in the orthogonal complement of ${\bf E} ([a,b]) {\cal H}$. \par Let $f \in {\cal H}$ and suppose that $({\tilde \phi}^\lambda_n, f) = 0 $ for all $n \in \hbox{\BB Z} [i]$ and all $\lambda \in [a,b]$. Then since $(G^\lambda_0 f) (n) = ({\tilde \phi}^\lambda_n, f)$, $G^\lambda f = G^\lambda_0 f$. Therefore ${\bf E} ([a,b]) G^\lambda f = {\bf E} ([a,b]) G^\lambda_0 f$ and thus \begin{equation} \sup_{\lambda \in [a,b]} || {\bf E} ([a,b]) G^\lambda f || \leq \sup_{\lambda \in [a,b]} || G^\lambda_0 || || f || < \infty. \end{equation} Let $ \mu_1 (A) = (f, {\bf E} ([a,b] \cap A) f)$. Then \begin{equation} || {\bf E} ([a,b])G^\lambda f ||^2 = \int_{[a,b]} {{\mu_1 (d \lambda')} \over {| \lambda - \lambda' |^2}} . \end{equation} Let $x_i = a + (b - a)i/M$, $i = 0,\ldots, M $ and $\lambda_i = {1\over 2} (x_i + x_{i+1})$. Then \begin{equation} \int_{[a,b]} {{\mu_1 (d \lambda')} \over {| \lambda' - \lambda_i |^2}} \geq \int_{[x_i, x_{i + 1}]} {{\mu_1 (d\lambda')} \over {| \lambda' - \lambda_i |^2}} \geq {{4M^2} \over {(b - a)^2}} \mu_1 ([x_i, x_{i + 1}]) \end{equation} Therefore \begin{equation} \sup_{\lambda \in [a,b]} \int_{[a,b]} {{\mu_1(d \lambda')} \over {| \lambda' - \lambda |^2}} \geq {{4M^2} \over {(b-a)}} \mu_1([x_i, x_{i+1}]) \end{equation} for all $i$, and so \begin{equation} \sup_{\lambda \in [a,b]} \int_{[a,b]} {{\mu_1 (d \lambda')} \over {| \lambda -\lambda'|^2}} \geq {{4M^2} \over {(b-a)^2}} {1 \over M} \sum^M_{i=0} \mu_1 ([x_i, x_{i+1}] \geq {{4M} \over {(b-a)^2}} \mu_1 ([a,b]). \end{equation} Since $M$ is arbitrary $\sup_{\lambda \in [a,b]} || {\bf E} ([a,b]) G^\lambda f || = \infty$ unless $\mu_1 ([a,b]) = 0$. But $\mu_1 ([a,b]) = || {\bf E} ([a,b]) f ||^2$. \par\noindent\rightline{$\square$}\par Let ${\cal F}$ be the $\sigma$-algebra generated by $\{\omega_{n'}: n'\in \hbox{\BB Z}[i]\}$ and let ${\cal F}_n$ be the sub $\sigma$-algebra generated by $\{\omega_{n'}: n'\neq n\}$. Let ${\cal B}_N$ be the Borel sets of $I_N$. \par\noindent {\bf Lemma 5.7}\ {\sl Let $ B\mapsto {\bf E}(B) $ be the spectral measure of $H$ and \break\hfill $ A\in \cap_{n\in \hbox{\BB Z}[i]}({\cal F}_n\otimes{\cal B}_N)$. If for a.e $\lambda \in I_N $ with respect to Lebesgue measure $\hbox{\BB E} \left \{ {\bf 1}_A(\ \cdot\ ,\lambda)\right \}=0$, then $\hbox{\BB E} \left \{{\bf E}(\{\lambda:\ ( \ \cdot\ ,\lambda)\in A\})\right \}=0$.} \par\noindent {\bf Proof:}\ If for a.e $\lambda$ with respect to Lebesgue measure $\hbox{\BB E} \left \{ {\bf 1}_A(\ \cdot\ ,\lambda)\right \}=0$, then by Fubini's Theorem $\hbox{\BB E} \left \{ \int_{I_N} d\lambda {\bf 1}_A(\ \cdot\ ,\lambda)\right \}=0$. \par Let $\Lambda$ be a bounded subset of $\hbox{\BB Z}[i]$ and let $H_\Lambda$ be defined in the same way as $H$ with ${\cal M}$ replaced by ${\cal M}_\Lambda = l^2(\Lambda)$. By the same argument as in Proposition 4.1 $\lambda \notin \hbox{\BB N}_0$ is an eigenvalue of $H_\Lambda$ if and only if there exists $v \in {\cal M}_\Lambda $ such that $M_\Lambda^\lambda v = 0$, where $M_\Lambda^\lambda$ is the restriction of $M^\lambda $ to $ {\cal M}_\Lambda $. Then the corresponding eigenfunction is $U^*_\lambda v $. Since in the interval $I_N$, $\psi$ is bijective it is clear that there are $| \Lambda |$ eigenvalues in $I_N$. \par Let $\lambda_1, \ldots , \lambda_{| \Lambda |}$ be the eigenvalues in $I_N$, say, and let $v_1, \ldots, v_{| \Lambda |}$ be the corresponding vectors such that $M_\Lambda^{\lambda_k} v_k = 0$. Let $u_n = 1/\omega_n$. Then for $n \in \Lambda $ we get \begin{equation} {{d \lambda_k} \over {d u_n}} = -{{| \langle v_k | n \rangle |^2} \over {|| U^*_{\lambda_k} v_k ||}} \label{e.8} \end{equation} If $M_\Lambda^{\lambda_k} v_k = 0$ and $\langle v_k | n \rangle = 0$ for a particular value of $u_n$ then $M_\Lambda^{\lambda_k} v_k = 0$ for all values of $u_n$. We shall see later that we can ignore these eigenvalues. \par We see from equation (\ref{e.8}) that each $\lambda_k$ is a monotonic decreasing function of $u_n$. Moreover as $u_n \to \pm \infty$, the $\lambda_k$'s become identical, except the value of $\lambda_k$ corresponding to the $v_k$ which tends to $| n \rangle$ and this latter value of $\lambda_k$ decreases from $N$ to $N-1$ (respectively $-\infty $ if $N=0$) as $u_n$ increases from $-\infty$ to $+\infty$. Therefore \begin{equation} \sum_k \int_{-\infty}^\infty f(\lambda_k) {{d \lambda_k} \over {d u_n}} du_n = -\int_{I_N} f(\lambda) d \lambda. \label{e.9} \end{equation} Let $\psi_k = {{U^*_{\lambda_k} v_k} \over {|| U^*_{\lambda_k} v_k ||}}$ so that $H_\lambda \psi_k = \lambda_k \psi_k$. Let $\lambda \in J_N\setminus I_N$ and $n \in \hbox{\BB Z} [i]$. For $B\subset I_N$ let $\mu^{n,\lambda}_\Lambda (B) = (\phi^\lambda_n, {\bf E}_\Lambda (B) \phi^\lambda_n)$ where ${\bf E}_\Lambda$ is the spectral measure of $H_\Lambda$. Then for $\Lambda$ sufficiently large \begin{eqnarray} \mu^{n,\lambda}_\Lambda (B) & = &\sum_{\lambda_k \in B} | (\phi^\lambda_n, \psi_k) |^2 \nonumber \\ & = &|| \Gamma^\lambda | n \rangle ||^{-2} \sum_{\lambda_k \in B} {1 \over {| \lambda - \lambda_k |^2}} {{| \langle n | v_k \rangle |^2} \over {|| U^*_{\lambda_k} v_k ||^2}} \nonumber \\ & = & - || \Gamma^\lambda | n \rangle ||^{-2} \sum_{\lambda_k \in B} {1 \over {| \lambda - \lambda_k|^2}} {{d \lambda_k} \over {d u_n}}. \label{e.10} \end{eqnarray} Note that if $\langle n | v_k \rangle=0$ then the corresponding term in (\ref{e.10})is absent. Also if $\lambda_k$ is degenerate, we can choose the corresponding orthogonal set of eigenvectors so that only one satisfies $\langle n | v_k \rangle \neq 0 $. Therefore there is only one term corresponding to such $\lambda_k$ in the sum (\ref{e.10}). From (\ref{e.10}) and (\ref{e.9}) we get \begin{equation} \int_{-\infty}^\infty du_n \mu^{n,\lambda}_\Lambda (B) = || \Gamma^\lambda | n \rangle ||^{-2} \int_B {{d \lambda'} \over {| \lambda - \lambda' |^2}}, \end{equation} and thus \begin{equation} \int_{-\infty}^\infty du_n \rho(u_n) \mu^{n,\lambda}_\Lambda (B) \leq || \Gamma^\lambda | n \rangle ||^{-2} || \rho ||_\infty \int_B {{d \lambda'} \over {| \lambda - \lambda' |^2}}. \label{e.11} \end{equation} If $\mu^{n,\lambda}(B) = (\phi^\lambda_n, {\bf E}(B) \phi^\lambda_n)$ then by the weak convergence of $\mu^{n,\lambda}_\Lambda $ to $\mu^{n,\lambda}$ we have the bound (\ref{e.11}) for $\mu^{n,\lambda}$. By Kotani's argument \cite{K}, we have that \hfill\break $\hbox{\BB E} \left \{ \int_{I_N} d\mu^{n,\lambda}(\lambda') {\bf 1}_A(\ \cdot\ ,\lambda')\right \}=0$. By Lemmas 5.4 and 5.6 we get that \hfill\break $\hbox{\BB E} \left \{{\bf E}(\{\lambda:\ ( \ \cdot\ ,\lambda)\in A\})\right \}=0$.} \par\noindent\rightline{$\square$} \par By combining Lemmas 5.3 and 5.7 we obtain our final theorem. \par\noindent {\bf Theorem 5.8}\ {\sl For each $N\in\hbox{\BB N}$ there exists $\kappa_0 > 0$ such that for $\kappa >\kappa_0$, with probability one, $\sigma_{{\rm cont}}(H)\cap (-\infty, N)=\emptyset $, and if $\lambda \in \sigma(H)\cap (-\infty,N)\setminus \hbox{\BB N}_0$, is an eigenvalue of $H$ and the corresponding eigenfunction is $\phi_\lambda $, then for any compact subset $B$ of $\hbox{\BB C}$, $\int_B|\phi_{\lambda}(z-z')|^2dz'$ decays exponentially in $z$ with exponential length less than or equal to $ 2/\kappa$.} \vskip.25cm \noindent {\bf Acknowledgements} \par This work was supported by the Forbairt (Ireland) International Collaboration Programme 1997. J.V.P. and T.C.D. would like to thank the Institut de physique th\'eorique of the Ecole Polytechnique F\'ed\'erale de Lausanne for their hospitality and financial support. T.C.D. and N.M. would like to thank University College Dublin for their hospitality. \section{ Appendix A.\ Bounds for the Green's Function} In this appendix we shall obtain bounds on the Green's function $G^\lambda_0(z,z')$. Our basic tools are the the integral representation (\cite{AS} 13.2.5) \begin{equation} \Gamma (a) U(a, b, \rho) = \int^\infty_0 dt e^{- \rho t } t^{a-1} (1 + t)^{b-a-1}, \label{f.1} \end{equation} which is valid for ${\cal R}a > 0$ and $ \rho > 0$ and the recurrence relation (\cite{AS} 13.4.18) \begin{equation} U(a,b,\rho) = \rho U(a+1, b+1, \rho) - (b-a-1) U(a+1, b,\rho). \label{f.2} \end{equation} We first obtain bounds for $| G_0^\lambda (z,z') | $ when ${\cal R}\lambda <0$. \par\noindent {\bf Proposition 6.1} \ {\sl There exists a constant $C<\infty$, such that for ${\cal R}\lambda \in I_0$, \begin{equation} | G^\lambda_0(z,z') | \leq C \kappa e^{- \kappa | z - z' |^2} \left \{ {1\over{|{\cal R }\lambda|}}+ e^{-\sqrt{2\kappa |{\cal R}\lambda|} |z-z'|} \(1+|\ln \(2\kappa | z - z' |^2\)|\)\right \}, \end{equation} if $2\kappa | z - z' |^2 \leq 1$, and \begin{equation} | G^\lambda_0(z,z') | \leq {{C \kappa}\over{|{\cal R }\lambda |}} e^{- \kappa | z - z' |^2}, \end{equation} if $2\kappa | z - z' |^2\ > 1$. } \par \noindent {\bf Proof:}\ Let $\lambda = x + iy $ with $ x < 0$. Then from (\ref{f.1}) we get \begin{eqnarray}|\Gamma (-\lambda) U(-\lambda, 1, \rho)| &\leq &\int^{\infty}_0 dt e^{- \rho t} t^{|x|-1} (1 + t)^{-|x|} \nonumber \\ & = &\int^1_0 dt e^{- \rho t} t^{|x|-1}(1+t)^{-|x| } + \int^{\infty}_1 dt e^{- \rho t} t^{|x|-1}(1 + t)^{-|x|} \nonumber \\ & \leq &\int^1_0 dt\ t^{|x|-1} + \int^{\infty}_1 dt {{e^{-\rho t}} \over t} \left({t \over {1+t}}\right)^{|x|} \nonumber \\ &\leq & {1 \over |x|} + \int^{\infty}_1 dt {{e^{- (\rho t +{{|x|}\over {2 t}})}} \over t}. \end{eqnarray} If $\rho \leq 1$, we have \begin{eqnarray} |\Gamma (-\lambda) U(-\lambda, 1, \rho)| &\leq & {1 \over |x|} + \int^{\infty}_1 dt {{ e^{- {{1\over 2}\rho t}} e^{- {1\over 2}(\rho t+{{|x|}\over { t}})}} \over t}\nonumber \\ &\leq & {1 \over |x|} + e^{- (\rho |x|)^{1\over 2} } \int^{\infty}_1 dt {{ e^{- {{1\over 2}\rho t}}}\over t} \nonumber \\ & = & {1 \over |x|} + e^{- (\rho|x|)^{1\over 2} }\int^{\infty}_{{1\over 2}\rho} dt {{e^{-t}} \over t} \nonumber \\ & \leq & {1 \over |x|} + e^{- (\rho|x|)^{1\over 2} }\int^{1}_{{1\over 2}\rho} {{dt} \over t} + e^{- (\rho|x|)^{1\over 2} }\int^{\infty}_{1} dt {{e^{-t}} \over t} \nonumber \\ & \leq & {1 \over |x|} - e^{-(\rho |x|)^{1\over 2} }\ln (\rho/2) + e^{- (\rho|x|)^{1\over 2} }\int^{\infty}_1 dt e^{-t} \nonumber \\ & = & {1 \over |x|} + e^{- (\rho |x|)^{1\over 2} }| \ln (\rho/2) | + {{ e^{- (\rho|x|)^{1\over 2} }} \over e}. \nonumber \end{eqnarray} Thus, \begin{equation} |\Gamma(-\lambda) U(-\lambda, 1, \rho) | \leq C\left \{ {1 \over |x|}+\left (1 + | \ln \rho |\right ) e^{- (\rho |x|)^{1\over 2} }\right \}. \label{f.3} \end{equation} If $\rho > 1$, we have \begin{eqnarray} |\Gamma (-\lambda) U(-\lambda, 1, \rho)| &\leq & {1 \over |x|} + \int^{\infty}_1 dt {{e^{- (t +{{|x|}\over {2 t}})}} \over t}\nonumber \\ &\leq & {1 \over |x|} + \int^{\infty}_1 dt {{ e^{- {{1\over 2} t}} e^{- {1\over 2}(t+{{|x|}\over { t}})}} \over t}\nonumber \\ &\leq & {1 \over |x|} + e^{- |x|^{1\over 2} } \int^{\infty}_1 dt e^{- {{1\over 2} t} } \nonumber \\ &= & {1 \over |x|} + 2 e^{- |x|^{1\over 2} } \nonumber \end{eqnarray} Therefore, \begin{equation} |\Gamma(-\lambda) U(-\lambda, 1, \rho) | \leq {C \over |x|}. \label{f.4} \end{equation} Inserting the inequalities (\ref{f.3}) and (\ref{f.4}) into (\ref{b.0}) we get Proposition 6.1. \par\noindent\rightline{$\square$} \par Now we shall obtain bounds for ${\cal R}\lambda >0$. \par\noindent {\bf Proposition 6.2} \ {\sl There exists a constant $C<\infty$, such that for ${\cal R}\lambda \in I_N$, $N\in \hbox{\BB N}$,$|{\cal I}\lambda |\leq 1$, \begin{equation} | G^\lambda_0 (z,z') | \leq \kappa C^N N^N | \Gamma (-{\cal R}\lambda)| (1 + | \ln (2\kappa | z - z' | ^2)|) e^{- \kappa | z - z' |^2}, \end{equation} if $2\kappa | z - z' |^2 \leq 1$, and \begin{equation} | G^\lambda_0(z,z') | \leq(C\kappa)^{N+1} N^N | \Gamma (-{\cal R}\lambda)| | z - z' |^{2N} e^{- \kappa | z - z' |^2}, \end{equation} if $2\kappa | z - z' |^2\ > 1$. } \par Let $\lambda =x+iy$. We shall prove that if $N-1 < x < N $, $N\in \hbox{\BB N}_0$, $b \in \hbox{\BB N}$ and $ \rho > 1$, then \begin{equation} | U(-\lambda,b,\rho) | \leq 2^{b + N - 1}\rho^x (b + N+|y|)^N \left | {{\Gamma(-x)}\over {\Gamma(-\lambda)}}\right | + e^{-(\rho-2)} (\rho+|y|+1)^N {{(b+N)!} \over {|\Gamma (N-\lambda)|}}. \label{f.6} \end{equation} We shall do this by induction on $N$. We first prove (\ref{f.6}) for $N=0$ by using (\ref{f.1}) which gives \begin{eqnarray} |\Gamma (-\lambda) U(-\lambda, b, \rho)| & \leq &\int^1_0\!\!\!\! dt e^{-\rho t} t^{-(x+1)} (1+t)^{b+x-1} + \int^\infty_1\!\!\!\!\!\! dt e^{-\rho t} t^{-(x+1)} (1 +t)^{b+x-1} \nonumber \\ & = &I_1 + I_2. \nonumber \end{eqnarray} We now take $\rho > 1$, $-1<x<0$ and $b\geq 1$. For $I_1$, since $t<1$, we get \begin{eqnarray} I_1 & \leq &2^{b-1}\int^1_0 dt e^{-\rho t} t^{-(x+1)} = 2^{b-1}\rho^x \int^\rho_0 dt e^{-t} t^{-(x+1)} \nonumber \\ & \leq & 2^{b-1}\rho^x \int^\infty_0 dt e^{-t} t^{-(x+1)} = 2^{b-1}\rho^x \Gamma (-x). \nonumber \end{eqnarray} On the other hand, using $t > 1$, we get \begin{eqnarray} I_2 & = &\int^\infty_1 \!\!\!\!\!\! dt e^{-(\rho-1)t} e^{-t} t^{-(x+1)} (1+t)^{b+x-1} \leq e^{-(\rho-1)} \int^\infty_1 \! \!\!\!\!\! dt e^{-t} t^{-(x+1)} (1 +t)^{b+x-1} \nonumber \\ & \leq & e^{-(\rho-1)} \int^\infty_1 dt e^{-t} (1 + t)^{b-1} . \nonumber \end{eqnarray} Therefore \begin{equation} I_2 \leq e^{- (\rho-1)} \int^{\infty}_2\!\! ds e^{- s+1} s^{b-1} \leq e^{-(\rho-2)} \int^\infty_0 \!\!ds e^{-s} s^{b-1} = e^{-(\rho-2)} \Gamma (b)\leq e^{-(\rho-2)}b!. \end{equation} Thus we have \begin{equation} |\Gamma (-\lambda) U(-\lambda, b, \rho) | \leq 2^{b-1}\rho^x \Gamma(-x) + e^{-(\rho-2)} b! \label{f.7} \end{equation} for $ \rho > 1$ and $ -1 < x < 0 $, or \begin{equation} | U(-\lambda,b,\rho) | \leq 2^{b-1}\rho^x \left | {{\Gamma(-x)}\over {\Gamma(-\lambda)}}\right | + e^{-(\rho-2)} {{b!} \over {|\Gamma(-\lambda)|}}. \label{f.8} \end{equation} Suppose that (\ref{f.6}) is true for $ N-1 < x < N$. Then by using the recurrence relation (\ref{f.2}), we get for $ N < x < N+1$ \begin{eqnarray} | U(\!\!\!\!\!&-&\!\!\!\!\! \lambda ,b,\rho)|\nonumber \\ &\leq & \rho \Bigl\{2^{b+N}\rho^{(x-1)} (b + N + |y|+1)^N \left | {{\Gamma(-x+1)}\over {\Gamma(-\lambda +1)}}\right | \nonumber \\ & & \ \ \ + e^{-(\rho-2)} (\rho+|y|+1)^N {{(b+N+1)!} \over {|\Gamma(N -\lambda+ 1)|}}\Bigr\}\nonumber \\ & & \ \ \ + |b +\lambda-1| \Bigl \{2^{b + N -1}\rho^{(x-1)}(b + N+|y|)^N \left | {{\Gamma(-x+1)}\over {\Gamma(-\lambda +1)}}\right | \nonumber \\ & & \ \ \ \ \ + e^{-(\rho-2)} (\rho+1+|y|)^N {{(b+N)!} \over {|\Gamma(N-\lambda+ 1)|}}\Bigr \}.\nonumber \end{eqnarray} The identity $\Gamma(x+1)=x\Gamma(x) $ gives $$ \left |{{\Gamma(-x+1)}\over{\Gamma(-\lambda+1)}}\right| = \left |{{x}\over{\lambda}}\right| \left |{{\Gamma(-x)}\over{\Gamma(-\lambda)}}\right| \leq \left |{{\Gamma(-x)}\over {\Gamma(-\lambda)}}\right|. $$ Therefore \begin{eqnarray} | U(-\lambda,b,\rho)|\!\! &\leq &\!\! 2^{b+N}\rho^x \left\{(b+ N + |y|+1)^N + {{|b+\lambda -1|} \over {2\rho}} (b + N+|y|)^N\right\} \left | {{\Gamma(-x)}\over {\Gamma(-\lambda)}}\right | \nonumber \\ & & \ \ \ \ \ \ \ \ \ \ + e^{-(\rho-2)} (\rho+|y|+1)^N {{(b+N+1)!} \over {|\Gamma(N-\lambda+1)|}} \left \{ \rho +{{|b+\lambda -1|}\over {b+N+1}}\right\} \nonumber \\ & \leq & \!\! 2^{b + N}(b + N + |y| +1)^N \rho^x\left\{1 + {{|b+\lambda-1|} \over {2\rho}} \right\} \left | {{\Gamma(-x)}\over {\Gamma(-\lambda)}}\right | \nonumber \\ & & \ \ \ \ \ \ \ \ \ \ + e^{-(\rho-2)} (\rho+|y|+1)^{N+1} {{(b+N+1)!} \over {\Gamma(N-\lambda+1)}}. \nonumber \end{eqnarray} Therefore since $2 \leq b + N + |y|+1$ and $|b+\lambda-1 |\leq b + N + |y|$ we get the required bound $$ | U (-\lambda, b, \rho) | \leq 2^{b+ N}\rho^x (b + N + |y|+1)^{N+1} \left | {{\Gamma(-x)}\over {\Gamma(-\lambda)}}\right | + e^{-(\rho-2)} (\rho +1)^{N+1} {{(b+N+1)!} \over {\Gamma(N-\lambda+1)}}. $$ This gives the following bound on the Green's function for $2 \kappa | z - z' |^2 \geq 1$ and $N - 1 < x < N $ and $|y|\leq 1$ \begin{eqnarray} |\!\!\!\!\!\!\! & & \!\!\!\!\!\!\! G^\lambda_0(z,z') | \leq {{2 \kappa} \over \pi} e^{- \kappa | z - z' |^2} | \Gamma (- \lambda) | \nonumber \\ &\times & \Bigl \{(2 \kappa)^x (2(2 + N))^N | z - z' |^{2x} \left | {{\Gamma(-x)}\over {\Gamma(-\lambda)}}\right | \nonumber \\ & & \ \ \ \ \ + e^2(2 + 2 \kappa | z - z' |^2)^N {{(N+ 1)!} \over {|\Gamma (N - \lambda)|}} e^{- 2 \kappa | z - z' |^2}\Bigr \}. \label{f.9} \end{eqnarray} From (\ref{f.9}) we get for $N - 1 < x < N $ with $N\in \hbox{\BB N}_0$ \begin{equation} | G^\lambda_0(z,z') | \leq(C\kappa)^{N+1} N^N | \Gamma (-x)| | z - z' |^{2N} e^{- \kappa | z - z' |^2} \end{equation} since $| \Gamma(-\lambda)|\leq | \Gamma(-x)|$ and $ \Gamma (N - \lambda)$ is bounded below. \par We shall prove, again by induction, that for $N-1 < x <N$, $N\in \hbox{\BB N}_0$, $b\in \hbox{\BB N} $ and $\rho\leq 1$ \begin{equation} |U(-\lambda,b,\rho)| \leq 2^{N+4} (b+N)! \left | {{\Gamma (-x)}\over {\Gamma (-\lambda)}}\right | {{(1 + | \ln \rho |)} \over {\rho^{b-1}}}. \label{f.13} \end{equation} We first prove (\ref{f.13}) for $N=0$. From (\ref{f.4}) we have for $\rho \leq 1$, and $x < 0$, \begin{equation} |\Gamma(-\lambda) U(-\lambda, 1, \rho)| \leq {1 \over {|x|}} + {1 \over e} + | \ln \rho |. \end{equation} Thus \begin{equation} |U(-\lambda, 1 ,\rho)| \leq \left | {{\Gamma (-x)}\over {\Gamma (-\lambda)}}\right | \left \{{1 \over {{|x|} \Gamma (-x)}} + {1 \over {e \Gamma (-x)}} + {1 \over {\Gamma (-x)}} | \ln \rho |\right \}. \end{equation} Since $ -1 < x < 0$, $\Gamma (-x) > 1$ and $|x| \Gamma (-x) = \Gamma (-x+1)> (e-1)/e $, this gives \begin{eqnarray} \label{f.14} |U(-\lambda,1, \rho)| &\leq & \left | {{\Gamma (-x)}\over {\Gamma (-\lambda)}}\right | \left \{ {{e^2+e-1}\over {e^2-e}} + | \ln \rho | \right \}\\ \nonumber &\leq & (2 + | \ln \rho |) \left | {{\Gamma (-x)}\over {\Gamma (-\lambda)}}\right| \leq 2^4\left | {{\Gamma (-x)}\over {\Gamma (-\lambda)}}\right | (1 + | \ln \rho |). \end{eqnarray} \par Now we take $b \geq 2$. \begin{eqnarray} |\Gamma (-\lambda) U(-\lambda,b,\rho) | & \leq &\int^{\infty}_0 dt e^{-\rho t} t^{-(x+1)} (1 +t)^{b+x-1}\nonumber \\ & = & {1 \over {\rho^{b-1}}} \int^{\infty}_0 dt e^{-t} t^{-(x+1)} (\rho +t)^{b+x-1}\nonumber \\ & \leq &{1 \over {\rho^{b-1}}} \int^{\infty}_0 dt e^{-t} t^{-(x+1)} (1 + t)^{b+x-1}\nonumber \end{eqnarray} since $b +x- 1 \geq 0$. Thus we have for $b \geq 2$, and $-1 < x < 0$ \begin{equation} |\Gamma (-\lambda) U(-\lambda,b,\rho)| \leq {1 \over {\rho^{b-1}}} \Gamma (-x) U(-x, b, 1). \end{equation} By inserting the bound obtained from (\ref {f.7}) by letting $\rho$ tend to $1$, \begin{equation} \Gamma (-x)U(-x, b, 1) \leq 2^{b-1}\Gamma (-x) + e b!, \end{equation} into this inequality we get \begin{equation} |U(-\lambda,b,\rho)| \leq {1 \over {\rho^{b-1}}} \left\{2^{b-1} + {{e b!}\over{\Gamma (-x)}} \right\}\left | {{\Gamma (-x)}\over {\Gamma (-\lambda)}}\right | \leq 2^4 b! \left | {{\Gamma (-x)}\over {\Gamma (-\lambda)}}\right | {{(1 + | \ln \rho |)} \over {\rho^{b-1}}}. \end{equation} We can combine this with the inequality (\ref{f.14}) to get for $b\in \hbox{\BB N}$, $-1 < \lambda < 0$ and $ \rho < 1$, \begin{equation} |U(-\lambda,b,\rho)| \leq 2^4 b! \left | {{\Gamma (-x)}\over {\Gamma (-\lambda)}}\right | {{(1 + | \ln \rho |)} \over {\rho^{b-1}}} \end{equation} Using the recurrence relation (\ref{f.2}) and the induction hypothesis we get for $N< \lambda < N+1$, \begin{eqnarray} |U(-\lambda,b,\rho)| &\leq &\rho \left\{\leq 2^{N+4} (b+N+1)! \left | {{\Gamma (-x+1)}\over {\Gamma (-\lambda+1)}}\right | {{(1 + | \ln \rho |)} \over {\rho^b}} \right\}\nonumber \\ & & \ \ \ \ \ \ \ \ \ \ \ \ \ \ + |b+\lambda -1|2^{N+4} (b+N)! \left | {{\Gamma (-x+1)}\over {\Gamma (-\lambda+1)}}\right | {{(1 + | \ln \rho |)} \over {\rho^{b-1}}}\nonumber \\ &\leq & 2^{N+4} (b+N)! \left | {{\Gamma (-x+1)}\over {\Gamma (-\lambda+1)}}\right | {{(1 + | \ln \rho |)} \over {\rho^{b-1}}} (b+N+1+|b+\lambda -1|)\nonumber \\ &\leq & 2^{N+5} (b+N+1)! \left | {{\Gamma (-x)}\over {\Gamma (-\lambda)}}\right | {{(1 + | \ln \rho |)} \over {\rho^{b-1}}} \end{eqnarray} In particular with $b = 1$, (\ref{f.13}) gives for $N-1 < x < N $ and $\rho \leq 1$ \begin{equation} |U(-\lambda, 1, \rho) | \leq 2^{N+4} (N+1)! \left | {{\Gamma (-x)}\over {\Gamma (-\lambda)}}\right | (1 + | \ln \rho |). \end{equation} This gives us the required bound on the Green's function for $ N - 1 < x < N$, and $2 \kappa | z - z' |^2 \leq 1$ \begin{eqnarray} | G^\lambda_0 (z,z') | & \leq & {{2 \kappa} \over \pi} |\Gamma (- x)| 2^{N+4} (N+1)! (1 + | \ln (2\kappa| z - z' |^2) |) e^{- \kappa | z - z' |^2}\nonumber \\ & \leq & \kappa C^N N^N | \Gamma (- x)| (1 + | \ln (2\kappa | z - z' | ^2)|) e^{- \kappa | z - z' |^2}. \end{eqnarray} \par \section{Appendix B. Regularity of $\mu$} {\bf Definition} {\sl A probability measure $\mu$ on $\hbox{\BB R}$ is said to be $\tau$-regular, with $\tau\in (0,1]$, if there exists $\nu >0$ and $C<\infty$ such that \begin{equation} \mu([x - \delta, \ x + \delta]) \leq C \delta^\tau \mu([x - \nu, x+\nu]) \end{equation} for all $x \in \hbox{\BB R}$ and $0 < \delta < 1$.} \par\noindent Note that it is equivalent to requiring that there exists $\nu >0$ and $C<\infty$ such that \begin{equation} \mu([x - \delta, \ x + \delta]) \leq C \delta^\tau \mu([x - \nu, x+\nu]) \label{G.2} \end{equation} for all $x \in \hbox{\BB R}$ and $0 < \delta < \nu$. We shall prove this with $\tau=1$. Recall that the probability measure $\mu_0$ has support is an interval $[-a,a]$ with $a <\infty $. $\mu_0$ is symmetric about the origin and that its density $\rho_0$ is differentiable on $(-a,a)$ and satisfies the following condition \begin{equation} A \equiv \sup_{\zeta\in (0,a)} \frac{\rho_0'(\zeta)} {\rho_0(\zeta)}< \infty. \end{equation} If $B\subset \hbox{\BB R}$, $\mu(B)=\mu_0(\{\omega\ :\ 1/\omega \in B\})$ and the density of $\mu$, $\rho$, is given by \begin{equation} \rho (x)=\frac{\rho_0(1/x)}{x^2}. \end{equation} Since in our case $\mu $ is symmetric about the origin, it is sufficient to prove (\ref{G.2}) for $x\geq 0$. Also it is easy to see that the following condition is sufficient for (\ref{G.2}) with $\tau=1$. \par\noindent {\sl There exists $\nu >0$ and $C<0$ such that \begin{equation} \rho (x+t')\leq C\rho (x+t) \label{G.3} \end{equation} for all $x\in R_{+}$, $-\nu \leq t' \leq t \leq \nu $.} \par\noindent Then \begin{eqnarray} \mu ([x-\delta ,x+\delta ]) & = & \frac \delta \nu \int_{-\nu }^\nu \rho (x+\frac \delta \nu t)dt \nonumber \\ & \leq & \frac{\delta (C+1)}\nu \int_0^\nu \rho (x+\frac \delta \nu t)dt \nonumber \\ & \leq & \frac{\delta (C+1)C}\nu \int_0^\nu \rho (x+t)dt \nonumber \\ & = & \frac{\delta (C+1)C}\nu \mu ([x,x+\nu ]) \end{eqnarray} Let $b=1/a$. If $0\leq x\leq b-t'$, then $\rho (x+t')=0$. If $x>b-t'$, then \begin{equation} \ln\rho_0(1/(x+t'))-\ln\rho_0(1/(x+t)) =\frac {t-t'}{(x+t)(x+t')} \frac{\rho'_0(\zeta)}{\rho_0(\zeta)}, \end{equation} where $\zeta\in (1/(x+t),1/(x+t'))$. Thus \begin{equation} \ln\rho_0(1/(x+t'))-\ln\rho_0(1/(x+t))\leq \max(0,\frac {2A\nu}{b^2}). \end{equation} Therefore, with $C'=\exp(\max(0,\frac {2A\nu}{b^2}))$, \begin{equation} \rho_0(1/(x+t'))\leq C'\rho_0(1/(x+t)). \end{equation} But \begin{equation} \left(\frac{x+t}{x+t'}\right)^2 \leq\left(\frac{b+2\nu}b\right)^2. \end{equation} Thus the inequality (\ref{G.3}) is satisfied with $C= C'\left(\frac{b+2\nu}b\right)^2$. Finally note that \begin{equation} \int |x|^\epsilon \mu(dx)= \int |x|^{-\epsilon}\rho_0(x)dx< \infty \end{equation} for all $\epsilon<1$ since $\rho_0$ is continuous at the origin. \newpage
\section{Introduction} Two-particle correlations provide information about the space-time structure and dynamics of the emitting source \cite{boal}. Considering the correlations that occur in relativistic heavy-ion collisions one usually assumes that: (i) the particles are emitted independently (or the source is completely chaotic), and (ii) finite multiplicity corrections can be neglected. Then the correlations reflect a) the effects from symmetrization (antisymmetrization) of the wave function and b) the effects that are generated by the final-state interactions of the detected particles between themselves and with the source. At first sight one can regard the final-state interactions (FSI) as a contamination of the `pure' particle correlations. It should be pointed out, however, that the FSI depend on the structure of the emitting source and thus provide information about the source dynamics as well. In fact, this was proved by intensive investigations during the last twenty years of two-particle and source-particle FSI \cite{boal,GKW} (for recent publications see, for example, on two-particle FSI, \cite{anch98,anch95} and references therein; on source-particle FSI, \cite{barz96,barz97} and references therein). Actually, the former approaches accounting for FSI deal with the secondaries in empty post-freeze-out space. Meanwhile, in recent SPS experiments, for instance ${\rm Pb+Pb}$ at $160\ {\rm GeV}\times A$, some 800--900 secondary charged pions are created which form obviously a plasma-like post freeze-out medium. Therefore, one might expect that the FSI of two separate pions, at this collision energy, and of course at the energies of the forthcoming RHIC and LHC colliders, would be strongly influenced by the environment formed by other particles. The goal of the present paper is to estimate how large are the consequences on the Coulomb final-state interactions due to the presence of a large number of secondary charged particles. To be as independent as possible from source models, we chose the estimation of the Gamov factor as a standard quantity, which serves as a measure of the Coulomb FSI. The fundamental observable for intensity interferometry in hadron physics is the relative momentum spectrum of identical particles. For two like-sign pions, the modifications to the spectrum by the final-state Coulomb interaction result in a correction that has typically been considered as tractable with high accuracy. This assumption was based on the significantly different length scales between strong ($\propto 1/m_\pi $) and Coulomb ($\propto 1/m_\pi \alpha$) interactions \cite{Sakharov,bf} (here $m_\pi$ is the pion mass and $\alpha $ represents the fine structure constant). The correction may then be treated on the basis of the Schr\"{o}dinger equation, resulting in the well-known Gamov factor $G({\bf q})$ \begin{equation} G(|{\bf q}|)= \mid \psi_{\bf q} ({\bf r}=0)\mid ^2= \frac{2\pi \eta }{e^{2\pi \eta }-1} \ , \label{eq10} \end{equation} where $\psi_{\bf q} ({\bf r})$ is the two-particle wave function, $\eta = \alpha m_{\pi }/|{\bf q}|$, and ${\bf q}$ is the relative momentum of the particles. The nominal quantity expressing the correlation function in terms of experimental distributions \cite{boal} is \begin{equation} C({\bf k}_1,{\bf k}_2)= \frac{\displaystyle P_2\left({\bf k}_1, {\bf k}_2\right) } {\displaystyle P_1\left({\bf k}_1\right) \, P_1\left({\bf k}_2\right) } \ , \label{eq1} \end{equation} \noindent where $P_1\left({\bf k}\right) =E\, d^3N /d^3k$ and $P_2\left({\bf k}_1, {\bf k}_2\right) = E_1 \, E_2 \ d^6N /(d^3k_1d^3k_2)$ are single- and two-particle cross-sections. For point-like emitters it can be expressed (due to the factorization of the corresponding matrix element) in terms of a product of the Gamov factor to the model correlations \cite{GKW,anch98}: \begin{equation} C({\bf k}_1,{\bf k}_2)=G\left(|{\bf k}_1-{\bf k}_2|\right) \, C_{\rm model}({\bf k}_1,{\bf k}_2) \ . \label{eq6.0} \end{equation} It should be mentioned that, in the centre of mass of the pair the relative pion momentum $|{\bf q}|=2|{\bf k}|$, with ${\bf k}={\bf k}_1=-{\bf k}_2$, coincides with invariant relative momentum $q_{\rm inv}\equiv [(k_{1}+k_{2})^{2}-4m_{\pi }^{2}]^{1/2}$, where $k_{1}$ and $k_2$ are the pion four-momenta in an arbitrary frame. We consider corrections due to the Coulomb final-state interactions in a common scheme in which we do not take into account the finite-size of the fireball (for a consistent treatment of the source finite-size effects, see for example \cite{anch98}). This means that we calculate the correction factor in accordance with the formula $G_{\rm corr}(|{\bf q}|)=\mid \psi _{\bf q}({\bf r}=0)\mid ^2$, where now $\psi _{\bf q}(\bf r)$ is obtained from a numerical solution of the Schr\"odinger equation with the two-particle Coulomb potential {\it distorted by a multipion environment}. Since, as shown in \cite{anch98}, the finite size of the emission source softens the manifestation of the FSI, the `Gamov factor' tends to overestimate the FSI effects and therefore it is the most sensitive quantity for deviations from the standard two-particle Coulomb interaction. \section{Static multiparticle environment} In the high multiplicity case, when a post-freeze-out multipion environment cannot be neglected, the relation between the two-particle electromagnetic potential $\phi ({\bf r})$ and the local charge density is given by \begin{equation} \nabla^2 \phi ({\bf r})=-4\pi e (n^{(+)}-n^{(-)})\ , \label{15} \end{equation} where $e=\sqrt{\alpha }$ is the elementary charge; the density of charged pions $n^{(\pm )}$ is related to that of neutral pions $n^{(0)}$ via a Boltzmann factor: \begin{equation} n^{(\pm )}=n^{(0)}\, \exp{\left(\mp \frac{{\it e}\phi }{T_{\rm f}}\right)} \ , \label{16} \end{equation} The density $n^{(0)}$ of $\pi ^{0}$-mesons at the freeze-out temperature $T_{\rm f}$ coincides with the equilibrium density of charged pions in the absence of Coulomb interactions (we consider symmetrical nuclear matter). In the limit $ e\phi \ll T_{\rm f}$, Eq.(\ref{16}) can be rewritten as \begin{equation} n^{(\pm )}=n^{(0)}\, \left( 1 \mp \frac{{\it e}\phi }{T_{\rm f}}\right)\ \label{17} \end{equation} (this requires that the pions are not closer than $\sim10^{-2}$~fm to one another at $T_{\rm f}\approx 200\ {\rm MeV}$), so that \begin{equation} \nabla^2 \phi ({\bf r})=\frac{4\pi e^2}{T_{\rm f}} ( 2n^{(0)} ) \phi ({\bf r})\ . \label{18} \end{equation} The solution of this equation is well known and given by a screened Coulomb potential \begin{equation} \phi _{\pi ^{\pm }}(r)=\pm {\it e}\frac{e^{-r/R_{\rm scr}}}{r}\ , \label{19} \end{equation} where \begin{equation} \frac{1}{R_{\rm scr}} = \sqrt{\frac{8\pi }{3} \alpha } \cdot \sqrt{\frac{n_{\pi }}{T_{\rm f}} } \label{20} \end{equation} with the total pion density $n_{\pi }=3\, n^{(0)}$. Thus, for like-sign pions the potential energy reads \begin{equation} U_{\pi \pi }(r)=\frac{\alpha }{r} \exp{\left( -\frac{r}{R_{\rm scr}} \right) } \ . \label{20a} \end{equation} To evaluate the correction factor we use the screened Coulomb potential (\ref{20a}) to solve the Schr\"{o}dinger equation numerically, for the two choises: 1) $n_{\pi}=0.25\, {\rm fm}^{-3}$ and 2) $n_{\pi}=0.03\, {\rm fm}^{-3}$. (From now on we shall quote these two cases as ``{\it LHC}'' and ``{\it SPS}'' freeze-out conditions, respectively.) Taking $T_{\rm f}=190\ {\rm MeV}$, for example, we obtain from Eq.~(\ref{20}) screening radii ~$R_{\rm scr} \approx 7.9$ fm ({\it LHC}), ~$R_{\rm scr} \approx 22.4$ fm ({\it SPS}), respectively. For completeness of illustration we also consider the intermediate case $R_{\rm scr}\approx 19.3\ {\rm fm}$. The results of these calculations are plotted in Fig.~1 together with the standard Gamov factor. We see that a substantial correction to the standard Gamov factor would be required even for the existing experimental data if one adopts the idealized picture of a uniform post-freeze-out density of the environment. It is interesting to point out that the correction factor $G_{\rm cor}(q)$, which was evaluated for the same screened potential using the quasi-classical approximation \cite{anch96}, does not make a big difference with the correction factor in Fig.~1. Actually, what we really learned from the consideration of this idealized scenario is that a cancellation of the Coulomb potential tail (by screening) results in an increase of the correction factor in the region of small relative momentum ($q \le 50~\rm{MeV}$), as can be seen from Fig.~1. Thus the long-distance behaviour of the potential is responsible for the dramatic deviation of the correction factor from the standard Gamov factor at small relative momenta. On the other hand, the tail of the two-particle potential gives the main contribution to interactions when the particles are at large distances from one another, where in turn the density of secondary particles is small in the realistic picture. To keep this qualitative speculation as a thread we next turn to a more realistic calculation, explicitly incorporating expansion. \section{Expansion scenario} In the previous scenario the whole position space was filled by particles with a constant density; this is certainly an unrealistic approximation (idealization). In order to take into account post-freeze-out expansion of the pion system, we parametrise the pion density as \begin{equation} n(R)=n_{\rm f}\frac{R_{\rm f}^2}{R^2}\ , \label{24} \end{equation} where $n_{\rm f}$ is the freeze-out pion density and $R_{\rm f}$ is the freeze-out radius. Indeed, the spatial volume of the expanding pion system in the solid angle $\Omega $ increases as $\Delta V=\Omega \cdot R^2 \cdot \Delta R$, where $R$ is the distance from the centre of the fireball and $\Delta R$ is the thickness of the layer, which we keep constant. Then, if the number of particles $\Delta N$ in this volume is constant, the density reads: $n(R)=\Delta N/(\Omega \cdot R^2 \cdot \Delta R)={\rm const}/R^2$. Thus, the model (\ref{24}) implies that all particles have the same modulus of radial velocity. As we shall see further, such a model still means an overestimation of the particle density. To support our assumption (\ref{24}) let us consider a classical pion phase-space distribution. After freeze-out it satisfies the collisionless Boltzmann kinetic equation \begin{equation} \frac{\partial f(x,p)}{\partial x_0} + {\bf v}\cdot \nabla f(x,p) =0 \ , \label{1} \end{equation} where ${\bf v}={\bf p}/p_0$ is the velocity of the particle and $p_0=\omega ({\bf p}) \equiv \sqrt{m^2_{\pi }+{\bf p}^2} $. We look for the expansion solution of this equation, which can be fixed by asymptotic condition, for instance $ \lim_{t \to \infty} f(t, {\bf x}=0;{\bf p})=0$. A solution of this type can be written in the form \begin{equation} f(t,{\bf R},{\bf p}) = f_{0}({\bf R}-{\bf v}t,{\bf p})\ , \label{104} \end{equation} where $f_0({\bf R},{\bf p})$ is the initial distribution ($t=0$). For the sake of simplicity, we take an isotropic initial distribution in position and momentum spaces: just before freeze-out the particles were distributed (a) in accordance with Boltzmann's law in momentum space, and (b) in accordance with a Gaussian distribution in position space. The classical kinetic equation is quite sufficient to describe the {\it collective} (!) behaviour of the pion system after freeze-out. Hence, we assume that the system, at time $t=0$, occupies the phase space according to the distribution function \begin{equation} f_0({\bf R},{\bf p}) = n_0({\bf R}) g_0({\bf p}) \ , \label{101} \end{equation} where \begin{equation} n_0({\bf R}) = \frac{N_{\pi }}{\left( 2\pi R_{\rm f}^2 \right) ^{3/2}} \exp{\left( -\frac{R^2}{2R_{\rm f}^2} \right)} \label{102} \end{equation} with $\int d^3 R\, n_0 ({\bf R}) = N_{\pi }$, and \begin{equation} g_0 ({\bf p}) = \frac{\displaystyle 2\pi ^2} {\displaystyle m^2 _{\pi } T_{\rm f} K_2\left( \frac{m}{T_{\rm f}} \right)} \exp{\left( -\frac{\sqrt{m^2_{\pi }+{\bf p}^2}}{T_{\rm f}} \right)} \label{103} \end{equation} with $\int [d^3 p/(2\pi )^3]\, g_0({\bf p}) = 1$. Here $N_{\pi }$ is the total number of pions, $T_{\rm f}$ and $R_{\rm f}$ are the temperature and the mean radius of the system at time $t=0$ (freeze-out), respectively, $m_{\pi }$ is the pion mass and $K_2$ is a Bessel function of imaginary argument. The spatial distribution of the particles at time $t$ is determined by integrating the distribution function (\ref{104}) over the momentum variable \begin{equation} n(t,{\bf R}) = \int \frac{d^3 p}{(2\pi )^3 } \, n_0 \! \left({\bf R}- \frac{\small {\bf p}}{\small \omega ({\bf p})} t \right) g_0 ({\bf p}) \ . \label{105} \end{equation} Because of the spherical symmetry, it is reasonable to look at the radial density of pions \begin{equation} n_{\rm sph} (t,R) \equiv 4\pi R^2 n(t,R) \, . \label{106} \end{equation} This quantity may be treated as the number of pions in the shell with unit thickness at time $t$ and at a distance $R=|{\bf R}|$ from the fireball centre. Hence, $n_{\rm sph} (t,R)$ is a one-dimensional spatial distribution function and, evidently, the area under this curve at any time is equal to the particle number $N_{\pi }$, because of its normalization $\int _0^{\infty }dR n_{\rm sph} (t,R) =N_{\pi }$. We evaluate this function in accordance with Eq.~(\ref{105}) for {\it SPS} freeze-out conditions: $n_{\rm f }=0.03\, {\rm fm}^{-1/3}$, $T_{\rm f} = 190~{\rm MeV}$ and $R_{\rm f}=7.1~{\rm fm}$. The results of this calculation at different times $t$ are given in Fig.~2. It shows that the spherical distribution is always almost Gaussian-like and the velocity of the distribution maximum is very close to the velocity of light. The horizontal line (see Fig.~2) denotes a constant spherical density $4\pi R^2 n(R)={\rm const}$. This line is nothing more than the 3-dimensional spatial density $n(R)={\rm const'}/R^2$ (for the {\it SPS} freeze-out conditions ${\rm const}' \approx 85/4\pi $). It follows that, at any time $t$, for the post-freeze-out particle density $n(t,R)$ the upper limit is valid \begin{equation} n(t,R) \le n_{\rm f} \frac{R_{\rm f}^2}{R^2} \ , \label{106a} \end{equation} where $n_{\rm f}={\rm max}\{ n(t=0,R)\} $ and the equality is reached for an expanding system where the radial particle velocities are equal. Because of momentum dispersion in the post-freeze-out pion system, the pion density (\ref{24}) is an {\it essential overestimation} of the real density, which is formed by the comoving multipion environment. Hence, adopting the stationary density dependence (\ref{24}) we overestimate the influence of the environment, and consequently overestimate a distortion of the Coulomb FSI in the expansion scenario. On the other hand, this approach provides us with a possibility to consider a stationary post-freeze-out environment even in the frame of the non-stationary expansion scenario (see Fig.~2). Adopting the parametrization (\ref{24}), we get Eq.~(\ref{18}), where the pion density $n_{\pi }$ now depends on $R$: \begin{equation} \nabla ^2 \phi ({\bf r})= \frac{8\pi \alpha }{3\, T_{\rm f}} n(R)\, \phi ({\bf r})\ , \label{51} \end{equation} where we put $n^{(0)} \approx n_{\pi }/3$ as before ($n_{\rm f}=n_\pi$). Inserting the pair centre of mass distance $R$ together with the distance $r$ between two detected particles, we have, in the classical approach: \begin{equation} R \approx R_{\rm f} + v_{\rm cm}\cdot t \ , \hspace{1.5cm} r \approx v_{\rm rel}\cdot t \ , \label{26} \end{equation} where $v_{\rm cm}$ is the velocity of the two-particle centre of mass in the fireball rest frame and $v_{\rm rel}$ is the relative velocity of the particles ($v_{\rm rel}=q/m$, $t=0$ is fixed on the freeze-out hyper-surface). Eliminating the time from the approximate equalities (\ref{26}) one has \begin{equation} R=R_{\rm f} + \frac{v_{\rm cm}}{v_{\rm rel}} \cdot r\ , \label{27} \end{equation} which means that we parametrize the time evolution by the mean distance $r$ between two detected particles. We can thus rewrite the dependence of the pion density on $r$ as \begin{equation} n\big( R(r) \big) =\left(\frac{v_{\rm rel}}{v_{\rm cm}} \right)^2 \frac{n_{\rm f }R_{\rm f}^{2}}{(r+\overline{r} )^2} \ , \label{27a} \end{equation} where we define the dynamical freeze-out radius \begin{equation} \overline{r}\equiv R_{\rm f} \frac{v_{\rm rel}}{v_{\rm cm}} \ . \label{27b} \end{equation} It is time to recapitulate what has been done. First, we eliminate the time dependence of the particle density, bounding it from above by a stationary non-uniform spatial distribution, in accordance with inequality (\ref{106a}). Then, we connect the mean distance $R$ of the pair c.m.s. from the fireball and the mean distance $r$ between pions, thus eliminating again the time dependence. This means that all time and $R$ dependences are now parametrized by the variable $r$. If the relative velocity of the separate pions is small, then the partial time derivative $\frac{\partial}{\partial t} =v_{\rm rel}\, \frac{\partial}{\partial r} $ can be neglected. Hence, the problem is approximately reduced to a stationary one. Solving Eq.~(\ref{51}) with particle density from (\ref{27a}), one obtains the two-particle potential energy $U(r)=e\, \phi (r)$, which will then be exploited in the stationary Schr\"odinger equation to find a wave function. This will be the further strategy of our estimations. So, the problem of a time-dependent (expansion scenario) is reduced to a stationary one by the price of somewhat overestimating the Coulomb corrections. Equation (\ref{51}) may be rewritten as \begin{equation} \nabla^2 \phi ({\bf r}) = \frac{c^2(q)}{(r+\overline{r} )^2 } \, \phi ({\bf r})\ , \label{52} \end{equation} where we note explicitly that when the particles are separated by large distances $r$ the density of the multiparticle environment goes down in accordance with (\ref{27a}). The quantity $c^{2}(q)$ is defined in the following way: \begin{equation} c^2 (q) = \frac{8\pi \alpha }{3} \frac{R_{\rm f}^{2}n_{\rm f }}{T_{\rm f}} \frac{v_{\rm rel}^2}{v_{\rm cm}^2}\ . \label{53} \end{equation} Fixing the screening radius at freeze-out \begin{equation} R_{\rm scr}^f= \sqrt{\frac{3T_{\rm f}}{8\pi \alpha n_{\rm f }} } \ , \label{53a} \end{equation} we have \begin{equation} c(q) = \frac{R_{\rm f}}{R_{\rm scr}^f} \frac{v_{\rm rel}(q)}{v_{\rm cm}} \ . \label{53b} \end{equation} In spherical coordinates, Eq. (\ref{52}) can be rewritten as \begin{equation} \frac{d^2\phi (r)}{dr^2} + \frac{2}{r} \frac{d\phi (r)}{dr} - \frac{c^2(q)}{(r+\overline{r} )^2 } \, \phi (r)\, =\, 0 \ . \label{54} \end{equation} This equation may be solved by \begin{equation} \phi (r)= \frac{e}{r} \left( \frac{\overline{r}}{r+\overline{r}} \right) ^b \ , \label{55} \end{equation} where $\phi (r)$ satisfies the boundary condition: $\phi (r) \to e/r$ when $r \to 0$. The exponent $b$ is a solution of the quadratic equation resulting from the substitution of the ansatz (\ref{55}) into Eq.~(\ref{54}) and takes the form (assuming proper asymptotic behaviour of the potential $\phi $): \begin{equation} b(q)=-\frac{1}{2} +\frac{1}{2}\sqrt{1+4c^{2}(q)} \, . \label{56a} \end{equation} One has $b \rightarrow 0$ when $n_{\rm f} \rightarrow 0$ and hence the potential $\phi (r)$ transforms into the Coulomb potential in this case. It is shown in Figs.~3a and 3b for {\it LHC} and {\it SPS} freeze-out conditions, respectively. In the interval of interest, $b$ increases with increasing relative velocity, and so do deviations from a pure Coulomb field. This intriguing behaviour is directly related to the ``Hubble-like'' expansion implied by Eq.~(\ref{27}): it becomes clear if we remember that the pion density decreases (hence $R_{\rm scr} \rightarrow \infty $) with increasing distance $R$. The expansion thus results in modifications to the Coulomb potential that are of power-law, not exponential, form, in contrast to the static result given by Eq.~(\ref{19}). The connection of the potential (\ref{55}) and screened one can be found in the following way. One may treat the corrected potential obtained in Eq.~(\ref{55}) as an effective charge distribution \begin{equation} e_{\rm eff}=e\left(\frac{\overline{r} }{r+\overline{r} } \right)^b \ , \label{58} \end{equation} which we are going to average. Equation (\ref{52}) can be rewritten as $ (\nabla^2 - \kappa ^{2})\phi ({\bf r})=0$, with $\kappa \equiv c(q)/(r+\overline{r} ) $. This is equivalent to an $r$-dependent screening radius, i.e. \begin{equation} R_{\rm scr}(r)=\frac{r+\overline{r} }{c(q)} \ . \label{25} \end{equation} As shown in Figs.~3, the deviation of the potential (\ref{55}) from the pure Coulomb form in the region of small relative pion momentum $q\le 30\ {\rm MeV}$, is small ($b \to 0$). Since, $\kappa \to 0$ when $v_{\rm rel} \ll 1$ (see Eq.~(\ref{53b})), we first ignore the $r$ dependence of $\kappa $ to obtain the solution for the electromagnetic $\pi \pi $ potential in the form of Eq.~(\ref{19}), then substitute the $r$-dependent $\kappa $ into Eq.~(\ref{19}), to find that \begin{equation} U_{\pi \pi } = \frac{\alpha e^{-c(q)r/(r+\overline{r} )}}{r} \label{59} \end{equation} no longer decreases exponentially with distance when $r$ is large enough $r\gg \overline{r} $. Instead, the numerator on the r.h.s. of Eq.~(\ref{59}) represents the averaged charge distribution (\ref{58}) squared (we are now considering the potential energy $U_{\pi\pi} $ rather than the electric potential $\phi$, hence the extra factor of charge leading the $\alpha$). It is clear from Eq.~(\ref{53b}) that if $v_{\rm rel}/v_{\rm cm} \ll 1$ ($R_{\rm f}$ and $R_{\rm scr}^{f}$ are of the same order for high multiplicities) the renormalized constant $\alpha _{\rm eff}=\alpha \exp{[-c(q)]} $ is close to the bare value of $\alpha $. Moreover, the same qualitative result comes from the $r$-behaviour of the screening radius (\ref{25}) when it approaches the asymptotic value $R_{\rm scr}=\infty $ (Coulomb law) with increasing $r$. The quantity $c(q)$ increases with relative pion momentum, leading to larger deviations from the Coulomb potential and agreement with the features of the potential (\ref{55}) (see discussion after Eq.~(\ref{56a})). \subsection{Evaluations} The numerical evaluations of the correction factor (Gamov factor) should be provided by the solution of the Schr\"{o}dinger equation with the potential \begin{equation} U_{\rm expan}(r) = \frac{\alpha }{r} \left( \frac{\overline{r} }{r+\overline{r} } \right) ^b \ , \label{60} \end{equation} where the exponent $b$ is momentum-dependent, in accordance with (\ref{56a}). Then, one can construct the correction factor $G_{\rm cor}(|{\bf q}|)=\mid \psi_{\bf q}({\bf r}=0)\mid ^2$. To compare it with the standard Gamov factor $G_0$ we calculate the ratio $G_0/G_{\rm cor}$ for the LHC freeze-out conditions. The results are depicted in Fig.~4 as dotted curves for three different values of the pair mean momentum $P_{\rm cm}$. We now correct the two-particle potential (\ref{60}) for small distances $r\leq a\equiv n_{\rm f}^{-1/3}$, where $n_{\rm f}$ is the freeze-out pion density, i.e. in the region where the distance $r$ between two pions is smaller than the mean distance between the particles in the gas. Indeed, there is no `screening' effect for this distance, and hence the potential should be the Coulomb one. There is no such a problem for the statically screened potential $U(r)=\alpha \exp{(-r/R_{\rm scr})} /r$, because the screening radius obeys the condition $a\ll R_{\rm scr}$ (by definition in the sphere $4\pi R_{\rm scr}^3/3$ the number of particles is much larger than 1); for the region $r\leq a\ll R_{\rm scr}$, the potential $U(r)$ automatically transforms into the Coulomb one, $U(r)=\alpha /r$. In the expansion scenario we have in addition to $R_{\rm scr}$ another scale parameter, $\overline{r} =R_{\rm f}\left(v_{\rm rel}/v_{\rm cm}\right)$, where $R_{\rm f}$ is the freeze-out size of the system ($R_{\rm f}=7-10~{\rm fm}$ for {\it SPS} and {\it LHC} freeze-out conditions). By construction the potential $U_{\rm expan}(r)$ approaches the Coulomb one if (i) $r \ll \overline{r} $ (see (\ref{60})). On the other hand, the scale parameter $\overline{r} $ may be smaller than the mean distance between particles in the gas (ii) $\overline{r} < a$. In Fig.~5 we depict $\overline{r}$ for different mean momenta of the pair: $P_{\rm cm}=50,~150,~450~\rm{MeV/c}$ and for the mean distance between particles at freeze-out $ a= n_{\rm f}^{-1/3}$ (horizontal lines). For instance, as is seen in Fig.~5, for $P_{\rm cm}=450~\rm{MeV/c}$ and relative momentum $q<60~\rm{MeV/c}$, we have $\overline{r}<a$ for {\it SPS} freeze-out conditions. Hence, combining (i) and (ii) we get that the asymptotic regime, i.e. the Coulomb potential, can be achieved only for $r \ll a$. But we know that for separation distances which obey $r < a$, all distortions of the Coulomb potential vanish. Thus, the behaviour of the potential (\ref{60}) must be corrected, since it should be a Coulomb one at distances that are smaller than the mean distance between particles. To improve the behaviour of the potential for distances smaller than $a$, we use a smooth `potential switcher' \begin{equation} s(x)= \frac{1}{2} - \frac{1}{2} {\rm tanh}\, [2(x-2)] \ , \label{21c} \end{equation} which is depicted in Fig.~6. Then, incorporating the correct behaviour of the potential at small distances, we finally obtain the potential energy in the following form \begin{equation} U_{\pi \pi }(r) = s\left(\frac{r}{a} \right) \, \frac{\alpha }{r} + \left[ 1-s\left( \frac{r}{a} \right) \right] \, U_{\rm expan}(r) \ . \label{59b} \end{equation} Certainly, with increasing time the mean distance between particles increases, and the Coulomb potential is thus switched on at distances larger than the freeze-out particle mean distance $n_{\rm f}^{-1/3}$. But we restrict our calculation to the freeze-out mean distance. It means that in the competition between the two potentials $U_{\rm Coulomb}$ and $U_{\rm expan}$ we overestimate the contribution of the distorted potential $U_{\rm expan}$ and hence we overestimate the influence of the multiparticle environment. The results of the calculation with potential (\ref{59b}) are shown for the {\it LHC} freeze-out conditions as solid curves in Fig.~4. The same ratio $G_0/G_{\rm cor}$ for the {\it SPS} freeze-out conditions is given in Fig.~7. The correction factor we obtained reveals only small deviations from the standard Gamov factor. This result can be explained by a fast decrease of the density of secondary particles with increasing distance of the pair from the fireball, which in turn results in a very small distortion of the two-particle Coulomb potential. It is instructive to point out that evaluations that were made by the authors in \cite{anch96} on the basis of a quasi-classical approximation are in good qualitative agreement with the present results. \section{Summary and conclusions} In our first approach (static scenario) it was assumed that the whole position space is filled by secondary pions with constant density. This uniform environment of secondary pions results in a screened two-pion Coulomb potential. We showed that for future LHC and RHIC experiments the screening radius of the Coulomb interaction at the freeze-out density is of a size comparable with the source, and therefore the factorization of Eq.~(\ref{eq6.0}) \cite{GKW} is no longer valid. Moreover, solutions of the Schr\"odinger equation produce a correction factor $G_{\rm cor}=|\psi_{\bf q}({\bf r}=0)|^2$ which noticeably deviates from the standard Gamov factor (see Fig.~1). However, as we showed further, this model is a quite unrealistic approximation and is not relevant to the real picture of an expanding pion system after freeze-out. The conclusions reached with the first model change drastically after passaging to a more realistic model of an expanding pion system. In the second model, we first reduce the time evolution of the multipion post-freeze-out environment to a stationary one, parametrizing the density of the secondary pions $n_{\pi}(t,R)$ for all times $t$ as $n(R)= {\rm const}/R^2$, where $R$ is the distance from the fireball. The `constant' is determinded by the particular freeze-out conditions, namely it should be normalized on the real pion density $n_{\rm f}$ at the time of freeze-out. This parametrization results from the inequality $n_{\pi}(t,R)\le {\rm const}/R^2$, where equality is reached for an expanding system in which the radial particle velocities are all equal. Quite obviously, owing to the particle velocity dispersion, this parametrization is an overestimation of the real post-freeze-out pion density at any time $t$. Consequently, adopting this model of the pair environment, we overestimate the deviation from the Coulomb potential and thus overestimate the deviation from the Gamov factor. A further reduction results from the time evolution of the mean radii: $R(t)$, which is the distance of the separate pion pair c.m.s. from the fireball centre, and $r(t)$, which is a classical distance between these pions. Exploiting that reduction, we parametrize the relative motion, as well as the pair c.m.s. motion, by one variable $r$, thus eliminating the time dependence. After these reductions only one independent variable $r$ is left. It should be pointed out that the time derivatives can also be included into our consideration. From the $r$-parametrization, we obtain a proportionality of the time derivative to the pion--pion relative velocity $v_{\rm rel}$, namely $\frac{\partial}{\partial t} =v_{\rm rel}\, \frac{\partial}{\partial r} $. Since we are interested only in the small relative pion momenta this derivative can be neglected for a first-order estimate. So, we reduce the problem to a stationary one where, in contrast to the first scenario, the pion pair moves after freeze-out in a non-uniform environment. Effectively, this means that we consider the problem in the two-pion rest frame, where the relative radial motion of the particles that create the environment is slow enough with respect to the radial expansion. Practically it allows us to consider the stationary Schr\"{o}dinger equation instead of the time-dependent one. We show that the main contribution to the behaviour of the correction factor comes from the behaviour of the potential at the large distances that separate interacting particles. On the other hand, the interacting particles are separated by large distances when they are far enough from the fireball (this statement is a basis of the $r$-parametrization). Hence, because of a very fast decrease of the density with respect to the distance from the fireball $R$, namely $n(R)= {\rm const}/R^2$, the density of secondaries is small or even negligible at these distances. At this stage of the evolution, the long-range part of the potential is just the Coulomb one and it is responsible for the behaviour of the correction factor at small relative momentum $q\le 50~{\rm MeV/c} $. That is why the correction factor for this region of relative pion momentum practically coincides with the Gamov one. The short-range behaviour of the potential (when $r$ is smaller than a mean distance between the particles that form the environment), which is the Coulomb one for point-like pions, provides also only small deviations of the correction factor $G_{\rm cor}$ from the standard Gamov factor $G_0$ for a large pion relative momentum $q\ge $50--70~ MeV/c, as can be seen in Figs.~4 and 7. For the high ``{\it LHC}'' freeze-out pion density $n_{\rm f }=0.25\, {\rm fm}^{-1/3}$ the reduction of the correction factor $G_{\rm cor}$, as seen in Fig.~4, to the standard Gamov factor $G_0$ increases with decreasing $v_{\rm rel}/v_{\rm cm}$, the ratio of the relative velocity of the detected pions to their c.m. velocity in the fireball rest frame. When this parameter is much less than unity, the pion pair promptly escapes the initial high-density region so that slow relative motion of the two pions takes place in an approximately empty spatial region and the distortion of the mutual Coulomb potential is weak. This is not the case for ``{\it SPS}'' freeze-out conditions, as seen in Fig.~7. For the comparetively low pion density $n_{\rm f }=0.03\, {\rm fm}^{-1/3}$ the effect of small relative velocity is not pronounced because from the very beginning of the freeze-out the pions are not in such a dense environment and even small c.m. velocities are quite sufficient to bring the pair promptly into regions where the influence of the environment is negligible. \bigskip {\bf Acknowledgements} D.A. acknowledges helpful discussions with G.~Baym, D.~Ferenc, V.~Zasenko and would like to express special thanks to U.~Heinz, C.~Slotta and J.~Sollfrank who read the manuscript and made a number of remarks to improve the presentation. \bigskip
\section{INTRODUCTION} Galaxy formation by gravitational condensation out of the expanding cosmological background universe is affected by the complex interplay of gravitational dynamics of collisionless dark matter and the gas dynamics of the baryon-electron fluid, including the feedback on the latter which results when energy is released by the stars and AGN's that form. Results are presented here of 3D numerical gas dynamical simulations of the effect of this energy release, utilizing our anisotropic Adaptive Smoothed Particle Hydrodynamics (``ASPH'') method. Current computational limitations make it virtually impossible for numerical simulations of the initial value problem involving a realistic initial spectrum of small-amplitude, Gaussian random noise primordial density fluctuations to resolve the full range of length and mass scales necessary to form galaxies and the stars within them at the same time. We choose, instead, to focus here on an idealized model of structure formation which we can hope to resolve more accurately, thereby elucidating some of the most important aspects of the problem which may be relevant, as well, to more realistic initial conditions. We are particularly interested in studying the effects of nonspherical geometry and continuous cosmological infall on the problem of explosion-driven blow-out from galaxies during their formation. It is well-known that structure formation from Gaussian random noise proceeds in a highly anisotropic way, favoring pancake and filament formation over the formation of quasi-spherical objects. Our previous work (\cite{ref1},\cite{ref2}) has demonstrated, however, that a cosmological pancake, modeled as the nonlinear outcome of a single plane-wave density fluctuation, is subject to a linear gravitational instability which results in the formation of filaments and lumps in the central plane of the pancake. The lumps of collisionless dark matter that form in this way are quasi-spherical and develop a universal density profile which is reminiscent of the universal profiles found from N-body simulations of 3D Gaussian noise density fluctuations in hierarchical clustering models like the CDM model. As such, they provide an ideal test-bed for exploring the gas dynamics of structure formation and feedback effects without the troublesome complexity of Gaussian random noise initial conditions. The pancake problem, moreover, is completely scale-free, once all lengths are expressed in units of the pancake wavelength $\lambda_p$, time is expressed in terms of the cosmic scale factor $a$, divided by the scale factor $a_c$ at which the unperturbed pancake collapses to form density caustics in the dark matter and shocks in the gas, and the energy release is expressed in units of the total energy contained in a comoving cube of side $\lambda_p$. As such, each simulation of galaxy formation via 3D pancake gravitational instability serves to represent the generic behavior independent of the particular mass or collapse epoch of the object which forms. In order to preserve this universality and scale-free character of the problem, we will neglect radiative cooling in these first calculations. Once the key features of this scale-free problem are delineated, we will later consider the scale-dependent effects of radiative cooling and photoheating. Among the results we seek to quantify in this way are: \begin{itemize} \item The amount of energy release required to blow the gas out of a dark matter halo. \item The efficiency for ejecting the fraction of gas which is initially responsible for receiving the energy release (and, in the case of supernova explosions, the metallicity associated with the SN ejecta). \item The distinction between ``blowout,'' in which the energy release results in the escape of some energy and gas into the surrounding IGM but leaves the bulk of the gas in the object unaffected, and ``blowaway,'' in which most or all of the gas is ejected from the dark matter potential well. \item The energy release rate required to shock-heat the entire IGM by the overlapping effect of energy release from neighboring objects. \end{itemize} \section{PANCAKE INSTABILITY AND FRAGMENTATION AS A TEST-BED MODEL FOR GALAXY FORMATION} \subsection{Model and Initial Conditions} We consider an Einstein-de Sitter universe (density parameter $\Omega_0=1$, cosmological constant $\lambda_0=0$) with $\Omega_B=0.03$ and $\Omega_X=0.97$ (where $\Omega_B$ and $\Omega_X$ are the contributions of baryons and dark matter to $\Omega_0$, respectively). The initial conditions correspond to the growing mode of a single sinusoidal plane-wave density fluctuation of wavelength $\lambda_p$ and dimensionless wavevector ${\bf k}=\hat{\bf x}$ (length unit = $\lambda_p$). We adjust the initial amplitude $\delta_i$ such that a density caustic forms in the collisionless dark matter component at scale factor $a=a_c$. We perturb this system by adding to the initial conditions two transverse, plane-wave density fluctuations with equal wavelength $\lambda_s=\lambda_p$, wavenumbers ${\bf k}_s$ pointing along the orthogonal vectors $\hat{\bf y}$ and $\hat{\bf z}$, and amplitude $\epsilon_y\delta_i$ and $\epsilon_z\delta_i$, respectively, where $\epsilon_y\ll1$ and $\epsilon_z\ll1$. We use the notation $S_{1,\epsilon_y,\epsilon_z}$ to designate a pancake perturbed by two such transverse perturbation modes. All results presented here refer to the case $S_{1,0.2,0.2}$ unless otherwise noted. The presence of the two perturbation modes will result in the formation of two perpendicular filaments in the plane of the pancake, with a dense, quasi-spherical cluster at the intersection of the filaments. \subsection{Self-Similar Profiles for Dark Matter Halos} In 3D, our previous $\rm P^3M$ simulations of collisionless matter involving $64^3$ particles and $128^3$ grid cells in a comoving cubic box of side equal to $\lambda_p$, with gravitational softening parameter $\eta=0.3$~grid spacings demonstrated that the quasi-spherical lumps that form as one of the generic outcomes of pancake gravitational instability in 3D, evolve self-similar density profiles of universal shape (\cite{ref2},\cite{ref3}). The universal profile is well-approximated as a power-law $\rho\propto r^{-2.75}$ over a large range of density, which flattens somewhat at small radii, a shape which is independent of the details of the initial perturbations to the pancake. This self-similar profile is similar to the universal profile found to fit the results of 3D N-body simulations of the collisionless dark matter in a CDM model \cite{ref4}. This suggests that this 3D instability of cosmological pancakes which leads generically to the formation of such quasi-spherical dark matter halos may be used as an alternative to the details of the CDM model with its Gaussian random noise initial density fluctuations as a test-bed in which to study halo and galaxy formation further. To illustrate the universal density profile associated with this pancake instability, we show in Figure~1 results from \cite{ref2} and \cite{ref3} for one particular case, $S_{1,0.2,0.2}$, for various values of the expansion factor, as well as a summary of results for several different perturbation modes, at $a/a_c=7$. \begin{figure} \centering \vspace{-5cm} \psfig{figure=gravity.ps,width=15cm} \caption{Top Panel: Density profiles (spherically averaged, in units of $\langle\rho\rangle$, the average background density) versus radius from halo center (in units of $\lambda_p$) for mode $S_{1,0.2,0.2}$ at $a/a_c=1$ (dot-short dash), 2 (long dash), 3 (short dash), 4.5 (dotted), and 7 (solid); Bottom Panel: Density profiles at $a/a_c=7$ for several different modes of pancake perturbation: $S_{1,1,1}$ (solid), $S_{1,0.5,0.5}$ (dotted), $S_{1,0.4,0.4}$ (short dash), $S_{1,0.3,0.3}$ (long dash), $S_{1,0.2,0.2}$ (dot-short dash), $S_{1,0.1,0.1}$ (dot-long dash), $S_{1,0.5,0.25}$ (short dash-long dash)(\cite{ref2},\cite{ref3})}\label{fig1} \end{figure} \section{THE EFFECT OF EXPLOSIVE ENERGY RELEASE ON GALAXY FORMATION: BLOW-AWAY AMIDST CONTINUOUS INFALL} Galaxy formation which leads to star formation can be affected by feedback when stars evolve to the point of supernova (SN) explosions and the resulting shock-heating and outward acceleration of interstellar and intergalactic gas. Previous attempts to model this effect have typically been along one of three lines, that which adopts a smooth initial gas distribution in a galaxy-like, fixed dark matter gravitational potential well (e.g. \cite{ref5}), that which considers a single, isolated, but evolving, density fluctuation (i.e. without merging, infall or the effects of external tidal forces) (e.g. \cite{ref6}), and that in which the galaxy forms by condensation out of Gaussian random noise primordial density fluctuations such as in the CDM model (e.g. \cite{ref7},\cite{ref8},\cite{ref9}, \cite{ref10},\cite{ref11},\cite{ref12}). In the first case, the computational ability to resolve shocks which propagate away from the sites of explosive energy release is generally greater, while the last is perhaps more realistic in terms of the initial and boundary conditions, but the resolving of shocks is still generally quite poor. We compromise here between these two limits by using the pancake instability problem as the model of galaxy formation in which to explore the feedback effect of the explosive release of energy by SNe inside a protogalaxy. This affords some of the benefit of greater ease of the first approach mentioned above in resolving the explosion-driven shocks which are the crucial element in blowing gas out and away. It also provides a self-consistent cosmological origin and boundary condition for a protogalaxy or cluster, including the important effect of anisotropic gravitational collapse and continuous infall. Sharing the initial conditions described above for the formation of a dark matter halo via 3D pancake instability mode $S_{1,0.2,0.2}$ is an additional component of baryon-electron gas. We model the explosive release of energy due to SNe in terms of a single impulsive explosion which may represent a starburst or the collective effect of multiple SNe. We initiate the explosion by waiting until the first gas particles at the center of our dark matter halo reach a density contrast relative to the average background density, $\rho/\langle\rho\rangle$, exceeding $10^3$, at which point we suddenly multiply the thermal energy of these particles by a factor $\chi$ and share some of the explosion energy smoothly amongst their nearest neighbor particles via ASPH kernel smoothing as well. This occurs at $a/a_c=2.06$. While we have performed a series of simulations for different values of $\chi$, we shall report here only the results for two limiting cases: $\chi=0$ (no explosion) and $\chi=10^3$ (blowaway regime). Simulations end at an expansion factor $a/a_c=3$. (Note: The proper, numerical prescription for depositing the energy of explosions in the interstellar gas surrounding the explosion site is a complicated question, dependent as it is on explosion details which are unresolvable by current numerical treatments. Some recent discussion of this question is contained in \cite{ref13} and \cite{ref14}, including the question of the fraction of the energy of a given SN explosion which ends up as kinetic energy of the SN remnant rather than as thermal energy. However, since the simulations described here are adiabatic and neglect radiative cooling, it is self-consistent for us to deposit the entire explosion energy initially as thermal energy.) All our gas dynamical simulations are based upon the new 3D version of our ASPH method (\cite{ref15},\cite{ref16},\cite{ref17}), coupled to a $\rm P^3M$ gravity solver. The simulations reported here use $32^3$ gas particles, $32^3$ dark matter particles (with unequal particle masses, $m_{\rm dark}/m_{\rm gas}=\Omega_X/\Omega_B=32.3$), and a $\rm P^3M$ grid of $64^3$ cells with softening length $\eta=0.3$ grid spacings. \section{RESULTS} In the absence of explosion, the simulation produces two orthogonal filaments within the pancake central plane at whose intersection is located a denser, quasi-spherical ball of gas which sits in the gravitational potential well of a dense, quasi-spherical dark matter halo like that in Figure 1. In the case with explosion, we found that away from the central object, the pancake and the filaments within it are hardly affected. However, gas has been blown out of the center and some exterior gas which was infalling along directions perpendicular to the pancake plane has been swept back out, as well, some as far as to the outer edge of the box. In Figure 2, we show a shaded contour plot of the gas temperature at $a/a_c=3$, in a plane perpendicular to the pancake and intersecting the center of the central cluster. The explosion is confined by the gas in the plane of the pancake (seen edge-on on this figure) outside the central object, with the hottest gas at the very center. The filaments are hardly affected, however, nor is the shocked pancake gas far from the central object. This edge-on view of the pancake reveals that a major blow-out has occurred in which the explosion, led by an outer shock, has propagated all the way to the edge of the box, half-way to the nearest neighbor pancake, and collided there with the explosion shock expanding away from the neighboring pancakes' central object and toward the pancake in this box. The temperature plot reveals multiple shocks interior to the explosion, especially along the symmetry axis of the blow-out. Although the central object in which the explosion took place was quasi-spherical, the existence of the pancake plane and of the filaments which intersect at the location of the central dark halo ensure that a highly anisotropic explosion results and serves to channel the energy and mass ejection outward along the symmetry axis. \begin{figure} \centering \vspace{12cm}\hspace{-3cm} \psfig{figure=XP8_022_z05T.ps,width=9cm} \vspace{-12cm} \caption{Dimensionless Gas Temperature [i.e. internal energy per gram in units of $(2401/4)H_0^2\lambda_p^2(a/a_c)^{-4}$] in the plane perpendicular to the pancake, a cross section view which intersects the center of the central cluster.}\label{fig2} \end{figure} Velocity arrows for the gas particles at $a/a_c=3$ are displayed in Figures 3 and 4. Figure~3 shows a thin slice of the computational volume which contains the pancake central plane (i.e. a top view looking down on the pancake central plane). Figure~4 shows a slice perpendicular to the pancake (that is, the same plane as in Figure~2). These show that outflow is restricted to the symmetry axis, while infall continues within the pancake plane and especially inward along the filaments. \begin{figure} \centering \psfig{figure=vel_yz.ps,width=15cm} \caption{Velocity Field at $a/a_c=3$ in the pancake central plane, for the case with explosion. Each arrow corresponds to a simulation gas particle in a thin slice containing the pancake symmetry plane. Top panel: Full image; Bottom panel: zoom of the central region}\label{fig3} \end{figure} \begin{figure} \centering \psfig{figure=vel_xy.ps,width=15cm} \caption{Velocity Field at $a/a_c=3$ in the plane perpendicular to the pancake, for the case with explosion. Each arrow corresponds to a simulation gas particle in a thin slice centered on the image plane shown in Figure 2. Top panel: Full image; Bottom panel: zoom of the central region}\label{fig4} \end{figure} The effect of the explosion in blowing gas away is illustrated by Figure 5, in which different particle groups are distinguished according to their fate with and without the explosion. In the top panel, the solid dots show the particles which were the original recipients of the explosion energy and, by implication, the metal-enriched SN ejecta, which were previously located at the very center of the central dark halo at $a/a_c=2.06$. Gas initially outside this core region which was within the halo defined according to a mean overdensity $\rho/\langle\rho\rangle\geq200$ but which did not receive the initial explosion energy or ejecta directly is shown as open dots. All of the gas originally inside the halo when the explosion occurred has been blown away by $a/a_c=3.0$. The lower panel of Figure 5 shows {\it all} those gas particles which were found to be inside the central halo with mean overdensity $\rho/\langle\rho\rangle\geq200$ at $a/a_c=3$ in the case with {\it no} explosion but which are {\it not} inside this overdensity at $a/a_c=3.0$ in the case {\it with} the explosion. \begin{figure} \centering \psfig{figure=Split.ps,width=15cm} \caption{Gas particle positions projected onto the image plane of Figure 2. Case with explosion, at $a/a_c=3$. Top panel: particles that were located in the central cluster when the explosion occurred. The filled symbols indicate the particles that were metal-enriched by the explosion itself. The line and circle indicate respectively the location of the pancake plane and the location of the cluster {\it in the absence of explosion}. The radius of the circle is that which enclosed matter with average density $\rho/\langle\rho\rangle\geq200$ inside cluster in absence of the explosion. Bottom panel: particle with density smaller than $200\langle\rho\rangle$ whose density would be larger than $200\langle\rho\rangle$ in the absence of explosion}\label{fig5} \end{figure} The effect of the explosion on the build-up of the gas mass of the ``galaxy'' by continuous infall is illustrated by the plot in Figure 6 of the collapsed gas fraction in the box, defined as the gas of overdensity $\rho/\langle\rho\rangle\geq200$, for the cases with and without the explosion. With no explosion, the central mass grows continuously from $a/a_c=2$ to $a/a_c=3$ to encompass 10\% of all the mass in the box. With the explosion, however, all the gas in the central halo is blown away shortly after the explosion occurs at $a/a_c=2.06$, but by $a/a_c=2.7$, unobstructed infall within the pancake plane and along the filaments starts to resupply the central halo with gas at a significant rate. \begin{figure} \centering \psfig{figure=coll.ps,width=9cm} \caption{Fraction of gas with overdensity $\rho/\langle\rho\rangle\geq200$ versus $a/a_c$}\label{fig6} \end{figure} \section{SUMMARY} \begin{itemize} \item We find that blow-out and blow-away are generically anisotropic events which channel energy and mass loss outward preferentially along the symmetry axis of the local pancake and away from the intersections of filaments in the pancake plane. \item This means that metal ejection from dwarf galaxies at high redshift due to explosive energy release is less likely to pollute the local filaments and pancake in which the dwarf galaxies reside and more likely to channel the metals away from those denser regions. \item Despite the complete blow-away of gas initially in the dark matter potential well of the ``galaxy'' by the large explosion simulated here, continuous infall is not completely halted in the directions away from the preferred direction of blow-out, so infall partially replenishes the gas which is blown-away. \end{itemize} \section*{Acknowledgments} This work was supported by NASA Grants NAG5-2785, NAG5-7363, and NAG5-7821, NSF Grants PHY-9800725 and ASC-9504046, and the High Performance Computing Facility, University of Texas. \section*{References}
\section{Introduction} It is now well established that middle-aged solar-type stars show variability on a wide range of time scales, including the intermediate time scales of $\sim 10^0 - 10^4$ years (Weiss 1990). The evidence for the latter comes from a variety of sources, including observational, historical and proxy records. Many solar-type stars seem to show cyclic types of behaviour in their mean magnetic fields (e.g. Weiss 1994, Wilson 1994), which in the case of the Sun have a period of nearly 22 years. Furthermore, the studies of the historical records of the annual mean sunspot data since 1607 AD show the occurrence of epochs of suppressed sunspot activity, such as the {\it Maunder minimum} (Eddy 1976, Foukal 1990, Wilson 1994, Ribes \& Nesme-Ribes 1993, Hoyt \& Schatten 1996). Further research, employing $^{14}C$ (Eddy 1980, Stuiver \& Quey 1980, Stuiver \& Braziunas 1988, 1989) and $^{10} B$ (Beer {\em et al.} 1990, 1994a,b, Weiss \& Tobias 1997) as proxy indicators, has provided strong evidence that the occurrence of such epochs of reduced activity (referred to as {\it grand minima}) has persisted in the past with similar time scales, albeit irregularly. These latter, seemingly irregular, variations are important for two reasons. Firstly, the absence of naturally occurring mechanisms in solar and stellar settings, with appropriate time scales (Gough, 1990), makes the explanation of such variations theoretically challenging. Secondly, the time scales of such variations makes them of potential significance in understanding the climatic variability on similar time scales (e.g.\ Friis-Christensen \& Lassen 1991, Beer {\em at al.} 1994b, Lean 1994, Stuiver, Grootes \& Braziunas 1995, O'Brien {\em et al.} 1995, Baliunas \& Soon 1995, Butler \& Johnston 1996, White {\em et al.} 1997). In view of this, a great deal of effort has gone into trying to understand the mechanism(s) underlying such variations by employing a variety of approaches. Our aim here is to give a brief account of some recent results that may throw some new light on our understanding of such variations. \section{Theoretical frameworks} Theoretically there are essentially two frameworks within which such variabilities could be studied: stochastic and deterministic. Here we mainly concentrate on the deterministic approach and recall that given the usual length and nature of the solar and stellar observational data, it is in practice difficult to distinguish between these two frameworks (Weiss 1990). Nevertheless, even if the stochastic features play a significant role in producing such variations, the deterministic components will still be present and are likely to play an important role. The original attempts at understanding such variabilities were made within the linear theoretical framework. An important example is that of linear mean-field dynamo models (Krause \& R\"adler 1980) which succeeded in reproducing the nearly 22 year cyclic behaviour. Unfortunately such linear models cannot easily and naturally\footnote{It is worth bearing in mind that one can always produce complicated looking behaviour within the linear framework, by combining many simpler behaviours. The crucial point is that in this case complexity in behaviour requires a complicated underlying mechanism. Furthermore, there are qualitative differences, in terms of spectra and other dynamical indicators, between complicated dynamical behaviours produced by linearly complex and nonlinearly chaotic systems.} account for the complicated, irregular looking solar and stellar variability. The developments in nonlinear dynamical systems theory, over the last few decades, have provided an alternative framework for understanding such variability. Within this nonlinear deterministic framework, irregularities of the grand minima type are probably best understood in terms of various types of dynamical intermittency, characterised by different statistics over different intervals of time. The idea that some type of dynamical intermittency may be responsible for understanding the Maunder minima type variability in the sunspot record goes back at least to the late 1970's (e.g. Tavakol 1978, Ruzmaikin 1981, Zeldovich {\em et al.} 1983, Weiss {\em et al.} 1984, Spiegel 1985, Feudel {\em et al.} 1994). We shall refer to the assumption that grand minima type variability in solar-type starts can be understood in terms of some type of dynamical intermittency as the {\it intermittency hypothesis}. To test this hypothesis one can proceed by adopting either a quantitative or a quantitative approach. \subsection{Quantitative approach} Given the complexity of the underlying equations, the most direct approach to the study of dynamo equations is numerical. Ideally one would like to start with the full 3--D dynamo models with the least number of simplifying assumptions and approximations. There have been a great deal of effort in this direction over the last two decades (e.g. Gilman 1983, Nordlund {\em et al.} 1992, Brandenburg {\em et al.} 1996, Tobias 1998). The difficulty of dealing with small scale turbulence has meant that a detailed fully self-consistent model is beyond the range of the computational resources currently available, although important attempts have been made to understand turbulent dynamos in stars (e.g. Cattaneo, Hughes \& Weiss 1991, Nordlund {\em et al.} 1992, Moss {\em et al.} 1995, Brandenburg {\em et al.} 1996, Cattaneo \& Hughes 1996) and accretion discs (e.g. Brandenburg {\em et al.} 1995, Hawley {\em et al.} 1996). Such studies have had to be restricted to the geometry of a Cartesian box, which in essence makes them local dynamos, whereas magnetic fields in astrophysical objects are observed to exhibit large scale structure, related to the shape of the object, and thus can only be captured fully by global dynamo models (Tobias 1998). Furthermore, despite great advancements in numerical capabilities, these models still involve approximations and parametrisations and are extremely expensive numerically, specially if the aim is to make a comprehensive search for possible ranges of dynamical modes of behaviours as a function of control parameters\footnote{Which at times would require extremely long runs to transcend transients.}. An alternative approach, which is much cheaper numerically, has been to employ mean-field dynamo models. Despite their idealised nature, these models reproduce some features of more complicated models and allow us to analyse certain global properties of magnetic fields in the Sun. For example, the dependence of various outcomes of these models (such as parity, time dependence, cycle period, etc.) on global properties, including boundary conditions, have been shown to be remarkably similar to those produced by full three-dimensional simulations of turbulent models (Brandenburg 1999a,b). This gives some motivation for using these models for our studies below. A number of attempts have recently been made to numerically study such models, or their truncations, to see whether they are capable of producing the grand minima type behaviours. There are a number of problems with these attempts. Firstly, the developments in dynamical systems theory over the last two decades have uncovered a number of theoretical mechanisms for intermittency, each with their dynamical and statistical signatures. Secondly, the simplifications and approximations involved in these models, make it difficult to decide whether a particular type of behaviour obtained in a specific model is in fact generic. And finally, the characterisation of such numerically obtained behaviours as ``intermittent'' is often phenomenological and based on simple observations of the resulting time series (e.g. Zeldovich {\em et al.} 1983, Jones {\em et al.} 1985, Schmalz \& Stix 1991, Feudel {\em et al.} 1993, Covas {\em et al.} 1997a,b,c, Tworkowski {\em et al.} 1998, and references therein), rather than a concrete dynamical understanding coupled with measurements of the predicted dynamical signatures and scalings. There are, however, examples where the presence of various forms of intermittency has been established concretely in such dynamo models, by using various signatures and scalings (Brooke 1997, (Covas \& Tavakol 1997, Covas {\em et al.} 1997c, Brooke {\em et al.} 1998, Covas \& Tavakol 1998, Covas {\em et al.} 1999b). \subsection{Qualitative approach} Given the inevitable approximations and simplifications involved in dynamo modelling (specially given the turbulent nature of the regimes underlying such dynamo behaviours and hence the parametrisations necessary for their modelling in practice), a great deal of effort has recently gone into the development of approaches that are in some sense generic. The main idea is to start with various qualitative features that are thought to be commonly present in such settings and then to study the generic dynamical consequences of such assumptions. Such attempts essentially fall into the following categories. Firstly, there are the low dimensional ODE models that are obtained using the Normal Form approach (Spiegel 1994, Tobias {\em et al.} 1995, Knobloch {\em et al.} 1996). These models are robust and have been successful in accounting for certain aspects of the dynamos, such as several types of amplitude modulation of the magnetic field energy, with potential relevance for solar variability of the Maunder minima type. The other approach is to single out the main generic ingredients of such models and to study their dynamical consequences. For axisymmetric dynamo models, these ingredients consist of the presence of invariant subspaces, non-normal parameters and non-skew property. The dynamics underlying such systems has recently been studied in (Covas {\em et al.}, 1997c,1999b; Ashwin {\em et al.} 1999). This has led to a number of novel phenomena, including a new type of intermittency, referred to as {\it in--out intermittency}, which we shall briefly discuss in section \ref{intermittencies} \section{Models} The standard mean-field dynamo equation is given by \begin{equation} \label{dynamo} \frac{\partial {\bf B}}{\partial t}= \nabla \times \left( {\bf u} \times {\bf B} + \alpha {\bf B} - \eta_t \nabla \times {\bf B} \right), \end{equation} where ${\bf B}$ and ${\bf u}$ are the mean magnetic field and mean velocity respectively and the turbulent magnetic diffusivity $\eta_t$ and the coefficient $\alpha$ arise from the correlation of small scale turbulent velocities and magnetic fields (Krause \& R\"adler, 1980). In axisymmetric geometry, eq.\ (1) is solved by splitting the magnetic field into meridional and azimuthal components, ${\bf B} = {\bf B_{p}} + {\bf B_{\phi}}$, and expressing these components in terms of scalar field functions {${\bf B_{p}} = \nabla \times A \hat{\phi}$, ${\bf B}_{\phi} = B \hat{\phi}$}. In the following we shall also employ a family of truncations of the one dimensional version of equation (\ref{dynamo}), along with a time dependent form of $\alpha$, obtained by using a spectral expansions of the form: \begin{eqnarray} \frac{dA_n}{dt}&=&-n^2A_n+\frac{D}{2}(B_{n-1}+B_{n+1})+\nonumber \sum_{m=1}^{N}\sum_{l=1}^{N}{\cal F}(n,m,l)B_mC_l,\nonumber\\ \frac{dB_n}{dt}&=&-n^2B_n+\sum_{m=1}^{N}{\cal G}(n,m)A_m,\label{truncated} \\ \nonumber\frac{dC_n}{dt}&=&-\nu n^2 C_n -\sum_{m=1}^{N}\sum_{l=1}^{N}{\cal H}(n,m,l)A_mB_l. \end{eqnarray} where $A_n$, $B_n$ and $C_n$ are derived from the spectral expansion of the magnetic field ${\bf B}$ and $\alpha$ respectively, ${\cal F, H}$ and ${\cal G}$ are coefficients expressible in terms of $m,n$ and $l$, $N$ is the truncation order, $D$ is the dynamo number and $\nu$ is the Prandtl number (see Covas {\em et al.} 1997a,b,c for details). \section{Different forms of intermittency in ODE and PDE dynamo models} \label{intermittencies} Recent detailed studies of axisymmetric mean field dynamo models have produced concrete evidence for the presence of various forms of dynamical intermittency in such models. We shall give a brief overview of these results in this section. \begin{figure}[!htb] \centerline{\def\epsfsize#1#2{0.6#1} \epsffile{ND_r=0.2_Calph=25.5_f=0.0_p=0.0_Comega=-1.0e4_Angle=-0.7854_mesh=041x081_eps=0.01.eps}} \caption{\label{crisis1} Example of crisis induced intermittency in a shell dynamo with a cut, with $r_0=0.2$, $C_\alpha=25.5$, $C_\Omega=-10^4$, $\theta_0=45^{\circ}$. See Covas {\em et al.}, 1999a for details of the model.} \end{figure} \subsection{Crisis (or attractor merging) intermittency} A particular form of this type of intermittency, discovered by Grebogi, Ott \& Yorke (Grebogi {\em et al.} 1982, 1987), is the so called ``attractor merging crisis'', where as a system parameter is varied, two or more chaotic attractors merge to form a single attractor. There is both experimental and numerical evidence for this type of intermittency (see for example Ott (1993) and references therein). We have found concrete evidence for the presence of such a behaviour in a 6-dimensional truncation of mean-field dynamo model of the type (\ref{truncated}) (Covas \& Tavakol 1997) and more recently, in a PDE model of type (\ref{dynamo}) (see Covas \& Tavakol (1999) for details). Fig.\ \ref{crisis1} shows an example of the latter which clearly demonstrates the merging of two attractors, with different time averages for energy and parity. For a concrete characterisation and scaling, see Covas \& Tavakol (1999). \subsection{Type I-Intermittency} This form of intermittency, first discovered by Pomeau and Manneville in the early 1980's (Pomeau \& Manneville 1980), has been extensively studied analytically, numerically and experimentally (see Bussac \& Meunier 1982, Richter {\em et al.} 1994 and references therein). It is identified by long almost regular phases interspersed by (usually) shorter chaotic bursts. In particular, this type of intermittency has been found in a 12--D truncated dynamo model of type (\ref{truncated}) (Covas {\em et al.} 1997c), and more recently in a PDE dynamo model of type (\ref{dynamo}) (Covas \& Tavakol 1999). Fig.\ \ref{typeIa} gives an example of such time series, where the irregular interruptions of the laminar phases by chaotic bursts can easily be seen. For a concrete characterisation, including the scaling for the average length of laminar phases see Covas \& Tavakol (1999). \begin{figure}[!htb] \centerline{\def\epsfsize#1#2{0.6#1} \epsffile{ND_r=0.7_Calph=28.0_f=0.0_p=0.0_Comega=-1.0e4_Angle=-0.7854_mesh=041x081_eps=0.01.eps}} \caption{\label{typeIa} Example of Type-I intermittency in a shell dynamo with a cut, with $r_0=0.7$, $C_\alpha=28.0$, $C_\Omega=-10^4$, $\theta_0=45^{\circ}$. See Covas {\em et al.}, 1999a for details of the model. } \end{figure} \subsection{On-Off and In-Out Intermittency} An important feature of systems with symmetry (as in the case of solar and stellar dynamos) is the presence of invariant submanifolds. It may happen that attractors in such invariant submanifolds may become unstable in transverse directions. When this happens, one possible outcome could be that the trajectories can come arbitrarily close to this submanifold but also have intermittent large deviations from it. This form of intermittency is referred as {\it on-off intermittency} (Platt {\em et al.} 1993a,b). Examples of this type of intermittency have been found in dynamo models, both phenomenologically (Schmitt {\em et al.}, 1996) and concretely in truncated dynamo models of the type (\ref{truncated}) (Covas {\em et al.} 1997c). A generalisation of on-off intermittency, the in-out intermittency, discovered recently (Ashwin {\em et al.} 1999) is expected to be generic for axisymmetric dynamo settings. The crucial distinguishing feature of this type of intermittency is that, as opposed to on-off intermittency, there can be different invariant sets associated with the transverse attraction and repulsion to the invariant submanifold, which are not necessarily chaotic. This gives rise to identifiable signatures and scalings (Ashwin {\em et al.} 1999). Concrete evidence for the occurrence of this type of intermittency has been found recently in both PDE and truncated dynamo models of the types (\ref{dynamo}) and (\ref{truncated}) respectively (see Covas {\em et al.} (1999a,b) for details). \section{Intermittency hypothesis: theory and observation} In the previous section, we have summarised concrete evidence for the presence of four different types of dynamical intermittency in both truncated and PDE mean-field dynamo models. >From a theoretical point of view, the intermittency hypothesis may therefore be said to have been established, at least within this family of mean-field models. What remains to be seen is whether these types of intermittency still persist in more realistic models. An encouraging development in this connection is the discovery of a type of intermittency which is expected to occur generically in axisymmetric dynamo settings, independently of the details of specific models. Despite these developments, testing the intermittency hypothesis poses a number of difficulties in practice: \begin{enumerate} \item Observationally, all precise dynamical characterisation of solar and stellar variability are constrained by the length and the quality of the available observational data. This is particularly true of the intermediate (and of course longer) time scale variations. Such a characterisation is further constrained by the fact that some of the indicators of such mechanisms, such as scalings, require very long and high quality data. \item Theoretically, there is now a large number of such mechanisms, some of which share similar signatures and scalings, which could potentially complicate the process of differentiation between the different mechanisms. \item An important feature of real dynamo settings is the inevitable presence of noise. This calls for a theoretical and numerical study of effects of noise on the dynamics, on the one hand (e.g. Meinel \& Brandenburg 1990, Moss {\em et al.} 1992, Ossendrijver \& Hoyng 1996, Ossendrijver, Hoyng \& Schmitt 1996) and on the signatures and scalings of various mechanisms of intermittency on the other. \end{enumerate} These issues raise a number of interesting questions. Is, for example, the intermittency hypothesis operationally decidable at present? Will it be operationally decidable in foreseeable future? In this connection it is worth bearing in mind that some types of intermittency do possess signatures that are rather easily identifiable. Nevertheless, we believe the answer to these difficult questions can only be realistically contemplated once a more clear picture has emerged of all the possible types of intermittency that can occur in more realistic solar-type dynamo models (and ultimately real dynamos) and once their precise signatures and scalings, in presence of noise, have been identified. \acknowledgments We would like to thank the organisers of this meeting for their kind hospitality and for bringing about the opportunity for many fruitful exchanges. We would also like to thank Peter Ashwin, Axel Brandenburg, John Brooke, David Moss, Ilkka Tuominen and Andrew Tworkowski for the work we have done together and Edgar Knobloch, Steve Tobias, Alastair Rucklidge, Michael Proctor and Nigel Weiss for many stimulating discussions. \vspace{0.2cm} EC is supported by grant BD/5708/95 -- PRAXIS XXI, JNICT. EC thanks the Astronomy Unit at QMW for support to attend the conference. RT benefited from PPARC UK Grant No. L39094. This research also benefited from the EC Human Capital and Mobility (Networks) grant ``Late type stars: activity, magnetism, turbulence'' No. ERBCHRXCT940483.
\@startsection{section}{1}{\z@}{3.5ex plus 1ex minus .2ex{\@startsection{section}{1}{\z@}{3.5ex plus 1ex minus .2ex} {2.3ex plus .2ex}{\large\bf}} \def\@startsection{subsection}{2}{\z@}{2.3ex plus .2ex{\@startsection{subsection}{2}{\z@}{2.3ex plus .2ex} {2.3ex plus .2ex}{\bf}} \newcommand\Appendix[1]{\def\Alph{section}}{Appendix \Alph{section}} \@startsection{section}{1}{\z@}{3.5ex plus 1ex minus .2ex{\label{#1}}\def\Alph{section}}{\Alph{section}}} \begin{document} \begin{titlepage} \samepage{ \setcounter{page}{1} \rightline{ACT-2/99} \rightline{CTP-TAMU-12/99} \rightline{TPI-MINN-99/22} \rightline{UMN-TH-1760-99} \rightline{\tt hep-ph/9904301} \rightline{May 2000} \vfill \begin{center} {\Large \bf A Minimal Superstring Standard Model \{\bf I}: Flat Directions\\ } \vfill \vskip .4truecm \vfill {\large G.B. Cleaver,$^{1,2}$\footnote{<EMAIL>} A.E. Faraggi,$^{3}$\footnote{<EMAIL>} and D.V. Nanopoulos$^{1,2,4}$\footnote{<EMAIL>}} \\ \vspace{.12in} {\it $^{1}$ Center for Theoretical Physics, Dept.\ of Physics, Texas A\&M University,\\ College Station, TX 77843, USA\\} \vspace{.06in} {\it $^{2}$ Astro Particle Physics Group, Houston Advanced Research Center (HARC),\\ The Mitchell Campus, Woodlands, TX 77381, USA\\} \vspace{.06in} {\it$^{3}$ Department of Physics, University of Minnesota, Minneapolis, MN 55455, USA\\} \vspace{.025in} {\it$^{4}$ Academy of Athens, Chair of Theoretical Physics, Division of Natural Sciences,\\ 28 Panepistimiou Avenue, Athens 10679, Greece\\} \vspace{.025in} \end{center} \vfill \begin{abstract} Three family $SU(3)_C\times SU(2)_L\times U(1)_Y$ string models in several constructions generically possess two features: (i) an extra local anomalous $U(1)_A$ and (ii) numerous (often fractionally charged) exotic particles beyond those in the minimal supersymmetric model (MSSM). Recently, we demonstrated that the observable sector effective field theory of such a free fermionic string model can reduce to that of the MSSM, with the standard observable gauge group being just $SU(3)_C\times SU(2)_L\times U(1)_Y$ and the $SU(3)_C\times SU(2)_L\times U(1)_Y$--charged spectrum of the observable sector consisting solely of the MSSM spectrum. An example of a model with this property was shown. We continue our investigation of this model by presenting a large set of different flat directions of the same model that all produce the MSSM spectrum. Our results suggest that even after imposing the conditions for the decoupling of exotic states, there may remain sufficient freedom to satisfy the remaining phenomenological constraints imposed by the observed data. \end{abstract} \smallskip} \end{titlepage} \setcounter{footnote}{0} \def { } \def\begin{equation}{\begin{equation}} \def\end{equation}{\end{equation}} \def\begin{eqnarray}{\begin{eqnarray}} \def\end{eqnarray}{\end{eqnarray}} \def<{<} \def\eq#1{eq.\ (\ref{#1})} \def\dagger{\dagger} \def\qandq{\quad {\rm and} \quad} \def\qand{\quad {\rm and} } \def\andq{ {\rm and} \quad } \def\qwithq{\quad {\rm with} \quad} \def\qwith{ \quad {\rm with} } \def\withq{ {\rm with} \quad} \def\noindent {\noindent } \def\nonumber {\nonumber } \defi.e., {i.e., } \def{\it e.g.}{{\it e.g.}} \def\frac{1}{2}{\frac{1}{2}} \def\frac{1}{\sqrt{2}}{\frac{1}{\sqrt{2}}} \def\frac{1}{2}{{\textstyle{1\over 2}}} \def\frac{1}{3}{{\textstyle {1\over3}}} \def{\textstyle {1\over4}}{{\textstyle {1\over4}}} \def\frac{1}{6}{{\textstyle {1\over6}}} \def$\phantom{-}${{\tt -}} \def{\tt +}{{\tt +}} \def\phantom{+}{\phantom{+}} \def$Z_2\times Z_2$ {$Z_2\times Z_2$ } \def{\rm Tr}\, {{\rm Tr}\, } \def{\rm tr}\, {{\rm tr}\, } \defM_{P}{M_{P}} \def\bar{h}{\bar{h}} \def\slash#1{#1\hskip-6pt/\hskip6pt} \def\slash{k}{\slash{k}} \def\,{\rm GeV}{\,{\rm GeV}} \def\,{\rm TeV}{\,{\rm TeV}} \def$\ast${\,{\rm y}} \defStandard--Model {Standard--Model } \defsupersymmetry {supersymmetry } \defsupersymmetric standard model{supersymmetric standard model} \defminimal supersymmetric standard model{minimal supersymmetric standard model} \def $SU(3)_C\times SU(2)_L\times U(1)_Y$ { $SU(3)_C\times SU(2)_L\times U(1)_Y$ } \def\vev#1{\left\langle #1\right\rangle} \def\mvev#1{|\langle #1\rangle|^2} \def\epsilon{\epsilon} \defU(1)_{\rm A}{U(1)_{\rm A}} \defQ^{(\rm A)}{Q^{(\rm A)}} \def\mssm{SU(3)_C\times SU(2)_L\times U(1)_Y} \defKa\v c--Moody {Ka\v c--Moody } \def\langle{\langle} \def\rangle{\rangle} \def\o#1{\frac{1}{#1}} \def{\tilde H}{{\tilde H}} \def{\overline{\chi}}{{\overline{\chi}}} \def{\overline{q}}{{\overline{q}}} \def{\overline{\imath}}{{\overline{\imath}}} \def{\overline{\jmath}}{{\overline{\jmath}}} \def{\overline{H}}{{\overline{H}}} \def{\overline{Q}}{{\overline{Q}}} \def{\overline{a}}{{\overline{a}}} \def{\overline{\alpha}}{{\overline{\alpha}}} \def{\overline{\beta}}{{\overline{\beta}}} \def{ \tau_2 }{{ \tau_2 }} \def{ \vartheta_2 }{{ \vartheta_2 }} \def{ \vartheta_3 }{{ \vartheta_3 }} \def{ \vartheta_4 }{{ \vartheta_4 }} \def{\vartheta_2}{{\vartheta_2}} \def{\vartheta_3}{{\vartheta_3}} \def{\vartheta_4}{{\vartheta_4}} \def{\vartheta_i}{{\vartheta_i}} \def{\vartheta_j}{{\vartheta_j}} \def{\vartheta_k}{{\vartheta_k}} \def{\cal F}{{\cal F}} \def\smallmatrix#1#2#3#4{{ {{#1}~{#2}\choose{#3}~{#4}} }} \def{\alpha\beta}{{\alpha\beta}} \def{ (M^{-1}_\ab)_{ij} }{{ (M^{-1}_{\alpha\beta})_{ij} }} \def{(i)}{{(i)}} \def{\bf V}{{\bf V}} \def{\bf N}{{\bf N}} \def{\bf b}{{\bf b}} \def{\bf S}{{\bf S}} \def{\bf X}{{\bf X}} \def{\bf I}{{\bf I}} \def{\mathbf 1}{{\mathbf 1}} \def{\mathbf 0}{{\mathbf 0}} \def{\mathbf S}{{\mathbf S}} \def{\mathbf b}{{\mathbf b}} \def{\mathbf b}{{\mathbf b}} \def{\mathbf S}{{\mathbf S}} \def{\mathbf S}{{\mathbf S}} \def{\mathbf S}{{\mathbf S}} \def{\mathbf X}{{\mathbf X}} \def{\mathbf I}{{\mathbf I}} \def{\mathbf I}{{\mathbf I}} \def{\mathbf \alpha}{{\mathbf \alpha}} \def{\mathbf \beta}{{\mathbf \beta}} \def{\mathbf \gamma}{{\mathbf \gamma}} \def{\mathbf \xi}{{\mathbf \xi}} \def{\mathbf \alpha}{{\mathbf \alpha}} \def{\mathbf \beta}{{\mathbf \beta}} \def{\mathbf \gamma}{{\mathbf \gamma}} \def{\mathbf \xi}{{\mathbf \xi}} \def\overline{\Phi}{\overline{\Phi}} \def\epsilon{\epsilon} \def\t#1#2{{ \Theta\left\lbrack \matrix{ {#1}\cr {#2}\cr }\right\rbrack }} \def\C#1#2{{ C\left\lbrack \matrix{ {#1}\cr {#2}\cr }\right\rbrack }} \def\tp#1#2{{ \Theta'\left\lbrack \matrix{ {#1}\cr {#2}\cr }\right\rbrack }} \def\tpp#1#2{{ \Theta''\left\lbrack \matrix{ {#1}\cr {#2}\cr }\right\rbrack }} \def\langle{\langle} \def\rangle{\rangle} \def\f#1{$\Phi_{#1}$} \def\fb#1{$\overline{\Phi}_{#1}$} \def$.5$#1{$\Phi^{'}_{#1}$} \def\fbp#1{$\overline{\Phi}^{'}_{#1}$} \def\fpx#1{$\Phi^{(')}_{#1}$} \def\fbpx#1{$\overline{\Phi}^{(')}_{#1}$} \def\Hd#1{$H_{#1}$} \def\Vd#1{$V_{#1}$} \def\NC#1{$N^c_{#1}$} \def\mf#1{\Phi_{#1}} \def\mnfb#1{\overline{\Phi}_{#1}} \def\mnfp#1{\Phi^{'}_{#1}} \def\mfbp#1{\overline{\Phi}^{'}_{#1}} \def\mfpx#1{\Phi^{(')}_{#1}} \def\mfbpx#1{\overline{\Phi}^{(')}_{#1}} \def\mH#1{H_{#1}} \def\mV#1{V_{#1}} \def\mNC#1{N^c_{#1}} \def\Q#1{Q_{#1}} \def\dc#1{d^{c}_{#1}} \def\uc#1{u^{c}_{#1}} \def\h#1{h_{#1}} \def\bar{h}#1{{\bar{h}}_{#1}} \def\bh#1{{\bar{h}}_{#1}} \def\L#1{L_{#1}} \def\ec#1{e^{c}_{#1}} \def\Nc#1{N^{c}_{#1}} \def\H#1{H_{#1}} \def{\bf V}#1{V_{#1}} \def\Hs#1{H^{s}_{#1}} \def\Vs#1{V^{s}_{#1}} \def\sH#1{H^{s}_{#1}} \def\sV#1{V^{s}_{#1}} \def\P#1{\Phi_{#1}} \def{\tt +}#1{\Phi_{#1}} \def\pp#1{\Phi^{'}_{#1}} \def\pb#1{{\overline{\Phi}}_{#1}} \def\bp#1{{\overline{\Phi}}_{#1}} \def\pbp#1{{\overline{\Phi}}^{'}_{#1}} \def\ppb#1{{\overline{\Phi}}^{'}_{#1}} \def\bpp#1{{\overline{\Phi}}^{'}_{#1}} \def\bi#1{{\overline{\Phi}}^{'}_{#1}} \def{\rm observable}{{\rm observable}} \def{\rm singlets}{{\rm singlets}} \def{\rm hidden}{{\rm hidden}} \def{\rm mixed}{{\rm mixed}} \def\,\vrule height1.5ex width.4pt depth0pt{\,\vrule height1.5ex width.4pt depth0pt} \def\relax\hbox{$\inbar\kern-.3em{\rm C}$}{\relax\hbox{$\,\vrule height1.5ex width.4pt depth0pt\kern-.3em{\rm C}$}} \def\relax\hbox{$\inbar\kern-.3em{\rm Q}$}{\relax\hbox{$\,\vrule height1.5ex width.4pt depth0pt\kern-.3em{\rm Q}$}} \def\relax{\rm I\kern-.18em R}{\relax{\rm I\kern-.18em R}} \font\cmss=cmss10 \font\cmsss=cmss10 at 7pt \def\IZ{\relax\ifmmode\mathchoice {\hbox{\cmss Z\kern-.4em Z}}{\hbox{\cmss Z\kern-.4em Z}} {\lower.9pt\hbox{\cmsss Z\kern-.4em Z}} {\lower1.2pt\hbox{\cmsss Z\kern-.4em Z}}\else{\cmss Z\kern-.4em Z}\fi} \defA.E. Faraggi{A.E. Faraggi} \def\AP#1#2#3{{\it Ann.\ Phys.}\/ {\bf#1} (19#2) #3} \def\NPB#1#2#3{{\it Nucl.\ Phys.}\/ {\bf B#1} (19#2) #3} \def\NPBPS#1#2#3{{\it Nucl.\ Phys.}\/ {{\bf B} (Proc. Suppl.) {\bf #1}} (19#2) #3} \def\PLB#1#2#3{{\it Phys.\ Lett.}\/ {\bf B#1} (19#2) #3} \def\PRD#1#2#3{{\it Phys.\ Rev.}\/ {\bf D#1} (19#2) #3} \def\PRL#1#2#3{{\it Phys.\ Rev.\ Lett.}\/ {\bf #1} (19#2) #3} \def\PRT#1#2#3{{\it Phys.\ Rep.}\/ {\bf#1} (19#2) #3} \def\PTP#1#2#3{{\it Prog.\ Theo.\ Phys.}\/ {\bf#1} (19#2) #3} \def\MODA#1#2#3{{\it Mod.\ Phys.\ Lett.}\/ {\bf A#1} (#2) #3} \def\IJMP#1#2#3{{\it Int.\ J.\ Mod.\ Phys.}\/ {\bf A#1} (19#2) #3} \def\nuvc#1#2#3{{\it Nuovo Cimento}\/ {\bf #1A} (#2) #3} \def\RPP#1#2#3{{\it Rept.\ Prog.\ Phys.}\/ {\bf #1} (19#2) #3} \def{\it et al\/}{{\it et al\/}} \hyphenation{su-per-sym-met-ric non-su-per-sym-met-ric} \hyphenation{space-time-super-sym-met-ric} \hyphenation{mod-u-lar mod-u-lar--in-var-i-ant} \setcounter{footnote}{0} \@startsection{section}{1}{\z@}{3.5ex plus 1ex minus .2ex{Introduction} Deriving the Standard Model from heterotic string theory remains one of the strongly motivated endevours in theoretical physics. On the one hand, for over a quarter of a century now, the structure of the Standard model itself suggests the realization of grand unified structures, most appealing in the context of $SO(10)$ unification \cite{granda,grandb}. On the other hand supersymmetry, a key ingredient in grand and string unification, continues to be the only extension of the Standard Model still consistent with the experimental data, whereas theories with a low energy cutoff in general run into conflict with experiment once detailed models are considered. Finally, string theory remains the only known theoretical framework for the consistent unification of gravity and the gauge interactions. There are two main approaches regarding how to derive the Standard Model from string theory. The first approach asserts that we must first understand the nonperturbative formulation of string theory and then the true string vacuum will be uniquely revealed. The second approach argues that much can be learned about the string realization of the Standard Model by studying the features and properties of phenomenologically promising perturbative string models. Singling out such string models for study may be instrumental for learning about the nonperturbative dynamics of the theory. Following the second line of thought, a class of phenomenologically promising models has been identified. These models, constructed in the free fermionic formulation \cite{fff}, correspond to $Z_2\times Z_2$ orbifold models with nontrivial Wilson lines and background fields. Studying F--theory compactification on the $Z_2\times Z_2$ orbifold, which is related to the free fermionic models \cite{befnq}, indeed reveals new features that do not arise in other similar F--theory studies \cite{ftheory}. It is not unlikely that these new features will eventually prove to be important for understanding the issue of vacuum selection in string theory. Connecting string theory to low energy experimental data, parameterized by the Standard Model, remains vital. It is important to point out that of the semi--realistic orbifold \cite{orbi} and free fermionic models \cite{flipped,fny,eu,slm,so64,chl}, the NAHE--based $Z_2\times Z_2$ free fermionic models \cite{nahe,slm} (or more generally $Z_2\times Z_2$ orbifold models at the free fermionic point) are the only ones that naturally give rise to the $SO(10)$ unification structures. Thus, it seems of much value to continue improving our understanding of the more phenomenologically viable three generation $Z_2\times Z_2$ free fermionic models, as well as their embeddings in compactifications of M-- and F--theory. It was stressed in the past for a variety of phenomenological reasons that the canonical $SO(10)$ embedding of the weak--hypercharge in string models is highly preferred \cite{ibanez}. In fact, it was even suggested that this must be the case in the true string vacuum. We stress that this is precisely the embedding obtained in the $Z_2\times Z_2$ free fermionic models, in contrast to other quasi--realistic orbifold \cite{orbi} or free fermionic models \cite{chl}, which do not produce the standard $SO(10)$ embedding. The first phenomenological criterion that a string model must satisfy are three chiral generations with the standard $SO(10)$ embedding of the Standard Model gauge group. While the $Z_2\times Z_2$ free fermionic models naturally give rise to the $SO(10)$ unification structure, the $SO(10)$ symmetry has to be broken at the string level. Therefore, there is no explicit $SO(10)$ symmetry in the effective field theory level. Nevertheless, the $SO(10)$ symmetry is still reflected, for example, in some of the Yukawa coupling relations. In realistic heterotic string models, it is well known that modular invariance constraints impose that the string spectrum contains exotic fractionally charged states \cite{schellekens}. Such states indeed occur in all known string models built from level--one Ka\v c--Moody algebras, and may appear in the massive or massless sectors. In many of the more realistic examples such states arise in vector--like representations and therefore may obtain mass terms from cubic--level or higher order terms in the superpotential. It is clear that in the true string vacuum such fractionally charged states must be either confined \cite{flipped} or sufficiently massive \cite{fc,huet,nrt}. Thus, the next non--trivial phenomenological criteria on viable string models is that there should be no free fractionally charged states surviving to low energies. Furthermore, as the existence of fractionally charged states and other states beyond the minimal supersymmetric standard model will in general affect the unification of the the gauge couplings, an attractive scenario is that all the states beyond the MSSM decouple from the low energy spectrum at the string scale. Recently, we have demonstrated the existence of string models with the above properties \cite{cfn1}. Studying the flat directions in the string model of ref.\ \cite{fny} (referred to henceforth as the ``FNY model''), we showed the existence of one flat direction for which the massless spectrum below the string scale consists solely of the spectrum of the minimal supersymmetric standard model. Furthermore, it was found that in this particular model the $D$--flatness constraints necessarily impose that either of the $U(1)_{Z^\prime}$ or $U(1)_Y$ symmetries must be broken by the choices of flat directions. This is in fact an attractive situation in which in the phenomenologically viable case the surviving $SO(10)$ subgroup necessarily coincides with that of the Standard Model, and the $U(1)_{Z^\prime}$ is necessarily broken by the choices of flat directions. We remark that our model also contains, at the massless string level, a number of electroweak Higgs doublets and a color triplet/anti--triplet pair beyond the MSSM. We show that by the same suitable choices of flat directions that only one Higgs pair remains light below the string scale. The additional color triplet/anti--triplet pair receives mass from a fifth order superpotential term. This results in the triplet pair receiving a mass that is slightly below the string scale and is perhaps smaller than the doublet and fractional exotic masses by a factor of around $(1/10-1/100)$. We emphasize that the numerical estimate of the masses arising from the singlet VEVs should be regarded only as illustrative. The important result is the generation of mass terms for all the states beyond the MSSM, near the string scale. The actual masses of the extra fields may be spread around the $M_U$ scale, thus inducing small threshold corrections that are still expected to be compatible with the low energy experimental data. The string solution found in ref.\ \cite{cfn1} is the first known example of a minimal superstring derived standard model, in which, of all the $SU(3)_C\times SU(2)_L \times U(1)_Y$--charged states, only the MSSM spectrum remains light below the string scale. In this paper we expand the analysis of ref.\ \cite{cfn1}. The first obvious question is naturally whether the special solution found in ref.\ \cite{cfn1} is an isolated example or whether there exists an enlarged space of solutions giving rise solely to the MSSM spectrum below the string scale. We recall that in the FNY model of ref.\ \cite{fny} the condition for the decoupling of the exotic fractionally charged states from the massless spectrum is that a specific set of Standard Model singlet fields \cite{fc} acquire non--vanishing vacuum expectation values (VEVs) in the cancellation of the anomalous $U(1)_A$ Fayet--Iliopoulos (FI) $D$--term. Additionally we showed that the same set of VEVs produce mass terms that lead to the decoupling of the extra color triplets and electroweak doublets, beyond the MSSM. Thus, in any solution that incorporates those VEVs, the resulting spectrum below the string scale consists solely of the MSSM spectrum. Expanding the analysis of ref.\ \cite{cfn1} we show that there exists, in fact, an extended space of solutions which incorporate those VEVs. This is a very promising situation, for it demonstrates that there may still be sufficient freedom to allow the possibility of accommodating the various phenomenological constraints in one of these solutions. Our paper is organized as follows: Section 2 is a brief review of the FNY model \cite{fny}, including a discussion of the model's massless spectrum, prior to any states taking on VEVs. This is accompanied by Appendix A, which lists all string quantum numbers, including the non--gauge charges, of the massless states, and by Appendix B, which contains the complete renormalizable superpotential and all fourth through sixth order non--renormalizable superpotential terms. Section 3 reviews constraints on flat directions and presents a survey of viable flat directions that generate (near) string--scale mass to all $SU(3)_C\times SU(2)_L\times U(1)_Y$--charged MSSM exotic states in the FNY model, producing an effective MSSM field theory below the string scale. Accompanying Section 3 are Appendices C and D. Appendix C contains a SM--conserving basis set, of the ``maximally orthogonal'' class, for generating $D$--flat directions, while Appendix D contains the actual tables of classes of $D$-- and $F$--flat directions. These tables indicate the order to which $F$--flatness is retained, the respective superpotential terms that break $F$--flatness, the dimension of each flat direction, and the respective number of non--anomalous $U(1)_i$ that are broken by these directions. Section 4 then discusses some distinguishing features of the various flat directions. Phenomenological implications of these flat directions will be presented in \cite{cfn3}. \@startsection{section}{1}{\z@}{3.5ex plus 1ex minus .2ex{The FNY Model (A Review)} \@startsection{subsection}{2}{\z@}{2.3ex plus .2ex{Construction} The more realistic free fermionic models, which utilize the NAHE--set of boundary condition basis vectors, admit the $SO(10)$ embedding of the Standard Model gauge group. Aside from the phenomenological aspects, which motivate the need for the $SO(10)$ embedding, another advantage of utilizing the standard $SO(10)$ embedding and hence of the NAHE--set, is the decoupling \cite{decoup} of the exotic fractionally charged states from the massless spectrum. This should be contrasted with models, like the free fermionic models of \cite{chl}, which do not allow the $SO(10)$ embedding and contain exotic fractionally charged states which cannot decouple from the massless spectrum \cite{cceelw}. This distinction is an important one, as it severely limits the number of phenomenologically viable models, even among the three generation orbifold models that are traditionally viewed as semi--realistic. The more realistic NAHE--based free fermionic models represent one class of string models that still survives this requirement. For completeness we recall here the construction of the FNY model and its distinctive properties. The boundary condition basis vectors which generate the more realistic free fermionic models are, in general, divided into two major subsets. The first set consists of the NAHE set \cite{nahe,slm}, which is a set of five boundary condition basis vectors denoted $\{{\bf 1},{\mathbf S},{\mathbf b}_1,{\mathbf b}_2,{\mathbf b}_3\}$. With `${\mathbf 0}$' indicating Neveu--Schwarz boundary conditions and `${\mathbf 1}$' indicating Ramond boundary conditions, these vectors are as follows: \begin{eqnarray} &&\begin{tabular}{c|c|ccc|c|ccc|c} ~ & $\psi^\mu$ & $\chi^{12}$ & $\chi^{34}$ & $\chi^{56}$ & $\bar{\psi}^{1,...,5} $ & $\bar{\eta}^1 $& $\bar{\eta}^2 $& $\bar{\eta}^3 $& $\bar{\phi}^{1,...,8} $ \\ \hline \hline ${\mathbf 1}$ & 1 & 1&1&1 & 1,...,1 & 1 & 1 & 1 & 1,...,1 \\ ${\bf S}$ & 1 & 1&1&1 & 0,...,0 & 0 & 0 & 0 & 0,...,0 \\ \hline ${{\mathbf b}}_1$ & 1 & 1&0&0 & 1,...,1 & 1 & 0 & 0 & 0,...,0 \\ ${{\mathbf b}}_2$ & 1 & 0&1&0 & 1,...,1 & 0 & 1 & 0 & 0,...,0 \\ ${{\mathbf b}}_3$ & 1 & 0&0&1 & 1,...,1 & 0 & 0 & 1 & 0,...,0 \\ \end{tabular} \nonumber\\ ~ && ~ \nonumber\\ ~ && ~ \nonumber\\ &&\begin{tabular}{c|cc|cc|cc} ~& $y^{3,...,6}$ & $\bar{y}^{3,...,6}$ & $y^{1,2},\omega^{5,6}$ & $\bar{y}^{1,2},\bar{\omega}^{5,6}$ & $\omega^{1,...,4}$ & $\bar{\omega}^{1,...,4}$ \\ \hline \hline {${\mathbf 1}$} & 1,...,1 & 1,...,1 & 1,...,1 & 1,...,1 & 1,...,1 & 1,...,1 \\ ${\mathbf S}$ & 0,...,0 & 0,...,0 & 0,...,0 & 0,...,0 & 0,...,0 & 0,...,0 \\ \hline ${{\mathbf b}}_1$ & 1,...,1 & 1,...,1 & 0,...,0 & 0,...,0 & 0,...,0 & 0,...,0 \\ ${{\mathbf b}}_2$ & 0,...,0 & 0,...,0 & 1,...,1 & 1,...,1 & 0,...,0 & 0,...,0 \\ ${{\mathbf b}}_3$ & 0,...,0 & 0,...,0 & 0,...,0 & 0,...,0 & 1,...,1 & 1,...,1 \\ \end{tabular} \label{nahe} \end{eqnarray} with the following choice of phases which define how the generalized GSO projections are to be performed in each sector of the theory: \begin{equation} C\left( \matrix{{\mathbf b}_i\cr {\mathbf b}_j\cr}\right)~=~ C\left( \matrix{{\mathbf b}_i\cr {\mathbf S}\cr}\right) ~=~ C\left( \matrix{{\mathbf 1} \cr {\mathbf 1} \cr}\right) ~= ~ -1~. \label{nahephases} \end{equation} The remaining projection phases can be determined from those above through the self--consistency constraints. The precise rules governing the choices of such vectors and phases, as well as the procedures for generating the corresponding space--time particle spectrum, are given in refs.~\cite{fff}. After imposing the NAHE set, the resulting model has gauge group $SO(10)\times SO(6)^3\times E_8$ and $N=1$ space--time supersymmetry. The model contains 48 multiplets in the $16$ representation of $SO(10)$, 16 from each twisted sector ${\mathbf b}_1$, ${\mathbf b}_2$ and ${\mathbf b}_3$. In addition to the spin 2 multiplets and the space--time vector bosons, the untwisted sector produces six multiplets in the vectorial 10 representation of $SO(10)$ and a number of $SO(10)\times E_8$ singlets. As can be seen from Table (\ref{nahe}), the model at this stage possesses a cyclic permutation symmetry among the basis vectors ${\mathbf b}_1$, ${\mathbf b}_2$ and ${\mathbf b}_3$, which is also respected by the massless spectrum. The second stage in the construction of these NAHE--based free fermionic models consists of adding three additional basis vectors to the above NAHE set. These three additional basis vectors, which are often called $\lbrace {\mathbf \alpha},{\mathbf \beta}, {\mathbf \gamma}\rbrace$, correspond to ``Wilson lines'' in the orbifold construction. The allowed fermion boundary conditions in these additional basis vectors are of course also constrained by the string consistency constraints, and must preserve modular invariance and world--sheet supersymmetry. The choice of these additional basis vectors $\lbrace {\mathbf \alpha},{\mathbf \beta},{\mathbf \gamma}\rbrace$ nevertheless distinguishes between different models and determine their low--energy properties. For example, three additional vectors are needed to reduce the number of massless generations to three, one from each sector ${\mathbf b}_1$, ${\mathbf b}_2$, and ${\mathbf b}_3$, and the choice of their boundary conditions for the internal fermions ${\{y,\omega\vert{\bar y},{\bar\omega}\}^{1,\cdots,6}}$ also determines the Higgs doublet--triplet splitting and the Yukawa couplings. These low--energy phenomenological requirements therefore impose strong constraints \cite{slm} on the possible assignment of boundary conditions to the set of internal world--sheet fermions ${\{y,\omega\vert{\bar y},{\bar\omega}\}^{1,\cdots,6}}$. Consider the model in Table (\ref{fnymodel}) \begin{eqnarray} &\begin{tabular}{c|c|ccc|c|ccc|c} ~ & $\psi^\mu$ & $\chi^{12}$ & $\chi^{34}$ & $\chi^{56}$ & $\bar{\psi}^{1,...,5} $ & $\bar{\eta}^1 $& $\bar{\eta}^2 $& $\bar{\eta}^3 $& $\bar{\phi}^{1,...,8} $ \\ \hline \hline ${{\mathbf b}_4}$ & 1 & 1&0&0 & 1~1~1~1~1 & 1 & 0 & 0 & 0~0~0~0~0~0~0~0 \\ ${{\mathbf \beta}}$ & 1 & 0&0&1 & 1~1~1~0~0 & 1 & 0 & 1 & 1~1~1~1~0~0~0~0 \\ ${{\mathbf \gamma}}$ & 1 & 0&1&0 & ${1\over2}$~${1\over2}$~${1\over2}$~${1\over2}$~${1\over2}$ & ${1\over2}$ & ${1\over2}$ & ${1\over2}$ & ${1\over2}$~0~1~1~${1\over2}$~${1\over2}$~${1\over2}$~1 \\ \end{tabular} \nonumber\\ ~ & ~ \nonumber\\ ~ & ~ \nonumber\\ &\begin{tabular}{c|c|c|c} ~& $y^3{y}^6$ $y^4{\bar y}^4$ $y^5{\bar y}^5$ ${\bar y}^3{\bar y}^6$ & $y^1{\omega}^6$ $y^2{\bar y}^2$ $\omega^5{\bar\omega}^5$ ${\bar y}^1{\bar\omega}^6$ & $\omega^1{\omega}^3$ $\omega^2{\bar\omega}^2$ $\omega^4{\bar\omega}^4$ ${\bar\omega}^1{\bar\omega}^3$ \\ \hline \hline ${\mathbf b}_4$& 1 ~~~ 0 ~~~ 0 ~~~ 1 & 0 ~~~ 0 ~~~ 1 ~~~ 0 & 0 ~~~ 0 ~~~ 1 ~~~ 0 \\ ${\mathbf \beta}$ & 0 ~~~ 0 ~~~ 0 ~~~ 1 & 0 ~~~ 1 ~~~ 0 ~~~ 1 & 1 ~~~ 0 ~~~ 1 ~~~ 0 \\ ${\mathbf \gamma}$& 0 ~~~ 0 ~~~ 1 ~~~ 1 & 1 ~~~ 0 ~~~ 0 ~~~ 1 & 0 ~~~ 1 ~~~ 0 ~~~ 0 \\ \end{tabular} \label{fnymodel} \end{eqnarray} With the choice of generalized GSO coefficients: \begin{eqnarray} C\left(\matrix{{\mathbf b}_4\cr {\mathbf b}_j,{\mathbf \beta}\cr}\right)&&= -C\left(\matrix{{\mathbf b}_4\cr {{\mathbf 1}}\cr}\right)= -C\left(\matrix{{\mathbf \beta}\cr {{\mathbf 1}}\cr}\right)= C\left(\matrix{{\mathbf \beta}\cr {\mathbf b}_j\cr}\right)=\nonumber\\ -C\left(\matrix{{\mathbf \beta}\cr {\mathbf \gamma}\cr}\right)&&= C\left(\matrix{{\mathbf \gamma}\cr {\mathbf b}_2\cr}\right)= -C\left(\matrix{{\mathbf \gamma}\cr {\mathbf b}_1,{\mathbf b}_3,{\mathbf b}_4,{\mathbf \gamma}\cr}\right)= -1\nonumber \end{eqnarray} $(j=1,2,3),$ with the others specified by modular invariance and space--time supersymmetry. Several properties of the boundary conditions, eq.\ (\ref{fnymodel}), that generate the FNY model distinguish it from other standard--like models \cite{eu}. First the basis vector ${\mathbf \alpha}\equiv {\mathbf b}_4$ does not break the $SO(10)$ symmetry. In the models \cite{eu} both the basis vectors ${\mathbf \alpha}$ and ${\mathbf \beta}$ break the $SO(10)$ symmetry to $SO(6)\times SO(4)$. This has the consequence that the combination ${\mathbf \alpha}+{\mathbf \beta}$ gives rise to states transforming only under the observable ($SO(10)\times SO(6)^3$) part of the gauge group, which produce electroweak Higgs representations. Thus, we may anticipate that the structure of the Higgs mass matrix as well the fermion mass matrices in the FNY model will differ from those in the models of ref.\ \cite{eu}. Another distinction between the models is in the pairing of left--moving and right--moving fermions which produces Ising model operators \cite{slm}. \@startsection{subsection}{2}{\z@}{2.3ex plus .2ex{Gauge Group} Before cancellation of the FI term by an $F$-- and $D$--flat direction, the observable gauge group for the FNY model consists of the universal $SO(10)$ sub--group, $SU(3)_C\times SU(2)_L\times U(1)_C\times U(1)_L$, generated by the five complex world--sheet fermions ${\bar\psi}^{1,\cdots,5}$, and six observable horizontal, flavor--dependent, $U(1)$ symmetries $U(1)_{1,\cdots,6}$, generated by $\{{\bar\eta}^1,{\bar\eta}^2,{\bar\eta}^3,{\bar y}^3{\bar y}^6, {\bar y}^1{\bar\omega}^6,{\bar\omega}^1{\bar\omega}^3\}$, respectively. The hidden sector gauge group is the $E_8$ sub--group of $(SO(4)\sim SU(2)\times SU(2))\times SU(3)\times U(1)^4$, generated by ${\bar\phi}^{1,\cdots,8}$. The weak hypercharge is given by \begin{equation} U(1)_Y={1\over3}U(1)_C\pm{1\over2}U(1)_L, \label{weakhyper} \end{equation} which has the standard effective level $k_1$ of $5/3$, necessary for MSSM unification at $M_U$. As we noted in \cite{cfn1}, the sign ambiguity in eq.\ (\ref{weakhyper}) can be understood in terms of the two alternative embeddings of $SU(5)$ within $SO(10)$, that produce either the standard or flipped $SU(5)$ \cite{flipeq}. Switching signs in (\ref{weakhyper}) flips the representations, \begin{eqnarray} + &\leftrightarrow& -\\ e_L^c &\leftrightarrow& N_L^c\nonumber\\ u_L^c &\leftrightarrow& d_L^c\nonumber\\ h &\leftrightarrow& {\bar h}\,\, . \label{repflip} \end{eqnarray} In the case of $SU(5)$ string GUT models, only the ``--'' (i.e., flipped version) is allowed, since there are no massless matter adjoint representations, which are needed to break the non--Abelian gauge symmetry of the unflipped $SU(5)$, but are not needed for the flipped version. For MSSM--like strings, either choice of sign is allowed since the GUT non--Abelian symmetry is broken directly at the string level. The ``+'' sign was chosen for the hypercharge definition in \cite{fny}. In \cite{cfn1}, we showed that the choice of the sign in eq.\ (\ref{weakhyper}) has interesting consequences in terms of the decoupling of the exotic fractionally charged states. The other combination of $U(1)_C$ and $U(1)_L$, which is orthogonal to $U(1)_Y$, is given by \begin{equation} U(1)_{Z^\prime}=U(1)_C\mp U(1)_L\, . \label{u1prime} \end{equation} In Section 3 we will show that cancellation of the FI term by singlet fields along directions that are $D$--flat for all of the non--anomalous $U(1)$ requires that at least one of the $U(1)_Y$ and $U(1)_{Z'}$ be broken. Proof of this is found in Table C.II in Appendix C. Therefore, in the phenomenologically viable case we are forced to have only $SU(3)_C\times SU(2)_L\times U(1)_Y$ as the unbroken $SO(10)$ subgroup below the string scale, when only singlets take on VEVs. Under the constraint that only singlets take on VEVs, this is an interesting example of how string dynamics may force the $SO(10)$ subgroup below the string scale to coincide with the Standard Model gauge group. (Complete proof that only one of $U(1)_Y$ and $U(1)_{Z'}$ can survive would necessitate that non--Abelian state VEVs also be considered, which is currently being investigated \cite{cfn4}.) \@startsection{subsection}{2}{\z@}{2.3ex plus .2ex{Matter Spectrum and Superpotential} The full massless spectrum of the model, together with the quantum numbers under the right--moving gauge group, were first presented in ref.\ \cite{fny}. In our Tables A.I and A.II of Appendix A, we again list these states and their gauge and global charges. The gauge charges in Table A.I are expressed in the rotated $U(1)$ basis of eqs.\ (\ref{anomau1infny},\ref{nonau1}) (and of eqs.\ (18a-f) in \cite{fny}), rather than in the unrotated basis associated with the tables of \cite{fny}. In the FNY model, the three sectors ${\mathbf b}_1$, ${\mathbf b}_2$, and ${\mathbf b}_3$ correspond to the three twisted sectors of the $Z_2\times Z_2$ orbifold model, with each sector producing one generation in the 16 representation, ($\Q{i}$, $\uc{i}$, $\dc{i}$, $\L{i}$, $\ec{i}$, $\Nc{i}$), of $SO(10)$ decomposed under $SU(3)_C\times SU(2)_L\times U(1)_C\times U(1)_L$, with charges under the horizontal symmetries. In addition to the gravity and gauge multiplets and several singlets (see below), the untwisted Neveu--Schwarz (NS) sector produces three pairs of electroweak scalar doublets $\{h_1, h_2, h_3, {\bar h}_1, {\bar h}_2, {\bar h}_3\}$. Each NS electroweak doublet set $(h_i,\bar{h}_i)$ may be viewed as a pair of Higgs with the potential to give renormalizable (near EW scale) mass to the corresponding ${\mathbf b}_i$--generation of MSSM matter. Thus, to reproduce the MSSM states and generate a viable three generation mass hierarchy, two out of three of these Higgs pairs must become massive near the string/FI scale. The twisted sector provides some additional $SU(3)_C\times SU(2)_L$ exotics: one $SU(3)_C$ triplet/anti--triplet pair $\{\H{33},\,\, \H{40}\}$; one $SU(2)_L$ up--like doublet, $\H{34}$, and one down--like doublet, $\H{41}$; and two pairs of vector--like $SU(2)_L$ doublets, $\{{\bf V}{45},\,\, {\bf V}{46}\}$ and $\{{\bf V}{51},\,\, {\bf V}{52}\}$, with fractional electric charges $Q_e= \pm\frac{1}{2}$. The FNY model contains a total of 63 non-Abelian singlets. Three of these are the MSSM electron conjugate states ($\ec{1}$, $\ec{2}$, $\ec{3}$) and another three are the neutrino $SU(2)_L$ singlets ($\Nc{1}$, $\Nc{2}$, $\Nc{3}$). Of the remaining 57 singlets, 16 possess electric charge and 41 do not. The set of 16 are twisted sector states,\footnote{Vector--like representations of the hidden sector are denoted by a ``V'', while chiral representations are denoted by a ``H''. A superscript ``s'' indicates a non--Abelian singlet.} eight of which carry $Q_e= \frac{1}{2}$, \begin{equation} \{\Hs{3}, \Hs{5}, \Hs{7}, \Hs{9}, \Vs{41}, \Vs{43}, \Vs{47}, \Vs{49}\}, \label{set8a} \end{equation} and another eight of which carry $Q_e= -\frac{1}{2}$, \begin{equation} \{\Hs{4}, \Hs{6}, \Hs{8}, \Hs{10}, \Vs{42}, \Vs{44}, \Vs{48}, \Vs{50}). \label{set8b} \end{equation} Three of the 41 $Q_e= 0$ states, \begin{equation} \{\Phi_1,\Phi_2,\Phi_3 \}, \label{set3a} \end{equation} are NS sector singlets of the entire four dimensional gauge group. Another fourteen of these singlets, \begin{equation} \{{\tt +}{12},\,\, \pb{12},\,\, {\tt +}{23},\,\, \pb{23},\,\, {\tt +}{13}\,\, ,\pb{13}\,\, , {\tt +}{56},\,\, \pb{56},\,\, \pp{56},\,\, \ppb{56},\,\, {\tt +}{4},\,\, \pb{4},\,\, \pp{4},\,\, \ppb{4} \}, \label{set1a} \end{equation} form seven pairs, \begin{equation} ({\tt +}{12},\pb{12}),\,\, ({\tt +}{23},\pb{23}),\,\, ({\tt +}{13},\pb{13})\,\, , ({\tt +}{56},\pb{56}),\,\, (\pp{56},\ppb{56}),\,\, ({\tt +}{4} ,\pb{4} ),\,\, (\pp{4},\ppb{4}), \label{set1a2} \end{equation} that are vector--like, i.e., possessing charges of equal magnitude but opposite sign, for all local Abelian symmetries. (Note for later discussion that ${\tt +}{4}$ and $\pp{4}$ carry identical Abelian gauge charges, thereby resulting in six, rather than seven, {\it distinct} vector--like pairs.) The remaining 24 $Q_e= 0$ singlets, \begin{eqnarray} &&\{\Hs{15},\,\, \Hs{16},\,\, \Hs{17},\,\, \Hs{18},\,\, \Hs{19},\,\, \Hs{20},\,\, \Hs{21},\,\, \Hs{22},\,\, \Hs{29},\,\, \Hs{30},\,\, \Hs{31},\,\, \Hs{32},\,\, \Hs{36},\,\, \Hs{37},\,\, \nonumber \\ &&\phantom{\{} \Hs{38},\,\, \Hs{39},\,\, \Vs{1},\,\, \Vs{2},\,\, \Vs{11},\,\, \Vs{12},\,\, \Vs{21},\,\, \Vs{22},\,\, \Vs{31},\,\, \Vs{32} \}, \label{set1b} \end{eqnarray} are twisted sector states carrying both observable and hidden sector Abelian charges. The FNY model contains 34 hidden sector non--Abelian states, all of which also carry both observable and hidden $U(1)_i$ charges: Five of these are $SU(3)_H$ triplets, \begin{equation} \{\H{42},\,\, {\bf V}{4},\,\, {\bf V}{14},\,\, {\bf V}{24},\,\, {\bf V}{34} \}, \label{set3b} \end{equation} while another five are anti--triplets, \begin{equation} \{\H{35},\,\, {\bf V}{3},\,\, {\bf V}{13},\,\, {\bf V}{23},\,\, {\bf V}{33} \}. \label{set3c} \end{equation} The remaining 24 states include 12 $SU(2)_H$ doublets, \begin{equation} \{\H{1},\,\, \H{2},\,\, \H{23},\,\, \H{26},\,\, {\bf V}{5},\,\, {\bf V}{7},\,\, {\bf V}{15},\,\, {\bf V}{17},\,\, {\bf V}{25},\,\, {\bf V}{27},\,\, {\bf V}{39},\,\, {\bf V}{40} \} \label{set2ha} \end{equation} and a corresponding 12 $SU(2)^{'}_H$ doublets, \begin{equation} \{\H{11},\,\, \H{13},\,\, \H{25},\,\, \H{28},\,\, {\bf V}{9},\,\, {\bf V}{10},\,\, {\bf V}{19},\,\, {\bf V}{20},\,\, {\bf V}{29},\,\, {\bf V}{30},\,\, {\bf V}{35},\,\, {\bf V}{37} \}. \label{set2hb} \end{equation} The only hidden sector NA states with non--zero $Q_e$ are the four doublets $\H{1}$, $\H{2}$, $\H{11}$, and $\H{13}$, which carry $Q_e= \pm\frac{1}{2}$. The sector origins of all exotics is discussed in \cite{fny,fc,cfn1}, and a general classification of exotic states in the more realistic free fermionic models is discussed in ref.\ \cite{ssr}. For a string derived MSSM to result from the FNY model, the several exotic MSSM--charged states must be eliminated from the low energy effective field theory. Along with two linearly independent combinations of the $\h{i=1,2,3}$, and of the $\bar{h}{i=1,2,3}$, the entire set of states, \begin{eqnarray} &&\{ \H{33},\,\, \H{40},\,\, \H{34},\,\, \H{41},\,\, {\bf V}{45},\,\, {\bf V}{46},\,\, {\bf V}{51},\,\, {\bf V}{52},\,\, \label{minvevsa}\\ &&\phantom{\{} \H{1},\,\, \H{2},\,\, \H{11},\,\, \H{12},\,\, \label{minvevsb}\\ &&\phantom{\{} \Hs{3},\,\, \Hs{5},\,\, \Hs{7},\,\, \Hs{9},\,\, \Vs{41},\,\, \Vs{43},\,\, \Vs{47},\,\, \Vs{49},\,\, \label{minvevsc}\\ &&\phantom{\{} \Hs{4},\,\, \Hs{6},\,\, \Hs{8},\,\, \Hs{10},\,\, \Vs{42},\,\, \Vs{44},\,\, \Vs{48},\,\, \Vs{50} \}. \label{minvevsd} \end{eqnarray} must be removed. Examination of the MSSM--charged state superpotential\footnote{In Appendix B, we present the FNY superpotential up to sixth order. Terms in the superpotential belong to one of four classes, those containing, in addition to (possible) singlets: (i) nothing else, (ii) only MSSM--charged states, (iii) both MSSM and hidden sector charged--states, and (iv) only hidden sector charged--states.} shows that two out of three of each of the $h_i$ and $\bar{h}_i$ Higgs, and {\it all} of the (\ref{minvevsa}--\ref{minvevsd}) states, can, indeed, be decoupled from the low energy effective field theory via the terms, \begin{eqnarray} && {\tt +}{12} \h{1} \bar{h}{2} + {\tt +}{23} \h{3} \bar{h}{2} + \Hs{31} \h{2} \H{34} + \Hs{38} \bar{h}{3} \H{41}+\label{minveva}\\ && {\tt +}{4} [ {\bf V}{45} {\bf V}{46} + \H{1} \H{2} ] + \label{minvevb}\\ && \pb{4} [ \Hs{3} \Hs{4} + \Hs{5} \Hs{6} + \Vs{41} \Vs{42} + \Vs{43} \Vs{44} ] + \label{minvevc}\\ && \pp{4} [ {\bf V}{51} {\bf V}{52} + \Hs{7} \Hs{8} + \Hs{9} \Hs{10} ]+\label{minvevd}\\ && \ppb{4} [ \Vs{47} \Vs{48} + \Vs{49} \Vs{50} +\H{11} \H{13}]+ \label{minveve}\\ && {\tt +}{23} \Hs{31} \Hs{38} [ \H{33} \H{40} + \H{34} \H{41} ]\,\, .\label{minvevf} \end{eqnarray} This will occur if all states in the set \begin{equation} \{ {\tt +}{4},\,\, \pb{4},\,\, \pp{4},\,\, \ppb{4},\,\, {\tt +}{12},\,\, {\tt +}{23},\,\, \Hs{31},\,\, \Hs{38} \} \label{minvev} \end{equation} take on near string scale VEVs through FI term anomaly cancellation. All but two of the terms in (\ref{minveva}-\ref{minvevf}) are of third order and will result in unsuppressed FI scale masses, while the remaining two terms are of fifth order. The $\H{33} \H{40}$ fifth order term may result in $\H{33}$ and $\H{40}$ receiving a slightly suppressed mass. On the other hand, the $\H{34} \H{41}$ term has only a minor, perturbative effect on $\H{34}$ and $\H{41}$ masses, since $\H{34}$ and $\H{41}$ both appear in third order mass terms as well. In Section 3, we present $D$-- and $F$--flat directions that contain the required fields (\ref{minvev}) for decoupling all of the SM--charged exotics. Some of these directions are flat to all orders in the superpotential, while others are flat only to finite order. Before discussing these directions though, we review the process by which they were found. \subsubsection{Anomalous $U(1)_A$} All known chiral three generation $SU(3)_C\times SU(2)_L \times U(1)_Y$ models, of lattice, orbifold, or free fermionic construction, contain an anomalous local $U(1)_A$ \cite{anoma}. An anomalous $U(1)_A$ has non--zero trace of its charge over the massless states of the low energy effective field theory, \begin{equation} {\rm Tr}\, Q^{(A)} \neq 0\,\, . \label{audef} \end{equation} String models often appear to have not just one, but several anomalous Abelian symmetries $U(1)_{A,i}$ ($i= 1$ to $n$), each with ${\rm Tr}\, Q^{(A)}_i \neq 0$. However, there is always a rotation that places the entire anomaly into a single $U(1)_{A}$, uniquely defined by \begin{equation} U(1)_{\rm A} \equiv c_A\sum_{i=1}^n \{{\rm Tr}\, Q^{(A)}_{i}\}U(1)_{A,i}, \label{rotau} \end{equation} with $c_A$ a normalization coefficient. There are then $n-1$ traceless $U(1)'_j$ formed from linear combinations of the $n$ $U(1)_{A,i}$ and orthogonal to $U(1)_{\rm A}$. Prior to rotating the anomaly into a single $U(1)_{\rm A}$, six of the FNY model's twelve $U(1)$ symmetries are anomalous: Tr${\, U_1=-24}$, Tr${\, U_2=-30}$, Tr${\, U_3=18}$, Tr${\, U_5=6}$, Tr${\, U_6=6}$ and Tr${\, U_8=12}$. Thus, the total anomaly can be rotated into a single $U(1)_{\rm A}$ defined by \begin{equation} U_A\equiv -4U_1-5U_2+3U_3+U_5+U_6+2U_8. \label{anomau1infny} \end{equation} The five orthogonal linear combinations, \begin{eqnarray} U^{'}_1 &=& \hbox to 3.0truecm{$2 U_1 - U_2 + U_3$\,\, ;\hfill}\quad U^{'}_2= -U_1 + 5 U_2 + 7 U_3\,\, ;\nonumber \\ U^{'}_3 &=& \hbox to 3.0truecm{$U_5 - U_6$\,\, ;\hfill}\quad U^{'}_4= U_5 + U_6 - U_8\,\, \label{nonau1}\\ U^{'}_5 &=& 12 U_1 + 15 U_2 - 9 U_3 + 25 U_5 + 50 U_8\,\, . \nonumber \end{eqnarray} are all traceless. \@startsection{section}{1}{\z@}{3.5ex plus 1ex minus .2ex{Flat Directions} \@startsection{subsection}{2}{\z@}{2.3ex plus .2ex{Constraints on Flat Directions} \subsubsection{$D$--constraints} A set of vacuum expectations values (VEVs) will automatically appear in any string model with an anomalous $U(1)_{\rm A}$ as a result of the string theory anomaly cancellation mechanism \cite{u1a} . Following the anomaly rotation of \eq{rotau}, the universal Green--Schwarz (GS) relations, \begin{eqnarray} \frac{1}{k_m k_A^{1/2}}\mathop{{\rm Tr}\, }_{G_m}\, T(R)Q_A &=& \frac{1}{3k_A^{3/2}}{\rm Tr}\, Q_A^3 = \frac{1}{k_i k_A^{1/2}}{\rm Tr}\, Q_i^2 Q_A =\frac{1}{24k_A^{1/2}}{\rm Tr}\, Q_A \nonumber \\ &\equiv& 8\pi^2 \delta_{\rm GS} \, , \label{gsa}\\ \frac{1}{k_m k_i^{1/2}}\mathop{{\rm Tr}\, }_{G_m}\, T(R)Q_i &=& \frac{1}{3k_i^{3/2}}{\rm Tr}\, Q_i^3 = \frac{1}{k_A k_i^{1/2}}{\rm Tr}\, Q_A^2 Q_i = \frac{1}{(k_i k_j k_A)^{1/2}}{\rm Tr}\, Q_i Q_{j\ne i} Q_A \nonumber \\ &=&\frac{1}{24k_i^{1/2}}{\rm Tr}\, Q_i = 0 \, , \label{gsna} \end{eqnarray} where $k_m$ is the level of the non--Abelian gauge group $G_m$ and $2 T(R)$ is the index of the representation $R$ of $G_m$, defined by \begin{equation} {\rm Tr}\, \, T^{(R)}_a T^{(R)}_b = T(R) \delta_{ab}\, , \label{tin} \end{equation} removes all Abelian triangle anomalies except those involving either one or three $U_A$ gauge bosons. (The GS relations are a by--product of modular invariance constraints.) The standard anomaly cancellation mechanism breaks $U_A$ and, in the process, generates a FI $D$--term, \begin{equation} \epsilon\equiv \frac{g^2_s M_P^2}{192\pi^2}{\rm Tr}\, Q^{(A)}\, , \label{fidt} \end{equation} where $g_{s}$ is the string coupling and $M_P$ is the reduced Planck mass, $M_P\equiv M_{Planck}/\sqrt{8 \pi}\approx 2.4\times 10^{18}$ . The form of the FI--term was determined from string theory assumptions. Therefore, a more encompassing $M$--theory \cite{mth} might suggest modifications to this FI--term. However, recently it was argued that $M$--theory does not appear to alter the form of the FI--term \cite{jmr}. Instead an $M$--theory FI--term should remain identical to the FI--term obtained for a weakly--coupled $E_8\times E_8$ heterotic string, independent of the size of $M$--theory's 11$^{\rm th}$ dimension. Spacetime supersymmetry is broken near the string scale by the FI $D_A$--term, unless a set of scalar VEVs, $\{\vev{\varphi_m}\}$, carrying anomalous charges $Q^{(A)}_m$ can contribute a compensating $\vev{D_{A}(\varphi_m)} \equiv \sum_\alpha Q^{(A)}_m |\vev{\varphi_{m}}|^2$ term to cancel the FI--term, i.e., \begin{equation} \vev{D_{A}}= \sum_m Q^{(A)}_m |\vev{\varphi_{m}}|^2 + \epsilon = 0\,\, , \label{daf} \end{equation} thereby restoring supersymmetry. The actual set of VEVs accomplishing this cancellation will be dynamically determined by non--perturbative effects. Whichever VEV combination this may be, the phenomenology of the model will be drastically altered from that which exists before the VEV is applied. Any set of scalar VEVs satisfying eq.\ (\ref{daf}) must also retain $D$--flatness for all of the non--anomalous Abelian $U_i$ symmetries as well,\footnote{Here we consider flat directions involving only non--Abelian singlet fields. In cases where non--trivial representations of the non--Abelian gauge groups are also allowed to take on VEVs, generalized non--Abelian $D$--flat constraints must also be imposed. See, for example, \cite{cfn4}.} \begin{equation} \vev{D_i}= \sum_m Q^{(i)}_m |\vev{\varphi_{m}}|^2 = 0\,\, . \label{dana} \end{equation} \subsubsection{$F$--constraints} Each superfield $\Phi_{m}$ (containing a scalar field $\varphi_{m}$ and chiral spin--$\frac{1}{2}$ superpartner $\psi_m$) that appears in the superpotential imposes further constraints on the scalar VEVs. $F$--flatness will be broken (thereby destroying spacetime supersymmetry) at the scale of the VEVs unless, \begin{equation} \vev{F_{m}} \equiv \vev{\frac{\partial W}{\partial \Phi_{m}}} = 0; \,\, \vev{W} =0. \label{ff} \end{equation} $F$--flatness of a given set of VEVs can be broken by two types of superpotential terms: (i) those composed only of the VEV'd fields, and (ii) those composed, in addition to the VEV'd fields, of a single field without a VEV. (These superpotential terms were called ``type A'' and ``type B'' respectively in \cite{cceel2}.) Any power of a type A term appearing in a superpotential will break $F$. For example if generic fields $\varphi_1$, $\varphi_2$ and $\varphi_3$ take on VEVs in a $D$--flat direction, then any superpotential term of the form \begin{equation} \left( (\Phi_1)^{n_1} (\Phi_2)^{n_2} (\Phi_3)^{n_3} \right)^m \label{atypea} \end{equation} (with $\{n_{i=1,2,3}\}$ being a set of relative primes) eliminates $F$--flatness at respective order $m(n_1+n_2+n_3)$. Thus, all order $F$--flatness, requires that no terms of this form appear in the superpotential for any combination of $m$, $n_{i=1,2,3}$ values. On the other hand, when $\Phi_x$ lacks a VEV, we see from (\ref{ff}) that superpotential terms $((\Phi_1)^{n_1} (\Phi_2)^{n_2} (\Phi_3)^{n_3}(\Phi_x)^{n_x})^m$, (with $\{n_{i=1,2,3},n_x\}$ similarly a set of relative primes) $F$--flatness breaking type B terms only when $n_x=m=1$. Generically, there are many more $D$--flat directions that are simultaneously $F$--flat to a given order in the superpotential for the effective field theory of a string model than for the field--theoretic counterpart. In particular, there are usually several $D$--flat directions that are $F$--flat to all order in a string model, but only flat to some finite, often low, order in the corresponding field--theoretic model. This may be attributed to string world--sheet selection rules \cite{nahew5,wsc} which impose strong constraints on allowed superpotential terms beyond gauge invariance \cite{cceelw}. For example, for a given set of states, a comparison between the allowed terms in a stringy superpotential and the corresponding gauge invariant field--theoretic superpotential is performed in \cite{lr}. \@startsection{subsection}{2}{\z@}{2.3ex plus .2ex{$D$--flat Basis Sets} We continue our flat direction discussion by considering basis sets of $D$--flat directions. Let $\{ \varphi_{m=1\,\, {\rm to}\,\, n} \}$ denote the $n$--dimension set of fields of that are allowed to possibly take on VEVs. Further, let \begin{eqnarray} C_j &=& \{ |\vev{\varphi_{1}}_j|^2,\, |\vev{\varphi_{2}}_j|^2, \cdots , |\vev{\varphi_{n}}_j|^2 \} \label{cdf1}\\ &\equiv& \{ a_{j,1}, a_{j,2}, \cdots , a_{j,n} \}\,\, , \label{cdf12} \end{eqnarray} with \begin{equation} a_{j,x}\equiv |\vev{\varphi_{x}}_j|^2 \ge 0\,\, , \label{psd} \end{equation} be a generic set of the norms of VEVs \footnote{Hereon, we will often, for sake of brevity, refer to the norms of the VEVs simply as the VEVs when it is clear that the norms are implied.} that satisfy all non--anomalous $D$--flat constraints \eq{dana}. A set $C_j$ of such VEVs corresponds to a polynomial of fields $\varphi_{1}^{a_{j,1}} \varphi_{2}^{a_{j,2}} \cdots \varphi_{n}^{a_{j,n}}$, invariant under all non--anomalous gauge symmetries \cite{dfset,dfset2,cceel2}. A non--anomalous $D$--flat direction (\ref{cdf1}) possesses some number of overall scale degrees of freedom (DOFs), which is the dimension of the direction. For example, the norms of the VEVs of a one--dimensional $D$--flat direction can be expressed as a product of a single positive real overall scale factor $\alpha$ and positive semi--definite integral coefficients, $c^{\alpha}_{j,x} \ge 0$, which specify the ratios between the VEVs of the various $n$ fields, \begin{equation} C^{1{\rm -dim}}_j = \{ c^{\alpha}_{j,1} \alpha,\, c^{\alpha}_{j,2}\alpha,\cdots , c^{\alpha}_{j,n} \alpha \}\,\, . \label{cdf3} \end{equation} Similarly, a two-dimensional flat direction involves independent scales $\alpha$ and $\beta$, i.e., \begin{equation} C^{2{\rm -dim}}_j = \{ c^{\alpha}_{j,1} \alpha + c^{\beta}_{j,1} \beta ,\, c^{\alpha}_{j,2} \alpha + c^{\beta}_{j,2} \beta, \cdots , c^{\alpha}_{j,n} \ + c^{\beta}_{j,n} \beta \}\,\, . \label{cdf4} \end{equation} Any physical $D$--flat direction $C_j$ can be expressed as a linear combination of a set of $D$--flat basis directions, $\{B_1, B_2, B_3, \cdots , B_k \}$: \begin{equation} C_j = \sum_k w_{j,k} B_{k}\,\, , \label{cfb1} \end{equation} where $w_{j,k}$ are real weights. A basis can always be formed in which each element \begin{eqnarray} B_k& = &\{ |\vev{\varphi_{1}}_k|^2,\, |\vev{\varphi_{2}}_k|^2, \cdots , |\vev{\varphi_{n}}_k|^2 \} \label{bs2b}\\ &\equiv& \{ b_{k,1}, b_{k,2}, \cdots , b_{k,n} \}\,\, , \label{bs2b2} \end{eqnarray} where $b_{k,x}\equiv |\vev{\varphi_{x}}_k|^2$ is an integer, is one--dimensional, i.e., \begin{equation} B^{1{\rm -dim}}_k = \{ b^{'}_{k,1} \gamma_k, b^{'}_{k,2} \gamma_k, \cdots , b^{'}_{k,n} \gamma_k \}\,\, , \label{bs3b1} \end{equation} where $\gamma_k$ is the overall scale factor and $b^{'}_{k,x}$ are the relative ratios of the VEVs. Further, the scale factor of a basis element can always be normalized to 1, thereby leaving $B_k$ to be defined solely by the $b^{'}_{k,x}$, \begin{equation} B^{1{\rm -dim}}_k = \{ b^{'}_{k,1}, b^{'}_{k,2}, \cdots , b^{'}_{k,n} \}\,\, . \label{bs3bd2} \end{equation} Neither the coefficients $b^{'}_{k,x}$ \cite{gcm} of a basis element $B_k$, nor the weights $w_{j,k}$ \cite{cceel2} need all be non--negative, so long as the {\it total} contribution of all basis elements to an individual norm of a VEV, $a_{j,x} \equiv \sum_k w_{j,k} b^{'}_{k,x}$, in a flat direction $C_j$ is non--negative \cite{cceel2}. However, a basis vector $B_k$ that contains at least one negative coefficient $b^{'}_{k,x}< 0$ cannot be viewed as a physical one--dimensional $D$--flat direction. Instead it corresponds to a monomial of fields containing at least one field with a negative power, $ \varphi_{1}^{b^{'}_{k,1}} \varphi_{2}^{b^{'}_{k,2}} \cdots \varphi_{n}^{b^{'}_{k,n}}/\varphi_{x}^{|b^{'}_{k,x}|}$ Two types of basis sets were primarily discussed in \cite{cceel2}: (i) a basis composed of a maximal set of linearly independent {\it physical} (i.e., all $b^{'}_{k,x}\ge 0$) one--dimensional $D$--flat directions\footnote{See Table IV of \cite{cceel2} for an FNY non-anomalous flat direction basis of this type.} and (ii) a ``superbasis'' composed of the set of all (and therefore not all linearly independent) one--dimensional flat directions. The dimension $d_B$ of a linearly independent physical basis set is less than or equal to $N_{VEV} - N_{d}$, where $N_{VEV}=n$ is the number of fields allowed VEVs and $N_{d}$ is the number of independent non--anomalous $D$--constraints from the set of $N_U$ non--anomalous $U(1)_i$.\footnote{For a proper subset of states in a model, all $U(1)_i$ charges may not be independent.} A necessity for saturation of the upper bound for $d_B$, is that there is at least one pair of states with charges of opposite sign for each non--anomalous $U(1)_i$. When there is some $U(1)_s$ under which $N_s$ states all have charges of the same sign and the remaining $N_{VEV} - N_s$ states are uncharged, none of the $N_s$ states may take on a VEV and the number of relevant independent $U(1)_i$ is reduced by one, $d_B \le N^{'}_{VEV} - N^{'}_{d} = (N_{VEV}- N_s) - (N_{d} - 1) \le N_{VEV} - N_{d}$. In general, determining the effective $N^{'}_{VEV}$ and $N^{'}_{d}$ is an iterative process: After the $N_s$ states are removed and $N_{d}$ decreased by one for $U(1)_s$, then we must check again that there are no other $U(1)_i$ without a pair of states of oppositely sign charges, etc $\cdots$. Let $N^{'}_{VEV}$ and $N^{'}_d$ be the number of states allowed VEVs and number of independent $D$--constraints after ``eliminating'' all $U(1)_i$ that lack pairs of states with charges of opposite sign. As we discuss in greater detail in the next subsection, $d_B$ will generally be less than $N^{'}_{VEV} - N^{'}_{d}$ because of the very non--trivial positive semi--definite requirement (\ref{psd}) on the $\mvev{\phi_k}$ in (\ref{cdf1},\ref{cdf12}). There is one possible disadvantage to choosing a basis set of linearly independent physical VEVs to generate higher dimensional physical flat directions $C^{n{\rm -dim}}_j$. That is, there are usually physical flat directions $C_j$ that require some negative $w_{j,k}$ coefficients in (\ref{cfb1}) when $C_j$ is expressed in terms of physical basis directions. This can complicate the systematic generation of $D$--flat directions. On the other hand, using a superbasis of not all linearly independent physical VEVs overcomes this possible difficulty, since all higher dimensional physical $D$--flat directions, $C^{n\ge 2{\rm -dim}}_i$, can be constructed from superbasis elements $B_k$ using only non--negative coefficients $|w_{j,k}|$ \cite{cceel2}, \begin{equation} C_j= \sum_k |w_{j,k}| B_k\,\, . \label{gdf} \end{equation} In practice, however, application of the superbasis approach can, at times, have its own complication: the number of one--dimensional flat directions composing a given superbasis can be extremely large (on the order of several hundred or more) in some string models such as the FNY model. Extremely large dimensions of a superbasis can make systematic generation of $D$--flat directions unwieldy. An alternative to these two types of $D$--flat basis sets is what we term the ``maximally orthogonal'' basis set. This type of basis may, in fact, coincide with a linearly independent physical basis in some models, but more commonly is slightly larger in dimension. In a maximally orthogonal basis, each $D$--flat basis element $B^{mo}_k$ has one non--zero {\it positive} coefficient $b^{mo'}_{k,x}$ for which the corresponding $b^{mo'}_{k',x}$ are zero for all other basis elements $B^{mo}_{k'\neq k}$. Each $B^{mo}_k$ becomes associated with a particular field $\varphi_{x}$. Hence, all $w_{j,k}$ defining a physical non--anomalous $D$--flat direction $C_j$ must be non--negative since \begin{equation} a_{j,x}\equiv |\vev{\varphi_{x}}_j|^2 = w_{j,k} b^{mo'}_{k,x} \ge 0\,\, . \label{mopf} \end{equation} Otherwise a flat direction involving a negative $w_{j,k}$ weight would have at least one VEV with a negative norm and would, therefore, be unphysical. Associating one component $b^{mo'}_{k,x}$ of a maximally orthogonal basis element $B^{mo}_k$ with the field $\varphi_{x}$ often results in a few of the other components $b^{mo'}_{k,x'\ne x}$ being negative. Hence, several $B^{mo}_k$ may correspond to unphysical directions. This does not present any real difficulty though. Rather, production of only physical directions $C_j$ then simply places some constraints on linear combinations of the $B_k$. A ``maximally orthogonal'' basis set of flat directions has essentially the same advantage as a superbasis, yet can keep the dimension of the basis set reasonable when the dimension of the superbasis may be unfeasibly large. There is one case in which the positivity constraint (\ref{psd}) discussed above is effectively relaxed. The exception to (\ref{psd}) occurs when two states can be combined into vector--like pairs. Let us denote a generic pair as $\varphi_{m}$ and $\varphi_{-m}$, For this pair, $a_{j,m}$ in (\ref{cdf12}) may take on negative values in a physical flat direction, because a vector--like pair of states acts effectively like a single state with regard to all $D$--constraints. For a flat direction, $C_j$, the contribution of this pair to each $D$--term is \begin{eqnarray} \vev{D_i}(\varphi_{m},\varphi_{-m}) &=& Q^{(i)}_{m_{n}} \mvev{\varphi_{m}}_j + Q^{(i)}_{-m} \mvev{\phi_{-m}}_j \label{dppb1}\\ &\equiv& Q^{(i)}_{m} a_{j,m} + Q^{(i)}_{-m} a_{j,-m}\,\, . \label{dppb2} \end{eqnarray} Since by definition $Q^{(i)}_{-m} = - Q^{(i)}_{m}$, eq.\ (\ref{dppb2}) can be rewritten as \begin{eqnarray} \vev{D_i}(\varphi_{m},\varphi_{-m}) &=& Q^{(i)}_{m} (a_{j,m} - a_{j,-m}) \label{vdppb4}\\ &\equiv& Q^{(i)}_{m} a^{'}_{j,m}\,\, , \label{vdppb5} \end{eqnarray} where $a^{'}_{j,m}\equiv (a_{j,m} - a_{j,-m})$ may be positive, negative, or zero. We can consider $a^{'}_{j,m}$ as originating from a single field $\vev{\varphi^{'}_{m}}$. This allows us to reduce the effective number of nontrivial states, $N_{VEV}$, by one for each vector pair \cite{slm,gcm}. This reduction is compensated by an additional {\it trivial} $D$--flat direction basis element, $\vev{\varphi_{m}} = \vev{\varphi_{-m}}$, (corresponding to the binomial $\varphi_{m} \varphi_{-m}$) formed from both components of a vector pair. The FNY model contains several vector--like pairs and provides an excellent example of reduction of effective nontrivial states. Having all states in vector--like pairs is equivalent to totally relaxing the ``positivity'' constraint. As a general rule, the more vector pairs of non--Abelian singlets there are, the more likely a $D$--flat FI--term cancelling direction can be formed. \subsubsection{Maximally Orthogonal Basis Sets Via Singular Value Decomposition} One method for generating a maximally orthogonal basis set of $D$--flat directions for the non--anomalous $U(1)_i$ involves singular value decomposition (SVD) of a matrix \cite{pftv}. While the matrix \cite{gcm} method and the more standard monomial approach \cite{dfset,dfset2,cceel2} are essentially different languages for the same process, a strength of the matrix decomposition method is that it generates a complete basis of $D$--flat directions for non--Abelian singlet states {\it en masse}. We briefly discuss the SVD approach here, since it provides a somwhat new interpretation to flat directions. SVD is based on the mathematical fact that any $(M\times N)$--dimensional matrix $\bf D$ whose number of rows $M$ is greater than or equal to its number of columns $N$, can be written as the product of an $M\times N$ column--orthogonal matrix $\bf U$, an $N\times N$ diagonal matrix $\bf W$ containing only semi--positive--definite elements, and the transpose of an $N\times N$ orthogonal matrix $\bf V$ \cite{pftv}, \begin{equation} {\bf D}_{M\times N} = {\bf U}_{M\times N}\cdot {\bf W}^{\rm diag}_{N\times N} \cdot {\bf V}^T_{N\times N}, \quad {\rm for~} M\ge N\, . \label{decomp} \end{equation} This decomposition is always possible, no matter how singular the matrix is. The decomposition is also nearly unique, up to (i) making the same permutation of the columns of $U$, diagonal elements of $W$, and columns of $V$, or (ii) forming linear combinations of any columns of $U$ and $V$ whose corresponding elements of $W$ are degenerate. If initially $M < N$, then a $(N-M)\times N$ zero--matrix can always be appended onto $\bf D$ so this decomposition can be performed: ${\bf D}_{(M<N),N}\rightarrow {\bf D}^{'}_{(M=N),N}$. SVD is extremely useful when the matrix $\bf D$\footnote{We assume from hereon that $\bf D$ has been enhanced by a zero submatrix if necessary so that $M\ge N$.} is associated with a set of $M$ simultaneous linear equations expressed by, \begin{equation} {\bf D}\cdot \vec{x} = \vec{b}, \label{sle} \end{equation} where $x$ and $b$ are vectors. Eq.\ (\ref{sle}) defines a linear mapping from $N$--dimensional vector space $x$ to $M$--dimensional vector--space $b$. For $M=N$, $\bf D$ is singular when at least one of the $M$ constraints is not linearly independent. Associated with a singular $\bf D$ is a subspace of $\vec x$ termed the {\it nullspace}, ${\cal M}_{null}$, that is mapped to $\vec 0$ in $b$--space by $\bf D$. The dimension of this nullspace is referred to as the {\it nullity}. The subspace of $\vec b$ that {\it can} be reached by the matrix $\bf D$ acting on $\vec x$ is called the {\it range} of $\bf D$. The dimension of the range is denoted as the {\it rank} of $\bf D$ and is equal to the number of independent constraint equations $\equiv M'\le M$. Clearly \begin{equation} {\rm rank}\,{\bf D} + {\rm nullity}\,{\bf D} = N\,\, . \label{svd1} \end{equation} In the decomposition of $\bf D$ in (\ref{decomp}), the columns of $\bf U$ corresponding to the non--zero diagonal components of $\bf W$ form an orthonormal set of basis vectors that span the range of $D$. Alternately, the columns of $\bf V$ corresponding to the zero diagonal components of $\bf W$ form an orthonormal basis for the nullspace. As is perhaps obvious, this method is directly applicable to constructing $D$--flat directions, especially when only non--Abelian singlet states are allowed VEVs. Let $M= N_U$ ($M_I= N_d$) denote the number of (independent) $D$--flat constraints and $N= N_{VEV}$, the number of fields allowed to take on VEVs. Then the $D_{i,j}$ component of the matrix $\bf D$ is the $Q^{(i)}_j$ charge of the state $\varphi_j$. ($i$ takes on the value $A$ for the anomalous $U(1)_A$ and values $\{a=1 {\rm ~to~} M-1\}$ for the set of non--anomalous $U(1)_i$.) The components of the vector $x$ are the values of $\mvev{\phi_j}$, and $b$ has all zero--components except in its $U(1)_A$ position. The value of $b$ in the anomalous position is $-\epsilon$ from eq.\ (\ref{daf}). Let $\bf D'$ be the matrix that excludes the row of anomalous charges in $\bf D$. In this language, the dimension $d_B$ of the moduli space of flat directions (not necessarily all physical) for the $M'\equiv M-1 = N_U - 1 $ non--anomalous $U(1)_{i= 1 {\rm ~to~} M'}$ is the nullity of matrix ${\bf D}'$, denoted as dim ${\cal M}^{'}_{null}$, formed from the $M'$ non--anomalous $D$--flat constraints. In other words, the nullity of ${\bf D}'$ is all VEVs formed from combinations of states that have zero net charge in each non--anomalous direction. dim ${\cal M}^{'}_{null}$ is in the range \begin{equation} N-M'\le {\rm dim}\,{{\cal M}^{'}_{null}}= N-M'_I\le N, \label{mrange} \end{equation} where $N$ is the number of states allowed VEVs and $M'_I$ is the number of independent non--anomalous constraints. For the matrix $\bf D$, which contains the anomalous charges, the elements of the range corresponding to anomaly cancelling $\bf D$--flat directions are those formed solely from linear combinations of elements of the nullity of ${\bf D}'$ that generate an anomalous component for $\vec{b}$ of opposite sign to the FI term, $\epsilon$. The nullspace of $\bf D$ will likewise be formed from linear combinations of ${\bf D}'$s nullity elements that generate a zero anomalous component for $\vec{b}$. The dimension of the ${\bf D}'$ nullity subset that projects into the range of $\bf D$, denoted by ${\rm dim}\,{\cal M}^{'}_R$, is 1 (since the anomalous constraint must necessarily be independent of the non--anomalous constraints). An $(N-M'_I-1)$--dimensional subset of ${\cal M}^{'}_{null}$ forms the nullity of $\bf D$. Our maximally orthogonal basis set for FNY was obtained using the SVD routine in \cite{pftv}. After eliminating the row of the ${\bf D}_{(10+23)\times 33}^{fny}$ charge matrix corresponding to the anomalous $U(1)_A$ $D$--constraint, singular value decomposition was performed on the reduced matrix ${\bf D}_{9\times 33}^{fny'}$, \begin{equation} {\bf D}_{(9+24)\times 33}^{fny'} = {\bf U}_{(9+24)\times 33}^{fny'} \cdot {\bf W}_{33\times 33}^{fny'}\cdot {{\bf V}^{fny'}}^{T}_{33\times 33}\,\, . \label{fnysvg} \end{equation} As discussed in the following subsections, 33 is the number of FNY nontrivial singlets allowed VEVs and 9 is the number of effective $D$--constraints. An initial basis set of $D$--flat directions of dimension $d_b = 24 = {\rm dim}\,\, {\cal M}^{'}_{null} = 33 - 9$, was obtained. The components of these 24 basis directions are the components of the 24 columns of $\bf V$ for which the diagonal components of $\bf W$ is zero. While these 24 basis directions were not initially in maximally orthogonal form, a simple rotation transformed them into this. \@startsection{subsection}{2}{\z@}{2.3ex plus .2ex{FNY Flat Directions} \subsubsection{$D$--flat Basis} The FNY model contains a total of 63 non--Abelian singlets. These fields, along with their local $U(1)$ and global world--sheet charges, are listed in Appendix A. Of the 63 singlets, 14 can be used to form seven pairs of vector--like singlets: $({\tt +}{12},\pb{12})$, $({\tt +}{13},\pb{13})$, $({\tt +}{23},\pb{23})$, $({\tt +}{56},\pb{56})$, $(\pp{56},\ppb{56})$, $({\tt +}{4},\pb{4})$, $(\pp{4},\ppb{4})$. The two ${\tt +}{4}$--related pairs possess identical gauge charges, so we will refer to the model as having six {\it distinct} vector--like pairs of singlets and 49 non--vector--like singlets. Since we wish the SM gauge group to survive after cancellation of the FI term, we investigate flat directions involving only singlets not carrying hypercharge. The set of hypercharge--free singlets is composed of the six distinct vector--like pairs and 27 non--vectors, $\Hs{15\,\, {\rm to }\,\, 22}$, $\Hs{29\,\, {\rm to }\,\, 32}$, $\Hs{36\,\, {\rm to }\,\, 39}$, $\Vs{ 1\,\, {\rm to }\,\, 2}$, $\Vs{11\,\, {\rm to }\,\, 12}$, $\Vs{21\,\, {\rm to }\,\, 22}$, $\Vs{31\,\, {\rm to }\,\, 32}$, and $\Nc{ 1\,\, {\rm to }\,\, 3}$. These 33 vector--like and non--vector--like singlets carry varying combinations of $U(1)_{\rm A}$ and nine other $U(1)_i$ charges.\footnote{While there are ten non--anomalous $U(1)_i$ besides hypercharge, all singlets charged under $U(1)_9$ of the hidden sector also carry non--zero hypercharge. Thus, none of the singlets we allow to take on a VEV carry $U(1)_9$ charge.} Hence, from these singlets we can form $33 - 9 = 24$ non--trivial $D$--flat basis directions (some physical, others not physical). A non--trivial basis set may be generated by several methods, which can give differing properties to the set. As discussed in the preceding subsection, we chose the singular value decomposition method to generate a maximally orthogonal basis set, for which there is a one--to--one correspondence between the 24 flat directions in the basis set and 24 of the 33 distinct states allowed VEVs. Eight more {\it trivial} basis elements are formed from vector--like pairs. There are eight rather than six vector--like pairs because of the gauge charge redundancy between the two vector--like pairs, $(\Phi_4,\overline{\Phi}_4)$ and $(\Phi^{'}_4,\overline{\Phi}^{'}_4)$. In Table C.I, the first six components of each basis element correspond to the norms of the VEVs of the six distinct vector--like pairs. The norms of the VEVs of these fields may be either positive or negative in a physical flat direction. A positive norm implies a field $\phi_m$ takes on the VEV, while a negative norm implies the vector partner field $\phi_{-m}$ of opposite charge takes on the VEV. The remaining 27 components of each basis element give the norms of the the non--vector--like singlet fields. A true physical flat direction formed from a combination of basis elements of $D$--flat directions must have positive semi--definite norms for all of its non--vector--like fields. However, as discussed earlier, corresponding components of the basis elements need not. A basis element with a negative norm of a non--vector field simply is not a physical flat direction. Our maximally orthogonal method associates as many basis directions with non--vector fields as possible. In the FNY model, this is possible to do for 23 of the 24 basis directions. Thus, for the FNY model, a given basis direction could have up to $33-23-6=4$ VEV non--vector--like negative component norms, any number of which would imply a flat direction is ``non--physical.'' The four non--vector--like fields with potentially negative norms are $\Nc{1}$, $\Hs{39}$, $\Hs{16}$, and $\Nc{3}$.\footnote{The VEVs of these four fields appear as the last components of the basis directions in Table C.I of Appendix C.} Thus, the physical constraints, \begin{equation} \sum_{k= 1}^{24} w_k b^{mo'}_{k,{\Nc{ 1}}}\ge 0\,\,;\quad \sum_{k= 1}^{24} w_k b^{mo'}_{k,{\Hs{39}}}\ge 0\,\,;\quad \sum_{k= 1}^{24} w_k b^{mo'}_{k,{\Hs{16}}}\ge 0\,\,;\quad \sum_{k= 1}^{24} w_k b^{mo'}_{k,{\Nc{ 3}}}\ge 0\,\, , \label{wkcon2} \end{equation} must be imposed upon the weight factors $w_k$ in (\ref{cfb1}) through which physical flat directions $C_j$ are formed from the basis elements $B^{mo}_k$. An additional constraint on physical flat directions is, of course, that the net anomalous charge $Q^{(A)}$ must be negative, \begin{equation} Q^{(A)}(C_j) =\sum_{k= 1}^{24} w_k Q^{(A)}(B_k) < 0 \,\, , \label{wkcon3} \end{equation} since $\epsilon > 0$ in eq.\ (\ref{daf}). Of the 24 non--trivial basis directions, nine carry negative anomalous charge, while eight carry positive anomalous charge and seven do not carry this charge. The SVD approach was also used to generate a set of basis directions that simultaneously conserve both $U(1)_Y$ and $U(1)_{Z'}$. The elements of this five--dimensional set are presented in Table C.II. Interestingly, none of the these basis directions have negative anomalous charge: all are, in fact, chargeless under $U(1)_A$! Thus, FI--term cancelling $D$--flat directions can never be formed by non--Abelian singlets if both $U(1)_Y$ and $U(1)_{Z'}$ are to survive. In other words, singlet flat directions imply the reduction of observable $SO(10)$ to exactly the SM, $SU(3)_C\times SU(2)_L \times U(1)_Y$. \subsubsection{$F$--flat Directions} In Table D.I of Appendix D, we present several classes of directions in the parameter space of VEVs that are flat up to at least sixth order in the FNY superpotential. These classes are defined solely by the states that take on VEVs, rather than the specific ratios of the VEVs. The first three classes of directions are flat to {\it all} order in the superpotential. The fourth class is broken by two twelfth order type A terms, denoted as ``12--1'' and ``12--2'' in Table D.II, while the fifth through eighth classes are all broken at seventh order by a single term, denoted as ``7--1'' in Table D.II. Numerous (specifically 22) classes are broken at sixth order, again all by a single term, ``6--1.'' Three of the sixth order classes are simultaneously broken by an additional sixth order term, ``6--2.'' Thus, it appears likely that FNY $D$--flat directions producing the MSSM spectrum are either $F$--flat to all finite orders or experience $F$--breaking at twelfth order or lower. Our MSSM flat directions were found via a computer search that generated combinations of the maximally orthogonal $D$--flat basis directions $B^{mo}_k$ given in Table C.I of Appendix C. Linear combinations of up to nine $L_j$--class basis directions were surveyed, with the range of the non--zero integer weights, $w_k$, in (\ref{cfb1}) being from one to ten. To eliminate redundancy of flat directions, the set of non--zero $w_k$ in a given linear combination was required to be relatively prime, i.e., the greatest common factor among any permitted set of $w_k$ was 1. The combinations of basis directions producing our 30 classes of flat directions is given in Table D.IV. The examples in each of the 30 flat direction classes were formed from five, six, or seven $L_j$--class basis directions. We found that all $D$--flat directions involving eight or more maximally orthogonal basis directions experienced breaking of $F$--flatness at fifth order or lower. In our computer search, we required flat directions to minimally contain VEVs for the set of states, \begin{equation} \{ {\tt +}{4},\,\, \pb{4},\,\, \pp{4},\,\, \ppb{4},\,\, {\tt +}{12},\,\, {\tt +}{23},\,\, \Hs{31},\,\, \Hs{38} \}\,\, , \nonumber \end{equation} necessary for decoupling of all 32 SM--charged MSSM exotics, comprised of the four extra Higgs doublets and the 28 exotics identified in (\ref{minvevsa}--\ref{minvevsd}). Imposing this, we found VEVs of $\Hs{15}$ and $\Hs{30}$ were also always present. We refer to the set of VEVs of these ten fields in Appendix $D$ as ``$\{ VEV_1\}$.'' All of the eight directions broken at seventh order or higher additionally contained the VEV of $\pb{56}$. The three classes of directions $F$--flat to all finite order also involved (i) no other VEVs, (ii) the VEV of $\pbp{56}$, and (iii) the VEV of $\Hs{19}$, respectively. The class broken at twelfth order additionally included the VEV of $\Hs{20}$. Note that the trilinear superpotential term $\pbp{56}\Hs{19}\Hs{20}$ allows only one of $\pbp{56}$, $\Hs{19}$, and $\Hs{20}$ to receive a VEV in any flat direction. In all directions with seventh order flatness, the sneutrino $SU(2)_L$ singlet $\Nc{1}$ takes on a VEV, as do one of $\pbp{56}$, $\Hs{19}$, or $\Hs{20}$ and/or $\Vs{31}$. For the 22 sixth order classes, subsets of $\{\Nc{3}$, $\Hs{17}$, $\Hs{18}$, $\Hs{21}$, $\Hs{39}$, $\Vs{12}\}$ obtain VEVs, along with various combinations from $\{\pbp{56}$, $\pp{56}$, $\Hs{19}$, $\Hs{20}$, $\Vs{31}$, $\Nc{1}\}$. Systematic generation of flat directions of the FNY model is efficiently performed using a maximally orthogonal basis. However, the dimension (i.e., the number of VEV scale degrees of freedom) $Dim$ of a given direction is not always apparent from this approach. To determine $Dim$, we also express each flat direction class in terms of its embedded {\it physical} one--dimensional $D$--flat directions. (See Tables D.IV and D.V.) Prior to FI term cancellation, the dimension of a given MSSM flat direction equals the number of embedded physical dimension--one $D$--flat directions. Cancellation of the FI term removes one degree of freedom, so the dimension after FI term cancellation, $Dim_{FI}$, is one less than $Dim$. The number $N_B$ of non--anomalous $U(1)_i$ broken along a given direction is the difference between the number of independent VEVs, $N_{VEV}$, and the dimension $Dim$, \begin{equation} N_{VEV} - Dim = N_{B}. \label{nbr1} \end{equation} Or, equivalently, \begin{equation} N_{VEV} - Dim_{FI} = N_{B} + 1. \label{nbr2} \end{equation} As Table D.VI shows, the all--order and twelfth order $F$--flat directions break seven non--anomalous $U(1)_i$, while the seventh order flat directions break eight. The sixth order directions remove anywhere from eight to ten non--anomalous $U(1)_i$. \setcounter{footnote}{0} A pair of physical one--dimensional flat directions, denoted ``$X$'' and ``$Y$'', are at the heart of 27 (out of 30) of our MSSM directions. $X$ is identified with the class 1 all--order flat direction\footnote{The class 1 flat direction was first formed in \cite{cceel2} from the combination of VEVs denoted $M_6$, $M_7$ and $R_{10}$. Table V of \cite{cceel2} also contains several FNY $Dim_{FI}=0$ non--MSSM flat directions.} and is the root of the three all--order, the one twelfth order, the four seventh order, and five of the sixth order flat directions. The seventh and higher order directions contain anywhere from zero to three additional physical dimension--one directions. At the root of 14 sixth order directions is the direction $Y$. $X$ and $Y$ are simultaneously embedded in five of the sixth order directions, while three of the sixth order directions contain neither. Our MSSM flat direction classes have a property not common to generic stringy flat directions. Specifically, $F$--flatness of the MSSM directions is broken by the stringy superpotential at exactly the same level as it would be in a field--theoretic gauge invariant superpotential. That is, stringy world sheet constraints do not remove all of the lowest order dangerous gauge--invariant terms. The equivalent string and gauge--invariant seventh and twelfth order $F$--breaking essentially results from all of the ${\tt +}{4}$, $\pp{4}$, $\pb{4}$, and $\ppb{4}$ states necessarily taking on VEVs. \@startsection{section}{1}{\z@}{3.5ex plus 1ex minus .2ex{Discussion} We have investigated $D$-- and $F$--flat directions in the FNY model of \cite{fny,fc}. The FNY model possesses two aspects generic to many classes of three family $SU(3)_C\times SU(2)_L\times U(1)_Y$ string models: both an extra local anomalous $U(1)_A$ and numerous (often fractionally charged) exotic particles beyond the MSSM. We found several flat directions involving only non--Abelian singlet fields that near the string scale can simultaneously break the anomalous $U(1)_A$ and give mass to {\it all} exotic SM--charged observable particles, decoupling them from the low energy spectrum. We were thus able to produce the first known examples of Minimal Superstring Standard Models. Some of our flat directions were shown to be flat to {\it all} finite orders in the superpotential. The models produced by our flat directions are consistent with, and may in fact offer the first potential realizations of, the recent conjecture by Witten of possible equivalence between the string scale and the minimal supersymmetric standard model unification scale $M_U\approx 2.5\times 10^{16}$ GeV. This conjecture indeed suggests that the observable gauge group just below the string scale should be $SU(3)_C\times SU(2)_L\times U(1)_Y$ and that the $SU(3)_C\times SU(2)_L\times U(1)_Y$--charged spectrum of the observable sector should consist solely of the MSSM spectrum. We have also discovered that the FNY model provides an interesting example of how string dynamics may force the $SO(10)$ subgroup below the string scale to coincide with the SM gauge group. When only non--Abelian singlets take on VEVs, we have shown that $U(1)_Y$ or $U(1)_{Z'}$ of $SU(3)_C\times SU(2)_L\times U(1)_Y\times U(1)_{Z'}\in SO(10)$ is necessarily broken. Reversing the roles of $U(1)_Y$ and $U(1)_{Z'}$ corresponds to flipping the components in each $SU(2)_L$ doublet. The phenomenology obtained from our various flat directions will be studied in \cite{cfn3}. In particular, we will examine the mass hierarchies of the three generations of SM quarks and leptons, the hidden sector dynamics, and issues such as proton decay. For each flat direction, we will present the resulting superpotential after decoupling of the states turned massive via FI--term cancelling VEVs. The rich space of flat directions that we found in the present paper suggests the exciting prospect that one of these flat directions may accommodate all of the phenomenological constraints imposed by the Supersymmetric Standard Model phenomenology. Furthermore, the rich space of solutions may be even further enlarged by adding VEVs of the non--Abelian fields. It ought to be emphasized that it is the promising structure, afforded by the NAHE set, which enables this promising scenario. To highlight this important fact, NAHE--based models should be contrasted with the non--NAHE based models, which although having three generations with the Standard Model gauge group, do not allow the standard $SO(10)$ embedding of the Standard Model spectrum and contain massless exotic states that cannot be decoupled. Thus, we emphasize once again, that although suggesting a specific three generation model as the true string vacuum, seems still premature, the concrete results, obtained in the analysis of specific models, highlight the underlying, phenomenologically successful, structure generated by the NAHE set. Therefore, it suggests that the true string vacuum could be in the vicinity of these models. That is, it is a $Z_2\times Z_2$ model in the vicinity of the free fermionic point in the Narain moduli space, also containing several, perhaps still unknown, additional Wilson lines. Such Wilson lines correspond in the fermionic language to the boundary condition basis vectors beyond the NAHE set. \@startsection{section}{1}{\z@}{3.5ex plus 1ex minus .2ex{Acknowledgments} This work is supported in part by DOE Grants No. DE--FG--0294ER40823 (AF) and DE--FG--0395ER40917 (GC,DVN). GC wishes to thank Alexander Maslikov for helpful discussions. \newpage
\section{INTRODUCTION} \begin{table*}[tb] \small \renewcommand{\arraystretch}{1.3} \renewcommand{\tabcolsep}{2mm} \begin{center} TABLE 1 {\sc Cluster Parameters$^{\rm a}$} \vspace{1mm} \begin{tabular}{p{1.8cm}cccccccccc} \hline \hline Cluster& $a_x^{\rm b}$ & $\beta^{\rm b}$ &$\rho_{\rm gas,0}^{\rm b}$ % & $T_e^{\rm c}$ & $T_X^{\rm d}$ % & $M(0.2\;{\rm Mpc})$ & $M(1\;{\rm Mpc})$ & $f_{\rm gas}(1\;{\rm Mpc})$ % & $M(r_{500})^{\rm e}$ & $r_{200}/r_s^{\rm ~~e}$ \\ & kpc & & $M_\odot {\rm Mpc}^{-3}$ % & keV & keV % & $10^{14}$~$M_{\odot}$ & $10^{14}$~$M_{\odot}$ & & $10^{14}$~$M_{\odot}$ & \\ \hline A2199 \dotfill&134 &0.636 & $2.24\times10^{14}$& $4.4\pm0.2$ &$4.8\pm0.2$ % & $0.65\pm0.11$ &$2.9\pm0.3$ & $0.161\pm 0.014$ & $3.6\pm0.5$ & 10 \\ A496 \dotfill &249 &0.700 & $1.02\times10^{14}$& $4.3\pm0.2$ &$4.7\pm0.2$ % & $0.47\pm0.10$ &$3.1\pm0.3$ & $0.158\pm 0.017$ & $3.9\pm0.6$ & 6 \\ \hline \end{tabular} \vspace{1mm} \begin{minipage}{17.5cm} $^{\rm a}$All values are for $h=0.5$. $^{\rm b}$Best-fit values for the {\em ROSAT}\/\ PSPC brightness profile excluding the central $r=3'$. The gas density value is an extrapolation of this $\beta$-model to the center. $^{\rm c}$Wide-beam single temperature fit. $^{\rm d}$Emission-weighted temperature excluding cooling flows. $^{\rm e}$Involves extrapolation to the area not covered by the temperature profile. \end{minipage} \end{center} \vspace*{-2mm} \end{table*} Under a reasonable, but as yet not directly tested, set of assumptions that the hot intracluster gas is supported by its own thermal pressure and is in hydrostatic equilibrium in the cluster gravitational well, one can determine the total mass of a cluster, including its dominant dark matter component (Bahcall \& Sarazin 1977; Mathews 1978). Because clusters are the largest collapsed objects in the Universe, their mass values are of great importance for cosmology. The cluster mass function and its evolution with redshift constrain the spectrum of the cosmological density fluctuations and the density parameter $\Omega_0$ (e.g., Press \& Schechter 1974; White \& Rees 1978; Bahcall \& Cen 1992; Viana \& Liddle 1996). If the cluster matter inventory is representative of the Universe as a whole, as is expected, then by measuring the cluster total and baryonic mass and comparing it to the predictions of primordial nucleosynthesis, one can constrain $\Omega_0$ (White et al.\ 1993). Comparison of independent cluster mass estimates, for example, by X-ray and gravitational lensing (e.g., Bartelmann \& Narayan 1995) methods, provide unique insights into cluster structure and physics. A discrepancy between the different estimates may indicate significant turbulence or nonthermal pressure in the intracluster gas (e.g., Loeb \& Mao 1994), or the effect of line of sight projections. For an X-ray measurement of the cluster mass, one needs accurate radial profiles of the gas density and temperature, as well as confidence that the cluster is in hydrostatic equilibrium. The gas density profile for a symmetric cluster can readily be obtained with an imaging instrument, such as {\em Einstein}\/\ or {\em ROSAT}\/. Obtaining temperature distributions has proven to be more problematic, especially for hotter, more massive clusters. {\em ASCA}\/\ (Tanaka, Inoue, \& Holt 1994) now provides spatially resolved temperature data for nearby hot clusters (e.g., Ikebe et al.\ 1997; Loewenstein 1997; Donnelly et al.\ 1998; Markevitch et al.\ 1998 [hereafter MFSV] and references therein), although their accuracy is still limited. Outside the central cooling flow regions, the temperature decreases with radius in most studied clusters. For a few clusters with more accurate temperature profiles, accurate mass profiles were already obtained (e.g., for A2256 by Markevitch \& Vikhlinin 1997, hereafter MV). MFSV found that gas temperature profiles of nearby symmetric clusters outside the cooling flow regions are similar when scaled by the virial radius and average temperature. The gas density profiles also are rather similar (e.g., Jones \& Forman 1984; Vikhlinin, Forman, \& Jones 1999). This suggests that the underlying dark matter profiles are similar. Indeed, analytical work and cosmological cluster simulations (e.g., Bertschinger 1985; Cole \& Lacey 1996; Navarro, Frenk, \& White 1995, 1997, hereafter NFW) predict that the dark matter radial profiles of most clusters in equilibrium should be similar in units of the virial radius. It is interesting to see whether their predicted ``universal'' dark matter profile agrees with the observations. For all but a few clusters in the MFSV sample, the temperature data have insufficient accuracy for such a test. We therefore selected two additional typical, relaxed, but less distant clusters, A2199 ($z=0.030$) and A496 ($z=0.033$), for a more accurate temperature profile and mass derivation using {\em ASCA}\/. These clusters are very ordinary in their X-ray luminosities and temperatures ($T\simeq 4.5$ keV) and, similarly to most clusters, have moderate cooling flows (170 and 95 $M_{\odot}\;$yr$^{-1}$, respectively; Peres et al.\ 1998). The presence of cooling flows is suggestive of a relaxed cluster, while at the same time these flows are not so strong as to prevent accurate resolved temperature measurements with ASCA (see MFSV). A subset of the data presented here (observations of the central regions) was already analyzed by Mushotzky et al.\ (1995). We have since obtained offset observations, and include the {\em ASCA}\/\ PSF correction in our analysis. Below we use {\em ASCA}\/\ and {\em ROSAT}\/\ data on these two clusters to derive their total mass profiles. We use $H_0$=50~km$\;$s$^{-1}\,$Mpc$^{-1}$\ ($h=0.5$); the error intervals are 90\%. \begin{figure*}[tb] \begingroup\pst@ifstar\pst@picture(0,9.8)(18.5,22.9) \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[tl]{0}(0.5,23.2){\epsfxsize=8.5cm \epsffile{2199map_gray_bold.ps}} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[tl]{0}(0.75,16.2){\epsfxsize=7.75cm \epsffile[30 470 530 678]{2199temps_bold.ps}} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[tl]{0}(9.5,23.2){\epsfxsize=8.5cm \epsffile{496map_gray_bold.ps}} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[tl]{0}(9.75,16.2){\epsfxsize=7.75cm \epsffile[30 470 530 678]{496temps_bold.ps}} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[l]{0}(2.3,22.5){\small A2199} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[l]{0}(11.3,22.5){\small A496} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[tl]{0}(0,12.0){ \begin{minipage}{18cm} \small\parindent=3.5mm {\sc Fig.}~1.---{\em ASCA}\/\ projected temperature maps (color) overlaid on the {\em ROSAT}\/\ PSPC brightness contours spaced by a factor of 2. Sectors in which the temperature was derived are numbered in upper panels and the temperatures with 90\% errors are given in lower panels. Different colors correspond to significantly different temperatures. Vertical dotted lines separate regions belonging to different annuli. Dotted horizontal lines show average temperature inside the annulus. For the central cooling flow regions, a single-temperature fit is shown. A white circle in the map shows a point source excluded from the fit. \end{minipage} } \endpspicture \end{figure*} \begin{figure*}[tb] \begingroup\pst@ifstar\pst@picture(0,10.2)(18.5,21) \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[tl]{0}(0.4,20.7){\epsfxsize=8.5cm \epsffile{2199tprof_bold.ps}} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[tl]{0}(9.5,20.7){\epsfxsize=8.5cm \epsffile{496tprof_bold.ps}} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[bl]{0}( 2.3,19.4){\small A2199} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[bl]{0}(11.4,19.4){\small A496} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[tl]{0}(0,12.4){ \begin{minipage}{18cm} \small\parindent=3.5mm {\sc Fig.}~2.---Radial projected temperature profiles. Crosses are centered on the emission-weighted radii. Vertical errors are for 90\% confidence; horizontal error bars show the boundaries of the annulus. Gray bands denote a continuous range of temperatures in a cooling flow, and the central cross corresponds to the upper (ambient) temperature of the cooling flow. Dotted and dashed horizontal lines show wide beam single-temperature fits and average emission-weighted temperatures excluding the cooling flow. Smooth lines show polytropic fits to the values outside the central bin, $\gamma=1.17$ and 1.24 for A2199 and A496, respectively. Values of $r_{1000}$ and $r_{500}$ are calculated from the mass profiles obtained from these temperature data (see Figs.\ 5 and 6 below). \end{minipage} } \endpspicture \end{figure*} \section{{\em ROSAT}\/\ PSPC DATA} \label{rossec} To derive the gas density distribution, we use {\em ROSAT}\/\ PSPC data. The archival observations of A2199 and A496 were analyzed as prescribed by Snowden et al.\ (1994) and using S. Snowden's code. To optimize the signal to noise ratio, we used Snowden bands 5--7 that correspond to 0.7--2.0 keV. For A2199, two observations of the same field were combined. The radial brightness profiles were then fit with a $\beta$-model $S_X(r) \propto (1+r^2/a_x^2)^{-3\beta+\frac{1}{2}}$ plus a uniform X-ray background within a radial range of $3'-50'$. The inner radius of 3$^\prime$\ approximately corresponds to the cooling radius for both clusters (e.g., Peres et al.\ 1998) and encompasses all of the X-ray brightness excess due to the moderate cooling flows in the cluster centers. The resulting parameters of the gas density profile are given in Table 1 and are typical (e.g., Jones \& Forman 1984). The $\beta$-model values for A2199 are similar to the results of Siddiqui, Stewart, \& Johnstone (1998) using the same data. \section{{\em ASCA}\/\ DATA} A2199 and A496 were each observed by {\em ASCA}\/\ with one central pointing and two different 14--15$^\prime$\ offsets from the cluster centers. Such a configuration has been chosen to cover the cluster to a radius where the mean overdensity is 500 ($r_{500}$), while at the same time keeping the cluster brightness peak within the {\em ASCA}\/\ field of view to avoid stray light contamination. Observing the clusters at different positions in the focal plane also reduces the {\em ASCA}\/\ systematic uncertainties that dominate in the temperature estimates. The offset positions were chosen to avoid bright foreground sources and also to cover representative regions of these slightly elliptical clusters. After the standard data screening (ABC Guide% \footnote{http://heasarc.gsfc.nasa.gov/docs/asca/abc/abc.html}% ), useful GIS exposures for the A2199 central and offset pointings were 31 ks, 19 ks, and 22 ks, and for A496, they were 37 ks, 24 ks, and 24 ks, respectively (the corresponding SIS exposures were about a factor of 0.8 of the GIS exposures). For the temperature fits, all pointings for both GIS and SIS were used simultaneously; different pointings and instruments fitted separately give consistent results. To derive the spatial temperature distributions, we used the method described in detail in MFSV and references therein. This method accounts for the {\em ASCA}\/\ PSF and assumes that outside the cooling flow regions, the {\em ROSAT}\/\ PSPC image provides an accurate description of the relative spatial distribution of the projected gas emission measure, after a correction of the PSPC brightness for any gas temperature variations. It should be mentioned here that a recent discovery of the possibly nonthermal EUV and soft X-ray ($E<0.2$ keV) emission in A2199 should not affect the latter assumption in any significant way, since we use a relatively hard ($0.7-2.0$ keV) PSPC band where this excess is absent (Lieu, Bonamente, \& Mittaz 1999). The absorption column was assumed uniform at the Galactic values ($N_H=0.9\times 10^{20}$~cm$^{-2}$\ and $4.6\times 10^{20}$~cm$^{-2}$\ for A2199 and A496); for our $E>1.5$ keV spectral fitting band, any expected variations are unimportant. The analysis method propagates all known calibration and other systematic uncertainties, including those of the {\em ASCA}\/\ PSF, effective area, {\em ROSAT}\/\ and {\em ASCA}\/\ backgrounds etc., to the final temperature values. All reported confidence intervals are one-parameter 90\% and are estimated by Monte-Carlo simulations. \section{RESULTS} \subsection{Temperature Maps} The resulting two-dimensional projected temperature maps are shown in Fig.\ 1, overlaid on the {\em ROSAT}\/\ images. We show only sectors in which the temperature is accurately constrained. The maps show no significant azimuthally asymmetric variations, and together with the brightness contours suggest that these clusters are well relaxed. These maps may be contrasted to the similarly derived, but highly irregular, temperature maps of merging clusters, e.g., A754 (Henriksen \& Markevitch 1996) and Cygnus-A and A3667 (Markevitch, Sarazin, \& Vikhlinin 1999). In the central regions, the maps clearly show low temperature regions that correspond to the previously known cooling flows (e.g., Stewart et al.\ 1984; Edge, Stewart, \& Fabian 1992). \subsection{Radial Temperature Profiles} \label{tprof} Figure 2 shows the cluster projected temperature profiles in five annuli. For the central radial bin, we used a model consisting of a thermal component and a cooling flow with the upper temperature tied to that of the thermal component, both with free normalizations. The figure also shows wide-beam, single-temperature fits ($T_e=4.4\pm0.2$ keV and $4.3\pm0.2$ keV for A2199 and A496, respectively) and emission-weighted average temperatures excluding the cooling flow component ($T_X=4.8\pm0.2$ keV and $4.7\pm0.2$ keV). The latter are calculated from these temperature profiles as described in MFSV. The data indicate a higher temperature in the central cluster regions (outside the cooling flows) compared to the average temperature, and a temperature decline with radius. This is similar to other clusters; in fact, when the profiles for A2199 and A496 are plotted in units of $T_X$ and virial radius, they lie within the composite profile obtained by MFSV for other nearby, relatively symmetric clusters (Fig.\ 3). Such typical temperature profiles, together with the typical gas density profiles and the presence of cooling flows, make A2199 and A496 representative examples of relaxed clusters. Outside the central cooling flow bin, the profiles in Fig.\ 2 are described remarkably well by a polytrope, $T_{\rm gas}\propto \rho_{\rm gas}^{\gamma-1}$ (both temperature profiles appear slightly more concave than the polytropic fits, which is probably nothing more than a coincidence; note that they differ in a similar way from the composite profile in Fig.\ 3). Assuming the {\em ROSAT}\/-derived $\beta$-models for $\rho_{\rm gas}$, we find $\gamma=1.17\pm0.07$ and $\gamma=1.24_{-0.11}^{+0.08}$ for A2199 and A496, respectively. Regardless of whether this fact has any physical meaning or is purely fortuitous, it simplifies the total mass derivation by providing a convenient functional form for the observed temperature profile. We will use it in the next section, but first note that for the mass derivation, one needs a real (three-dimensional) gas temperature profile as opposed to projected on the plane of the sky that we have obtained. We show in Appendix that as long as the gas density follows a $\beta$-model and the temperature is proportional to a power of density, a projected polytropic temperature profile differs from the three-dimensional profile only by a normalization. For the best-fit $\beta$ and $\gamma$ values for A2199 and A496, the projected temperature profiles are factors of 0.94 and 0.92 lower than the three-dimensional profiles, respectively. \section{TOTAL MASS PROFILES} \label{masssec} For the total mass determination, we will take advantage of the fact that the temperature profiles can be described by a polytropic functional form. >From the hydrostatic equilibrium equation for a spherically symmetric gas distribution $\rho_{\rm gas}(r)\propto (1+r^2/a_x^2)^{-\frac{3}{2}\beta}$ and a temperature profile $T\propto \rho_{\rm gas}^{\gamma-1}$, the total mass within a radius $r=xa_x$ is given by \begin{equation} \label{eq:m} M(r) = 3.70\times 10^{13} M_\odot\, \frac{0.60}{\mu}\, \frac{T(r)}{\rm 1\;keV}\, \frac{a_x}{\rm 1\;Mpc}\, \frac{3 \beta \gamma x^3}{1+x^2} \end{equation} (see, e.g., Sarazin 1988). A polytropic temperature decline thus corresponds to the following correction to an isothermal mass estimate $M_{\rm iso}$: \begingroup\pst@ifstar\pst@picture(0,-1.8)(8.5,9.8) \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[tl]{0}(-0.5,9.7){\epsfxsize=8.8cm \epsffile{2199evr2_bold.ps}} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[tl]{0}(-0.45,1){ \begin{minipage}{8.75cm} \small\parindent=3.5mm {\sc Fig.}~3.---Temperature profiles for A2199 and A496 (symbols) overlaid on the gray band representing a composite profile for a sample of 19 nearby relatively symmetric clusters from MFSV (their Figs.\ 7 and 8). For this comparison, the profiles are normalized by their cooling flow-corrected average temperatures $T_X$ and plotted in units of $r_{180}$ estimated from $T_X$ using the relation of Evrard et al.\ (1996). Cooling flow bins are not shown. The darker band corresponds to a scatter of best-fit temperature values of the MFSV sample and the lighter band covers most of their 90\% intervals. \end{minipage} } \endpspicture \begin{equation} \label{eq:mcor} \frac{M(r)}{M_{\rm iso}(r)}=\frac{T(r)}{\overline{T}}\,\gamma, \end{equation} where $\overline{T}$ is the average temperature (see also Ettori \& Fabian 1999). To calculate the 90\% confidence bands on mass profiles (as well as the confidence intervals on the values of $\gamma$ above), we have fitted the polytropic model to the same simulated temperature values in those annuli that were used to calculate the temperature error bars (see MFSV). These fitted polytropic models were substituted into equations (1) and (2) above and 90\% confidence intervals of the resulting values were calculated at each radius. The resulting correction factor to the isothermal mass estimate is shown in Fig.\ 4 as a function of radius. The corresponding profiles of the total mass, $M$, and the gas mass fraction, $f_{\rm gas}\equiv M_{\rm gas}/M$, are shown in Fig.\ 5. The corresponding ratio of the mean total density within a given radius to the critical density at the cluster's redshift [$\rho_c=3 H_0^2 (1+z)^3 /8\pi G$] is shown as a function of radius in Fig.\ 6. For both clusters, our mass profiles correspond to $r_{1000}\approx 1.0$ Mpc and $r_{500}\approx 1.3-1.4$ Mpc (the latter involves extrapolation to a region not covered by the temperature profile, see Fig.\ 2). Masses and gas fractions at several interesting radii are also given in Table 1. At $r=1$ Mpc, Mushotzky et al.\ (1995) obtained mass estimates of $2.55\times 10^{14}$~$M_{\odot}$\ for A2199 and $3.05\times 10^{14}$~$M_{\odot}$\ for A496. These estimates are close to ours, even though Mushotzky et al.\ did not apply the {\em ASCA}\/\ PSF correction in their analysis. From the galaxy velocity data, Girardi et al.\ (1998) obtained, at $r=1$ Mpc, masses of $5.4^{+2.4}_{-1.9}\times 10^{14}$~$M_{\odot}$\ for A2199 and $3.5\pm 1.8\times 10^{14}$~$M_{\odot}$\ for A496 (their 68\% errors are multiplied by 1.65 to obtain 90\% intervals). These are consistent (although for A2199, only marginally) with our more accurate values. \begin{figure*}[tb] \begingroup\pst@ifstar\pst@picture(0,11.2)(18.5,21) \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[tl]{0}(0.4,20.7){\epsfxsize=8.5cm \epsffile{2199polycor_bold.ps}} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[tl]{0}(9.5,20.7){\epsfxsize=8.5cm \epsffile{496polycor_bold.ps}} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[bl]{0}( 6.9,19.4){\small A2199} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[bl]{0}(16.0,19.4){\small A496} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[tl]{0}(0,12.4){ \begin{minipage}{18cm} \small\parindent=3.5mm {\sc Fig.}~4.---Polytropic correction to the mass profiles derived under the isothermality assumption (see \S\ref{masssec}). Gray bands correspond to 90\% temperature errors of the observed profiles shown in Fig.\ 2 (excluding the central cooling flow bins). The constant temperature is taken to be equal to the cooling flow-corrected values. X-ray core radii $a_x$ are from the {\em ROSAT}\/\ PSPC data (\S\ref{rossec}), and $r_{1000}$ and $r_{500}$ are same as in Fig.\ 2. \end{minipage} } \endpspicture \end{figure*} \begin{figure*}[tb] \begingroup\pst@ifstar\pst@picture(0,3.1)(18.5,21) \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[tl]{0}(0.4,20.7){\epsfxsize=8.5cm \epsffile{2199polymass_bold.ps}} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[tl]{0}(9.5,20.7){\epsfxsize=8.5cm \epsffile{496polymass_bold.ps}} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[bl]{0}( 2.3,19.4){\small A2199} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[bl]{0}(11.4,19.4){\small A496} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[tl]{0}(0.4,12.9){\epsfxsize=8.5cm \epsffile{2199gasfrac_bold.ps}} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[tl]{0}(9.5,12.9){\epsfxsize=8.5cm \epsffile{496gasfrac_bold.ps}} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[bl]{0}( 2.3,11.6){\small A2199} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[bl]{0}(11.4,11.6){\small A496} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[tl]{0}(0,4.6){ \begin{minipage}{18cm} \small\parindent=3.5mm {\sc Fig.}~5.---Total enclosed mass profiles and the corresponding gas fraction profiles. All values are for $h=0.5$ (the total mass scales as $h^{-1}$ and the gas fraction as $h^{-3/2}$). The 90\% confidence bands (gray) correspond to those in Fig.\ 4. Best-fit profiles of the Navarro, Frenk, \& White form are shown as solid lines in the mass panels. In all panels, dashed lines show the profiles obtained assuming a constant temperature (at its average, cooling flow-corrected value). The $r_{1000}$ values correspond to the best-fit NFW profiles shown here (see also Fig.\ 6). \end{minipage} } \endpspicture \end{figure*} \begin{figure*}[tb] \begingroup\pst@ifstar\pst@picture(0,11.6)(18.5,21) \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[tl]{0}(0.4,20.7){\epsfxsize=8.5cm \epsffile{2199overd_bold.ps}} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[tl]{0}(9.5,20.7){\epsfxsize=8.5cm \epsffile{496overd_bold.ps}} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[bl]{0}( 6.9,19.4){\small A2199} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[bl]{0}(16.0,19.4){\small A496} \def\pst@par{}\pst@ifstar{\@ifnextchar[{\rput@i}{\rput@ii}}[tl]{0}(0,12.4){ \begin{minipage}{18cm} \small\parindent=3.5mm {\sc Fig.}~6.---Average total overdensity within a given radius with respect to the critical density at the clusters' redshifts, calculated for the profiles shown in Fig.\ 5. The overdensity at a given angular distance does not depend on $h$. The {\em ASCA}\/\ data covers $r<25'$. \end{minipage} } \endpspicture \end{figure*} \section{DISCUSSION} \subsection{Uncertainty of Our Mass Values} Although the polytropic model is a good representation of the observed temperature profiles, it does not necessarily cover all possibilities that could be consistent with the data. Since our best-fit model adequately represents the temperature and its gradient, the best-fit mass values should be unbiased (as long as the hydrostatic equilibrium assumption is valid). However, at the extremes of the confidence intervals, some other acceptable temperature models may result in slightly different mass profiles. Therefore, our confidence intervals on mass may be underestimated. A more exhaustive method of mass modeling would be to assume a certain functional form of the dark matter profile and find parameter ranges consistent with the data (e.g., Hughes 1989; Henry, Briel, \& Nulsen 1993; Loewenstein 1994; MV; Nevalainen et al.\ 1999a,b). However, for the relatively high-quality data on A2199 and A496, we have chosen a simpler approach with the polytropic models, without giving undue importance to the formal error estimates. Indeed, given the present accuracy of the temperature and density profiles, the formal statistical uncertainties are already too small to be physically meaningful (e.g., MV). Hydrodynamic simulations suggest that systematic uncertainties in the method itself, such as the possible deviations from spherical symmetry and hydrostatic equilibrium (in the form of significant gas bulk motions), can give rise to {\em rms} mass errors of about 15--30\% (e.g., Evrard, Metzler, \& Navarro 1996; Roettiger et al.\ 1996; see a more detailed discussion in \S4.2 of MV). Those simulations included merging clusters in the statistical sample, so the relaxed clusters A2199 and A496 should have mass errors on the lower side of these estimates. Another source of systematic mass uncertainty is the possible deviation of the measured electron temperature from the local mean plasma temperature that enters the hydrostatic equilibrium equation. Markevitch et al.\ (1996) proposed such nonequality as a possible explanation for an unusually sharp observed temperature gradient in A2163. Later theoretical work (Fox \& Loeb 1997; Ettori \& Fabian 1998; Chi\`eze, Alimi, \& Teyssier 1998; Takizawa 1998) concluded that for relaxed clusters, this effect should not be significant within the radial distances presently accessible for accurate X-ray temperature measurements (about half the virial radius). Therefore, it is safe to assume that the mass values within $r_{1000}$ obtained in this paper are unaffected by this complication. Finally, at these low redshifts, the unknown cluster peculiar velocity may introduce a noticeable distance and mass error (e.g., a 1000 ~km$\;$s$^{-1}$\ velocity would correspond to a 10\% error in the calculated mass). To summarize all of the above, the true uncertainty of the masses of A2199 and A496 is probably greater than our formal $\pm10$\% estimates and perhaps closer to 20--25\% (90\% confidence at $r_{1000}$), and is dominated by systematics. \subsection{The ``Universal'' Mass Profile} NFW have found that radial density profiles of equilibrium clusters in their cosmological simulations can be approximated by a functional form \mbox{$\rho(r) \propto (r/r_s)^{-1}(1+r/r_s)^{-2}$}. This form is a very good description of our observed total mass profiles in the range of radii covered by the temperature data, as shown in Fig.\ 5. Normalizations and scale radii $r_s$ of the NFW profiles were selected to fit the observed mass profiles. For A2199, $r_s=0.18$ Mpc, and for A496, $r_s=0.36$ Mpc.% \footnote{NFW included only dark matter in their simulations and it is unclear whether the inclusion of gas would significantly change the shape of the mass profiles. If we subtract the gas mass from our total mass profiles (assuming, for example, the currently favored value of $h=0.65$), the resulting dark matter profiles are well described by the same functional form with slightly different parameter values.} Extrapolating the best-fit NFW profiles to greater radii, we obtain the NFW's concentration parameter $c\equiv r_{200}/r_s$ of about 10 and 6 for the two clusters, respectively. According to the NFW's simulations, $c$ and the total mass within $r_{200}$ are strongly correlated for a given cosmological model. Our $c$ and $M_{200}$ values agree well with those for several cosmological models considered by NFW, including CDM$\Lambda$ (for that model, our observed masses correspond to $M_{200}/M_*\approx 5-6$ as defined in NFW). An isothermal profile would imply a less concentrated dark matter distribution than that suggested by the NFW simulations (see also Makino, Sasaki, \& Suto 1998). The outer regions of other clusters for which relatively accurate mass profiles were derived from the {\em ASCA}\/\ temperature profiles (e.g., A2256, MV; A3571, Nevalainen et al.\ 1999b) are consistent with the NFW profiles as well, although the constraints are poorer. Thus, the similarity of the gas temperature profiles found by MFSV for nearby relaxed clusters (outside the cooling flow regions) appears to be due to the underlying ``universal'' dark matter profile of the NFW form. Note that we do not consider cooling flow regions due to their unknown temperature structure. As noted in MV and Nevalainen et al.\ (1999a), in those few relaxed clusters without cooling flows where the gas density profile exhibits a flat core all the way to the center (e.g., A401), the dark matter cannot have an NFW central cusp because the gas halo would be convectively unstable (see also Suto, Sasaki, \& Makino 1998). On the other hand, it is likely that in clusters that do have central dark matter cusps, the corresponding dip of the gravitational potential causes the gas density peak and acts as a focus for a cooling flow. \subsection{Mass -- Temperature Scaling} Our mass values within our $r_{1000}$ are a factor of $1.6-1.8$ below the scaling relation between the total mass and emission-weighted average gas temperature derived from the simulations by Evrard et al.\ (1996). The same is true for other clusters (e.g., A2256, MV; A2029, Sarazin, Wise, \& Markevitch 1998; A401, A3571, Nevalainen et al.\ 1999a,b). Note that isothermal mass estimates are also lower than the Evrard et al.\ $M-T$ relation predicts. Given the agreement of our observed total (or dark) mass profile with the ``universal'' profile from NFW as well as from Evrard et al., the main source of this discrepancy apparently lies in the gas density and temperature distributions. Indeed, as noted by MFSV, simulations predict a less steep temperature decline than observed, and steeper gas density profiles than observed (e.g., Vikhlinin et al.\ 1999). Both these effects have the sign needed to cause the $M-T$ discrepancy. Because most of the cluster X-ray emission originates in the central region, simulations that do not sufficiently resolve the cluster core may predict an incorrect (apparently too low) emission-weighted gas temperature. For example, the Evrard et al.\ (1996) simulations have a resolution of 0.2--0.3 Mpc, comparable to a typical cluster core radius. On the other hand, the hydrostatic mass measurement can underestimate the true total mass if, for example, there is significant gas turbulence. Indeed, simulations suggest that there may be residual turbulence even in an apparently relaxed cluster (e.g., Evrard et al.\ 1996; Norman \& Bryan 1998) resulting in a 10--15\% underestimate of the mass within $r\sim r_{1000}$. \subsection{Implications for the X-ray--Lensing Mass Discrepancy} A2199 and A496 are representative examples of relaxed clusters and the difference in their X-ray mass estimates using the measured temperature profile from the isothermal estimates, shown in Fig.\ 4, generally applies to other such clusters (see MV and MFSV). The upward revision of the mass estimate in the inner part has one important implication, the convergence of the X-ray and gravitational lensing mass estimates in the cluster central regions. The strong lensing mass values (that usually correspond to $r\lesssim 0.2$ Mpc) often exceed by a factor of 2--3 the X-ray estimates made under the assumptions of isothermality and a typical $\beta$-model density profile (e.g., Loeb \& Mao 1994; Miralda-Escud\'e \& Babul 1995; Tyson \& Fischer 1995). In many cases, the lensing analysis is likely to overestimate the mass as a result of substructure or projection (e.g., Bartelmann 1995). On the X-ray side, some clusters are undergoing mergers and the hydrostatic equilibrium method may give a wrong mass value. For those clusters which are relaxed, the low-resolution X-ray image analysis may underestimate the gas density gradient at small radii typical of the lensing measurements and may be responsible for part of the disagreement (e.g., Markevitch 1997; Allen 1998). If the cluster has a strong cooling flow, the overall temperature can be significantly underestimated if no allowance for the cool component is made (Allen 1998), although for most clusters this correction is within $\sim 20$\% (MFSV). Still, in many cases these effects alone are not sufficient to account for the mass discrepancy. It has been suggested, e.g., by Miralda-Escud\'e \& Babul (1995) that a gas temperature gradient could explain the discrepancy for the distant non-cooling-flow cluster A2218. A temperature decline with radius has indeed been observed in A2218 by Loewenstein (1997) and Cannon, Ponman, \& Hobbs (1999) using {\em ASCA}\/, while MFSV find that such a declining profile is common among nearby clusters. The analysis in \S\ref{masssec} has shown that within the core radius, the commonly observed temperature gradient implies a mass that is higher than the isothermal estimate by a factor of $\gtrsim 1.5$. The reference isothermal estimate uses a cooling flow-corrected temperature, so this effect is in addition to the cooling flow-related mass correction of Allen (1998). If a similar temperature gradient is common in more distant clusters, this effect, together with others mentioned above, effectively resolves the mass discrepancy. This seems to obviate the need for more exotic causes, such as a significant magnetic field pressure or strong turbulence within cluster cores (Loeb \& Mao 1994). \subsection{Gas Fraction} The lower panels in Fig.\ 5 show the gas mass fraction a function of radius for the two clusters. At $r_{1000}$, we obtain similar values of $f_{\rm gas}=0.161\pm 0.014$ and $0.158\pm 0.017$ for the two clusters, respectively. These values are consistent with those for A2256, $0.14\pm 0.01$ at $r_{1000}$, obtained by MV using an {\em ASCA}\/\ temperature profile, and for A401 ($0.18^{+0.02}_{-0.04}$) and A3571 ($0.16^{+0.03}_{-0.01}$) from Nevalainen et al.\ (1999a,b). Our values are also similar to the median values for large samples of clusters analyzed using the isothermal assumption: $f_{\rm gas}=0.168$ from Ettori \& Fabian (1999, scaled to $r_{1000}$), and $f_{\rm gas}=0.160$ from Mohr, Mathiesen, \& Evrard (1999, for clusters cooler than 5 keV). The latter similarity is due to the fact that the effect of the radial temperature decline on mass is small at this radius; at greater radii, the isothermal analysis underestimates the gas fraction as Fig.\ 5 shows. The values of the cluster gas fraction from X-ray analysis are often used to place constraints on the cosmological density parameter ($\Omega_0\lesssim 0.3$), under the assumption that $f_{\rm gas}$ in clusters is representative of the Universe as a whole (White et al.\ 1993 and many later works). However, $f_{\rm gas}$ increases with radius even if one assumes a constant gas temperature (e.g., David, Jones, \& Forman 1995; Ettori \& Fabian 1999), and the true increase is steeper as seen in clusters with measured temperature profiles. Fig.\ 5 shows that $f_{\rm gas}$ increases by a factor of 3 between the X-ray core radius and $r_{1000}$ and does not show any evidence of flattening at large radii and hence asymptotically reaching a universal value. Although at some radius, the cluster must merge continuously into the infalling matter with a cosmic mix of components, that presumably happens at the infall shock radius of $\sim 2-3\, r_{1000}$, well beyond the region presently accessible to accurate measurements. Note that both the dark matter and the gas mass within a given radius, and thus $f_{\rm gas}$, are dominated by the contribution at large radii. Cosmological simulations suggest that at smaller radii, a deviation from the universal $f_{\rm gas}$ value is not large (e.g., Frenk et al.\ 1996). However, at the present stage, the simulations do not accurately reproduce the observed gas density, temperature and $f_{\rm gas}$ profiles (e.g., MFSV; Vikhlinin et al.\ 1999). We conclude that our $f_{\rm gas}$ values for A2199 and A496 are consistent with the constraints on $\Omega_0$ derived in earlier works (e.g., Ettori \& Fabian 1999; Mohr et al.\ 1999), but caution that such estimates at present involve a large extrapolation. Future observatories {\em Chandra}\/\ and {\em XMM}\/\ will be capable of studying the cluster outermost regions and possibly determining the asymptotic value of $f_{\rm gas}$. \section{SUMMARY} The {\em ASCA}\/\ gas temperature maps and radial profiles for A2199 and A496 indicate that these systems are representative examples of relaxed, moderately massive clusters. Our high quality temperature data imply total mass profiles that are in good agreement with the NFW simulated ``universal'' profile over the range of radii covered by the data ($0.1\,{\rm Mpc}< r < r_{1000}\approx 1\,{\rm Mpc}$). Because the temperature profiles of these two clusters are similar to the average profile for a large sample of nearby clusters in MFSV, this agreement indicates that the NFW profile is indeed common in nearby clusters. The upward revision of the total mass at small radii, by a factor of $\gtrsim 1.5$ compared to an isothermal analysis, may reconcile X-ray and strong lensing mass estimates in distant clusters. The observed mass profile also implies a gas mass fraction profile steeply rising with radius. While our $f_{\rm gas}$ values at $r_{1000}$ support earlier upper limits on $\Omega_0$, the steep increase of $f_{\rm gas}$ with radius, not anticipated by most cluster simulations, suggests that we may not yet have correctly determined the universal baryon fraction and caution is needed in such analysis. \acknowledgments This work was supported by NASA contracts and grants NAS8-39073, NAG5-3057, NAG5-4516, NAG5-8390, and by the Smithsonian Institution.
\section{Introduction} \label{sec:intro} Several properties of gamma-ray burst (GRB) afterglows can be well explained by `fireball' models, in which a relativistically expanding shock front, caused by an energetic explosion in a central compact region, sweeps up the surrounding medium and accelerates electrons in a strong synchrotron emitting shock (M\'eszar\'os and Rees 1994, Wijers, Rees and M\'eszar\'os 1997; Sari, Piran and Narayan 1998; Galama et al. 1998a). The emission shows a gradual softening with time, corresponding to a decrease of the Lorentz factor of the outflow. Most X-ray and optical/infrared (IR) afterglows display a power law decay (except GRB\,980425, which is most likely associated with the peculiar supernova SN\,1998bw; Galama et al. 1998b). Spectral transitions in the optical/IR have been detected for the afterglows of GRB\,970508 (Galama et al. 1998a; Wijers and Galama 1998) and GRB\,971214 (Ramaprakash et al. 1998). These have been explained by the passage through the optical/IR waveband of the cooling break, at $\nu_{\rm c}$ (for GRB\,970508) and the peak of the spectrum, at $\nu_{\rm m}$ (for GRB\,971214); see Sari, Piran \& Narayan (1998) and Wijers \& Galama (1998) for their definition. For GRB\,970508 the observed break is in excellent agreement with such `fireball' models, while for GRB\,971214 an exponential extinction has been invoked to explain the discrepancy between the expected and observed spectral index. GRB\,980703 was detected on July 3.182 UT with the All Sky Monitor (ASM) on the {\it Rossi X-ray Timing Explorer} (RXTE; Levine et al. 1998), the {\it Burst And Transient Source Experiment} (BATSE, trigger No. 6891; Kippen et al. 1998), {\it Beppo}\,SAX (Amati et al. 1998) and {\it Ulysses} (Hurley et al. 1998). The burst as seen by BATSE consisted of two pulses, each lasting approximately 100 sec., with a total duration of about 400 sec. (Kippen et al. 1998). The first pulse had significant sub-structure, whereas the second, weaker episode was relatively smooth. This double-peak morphology has also been seen with the {\it Beppo}\,SAX {\it Gamma Ray Burst Monitor} (Amati et al. 1998). BATSE measured a peak flux of (1.9 $\pm$ 0.1) $\times$ 10$^{-6}$ erg cm$^{-2}$ s$^{-1}$ (25 - 1000 keV) and fluence of (4.6 $\pm$ 0.4) $\times$10$^{-5}$ erg cm$^{-2}$ ($>$ 20 keV), consistent with the {\it Beppo}\,SAX GRBM measurement. A time resolved spectral analysis of the burst will be presented elsewhere (Koshut et al. 1999). Observations with the Narrow-Field Instruments (NFIs) of {\it Beppo}\,SAX showed a previously unknown X-ray source (Galama et al. 1998c) inside both the ASM error box and the InterPlanetary Network annulus (Hurley et al. 1998). Frail et al. (1998a; see also Zapatero Osorio et al. 1998) subsequently reported the discovery of a radio (6 cm) and optical (R-band) counterpart to GRB\,980703. Here we report X-ray (0.1-10 keV), optical (VRI), and infrared (JHK) follow-up observations of GRB\,980703. In \S \ref{sec:nfi} we report our NFI X-ray observations of the ASM error box, and \S \ref{sec:opir} is devoted to the description and results of our spectroscopic and photometric optical/IR monitoring campaign. We discuss the results of these observations in \S \ref{sec:dis}. \section{X-ray observations} \label{sec:nfi} We observed the ASM error box of GRB\,980703 with the {\it Beppo}\,SAX Low- and Medium Energy Concentrator Spectrometers (LECS, 0.1-10 keV, Parmar et al. 1997; MECS, 2-10 keV, Boella et al. 1997) on July 4.10-5.08 UT (starting 22 hrs after the burst) and on July 7.78-8.71 UT. The LECS and MECS data show a previously unknown X-ray source 1SAX\,J2359.1+0835 at R.A. = 23$^{\rm h}$59$^{\rm m}$06\fs8, Decl. = +08\degr35\arcmin45\hbox{$^{\prime\prime}$} (equinox J2000.0), with an error radius of 50\hbox{$^{\prime\prime}$}. The field also contains the sources 1SAX J2359.9+0834 at R.A. = 23$^{\rm h}$59$^{\rm m}$59$^{\rm s}$, Decl.= +08\degr34\arcmin03\hbox{$^{\prime\prime}$}, and 1SAX J0000.1+0817 at R.A. = 00$^{\rm h}$00$^{\rm m}$04$^{\rm s}$, Decl.=+08\degr17\arcmin14\hbox{$^{\prime\prime}$}. Both are outside the ASM error box, do not show any variability and coincide with the known ROSAT sources, 1RXS J235959.1+083355 and 1RXS J000007.0+081653, respectively. We extracted 1SAX J2359.1+0835 data at the best fit centroids with radii of 8$^{'}$ (LECS) and 2$^{'}$ (MECS). To analyze the spectrum we binned the data into channels, such that each contained at least 20 counts. Using the separate standard background files of the spectrometers, we simultaneously fitted the LECS and MECS data of July 4-5 UT, with a power law spectrum and a host galaxy absorption cut-off, using a redshift of $z = 0.966$ (Djorgovski et al. 1998; see also \S \ref{sec:opir} of this paper). In the fit we fixed the Galactic foreground absorption at $N_{\rm H}$ = 3.4 $\times$ 10$^{20}$ cm$^{-2}$ (A$_V$ = 0.19, as inferred from the dust maps of Schlegel, Finkbeiner \& Davis 1998\footnote{see http://astro.berkeley.edu/davis/dust/index.html}, and the A$_V$-N$_{\rm H}$ relation of Predehl \& Schmitt 1995). The average spectrum can be modelled with a photon index $\Gamma = 2.51 \pm 0.32$ and a host galaxy column density $N_{\rm H}({\rm host})$ = \gpm{3.6}{2.2}{1.3} $\times$ 10$^{22}$ cm$^{-2}$, corresponding to A$_{V}({\rm host})$ = \gpm{20.2}{12.3}{7.3} (local to the absorber). Modelling the spectrum without $N_{\rm H}({\rm host})$ results in a very poor fit. We did not account for a possible small position dependent error in the relative flux normalizations between the LECS and MECS, which is only a few percent near the center of the image. A change of 25\% in the assumed Galactic foreground absorption does not affect the output values of the fit parameters. For the second epoch of NFI observations we kept the position fixed at the position determined from the first epoch; we find F$_{X} <$ 1.1 $\times$ 10$^{-13}$ erg cm$^{-2}$ s$^{-1}$ (3$\sigma$). Fitting a power law model to the light curve including the upper limit, we obtain $\alpha <$ --0.91 for the decay index. The X-ray light curve is shown in Fig \ref{fig:xray}. We checked for the presence of the 6.4 keV K line at the redshifted energy of 3.26 keV, but do not detect it. The upper limit on its flux is 8.3 $\times$ 10$^{-6}$ photons cm$^{-2}$ s$^{-1}$ (90\% confidence level), corresponding to an equivalent width of 532 eV in the observer's frame. \section{Optical and infrared observations} \label{sec:opir} We observed the field of GRB\,980703 with the Wise Observatory 1-m telescope (in I); the 3.5-m New Technology Telescope (NTT; in V, I and H), the 2.2-m (in H and Ks) and Dutch 90-cm (in gunn i) telescopes of ESO (La Silla); the CTIO 0.9-m telescope (in R) and UKIRT (in H, J and K). The optical images were bias-subtracted and flat-fielded in the standard fashion. The infrared frames were reduced by first removing bad pixels and combining about five frames around each object image to obtain a sky image. This sky image was then subtracted after scaling it to the object image level; the resulting image was flat-fielded. Four reference stars were used to obtain the differential magnitude of the optical transient (OT) in each frame. These stars were calibrated by observing the standard stars PG2331+055 (in V and I; Landolt 1992), FS2 and FS32 (in J, H and K; Casali \& Hawarden 1992). We used the R-band calibration of Rhoads et al. (1998). The offsets in right ascension and declination from the OT, and the apparent standard magnitudes outside the Earth's atmosphere of the reference stars are listed in Table \ref{tab:refstars}. The light curves of the OT are shown in Fig. \ref{fig:mags.ps} (see Table \ref{tab:log} for a list of the magnitudes). In view of the flattening of the light curves after $t \sim 5$ days we fitted a model $F_{\nu} = F_{0} \cdot t^{\alpha} + F_{\rm gal}$ to our own I and H band light curves (in these bands we have sufficient data for a free parameter fit). Here $t$ is the time since the burst in days, and $F_{\rm gal}$ is the flux of the underlying host galaxy. The values for $m_{0}= -2.5 \cdot {\rm log}\,(F_{0})+C$, the decay index $\alpha$ and $m_{\rm gal}= -2.5 \cdot {\rm log}\,(F_{\rm gal})+C$, are listed in Table \ref{tab:mag}. The photometric calibration, which determines $C$, has been taken from Bessell (1979) for V, R and I and Bessell \& Brett (1988) for J, H and K. The weighted mean value of $\alpha$ for the I and H-bands equals $-1.61 \pm 0.12$, while an I and H-band joint fit, with a single power law decay index, gives $\alpha = -1.63 \pm 0.12$ ($\chi$ = 15.4/16). For the V, R, J and K bands we fixed $\alpha$ at --1.61, included also data from the literature, and fitted $m_0$ and $m_{\rm gal}$. The fits are shown as solid lines in Fig. \ref{fig:mags.ps}. We note that when we fit all three parameters to the R band data (mainly data from the literature) we obtain a temporal slope of $-1.94 \pm 0.22$, consistent with the adopted value of $-1.61 \pm 0.12$. However, we do not include this value in the average, since the R band magnitudes are taken from the literature and thus not consistently measured. For four epochs we have reconstructed the spectral flux distribution of the OT (times t$_1$, t$_2$, t$_3$ and t$_4$ in Fig. \ref{fig:mags.ps}, corresponding to 1998 July 4.4, 6.4, 7.6 and 8.4 respectively). The host galaxy flux, obtained from the fits to the light curves, was subtracted. We corrected the OT fluxes for Galactic foreground absorption (A$_V$ = 0.19). If more than one value per filter was available around the central time of the epoch, we took their weighted average. All values were brought to the same epoch by applying a correction using the slope of the fitted light curve. For the first epoch (t$_1$) we fitted the resulting spectral flux distribution with a power law, $F_\nu \propto \nu^{\beta}$, and find $\beta = -2.71 \pm 0.12$. We took three 1800 sec. spectra of the OT with the NTT, around July 8.38 UT. The \#3 grism that was used has a blaze wavelength of 4600 $\rm\AA$ and a dispersion of 2.3 \AA/pixel. The slit width was set at 1\hbox{$^{\prime\prime}$}. The three spectra were bias-subtracted and flat-fielded in the usual way. The co-added spectrum was then extracted the same way as the standard star Feige 110. We wavelength calibrated the spectrum with a Helium/Argon lamp spectrum, with a residual error of 0.03 $\rm\AA$. The spectrum was flux calibrated with the standard star Feige 110 (Massey et al. 1988). We estimate the accuracy of the flux calibration to be 10\%. The wavelength and flux calibrated spectrum shows one clear emission line at 7330.43 $\pm$ 0.14 $\rm\AA$ with a flux of 3.6 $\pm$ 0.4 $\times 10^{-16}$ erg cm$^{-2}$ s$^{-1}$. At the redshift $z=0.966$ determined by Djorgovski et al. (1998) this is the $\lambda$\,3727 line of [O\,II]; our wavelength measurement corresponds to $z=0.9665\pm 0.0005$. \section{Discussion} \label{sec:dis} From the optical/IR light curves presented in \S 3 we have obtained an average power law decay constant $\alpha$ = $-1.61\pm 0.12$. This value is consistent with the ones derived by Bloom et al. (1998) ($\alpha_{\rm R} = -1.22 \pm 0.35$, and $\alpha_{\rm I} = -1.12\pm 0.35$) and Castro-Tirado et al. (1999) ($\alpha_{\rm R} = -1.39 \pm 0.3$, and $\alpha_{\rm H} = -1.43\pm 0.11$). If we make the assumption that the OT emission is due to synchrotron radiation from electrons with a power law energy distribution (with index $p$), one expects a relation between $p$, the spectral slope $\beta$, and the decay constant $\alpha$ (Sari, Piran and Narayan 1998). We assume that our observations are situated in the slow cooling, low frequency regime (e.g. for GRB\,970508 this was already the case after 500 sec.; Galama et al. 1998a). One must distinguish two cases: (i) both the peak frequency $\nu_{\rm m}$ and the cooling frequency $\nu_{\rm c}$ are below the optical/IR waveband. Then $p = (-4\alpha +2)/3 = 2.81 \pm 0.16$ and $\beta = -p/2 = -1.41 \pm 0.08$, (ii) $\nu_{\rm m}$ has passed the optical/IR waveband, but $\nu_{\rm c}$ has not yet. In that case $p = (-4\alpha +3)/3 = 3.15 \pm 0.16$ and $\beta = -(p-1)/2 = -1.07 \pm 0.08$. In both cases the expected value of $\beta$ is inconsistent with the observed $\beta = -2.71 \pm 0.12$. Following Ramaprakash et al. (1998) we assume that the discrepancy is caused by host galaxy extinction (note that we have already corrected the OT fluxes for Galactic foreground absorption). To determine the host galaxy absorption we first blueshifted the OT flux distribution to the host galaxy rest frame (using $z$ = 0.966), and then applied an extinction correction using the Galactic extinction curve of Cardelli, Clayton \& Mathis (1989), to obtain the expected spectral slope $\beta$. For epoch t$_1$ (July 4.4 UT), we obtain $A_V = 1.15 \pm 0.13$ and $A_V = 1.45 \pm 0.13$ for the cases (i) and (ii), respectively (see Fig. \ref{fig:bbspec14}). In case (i) we find that an extrapolation of the optical flux distribution to higher frequencies predicts an X-ray flux that is significantly below the observed value, whereas in case (ii) the extrapolated and observed values are in excellent agreement. The mismatch in case (i) is in a direction that cannot be interpreted in terms of the presence of a cooling break between the optical and X-ray wavebands. When we include the X-ray data point in the fit to obtain a more accurate determination of A$_V$, we find A$_V = 1.50 \pm 0.11$, and $\beta = -1.013 \pm 0.016$. We conclude that the optical/IR range is not yet in the cooling regime, and so $p$ = 3.15 $\pm$ 0.16. Where would the cooling frequency, $\nu_{\rm c}$, be located? The X-ray photon index, $\Gamma = 2.51 \pm 0.32$, corresponding to a spectral slope of $\beta = -1.51 \pm 0.32$ suggests that perhaps the X-ray waveband is just in the cooling regime, in which case the expected local X-ray slope would be $\beta = -p/2 = -1.57 \pm 0.08$, while if not in the cooling regime, it would be $\beta = -(p-1)/2 = -1.07 \pm 0.08$. However, the large error on the measured X-ray spectral slope would also allow the cooling break to be above 2-10 keV. We estimate the (2$\sigma$) lower limit to the cooling frequency to be $\nu_{\rm c}$ $>$ 1.3 $\times$ 10$^{17}$ Hz ($h \nu_{\rm c} >$ 0.5 keV). We performed the same analysis for the other epoch (t$_4$) with X-ray data (see Fig. \ref{fig:bbspec14}). At this epoch, the X-ray upper limit does not allow us to discriminate between the two cases. However, we can still estimate a lower limit to the cooling break from its time dependence: $\nu_{\rm c} \propto t^{-1/2}$, which would allow the break to drop to $\nu_{\rm c}$ $>$ 6.3 $\times$ 10$^{16}$ Hz only, between epoch t$_1$ and t$_4$. On the basis of our analysis we conclude that there is no strong evidence for a cooling break between the optical/IR and the 2-10 keV passband before 1998 July 8.4 UT. This conclusion is at variance with the inference of Bloom et al. (1998), who infer from their fits that there is a cooling break at about $10^{17}$\,Hz. Upon closer inspection, there is no real disagreement: Bloom et al. found a slightly shallower temporal decay, and therefore a bluer spectrum of the afterglow, which causes their extrapolated optical spectrum to fall above the X-ray point. However, their error of 0.35 on the temporal decay leads to an error of 0.24 on their predicted spectral slope, and this means that a 1$\sigma$ steeper slope in their Fig. 2 would be consistent with no detected cooling break. Assuming that the spectral slope ($\beta = -1.013 \pm 0.016$) did not change during the time spanned by the four epochs t$_1$ - t$_4$ (as suggested by the lack of evidence for a break in the light curve during this timespan) we have derived the V-band extinction A$_V$ as a function of time: A$_V$= 1.50 $\pm$ 0.11, 1.38 $\pm$ 0.35, 0.84 $\pm$ 0.29 and 0.90 $\pm$ 0.25 for the epochs 1 through 4, respectively. Fitting a straight line through these, we obtain a slope of --0.16 $\pm$ 0.06, i.e. not consistent with zero at the 98.8\% confidence level. Such a decrease of the optical extinction, A$_V$, might be caused by ionization of the surrounding medium (Perna and Loeb 1997). The V-band extinction A$_V$ = \gpm{20.2}{12.3}{7.3}, derived from the host galaxy N$_H$ fit to the MECS and LECS data (July 4-5 UT) is not in agreement with A$_V = 1.50 \pm 0.11$ as derived from the fit from the optical spectral flux distribution. This may be due to a different dust to gas ratio for the host galaxy of GRB\,980703, or a higher abundance than normal of the elements that cause the X-ray absorption. With the above derived constraint on $\nu_{\rm c}$ we can partially reconstruct the broad-band flux distribution of the afterglow of GRB\,980703: from the radio observations of Frail et al. (1998b) at 1.4, 4.86 and 8.46 GHz, we determine the self-absorption frequency $\nu_{\rm a}$ and its flux $F_{\nu_{\rm a}}$ from the fit $F_{\nu} = F_{\nu_{\rm a}}(\nu/\nu_{\rm a})^{2}(1-\exp[-(\nu/\nu_{\rm a})^{-5/3}])$ to the low-energy part of the spectrum (e.g. Granot, Piran and Sari 1998). We have used averages of the 1.4 and 4.86 GHz observations to obtain best estimates of the radio flux densities as particularly those frequencies suffer from large fluctuations due to interstellar scintillation (Frail et al. 1998b). We find $\nu_{\rm a} = 3.68 \pm 0.33$ GHz and $F_{\nu_{\rm a}}$ = 789 $\pm$ 42 $\mu$Jy. (The fit is shown in Fig. \ref{fig:specfit}.) The intersection of the extrapolation from the low-frequency to the high-frequency fit gives a rough estimate of the peak frequency, $\nu_{\rm m} \sim 4\times10^{12}$ Hz, and of the peak flux, $F_{\nu_{\rm m}} \sim 8$ mJy (see Fig. \ref{fig:specfit}). By assuming such a simple broken power law spectrum the peak flux density will likely be overestimated (realistic spectra are rounder at the peak); it is clear from Fig. \ref{fig:specfit} that $1 < F_{\nu_{\rm m}} < 8$ mJy. Following the analysis of Wijers and Galama (1998) we have determined the following intrinsic fireball properties: (i) the energy of the blast wave per unit solid angle: ${\cal E} >$ 5 $\times$ 10$^{52}$ erg/(4$\pi$ sterad), (ii) the ambient density: $n > 1.1$ nucleons cm$^{-3}$, (iii) the percentage of the nucleon energy density in electrons: $\epsilon_{\rm e} > 0.13$, and (iv) in the magnetic field: $\epsilon_{B} < 6\times10^{-5}$. The very low energy in the magnetic field, $\epsilon_B$, is a natural reflection of the high frequency of the cooling break $\nu_{\rm c}$. We have compared this afterglow spectrum with that of GRB\,970508. Scaling the latter in time according to $\nu_{\rm a} \propto t^{0}$, $\nu_{\rm m} \propto t^{-3/2}$ and $\nu_{\rm c} \propto t^{-1/2}$, the results of GRB\,970508 (Galama et al. 1998a; see also Granot, Piran and Sari 1998) would correspond to $\nu_{\rm a} \sim 2.3$ GHz, $\nu_{\rm m} = 2.8\times10^{12}$ Hz, $\nu_{\rm c} = 4.8\times10^{14}$ Hz, and $F_{\nu_{\rm m}}$ = 1.3 mJy. In this calculation we have corrected for the effect of redshift (see Wijers and Galama 1998) such that the values represent GRB\,970508, were it at the redshift of GRB\,980703 and observed 1.2 days after the event. The greatest difference between the two bursts is in the location of the cooling frequency, $\nu_{\rm c}$. The observations on the United Kingdom Infrared Telescope, which is operated by the Joint Astronomy Centre on behalf of the U.K. Particle Physics and Astronomy Research Council, were carried out in Service mode by UKIRT staff. The BeppoSAX satellite is a joint Italian and Dutch programme. PMV is supported by the NWO Spinoza grant. TJG is supported through a grant from NFRA under contract 781.76.011. CK acknowledges support from NASA grant NAG 5-2560. TO acknowledges an ESA Fellowship. KH is grateful for support under JPL Contract 958056 for Ulysses, and under NASA grant NAG 5-1560 for IPN operations. \references Amati, L., et al. 1998, GCN circular, 146 \\ Bessell, M.S. 1979, PASP, 91, 589 \\ Bessell, M.S. \& Brett, J.M. 1988, PASP, 100, 1134 \\ Bloom, J.S., et al. 1998, \apjl, 508, L21 \\ Boella, G., et al. 1997, A\&AS, 122, 299 \\ Cardelli, J.A., Clayton, G.C. \& Mathis, J.S. 1989, \apj, 345, 245 \\ Castro-Tirado, A.J. et al. 1999, \apjl, 511, L85 \\ Casali, M.M. \& Hawarden, T.G. 1992, JCMT-UKIRT Newsletter, No. 3, 33\\ Djorgovski, S.G., et al. 1998, \apjl, 508, L17\\ Frail, D.A., et al. 1998a, GCN circular, 128 \\ Frail, D.A., et al. 1998b, GCN circular, 141 \\ Galama, T.J., et al., 1998a, \apjl, 500, L97 \\ Galama, T.J., et al. 1998b, \nat, 395, 670 \\ Galama, T.J., et al. 1998c, GCN circular, 127 \\ Granot, J., Sari, T. and Sari, R. 1998, submitted, astro-ph/9808007\\ Henden, A.A., et al. 1998, GCN circular, 131 \\ Hurley, K., et al. 1998, GCN circular, 125 \\ Kippen, M., et al. 1998, GCN circular, 143 \\ Koshut, T., et al. 1999, in preparation\\ Landolt, A.U. 1992, AJ 104, 340 \\ Levine, A., et al. 1998, \iaucirc, 6966 \\ Massey, P., et al. 1988, \apj, 328, 315\\ M\'eszar\'os, P. \& Rees, M.J. 1994, MNRAS, 269, L41 \\ Parmar, A.N., et al. 1997, A\&AS, 122, 309 \\ Pedersen, H., et al. 1998, GCN circular, 142 \\ Perna, R. \& Loeb, A. 1998, \apj, 501, 467\\ Predehl, P. and Schmitt, J.H.M.M. 1995, A\&A, 293, 889 \\ Ramaprakash, A.N., et al. 1998, \nat, 393, 43 \\ Rhoads, J., et al. 1998, GCN circular, 144 \\ Sari, R., Piran, T., Narayan, R. 1998, \apjl, 497, L17 \\ Schlegel, D.J., Finkbeiner, D.P. and Davis, M. 1998, \apj, 500, 525 \\ Sokolov, V., et al. 1998, GCN circular, 147 \\ Wijers, R.A.M.J. and Galama, T.J. 1998, in press, \apj \\ Wijers, R.A.M.J., Rees, M.J. and M\'eszar\'os, P. 1997, \mnras, 288, L51 \\ Zapatero Osorio, M.R., et al. 1998, \iaucirc, 6967 \\ \begin{figure} \caption[]{The 2-10 keV light curve of GRB\,980703. Time and flux are on a logarithmic scale. \label{fig:xray}} \centerline{\psfig{figure=FIG1.ps,angle=0,width=8.8cm}} \end{figure} \begin{figure} \caption[]{V, R, I, J, H and K light curves of GRB\,980703. The filled symbols denote our data, while the open symbols represent data taken from the literature (Zapatero Osorio et al. 1998; Rhoads et al. 1998; Henden et al. 1998; Bloom et al. 1998; Pedersen et al. 1998; Djorgovski et al. 1998; Sokolov et al. 1998 \& private communication). For each filter a power law model plus a constant: $F_{\nu} = F_{0} \cdot t^{\alpha} + F_{\rm gal}$ is fitted (solid lines). The fit parameters are listed in Table \ref{tab:mag}. The times t$_1$ - t$_4$, at which we have reconstructed the spectral flux distribution of the OT, are indicated by the dashed lines. \label{fig:mags.ps}} \centerline{\psfig{figure=FIG2.ps,angle=-90,width=8.8cm}} \end{figure} \begin{figure} \centerline{\psfig{figure=FIG3a.ps,angle=-90,width=8.8cm}\psfig{figure=FIG3b.ps,angle=-90,width=8.8cm}} \caption[]{Left figure: Broad-band spectrum of GRB\,980703 at July 4.4 UT (i.e., at t$_1$ in Fig. \ref{fig:mags.ps}). The open symbols are the R, I and H OT fluxes (interpolated to July 4.4, corrected for Galactic foreground absorption and the host galaxy flux) and the MECS (2-10 keV) de-absorbed flux (the absorption correction is 7\%). The filled symbols are obtained by invoking an interstellar extinction, A$_V$, to force the slope of the data points to take on the two possible theoretical spectral slopes. The two slopes $\beta$ and their 1$\sigma$ errors are indicated by the solid and dotted lines. Right figure: Broad-band spectrum of GRB\,980703 at July 8.4 UT (i.e., at t$_4$ in Fig. \ref{fig:mags.ps}). The open symbols are the V, R, I, J, H and K OT fluxes and the MECS (2-10 keV) de-absorbed 3$\sigma$ upper limit. \label{fig:bbspec14}} \end{figure} \begin{figure} \centerline{\psfig{figure=FIG4.ps,angle=-90,width=8.8cm}} \caption[]{Radio to X-ray spectrum of GRB\,980703 at July 4.4 UT (i.e., at t$_1$ in Fig. \ref{fig:mags.ps}). Shown are data from Fig. (\ref{fig:bbspec14}) as well as 1.4, 4.86 and 8.46 GHz observations from Frail et al. (1998b). The fit $F_{\nu} = F_{\nu_{\rm a}}(\nu/\nu_{\rm a})^{2}(1-\exp[-(\nu/\nu_{\rm a})^{-5/3}])$ to the low-energy part of the spectrum with $\nu_{\rm a} = 3.68 \pm 0.33$ GHz and $F_{\nu_{\rm a}}$ = 789 $\pm$ 42 $\mu$Jy is shown by the dotted line. The best fit to the optical/IR and X-ray data is also shown. \label{fig:specfit}} \end{figure} \normalsize \begin{table} \begin{minipage}{15cm} \begin{tabular}{lrrrr} \hline filter & star 1 & star 2 & star 3 & star 4 \\ \hline $\Delta$ R.A.($\hbox{$^{\prime\prime}$}$) & --18.0 & --11.7 & --8.0 & --13.9 \\ $\Delta$ Decl.($\hbox{$^{\prime\prime}$}$) & 3.9 & --9.7 & --21.7 & --31.8 \\ \hline V & 21.33 $\pm$ 0.06 & 17.02 $\pm$ 0.05 & 22.06 $\pm$ 0.08 & 22.68 $\pm$ 0.12 \\ R & 20.39 $\pm$ 0.04 & 16.64 $\pm$ 0.02 & 20.72 $\pm$ 0.05 & \\ I & 19.55 $\pm$ 0.07 & 16.30 $\pm$ 0.05 & 19.21 $\pm$ 0.06 & 20.01 $\pm$ 0.08 \\ J & 18.45 $\pm$ 0.13 & 15.75 $\pm$ 0.11 & 17.52 $\pm$ 0.12 & 18.25 $\pm$ 0.13 \\ H & 17.85 $\pm$ 0.12 & 15.44 $\pm$ 0.10 & 16.96 $\pm$ 0.11 & 17.69 $\pm$ 0.12 \\ K & 17.71 $\pm$ 0.13 & 15.41 $\pm$ 0.12 & 16.72 $\pm$ 0.12 & 17.52 $\pm$ 0.13 \\ \end{tabular} \end{minipage} \caption[]{The magnitudes and offset from the OT in arc seconds of the four comparison stars used. The error is the quadratic average of the measurement error (Poisson noise) and a constant offset, which we estimate to be 0.05 for the optical passbands and 0.1 for the infrared filters. \label{tab:refstars}} \end{table} \scriptsize \begin{table} \begin{minipage}{15cm} \begin{tabular}{rccrcl} \hline UT date & magnitude & filter & exp. time & seeing & telescope/reference \\ (1998 July) & & & (seconds) & (\hbox{$^{\prime\prime}$}) & \\ \hline 4.059 & 20.07 $\pm$ 0.19 & I & 2100 & 2.39 & Wise 1-m \\ 4.347 & 20.43 $\pm$ 0.04 & I & 900 & 1.16 & ESO NTT (EMMI) \\ 4.359 & 20.49 $\pm$ 0.03 & I & 900 & 1.10 & ESO NTT (EMMI) \\ 4.372 & 20.54 $\pm$ 0.03 & I & 900 & 1.14 & ESO NTT (EMMI) \\ 4.383 & 20.55 $\pm$ 0.04 & I & 900 & 1.02 & ESO NTT (EMMI) \\ 4.439 & 17.61 $\pm$ 0.04 & H & 810 & & ESO NTT (SOFI) \\ 5.059 & 20.73 $\pm$ 0.29 & I & 1800 & 3.09 & Wise 1-m \\ 5.339 & 21.84 $\pm$ 0.08 & R & 3600 & 1.86 & CTIO 0.9-m \\ 6.395 & 18.86 $\pm$ 0.14 & H & 540 & & ESO NTT (SOFI) \\ 7.609 & 19.25 $\pm$ 0.12 & H & 1200 & & UKIRT \\ 7.622 & 18.36 $\pm$ 0.13 & K & 600 & & UKIRT \\ 8.361 & 19.27 $\pm$ 0.22 & H & 2700 & & ESO 2.2m \\ 8.375 & 21.60 $\pm$ 0.06 & I & 900 & 0.92 & ESO NTT (EMMI) \\ 8.396 & 18.14 $\pm$ 0.35 & Ks & 2700 & & ESO 2.2m \\ 8.438 & 22.64 $\pm$ 0.08 & V & 900 & 1.93 & ESO NTT (EMMI) \\ 8.578 & 19.54 $\pm$ 0.08 & H & 1620 & & UKIRT \\ 8.608 & 20.28 $\pm$ 0.10 & J & 2160 & & UKIRT \\ 8.633 & 18.77 $\pm$ 0.24 & K & 600 & & UKIRT \\ 9.509 & 20.40 $\pm$ 0.12 & J & 2160 & & UKIRT \\ 9.554 & 19.76 $\pm$ 0.12 & H & 2160 & & UKIRT \\ 9.614 & 18.94 $\pm$ 0.09 & K & 2160 & & UKIRT \\ 10.353 & 21.62 $\pm$ 0.16 & gunn i & 4800 & 1.22 & ESO Dutch \\ 10.380 & 20.09 $\pm$ 0.20 & H & 3750 & & ESO 2.2m \\ 10.435 & 19.24 $\pm$ 0.16 & Ks & 3900 & & ESO 2.2m \\ 10.440 & 22.87 $\pm$ 0.34 & V & 900 & 1.06 & ESO NTT (EMMI) \\ 11.496 & 19.98 $\pm$ 0.21 & H & 2160 & & UKIRT \\ 11.527 & 19.47 $\pm$ 0.27 & K & 2160 & & UKIRT \\ 13.414 & 18.76 $\pm$ 0.30 & Ks & 4950 & & ESO 2.2m \\ 13.438 & 21.47 $\pm$ 0.41 & gunn i & 2400 & 1.98 & ESO Dutch \\ 13.558 & 20.42 $\pm$ 0.13 & J & 2160 & & UKIRT \\ 14.536 & 20.00 $\pm$ 0.15 & H & 2160 & & UKIRT \\ 14.545 & 19.36 $\pm$ 0.14 & K & 2160 & & UKIRT \\ 15.582 & 20.56 $\pm$ 0.12 & J & 3240 & & UKIRT \\ 17.359 & 22.68 $\pm$ 0.12 & V & 900 & 1.05 & ESO NTT (SUSI2) \\ 17.371 & 21.91 $\pm$ 0.12 & I & 900 & 0.75 & ESO NTT (SUSI2) \\ 23.501 & 20.04 $\pm$ 0.12 & H & 4860 & & UKIRT \\ 23.578 & 19.28 $\pm$ 0.11 & K & 4860 & & UKIRT \\ \end{tabular} \end{minipage} \caption[]{\tiny The log of the observations with the columns: UT Date, magnitude and error, filter, exposure time, seeing and the telescope. Instruments and CCDs used: NTT EMMI: red arm with TEK 2k $\times$ 2k CCD (\#36), 0.27\hbox{$^{\prime\prime}$} /pixel; NTT SUSI2: EEV 4k $\times$ 2k CCD (\#45 \& \#46), 0.08\hbox{$^{\prime\prime}$} /pixel; NTT SOFI: Hawaii 1k $\times$ 1k HgCdTe array, 0.29\hbox{$^{\prime\prime}$} /pixel; ESO Dutch: CCD Camera with TEK 512 $\times$ 512 CCD (\#33), 0.47\hbox{$^{\prime\prime}$} /pixel; Wise 1-m: TEK 1k $\times$ 1k CCD, 0.70\hbox{$^{\prime\prime}$} /pixel; CTIO 0.9-m: TEK 2k $\times$ 2k CCD, 0.38\hbox{$^{\prime\prime}$} /pixel; UKIRT: IRCAM3 with FPA42 256 $\times$ 256 detector, 0.29\hbox{$^{\prime\prime}$} /pixel; 2.2-m (IRAC2b): NICMOS-3 256 $\times$ 256 array, 0.507\hbox{$^{\prime\prime}$} /pixel. We note that we do not list an estimate of the seeing in case of infrared observations, since the real seeing is overestimated due to the process of co-adding the individual frames. \label{tab:log}} \end{table} \normalsize \begin{table} \caption[]{Fit parameters for the model $m$ = -2.5\,log$(10^{-0.4\,m_{0}}$ t$^{\alpha}$ + 10$^{-0.4\,m_{gal}})$ \label{tab:mag}} \begin{tabular}{lllll} \hline filter & $m_{0}$ & $\alpha$ & m$_{gal}$ & $\chi_{red}^{2}$ \\ \hline V & \gpm{21.22}{0.48}{0.33} & --1.61 & \gpm{23.04}{0.08}{0.08} & 5.5/5 \\ R & \gpm{21.18}{0.09}{0.08} & --1.61 & \gpm{22.58}{0.06}{0.05} & 14.7/10 \\ I & \gpm{20.60}{0.04}{0.04} & \gpm{-1.36}{0.27}{0.36} & \gpm{21.95}{0.25}{0.16} & 4.5/8 \\ J & \gpm{18.32}{0.33}{0.25} & --1.61 & \gpm{20.87}{0.07}{0.11} & 5.6/4 \\ H & \gpm{17.29}{0.06}{0.06} & \gpm{-1.67}{0.13}{0.15} & \gpm{20.27}{0.19}{0.15} & 6.5/7 \\ K & \gpm{16.48}{0.18}{0.15} & --1.61 & \gpm{19.62}{0.12}{0.11} & 11.3/9 \\ \hline \end{tabular} \end{table} \end{document}
\section{Introduction}\label{shapovalov:Intr} Soliton phenomena is an attractive f\/ield of present day research in nonlinear physics and mathematics. Essential ingredients in the soliton theory are the nonlinear Schr\"odinger equation (NLSE) and its variants appearing in a wide spectrum of problems. Examples are coupled non\-li\-near optics~\cite{shapo:Hasegawa,shapo:Mollenauer,shapo:Gordon,shapo:Nassar}, superconductivity~\cite{shapo:Tinkham,shapo:Floria}, and excitation in lattice systems~\cite{shapo:Christiansen}. More exactly, solitons are identif\/ied with a certain class of ref\/lectionless solutions of the equations integrable via the Inverse Scattering Transform method (see, for example,~\cite{shapo:Zakh1}). Such equations, including NLSE, are named soliton equations. At every instant a soliton is localized in a restricted spatial region with its centroid moving like a particle. The particle-like properties of solitons are also manifested in their elastic collisions. Soliton equations make up a narrow class of nonlinear equations, whereas a wider set of nonlinear equations, being nonintegrable in the framework of the IST, possess soliton- like solutions. They are localized in some sense, propagate with small energy losses, and collide with a varied extent of inelasticity. These solutions are termed soli\-ta\-ry waves (SWs), quasisolitons, soliton-like solutions, etc. to dif\/ferentiate them from the solitons in the above exact meaning. The stability of the localized form of solitons and SWs and their elastic collisions have led to interesting physical applications. It is of interest to study the inf\/luence of external f\/ields on the soliton propagation. To do that it is necessary to modify the original solitary equation by introducing variable coef\/f\/icients representing an external f\/ield potential that breaks the IST-integra\-bi\-lity. This problem was considered by variatio\-nal methods~\cite{shapo:Anderson}, the theory of soliton perturbations (see, for example, reviews~\cite{shapo:Kivshar,shapo:Maymistov}), or by using an appropriate ansatz~\cite{shapo:Nassar,shapo:Moura}. An investigation of soliton-like states is a separate problem for multi\-di\-men\-sional mo\-dels. The IST constructions are unsuitable for them except for the D4 self-dual Yang-Mills equations and their reductions in $1\le D \le 3$ (see, for example,~\cite{shapo:Ivanova} for details). The problem is more complicated in view of the singular behavior of NLSE in 2D~\cite{shapo:Zakh2}. Nevertheless, soliton-like solutions for NLSE on a plane can exist, as shown in~\cite{shapo:Jackiw} in terms of a suitable ansatz. Since the methods for constructing exact solutions for multidimensional models are restricted if compared to the cases of 2D models, approximate methods should be used. An ef\/fective approach to the problem can be developed based on the WKB-method~\cite{shapo:Maslov}. } Consider the generalized nonlinear Schr\"odinger equation for a ``matter'' f\/ield $\Psi (\vec x,t)$: \begin{equation} \label{shapovalov:1} \hspace*{-5pt}\begin{array}{l} \displaystyle -i\hbar \frac {\partial}{\partial t} \Psi+ \hat {\mathcal H}_{\rm nl}(\Psi ) = \Biggl \{ -i\hbar \frac {\partial}{\partial t} + \frac {1}{2m}\left(-i\hbar \nabla -\vec {\mathcal A}(\vec x,t)\right)^2 \vspace{3mm}\\ \displaystyle \hspace*{50mm} + V (\vec x,t) -2r|\Psi (\vec x,t)|^2 \Biggr\}\Psi (\vec x,t)=0. \end{array} \end{equation} Here, $\vec x \in {\mathbb R }^n$, $t\in {\mathbb R} ^1$; $V (\vec x,t) $, $\vec{\mathcal A}(\vec x,t)$ are given functions; $m$ is a real constant, $r$ is a real parameter of nonlinearity; $\hbar$ is Planck's constant playing the role of an asymptotic parameter; $|\Psi|^2=\Psi^*\Psi$, $\Psi^*$ is the complex conjugate of $\Psi$. In such a form equation~(\ref{shapovalov:1}) was studied in Refs.~\cite{shapo:Tinkham,shapo:Jackiw}. Since the f\/ield of application of~(\ref{shapovalov:1}) is wider, the physical meaning of the quantities entering in (\ref{shapovalov:1}) may be quite dif\/ferent from the quantum-mechanical one. In particular, in the one-dimensional problem of the propagation of an optical pulse~\cite{shapo:Mollenauer,shapo:Gordon} we have to assume for equation~(\ref{shapovalov:1}) $\vec x=x \in {\mathbb R} ^1$ with~$x$ being a normed temporal variable and $t$ being a normed space coordinate along which the pulse propagates. The function $\Psi (x,t)$ is an envelope of the pulse f\/ield. Equation (\ref{shapovalov:1}) takes the form \begin{equation} \label{shapovalov:2} \left \{-i\hbar \frac {\partial}{\partial t} + \frac {1}{2m}\left(i\hbar \frac {\partial }{\partial x} + {\mathcal A}(x,t)\right)^2 + V(x,t) - 2r|\Psi (x,t)|^2 \right\}\Psi (x,t)=0. \end{equation} The functions $V(x,t)$ and ${\mathcal A}(x,t)$ simulate the heterogeneity of the me\-dium. The symmetry of equation~(\ref{shapovalov:2}) was considered in Ref.~\cite{shapo:Broadbridge}, and asymp\-to\-ti\-cal solutions where studied for~(\ref{shapovalov:2}) with ${\mathcal A} =0$, $r=r(x,t)$ in Ref.~\cite{shapo:Subochev}. In the present work we have formulated the concept of semiclassi\-cally localized solutions for~ (\ref{shapovalov:1}), following the ideas of Ref.~\cite{shapo:Trifonov}. These solutions are the multidimensional analogues of the soliton, like the solutions for~(\ref{shapovalov:1}). The particle-like properties of a solitary wave are described in terms of the wave centroid. The latter is shown to move along with the bi\-charac\-teris\-tics of the basic symbol of the correspondent {\it linear} Schr\"o\-din\-ger equation. We construct asympto\-ti\-cal WKB-solutions for equation~(\ref{shapovalov:1}) and consider some examples. \renewcommand{\thefootnote}{\arabic{footnote}} \setcounter{footnote}{0} \section{Semiclassical concentraited solutions} \label{shapovalov:SCS} Soliton solutions are known to show particle properties. In classical mechanics, a particle is completely described by its phase orbit. Therefore, it is natural to introduce a similar concept for the soliton-like solutions of the nonlinear Schr\"odinger equation~(\ref{shapovalov:1}). The way by which to introduce the phase orbit seems to be obvious enough. It is based on the fact that in quantum mechanics, the f\/irst moments of a state $\Psi (\vec x,t)$ play the role of the phase orbit for the quantum system. Let now $\Psi (\vec x,t)$ be a solution of NLSE~(\ref{shapovalov:1}). The generalized position operators are $\hat{\vec x} (=\vec x)$ and their conjugate momentum variables are $\hat{\vec p}(=-i\hbar \nabla)$, \[ [{\hat x}_k, {\hat p}_s] = i\hbar \delta_{k,s}, \qquad k,s=\overline{1,n}. \] The mean value of an operator $\hat A$ by the solution $\Psi$ is def\/ined as \begin{equation} \label{shapovalov:3} \langle A \rangle =\frac{\langle\Psi|{\hat A}|\Psi\rangle}{\|\Psi\|^2}. \end{equation} Here, ${\|\Psi\|^2}=\langle\Psi|\Psi\rangle $; \[ \langle\Psi|{\hat A(t)}|\Psi\rangle = \int \Psi^*(\vec x,t){\hat A(t)}\Psi (\vec x,t) d\vec x, \] is a function of time for every operator $\hat A(t)$ and it parametrically depends on $\hbar$, \begin{equation} \label{shapovalov:4} \langle \vec x \,\rangle = \vec x(t,\hbar ), \qquad \langle \vec p \,\rangle = \vec p(t,\hbar ). \end{equation} If there exist the limits: \begin{equation} \label{shapovalov:5} \lim_{\hbar \to 0}\vec x(t,\hbar ) =\vec x(t), \qquad \lim_{\hbar \to 0}\vec p(t,\hbar ) =\vec p(t), \end{equation} then $\vec x(t)$ and $\vec p(t)$ are natural to be named the phase orbit of the classical system corresponding to the given solution $\Psi$. It is obvious that both the mean values~(\ref{shapovalov:4}) and the limit values~ (\ref{shapovalov:5}) depend on the solution $\Psi$ in the general case. Hence, the choice of the solution $\Psi$ meets the requirement for the expressions~(\ref{shapovalov:5}) to be a solution of the classical equations of motion. By analogy with quantum mechanics (see Ref.~\cite{shapo:Trifonov}), we can def\/ine the soliton-like solutions asymptotic in $\hbar \to 0$ as follows. \begin{Def}\label{shapovalov:D1} Let $z(t)=\{(\vec x(t),\vec p(t)),\,0\leqslant t\leqslant T\}$ be an arbitrary phase orbit in ${\mathbb R}^{2n}$. We name the solution $\Psi(\vec x,t,\hbar )$ of equation~(\ref{shapovalov:1}) as a semiclassically concentrated solution ({\sl SCS}) of the class ${\mathbb C}{\mathbb S}_S(z(t),N)$ $(\Psi\in{\mathbb C}{\mathbb S}_S(z(t),N))$ if: {\rm (i)} there exist the generalized limits \footnote{By generalized limit we mean the passage to the limit standardly defined in the distribution theory (see, for example, Ref. \cite{shapo:Rudin}).} \[ \lim_{\hbar \to0}\frac{|\Psi(\vec x,t,\hbar )|^2}{\|\Psi\|^2}=\delta (\vec x-\vec x(t)), \qquad \lim_{\hbar \to0}\frac{|\tilde\Psi(\vec p,t,\hbar )|^2}{\|\Psi\|^2}=\delta (\vec p-\vec p(t)); \] {\rm (ii)} there exist the centered moments \[ \Delta ^{(k)}_{\alpha,\beta}(t,\hbar ) = \langle \hat \Delta^{(k)}_{\alpha,\beta} \rangle, \qquad 0\leqslant k\leqslant N. \] \end{Def} Here, $\alpha$ and $\beta$ are multiindices, $|\alpha|+|\beta|=k$, $0\leqslant k\leqslant N$, and $\hat\Delta^{(k)}_{\alpha,\beta}$ is an operator with the symmetrized (Weyl) symbol $\Delta^{(k)}_{\alpha,\beta}(\vec p,\vec x)=$ $(\vec p-\vec p(t))^\alpha (\vec x-\vec x(t))^\beta$. Recall that the multiindex $\alpha$ is a vector of the form $\alpha=(\alpha_1,\dots,\alpha_n)$, where $\alpha_j\geqslant0$ are integer numbers. In addition, $|\alpha|=\sum\limits_{j=1}^n\alpha_j$ and for a vector $\vec\zeta=(\zeta_1,\dots,\zeta_n) \in{\mathbb R}^n$ we suppose $\vec\zeta^\alpha =\prod\limits_{j=1}^n \zeta_j^{\alpha_j}$. Note that the vectors $\vec x(t)$ and $\vec p(t)$ in Def\/inition~\ref{shapovalov:D1} are by no means connected with each other. The vector $z(t)=(\vec p(t),\vec x(t))$ is named the classical phase orbit of the system. Def\/inition~\ref{shapovalov:D1} specif\/ies the concept of solitary waves for asymptotic solutions of NLSE~(\ref{shapovalov:1}). \begin{theo} If a solution $\Psi$ of (\ref{shapovalov:1}) is semiclassically concentrated $(\Psi$ $\in $ ${\mathbb C}{\mathbb S}_S(z(t),N))$, then $z(t)=(\vec p(t)$, $\vec x(t))$ is a solution of a classical Hamilton system with the Hamiltonian \[ {\cal H}_{\rm cl}(\vec p, \vec x, t) = \frac {1}{2m}\left(\vec p -\vec{\mathcal A}(\vec x,t)\right)^2 + V (\vec x,t). \] \end{theo} \noindent {\bf Proof}. Let $\hat A$ be an operator, then the Ehrenfest theorem~\cite{shapo:Ehrenfest} is true for the mean value of $\hat A$: \[ \frac {d}{d t} \langle A \rangle =\langle \frac {\partial \hat A}{\partial t} \rangle + \frac {i}{\hbar } \langle [\hat{\mathcal H}_{nl} ,\hat A]\rangle . \] In particular, for the operators $\hat {\vec p}$ and $\hat {\vec x}$ we have: \begin{equation}\label{shapovalov:6} \frac {d}{d t} \langle {\vec p}\,\rangle = \frac {i}{\hbar } \langle [\hat{\mathcal H}_{\rm nl} ,\hat {\vec p}\,] \rangle,\qquad \frac {d}{d t} \langle {\vec x} \,\rangle = \frac {i}{\hbar } \langle [\hat{\mathcal H}_{\rm nl} ,\hat {\vec x}\,] \rangle . \end{equation} Using the obvious relations \[ \langle [ |\Psi (\vec x,t)|^2, \hat{\vec p}\,]\rangle = \langle [ |\Psi (\vec x,t)|^2, \hat{\vec x}\,]\rangle =0, \] we have from (\ref{shapovalov:6}): \begin{equation}\label{shapovalov:7} \frac {d}{d t} \langle {\vec p}\,\rangle = \frac {i}{\hbar} \langle [\hat{\mathcal H}_{\rm l} ,\hat {\vec p}\,]\rangle ,\qquad \frac {d}{d t} \langle {\vec x} \,\rangle = \frac {i}{\hbar} \langle [\hat{\mathcal H}_{\rm l} ,\hat {\vec x}\,]\rangle, \end{equation} where \[ \hat{\mathcal H}_{\rm l}= \frac {1}{2m}\left(i\hbar \nabla +\vec {\mathcal A}(\vec x,t)\right)^2 +V(\vec x,t). \] With Def\/inition~\ref{shapovalov:D1} of the SCS further proof coincides with a similar one for the linear case~\cite{shapo:Trifonov}. \medskip \noindent {\bf Remark.} Emphasize that, as follows from the theorem, the centroid of the SCS $\Psi$ moves along the bicharacteristics of the linear Schr\"odinger equation. \section{Asymptotic solutions } \label{shapovalov:Sol} In Section~\ref{shapovalov:SCS} we have discussed the def\/inition and the basic features of SCS. Here, we study a theoretical possibility of construction of the WKB-asymptotic semiclassically concentrated solutions of equation~(\ref{shapovalov:1}) on a limited time domain $0<t<T$ with $\hbar$-independent~$T$. Taking into account the form of the one-soliton solution of NLSE (see, for example,~\cite{shapo:Zakh1}), let us try solution of equation~(\ref{shapovalov:1}) in the form \begin{equation}\label{shapovalov:8} \Psi =\rho (\theta, \vec x,t,\hbar ) \exp \left[\frac {i}{\hbar }S(\vec x,t,\hbar )\right]. \end{equation} Here, $\theta =\hbar ^{-1}\sigma(\vec x,t,\hbar )$ is a ``fast'' variable; $\sigma(\vec x,t,\hbar )$, $\rho (\theta, \vec x,t,\hbar )$, and $S(\vec x,t,\hbar)$ are real functions regular in $\hbar$, that is: \[ S(\vec x,t,\hbar)= S(\vec x,t)+\hbar S_1(\vec x,t)+\cdots . \] The solution (\ref{shapovalov:8}) is assumed to be localized according to Def\/inition~(\ref{shapovalov:D1}). The derivative operators $\partial/\partial t$ and $\nabla $ act on the function (\ref{shapovalov:8}) as follows: \[ -i\hbar \frac \partial{\partial t}= -i\hbar \frac\partial{\partial t}\Big|_{\theta ={\rm const}}- i\sigma_{,t} \frac\partial{\partial \theta},\qquad -i\hbar\nabla= -i\hbar\nabla|_{\theta ={\rm const}}- i(\nabla \sigma ) \frac\partial{\partial \theta}, \] where $\sigma _{,t}=\partial \sigma/\partial t$. Henceforth we put ${\partial }/{\partial t}|_{\theta ={\rm const}} \equiv \partial _t$, $\nabla |_{\theta ={\rm const}}\equiv \nabla $, $\partial /\partial \theta \equiv \partial _\theta$. Substituting (\ref{shapovalov:8}) into (\ref{shapovalov:1}) we f\/ind: \[ \hspace*{-5pt}\begin{array}{l} \displaystyle \exp \left(\frac {i}{\hbar}S\right)\Biggl\{-i\hbar {\partial}_t + S_{,t} - i\sigma _{,t} \partial _\theta + V - \frac {\hbar ^2}{2m}\nabla ^2 - \frac {\hbar }{2m}(\nabla ^2\sigma )\partial _\theta -\frac {\hbar }{m}(\nabla \sigma \cdot \nabla )\partial _\theta \vspace{3mm}\\ \displaystyle \quad - i\frac {\hbar }{2m}(\nabla ^2 S) - \frac {1}{2m}(\nabla \sigma )^2 {\partial ^2}_{\theta \theta} - i\frac {\hbar}{m}(\nabla S \cdot \nabla )- i\frac {1 }{m}(\nabla S\cdot \nabla \sigma )\partial _\theta+ \frac {1}{2m}(\nabla S)^2 \vspace{3mm}\\ \displaystyle \quad + i\frac {\hbar}{m}({\vec A}\cdot \nabla ) + i\frac {1}{m}({\vec A}\cdot \nabla \sigma )\partial _\theta - \frac {1}{m}({\vec A}\cdot \nabla S) + i\frac {\hbar }{2m}(\nabla {\vec A}) + \frac {1}{2m}{\vec A}^2 - 2r \rho ^2 \Biggr\}\rho =0. \end{array} \] Let us gather $\hbar $-free terms in this equation and put their sum to zero. In the obtained equation we separate real and imaginary parts and then separate the ``fast'' variable $\theta$ from others. As a result we come to the following system of equations which determines the leading term of the asymptotic solution: \begin{equation}\label{shapovalov:9} \sigma _{,t} +\frac 1m\langle\left(\nabla S-\vec{\mathcal A}\right),\nabla \sigma \rangle=0, \end{equation} \begin{equation}\label{shapovalov:10} S _{,t} + V+ \frac 1{2m}\left(\nabla S-\vec{\mathcal A}\right)^2=r \tilde b(t,\vec x), \end{equation} \begin{equation}\label{shapovalov:11} \frac {1}{2m}(\nabla \sigma)^2 \rho _{,\theta \theta} +2 \rho^3=r\tilde b\rho. \end{equation} Here, $\langle \vec a, \vec b\rangle $ denotes the Euclidean scalar product of the vectors: $\sum\limits^n _{j=1} a_j b_j$; the function $\tilde b(t,\vec x)$ appears as a ``separation parameter'' in separating the ``fast'' variable $\theta$, and $\tilde b(t,\vec x)$ is to be determined in what follows. Let us look for $\rho (\theta ,\vec x,t, \hbar )$ in the class of functions satisfying the conditions: \begin{equation}\label{shapovalov:12} \lim_{\theta \to \infty }\rho (\theta , \vec x, t, \hbar)= \lim_{\theta \to \infty }\rho _{,\theta } (\theta , \vec x, t, \hbar)=0. \end{equation} Integrating equation~(\ref{shapovalov:11}) in view of (\ref{shapovalov:12}) we obtain: \begin{equation}\label{shapovalov:13} \rho _{,\theta}= \sqrt {\frac {2m r}{(\nabla \sigma )^2}} \sqrt{\tilde b -\rho^2} \rho. \end{equation} Let us put $r=\varkappa^2>0$ that corresponds to the existence of soliton solutions in the case when equation~(\ref{shapovalov:1}) is reduced to the standard NLSE (one-dimensional with $V (\vec x,t) =$, $\vec{\mathcal A}(\vec x,t)=0$)~\cite{shapo:Zakh1}. Then $\tilde b=b^2>0$ and further integration of~(\ref{shapovalov:13}) results in \begin{equation}\label{shapovalov:14} \rho =\frac {b}{\cosh \left\{b\varkappa\sqrt {2m (\nabla \sigma)^{-2}} \left(\frac {1}{\hbar }\sigma (\vec x,t)+\sigma _1(\vec x,t)\right)\right\}}, \end{equation} where the function $\sigma _1(\vec x,t)$ appears as a ``separation constant'' which is to be determined later. By def\/inition, the ``fast'' variable $\theta$ must have the structu\-re $\frac {1}{\hbar} \sigma (\vec x,t)$. For equation~(\ref{shapovalov:14}) to correspond this condition it is necessary to set \[ b\varkappa \sqrt {\frac {2m}{(\nabla \sigma )^2}}= {\rm const}=\alpha. \] Without loss of generality we can put $\alpha =1$ and then we have \begin{equation}\label{shapovalov:15} \rho = \sqrt {\frac {(\nabla \sigma )^2}{2m \varkappa ^2}} \; \frac {1}{\cosh \left(\frac {1}{\hbar }\sigma (\vec x,t)+\sigma _1(\vec x,t)\right)}. \end{equation} Equation (\ref{shapovalov:10}) takes the form \[ S _{,t} + V+ \frac 1{2m}\left(\nabla S-\vec{\mathcal A}\right)^2=\frac {1}{2m}(\nabla \sigma)^2, \] that, together with (\ref{shapovalov:9}), is equivalent to the single comp\-lex Ha\-mil\-ton-Jaco\-bi equat\-ion \begin{equation}\label{shapovalov:16} (S+i\sigma )_{,t} + V + \frac {1}{2m}\left[\nabla (S+i\sigma ) -\vec{\mathcal A}\right]^2=0. \end{equation} Thus, for the leading term of the asymptotic expansion~(\ref{shapovalov:8}), \[ \Psi (\vec x,t,\hbar )=\Psi ^0(\vec x,t,\hbar ) + O(\hbar ), \] we have: \begin{equation}\label{shapovalov:17} \Psi ^0 =\rho (\theta, \vec x,t,\hbar) \exp \left[\frac {i}{\hbar}S(\vec x,t)+S_1(\vec x,t)\right], \end{equation} where $\rho$ has the form (\ref{shapovalov:15}) and the functions $\sigma _1(\vec x,t), S_1(\vec x,t)$ are determined from successive approximations. The function~(\ref{shapovalov:17}) can be represented in the form: \begin{equation} \label{shapovalov:18} \Psi (\vec x, t, \hbar)= 2 \sqrt {\frac {(\nabla \sigma )^2}{2m \varkappa ^2}}\; \frac {\Psi _0(\vec x, t, \hbar)}{1+ |\Psi _0(\vec x, t, \hbar)|^2}, \end{equation} where \[ \Psi _0(\vec x, t, \hbar)=\exp \left \{\frac {i}{\hbar} \left[S(\vec x,t)+ i\sigma (\vec x,t)+ \hbar (S_1(\vec x,t)+i\sigma _1(\vec x,t))\right] \right\}. \] It can easily to show that $\Psi _0$ is an asymptotic solution of the {\it linear} Schr\"odinger equation: \begin{equation} \label{shapovalov:19} \left \{ -i\hbar \frac {\partial}{\partial t} + V (\vec x,t) +\frac {1}{2m}\left(-i\hbar \nabla -\vec{\mathcal A}(\vec x,t)\right)^2 \right\} \Psi _0 (\vec x,t)=O(\hbar ^\alpha ), \end{equation} where $\alpha =1$. Let us write a function $\Psi$, which satisf\/ies equation~(\ref{shapovalov:1}) to an accuracy of~$O(\hbar ^2)$, in the form \begin{equation} \label{shapovalov:20} \Psi = \Psi ^0(1+\hbar \Psi ^1). \end{equation} Here, $\Psi ^0$ is determined by expression (\ref{shapovalov:17}) and the function $\Psi ^1(\theta ,\vec x, t)$ is to be de\-ter\-mined. As can be seen from (\ref{shapovalov:15}), it is convenient to take the variable $\theta$ in the form \begin{equation} \label{shapovalov:21} \theta = \frac 1\hbar \sigma (\vec x,t)+\sigma _1(\vec x,t). \end{equation} Let us denote ${\rm Re}\, \Psi^1(\theta,\vec x,t)=u(\theta , \vec x,t)$, ${\rm Im}\,\Psi ^1(\theta , \vec x,t)=v(\theta , \vec x,t)$. Substituting (\ref{shapovalov:20}) into~(\ref{shapovalov:1}) and setting to zero sum\-mands at the equal powers of $\hbar $. Extract real and imaginary parts in the obtained equations and f\/ind a system of equations for function~(\ref{shapovalov:20}). This system includes the Hamilton-Jacobi equation~(\ref{shapovalov:16}) for the functions $S(\vec x,t)$, $\sigma (\vec x, t)$ and also the following equations determining the functions $S _1(\vec x,t)$ and $\sigma _1(\vec x, t)$: \begin{equation} S_{1,t}+\frac 1m\langle \left(\nabla S-\vec{\mathcal A}\right),\nabla S_1 \rangle - \frac 1m\langle \nabla \sigma,\nabla \sigma _1\rangle + \frac {1}{2m}\Delta \sigma+ \frac {1}{2m}\langle \nabla \sigma,\nabla \rangle \log (\nabla \sigma )^2 =0, \label{shapovalov:22} \end{equation} \begin{equation} \hspace*{-5pt}\begin{array}{l} \displaystyle \sigma _{1,t}+\frac 1m\langle \left(\nabla S-\vec{\mathcal A}\right), \nabla \sigma_1\rangle+ \frac 1m \langle \nabla \sigma, \nabla S_1\rangle \vspace{3mm}\\ \displaystyle \qquad - \frac 12\left [\frac 1m\langle \nabla ,\left(\nabla S -\vec{\mathcal A}\right)\rangle + \left[\log (\nabla \sigma )^2\right]_{,t}+ \frac 1m\langle \left(\nabla S-\vec{\mathcal A}\right),\nabla \log (\nabla \sigma)^2 \rangle \right]=0. \end{array} \hspace{-4.1pt}\label{shapovalov:23} \end{equation} The functions $u(\theta , \vec x,t)$, $v(\theta , \vec x,t)$ are given by the expressions \begin{equation} \hspace*{-5pt}\begin{array}{l} \displaystyle \rho (\theta ,\vec x,t)u(\theta ,\vec x,t)= \sqrt { \frac {2m}{\varkappa ^2(\nabla \sigma )^2}}\; \frac {1}{\cosh \theta } \Biggl\{ C_1(\vec x,t) \tanh \theta+ \frac {1}{2m}\langle \nabla \sigma , \nabla \sigma _1\rangle \vspace{3mm}\\ \displaystyle \qquad + \frac {1}{12m}\left[\Delta \sigma + \langle \nabla \sigma , \nabla \log (\nabla \sigma )^2\rangle \right] \left(\sinh \theta \cosh \theta- \epsilon \cosh ^2 \theta \right)\Biggr\}, \end{array} \label{shapovalov:24} \end{equation} \begin{equation} \hspace*{-5pt}\begin{array}{l} \displaystyle \rho (\theta ,\vec x,t)v(\theta ,\vec x,t)= \sqrt { \frac {2m}{\varkappa ^2(\nabla \sigma )^2}} \Biggl\{ \frac {C_1(\vec x,t)}{ \cosh \theta }+ \frac {1}{4}\left[\frac 1m \langle \nabla ,\nabla S - \vec{\mathcal A} \rangle \right. \vspace{3mm}\\ \displaystyle \left. \qquad + \left(\partial _t + \frac {1}{m}\langle \left(\nabla S-\vec{\mathcal A}\right), \nabla \rangle \right) \log (\nabla \sigma )^2\right] (\epsilon \sinh \theta -\cosh \theta ) \Biggr\}. \end{array} \label{shapovalov:25} \end{equation} Here, $\epsilon ={\rm sign} (\sigma)$ and the function $C_1(\vec x, t)$ is determined by successive approximations. \medskip \noindent {\bf Remark.} From relations (\ref{shapovalov:22}) and (\ref{shapovalov:23}) it follows that (\ref{shapovalov:19}) is accurate to $O(\hbar ^2)$. \medskip Relation (\ref{shapovalov:18}) can be considered as a transformation con\-nec\-ting the asymptotic solutions of the nonlinear equation~(\ref{shapovalov:1}) with the solution of the linear equation (\ref{shapovalov:19}) for $\alpha=2$. So, the leading term $\Psi^0$ of the asymptotic solution of equation~(\ref{shapovalov:1}) is completely f\/ixed by expressions (\ref{shapovalov:15}), (\ref{shapovalov:17}), (\ref{shapovalov:20}), and~(\ref{shapovalov:21}), where the functions $S(\vec x,t)$, $\sigma (\vec x,t)$, $S_1 (\vec x,t)$, and $\sigma _1(\vec x,t)$ are determined by the system of equations~(\ref{shapovalov:16}), (\ref{shapovalov:22}), and~(\ref{shapovalov:23}). \section{Special solutions }\label{shapovalov:Special} For a more detailed study of the above asymptotic solutions~(\ref{shapovalov:17}), let us consider some special cases of these solutions. \subsection{One-dimensional NLSE }\label{shapovalov:NS} Let us put $\vec x=x \in {\mathbb R}^1$, $\Delta =\partial ^2/\partial x^2$ and $\vec{\mathcal A}=V=0$ in (\ref{shapovalov:1}); then equation~(\ref{shapovalov:1}) takes the form \begin{equation}\label{shapovalov:26} \left[i\hbar \frac{\partial}{\partial t} + \frac {\hbar ^2}{2m} \frac{\partial ^2}{\partial x^2}+ 2\varkappa ^2 |\Psi (x,t,\hbar)|^2\right]\Psi (x,t,\hbar )=0. \end{equation} The leading term of the asymptotic solution (\ref{shapovalov:20}) is as follows: \begin{equation} \label{shapovalov:27} \Psi (x,t)=\rho (x,t,\hbar ) \exp \left[\frac {i}{\hbar }(S(x,t) + \hbar S_1(x,t))\right], \end{equation} where \begin{equation} \label{shapovalov:28} \rho (x,t,\hbar )=\sqrt {\frac {\sigma _{,x}^2}{2m \varkappa^2}} \; \frac {1}{\cosh \left(\frac {1}{\hbar } \sigma (x,t)+ \sigma _1(x,t)\right)}. \end{equation} Equations (\ref{shapovalov:16}), (\ref{shapovalov:22}), and (\ref{shapovalov:23}) become: \begin{equation} (S+i \sigma )_{,t}+\frac {1}{2m} (S+i \sigma)_{,x}^2=0,\label{shapovalov:29} \end{equation} \begin{equation} S_{1 ,t}+\frac {1}{m} S_{,x}S_{1,x}-\frac 1m\sigma _{,x}\sigma _{1,x} +\frac {3}{2m}\sigma _{,xx}=0,\label{shapovalov:30} \end{equation} \begin{equation} \sigma _{1,t} +\frac {1}{m}S_{,x}\sigma _{1,x}+ \frac {1}{m} \sigma _{,x}S_{1,x}-\frac 12\left[\frac 1m S_{,xx}+ \frac {2}{\sigma _{,x}}\sigma _{,xt}+ \frac {2}{m}\frac {S_{,x}}{\sigma _{,x}}\sigma _{,xx}\right]=0.\label{shapovalov:31} \end{equation} To construct solution (\ref{shapovalov:27}) and (\ref{shapovalov:28}), let us look for a special solution of equation~(\ref{shapovalov:29}) in the form \begin{equation}\label{shapovalov:32} S=\alpha _1t+\alpha _2x +\varphi _0,\qquad \sigma =\beta _1 t+\beta _2 (x-x_0), \end{equation} where $\alpha _1$, $\alpha _2$, $\beta _1$, $\beta _2$, $\varphi _0$, and $x_0$ are real constants. Substitution (\ref{shapovalov:32}) into (\ref{shapovalov:29}) gives \[ \alpha _1= \frac {1}{2m}(\beta _2^2-\alpha _2^2), \quad \beta _1=-\frac {1}{m}\alpha _2\beta _2. \] Substituting (\ref{shapovalov:32}) into (\ref{shapovalov:30}) and (\ref{shapovalov:31}), we easily obtain the following complex equation for the functions $S _1$ and $\sigma _1$: \begin{equation} \label{shapovalov:33} (S_1 +i \sigma _1)_{,t}+ \frac {\alpha _2+ i\beta _2}{m}(S_1 +i \sigma _1)_{,x} =0. \end{equation} Equation (\ref{shapovalov:33}) is a complex wave equation the solution of which can be written as \begin{equation}\label{shapovalov:34} w (x,t)\equiv S_1 (x,t)+i \sigma _1(x,t)=f(x- at). \end{equation} Here, $a\equiv (\alpha _2+ i\beta _2)/m$ and $f(\zeta )$ are analytical functions of the complex variable $\zeta= x-at$. Denote $\beta _2=2\eta$ and $\alpha _2=2 \xi$, then solution (\ref{shapovalov:27}) takes the form \begin{equation} \label{shapovalov:35} \hspace*{-5pt}\begin{array}{l} \displaystyle \Psi =- \frac {2\eta }{\varkappa \sqrt 2m}\; \frac {1}{\cosh \left[\frac {2\eta }{\hbar }\left(x-x_0-\frac {2\xi}{m}t\right)+ {\rm Im}f(x-at)\right]} \vspace{3mm}\\ \displaystyle \qquad \times \exp \left [\frac {i}{\hbar }(2\xi x-\frac {2}{m} (\xi ^2-\eta ^2)t+\varphi _0+\hbar {\rm Re}\,f(x-at) )\right]. \end{array} \end{equation} If $f (x-at)=0$ in (\ref{shapovalov:34}), then (\ref{shapovalov:35}) takes the form of the exact one-soliton solution of the nonlinear Schr\"odinger equation (\ref{shapovalov:26}) which is reduced to the standard form~\cite{shapo:Zakh1} when $\hbar =m=1$, $\varkappa ^2=1/2$. If $f(x-at)\neq 0$, we have an asymptotic solution similar to the one-soliton solution. \subsection{NLSE with a separated potential}\label{shapovalov:NSV} Consider equation~(\ref{shapovalov:26}) with a potential $V(x,t)$: \begin{equation} \label{shapovalov:36} \left[i\hbar \frac {\partial}{\partial t} + \frac {\hbar ^2}{2m} \frac{\partial^2}{\partial x^2} + 2\varkappa ^2 |\Psi (x,t,\hbar )|^2-V(x,t)\right]\Psi (x,t,\hbar )=0. \end{equation} Solution (\ref{shapovalov:27}) and (\ref{shapovalov:28}) is determined by (\ref{shapovalov:30}), (\ref{shapovalov:31}), and \begin{equation} \label{shapovalov:37} (S+i \sigma )_{,t}+\frac {1}{2m} (S+i \sigma )_{,x}^2+V=0. \end{equation} For the potential $V$ of the separated form, \[ V(x,t)=v_0(t)+v_1(x), \] we can easily f\/ind for the system (\ref{shapovalov:37}), (\ref{shapovalov:30}), and (\ref{shapovalov:31}) two classes of separated solutions determining~(\ref{shapovalov:27}) and~(\ref{shapovalov:28}). The f\/irst class is described by the expressions: \[ \hspace*{-5pt}\begin{array}{l} \displaystyle S(x,t)=c_1 t- \int v_0(t) dt, \qquad \sigma (x,t)=\int \sqrt {2m(c_1+v_1(x))} dx, \vspace{3mm}\\ \displaystyle S_1(x,t)=c_2t+c_3, \qquad \sigma _1=(3/2)\log|\sigma_{,x}|+ mc_2 \int \sigma _{,x}^{-1} dx +c_4, \end{array} \] where $c_1,\dots ,c_4= \mbox{const}$. The second class is presented by \[ \hspace*{-5pt}\begin{array}{l} \displaystyle S(x,t)=c_3 t- \int v_0(t) dt +c_4, \qquad \sigma (x,t)=c_1 \left[t-m\int \frac {dx}{p'(x)}\right]+c_2, \vspace{3mm}\\ \displaystyle S_1(x,t)=a_1t+a_3+f(x), \qquad \sigma _1=a_2t+a_4+g(x), \end{array} \] Here, $c_1,\dots ,c_4; a_1,\dots ,a_4 = \mbox{const}$ and the functions $p(x)$, $f(x)$, $g(x)$ are determined in quadratures from the following equations: \[ \hspace*{-5pt}\begin{array}{l} \displaystyle \frac 1m (p'(x))^2=-(v_1(x)+c_3)+\sqrt {(v_1(x)+c_3)^2+c_1^2}, \vspace{3mm}\\ \displaystyle g'(x)= \frac {m}{p'(x)}\left [\frac {c_1}{p'(x)} f'(x)- \frac {1}{2m}p''(x)-a_2\right], \vspace{3mm}\\ \displaystyle [(p'(x)]^4+c_1^2m^2)f'(x)=c_1m^2a_2p'(x)-ma_1[p'(x)]^3- c_1mp'(x)p''(x). \end{array} \] The above solutions show the potentialities of separation of variables as applied to the solution of the general system of equations (\ref{shapovalov:16}), (\ref{shapovalov:22}), and (\ref{shapovalov:23}). \subsection{Cylindrical coordinates}\label{shapovalov:NSC} The self-focusing ef\/fect of a beam of high power optical radiation propa\-ga\-ting along the $z$-axis is described by a nonlinear Schr\"odinger equation of the form~\cite{shapo:Landau}: \begin{equation} \label{shapovalov:38} \left[i\hbar \frac{\partial }{\partial t} + \frac {\hbar ^2}{2m} \Delta _{\perp} + 2\varkappa ^2 |\Psi (x,y,t,\hbar )|^2\right]\Psi (x,y,t,\hbar )=0. \end{equation} Here, $t$ is the time coordinate in the coordinate system moving with a group velocity along the direction of the radiation propagation. $\Delta _{\perp} =\partial ^2/\partial x^2+\partial ^2/\partial y^2$ is the Laplace operator in a plane orthogonal to the radiation direction. A function $V(t,x,y)$ simulates some nonstationarity and heterogeneity of the medium where the pulse is propagating. In the stationary statement of the problem, the variable $z$ takes the place of $t$. Let us f\/ind asymptotic solutions of the form (\ref{shapovalov:20}) for equation~(\ref{shapovalov:38}) in the cylindrical coordinates \begin{equation}\label{shapovalov:39} x=r\cos \varphi, \qquad y=r\sin \varphi. \end{equation} The leading term of the asymptotic solution (\ref{shapovalov:20}) takes the form \begin{equation} \label{shapovalov:40} \Psi (r,\varphi ,t,\hbar )= \rho (r,\varphi ,t,\hbar ) \exp \left\{\frac {i}{\hbar}[S(r,\varphi ,t) + \hbar S_1(r,\varphi ,t)]\right\}, \end{equation} where \begin{equation} \label{shapovalov:41} \rho (r,\varphi ,t,\hbar )=\sqrt { \frac {(\nabla _{\perp }\sigma )^2}{2m \varkappa ^2}} \, \frac {1}{\cosh \left(\frac {1}{\hbar} \sigma (r,\varphi ,t)+ \sigma _1(r,\varphi ,t)\right)}, \end{equation} $\nabla_{\perp}$ is the gradient operator with respect to the variables $r$ and $\varphi$ (\ref{shapovalov:39}). The special solution, illustrating a transverse heterogeneity of the optical pulse, can be written as follows: \[ \hspace*{-5pt}\begin{array}{l} \displaystyle \Psi (r,\varphi ,t,\hbar )=\frac {c_1}{\varkappa \sqrt {2m}} \cosh ^{-1} \left[\left(\frac {c_1}{\hbar }+a_2\right)r+c_1b_1t+ \frac 12 \log r + \frac {a_1}{\hbar }+a_3\right] \vspace{3mm}\\ \displaystyle \qquad \times \exp \left \{i \left[\left(\frac{c_1^2}{2m \hbar }+ \frac {a_2c_1}{m}\right)t - mb_1r+\frac {c_2}{\hbar }+c_3\right] \right\}, \end{array} \] where $c_1,c_2,c_3,a_1,a_2,a_3,b_1=\mbox{const}$; $c_1\neq 0$. \section{Conclusion}\label{shapovalov:Conc} An outcome of the work are the asymptotic solutions of the solitary wave type for NLSE~(\ref{shapovalov:1}) obtained for a f\/inite time interval $0<t<T$. The question of the validity of long-time asymptotics requires a special consideration, since, even for the linear case, this problem has received rather much attention in the literature~\cite{shapo:Voros}. Solitary wave solutions are considered here similar to quantum wavepackets. This permits one to investigate the particle-like properties of the SWs in terms of the Def\/inition~\ref{shapovalov:D1} based on the Ehrenfest theorem. The equations of motion for the centroid of an SW are found to be independent of the nonlinearity factor $r$ in~(\ref{shapovalov:1}), while the solution itself depends essentially on $r$. In such an approach, the centroid of the solitary wave solution moves like a classical particle in the external f\/ield described by the potentials $(V, \vec {\mathcal A})$ in~(\ref{shapovalov:1}). This completely corresponds to the well-known soliton properties for the one-dimensional case~\cite{shapo:Moura}. Such a situation takes place only when the nonlinearity factor is constant. If $r=r(\vec x,t)$, the f\/ield $\Psi (\vec x,t,\hbar )$ results in the appearance of additional classical variables, and the correspondent equations of motion become much more complicated. The problems discussed are beyond the scope of this work and requires a special investigation. \subsection*{Acknowledgements} This work was partially supported by the Russian Foundation for Basic Research, Grant No 98-02-16195.
\section{Introduction} The large amount of dust and gas in the direction of the galactic bulge is known to be responsible for the large extinction which makes very difficult the observation of the bulge in several bands. However, it has long been known that the irregular distribution of dust and gas presents some "holes" where the extinction is considerably lower. The best known of these so-called windows has been found by Baade (1951) and extensively studied since then. Generally speaking, these low-extinction windows offer a unique opportunity to obtain information about the bulge in the optical range. Several groups (MACHO, OGLE, for example) have been recently monitoring the bulge in search of microlensing events (see Paczy\'nski (1996) for a review). Since April 1997 we have been conducting a small monitoring project of 12 selected low-extinction windows (see Table 1), taken from Blanco \& Terndrup (1989) and Blanco (1988), using the recently refurbished Meridian Circle of the Abrah\~ao de Moraes Observatory (see below). The main objective of the project is the discovery and classification of variable stars. The final goal will be to have an on-line, real time processing data to stimulate the study of potentially interesting events by other observing facilities. Meanwhile, we have faced the problem that the total number of reference stars of the usual catalogues inside the observed low-extinction fields (13' in declination and between 3 and 6 minutes in right ascension) is small. This posed a serious difficulty for the reduction of data and the search for variable objects. As a first step towards a comprenhensive study of the windows, we have attempted to construct dense (secondary) reference catalogues intended to be of general use. The publication of the Tycho Catalog (ESA, 1997) has added an additional motivation to our task, since the refinement of the latter beyond the completeness limit ($V \, \sim \, 11$) was already foreseen as an useful tool for future observations. We shall describe in the next Section the selected fields. Section III is devoted to a presentation of the instrumental facilities and data reduction. The criteria of the selection of the set of standards is discussed in Section IV. A general discussion in Section V, the light curves in Appendix A and the new catalogues in Appendix B closes the present work. \section{Selection of the fields} Table I summarizes the list of low-extinction windows in which we have defined a relatively dense set of standards for further use. Due to specific needs of our project, but also in view of more general applications, we have proceeded to select standard stars up to $m_{Val} \, = \, 13$ (we will use "$m_{Val}$" to designate the magnitudes obtained with the Valinhos system ($V_{V}$ filter) and "$V$" for magnitudes in the standard Jonhson system ($V_{J}$ filter)). We should emphasize that, since the primary objective was to construct an astrometric and photometric reference set, the selected stars do {\it not} constitute a {\it complete} sample up to the stated magnitude limit. It is also clear that, due to instrumental limitations and sky quality of the site, the extension of quality observations to fainter magnitudes is not devoided of problems. However, according to the performed tests and preliminary results (Dominici, 1998), we are be able to monitor variable stars up to $m_{Val} \, =\, 14.5$; and construct complete stellar catalogues up to $m_{Val} \, \sim \, 15.5$, the latter limit being slightly dependent on the field declination (see below). This relatively shallow sampling naturally means that most of our standards do {\it not} reside in the bulge but are just foreground stars. \begin{table}[tb] \caption{Coordinates of the 12 low-extinction windows for wich secondary catalogs based on Tycho data are presented (Blanco \& Terndrup, 1989 and Blanco, 1988) } \label{tabrmm} \begin{flushleft} \begin{tabular}{@{}lcrcrr@{}} \hline \rule{0pt}{1.2em}Window& \# & $\alpha$ (J2000) &$\delta$ (J2000) & b ($^{o}$)&l ($^{o}$)\\ \hline BE & 5 & 18 10 17 & -31 45 49 & -6.02 & 0.20 \\ BG & 7 & 18 18 06 & -32 51 24 & -7.99 & 0.00 \\ BJ & 10 & l8 40 49 & -34 49 48 & -13.10 & 0.27 \\ LA & 11 & 16 56 57 & -52 44 53 & -6.02 & -24.77 \\ LB & 12 & 17 18 27 & -48 05 40 & -6.01 & -18.96 \\ LC & 13 & 17 29 22 & -45 20 05 & -6.03 & -15.61 \\ LD & 14 & 17 44 59 & -40 41 37 & -5.99 & -10.14 \\ LI & 19 & 18 04 48 & -33 47 46 & -5.96 & -2.11 \\ LR & 27 & 18 23 32 & -26 10 16 & -5.98 & +6.54\\ LT & 29 & 18 34 09 & -21 23 35 & -6.04 & +9.87 \\ LU & 30 & 18 51 26 & -13 18 41 & -6.03 & +21.04 \\ LV & 31 & 18 57 37 & -10 23 35 & -6.08 & +24.35 \\ \hline \end{tabular} \end{flushleft} \end{table} \section{Observations and data reduction} The Askania-Zeiss Abrah\~ao de Moraes Meridian Circle, operated at the IAG/USP Valinhos Observatory ($\phi$ = $46^{o}$ 58' 03'', $\lambda$ = $-23^{o}$ 00' 06''), is a 0.19 m refractor instrument and 2.6 m focal distance. Recently, a CCD detector was installed as a part of the continuing Bordeaux Observatory-IAG/USP collaboration (Viateau et al., 1998). The capabilities of the Meridian Circle have thus been greatly increased and allowed reliable photometric (as well as the astrometric) measurements to be performed. The CCD detector Thomson 7895A installed at Valinhos has a $512 \; \times \, 512$ pixel matrix, with a pixel scale of $1.5"/pixel$. It is cooled down to $- 40^{o} \, C$ by a two stage termocouple system. The observations are performed in a drift-scanning mode, that is, the speed of charge transfer by the CCD is the same as the speed of stellar transit for a fixed instrument position. Therefore the integration time for a given declination $\delta$ is $$ t_{int} \, = \, 51 \, \sec \delta \, s \, \; \; (1)$$ The observed field has a width of $13'$ in declination by an arbitrary time in right ascention (some minutes or hours). After the installation of the CCD detector, it became possible to obtain images of several thousands of stars per night, up to magnitudes as faint as $m_{Val} \, = \, 16.5$, depending on the transit time as reflected by eq.(1). Although the accuracy of the position and magnitude measurements depend on several factors, like the magnitude of the observed star, they have been checked to be better than 0.05 of arcsec for the position and 0.05 for the magnitude (inside the optimal magnitude interval $9 \, < \, m_{Val} \, < \, 14$, see Figure 1). \bigskip Figure 1:Precision of the positions and magntudes for one typical observation night. \bigskip The filter we have used in this and other observational programmes is somewhat wider than the standard Johnson filter $V_{J}$; allowing a larger coverage towards the infrared band in order to maximize the number of objects by taking advantage of the better quantum efficiency of the CCD in that region. Figure 2 shows the response of the Valinhos filter $V_{V}$ together with the standard Johnson filter $V_{J}$, both curves already weighted with the CCD detector efficiency. \bigskip Figure 2:The response of the Valinhos and Jonhson filters, already weighted with detector quantum eficiency. \bigskip In order to evaluate the difference between our magnitude system and Johnson's visual band, we selected a filter (kindly provided to us by the Laborat\'orio Nacional de Astrof\'\i sica - LNA) capable to roughly reproducing the necessary spectral characteristics to mimic the $V_{J}$ band. We used this filter to perform a full observation night, aiming at several fields for wich we had good data with the broader filter. For the photometric reduction we made use of the Tycho catalogue $V$ magnitude. A correlation of the magnitude difference $V \, - \, m_{Val}$ with the (poorly determined) color index $(B - V)$ as reported in the Tycho catalog is shown in Figure 3 and can be fitted by $$ V \, - \, m_{Val} \, = 0.14 \, - \, 0.07 (B - V) \, \, \, (2)$$ However, a large scattering of the data having various origins is present and precludes a straightforward utilization of eq.(III.2). Moreover, given the intrinsic capabilities of the Meridian Circle we are not able to measure $B$ band magnitudes properly because of focusing and detector limitations. This means that even though our claims about the photometric quality and long-term stability of the selected standards are quite secure, those stars should be recalibrated whenever other system is needed for purposes other than differential photometry. With this shortcome in mind, an approximate relationship between filters may be obtained as $m_{Val} \, = \, 1.02 V \, - \, 0.28 $ for quick references purposes only. \bigskip Figure 3:Correlation of the magnitude difference ($V - m_{Val}$) with the color index $(B-V)$ \bigskip To select a set of standard stars we have performed 5 to 6 observations of $\sim \, 1^{h}$ in right ascention each, centered in the selected windows. The main objective was to include as many Tycho stars as possible in each field to link the selected set to them. In addition, 10 short observations centered around the field coordinates given in Table 1, with a total duration of 3 to 6 minutes each (this is the standard duration of the observations), were employed to construct light curves for the candidates to reference stars and to check the stability of their magnitudes. The employed data reduction method requires a preliminar reduction, with the sky background subtracted by a linear polynomial fitted to each pixel column. Objects are identified when 3 consecutive pixels with a $2\sigma$ confidence level are detected, where $\sigma$ is the standard deviation of the mean count rate in each column. A two-dimensional Gaussian curve is adjusted to the flux distribution of the objects, to obtain the {\it x} and {\it y} centroid coordinates, the flux and respective errors. In the following step, the celestial positions and magnitudes are calculated by a global reduction, using the field overlap among all observation nights (Eichhorn, 1960, Benevides-Soares \& Teixeira, 1992, and Teixeira et al., 1992). In this procedure, first of all, each night is reduced independently of the remaning data by solving the following by least squares system, with respect to reference catalogue stars: $$ \alpha \, = \, a_{0} \, + \, a_{1} x \, + \, a_{2} y \; \; (3)$$ $$ \delta \, = \, b_{0} \, + \, b_{1} y \, + \, b_{2} x \; \; (4)$$ $$ m_{Val} \, = \, m_{Val0} \, - \, 2.5 \log F \; \; \; (5) $$ where $a_{0}$ is the initial sidereal time of the observation of the field, $b_{0}$ is the central declination of the field and $x$ and $y$ are the relative coordinates. The $x$ coordinate is measured in the diurnal motion direction while the $y$ coordinate is in the perpendicular direction. The terms $a_{1}$ and $b_{1}$ are the scale in $x$ and $y$, respectively, and $a_{2}$ and $b_{2}$ are corrections in the CCD orientation. A more robust model for the magnitude computation, using color index reported in Tycho catalog, are in development, but our tests demonstrated that actual results are confiable. After this first reduction, the system of eqs.(3-5) is again solved by using an iterative process, now for all stars detected in the field. At each step of iteration, the system is solved by least squares (Benevides-Soares \& Teixeira, 1992 and Teixeira et al., 1992). The process converges in a few iterations (typically less than 10 steps). The Tycho Catalog is presented in the form of a multicolumn file in which a quality criterion is given for each object. This criterion guides the further utilization of the stars as standards for any purpose. We have concluded that the construction of secondary catalogues referred to Tycho stars of quality index worse than 5 can compromised the desired accuracy, and therefore we excluded such Tycho stars. Stars brighter than $m_{Val} \, = \, 8.5$ have been also excluded because they have a saturated image in the CCD and their use would spoil the magnitude mesurement of all the stars in the field. All the known variable stars were also taken out according to the GCVS (Khopolov et al, 1988) and NSV (Kukarkin et al, 1982) Catalogues, and the new variables found by the HIPPARCOS mission. With this criteria the final number of Tycho stars present in the "short" exposure lies between 1 and 8, and in the "long" exposures, between 31 and 89 stars. \section{Construction of the Reference Catalogue} After reduction of the "long" images, as described in the previous Section, we have separated and relabeled all the stars brighter than $m_{Val} \, =\, 13$ that appeared in all the observational runs. They constitued the first set of candidates for a reference catalogue. A visual inspection was performed to eliminate those stars located near the edges of the images and those with very near apparent neighbours (which may have complicated the magnitude determination and introduce errors in position coordinates as well). The number of "long" observations was certainly enough to guarantee the astrometric accuracy of the candidates, but not their photometric stability. Thus, to detect possible variables among our reference candidates, we have employed 10 "short" exposure observations (in principle obtained for the monitoring project) in addition to the "long" ones to construct the appropriate light curves. For a few fields, the "short" observations did not contain the minimum of four Tycho stars, necessary for the reduction procedure. In these cases, the number of reference stars was increased with candidates stars from the "long" observations. On the other hand, this procedure allowed us to check the sensitivity of the results to a change in the input standard stars. Once we had constructed all the light curves for the remaining candidates we proceeded to eliminate those which had large magnitude and positions variations based on the derived parameters: position and magnitude average values, their corresponding standard deviations($\sigma_{\alpha}$, $\sigma_{\delta}$, $\sigma_{m_{Val}}$), the average computed for the magnitude weighted with the error bars in each night and weighted error $\sigma_{w}$. This selection eliminated stars having a magnitude standard deviation larger than 0.1; right ascention errors larger than $0.01^{s}$ and declination errors larger than $0.1"$. In addition, stars for which the mean value of the magnitude was significantly different from the average weighted with the nightly error bars were also discarded (see examples of the light curves in Appendix A, Figs.4-6). In Table 3 we present the average values of the errors in position and magnitude and the number of stars in the final catalog. \begin{table}[tb] \caption{Average position and magnitude precisions for reference stars in the low-extinction windows. The last colum indicate the number of stars in the final catalogs} \label{tabrmm} \begin{flushleft} \begin{tabular}{@{}lcrccc@{}} \hline \rule{0pt}{1.2em}Window& $\sigma_\alpha(^{s})$& $\sigma_\delta('')$& $\sigma_{m_{Val}}$& $\sigma_{w}$& N\\ \hline BE & 0.0017 & 0.015 & 0.028 & 0.0043 & 60\\ BG & 0.0021 & 0.014 & 0.026 & 0.0040 & 41\\ BJ & 0.0021 & 0.013 & 0.027 & 0.0041 & 43\\ LA & 0.0019 & 0.011 & 0.035 & 0.0038 & 55\\ LB & 0.0019 & 0.011 & 0.033 & 0.0039 & 36\\ LC & 0.0014 & 0.011 & 0.029 & 0.0035 & 61\\ LD & 0.0015 & 0.011 & 0.027 & 0.0036 & 47\\ LI & 0.0010 & 0.011 & 0.031 & 0.0038 & 20\\ LR & 0.0025 & 0.014 & 0.031 & 0.0048 & 37\\ LT & 0.0021 & 0.014 & 0.028 & 0.0049 & 32\\ LU & 0.0017 & 0.013 & 0.028 & 0.0045 & 31\\ LV & 0.0015 & 0.016 & 0.034 & 0.0048 & 44\\ \hline \end{tabular} \end{flushleft} \end{table} \section{Concluding remarks} The data from standard stars appearing in this work are the result of the best quality images obtained in April-September 1997 (approximately 54 nights in total, average of 30 observations per window), reduced and processed according to the "global" procedures discussed in Section II. Some stars preliminary selected as standards were in fact eliminated after the end of 1997 campaign because they did not fully satisfy the selection criteria in the long run. We note that all these standards are being observed daily (whenever observational conditions are satisfactory) as a byproduct of our monitoring programme, and therefore the results of our catalogues are continuously being improved. Upgraded versions and the finding charts of the catalogues can be requested electronically. \begin{acknowledgements} We would like to thanks the financial support by FAPESP, CAPES, CNPq, Bordeaux Observatory and IAG/USP, R.D.D da Costa and the referee, F.Mignard, for useful suggestions that helped to improve the first version of the present work. \end{acknowledgements}